18.01.2013 Views

Human-like Control of Dynamically Walking Bipedal Robots

Human-like Control of Dynamically Walking Bipedal Robots

Human-like Control of Dynamically Walking Bipedal Robots

SHOW MORE
SHOW LESS

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

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

<strong>Human</strong>-<strong>like</strong> <strong>Control</strong> <strong>of</strong><br />

<strong>Dynamically</strong> <strong>Walking</strong> <strong>Bipedal</strong> <strong>Robots</strong><br />

Tobias Luksch<br />

Vom Fachbereich Informatik der<br />

Technischen Universität Kaiserslautern<br />

zur Verleihung des akademischen Grades<br />

Doktor der Ingenieurwissenschaften (Dr.-Ing.)<br />

genehmigte Dissertation.


Zur Begutachtung eingereicht am: 26. Oktober 2009<br />

Datum der wissens. Aussprache: 18. Mai 2010<br />

Vorsitzender: Pr<strong>of</strong>. Dr. rer. nat. Rolf Wiehagen<br />

Erster Berichterstatter: Pr<strong>of</strong>. Dr. rer. nat. Karsten Berns<br />

Zweiter Berichterstatter: Pr<strong>of</strong>. Dr.-Ing. Rüdiger Dillmann<br />

Dekan: Pr<strong>of</strong>. Dr. rer. nat. Karsten Berns<br />

Zeichen der TU im Bibiliotheksverkehr: D 386


Acknowledgments<br />

It would have been a tremendous task indeed to write a doctoral thesis without the help<br />

and support <strong>of</strong> many people around me, to only some <strong>of</strong> whom it is possible to give<br />

particular mention here.<br />

In the first place, I wish to express my gratitude to my supervisor Pr<strong>of</strong>. Dr. Karsten Berns,<br />

who made it possible for me to continue my interests in the control <strong>of</strong> walking machines.<br />

He granted all the freedom and support one could wish for with regard to deciding on and<br />

to following the own research directions. During many sessions at late evening hours, he<br />

spared his time for scientific discussions, with topics ranging from specific problems to<br />

future visions. Special thanks go to my second referee Pr<strong>of</strong>. Dr. Rüdiger Dillmann, who<br />

was willing to take the job <strong>of</strong> reading and judging yet another doctoral thesis despite his<br />

tight schedule.<br />

Furthermore, I would <strong>like</strong> to thank several students who helped me in developing and<br />

implementing some ideas <strong>of</strong> this work: the mechanical design <strong>of</strong> the single leg prototype<br />

was done by Florian Flörchinger, Matthias Roth, and Jan Schumann. Zornitza Chiderova,<br />

Dominik Lamp, and Boris Gilsdorf did the first evaluations on various aspects <strong>of</strong> the<br />

simulation <strong>of</strong> bipedal walking. In particular, I am very grateful to Max Steiner, who<br />

did substantial work on the stable standing control and on preliminary integration <strong>of</strong><br />

machine learning methods. In this context, I wish to thank Pr<strong>of</strong>. Dr. Katja Mombaur and<br />

Gerrit Schultz, then her diploma student, for their cooperation in the bmbf project on<br />

optimization-based generation <strong>of</strong> natural fast and stable motion patterns <strong>of</strong> bio-inspired<br />

walking machines, the results <strong>of</strong> which were included in the prototype development.<br />

My infinite gratitude must be expressed to all the colleagues at the Robotics Research Lab:<br />

Christopher Armbrust, Sebastian Blank, Tim Braun, Tobias Föhst, Carsten Hillenbrand,<br />

Jochen Hirth, Yasir Niaz Khan, Lisa Kiekbusch, Jan Koch, Syed Atif Mehdi, Martin<br />

Proetzsch, MaxReichardt, AlexanderRenner, DanielSchmidtandDanielSchmidt, Norbert<br />

Schmitz, Helge Schäfer, Thomas Wahl, Jens Wettach, and Gregor Zolynski. They created a<br />

truly inspiring environment and enjoyable atmosphere to work at. Regarding the realization<br />

<strong>of</strong> this thesis, special thanks go to Martin for valuable discussions on even the most obscure<br />

facets <strong>of</strong> behavior-based control. To Thomas, for solely taking on the research on bipedal<br />

walking and for his help with the prototype. To Jens and to Tim for their great work on the<br />

simulation environment. And to Carsten, for his priceless help in all questions concerning<br />

electronics and mechanics, and many fruitful and enjoyable discussions regarding all aspects<br />

<strong>of</strong> robot design and development.<br />

I also wish to thank Rita Broschart for her support concerning all kinds <strong>of</strong> administrative<br />

endeavor. Further thanks go to Lothar Gauss; since his joining the group, producing parts<br />

in the workshop became a much easier and less time-consuming task.


ii<br />

Finally, I thank my parents for their support, especially my father for taking up the tedious<br />

task <strong>of</strong> pro<strong>of</strong>-reading this whole thesis and supplying me with countless corrections and<br />

suggestions regarding my flawed technical writing. Last but not least, my thanks and love<br />

go to my wife Birgit. Only her encouragement, understanding, and insistence enabled me<br />

to finish this thesis.


Abstract<br />

Despite several decades <strong>of</strong> research, locomotion <strong>of</strong> bipedal robots is still far from achieving<br />

the graceful motions and the dexterity observed in human walking. Most <strong>of</strong> today’s bipeds<br />

are controlled by analytical approaches based on multibody dynamics, pre-calculated<br />

joint trajectories, and Zero-Moment Point considerations to ensure stability. However,<br />

beside their considerable achievements these methods show several drawbacks <strong>like</strong> strong<br />

model dependency, high energetic and computational costs, and vulnerability to unknown<br />

disturbances. In contrast to this, human locomotion is elegant, highly robust, fast, and<br />

energy efficient. These facts gave rise to the main hypothesis <strong>of</strong> this thesis, namely that a<br />

control system based on insights into human motion control can yield human-<strong>like</strong> walking<br />

capabilities in two-legged robots.<br />

This thesis thus presents a control methodology for bipeds relying heavily on the transfer<br />

<strong>of</strong> concepts found in the locomotion control <strong>of</strong> humans. Based on a thorough review on<br />

biomechanics and neuroscience literature, a control approach is derived that can achieve<br />

dynamic, efficient, and robust walking <strong>of</strong> three-dimensional and fully articulated bipeds.<br />

Being located above the neural level, the control system is structured as a hierarchical<br />

network <strong>of</strong> local feed-forward and feedback units, without using a complete dynamic model<br />

or pre-calculated joint trajectories. Sensor event-based spinal pattern generators coordinate<br />

the stimulation and synchronization <strong>of</strong> control units and the compliance <strong>of</strong> passive joints.<br />

By applying local torque commands instead <strong>of</strong> joint angle control, passive dynamics and<br />

self-stabilizing effects <strong>of</strong> elasticities can be exploited. Postural control is achieved by the<br />

phase-dependent activity <strong>of</strong> several reflexes <strong>of</strong> various complexity.<br />

The suggested approach is tested in a full-featured dynamics simulation framework on an<br />

anthropomorphic biped with 21 degrees <strong>of</strong> freedom and human-<strong>like</strong> morphology, weight,<br />

and actuation. The control system can achieve three-dimensional dynamic walking <strong>of</strong><br />

variable velocity as well as balanced standing. It is able to cope with the high complexity<br />

and the mechanical elasticities <strong>of</strong> the modeled biped. The emerging, naturally looking<br />

walking gait shows remarkable similarities to human walking. Simultaneous control <strong>of</strong><br />

only a subset <strong>of</strong> joints is sufficient to direct the passive dynamics <strong>of</strong> the robot towards a<br />

walking gait. The achieved walking velocity <strong>of</strong> up to 5 km/h can compete with even the<br />

most advanced <strong>of</strong> today’s bipeds. At the same time, energy efficiency is much better than<br />

in joint angle controlled robots. The control system shows considerable robustness against<br />

unknown and unexpected disturbance <strong>like</strong> steps, slopes, or external forces.


Contents<br />

1 Introduction 1<br />

1.1 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3<br />

1.2 Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5<br />

2 Technical <strong>Control</strong> Methods for <strong>Bipedal</strong> <strong>Robots</strong> 7<br />

2.1 Challenges in <strong>Bipedal</strong> <strong>Walking</strong> . . . . . . . . . . . . . . . . . . . . . . . . . 7<br />

2.1.1 <strong>Dynamically</strong> vs. Statically Stable <strong>Walking</strong> . . . . . . . . . . . . . . 7<br />

2.2 Common Methods in <strong>Control</strong>ling Bipeds . . . . . . . . . . . . . . . . . . . 10<br />

2.2.1 Zero-Moment Point . . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

2.2.2 Virtual Model <strong>Control</strong> . . . . . . . . . . . . . . . . . . . . . . . . . 12<br />

2.3 Examples for Technically <strong>Control</strong>led Bipeds . . . . . . . . . . . . . . . . . 13<br />

2.4 Assessment <strong>of</strong> Technical <strong>Control</strong> Approaches . . . . . . . . . . . . . . . . . 20<br />

3 <strong>Human</strong> Locomotion <strong>Control</strong> 23<br />

3.1 Structural Organization <strong>of</strong> Motion <strong>Control</strong> . . . . . . . . . . . . . . . . . . 23<br />

3.1.1 Functional Morphology . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

3.1.2 Neurological Basics . . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

3.1.3 Sensorimotor Interaction . . . . . . . . . . . . . . . . . . . . . . . . 30<br />

3.1.4 Hierarchical Layout <strong>of</strong> Motion <strong>Control</strong> . . . . . . . . . . . . . . . . 32<br />

3.2 Normal <strong>Walking</strong> in <strong>Human</strong>s . . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

3.2.1 Biomechanical Gait Analysis . . . . . . . . . . . . . . . . . . . . . . 36<br />

3.2.2 Phases <strong>of</strong> <strong>Walking</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . 47<br />

3.2.3 Reflex Function during <strong>Walking</strong> . . . . . . . . . . . . . . . . . . . . 48<br />

3.2.4 Postural <strong>Control</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50<br />

3.3 Key Aspects <strong>of</strong> Biological <strong>Walking</strong> <strong>Control</strong> . . . . . . . . . . . . . . . . . . 53<br />

3.4 Biologically Inspired <strong>Control</strong> <strong>of</strong> <strong>Bipedal</strong> <strong>Robots</strong> . . . . . . . . . . . . . . . 55<br />

3.4.1 Exploitation <strong>of</strong> Inherent Dynamics and Elasticities . . . . . . . . . 55<br />

3.4.2 Neuro-, Reflex-, and Oscillator-based <strong>Control</strong> . . . . . . . . . . . . 60<br />

4 A Biologically Supported Concept for <strong>Control</strong>ling Bipeds 67<br />

4.1 Passive <strong>Control</strong> Aspects . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

4.2 <strong>Control</strong> Unit Classes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69<br />

4.2.1 Locomotion Modes . . . . . . . . . . . . . . . . . . . . . . . . . . . 69<br />

4.2.2 Spinal Pattern Generators . . . . . . . . . . . . . . . . . . . . . . . 70<br />

4.2.3 Motion Phases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

4.2.4 Motor Patterns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

4.2.5 Local Reflexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71<br />

4.2.6 Postural Reflexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71<br />

4.3 Hierarchical Layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72


vi Contents<br />

5 <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped 75<br />

5.1 System Premises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

5.1.1 Kinematic Layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

5.1.2 Assumed Actuation and Sensor System . . . . . . . . . . . . . . . . 77<br />

5.1.3 The Behavior-based <strong>Control</strong> Architecture iB2C . . . . . . . . . . . 77<br />

5.2 Guidelines for Designing <strong>Control</strong> Units . . . . . . . . . . . . . . . . . . . . 81<br />

5.2.1 Feed-Forward <strong>Control</strong> Units . . . . . . . . . . . . . . . . . . . . . . 81<br />

5.2.2 Feedback <strong>Control</strong> Units . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

5.2.3 Guidelines for Implementing iB2C Features . . . . . . . . . . . . . . 83<br />

5.3 Stable Standing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83<br />

5.3.1 Balanced Standing <strong>Control</strong> in <strong>Human</strong>s . . . . . . . . . . . . . . . . 83<br />

5.3.2 Stabilized Standing <strong>Control</strong> . . . . . . . . . . . . . . . . . . . . . . 85<br />

5.3.3 Stable Standing State Machine . . . . . . . . . . . . . . . . . . . . 85<br />

5.3.4 Ground Adaptation . . . . . . . . . . . . . . . . . . . . . . . . . . . 86<br />

5.3.5 Posture Stabilization . . . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

5.3.6 Posture Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

5.3.7 Relaxed Stance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89<br />

5.4 Dynamic <strong>Walking</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90<br />

5.4.1 Interrelation between Locomotion Modes . . . . . . . . . . . . . . . 90<br />

5.4.2 <strong>Walking</strong> Initiation . . . . . . . . . . . . . . . . . . . . . . . . . . . 91<br />

5.4.3 Spinal Pattern Generator for <strong>Walking</strong> . . . . . . . . . . . . . . . . . 95<br />

5.4.4 <strong>Walking</strong> Phase 1: Weight Acceptance . . . . . . . . . . . . . . . . . 97<br />

5.4.5 <strong>Walking</strong> Phase 2: Propulsion . . . . . . . . . . . . . . . . . . . . . 99<br />

5.4.6 <strong>Walking</strong> Phase 3: Stabilization . . . . . . . . . . . . . . . . . . . . 101<br />

5.4.7 <strong>Walking</strong> Phase 4: Leg Swing . . . . . . . . . . . . . . . . . . . . . . 102<br />

5.4.8 <strong>Walking</strong> Phase 5: Heel Strike . . . . . . . . . . . . . . . . . . . . . 104<br />

5.4.9 Posture <strong>Control</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105<br />

6 Implementation <strong>of</strong> <strong>Bipedal</strong> <strong>Walking</strong> <strong>Control</strong> 113<br />

6.1 Simulation Environment . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113<br />

6.1.1 Embedding the Physics Engine . . . . . . . . . . . . . . . . . . . . 113<br />

6.1.2 Actuators and Joint <strong>Control</strong> . . . . . . . . . . . . . . . . . . . . . . 115<br />

6.1.3 Simulation <strong>of</strong> Sensors . . . . . . . . . . . . . . . . . . . . . . . . . . 118<br />

6.1.4 Model <strong>of</strong> Simulated Biped . . . . . . . . . . . . . . . . . . . . . . . 119<br />

6.2 Notes on the Implementation . . . . . . . . . . . . . . . . . . . . . . . . . 121<br />

6.2.1 Group Layout and Phase Representatives . . . . . . . . . . . . . . . 121<br />

6.2.2 Application <strong>of</strong> iB2C Features . . . . . . . . . . . . . . . . . . . . . 122<br />

7 Balanced Standing Experiments 125<br />

7.1 Adaptation to the Ground Geometry . . . . . . . . . . . . . . . . . . . . . 125<br />

7.2 External Forces and Platform Movements . . . . . . . . . . . . . . . . . . . 126<br />

7.3 Posture Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133


Contents vii<br />

8 Dynamic <strong>Walking</strong> Experiments 137<br />

8.1 Normal <strong>Walking</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137<br />

8.1.1 Locomotion Modes . . . . . . . . . . . . . . . . . . . . . . . . . . . 137<br />

8.1.2 Switching <strong>Walking</strong> Phases . . . . . . . . . . . . . . . . . . . . . . . 139<br />

8.1.3 Behavioral Activity . . . . . . . . . . . . . . . . . . . . . . . . . . . 142<br />

8.1.4 Kinematic Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 148<br />

8.1.5 Kinetic Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158<br />

8.2 <strong>Walking</strong> under Disturbances . . . . . . . . . . . . . . . . . . . . . . . . . . 168<br />

8.2.1 Sloped Terrain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169<br />

8.2.2 <strong>Walking</strong> over Steps . . . . . . . . . . . . . . . . . . . . . . . . . . . 174<br />

8.2.3 Constant External Forces . . . . . . . . . . . . . . . . . . . . . . . 175<br />

9 Conclusion and Outlook 177<br />

9.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178<br />

9.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182<br />

A Biomechanical and Anatomical Terms 185<br />

B Dynamics Simulation Framework 191<br />

B.1 MCA2 and SimVis3D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191<br />

B.2 Embedding <strong>of</strong> the Physics Engine . . . . . . . . . . . . . . . . . . . . . . . 193<br />

B.3 Interface to <strong>Control</strong> System and Simulation . . . . . . . . . . . . . . . . . 195<br />

B.4 Simulation Scenario for Biped Experiments . . . . . . . . . . . . . . . . . . 198<br />

Bibliography 203


viii Contents


1. Introduction<br />

Since many centuries, man is dreaming <strong>of</strong> creating an artificial servant as assistant in<br />

daily life or as laborer for exhausting work. Before the advent <strong>of</strong> industrial robotics,<br />

these moving machines built <strong>of</strong> inanimate matter have been imagined as <strong>like</strong>nesses <strong>of</strong><br />

man, working in humanly environment. The golem described in the Jewish folklore can<br />

be seen as an early example, a figure <strong>of</strong> human shape being formed from clay. The tale<br />

<strong>of</strong> Rabbi Judah Loew ben Bezalel tells <strong>of</strong> his creation <strong>of</strong> a golem to defend the Jewish<br />

ghetto <strong>of</strong> the 16th century Prague against anti-Semitic attacks. Figure 1.1 shows Paul<br />

Wegener’s interpretation <strong>of</strong> a golem in his silent movie series from 1920. The character<br />

<strong>of</strong> the golem inspired many writers, <strong>like</strong> Johann Wolfgang von Goethe in his poem“The<br />

Sorcerer’s Apprentice”, Mary Shelley and her novel“Frankenstein”, or the Czech writer<br />

Karel ˘ Capek. His science fiction play“R.U.R.”(Rossum’s Universal <strong>Robots</strong>) from 1921<br />

introduced clone-<strong>like</strong> artificial creatures working for and then rebelling against their human<br />

creators. These androids are called <strong>Robots</strong> based on the Czech work robota meaning labor,<br />

thus making the term known and popular.<br />

While the golems, ˘ Capek’s robots, or Frankenstein’s creature are made <strong>of</strong> animated clay<br />

or human tissue, the idea <strong>of</strong> an artificial laborer also infected inventors and engineers. As<br />

early as 1886 Zadock Dederick built and was granted a patent for his Steam Man, a 250kg<br />

heavy machine powered by a steam engine and pulling a rockaway carriage (Figure 1.1).<br />

In 1893, George Moore adopted the idea and built a similar mechanical man, supported<br />

by a horizontal bar and able to walk in circles. At the New York World’s Fair in 1939, the<br />

robot Elektro was exhibited (Figure 1.1). Build by J.M. Barnett <strong>of</strong> the Pittsburgh-based<br />

Westinghouse Electric Corporation as an elaborate marketing tool, the 120kg machine <strong>of</strong><br />

about 2.1m in height could move its mouth, fingers, and limbs. Triggered by simple voice<br />

commands, it would replay recorded speech samples, walk, or smoke cigarettes. Similar<br />

projects <strong>of</strong> that time include the radio controlled robot George by Pilot Officer Sale (1950),<br />

Garco by Harvey Chapman (1953), Gygan by Piero Fiorito (1957), or MM47 by Claus<br />

Scholz (1961).<br />

The walking mechanisms <strong>of</strong> the machines mentioned above were <strong>of</strong> rather primitive form,<br />

the leg movement mostly induced by a single rotating motor and straight-line linkages. To<br />

enable more robust locomotion on uneven ground, more sophisticated constructions became


2 1. Introduction<br />

Figure 1.1: The Golem in Paul Wegener’s silent movies (1920), Steam Man created by Zadock<br />

Dederick (1886), Elektro built by the Pittsburgh-based Westinghouse Electric Corporation (1938),<br />

WL-1 (1967) and WABOT-1 (1973), developed by Ichiro Kato, Waseda University.<br />

necessary. The development <strong>of</strong> walking machines aroused the interest <strong>of</strong> the scientific<br />

community. Ichiro Kato <strong>of</strong> the Waseda University in Japan started his pioneer work on<br />

bipedal robots in 1966. His first robot WL-1 from 1967 is shown in Figure 1.1, WL-3<br />

developed in 1969 could already perform preliminary static walking as well as standing<br />

and sitting motions [Lim 06]. In 1973, Kato’s group built the robot WABOT-1, which is<br />

commonly accepted as the first fully articulated anthropomorphic biped [Kato 73].<br />

With the emergence <strong>of</strong> complex machines <strong>like</strong> those developed by Kato and others, and<br />

the availability <strong>of</strong> microprocessors, it soon became apparent that not only the mechanical<br />

design but also the control <strong>of</strong> bipedal robots poses challenging problems. While Rabbi<br />

Loew could simply write the instructions for his golem on a piece <strong>of</strong> paper and place it<br />

in the golem’s mouth, controlling two-legged locomotion turned out to be much more<br />

difficult. The high center <strong>of</strong> mass, the small support area spanned by only two feet,<br />

and an essentially dynamic gait make preserving balance and stability a tough job. As<br />

robotics scientists mainly emerged from mechanical engineering, the mathematical tools<br />

and concepts <strong>of</strong> this research field where applied to find fitting control strategies, the<br />

most prominent one being Vukobratović’s Zero-Moment Point concept. It found its first<br />

application in 1984 on Kato’s robot WL-10RD and will be described in detail later in<br />

this thesis. Since then, the majority <strong>of</strong> research projects on bipedal robots complied with<br />

these ideas and methods. Mainly located in Japan and other Asian countries, these efforts<br />

yielded impressive machines <strong>like</strong> Honda’s Asimo or Kaist’s Hubo.<br />

But in spite <strong>of</strong> over four decades <strong>of</strong> research, the problem <strong>of</strong> mechanical bipedal locomotion<br />

is still far from being solved. Compared to the elegant and efficient walking motions <strong>of</strong><br />

humans, even the most sophisticated bipedal robots appear clumsy and slow. One possible<br />

way <strong>of</strong> improving the performance <strong>of</strong> two-legged walking machines could be the transfer <strong>of</strong><br />

ideas from biology, not only in respect to the mechanical setup, but also to the control<br />

concepts.<br />

The knowledge necessary for this approach can be found in research on biomechanics<br />

and neurosciences. As with the urge for an artificial servant, man’s aspiration to know<br />

the working <strong>of</strong> his own body has a history <strong>of</strong> many centuries. Aristotle (384–322 BC) is<br />

commonly seen as founder <strong>of</strong> kinesiology, the science <strong>of</strong> movement <strong>of</strong> the body. In his book


1.1. Objectives 3<br />

“De Motu Animalium”he was the first to treat the action <strong>of</strong> muscles and their geometrical<br />

relationship to extremities, and to analyze walking as cyclic motion. The studies <strong>of</strong><br />

Archimedes (287–212 BC) on gravity and leverage laid the foundation to mechanical<br />

analysis <strong>of</strong> motion. The Roman physician Galen (AD 131–201) introduced the principle<br />

<strong>of</strong> antagonistic muscle work or the distinction between muscle and sensory nerves in his<br />

treatise“De Motu Musculorum”, considered to be the first textbook on biomechanics.<br />

It was up to Leonardo da Vinci (1452–1519) to further the field <strong>of</strong> kinesiology during the<br />

Renaissance after many centuries <strong>of</strong> stagnation. The artist and scientist had an interest in<br />

the structure <strong>of</strong> the human body and tried to identify muscles, nerves, and the mechanics<br />

<strong>of</strong> the body during various motions. In 1543, the Flemish physician Andreas Vesalius<br />

published his text“On the Structure <strong>of</strong> the <strong>Human</strong> Body”, correcting some <strong>of</strong> the error<br />

made by Galen. In the late 16th and the following century, important contributions can<br />

be attributed to Galileo Galilee, Marcello Malpighi, and especially to Giovanni Alfonso<br />

Borelli. He was the first to understand that the forces produced by muscles have to be<br />

larger than those acting against the skeletal motion, as the levers <strong>of</strong> the musculoskeletal<br />

system magnify motion rather than force. This was even before Isaac Newton (1642–1727)<br />

wrote his famous work“Principia Mathematica Philosophiae Naturalis”containing the<br />

three laws <strong>of</strong> rest and movement, being essential for the analysis <strong>of</strong> body dynamics. In<br />

the 18th century, electricity replaced the notion <strong>of</strong>“animal spirits”as origin <strong>of</strong> muscle<br />

activation. Luigi Galvani (1737–1798) first stated the presents <strong>of</strong> electrical potentials in<br />

muscles and nerves. In 1748 David Hartley established the term reflex as an automatic<br />

response to a stimulus.<br />

The 19th century saw the beginning <strong>of</strong> modern gait analysis. The Weber brothers<br />

published their book “Die Mechanik der Menschlichen Gehwerkzeuge”, describing the<br />

natural frequency <strong>of</strong> a pendulum-<strong>like</strong> leg swing and establishing the mechanism <strong>of</strong> muscular<br />

action on a scientific basis. The arise <strong>of</strong> photographic techniques applied by Eadweard<br />

Muybridge or Jules Marey allowed more detailed analysis <strong>of</strong> motion. Christian Wilhelm<br />

Braune and Otto Fischer experimentally determined the body’s center <strong>of</strong> gravity. Adolf<br />

Eugen Fick (1829–1901) introduced the terms isometric and isotonic during his analysis <strong>of</strong><br />

the mechanics <strong>of</strong> muscular movement and energetics. Sherrington’s work“The Integrative<br />

Action <strong>of</strong> the Nervous System”is published in 1906. Sixteen years later, Archibald V.<br />

Hill (1886–1977) receives the Nobel Prize for his studies on the oxygen consumption<br />

during muscle work. In the middle <strong>of</strong> the 20th century, analyzing dynamic motions and<br />

scrutinizing the“why”<strong>of</strong> human movement began to dominate research on biomechanics<br />

and neurosciences. This objective has not changed up to now, and still human motion<br />

control is far from being completely understood. Nevertheless, recent results in these fields<br />

encourage the transfer <strong>of</strong> biological concepts to the design and control <strong>of</strong> technical systems.<br />

Combined with the advances made in the development and control <strong>of</strong> complex robotic<br />

systems, this strategy acted as motivation for the work at hand.<br />

1.1 Objectives<br />

The goal <strong>of</strong> this thesis is the derivation <strong>of</strong> a control methodology for dynamical locomotion<br />

<strong>of</strong> bipedal robots based on concepts found in human motion control.<br />

This goal is based on the hypothesis that a control system similar to the human one will<br />

enable a robot to perform locomotion in a <strong>like</strong>wise fashion. Obviously, neither all aspects


4 1. Introduction<br />

<strong>of</strong> human motion control nor all capabilities <strong>of</strong> human locomotion can be expected to be<br />

considered or included. A reasonable subset needs to be regarded that is tailored to be<br />

applicable to bipedal robots with today’s technical means, but still is as general as possible<br />

to allow a broad range <strong>of</strong> motions and skills. Balanced standing and three-dimensional<br />

dynamic walking <strong>of</strong> a fully articulated anthropomorphic biped shall be the target forms <strong>of</strong><br />

locomotion for this work. This allows to analyze the control performance during both more<br />

static and dynamic motions <strong>of</strong> a complex robotic system. Further modes <strong>of</strong> movement <strong>like</strong><br />

running or jumping should potentially be viable.<br />

While the results <strong>of</strong> a single thesis surely cannot fully compete with the abilities <strong>of</strong> machines<br />

developed over decades, it at least should be shown that some <strong>of</strong> their drawbacks can be<br />

overcome. As in the natural example, a walking style <strong>of</strong> less clumsiness, higher speed, or<br />

better efficiency should be expected. In case a comparison to human walking should result<br />

in substantial similarities, e.g. regarding joint angle trajectories, even a step towards the<br />

confirmation <strong>of</strong> biomechanical or neuroscientific hypotheses could be ventured.<br />

To design a biologically inspired control concept in correspondence to the goal just<br />

formulated several issues have to be addressed. The following topics are regarded as being<br />

crucial and thus will be approached in detail within this thesis:<br />

Analysis <strong>of</strong> Recent Results from Biomechanics and Neuroscience<br />

As mentioned on the previous pages, biomechanics and neuroscience have a long lasting<br />

history. Accordingly the state-<strong>of</strong>-the-art on these topics is extensive and still growing.<br />

Nevertheless it needs to be examined to identify the key aspects <strong>of</strong> human motion control<br />

that have the potential to be transferred to a robotic control system. Fortunately, some<br />

researchers <strong>of</strong> these disciplines have already <strong>of</strong>fered some advice for the robotics community<br />

that can serve as starting point for the analysis.<br />

Dependence <strong>of</strong> Mechanics and <strong>Control</strong> in <strong>Human</strong> Motion Generation<br />

It can be assumed that during the course <strong>of</strong> evolution, the human morphology has been<br />

optimized towards bipedalism. Research on functional morphology can help to identify<br />

the properties <strong>of</strong> the musculoskeletal system essential for locomotion. While this work’s<br />

objectives cannot include the development <strong>of</strong> mechanical solutions, still properties <strong>like</strong><br />

muscle characteristics, mass distribution, or limb geometry need to be analyzed as control<br />

aspects are depending on them. Especially potential exploitation <strong>of</strong> passive dynamics<br />

should be considered.<br />

Granularity and Classes <strong>of</strong> Suitable <strong>Control</strong> Units<br />

The selected biological key aspects have to be transferred to a technically feasible control<br />

approach. As this thesis follows the concept <strong>of</strong> behavior-based robot control, a suitable<br />

fusion <strong>of</strong> conceptual ideas needs to be found. This implies that the granularity <strong>of</strong> control<br />

units should be above the level <strong>of</strong> biological neurons. A set <strong>of</strong> classes <strong>of</strong> such control units<br />

should be derived from the functional units found in natural motion control, including<br />

feedback as well as feed-forward mechanisms. To arrange such control units and to manage<br />

their communication, a structural layout needs to be defined serving as a design guideline.


1.2. Structure 5<br />

Implementation <strong>of</strong> Dynamic <strong>Walking</strong> within the <strong>Control</strong> Concept<br />

Having determined relevant classes and a control structure, the control units necessary<br />

to synthesize dynamic walking need to be designed. Functional units identified in the<br />

analysis <strong>of</strong> human walking have to be mapped to the suggested control concept. Solutions<br />

for the implementation <strong>of</strong> the derived control units must be developed. As a bipedal robot<br />

will never be an exact copy <strong>of</strong> the human morphology, adaptation <strong>of</strong> these units as well as<br />

the design <strong>of</strong> new units will be necessary.<br />

Test Environment for Validation<br />

Lacking an actual robot featuring the properties relevant for human-<strong>like</strong> dynamic walking,<br />

a suitable model needs to be developed within a simulation environment. The simulation<br />

has to reflect physical effects <strong>like</strong> gravity, mass distribution, or collision, otherwise dynamic<br />

locomotion and exploitation <strong>of</strong> passive dynamics could not be observed.<br />

Naturally, there already has been and still is research on the transfer <strong>of</strong> control aspects from<br />

biology to walking machines, as will be presented in Chapters 3. Hence this thesis aims at<br />

differing from previous work regarding the extend <strong>of</strong> including biological analysis and the<br />

resulting applicable control aspects, the manner in which these aspects are transformed<br />

into a robot control system, and the complexity <strong>of</strong> the considered robotic target platform.<br />

Only a small minority <strong>of</strong> the research projects on biologically motivated robot control has<br />

been applied to the dynamic walking <strong>of</strong> a fully articulated biped. Finally, as this work<br />

introduces behavior-based control concepts to bipedal robot control, it can be examined if<br />

this harmonizes with the biologically motivated approach or can bring additional benefits<br />

compared to other control methods.<br />

1.2 Structure<br />

This thesis is structured as follows:<br />

Chapter 2 begins by defining the main challenges <strong>of</strong> dynamic bipedal walking that have to<br />

be handled by the control system. Then common control methods <strong>like</strong> the Zero-Moment<br />

Point approach are introduced. These methods are based on techniques from mechanical<br />

engineering or industrial robotics, e.g. multibody dynamics, and will be called technical<br />

control methods in contract to biologically inspired ones for the remainder <strong>of</strong> this work.<br />

The state-<strong>of</strong>-the-art <strong>of</strong> these approaches is reviewed and their advantages and drawbacks<br />

are assessed.<br />

In order to develop a human-<strong>like</strong> biped control system, Chapter 3 introduces basic and<br />

advanced aspects <strong>of</strong> human locomotion control, both regarding its structural organization<br />

and its functional analysis. Based on these results from biomechanical and neuroscientific<br />

research, key aspects <strong>of</strong> biological walking control and implications for robotics are derived.<br />

The chapter concludes with an evaluation <strong>of</strong> literature on biologically inspired control <strong>of</strong><br />

two-legged robot locomotion.<br />

In Chapter 4, the insights just gained are used to design a control concept for dynamic<br />

bipedal locomotion. It is discussed how passive dynamics can be exploited and in what<br />

way they influence the control system. Classes <strong>of</strong> control units are suggested based on<br />

active feed-forward and feedback mechanisms observed in human walking. These control


6 1. Introduction<br />

units are arranged in a hierarchical layout similar to the structure that is assumed to exist<br />

in biological locomotion control.<br />

Chapter 5 describes the design and implementation <strong>of</strong> stable standing and dynamic<br />

walking control applying the suggested methodology. First, requirements regarding the<br />

mechatronics and the control system <strong>of</strong> the target platform are given. Then it is discussed<br />

how suitable control units can be found, e.g. by consulting gait analysis research. Next,<br />

the control units and the resulting network for balanced standing on uneven surfaces and<br />

under external disturbances as well as for dynamic walking are introduced.<br />

The simulation environment and the model <strong>of</strong> the bipedal robot used for the experiments<br />

are presented in Chapter 6. Additionally, it contains notes on particularities <strong>of</strong> the<br />

implementation regarding the behavior-based architecture and the control framework.<br />

The experiments on stable standing are described in Chapter 7. It is shown how the robot<br />

adapts to the ground geometry and how it reacts to external disturbances. Furthermore,<br />

the results <strong>of</strong> the posture optimization method are presented.<br />

Finally, Chapter 8 analyzes the performance <strong>of</strong> the suggested control system during<br />

dynamic walking. Behavioral activity as well as kinematic and kinetic progresses are<br />

discussed. The resulting motions are compared to data from human gait analysis. The<br />

robustness <strong>of</strong> the control system is evaluated by applying different disturbances during<br />

walking.<br />

The thesis is concluded in Chapter 9. The main aspects and contributions <strong>of</strong> the suggested<br />

control methodology are summarized and an outlook on possible future research is given.<br />

A short glossary on biomechanical terms, some anatomical guidance, and details on the<br />

implementation <strong>of</strong> the simulation framework are given in the appendices.


2. Technical <strong>Control</strong> Methods for<br />

<strong>Bipedal</strong> <strong>Robots</strong><br />

Looking at control methods for two-legged robots two fundamentally different approaches<br />

can be observed: the technical and the biological approach. In this thesis technical control<br />

approaches are understood as the class <strong>of</strong> methods mainly based on insights from industrial<br />

robotics and mechanical engineering. They draw on sound mathematical concepts <strong>like</strong><br />

multibody dynamics, linear and nonlinear control theory, or established joint constructions<br />

and materials from industrial robots. Biologically inspired approaches try to transfer<br />

results from human motion analysis, biomechanics, or neuroscientific research to technical<br />

system. While not totally dismissing approved mathematical and engineer’s resources,<br />

they focus more on new materials and actuation systems, intelligent mechanics, and, most<br />

notably, on adopting control concepts found in nature.<br />

This chapter will review the technical control approaches as they are most commonly used<br />

for bipeds. Then it will be argued that it is well worth looking at how nature is creating<br />

and controlling bipedal locomotion, as done in the subsequent chapter. But to start with,<br />

the main challenges in controlling two-legged locomotion will be introduced. These apply<br />

to the technical as well as the biological control methods.<br />

2.1 Challenges in <strong>Bipedal</strong> <strong>Walking</strong><br />

The task <strong>of</strong> getting a bipedal robot to walk can be divided into several separate problems.<br />

For each <strong>of</strong> these challenges a solution must be found to achieve stable, cyclic locomotion.<br />

This section will shortly introduce this set <strong>of</strong> problems to enhance the understanding on<br />

the subject matter <strong>of</strong> this thesis. However, first there must be distinguished between two<br />

fundamentally different ways <strong>of</strong> walking.<br />

2.1.1 <strong>Dynamically</strong> vs. Statically Stable <strong>Walking</strong><br />

<strong>Walking</strong> can be performed in two entirely different ways: statically or dynamically stable.<br />

This holds true not only for two-legged walking, but also for other multipeds <strong>like</strong> four-legged


8 2. Technical <strong>Control</strong> Methods for <strong>Bipedal</strong> <strong>Robots</strong><br />

support<br />

CoM<br />

support<br />

Figure 2.1: Statically stable walking keeps the projection <strong>of</strong> the center <strong>of</strong> mass inside the<br />

support area, dynamic walking allows for temporarily instability.<br />

mammals. In statically stable walking it is possible to freeze the motion at any time<br />

without risking instability. This can only be achieved by always keeping the projection <strong>of</strong><br />

the center <strong>of</strong> mass within the support area <strong>of</strong> the feet with ground contact. For bipedal<br />

walking this creates the necessity to first fully move the weight <strong>of</strong> the body over the foot<br />

<strong>of</strong> the next stance leg before the swing foot can leave the ground. As a result statically<br />

stable walking, while guaranteeing stability, is comparatively slow and tedious, consumes<br />

more energy, and limits the possible step length.<br />

In contrast, dynamically stable walking, or dynamic walking in short, does not pose<br />

limitations that hard on the trajectory <strong>of</strong> the center <strong>of</strong> mass. Rather, it must be ensured<br />

thatinstabilityremainstemporarily. Toputitsimply, dynamicwalkingmeanstoconstantly<br />

fall, but to bring forward the swing leg in time to prevent tilting over. This far more<br />

efficient kind <strong>of</strong> walking allows for higher velocity by larger step size, shorter double<br />

support phase, and less time consuming adjustment <strong>of</strong> weight distribution. Figure 2.1<br />

illustrates the difference by showing projection <strong>of</strong> the center <strong>of</strong> mass during a statically<br />

and dynamically stable step.<br />

Nearly all walking found in two- or four-legged vertebrates is dynamic. If at all, statically<br />

stable walking can be observed in very slow or cautious locomotion, or in case highly<br />

deliberate foot placement is necessary. In bipedal robots, both ways <strong>of</strong> walking can and<br />

have been realized with dynamic walking being the considerably more challenging control<br />

problem. Possible approaches to achieve this behavior will be discussed later in this chapter<br />

and at the end <strong>of</strong> the next one.<br />

As this thesis focuses on dynamic walking, from now on this way <strong>of</strong> locomotion is implied,<br />

and the distinction is explicitly made when talking about the statically stable case. The<br />

challenges <strong>of</strong> walking described in the following also deal with dynamic locomotion, even if<br />

several <strong>of</strong> these problems must also be tackled for the statically stable alternative.<br />

Transition between Stance and Swing<br />

<strong>Walking</strong> is a process <strong>of</strong> discrete phases and transitions: each leg switches its role from<br />

stance to swing, and the biped alternates between single and double support. The challenge<br />

here is to find the proper point in time when to switch from double to single support and<br />

CoM


2.1. Challenges in <strong>Bipedal</strong> <strong>Walking</strong> 9<br />

Figure 2.2: A typical step during normal dynamic walking <strong>of</strong> a human subject from one heel<br />

strike to the next.<br />

back again. If the rear leg is lifted too early, energy to transfer the body over the stance<br />

leg might be insufficient. Lifted too late, the time to swing the leg forward to support<br />

the body will be too short. Also, the objectives <strong>of</strong> the control units involved may change<br />

depending on the walking phase. Switching from one phase to the next can be triggered by<br />

sensory events or after a certain amount <strong>of</strong> time. Figure 2.2 illustrates the phases during<br />

normal dynamic walking by showing a typical step <strong>of</strong> a human subject from one heel strike<br />

to the next. The swing phase lasts for about one third <strong>of</strong> the whole stride’s time. The key<br />

events normally referenced in bipedal robotics as well as in biomechanical gait analysis are<br />

foot strike, midsupport, toe-<strong>of</strong>f, forward swing, and deceleration.<br />

Support <strong>of</strong> the Trunk<br />

During its stance phase, a leg must support the weight <strong>of</strong> the upper body. The trunk<br />

height is a result <strong>of</strong> the stance leg’s length, which can be kept straight or act as virtual<br />

spring. If the leg is kept stiff, more potential energy must be used to lift the upper body<br />

as it moves forward <strong>like</strong> an inverted pendulum, but the torque in the knee can remain low.<br />

In case the knee is bended, the leg length must be synchronized with the translation <strong>of</strong><br />

the trunk.<br />

Leg Swing<br />

The forward swing <strong>of</strong> the leg has to meet two conditions: the foot must clear the ground,<br />

and the leg must arrive in front <strong>of</strong> the body in time to act as next swing leg. Ground<br />

clearance can be achieved through shortening the leg by bending the knee, and by keeping<br />

the foot oriented level to the ground. Uneven terrain can make this task even more difficult.<br />

The swing duration can be shortened by accelerating in the hip joint, but the maximum<br />

hip torque confines the minimal duration and as such the maximum walking velocity given<br />

a maximum step length.<br />

<strong>Control</strong> <strong>of</strong> Forward Velocity<br />

The appropriate forward velocity <strong>of</strong> the upper body is crucial for stable dynamic walking.<br />

As the system acts <strong>like</strong> an inverted pendulum while traversing over the stance leg, enough<br />

kinetic energy must be present to be transferred to potential energy <strong>of</strong> the upper body.<br />

In contrast, if the velocity is too high, the time for the swing leg to travel before the<br />

body might be too short and the biped tumbles. During walking, energy is consumed by<br />

damping and by foot impact. Measuring or estimating the proper forward velocity is not<br />

easily achieved.


10 2. Technical <strong>Control</strong> Methods for <strong>Bipedal</strong> <strong>Robots</strong><br />

The forward velocity can be influenced by many factors: the further back the center <strong>of</strong><br />

pressure is located in the stance foot during single support, the faster is the resulting<br />

forward movement. Bending the trunk forward accelerates the biped, leaning backwards<br />

slows it down. During double support, the center <strong>of</strong> mass can be transferred forward<br />

or backwards within certain limitations. A push-<strong>of</strong>f motion using the ankle <strong>of</strong> the rear<br />

leg before lifting it inserts energy into the system, too. Finally, the timing <strong>of</strong> the phase<br />

transition events as well as the variation <strong>of</strong> the step length change the amount <strong>of</strong> time the<br />

body mass is in front <strong>of</strong> and behind the stance leg, and thus is accelerating or decelerating<br />

the robot.<br />

Stability <strong>of</strong> the Trunk in the Sagittal Plane<br />

Keeping the upper body erect or purposefully leaning it forward or backwards as just<br />

mentioned is a further objective during walking. Trunk pitch is disturbed by inertia during<br />

foot impact as well as by gravity as its center <strong>of</strong> mass lies above the hip joint. It is further<br />

influenced by hip torques generated during leg swing.<br />

Lateral Stability<br />

While the challenges mentioned so far also hold true for planar walking, i.e. walking<br />

confined to the sagittal plane, lateral movement is only introduced in three-dimensional<br />

locomotion. Foot impact can heavily disturb lateral stability. Keeping the biped from<br />

falling to one side can be reached by leaning the upper body sidewards, by applying<br />

torque in the ankle joints, or by shifting the foot position <strong>of</strong> the swing leg to the left or<br />

right. While the first two methods only allow marginal influence, the last one can even<br />

compensate strong sidewards motions, however it must potentially act over several strides.<br />

Introducing feedback control for both lateral and anterior/posterior stability requires an<br />

estimation <strong>of</strong> the upper body’s orientation. Given an unknown ground orientation, this<br />

cannot be calculated by kinematics. Sensor systems <strong>like</strong> inertial measurement units or<br />

vision systems gauging the horizon can provide the necessary information.<br />

2.2 Common Methods in <strong>Control</strong>ling Bipeds<br />

Nature was not the first choice when biped engineers first looked for inspiration on how<br />

to control two-legged walking machines. Rather, established methods from electrical and<br />

mechanical engineering <strong>like</strong> multibody dynamics or classical control theory served as basis<br />

for the development <strong>of</strong> control systems. This section will review the most prominent and<br />

more recent developments <strong>of</strong> biped research to illustrate the technical control approach.<br />

Most <strong>of</strong> the control systems for walking bipedal robots are based on Vukobratović’s Zero-<br />

Moment Point (zmp) approach [Vukobratovic 72, Vukobratovic 04] or the derived Center<br />

<strong>of</strong> Pressure (cop) approach [Sardain 04]. Regarding the prominent role <strong>of</strong> the zmp, the<br />

next section will give a short overview on its ideas.<br />

2.2.1 Zero-Moment Point<br />

The zmp describes a criterion for dynamic balance first formulated by Miomir Vukobratović<br />

in 1968. It is defined as that point on the ground at which the net moment <strong>of</strong> the inertial<br />

forces and the gravity forces has no component along the horizontal axes.


2.2. Common Methods in <strong>Control</strong>ling Bipeds 11<br />

Figure 2.3: Visualization <strong>of</strong> the Zero-Moment Point. From [Vukobratovic 04], p160.<br />

Let us consider a biped robot in single support phase. The influence <strong>of</strong> the dynamics <strong>of</strong> the<br />

segments above the ankle is represented by the force FA and the moment MA (Figure 2.3b).<br />

The ground reaction in point P consists <strong>of</strong> the force R = (RX,RY,RZ) and the moment<br />

M = (MX,MY,MZ). The friction <strong>of</strong> the non-sliding foot on the ground compensates the<br />

horizontal components <strong>of</strong> FA and the vertical part <strong>of</strong> MA (Figure 2.3c) and can therefore<br />

be represented by (RX,RY,MZ). RZ represents the ground reaction that balances vertical<br />

forces. The remaining horizontal components <strong>of</strong> active moments can only be compensated<br />

by shifting the position P <strong>of</strong> the reaction force R within the support polygon (Figure 2.3d).<br />

If the support polygon <strong>of</strong> the foot is not large enough for an appropriate position <strong>of</strong> R, the<br />

force acts at the edge <strong>of</strong> foot and a uncompensated component <strong>of</strong> the reaction moment<br />

remains. This results in a rotation about the foot edge and the robot stumbles. Therefore<br />

the condition for the robot to be in dynamic equilibrium can be given for the point P on<br />

the foot sole where the ground reaction force is acting:<br />

MX = 0 and MY = 0 (2.1)<br />

The point P is then called the Zero-Moment Point. The static equilibrium equations for<br />

the supporting foot can be given as<br />

R+FA +msg = 0, (2.2)<br />

−→<br />

OP ×R+ −→<br />

OG×msg +MA +MZ + −→<br />

OA×FA = 0, (2.3)<br />

where O is the origin <strong>of</strong> the coordinate system, P the ground reaction force acting point,<br />

G the foot’s center <strong>of</strong> mass, A the ankle joint, and ms the foot mass.<br />

Projecting Equation 2.3 onto the horizontal plane gives<br />

( −→<br />

OP ×R) H<br />

+ −→<br />

OG×msg +M H A +MZ +( −→<br />

OA×FA) H<br />

= 0 (2.4)<br />

which is the basis for computing the ground reaction force acting point P. To ensure<br />

dynamic equilibrium, the point P must be within the support polygon. If the computed


12 2. Technical <strong>Control</strong> Methods for <strong>Bipedal</strong> <strong>Robots</strong><br />

Figure 2.4: The ground reaction force acting point P must be within the support polygon.<br />

From [Vukobratovic 04], p164.<br />

point is outside the support polygon (called fictitious, fzmp then), the zmp does not exist<br />

as Equation 2.1 is not met and the ground reaction force acting point P is actually on<br />

edge <strong>of</strong> support polygon. This results in a rotation around the foot edge and the loss <strong>of</strong><br />

balance (Figure 2.4).<br />

The zmp can be used for the task <strong>of</strong> <strong>of</strong>fline gait synthesis or as a key indicator for online<br />

gait control. It can be measured approximately by force sensors in the foot sole <strong>of</strong> the robot.<br />

An extension to the zmp and cop can be found in the foot-rotation-indicator [Goswami 99].<br />

It is identical to the cop when located within the foot’s support area, but when outside<br />

it gives information about the degree and direction <strong>of</strong> postural instability. Most <strong>of</strong> the<br />

bipeds presented in the following are using zmp calculation as part <strong>of</strong> their control system.<br />

2.2.2 Virtual Model <strong>Control</strong><br />

Another methodology for the control <strong>of</strong> bipedal robots is the Virtual Model <strong>Control</strong><br />

developed by Jerry Pratt [Pratt 95b, Pratt 01]. The goal <strong>of</strong> this approach is to keep the<br />

control algorithms easy to understand and intuitive.<br />

The idea <strong>of</strong> Virtual Model <strong>Control</strong> is to apply forces to the robot via virtual components<br />

that are attached within the robot or between the robot and the environment. The<br />

resulting actual joint torques and forces create the same effect that the force <strong>of</strong> the virtual<br />

component would create. These components, which have to create a force based on their<br />

state, might include springs, dampers, masses, potential field, or others. The placement <strong>of</strong><br />

the components remains with the designer and requires physical intuition. No complete<br />

dynamic model <strong>of</strong> the robot is required. The equation describing the static dynamics for a<br />

serial link chain is given as<br />

τ = J T F (2.5)<br />

where τ are the joint torques, J is the Jacobian relating the two attached frames <strong>of</strong> the<br />

virtual components, and F is the force produced by it.<br />

<strong>Robots</strong> controlled by Virtual Model <strong>Control</strong> include a simulated six legged walking<br />

machine [Torres 96]. The 18 degree <strong>of</strong> freedom robot can walk in any direction, turn, and<br />

balance an inverted pendulum on its back. The bipedal robot Spring Turkey (Figure 2.5a)<br />

can walk in the sagittal plane [Pratt 97]. Each leg has an active degree <strong>of</strong> freedom in its<br />

knee and hip, but not feet or ankle. The robot is supported and balanced by a virtual<br />

“granny walker”composed <strong>of</strong> two springs and dampers (Figure 2.5b). <strong>Control</strong> <strong>of</strong> forward<br />

speed during double support is realized by a virtual“dog-track bunny”connected to the


2.3. Examples for Technically <strong>Control</strong>led Bipeds 13<br />

(a) (b)<br />

Figure 2.5: (a) The bipedal robot Spring Turkey. (b) Virtual components are attached to the<br />

robot. From [Pratt 97], p197.<br />

robot by a damper. The coordination <strong>of</strong> the legs is realized by a virtual“reciprocating gait<br />

orthosis”. A state machine selects the components during the different phases <strong>of</strong> walking.<br />

Spring Turkey can walk in a not so robust manner at a speed <strong>of</strong> up to 0.5m. The robot s<br />

Spring Flamingo is also partly controlled by using the Virtual Model <strong>Control</strong> and discussed<br />

in Section 3.4.1.<br />

2.3 Examples for Technically <strong>Control</strong>led Bipeds<br />

Since four decades, reseach institutes throughout the world have been developing bipedal<br />

robots. Despite their anthropomorphic appearance, most <strong>of</strong> the efforts follow a more industrial<br />

approach in the design and control <strong>of</strong> their machines and apply the aforementioned<br />

zmp calculation for generating joint trajectories. The most prominent representatives <strong>of</strong><br />

this kind <strong>of</strong> robots are described in the following section, two <strong>of</strong> them in more detail.<br />

The H7 Robot by the JSK Laboratory<br />

The Jouhou System Kougaku (jsk) Laboratory <strong>of</strong> the University <strong>of</strong> Tokyo has a long<br />

tradition <strong>of</strong> building humanoid robots, some <strong>of</strong> which are shown in Figure 2.7. The aim<br />

<strong>of</strong> its work is to develop an experimental research platform for walking, autonomous<br />

behavior and human interaction. The design <strong>of</strong> their latest robot H7 focused on additional<br />

degrees <strong>of</strong> freedom (resulting in 30), extra joint torques, high computing power, real-time<br />

support, power autonomy, dynamic walking trajectory generation, full body motions,<br />

and three-dimensional vision support. Being 1.5m tall and weighting 57kg, the robot<br />

features 7 degrees <strong>of</strong> freedom per leg including an active toe joint. A real-time capable<br />

on-board computer, four lead-acid batteries, wireless lan, two ieee1394 high resolution<br />

cameras, 6-axis forces sensors and an inertial measurement unit complete the robot’s<br />

equipment [Kuffner 01, Chestnutt 03, Nishiwaki 06].<br />

The online walking control system <strong>of</strong> H7 allows to generate walking trajectories satisfying<br />

a given robot translation and rotation as well as an arbitrary upper body posture. It<br />

is composed <strong>of</strong> several hierarchical layers as shown in Figure 2.6. Each layer represents<br />

a different control cycle and passes its processed results to the next, lower layer which<br />

usually runs at a higher frequency.<br />

The gait decision layer chooses the gait and calculates the footstep locations. The algorithm<br />

proposed by the authors determines the next swing leg’s foot point relative to the foot


14 2. Technical <strong>Control</strong> Methods for <strong>Bipedal</strong> <strong>Robots</strong><br />

Figure 2.6: Hierarchical layers <strong>of</strong> the dynamic walking control system <strong>of</strong> the robot H7.<br />

From [Nishiwaki 06], p83.<br />

<strong>of</strong> the supporting leg. Given a desired torso motion per step <strong>of</strong> (x,y,θ) with x being the<br />

forward, y the sideward and θ the rotational <strong>of</strong>fset, the next foot point is set at (x,2y±w,θ)<br />

from the support foot, with w being the normal foot distance. Furthermore, geometrical<br />

collision conditions are considered.<br />

Generating dynamically stable walking trajectories in the next layer is based on Zero-<br />

Moment Point considerations. Rather than analytically deriving the zmp trajectory from<br />

the robots motion, the authors suggest a method with the walking trajectories following a<br />

given zmp trajectory by horizontally shifting the torso. The starting point is an initial<br />

robot trajectory<br />

A(t) = (...,xi(t),yi(t),zi(t),θi(t),...). (2.6)<br />

describing the position and orientation for each <strong>of</strong> the robot’s segments. Given the gravity<br />

g and the position (xi,yi,zi), the mass mi, the inertia tensor Ii, and the angular velocity<br />

ωi <strong>of</strong> each robot link i, the x component (y similar) <strong>of</strong> the zmp P = (xp,yp,0) T can be<br />

calculated as �<br />

mizi¨xi −<br />

xp =<br />

� {mi(¨zi +g)xi +(0,1,0) T Ii, ˙ωi}<br />

− � (2.7)<br />

m1(¨zi +g)<br />

resulting in a zmp trajectory PA(t) = (xpa(t),ypa(t),0) T for the robot trajectory A(t). A<br />

(t),0) can be followed by finding horizontal and<br />

desired zmp trajectory P∗ A (t) = (x∗pa (t),y∗ pa<br />

vertical shifts x ′ i(t) and y ′ i(t) for each link <strong>of</strong> the robot. For this, the following equation<br />

must hold true, where xe p = x∗ p −xpa,x e i = x ′ i −xi<br />

x e p =<br />

� mizi¨x e i − � mi(¨zi +g)x e i<br />

− � m1(¨zi +g)<br />

. (2.8)<br />

If a uniform horizontal body shift is assumed, i.e. the whole upper body moves consistently<br />

(x e i = x e ), equation 2.8 can be transformed to<br />

x e p =<br />

� mizi<br />

− � m1(¨zi +g) +xe . (2.9)


2.3. Examples for Technically <strong>Control</strong>led Bipeds 15<br />

Figure 2.7: <strong>Human</strong>oids developed by the jsk Lab <strong>of</strong> the University <strong>of</strong> Tokyo: H5, H6, and H7.<br />

After discretizing time and assuming certain boundary conditions, xe (i) can be obtained<br />

from equation 2.9 in trinomial form, resulting in a robot trajectory following the desired<br />

zmp trajectory P∗ A (t). This method requires that mass, inertia and pose are known for all<br />

robot links.<br />

A walking trajectory is then calculated online for each footstep, but already for three<br />

steps in advance. Although normally only the first <strong>of</strong> these steps is used, the next two<br />

pre-calculated steps would bring the robot to a stop, thus improving the systems safety.<br />

Furthermore it is checked for each generated trajectory whether it remains within joint<br />

angle and velocity limits, and whether it is free <strong>of</strong> self-collision.<br />

Inthelastlayerbeforethemotorservolayer, thegeneratedwalkingtrajectoriesaremodified<br />

based on sensor feedback. This step aims at compensating disturbances, modelling errors or<br />

changes in the environment. Three controllers are used to achieve this: the horizontal torso<br />

position is shifted according to the difference between the desired and the measured zmp;<br />

the roll deflection <strong>of</strong> the upper body is compensated based on the inertial measurements;<br />

and the servo gain is lowered before foot contact to lower the impact <strong>of</strong> ground reaction<br />

forces.<br />

Higher levels <strong>of</strong> H7’s control system deal with foot point planing in environments containing<br />

obstacles, with object manipulation, or with full-body motion planing. Furthermore,<br />

autonomous behavior is improved using the visual system by object tracking or 3D<br />

reconstruction <strong>of</strong> the environment.<br />

The <strong>Human</strong>oid Robotics Project (HRP) by AIST<br />

The National Institute <strong>of</strong> Advanced Industrial Science and Technology (aist) in Tsukuba,<br />

Japan and Kawada Industries are developing the HRP robots since 1998 supported by<br />

the <strong>Human</strong>oid Robotics Project <strong>of</strong> the Ministry <strong>of</strong> Economy, Trade and Industry (meti).<br />

Figure 2.8 shows the latest models <strong>of</strong> this research project. The robot HRP-2 features 35<br />

degrees <strong>of</strong> freedom and is able to walk, lay down and stand up again [Kaneko 04]. Again,<br />

the locomotion control system is based on zmp calculation [Kaneko 02b, Kajita 03].<br />

The approach used a simplified dynamic model <strong>of</strong> the robot to enable online reference zmp<br />

trajectory calculation. Assuming an inverted pendulum constrained to a horizontal plane


16 2. Technical <strong>Control</strong> Methods for <strong>Bipedal</strong> <strong>Robots</strong><br />

Figure 2.8: The HRP robot series developed by the National Institute <strong>of</strong> Advanced Industrial<br />

Science and Technology in Tsukuba and Kawada Industries: HRP-2LR, HRP-2P, HRP-2,<br />

HRP-3P, and HRP-3<br />

<strong>of</strong> height zc as replacement model <strong>of</strong> the robot moving in the x/y-plane, the dynamics <strong>of</strong><br />

the system can be described as (similar in y-direction)<br />

¨x = g<br />

zc<br />

+ 1<br />

τy, (2.10)<br />

mzc<br />

where g denotes gravity, x the position <strong>of</strong> the body, m its mass, and τy the torque about<br />

the y-axis. This linear dynamics is called the Three-Dimensional Linear Inverted Pendulum<br />

Mode (3d-lipm). From this relation the position <strong>of</strong> the zmp can be calculated as<br />

px = − τy<br />

. (2.11)<br />

mg<br />

When substituted in equation 2.10 and rewritten as needed for the control <strong>of</strong> the zmp<br />

position , it results in<br />

px = x− zc<br />

¨x.<br />

g<br />

(2.12)<br />

Assuming the robot’s center <strong>of</strong> mass (CoM) position is given as (x,y) in the equations<br />

above, the resulting zmp can easily be calculated. But generating a walking pattern is the<br />

inverse problem: given a reference zmp trajectory, the movement <strong>of</strong> the center <strong>of</strong> mass is to<br />

be found. The authors propose to formulate the zmp control as a servo problem. Defining<br />

the time derivative ux <strong>of</strong> the horizontal acceleration ¨x as output <strong>of</strong> a servo controller<br />

working on the zmp error pref −p and as input <strong>of</strong> a reformulation <strong>of</strong> equation 2.12, the<br />

system shown in Figure 2.9 can calculate the CoM trajectory.<br />

But in this way the CoM is only moved as soon as the zmp reference value is shifted, but it<br />

should start moving before this to achieve the desired zmp. Thus a control concept known<br />

as“previewcontrol”isused, previouslydescribedbyKatayamaetal.in1985[Katayama 85].<br />

By first discretizing equation 2.12, an optimal preview servo controller is designed resulting<br />

inacontrollaw foru(k) containing anerrortermbasedonthezmp error, a state termbased<br />

on the CoM position, and a preview term based on the future zmp reference trajectory.<br />

Using a preview length <strong>of</strong> 1.6s, good matching <strong>of</strong> the zmp trajectory can be achieved.<br />

If the multibody model <strong>of</strong> HRP-2P is used instead <strong>of</strong> the simplified one, tracking errors <strong>of</strong><br />

the reference zmp can be observed. In order to compensate this, the authors suggest using


2.3. Examples for Technically <strong>Control</strong>led Bipeds 17<br />

Figure 2.9: Generating a CoM trajectory by tracking the zmp. From [Kajita 03], p1622.<br />

a second preview control using the tracking error from the multibody model as input. In<br />

doing so, the maximum zmp error can be reduced sufficiently. Using walking patterns<br />

generated by this approach, spiral stair climbing <strong>of</strong> the simulated HRP-2P could be shown.<br />

Several further improvements to this walking algorithm have been proposed. By adding<br />

an additional zmp control based on sensor feedback and virtual time shift <strong>of</strong> the reference<br />

zmp trajectory, walking on uneven terrain can be achieved with the simulated HRP-<br />

2 [Kajita 06]. Further enhancements to the gait can be reached by adding a passive toe<br />

joint [Sellaouti 06]. Higher walking speed and longer steps are made possible by adding an<br />

under-actuated phase during the single support phase before heel strike and by adapting<br />

the reference zmp trajectory accordingly. Recently, additional skills <strong>like</strong> sitting on a chair<br />

or opening and passing through a door have been shown [Arisumi 09, Escande 09].<br />

Besides the fully articulated HRP robots, advanced leg modules without an upper body<br />

have been developed to investigate more sophisticated ways <strong>of</strong> locomotion. Using these<br />

robot (HRP-2L, HRP-2LR and HRP-2LT) running-<strong>like</strong> motion could be achieved using<br />

inverted pendulum based control trajectories [Kaneko 02a, Nagasaki 03, Kajita 05]. The<br />

elasticity necessary for running is simulated by an active control <strong>of</strong> the dc motors instead<br />

<strong>of</strong> exploiting passive, mechanical elasticity. Adding a toe spring increases running speed up<br />

to 3 km/h in simulation [Kajita 07]. Currently, a dust and water pro<strong>of</strong> version <strong>of</strong> the HRP<br />

robot series is being developed [Akachi 05, Kaneko 08]. It can keep a balanced posture<br />

while handling tools <strong>like</strong> an electric screwdriver.<br />

Further Technically <strong>Control</strong>led Bipeds<br />

TheWasedaUniversityisdevelopingbipedalrobotssince1966. Themostrecentinstallment<br />

<strong>of</strong> their research is the Wabian-2R robot (WAseda BIped humANoid) featuring a total<br />

<strong>of</strong> 41 degrees <strong>of</strong> freedom [Ogura 06a]. Based on Zero-Moment Point calculations, a more<br />

human-<strong>like</strong> walking with stretched knees and toe-<strong>of</strong>f motion could be achieved [Ogura 06b].<br />

The pattern generation approach generates trajectories for the feet using inverse kinematics<br />

and Newton-Euler dynamics calculation. The knee trajectory is predetermined using cubic<br />

splines rather then exploiting the natural dynamics <strong>of</strong> the system. Compensatory motions<br />

in the waist maintain the robot’s balance according to the zmp trajectory. An iterative,<br />

genetic algorithm optimizes the calculated trajectories in order to minimize waist rolling,<br />

hip joint pitching, and ankle pitching.<br />

In 1984 the company Honda began to work on humanoid robots. The P2 robot was<br />

introduced in 1996 and can statically walk in any direction as well as climb stairs. Its six<br />

degree <strong>of</strong> freedom legs feature damping to reduce the impact on the joints while walking.<br />

The control system uses a detailed model <strong>of</strong> the robot and its foreknown environment


18 2. Technical <strong>Control</strong> Methods for <strong>Bipedal</strong> <strong>Robots</strong><br />

Figure 2.10: Further technically controlled humanoids: Asimo, KHR-3 (HUBO), Johnnie, and<br />

Wabian-2.<br />

and calculates stable trajectories with the help <strong>of</strong> the zmp method [Hirai 98]. The newest<br />

Honda robot called Asimo is equipped with 28 degrees <strong>of</strong> freedom at the height <strong>of</strong> a<br />

child [Sakagami 02]. While recent research focuses on its cognition skills and interaction<br />

with humans [Mutlu 06], the robot possesses an elaborate locomotion system. As the<br />

zmp-based walking approach is already working rather solid, further work is done on e.g.<br />

footstep planing [Chestnutt 05]. With a specialized model <strong>of</strong> Asimo, running motions <strong>of</strong><br />

up to 10 km/h can be achieved [Takenaka 09b].<br />

TherobotJohnnie hasbeendevelopedattheUniversity<strong>of</strong>Munichstartinginthelate1990s<br />

with the aim to achieve jogging motions [Gienger 00, Lohmeier 04]. It is equipped with 17<br />

joints and a sophisticated sensor setup. Using acceleration and gyroscopic sensors, the<br />

attitude <strong>of</strong> the trunk is calculated including acceleration compensation to balance the robot<br />

throughout the walking cycle. Cartesian trajectories for the center <strong>of</strong> gravity, the rotation<br />

<strong>of</strong> the upper body, and the foot pose are derived as fifth-order polynomials from reduced<br />

dynamic models and zmp positions. The computed torque method used for the modeling<br />

allows to consider the entire system dynamics for the control <strong>of</strong> the robot [Löffler 03].<br />

This model-based control poses high demands on the computational performance, the<br />

communication bandwidth, and the frequency <strong>of</strong> the sensor data. Therefore, the intended<br />

running motion could not be achieved. This goal is hoped to be reached with the successor<br />

Lola by reducing the weight, introducing new sensor systems, and revising the control<br />

concept [Lohmeier 06, Buschmann 09].<br />

Further robots <strong>of</strong> similar design and control approach are the KHR-2 and KHR-3 (aka.<br />

HUBO) by Kaist in Korea [Kim 05, Park 05], BHR-2 <strong>of</strong> the Beijing University <strong>of</strong> Science<br />

and Engineering [Peng 06], the Toyota Partner <strong>Robots</strong> [Soya 06], or the robot developed<br />

by the French inria project bip [Bourgeot 02].<br />

Beside the Zero-Moment Point method there exist other technical control systems using<br />

trajectories calculated on the basis <strong>of</strong> a dynamic model. One example <strong>of</strong> this approach can<br />

be found in the control <strong>of</strong> the robot Rabbit developed as joint project under the French<br />

Institute for Research in Computer Science and <strong>Control</strong> (inria). The project aims at the<br />

investigation <strong>of</strong> different control concepts <strong>of</strong> walking and running, stability <strong>of</strong> posture or<br />

robustness against external forces [Sabourin 05, Morris 06].


Robot Group DoF per Leg Height Weight Literature<br />

Wabian-2R Uni. Waseda, Japan 41 6+1p 1.5 65 [Ogura 06a]<br />

H6 Uni. Tokyo, Japan 35 7 1.3 55 [Kuffner 01]<br />

H7 Uni. Tokyo, Japan 30 7 1.5 57 [Chestnutt 03, Nishiwaki 06]<br />

P2 Honda, Japan 30 6 1.8 210 [Hirai 98]<br />

P3 Honda, Japan 30 6 1.6 118 [Hirai 99]<br />

Asimo Honda, Japan 26 6 1.2 52 [Sakagami 02]<br />

Asimo (research) Honda, Japan 34 6 1.3 54 [Chestnutt 05, Bolder 07, Takenaka 09b]<br />

HRP-2 AIST, Japan 30 6 1.5 58 [Kaneko 04, Arisumi 09, Escande 09]<br />

HRP-2L AIST, Japan 12 6 1.4 58 [Kaneko 02a, Nagasaki 03]<br />

HRP-3 AIST, Japan 42 6 1.6 68 [Akachi 05, Kaneko 08]<br />

SDR-4X (QRIO) Sony, Japan 38 6 0.6 6.5 [Fujita 03]<br />

Partner <strong>Robots</strong> Toyota, Japan 31 6 1.4 40 [Soya 06]<br />

KHR-2 Kaist, Korea 41 6 1.2 56 [Kim 05]<br />

KHR-3 (HUBO) Kaist, Korea 41 6 1.3 55 [Park 05]<br />

BHR-2 Uni. Beijing, China 32 6 1.6 63 [Peng 06]<br />

THBIP-I Uni. Beijing, China 32 6 1.7 130 [Zhao 02]<br />

Johnnie TU Munich, Germany 17 6 1.8 49 [Gienger 00, Löffler 03]<br />

Lola TU Munich, Germany 22 7 1.8 55 [Lohmeier 06, Buschmann 09]<br />

BARt-UH Uni. Hanover, Germany 13 3 1.3 25 [Albert 01]<br />

Lisa Uni. Hanover, Germany 12 6 1.3 38 [H<strong>of</strong>schulte 04]<br />

BIP INRIA, France 15 6 1.8 105 [Bourgeot 02]<br />

Rabbit INRIA, France 4 2 1.4 36 [Sabourin 05, Morris 06]<br />

Rh-1 Uni. Madrid, Spain 21 6 1.4 40 [Arbulu 07]<br />

Spring Turkey MIT, USA 4 2 0.6 10 [Pratt 97]<br />

Table 2.1: Recent examples for bipeds following the technical control approach (height in [m], weight in [kg]).<br />

2.3. Examples for Technically <strong>Control</strong>led Bipeds 19


20 2. Technical <strong>Control</strong> Methods for <strong>Bipedal</strong> <strong>Robots</strong><br />

Table 2.1 summarizes recent developments <strong>of</strong> bipedal walking machines following technical<br />

control approaches <strong>like</strong> the Zero-Moment Point method. The state-<strong>of</strong>-the-art on biped<br />

control concepts being inspired more by biological ideas are covert at the end <strong>of</strong> the next<br />

chapter in section 3.4.<br />

2.4 Assessment <strong>of</strong> Technical <strong>Control</strong> Approaches<br />

Technical control approaches as those mentioned above rely on concepts developed for<br />

industrial robotics and sound mathematical calculations <strong>like</strong> the zmp stability concept<br />

and joint angle control. While these methods have been shown to work for a multitude <strong>of</strong><br />

applications, never-the-less several shortcomings can be observed. The following section<br />

will discuss the advantages (Plus) and drawbacks (Minus) <strong>of</strong> this class <strong>of</strong> control systems<br />

and machines.<br />

Plus: Analytical Approach based on Multibody Dynamics<br />

Technical control approaches describe the process to be controlled as a dynamical system<br />

in an analytical way. The complexity varies from simplified inverted pendulum models to<br />

full multibody dynamics models <strong>of</strong> the robot and its environment. Physical laws build<br />

the foundation <strong>of</strong> control algorithms that are analytically deduced from these models.<br />

Assuming a correct and exhaustive model, the algorithms can derive a stable solution<br />

within the considered situation.<br />

Plus: Mathematically Sound<br />

The calculations needed for these approaches rely on mathematical and mechanical tools<br />

and strategies that have shown their soundness and correctness over the last decades or<br />

centuries. Even methods developed explicitly for biped walking <strong>like</strong> the Zero-Moment Point<br />

calculation are already applied for more that 30 years [Vukobratovic 04]. Considering this<br />

fact and the analytical nature <strong>of</strong> the concepts, the confidence in their feasibility is well<br />

justifiable.<br />

Plus: Exact and Purposeful Motions<br />

As exact joint trajectories are generated and – at least considering the used models –<br />

stability is ensured, highly purposeful motions <strong>of</strong> the controlled robots are possible. This<br />

allows to implement skills going beyond mere walking. Climbing stairs, crouching while<br />

walking, or performing martial arts exercises are just a few examples <strong>of</strong> what has been<br />

shown with state-<strong>of</strong>-the-art humanoids.<br />

Plus: Tried and Tested Mechanics<br />

In nearly all cases, the robots described above can be designed based on approved industrial<br />

mechanics as their control methods do not create novel requirements for e.g. the actuation<br />

system. The use <strong>of</strong> high-end dc motors or gear boxes <strong>like</strong> harmonic drives allows precise<br />

movements. This, in fact, is a mandatory requirement <strong>of</strong> technically controlled machines,<br />

as tracking errors <strong>of</strong> the given joint trajectories would introduce a discrepancy from the<br />

model that could endanger stability. High frequency and high gain controllers as used in<br />

industrial robotics are necessary to guarantee exact reproductions <strong>of</strong> trajectories.


2.4. Assessment <strong>of</strong> Technical <strong>Control</strong> Approaches 21<br />

Minus: Slow and Unnatural Looking Motions<br />

But despite the advantages just mentioned and the long lasting history <strong>of</strong> technically<br />

controlled bipeds, the resulting motions are still far behind those observed in humans or<br />

othertwo-leggedanimals. <strong>Walking</strong>and, whereevenpossible, runningspeediscomparatively<br />

low and motions do not look naturally. One reason for this can be found in the fact<br />

that those machines only do morphological and kinematic imitation <strong>of</strong> human walking.<br />

However, the dimensions and trajectories do not fit to the masses and actuation concepts<br />

<strong>of</strong> technical robots. The walking motions <strong>of</strong> the robot H7 has been compared to that<br />

<strong>of</strong> human subjects regarding CoM trajectory, joint angles, torques and ground reaction<br />

force [Kagami 03]. In nearly all cases the values differ substantially, with humans showing<br />

the more sophisticated and effective trajectories.<br />

Minus: High Energy Costs<br />

Locomotion <strong>of</strong> technically controlled bipeds has been shown to be very energy consuming.<br />

Several facts can be hold responsible for that: the necessity for high precision joint control<br />

as well as the heavy mechanics require high powered actuators. Following joint trajectories<br />

that do not match the actual movement <strong>of</strong> the robot’s segments oppose inertia and produce<br />

high peak torques. No energy is stored in elastic elements. The trajectories are optimized<br />

for stability resulting in unfavorable movements regarding efficiency. For instance the<br />

knee joint is bent much more than in human walking where an efficient semi-flat gait with<br />

lower impact forces and less head movement can be observed. It can be estimated that<br />

technically controlled robots <strong>like</strong> Asimo use at least 10 times the energy (normalized by<br />

weight and height) <strong>of</strong> a typical human during walking [Collins 05]. Similar results can be<br />

seen when comparing the simulated running motion <strong>of</strong> HRP-1 in simulation to human<br />

running [Kajita 02]. Peak torques <strong>of</strong> nearly 2000Nm and actuator power <strong>of</strong> up to 8500W<br />

are required for tracking the pre-calculated trajectory.<br />

Minus: High Computational Demands<br />

Computing joint trajectories based on a detailed dynamic model and stability constrains<br />

<strong>like</strong> zmp is computational expensive. As a consequence, many systems pre-calculate<br />

trajectories <strong>of</strong>fline, resulting in an inflexible control. On-line pattern generation <strong>like</strong> the<br />

one utilized in H7 take up to 5.2 seconds for calculating a three-step trajectory and<br />

still must be modified based on sensor feedback due to lacking accuracy <strong>of</strong> the dynamic<br />

multibody model [Nishiwaki 06]. Also the communication bandwidth can prove as bottle<br />

neck, as these control systems need high frequency sensor information for updating their<br />

models and equally fast control flow for precise position control <strong>of</strong> joint trajectories. This<br />

is the reason for changing the communication bus on LOLA from the slower canbus <strong>of</strong> its<br />

predecessor Johnnie to a sercos-based system [Lohmeier 06].<br />

Minus: Dependent on Dynamic Model<br />

Being based on dynamic models <strong>of</strong> the robot and the environment, technical control<br />

approaches consequently depend on the model’s correctness and level <strong>of</strong> detail. If not<br />

compensated, discrepancies from the model can lead to instability. Errors can occur by<br />

model inaccuracy concerning e.g. mass distribution, slip, gear backlash, or foot impact on<br />

uneven terrain. External forces and disturbances will lead to derivation, too.


22 2. Technical <strong>Control</strong> Methods for <strong>Bipedal</strong> <strong>Robots</strong><br />

Minus: No Exploitation <strong>of</strong> Natural Dynamics<br />

Exploiting the passive dynamics <strong>of</strong> the robot can help in creating natural motions, in<br />

saving energy, and in simplifying control. Technically controlled bipeds normally do<br />

not and cannot exploit the natural dynamics <strong>of</strong> the system as the joints are forced into<br />

fixed trajectories. Furthermore retarding gears will prohibit the natural motions to take<br />

place and make it difficult to store energy in parallel springs or to use the self-stabilizing<br />

properties <strong>of</strong> elastic actuators. It has been shown that walking can be achieved by purely<br />

relying on passive dynamics even without any actuation given cleverly designed mechanics,<br />

proper weight distribution and a small energy feed [McGeer 90]. Details on this principle<br />

<strong>of</strong> passive walking can be found in section 3.4.1.<br />

Minus: Difficult to Add Adaptability<br />

It can turn out to be difficult to add adaptability and to enhance the robustness <strong>of</strong> the<br />

control system in technically controlled bipeds. Very small disturbances can still be<br />

managed if safety stability margins are kept high enough. Otherwise feedback must be<br />

introduced, e.g. by comparing the measured zmp to the reference zmp and then adapting<br />

the model or the trajectories as for example done in the control <strong>of</strong> Rabbit [Wieber 06].<br />

More significant errors can prove to be hard to compensate in these monolithic systems.<br />

The models cannot reflect reality, and due to this fact, stiff mechanics, position based<br />

joint control, or even fully pre-calculated trajectories, adaptation to unknown terrain or<br />

external forces proves to be a tough problem.<br />

Beyond any doubt technical control approaches for bipeds have shown their feasibility and<br />

yielded impressive skills and machines. However, looking at the drawbacks just mentioned<br />

and comparing the results to what nature is capable, the following questions arise: How is<br />

nature managing to do so much better, are there fundamental differences in the control<br />

concepts, and can nature’s solutions be transferred to technical implementations? The<br />

next section is trying to answer these questions by sketching out how human motion<br />

control works, looking in particular at the processes during normal walking. Then selected<br />

literature on biologically motivated robotic approaches is reviewed.


3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

Despite many decades <strong>of</strong> research on two-legged walking machines, the performance <strong>of</strong> the<br />

developed robots is still far behind human capabilities. So it is not surprising that in recent<br />

years robotics researchers tried to find help in the results <strong>of</strong> biomechanics, neurosciences,<br />

human movement analysis, and other biological fields. Therefore, this chapter first presents<br />

selected work related to these topics serving as introduction to the research on human<br />

locomotion control. This will allow to better compare the technical approaches with the<br />

natural one and to show the differences that motivated this work. Finally, literature on<br />

biologically inspired biped control in robotics is reviewed.<br />

Throughout this chapter, terminology <strong>of</strong> anatomy, neuroscience, and biomechanics will<br />

appear. Thetermsoutsidecommonparlanceareexplainedatfirstappearance. Additionally,<br />

the reader may refer to Appendix A including several tableaux illustrating anatomical<br />

planes, directions and joint rotations, a figure <strong>of</strong> the human skeleton marking the locations<br />

<strong>of</strong> major bones, and a diagram and a list <strong>of</strong> skeletal muscles indicating their location,<br />

function, and common abbreviation.<br />

3.1 Structural Organization <strong>of</strong> Motion <strong>Control</strong><br />

When examining human locomotion control, two different aspects can be distinguished: the<br />

structure <strong>of</strong> the control and the actual functional units that are responsible for generating<br />

locomotion within this structure. This section introduces the structural organization <strong>of</strong><br />

motion control, whereas the next section will discuss how dynamic walking is produced.<br />

In the following, structural organization is to be understood as the architecture underlying<br />

the control <strong>of</strong> locomotion. Topics to be discussed include the role <strong>of</strong> the morphology<br />

and the muscle, communication and processing by the nerve pathways and neural cells,<br />

principles <strong>of</strong> perception and creation <strong>of</strong> action, the layout <strong>of</strong> groups <strong>of</strong> nerve cells, or the<br />

integration <strong>of</strong> feedback control.<br />

3.1.1 Functional Morphology<br />

It is <strong>of</strong>ten claimed that human morphology has evolved to be optimal regarding bipedal<br />

locomotion. This introduces the question if bipedalism emerged mainly because <strong>of</strong> ad-


24 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

Figure 3.1: Possible modes <strong>of</strong> locomotion from which bipedalism may have originated.<br />

From [Richmond 02], p72.<br />

vantages gained by two-legged walking. It could also be presumed that other benefits <strong>of</strong><br />

standing on two legs have been more prominent, thus guiding the optimization process in<br />

other directions.<br />

About a dozen hypotheses for the evolution <strong>of</strong> bipedality have been stated over the last<br />

century. Most <strong>of</strong> them are based on the assumed behavior <strong>of</strong> the earliest known hominids<br />

(primates that use a habitual upright bipedal gait) <strong>of</strong> about six million years ago. These<br />

hypotheses differ in which advantage <strong>of</strong> bipedalism is supposed to be decisive. Examples<br />

include greater viewing distance, higher efficiency than ape quadrupedalism, reducing<br />

skin exposure to lower heat stress, tools usage, or arm carrying. K. D. Hunt analyzes<br />

chimpanzee bipedal behavior and the anatomy <strong>of</strong> the Australopithecus africanus, and<br />

argues for a synthesis <strong>of</strong> two scenarios [Hunt 94]. A combination <strong>of</strong> the terrestrial and<br />

arboreal bipedal postural feeding hypotheses could explain the poor bipedal mechanics<br />

and the arboreal competence <strong>of</strong> early hominids. Other reasons stated might have well<br />

played a role in refining the locomotor bipedalism in Homo erectus.<br />

Richmond et al. review literature on possible modes <strong>of</strong> locomotion from which bipedalism<br />

could have originated [Richmond 02]. They evaluate these modes in the light <strong>of</strong> current<br />

evidence from comparative primate anatomy, biomechanics, and fossil hominid anatomy.<br />

Possible candidates are arboreal or terrestrial quadruped ancestors, gibbon-<strong>like</strong> brachiating<br />

ancestors, climbing ancestors, or knuckle-walking ancestors (Figure 3.1). It is stated that<br />

human evolved most <strong>like</strong>ly from an ancestor adapted to knuckle-walking and climbing and<br />

thus reject all hypotheses that are purely based on arboreal ancestors.<br />

Wang et al. argue that the development <strong>of</strong> erect walking and longer legs compared to<br />

the trunk length might partly be motivated by load-carrying. In the period <strong>of</strong> the first<br />

hominids with modern body proportions at around 1.8–1.5 Million years ago, the use <strong>of</strong>


3.1. Structural Organization <strong>of</strong> Motion <strong>Control</strong> 25<br />

stone tools and raw material transport increases. Computer simulations <strong>of</strong> loaded and<br />

unloaded walking show that indeed carrying loads is more effective in erect posture and<br />

thus could be the reason for today’s human proportions beside the advantage <strong>of</strong> persevering<br />

walking.<br />

Early bipedal walking seems to have been much more compliant than walking found in<br />

modern humans. D. Schmitt suggests that the compliant gait found in early hominids<br />

and most <strong>of</strong> today’s apes can be explained by their smaller bodies and poorly stabilized<br />

hindlimbs [Schmitt 03]. These modes <strong>of</strong> locomotion allow to achieve fast walking speeds<br />

with long strides <strong>of</strong> low frequency, low vertical peak forces, and good impact shock damping,<br />

but create significant energy costs. The increase <strong>of</strong> energy efficiency by using a stiff-legged,<br />

inverted pendulum-<strong>like</strong> gait did only take place more recently in early members <strong>of</strong> the<br />

genus Homo with their robust skeleton and joints.<br />

Research<strong>of</strong>Massaadetal.isgoingalongsimilarlines[Massaad 07]. Inwalkingexperiments<br />

with several subjects it is shown that humans are perfectly able to walk more flatly or<br />

bouncy, but still normal gait amounted to a semi-flat form in the course <strong>of</strong> evolution. While<br />

flat walking reduces the mechanical work on the center <strong>of</strong> mass, bouncy walking decreases<br />

the energy the muscles have to supply be relying on pendulum-<strong>like</strong> passive dynamics.<br />

Furthermore muscles seem to work in unfavorable conditions during flat walking. Thus the<br />

semi-flat walking gait <strong>of</strong> modern humans may well present an optimal solution in balancing<br />

efficiency against muscle work.<br />

The geometrical layout <strong>of</strong> the human leg in particular has been examined on its adequacy<br />

for walking and running. As just mentioned, walking with a fully straight leg is not the best<br />

solution, andcomplianceisidentifiedasanecessaryfeature, too. Seyfarthetal.investigated<br />

the stability <strong>of</strong> elastic, three-segment legs [Seyfarth 01]. They demonstrated that instability<br />

in form <strong>of</strong> counter-rotation <strong>of</strong> joints can be reduced in several ways: nonlinear springs,<br />

asymmetric segment lengths, biarticular structures, mechanical constraints <strong>like</strong> heels, or a<br />

bow-<strong>like</strong> mode <strong>of</strong> movement. Except for the last strategy, all <strong>of</strong> these can be found in the<br />

human leg, thus gaining internal self-stabilizing properties. More details on the dynamics<br />

and self-stabilization <strong>of</strong> tri-segmented elastic limbs in mammals are reviewed and discussed<br />

by Fischer and Blickhan [Fischer 06].<br />

The fact that appropriately designed mechanics can significantly reduce the control effort<br />

is discussed by Blickhan et al. [Blickhan 07]. Modelling the quasi-elastic leg operation<br />

as spring-mass model, it is demonstrated that such an“intelligent mechanics”can show<br />

attractive behavior and passively compensate external disturbances without sensing them.<br />

This holds true for running (Figure 3.2) as well as for walking. This passive control only<br />

works if the actuator features muscle-<strong>like</strong> properties such as inherent compliance, adjustable<br />

joint stiffness, and shock absorption, but also damping. Furthermore, for the system to fall<br />

back in an attractive cycle, the touchdown angle <strong>of</strong> the swing leg must be properly chosen,<br />

whereas leg retraction before touchdown can extend the range <strong>of</strong> acceptable angles.<br />

Besides the self-stabilizing properties <strong>of</strong> the elasticity <strong>of</strong> the muscle-tendon system, locomotion<br />

and other movements can also benefit from elastic energy storage. Witte et al.<br />

show that in fact energy is stored in the human locomotor apparatus [Witte 97]. The<br />

amount <strong>of</strong> energy accumulated as recorded in their experiments is considerable and should<br />

not be neglected in the consideration <strong>of</strong> human functional morphology and its role during<br />

locomotion.


26 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

Figure 3.2: Elastic operation <strong>of</strong> the leg can passively stabilize running in the presence <strong>of</strong><br />

external disturbances without changing the angle <strong>of</strong> attack or the stiffness. From [Blickhan 07],<br />

p211.<br />

Günther et al. examine leg geometry and joint axis alignment with respect to the minimization<br />

<strong>of</strong> the sum <strong>of</strong> joint torques as optimization criterion [Günther 04]. By including<br />

sensible geometrical and biomechanical constraints, the extended human leg configuration<br />

indeed proves to be an optimal solution with its short foot and unequal ankle and knee<br />

angles.<br />

In due consideration <strong>of</strong> what was said in the previous paragraphs, Witte et al. suggest to<br />

base the design (and control) <strong>of</strong> bipedal robots on human functional morphology, thereby<br />

moving from purely anthropomorphic to “anthrop<strong>of</strong>unctional” robots [Witte 04]. It is<br />

shown that human gait is energetically close to optimal by using resonance mechanisms:<br />

the arms and the swing leg move <strong>like</strong> an suspended pendulum, the stance leg <strong>like</strong> an<br />

inverted pendulum, and the trunk involves torsional springs. In addition, the proportions<br />

and the mass distribution in the trunk and in the swinging extremities compensate for<br />

undesirable torques [Witte 91]. Witte and his colleagues conclude that bipedal locomotion<br />

relies on the coordinated interaction <strong>of</strong> masses, gravity, elasticity, and muscle work, thus<br />

going beyond pure rigid body dynamics. They also point out that, with humans being<br />

vertebrates, the central role <strong>of</strong> the trunk and spinal cord as origin or at least initiator <strong>of</strong><br />

movement should not be neglected.<br />

Implications for Robotics:<br />

The human locomotion apparatus is highly optimized for efficient bipedal<br />

locomotion. To facilitate walking and to reduce the control effort, robot<br />

design should be based on functional morphology and its motions<br />

should exploit the inherent dynamics <strong>of</strong> the system. Key aspects<br />

are a semi-flat walking gait, proper mass distribution and geometry <strong>of</strong><br />

trunk and extremities, and low-resistant elastic actuators.<br />

3.1.2 Neurological Basics<br />

After the discussion <strong>of</strong> passive control through intelligent mechanics in the previous section,<br />

in the following the active control system <strong>of</strong> human motion control is considered. Unless


3.1. Structural Organization <strong>of</strong> Motion <strong>Control</strong> 27<br />

(a) (b)<br />

Figure 3.3: (a) The central nervous system consists <strong>of</strong> the brain and the spinal cord. (b) The<br />

peripheral nervous system <strong>of</strong> the lower extremities. From [Hamill 03], p102.<br />

noted otherwise, the information in this section is taken from the biomechanics book by<br />

Hamill and Knutzen [Hamill 03].<br />

Coordination and monitoring <strong>of</strong> human movements is an extensive task considering the<br />

number <strong>of</strong> muscles and sensory receptors involved. The nervous system is responsible for<br />

this job by conveying and processing the sensor information and generating an appropriate<br />

stimulation <strong>of</strong> muscle fibers.<br />

The nervous system is composed <strong>of</strong> two parts, the central nervous system and the peripheral<br />

nervous system. The central nervous system consists <strong>of</strong> the brain and the spinal cord,<br />

and can be seen as central processing unit initiating all movement. The remaining nerves<br />

outside the brain and spinal cord make up the peripheral nervous system. It connects the<br />

inner organs and muscles with the spinal cord via the spinal nerves that exit the spinal<br />

cord on the anterior, or ventral, side <strong>of</strong> the vertebral column. The nerves emerging from<br />

the sensory receptors enter the spinal cord on the posterior, or dorsal, side. By this means,<br />

31 pairs <strong>of</strong> spinal nerves enter and exit the spinal cord. For instance, the lower extremities<br />

are connected via the five parts <strong>of</strong> the lumbar and the five parts <strong>of</strong> the sacral region at the<br />

lower end <strong>of</strong> the spine. Figure 3.3 illustrates the central nervous system (a) and the lower<br />

part <strong>of</strong> the peripheral nervous system (b).<br />

The base unit <strong>of</strong> the nerve system is the neuron. Motor neurons, being connected to and<br />

stimulating muscle fibers, consist <strong>of</strong> the cell body, dentrites, and the axon (Figure 3.4).<br />

The cell body, or soma, usually resides within the gray matter <strong>of</strong> the spinal cord or in a<br />

ganglia, a group <strong>of</strong> neurons outside the cord, that can span up to three levels <strong>of</strong> the spinal<br />

cord. Information from other neurons is passed to the neuron via bundles <strong>of</strong> dendrites.


28 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

Figure 3.4: The cell body, dentrites and axon <strong>of</strong> a motor neuron. The axon connects to other<br />

neurons or to motor fibers. From [Hamill 03], p105.<br />

This allows cross-talk <strong>of</strong> neurons from all over the spinal cord and other ganglia. The<br />

axon, a large nerve fiber covered with an insulated shell, forms the output <strong>of</strong> a neuron. It<br />

connects to other neurons or to muscle fibers, where it fans out to form motor endplates<br />

and embeds into fissures near the center <strong>of</strong> the fibers. One single motor neuron connects to<br />

up to 2000 fibers as in the gluteus maximus, or only to 5 or 6 fibers as in the eye muscles.<br />

If sufficiently activated, the motor neuron innervates all its muscle fibers by chemical<br />

transmission and causes them to contract within a few milliseconds. Depending on the<br />

type <strong>of</strong> the motor unit (corresponding to the existing types <strong>of</strong> muscle fibers), this reaction<br />

will elapse slower or faster. The slow-twitch oxidative Type I units have reaction times<br />

greater than 70ms and transport impulses at approximately 80 m/s. Fast-twitch oxidative<br />

Type IIa motor units contract after about 30-50ms. The fast-twitch glycolytic Type IIb<br />

units are even faster with impulse transmission at more than 100 m/s resulting in reaction<br />

times <strong>of</strong> about 30-40ms. Still, viewed from the perspective <strong>of</strong> control theory, these delays<br />

would make feedback control difficult. Type IIb units develop high tension, but fatigue<br />

more quickly than Type IIa or Type I units. Muscles differ in their number <strong>of</strong> associated<br />

motor neurons and the ratio <strong>of</strong> Type I, IIa, and IIb units depending on the nature <strong>of</strong> their<br />

task, e.g. slow, precise movements, or repetitive high forces.<br />

The actual impulse traverses the nerve in form <strong>of</strong> an action potential. The electric potential<br />

<strong>of</strong> the membrane changes from its neutral state at about -70mV by rapid depolarization,<br />

repolarization and hyperpolarization as the impulse travels along the nerve fiber. When<br />

the impulse reaches the muscle fiber at the motor endplates, it generates cross-bridging<br />

and thus shortening within the muscle sarcomere (excitation-contraction coupling). A<br />

single action potential develops a single twitch respond <strong>of</strong> the muscle. But if the impulses


3.1. Structural Organization <strong>of</strong> Motion <strong>Control</strong> 29<br />

Figure 3.5: Interneurons can excite or inhibit other neurons. The Renshaw cell is considered to<br />

play an important role in muscle coordination. From [Hamill 03], p108.<br />

follow continuously and in short progression, the twitches will sum up and form a constant<br />

muscle tension, a tetanus. The tension will decline as the impulses thin down. A muscle<br />

will create the more force the more <strong>of</strong> its motor neurons are firing, i.e. the more neurons<br />

are recruited, and the higher the rate at which they are sending impulses.<br />

Besides being innervated, the action potential in a motor unit can also be inhibited by<br />

connected neurons and interneurons within the spinal cord. Interneurons can be excitatory<br />

or inhibitory, and are interconnecting branches. One type <strong>of</strong> interneuron is the Renshaw cell<br />

which is considered responsible for organizing coordinated muscular responses. Figure 3.5<br />

illustrates a possible inhibition <strong>of</strong> a motor neuron via a Renshaw interneuron.<br />

While the motor neurons are the means by which the central nervous system controls<br />

muscle action, the sensory neurons provide the necessary feedback on the state <strong>of</strong> the<br />

musculoskeletal system, the skin, and other sensory organs <strong>of</strong> the body. After processing<br />

this information, the central nervous system will initiate appropriate reactions.<br />

The so called proprioceptors supply information on the state <strong>of</strong> the musculoskeletal system.<br />

These include the main sensory receptors <strong>of</strong> the muscles, the muscle spindles, and the<br />

Golgi tendon organ. Muscle spindles and their corresponding neurons measure the stretch<br />

<strong>of</strong> a muscle. The spindles are located parallel to the muscle fibers within the belly <strong>of</strong> the<br />

muscle and are enclosed in a capsule, forming a spindle shape, hence the name. Each<br />

capsule contains up to 12 nuclear bags and nuclear chain fibers that are innervated by the<br />

so called gamma motor neurons. The gamma motor neurons are located at ventral exits<br />

<strong>of</strong> the spinal cord close to the alpha motor neurons that stimulate the muscle fibers (see<br />

above). The fibers again are connected to Type I or Type II sensory neurons that send<br />

information to the dorsal horn <strong>of</strong> the spinal cord where their cell body is located. Type I<br />

sensory neurons are more sensitive to muscle stretch whereas Type II sensory neurons only<br />

react when the stretch surpasses a certain threshold. The sensory neurons will fire at a<br />

higher rate if the muscle is elongated more rapidly. If the muscle remains in the stationary<br />

stretched state, Type I neurons will fire at a low, constant rate.<br />

In contrast to the muscle spindle, the Golgi tendon organ measures force or tension in<br />

the muscle. The Golgi tendon organ lies between the muscle fibers and the tendon that<br />

connects to the bone, but in parallel to the tension generated in the passive elements<br />

during stretch. The sensory neuron connected to the Golgi tendon organ create a response<br />

that is proportional to the load <strong>of</strong> the muscle and to the rate <strong>of</strong> change <strong>of</strong> the load.


30 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

Besides the two receptors just mentioned, there exist additional proprioceptors. A number<br />

<strong>of</strong> sensory receptors lie in proximity <strong>of</strong> the joints: the pacinian corpuscle responds to<br />

pressure, the Ruffini endings measure joint position and velocity, and free nerve endings<br />

create a pain sensation. The skin reacts to pressure, vibration, heat, cold, and damage.<br />

The vestibular system located in the inner ear is responsible for the sense <strong>of</strong> balance and<br />

spatial orientation. The cupula which lies within the ampulla arising in the semicircular<br />

canal <strong>of</strong> the vestibular organ, contains hair bundles and acts upon angular acceleration.<br />

The utricle and saccule in the vestibular sac again possess hair cells and are sensitive to<br />

linear acceleration, including gravity. The vision system also contributes to the sense <strong>of</strong><br />

balance by estimating the horizon and the optical flow. Contradiction <strong>of</strong> these two sources<br />

<strong>of</strong> information can cause nausea, e.g. when traveling in a swaying ship without the eyes<br />

perceiving this movement.<br />

Implications for Robotics:<br />

Biological control is organized as a network <strong>of</strong> basic units (neurons) and<br />

communication channels (nerve fibers). While the system is concentrated<br />

at the brain and spinal cord, it is still highly distributed and spatially<br />

related, e.g.motorneuronsforthefeetarelocatesatthelowerend<strong>of</strong>the<br />

spinal cord. The basic processing units can be gradually stimulated<br />

or inhibited. A distributed sensor system provides feedback on joint<br />

position, velocity, and torque, as well as a sense <strong>of</strong> balance. Motor control<br />

allows precise position control, adjustable elasticity, or pure torque<br />

control. Compared to technical systems, processing and communication<br />

in biological systems is relatively slow.<br />

3.1.3 Sensorimotor Interaction<br />

After having discussed how sensory information is transferred to the spinal cord and how<br />

motor action can be initiated from there, the question remains on how sensor input is<br />

processed to create an appropriate reaction. This section will introduce the foundation <strong>of</strong><br />

sensorimotor interaction in form <strong>of</strong> reflexes.<br />

The term reflex as“an automatic or involuntary response to a stimulus” 1 was first used in<br />

1748 by David Hartley. Its common scientific usage was established in Sherrington’s study<br />

<strong>of</strong> the nervous system from 1906 [Sherrington 06]. But while the concept <strong>of</strong> reflexes being<br />

a stereotyped response to nervous stimuli has persisted for long despite contradictionary<br />

evidence, nowadays it is agreed upon being false. Rather, reflex actions have been shown<br />

to be heavily affected by modulation and to exist in a wide range <strong>of</strong> complexity. Motor<br />

actions that conform the most with the former notion <strong>of</strong> a reflex are those“mediated by<br />

relatively straightforward circuitry in the spinal cord”[Nicholls 92]. But despite being<br />

in the focus <strong>of</strong> research for centuries, the function <strong>of</strong> reflexes especially during human<br />

locomotion is only slowly becoming clearer [Zehr 99].<br />

The most basic form <strong>of</strong> reflex arcs are monosynaptic reflexes. Sensory information is<br />

entering the spinal cord at the same level as the motor stimulation is leaving it [Hamill 03].<br />

On example for such a reflex is the stretch or myotatic reflex that creates a short muscle<br />

1 New Webster’s Dictionary and Thesaurus <strong>of</strong> the English Language


3.1. Structural Organization <strong>of</strong> Motion <strong>Control</strong> 31<br />

Figure 3.6: The flexor reflex as an example for a monosynaptic reflex arc. It causes quick<br />

withdrawal from a pain source. From [Hamill 03], p111.<br />

contraction after a rapid stretch. As the stretch is imposed on the muscle, the Type I<br />

sensory neurons <strong>of</strong> the muscle spindles are activated and stimulate interneurons in the<br />

spinal cord. If the activation is high enough, the interneurons will excite the motor neurons<br />

<strong>of</strong> the muscle just stretched and initiate its contractions. At the same time, additional<br />

connections to inhibitory interneurons will cause a reciprocal inhibition, or relaxation, <strong>of</strong><br />

the antagonistic muscle. Similarly, the Golgi tendon organs are responsible for the inverse<br />

stretch reflex. In case <strong>of</strong> a high velocity stretch, the reflex creates a relaxation <strong>of</strong> the<br />

muscle by reducing the alpha motor neuron output, thus relieving the strain <strong>of</strong> the muscle.<br />

Another example for monosynaptic reflexes is the flexor reflex (Figure 3.6). It triggers a<br />

rapid withdrawal movement due to a sensory information indicating pain.<br />

Reflexes processing information from different levels <strong>of</strong> the spinal cord and affecting muscles<br />

further apart are termed propriospinal. The crossed extensor reflex is an example for this<br />

group <strong>of</strong> reflexes. Similar to the flexor reflex, it reacts to pain stimulus, but it also initiates<br />

an extention <strong>of</strong> the limb opposite to the pained one. Another example for propriospinal<br />

reflexes is the tonic neck reflex. Here, a rotation <strong>of</strong> the head to one side is causing an<br />

extension <strong>of</strong> the arm on the same side and a flexion <strong>of</strong> the arm on the contralateral side.<br />

If the motor response results from information transmitted to and processed by the<br />

brain, a reflex is called supraspinal. Most postural reflexes belong to this group, e.g. the<br />

labyrinthine righting reflex. Incorporating upper levels <strong>of</strong> the nervous system and many<br />

levels <strong>of</strong> the spinal cord, it stimulates the neck and limb muscles to maintain or move to<br />

an upright position.<br />

However, reflexes do not always react to sensory stimulus in the same, recurring way. While<br />

monosynaptic reflexes show little variation, more complex reflexes seem to be modulated.<br />

Zehr et al. examined the function <strong>of</strong> sural nerve reflexes during human walking [Zehr 98].<br />

They show that the muscular response initiated by the reflexes depends on the part <strong>of</strong><br />

the step cycle in which the nerve was stimulated as well as on the intensity <strong>of</strong> stimulation.<br />

These cutaneous reflexes mainly seem to serve for postural stabilization during walking,<br />

e.g. lifting the leg during the first part <strong>of</strong> the swing phase by ankle dorsiflexion and knee<br />

flexion in case the foot collides with an obstacle.


32 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

Zehr and Stein also review research on the modulation <strong>of</strong> reflex responses during static<br />

tasks and locomotion [Zehr 99]. They show the existence <strong>of</strong> task-, phase- and intensity<br />

dependency. A simple modulation <strong>of</strong> the soleus H-reflex 2 depending on the subjects<br />

postural orientation can be observed. While keeping a constant peripheral stimulation, the<br />

amount <strong>of</strong> reflex inhibition changes [Rossi 88]. Similar modulation effects can be shown<br />

on the tonic motor output by cutaneous nerve stimulation while sitting or standing.<br />

Rossignol et al. review literature on the dynamic sensorimotor interactions in the spinal<br />

cord and at supraspinal levels [Rossignol 06]. It is postulated that locomotion emerges<br />

as a combination <strong>of</strong> central, feed-forward programs and feedback mechanisms. The<br />

central program originates in some kind <strong>of</strong> spinal circuitry, a central pattern generator<br />

that is genetically prescribed and generates basic motion patterns to trigger, stop, or<br />

steer locomotion. Feedback provoked by sensory afferents from spinal to supraspinal<br />

levels modulates the locomotor pattern. This interaction is a phase-dependent, corrective<br />

response not only due to heavy perturbations, but also present during normal, stable<br />

locomotion. Sensorimotor interaction can interfere with the descending pathways at various<br />

levels <strong>of</strong> the nervous system and results in a continuous adjustment <strong>of</strong> the overall emergent<br />

locomotion behavior.<br />

Implications for Robotics:<br />

Reflexes provide the basic method for feedback control in nature.<br />

They work independently, locally, and in parallel. Reflexes can occur<br />

in different degrees <strong>of</strong> complexity in terms <strong>of</strong> the number <strong>of</strong> sensory<br />

receptors and muscles that are involved and <strong>of</strong> the level <strong>of</strong> the central<br />

nervous system they are located on. Reflexes can be modulated by<br />

the intensity <strong>of</strong> the stimulus, by the current phase <strong>of</strong> motion, or by the<br />

current task in general (walking vs. sitting). Combined with central,<br />

feed-forward programs stable locomotion can be achieved.<br />

3.1.4 Hierarchical Layout <strong>of</strong> Motion <strong>Control</strong><br />

To produce a limb motion during grasping or locomotion, a multitude <strong>of</strong> muscles have to<br />

work together in a coordinated manner. For each <strong>of</strong> these muscles, again multiple muscle<br />

neurons have to be stimulated with suitable intensity. It does not seem <strong>like</strong>ly that for each<br />

motion, either voluntary or instinctive, this vast number for neurons is directly controlled<br />

by a region in the brain. Rather, one might suggest that motion control takes place on<br />

successive levels.<br />

Neuroscientific research results seem to support the assumption that neural motor control<br />

is <strong>of</strong> a hierarchical layout. Bizzi et al. found a spatial connection between the stimulation <strong>of</strong><br />

regions in the spinal cord <strong>of</strong> frogs and the kinematic reaction <strong>of</strong> its legs [Bizzi 91, Lemay 01].<br />

Exciting these interneurons at increasing strength produces similarly increasing force<br />

measured at the ankle <strong>of</strong> the frog’s limb. Co-stimulation <strong>of</strong> different regions in the spinal<br />

cord results in a vectorial sum <strong>of</strong> the corresponding ankle forces. This suggests the existence<br />

<strong>of</strong> movement primitives creating activities <strong>of</strong> whole groups <strong>of</strong> muscles. A limited number<br />

<strong>of</strong> these simple units can generate a rich variety <strong>of</strong> movements.<br />

2 H<strong>of</strong>fmann reflex: reaction <strong>of</strong> muscles after electrical stimulation <strong>of</strong> sensory fibers in their innervating<br />

nerves


3.1. Structural Organization <strong>of</strong> Motion <strong>Control</strong> 33<br />

Figure 3.7: Decomposition <strong>of</strong> 32 emg waveforms during walking resulting in five basic temporal<br />

components. Different statistical methods produce similar components. From [Ivanenko 06],<br />

p340.<br />

In order to analyze to what extent muscle activity is generated by centrally organized<br />

movement primitives (or muscle synergies), electromyographic 3 (emg) signals <strong>of</strong> intact<br />

and deafferented (limited sensor feedback) frogs during swimming and jumping are compared<br />

[Cheung 05]. Synergies can be found by a modified formulation <strong>of</strong> the non-negative<br />

matrix factorization algorithm. Most synergies are not specific to one but are shared<br />

by both data sets, thus suggesting a central origin and organization. However, with the<br />

reduced sensor feedback, the synergies are altered in amplitude and timing, indicating that<br />

some modification by feedback is taking place. It is further speculated that each shared<br />

synergy corresponds to a basic biomechanical function.<br />

Later results on the combination <strong>of</strong> such muscle synergies for movement show that the<br />

same synergies are even used for different modes <strong>of</strong> locomotion [Bizzi 07]. Each synergy<br />

can be seen as function unit in the spinal cord producing a specific pattern <strong>of</strong> muscle<br />

activation. By generating a more sophisticated motion as combination <strong>of</strong> these simpler<br />

modules, the complexity <strong>of</strong> locomotion control is reduced.<br />

Analysis <strong>of</strong> human locomotion control leads to similar findings. Earlier work by Patla,<br />

Vaughan, et al. examining emg patterns <strong>of</strong> the lower limbs during walking using patterns<br />

3 More information on electromyography will be given in Section 3.2.1


34 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

Figure 3.8: Mapping muscle activation during normal walking to the position <strong>of</strong> the corresponding<br />

motor neurons in the spinal cord. From [Ivanenko 06], p341.<br />

recognition algorithms already suggest variant and invariant features among these signals<br />

[Patla 85, Vaughan 92, Olree 95]. Ivanenko et al. include a larger set <strong>of</strong> muscles in<br />

their analysis [Ivanenko 04]. They use statistical methods <strong>like</strong> factor analysis or pca for<br />

the decomposition <strong>of</strong> 32 emg waveforms recorded during normal walking (Figure 3.7).<br />

Five basic temporal components can be identified whose weighted sum explains about<br />

90% <strong>of</strong> the total emg waveforms. This holds true even for different walking velocities<br />

and for various degrees <strong>of</strong> body weight support. Ivanenko and colleagues also apply their<br />

methods to data sets <strong>of</strong> other groups, e.g. Winter’s or Vaughan’s data, and produce similar<br />

results [Ivanenko 06]. Moreover, the identified components seem to be timed to kinematic<br />

or kinetic events.<br />

Mapping the emg activity <strong>of</strong> muscles back to the location <strong>of</strong> their corresponding motor<br />

neurons in the spinal cord also yields interesting results [Ivanenko 06]. Figure 3.8 shows<br />

such a spatiotemporal mapping <strong>of</strong> normal walking activity. These maps show common<br />

features over different walking speeds, e.g. bursts <strong>of</strong> activity at different locations that are<br />

temporally aligned. This could be seen as evidence that coordinated activity origins from<br />

a central pattern.


3.1. Structural Organization <strong>of</strong> Motion <strong>Control</strong> 35<br />

(a) (b)<br />

Figure 3.9: Hypothetical scheme <strong>of</strong> muscle activation. (a) An activation component creates<br />

a muscle synergy influenced by central control and feedback. (b) Activation components<br />

are triggered by central spinal pattern generator and/or higher levels <strong>of</strong> the nervous system.<br />

From [Ivanenko 06], p347.<br />

The concept <strong>of</strong> central motor patterns can also be tested by disturbing basic locomotion<br />

with voluntary movements. For example, when stepping over an obstacle during walking,<br />

an additional activation component can be identified alongside the five basic components<br />

<strong>of</strong> walking. When performing the same voluntary task while standing, in this case lifting<br />

one leg, again characteristic activation patterns are found. These can be superimposed on<br />

the walking components and timed with the corresponding kinematic events, resulting in<br />

overall activation similar to the one observed in the combined movement.<br />

This allows the assumption that motor patterns are not only used for one specific movement<br />

but can be called on for different tasks. Further evidence for this can be found when<br />

comparing the temporal components during walking and running [Cappellini 06]. The<br />

same five components can account for both modes <strong>of</strong> locomotion with only a phase shift<br />

<strong>of</strong> one <strong>of</strong> the components. Thus, despite their diverse biomechanical demands, walking<br />

and running mainly differ in the timing <strong>of</strong> motor programs.<br />

Considering the results discussed in the previous paragraphs, Ivanenko et al. suggest a<br />

hypothetical scheme <strong>of</strong> muscle activation as shown in Figure 3.9. Components or motor<br />

patterns create muscle activation via a weighting network <strong>of</strong> distributed interneurons<br />

connected to muscle neurons. Feedback and central control can alter the weights. Central<br />

control and/or rhythmic pattern generation in the spinal cord create a spatial and temporal<br />

sequence <strong>of</strong> activation and superposition <strong>of</strong> voluntary motor programs. Again, sensory<br />

feedback and downstream commands may modulate this process.<br />

In his review on central control <strong>of</strong> muscle activity during human walking, Nielsen states<br />

that neural control is located on many levels <strong>of</strong> the nervous system [Nielsen 03]. In the<br />

human more than in the animal kingdom, the supraspinal level gets involved, as becomes<br />

visible in positron emission tomography (pet) <strong>of</strong> the motor cortex. There even is evidence<br />

<strong>of</strong> a direct monosynaptic pathway having developed from the motor cortex to spinal motor<br />

neurons. But still the spinal circuitry is capable <strong>of</strong> generating basic rhythmic locomotion<br />

activity patterns.


36 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

Figure 3.10: Photographs taken by Eadweard Muybridge in the late 19th century for analyzing<br />

human locomotion.<br />

Implications for Robotics:<br />

Motion control in humans is <strong>of</strong> hierarchical layout. Motor programs<br />

provide an interlayer to create coordinated synergies <strong>of</strong> movement.<br />

Activation <strong>of</strong> these patterns can create locomotion if they are timed to<br />

correspond to kinematic or kinetic events. Feedback should be<br />

integrated to modulate the patterns according to outside influence and<br />

internal state.<br />

This section has discussed the structural organization and principles deployed in human<br />

motioncontrol. Thefollowingsectionwillspecificallycoverhumanwalkinganddemonstrate<br />

how the mechanisms above are employed during this form <strong>of</strong> locomotion.<br />

3.2 Normal <strong>Walking</strong> in <strong>Human</strong>s<br />

The study <strong>of</strong> human locomotion has a long history. Already in the 19th century series<br />

<strong>of</strong> photographs were taken to better understand the motions <strong>of</strong> animals and man. With<br />

the help <strong>of</strong> chief engineer for the Southern Pacific Railroad, John D. Isaacs, Eadweard<br />

Muybridge (1830–1904) developed a method for capturing a sequence <strong>of</strong> pictures using<br />

multiplecameras. Thatwayhetookthousands<strong>of</strong>images<strong>of</strong>animalsandhumansperforming<br />

various task (Figure 3.10).<br />

3.2.1 Biomechanical Gait Analysis<br />

Although Muybridge’s scientific interest in the topic <strong>of</strong> motion analysis is arguable, his<br />

contribution can still be seen as an important step during the development <strong>of</strong> the research


3.2. Normal <strong>Walking</strong> in <strong>Human</strong>s 37<br />

Figure 3.11: A typical setup for biomechanical gait analysis featuring kinematic tracking using<br />

markers, emg measurement, and force plates for capturing ground reaction forces (Source: bts<br />

Bioengineering, Italy).<br />

field called kinesiology, or in modern and more general terms, biomechanics. Today’s<br />

methods for analyzing human gait have advanced significantly, including high-speed<br />

cameras, or X-ray systems. Figure 3.11 shows a typical setup for capturing gait data<br />

with multiple sensor systems. The topics <strong>of</strong> biomechanical research can be divided in<br />

the examination <strong>of</strong> functional anatomy (see Section 3.1.1), the kinematics and kinetics <strong>of</strong><br />

motions, and the capturing <strong>of</strong> neural activity.<br />

Kinematics<br />

Kinematics describes the spatial and temporal elements <strong>of</strong> motion, i.e. the position,<br />

velocity, and acceleration <strong>of</strong> rigid bodies. In a biomechanical context, this refers to the<br />

translation and rotation <strong>of</strong> body segments relative to a fixed coordinate system or to<br />

one another. Quantitative methods to collect kinematic data during locomotion include<br />

accelerometers or other measurement devices attached to individual body parts, or most<br />

recently X-radiation as used for small mammals. The most common method for obtaining<br />

data uses motion capture systems based on high-speed cameras or optoelectric systems.<br />

These biomechanical capturing units usually operate at 60, 120, 180, or 200 frames per<br />

second. Active or passive markers are fixed on the subject, normally at the end points <strong>of</strong><br />

the body segments to be analyzed (Figure 3.11). The marker positions are then tagged<br />

manually or automatically in the digitized video data. Thus, the global marker trajectories<br />

are known given a calibrated capturing system. Including the kinematic connections <strong>of</strong> the<br />

markers, joint angles can be deduced for each frame. By numerical derivation, velocities<br />

and accelerations can be calculated. Beside the joint angles, parameters <strong>like</strong> stride length,<br />

walking speed, step rate or stance, and swing durations are <strong>of</strong> interest when kinematically<br />

analyzing walking motions.


38 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

Figure 3.12: Joint angles <strong>of</strong> the lower extremities during normal walking (solid line), running<br />

(broken line), and sprinting (dotted line) over the gait cycle. The vertical line represents the<br />

transition from stance (left side) to swing phase (right side). From [Hamill 03], p327.<br />

Many researchers have reported on the kinematics during normal walking in humans.<br />

Figure 3.12 illustrates typical hip, knee, and ankle joints during walking, running, and<br />

sprinting [Hamill 03]. While the role movement in the frontal and transverse plane should<br />

not be neglected, the major segment movement takes place in the sagittal plane. At<br />

ground contact, the body weight and downwards acceleration needs to be absorbed, which<br />

is achieved by hip flexion, knee flexion, and ankle dorsiflexion. These flexions continue<br />

while the body travels over the stance foot. At the end <strong>of</strong> the stance phase, the joint<br />

action reverses into hip extension, knee extension, and ankle plantarflexion.<br />

During walking, the hip flexion at touchdown lies in the range <strong>of</strong> 35 ◦ to 40 ◦ . Until toe-<strong>of</strong>f,<br />

the hip flexion reduces to about 0 ◦ to 3 ◦ . During the swing phase, the hip flexes again<br />

with a maximum <strong>of</strong> 25 ◦ to 50 ◦ .<br />

The knee is flexed to 10 ◦ to 15 ◦ at touchdown to reduce the impact forces. While accepting<br />

body weight at the beginning <strong>of</strong> the stance phase, the knee flexes further and reaches its<br />

maximum at midstance with 20 ◦ to 25 ◦ . The knee then stretches without reaching full<br />

extension, and flexes again during the propulsion phase <strong>of</strong> stance. At toe-<strong>of</strong>f, the knee<br />

angle ranges in between 10 ◦ to 40 ◦ , higher values being reached at faster walking speeds.<br />

The leg length is shortend during swing phase to guarantee ground clearance by further<br />

knee flexion <strong>of</strong> about 50 ◦ to 65 ◦ .<br />

The ankle shows a plantarflexion <strong>of</strong> 5 ◦ to 6 ◦ at heel strike. It then slowly moves to<br />

dorsiflexion <strong>of</strong> 10 ◦ to 12 ◦ until the whole foot touches the ground. During the stance phase,<br />

it gradually returns to plantarflexion <strong>of</strong> about 15 ◦ to 20 ◦ at toe-<strong>of</strong>f. Increasing dorsiflexion<br />

keeps the toes from touching the ground during the swing phase.<br />

The angular kinematics <strong>of</strong> the lower extremities adapt to changes <strong>of</strong> the environment. For<br />

example, walking on non-compliant ground can cause stronger knee flexion at heel strike.<br />

Uphill walking is compensated by additional dorsiflexion, knee flexion, and hip flexion at


3.2. Normal <strong>Walking</strong> in <strong>Human</strong>s 39<br />

Activity Relative Force (N/BW)<br />

Vertical compressive forces in the ankle joint 13<br />

Vertical reaction forces in the ankle joint 3.9–5.2<br />

Vertical reaction forces in the subtalar joint 2.4–2.8<br />

Achilles tendon force 3.9<br />

Peak forces acting in the hip 2.8–4.8<br />

Table 3.1: Maximum relative forces acting on the human body during walking. Forces are given<br />

in newton (N) per body weight (BW) [Hamill 03].<br />

touchdown as well as an increase <strong>of</strong> the motion range in the hip and ankle during stance<br />

phase. When walking downhill, the knees show up to 15 ◦ more flexion at heel strike.<br />

Rotation <strong>of</strong> the pelvis during walking only lies within a few degrees. The same holds true<br />

for the movement <strong>of</strong> the upper trunk. It moves laterally for about 2 to 3cm to the side <strong>of</strong><br />

the stance leg and slightly leans forward after heel strike and back again in the course <strong>of</strong><br />

the stance phase [Murray 64]. Vertical displacement <strong>of</strong> the trunk reaches its maximum <strong>of</strong><br />

about 5cm at midstance <strong>of</strong> each leg.<br />

Kinetics<br />

In biomechanics, kinetics describes the effects <strong>of</strong> internal and external forces and torques<br />

acting on the human body, <strong>of</strong>ten given relative to the body mass or body weight. Typical<br />

relative forces acting on the body during walking are given in Table 3.1.<br />

Forces acting on the human body can be classified into non-contact and contact forces.<br />

The most relevant non-contact forces during locomotion is gravity. Gravity acts at the<br />

center <strong>of</strong> mass <strong>of</strong> each body segment with the gravitational acceleration <strong>of</strong> about 9.81<br />

m/s 2 in vertical direction. Seven contact forces, i.e. forces acting between two bodies, are<br />

normally considered in biomechanics: ground reaction force, joint reaction force, friction,<br />

fluid resistance, inertial force, muscle force, and elastic force [Hamill 03].<br />

The ground reaction force (grf) acts between the ground and an individual body segment<br />

making contact with it. Figure 3.13a shows a typical behavior <strong>of</strong> the grf during the<br />

stance phase <strong>of</strong> a walking cycle. As the grf is a vector, it can be resolved into a vertical,<br />

an anteroposterior (forward-backward) and a mediolateral (side-to-side) component. The<br />

latter two forces act parallel to the ground surface and are referred to as shear forces.<br />

During walking, the vertical component <strong>of</strong> the grf is M-shaped with its first peak during<br />

weight acceptance, and the second peak during push-<strong>of</strong>f. The grf acts at point <strong>of</strong> pressure<br />

in the foot sole. As shown in Figure 3.13b, with the progress <strong>of</strong> the stance phase, the<br />

point <strong>of</strong> pressure travels along the sole from heel to toes [Rodgers 88]. In biomechanical<br />

research, grf is normally measured using a force platform embedded in the ground on<br />

which the subject is performing e.g. walking or jumping.<br />

Two adjoining body segments are connected by a joint at which point equal but opposing<br />

forces act on each segment. These forces are termed joint reaction forces and can be<br />

calculated based on kinematic and kinetic data and known body dimensions.<br />

Friction is mainly <strong>of</strong> interest during the ground contact <strong>of</strong> the foot or shoe. The force<br />

generated by friction is equal to the normal force onto the surface times a friction coefficient.


40 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

(a) (b)<br />

Figure 3.13: (a) Ground reaction forces during normal walking. The vertical ground reaction<br />

force(solidredline)showsthetypicalMshape. Anteroposterior(solidblackline)andmediolateral<br />

(dashed red line) are shear forces in the surface plane. From [Hamill 03], p365. (b) The center <strong>of</strong><br />

pressure travels along the foot sole during stance phase. From [Rodgers 88], p1825.<br />

For producing the necessary propulsion force in the later stance phase and to guarantee a<br />

safe landing at heel strike during walking, the foot friction must be high enough. Overly<br />

high friction can increase joint tension and cause instability or injuries.<br />

As the drag force in air or fluid resistance is proportional to the medium’s viscosity and<br />

to the square <strong>of</strong> the relative velocity, it does not substantially influence the process <strong>of</strong><br />

walking. Rather it is <strong>of</strong> interest in biomechanical studies <strong>of</strong> running, swimming, or other<br />

movement forms <strong>of</strong> faster speed or in viscose medium.<br />

In contrast, inertial forces do play a role in walking and cannot be neglected. An inertial<br />

force occurs between connected body segment as one segment is moving and thus causing<br />

a movement in the next without muscle action. For example during the swing phase, the<br />

tight segment exerts an inertial force on the leg.<br />

Obviously, muscle force is a force <strong>of</strong> major importance during locomotion. As already<br />

described in Section 3.1.2, muscle fibers contract due to stimulation by the motor neurons.<br />

On average, skeletal muscle can contract to about 57% <strong>of</strong> its resting length [Hamill 03].<br />

Only by contraction, a muscle can produce force, so a joint can only be controlled by<br />

opposing muscles, the agonist and antagonist. When a force is produced by an opposing<br />

muscle, or when an external force acts on a joint, a muscle can be extended beyond its<br />

resting length, storing elastic energy in the process.<br />

Depending on the type <strong>of</strong> fibers, muscles can quickly produce high forces, but then fatigue<br />

(Type II), or have slower contraction times and be more enduring (Type I). The force a<br />

muscle produces depends <strong>of</strong> its current length and the velocity <strong>of</strong> movement. The longer<br />

the muscle, the higher the force created. Thus, pre-stretching the muscle before e.g. weight<br />

lifting can increase the force output. The relationship between velocity and force is shown<br />

in Figure 3.14a. Regarding movement, the force produced by muscle action is used for<br />

joint movement, the maintenance <strong>of</strong> posture and positions, and joint stabilization.


3.2. Normal <strong>Walking</strong> in <strong>Human</strong>s 41<br />

(a) (b)<br />

Figure 3.14: (a) Increasing velocity reduces the force output <strong>of</strong> a muscle. This holds true for<br />

the lengthening muscle (eccentric muscle action) as well as the shortening muscle (concentric<br />

muscle action). (b) Hill’s mechanical muscle model in two equivalent forms. From [Hamill 03],<br />

p78f.<br />

Therehavebeenseveralattemptstodescribetheforceproducedbyamuscleinamechanical<br />

model, the most prominent being the model developed by A. V. Hill (Figure 3.14b). It<br />

consists <strong>of</strong> the contractile component (cc), a parallel elastic component (pec) and a series<br />

elastic component (sec). The sec describes the tendon and other elastic elements in series<br />

with the contracting units <strong>of</strong> the muscle. It shows highly nonlinear behavior, as does the<br />

pec expressing parallel elastic elements <strong>like</strong> the fascia surrounding the muscle and its<br />

compartments.<br />

As soon as the line <strong>of</strong> action <strong>of</strong> a force does not pass through the body’s center <strong>of</strong> mass, a<br />

torque, or moment <strong>of</strong> force, is induced. Torques occur when muscle contraction pulls via<br />

tendons at a certain distance away from the joint, when gravity acting on body segments<br />

causes a downward rotation, or when the direction on the ground reaction force does<br />

not pass through the center <strong>of</strong> gravity. The total center <strong>of</strong> mass <strong>of</strong> the body can be<br />

calculated from the segment centers <strong>of</strong> mass and masses. Table 3.2 gives the average<br />

values for both male and female subject as estimated by Plagenhoef et al. using segment<br />

immersion [Plagenhoef 83]. Other researchers examined cadavers for the same purpose.<br />

Using similar techniques, the average moment <strong>of</strong> inertia has been estimated for individual<br />

body segments. The moment <strong>of</strong> inertia necessary to calculate the angular acceleration <strong>of</strong> a<br />

body given an external torque amounts to<br />

T = I ¨ θ (3.1)<br />

where T is the torque, I the moment <strong>of</strong> inertia, and ¨ θ the angular acceleration.<br />

Static and dynamic gait analysis deals with the forces and torques just mentioned as they<br />

act on the human body. Leonard Euler created recursive equations for a dynamic analysis<br />

based on Newton’s laws <strong>of</strong> motion starting with the most distal body segments. Applying<br />

inverse dynamics calculations, the forces and torques acting on body segments can be<br />

concluded from their motions and masses. Motion capture data can provide information


42 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

Center <strong>of</strong> Mass Segment Weight<br />

Segment Male Female Male Female<br />

Head, neck 55.0 55.0 8.26 8.20<br />

Trunk 44.5 39.0 46.80 45.22<br />

Whole trunk 63.0 56.0 55.10 53.20<br />

Upper arm 43.6 45.8 3.25 2.90<br />

Forearm 43.0 43.4 1.87 1.57<br />

Hand 46.8 46.8 0.65 0.50<br />

Thigh 43.3 42.8 10.50 11.75<br />

Lower leg 43.4 41.9 4.75 5.35<br />

Foot 50.0 50.0 1.43 1.33<br />

Table 3.2: Average center <strong>of</strong> mass and weight <strong>of</strong> body segments. Center <strong>of</strong> mass location is<br />

given in percent <strong>of</strong> segment length from the proximal end. Weight is given in percent <strong>of</strong> the total<br />

body weight [Plagenhoef 83].<br />

on segment trajectories, masses and inertia can be estimated based on average mass<br />

distributions <strong>of</strong> the human body (Table 3.2). Further variables <strong>of</strong> interest in kinetics are<br />

work, power, or energy being mustered or consumed during motions. For example, the<br />

pendulum-<strong>like</strong> characteristics <strong>of</strong> walking exchanges potential with kinetic energy and vice<br />

versa. This mechanism allows energy savings <strong>of</strong> muscular work <strong>of</strong> up to 65%.<br />

Figure 3.15 illustrates angular kinematic and kinetic joint data during a walking cycle<br />

for the hip, the knee, and the ankle joint [Hamill 03]. At heel strike, the hip produces an<br />

extensor moment to stabilize the trunk. The leg swing is generated by hip flexion before<br />

toe-<strong>of</strong>f and is terminated by an opposing moment at the end <strong>of</strong> swing phase.<br />

Weight acceptance is controlled in the knee joint by an extensor moment. Leg swing is<br />

initiated by a small amount <strong>of</strong> flexion moment and controlled by the following extensor<br />

action. During the swing there is almost no power output in the knee, only towards the<br />

end <strong>of</strong> the swing a flexion is created in preparation for ground contact.<br />

The foot is slowly lowered to the ground after heel strike by a small amount <strong>of</strong> dorsiflexor<br />

moment. Thebody’smovementoverthefootduringstancephaseiscontrolledbyincreasing<br />

plantarflexor moment. The power output <strong>of</strong> the ankle shows its maximum during the<br />

propulsion before toe-<strong>of</strong>f. During swing phase, the ankle produces nearly no power.<br />

As described in the kinematics discussion <strong>of</strong> normal walking, the moment and power<br />

patterns also change due to ground condition or walking speed. For example, higher flexor<br />

and extensor moments can be observed in the knee during faster walking.<br />

Electromyography<br />

Besides the capturing <strong>of</strong> human motion and the kinematic and kinetic analysis <strong>of</strong> that<br />

data, the most frequently used method <strong>of</strong> motion analysis is the electromyography (emg)<br />

recording activation signals <strong>of</strong> muscles. More than the methods mentioned above, emg can<br />

provide inside into the control <strong>of</strong> reflexive or voluntary movements as the muscle action is<br />

directly connected to neural commands. However, analyzing emg plots has its limitations<br />

and thus requires some knowledge on this method, which will be provided in the next few


3.2. Normal <strong>Walking</strong> in <strong>Human</strong>s 43<br />

Figure 3.15: Kinetics during normal walking in the sagittal plane. The toe-<strong>of</strong>f event is<br />

represented by the solid vertical line. From [Hamill 03], p414.<br />

paragraphs. The main sources for this section are the books by Hamill et al. and Vaughan<br />

et al. [Hamill 03, Vaughan 92].<br />

To record an electromyogram, an electrode is placed on the muscle to detect an electrical<br />

signal. This emg signal is a complex combination <strong>of</strong> all action potentials generated by the<br />

muscle neurons connected to the muscle fibers that are gathered by the electrode. The raw<br />

signal is both positive and negative. While the amplitude <strong>of</strong> the signal is influenced by<br />

many factors (see below), in principle it increases with the strength <strong>of</strong> muscle contraction.<br />

However, as indicated in the discussion on muscle properties above, the amplitude is by no<br />

means relative to the produced muscle force or even the joint torque. A somehow direct<br />

correlation between emg signal amplitude and muscle force can only be seen in isometric<br />

contraction. Also, the so-called electromechanical delay between the electrical signal and<br />

the actual muscle contraction can be observed, as the action potential must travel along<br />

the muscle fibers before tension can be developed.<br />

Two types <strong>of</strong> electrodes can be used to detect the electrical signal: indwelling or surface<br />

electrodes. Indwelling electrodes are placed directly in deep muscle in form <strong>of</strong> needles or<br />

fine wire. More commonly used in biomechanics are surface electrodes placed on the skin<br />

and as such used for more superficial muscles. There are two possible arrangements for<br />

surface electrodes: in the monopolar arrangement a single electrode is placed over the


44 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

Figure 3.16: Common representations <strong>of</strong> an emg signal. From [Hamill 03], p84.<br />

muscle, a second one is attached over an electrically neutral place <strong>like</strong> a bony prominence<br />

to get a differential signal. The bipolar arrangement is more precise as another electrode<br />

is placed 1.5 to 2cm next to the first muscle electrode and a differential amplifier can be<br />

used to cancel out common error signals. Obviously, correct placement and orientation<br />

<strong>of</strong> the electrodes on properly prepared skin is important to produce a significant signal<br />

with low crosstalk. As emg signal intensity is very low ranging from 10 �V to 5mV, it<br />

must be amplified by high quality electronics to reduce additional noise or errors. It is<br />

then digitalized by an A/D converter working at least at 1kHz to take the characteristics<br />

<strong>of</strong> neural muscle activity into account.<br />

The technical and physiological factors influencing the emg signal are manifold. They<br />

include factors <strong>like</strong> muscle fiber diameter, the number <strong>of</strong> muscle fibers, the number <strong>of</strong><br />

active motor units, fiber type and location, muscle blood flow, and further effects directly<br />

concerning the muscle. Other factors <strong>like</strong> electrode spacing or electrode-skin interface also<br />

influence the collected data.<br />

Figure 3.16 illustrates different representations <strong>of</strong> an emg signal. The raw signal shows<br />

the composite <strong>of</strong> the multitude <strong>of</strong> action potentials <strong>of</strong> the motor units being active in the<br />

effective range <strong>of</strong> the electrodes. Commonly the raw signal is rectified, i.e. the absolute<br />

value <strong>of</strong> the signal is examined. Filtering out high frequency components, a smooth form <strong>of</strong><br />

the signal called linear envelope can be generated showing more clearly the volume <strong>of</strong> the<br />

activity. In some cases <strong>of</strong> analysis, the integration <strong>of</strong> the emg signal or the transformation<br />

in the frequency domain can be <strong>of</strong> interest.<br />

emg recordings can be used for a number <strong>of</strong> applications. There are attempts to define<br />

a relationship between muscle force and emg signal strength in isometric muscle action.<br />

Also, there is some evidence that frequency changes in emg signals can indicate muscle<br />

fatigue. In ergonomics, muscle activity analysis can help to point out improper sitting<br />

postures, movements, or working conditions. But most interesting for this work is the<br />

clinical gait analysis. Here, the emg signal can be interpreted to identify the contribution<br />

and timing <strong>of</strong> individual muscles or muscle groups during a gait cycle. This allows to<br />

draw certain conclusions regarding the underlying neural control. However, is must be<br />

mentioned that emg gives semiquantitative results at best regarding the strength <strong>of</strong> muscle<br />

contraction. Also it can prove difficult to obtain reliable emg reading during dynamic<br />

movements such as walking or running.


3.2. Normal <strong>Walking</strong> in <strong>Human</strong>s 45<br />

Muscles Footstrike Midsupport Toe-<strong>of</strong>f Forward Swing Deceleration<br />

Dorsiflexors +++ ++ ++ ++ ++<br />

Intrinsic Foot Muscles +++<br />

Gluteus Maximus + ++ +++<br />

Gluteus Medius ++ +++ ++ +<br />

Gluteus Minimus ++ +++ ++ +<br />

Hamstrings +++ ++ ++ + ++<br />

Iliopsoas +++<br />

Plantar Flexors + ++<br />

Quadriceps + +++ ++<br />

Sartorius ++ +<br />

Tensor Fascia Latae + ++ + +++<br />

Thigh Adductors ++ ++ + ++ +<br />

Table 3.3: Contribution <strong>of</strong> the main muscle groups in the lower extremities during normal<br />

walking. +/++/+++ is indicating low/moderate/high muscle activity [Hamill 03].<br />

Many researchers have traced emg signals during normal human walking. As shown in<br />

Figure 3.17, Winter et al. recorded relevant muscle group activities <strong>of</strong> the whole body<br />

during walking [Winter 91]. Table 3.3 lists the contribution to walking <strong>of</strong> the main muscles<br />

groups in the lower extremities [Hamill 03]. The meaning <strong>of</strong> muscle abbreviations, muscle<br />

locations, and their function can be found in Appendix A.<br />

The foot and ankle muscles control the ground contact during walking and create much <strong>of</strong><br />

the necessary propulsion. To keep the front <strong>of</strong> the foot from slapping down after heel strike,<br />

the dorsiflexor muscles <strong>like</strong> the tibialis anterior show maximum activity. This activity then<br />

decreases but nevertheless remains to control the pendulum-<strong>like</strong> tibia movement during<br />

stance phase, assisted by beginning activity <strong>of</strong> the gastrocnemius and soleus. The latter<br />

muscle group mainly contributes to propulsion just before toe-<strong>of</strong>f when it reaches peak<br />

activity. During swing phase, only the dorsiflexors are active to guarantee ground clearance<br />

<strong>of</strong> the foot.<br />

The knee musculature is mainly used for stabilization and force absorption during walking.<br />

At heel strike the hamstrings, biarticular muscles spanning the knee and the hip, get active<br />

to keep the trunk from swinging forward and at the same time to bend the knee to reduce<br />

impact forces. This contraction continues until the foot is flat on the ground. During<br />

weight acceptance, the quadriceps femores work to control the body load and to allow<br />

the knee to flex in a controlled way. As the leg extends towards midstance, this activity<br />

decreases. During propulsion, the knee muscles become active again to add to the forward<br />

force. At toe-<strong>of</strong>f, the hamstrings can flex the knee for increasing ground clearance and<br />

again before heel strike to reduce impact. The quadriceps femores enables leg extension<br />

towards the end <strong>of</strong> the swing phase.<br />

The hip joint and pelvis muscles mainly control trunk and pelvis posture and do not<br />

substantially contribute to propulsion. The gluteus medius provides lateral stability <strong>of</strong><br />

the pelvis and the trunk against the body weight from heel strike to midsupport as the<br />

contralateral leg is swinging and cannot contribute to support. This stabilization is assisted<br />

by activity <strong>of</strong> the gluteus minimus and the adductors. <strong>Control</strong> <strong>of</strong> the trunk in the frontal<br />

plane at heel strike and during stance phase is provided by the hamstrings, the gluteus


46 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

Rectus<br />

Femoris<br />

Tibealis<br />

Anterior<br />

prevent foot from slapping down lift foot for ground clearance<br />

Figure 3.17: emg analysis <strong>of</strong> 25 muscles during a cycle <strong>of</strong> normal walking. Exemplary, the<br />

name and location <strong>of</strong> two muscles is indicated, and two significant muscle activities interpreted.<br />

The meaning <strong>of</strong> the remaining muscle abbreviations can be found in Appendix A. emg data<br />

from [Winter 91], plot from [Ivanenko 06], p341.<br />

maximus and the tensor fascia latae. Towards toe-<strong>of</strong>f, the abductors support the propulsion<br />

process. The leg swing is initiated by activity <strong>of</strong> the iliopsoas, sartorius, and tensor fascia<br />

latae. The adductors control lateral movement <strong>of</strong> the swinging limb until it is decelerated<br />

by the hamstrings and the gluteus maximus before the next heel strike.<br />

The trunk muscles mainly control the lateral movement <strong>of</strong> the trunk as it flexes towards the<br />

side where ground contact is made. This is mainly achieved by activity <strong>of</strong> the multifidus<br />

and longissimus. At heel strike, first the contralateral muscle groups prevent the trunk<br />

from flexing to far, then the ipsilateral group stabilizes the trunk. The activity <strong>of</strong> the<br />

leg extensor muscles are synchronized with the activity <strong>of</strong> the erector spinae muscles.<br />

Excessive flexion <strong>of</strong> the trunk is also prevented by the lumbar muscles. An erect head<br />

posture is controlled by relatively low activity <strong>of</strong> the cervical muscles.<br />

Arm movement is coinciding with the leg movement [Wannier 01]. Activity <strong>of</strong> the flexor<br />

and extensor muscles <strong>of</strong> the arm create a contralateral motion contributing to stabilize<br />

trunk rotation. Figure 3.18 graphically recapitulates the contribution <strong>of</strong> the muscles in<br />

the lower extremities during one walking cycle in posterior and lateral views [Vaughan 92].<br />

Darker shading corresponds to higher muscle activity.<br />

Implications for Robotics:<br />

<strong>Human</strong> gait analysis provides valuable insight in bipedal walking mechanisms.<br />

Kinematic data is only partially helpful as direct mapping <strong>of</strong><br />

joint trajectories on a robot cannot work due to differences in mechanics.<br />

At the very least it can serve as benchmark or comparison<br />

basis. Kinetic data and emg recordings, however, allow to draw conclusion<br />

regarding the intention <strong>of</strong> control commands during individual<br />

walking phases or sensor events. Hence they can assist in conceiving<br />

feed-forward and feedback control components.


3.2. Normal <strong>Walking</strong> in <strong>Human</strong>s 47<br />

Figure 3.18: Illustration <strong>of</strong> active muscles in the lower extremities during normal walking in<br />

posterior and lateral view. From [Vaughan 92], p54.<br />

3.2.2 Phases <strong>of</strong> <strong>Walking</strong><br />

Phases <strong>of</strong> walking have been mentioned throughout this chapter, mainly connected to<br />

kinematic or kinetic event <strong>like</strong> heel strike <strong>of</strong> toe-<strong>of</strong>f. It has been shown in Section 3.1 that<br />

neural motor activity and reflex responses depend on the current phase <strong>of</strong> locomotion.<br />

The question <strong>of</strong> how the phases <strong>of</strong> walking are codes in the nervous system is discussed in<br />

the following.<br />

In insects as well as in lower vertebrates, strong evidence for the existence <strong>of</strong> a central<br />

pattern generator (cpg) can be found [Grillner 79, Pearson 93, Bueschges 95, Baessler 98].<br />

Even in the absence <strong>of</strong> sensory feedback and supraspinal commands, rhythmic walking<br />

behavior can be shown. The review by Frigon and Rossignol among others discusses how<br />

sensorimotor interaction affects the output <strong>of</strong> the cpg and thus the motor patterns in<br />

cats [Frigon 06].<br />

Unfortunately, the situation during humans walking is more complicated. While there is<br />

evidence for cpgs in humans, there seems to be much more influence from supraspinal<br />

levels and sensory information [Nielsen 03]. Still, as shown by Vaughan, Ivanenko and<br />

others and already introduced in Section 3.1.4, a substantial part <strong>of</strong> neural control during<br />

human walking is coordinated muscle action at purposefully timed phases <strong>of</strong> the walking<br />

cycle.<br />

Five different components <strong>of</strong> muscle activity can be identified during one cycle <strong>of</strong> human<br />

walking [Ivanenko 04]. Figure 3.19 illustrates the timing <strong>of</strong> the five patterns. The appearance<br />

<strong>of</strong> these activity patterns can be associated with kinematic and kinetic events during<br />

the walking cycle: (1) weight acceptance, (2) loading or propulsion, (3) trunk stabilization<br />

during double support, (4) toe lift-<strong>of</strong>f, and (5) heel strike<br />

The shape <strong>of</strong> the temporal components is Gaussian-<strong>like</strong>, its peak and width showing a<br />

precise timing and stability over the normalized gait cycle [Ivanenko 06]. This implies that<br />

the activity patterns are scaled regarding to the walking speed: at a high speed <strong>of</strong> 9 km/h<br />

the duration <strong>of</strong> on component is about 90ms, at very slow speeds <strong>of</strong> 1 km/h it increases up<br />

to about 250ms. If a central unit <strong>like</strong> a cpg is responsible for these bursts <strong>of</strong> activity, it


48 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

Figure 3.19: The muscle activity during one cycle in human walking can be explained by five<br />

components. From [Ivanenko 06], p342.<br />

only contributes to muscle action at certain fractions <strong>of</strong> the gait cycle. This contradicts the<br />

classic half-center oscillator being active for one half and inhibited for the other half <strong>of</strong> the<br />

cycle. Rather, this implies a stronger influence <strong>of</strong> sensory input, as this would naturally<br />

scale the activity according to biomechanical events.<br />

Neural walking control does also show signs <strong>of</strong> bilateral coordination <strong>of</strong> muscle activity<br />

[Olree 95, Ivanenko 06]. When observing the timing <strong>of</strong> the five components, four <strong>of</strong><br />

them appear in pairs being temporally synchronized: the patterns 1 and 3 and the patterns<br />

2 and 5 match in their onset and are one-half <strong>of</strong> the cycle apart (Figure 3.20). This fact<br />

could also account for the contralateral movement <strong>of</strong> the arms. The synchronized pairs<br />

<strong>of</strong> muscle activity take place at the heel strikes <strong>of</strong> both legs. Only component 4 being<br />

associated with the ipsilateral leg swing appears without contralateral partner. Again,<br />

the fact <strong>of</strong> bilateral synchronization suggests the existence <strong>of</strong> a central origin <strong>of</strong> muscle<br />

patterns and a timing influenced by sensory input.<br />

Implications for Robotics:<br />

<strong>Walking</strong>phasesshouldbecoordinatedbyacentralcontrolunitlocatedon<br />

aninterlayer<strong>of</strong>thecontrolnetwork. Dependingonthecurrentphase, this<br />

unit should stimulate or modulate motor patterns and reflexes.<br />

Phase transitions probably require sensory information and are not<br />

strictly timed. Certain bilateral phases should be synchronized.<br />

Patterns and reflexes should be weighted or scaled depending on the<br />

walking speed.<br />

3.2.3 Reflex Function during <strong>Walking</strong><br />

While muscle patterns in combination with a cpg-<strong>like</strong> structure as described in the<br />

previous section generate basic walking motions, the question remains how reflex action<br />

is incorporated in the neural control. What functions do spinal reflexes perform during<br />

walking? To generate hypotheses on this topic, adequate mechanical data has to be<br />

collected together with neural data, thus enabling a valid interpretation.<br />

Zehr and Stein review literature on the function <strong>of</strong> the stretch reflex, the load receptor<br />

reflexes, and the cutaneous reflexes during locomotion [Zehr 99]. They suggest that the


3.2. Normal <strong>Walking</strong> in <strong>Human</strong>s 49<br />

Figure 3.20: The five muscle patterns during walking show bilateral synchronization visualized<br />

as mirrored polar plots on the left and as superimposed sequence <strong>of</strong> activation timings on the<br />

right. From [Ivanenko 06], p343.<br />

stretch reflex functionally contributes to the force production during walking. By attaching<br />

mechanical devices to subjects during walking, artificial dorsi- and plantarflexion <strong>of</strong> the<br />

ankle can be produced during all phases <strong>of</strong> walking. Thus it is possible to excite the stretch<br />

reflex <strong>of</strong> e.g. the soleus and measure the response. It can be shown that the reflex action<br />

is at its maximum during stance, nearly disappearing in stance-swing transition, and at<br />

about 50% <strong>of</strong> the maximum during late swing. While it is difficult to separate the force<br />

produced by reflex response and by other possible causes, a functional contribution to<br />

walking is assumed.<br />

Only a few experiments on the role <strong>of</strong> load receptor reflexes in human walking have been<br />

carried out. There seems to be evidence <strong>of</strong> a functional contribution, even if somewhat<br />

smaller than what was observed in quadrupedal vertebrates. A phase dependency as well<br />

as significant reflex response in extensor muscles to translational load perturbations can<br />

be shown. Increasing the body load in splitbelt experiments where each leg walks on a<br />

surface <strong>of</strong> different velocity enhances the walking timing and prolongs stance phases.<br />

Some studies exist on the effect <strong>of</strong> cutaneous reflexes during human walking. Stimulation<br />

<strong>of</strong> the sural, tibial, and sp nerves corresponding to the lateral border, and the ventral and<br />

dorsal surface <strong>of</strong> the foot show phase dependent reaction. For instance, stimulating the<br />

tibial nerve during lift-<strong>of</strong>f causes dorsiflexion <strong>of</strong> the ankle. However, during heel strike<br />

it causes plantarflexion. Furthermore it can be shown that cutaneous reflexes generate<br />

kinematic relevant responses. These motions produce reactions to avoid destabilizing<br />

stumbles during the swing phase and adaptation to uneven terrain in stance to avoid<br />

excessive inversion <strong>of</strong> the ankle. Partly, the correcting movement involve muscle groups <strong>of</strong><br />

the whole lower limb.<br />

Zehr and Stein conclude that reflexes ensure balance and a stable walking pattern during<br />

normal walking and in face <strong>of</strong> perturbations. During swing phase, reflexes cause stumbling<br />

correction and trajectory stabilization <strong>of</strong> the swing limb. In the transition from swing to<br />

stance, reflexes work towards stable foot placement and ground adaption. Similar, during<br />

stance, reflexes alter the ankle trajectory depending on terrain structure, stabilize the<br />

body weight support, and enhance cycle timing. Finally in stance to swing transition,<br />

stretch and load reflexes react on unloading by extending the stance phase, and cutaneous<br />

reflexes generate withdrawal motions. Figure 3.21 illustrates the reflex functions during<br />

these four phases <strong>of</strong> walking.


50 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

Swing<br />

CUTANEOUS:<br />

stumble correction<br />

MUSCLE:<br />

trajectory stabilization<br />

Swing to Stance<br />

CUTANEOUS<br />

and<br />

MUSCLE:<br />

placing reactions<br />

Stance<br />

MUSCLE:<br />

weight support, stability,<br />

and cycle timing<br />

CUTANEOUS:<br />

stability<br />

Stance to Swing<br />

MUSCLE:<br />

unloading<br />

CUTANEOUS:<br />

withdrawal<br />

Figure 3.21: Function <strong>of</strong> stretch reflex, load receptor reflexes and cutaneous reflexes during<br />

locomotion. Reproduced from [Zehr 99].<br />

Implications for Robotics:<br />

Low-level reflexes help to finely regulate biped walking and to correct<br />

perturbations. Given the necessary sensor systems, control loops similar<br />

to the discussed reflexes can prevent stumbling and stabilize walking.<br />

The reflexes must be stimulated or modulated depending on the<br />

current phase <strong>of</strong> the walking cycle.<br />

3.2.4 Postural <strong>Control</strong><br />

While the reflexes discussed in the last section are mostly <strong>of</strong> the monosynaptic or spinal<br />

type, whole body posture control requires additional information. Supraspinal reflexes or<br />

high-level actions from the motor cortex are necessary to balance the upper body or to<br />

sensibly react to slope changes or lateral body tilts. Postural control must be considered<br />

to be a complex skill rather than a pure combination <strong>of</strong> simple reflexes [Horak 06].<br />

Two thirds <strong>of</strong> the human mass are located at two thirds <strong>of</strong> total height above the ground,<br />

aggregated in the head, arm, and trunk segments. Considering that during normal walking<br />

the trunk shows only an angular displacement <strong>of</strong> about 1-2 ◦ in the plane <strong>of</strong> progression, the<br />

accomplishment<strong>of</strong>thecentralnervoussystemtokeepthisbalanceissubstantial[Winter 95].<br />

Sensory information <strong>of</strong> the vision, the vestibular, and the somatosensory system is necessary.<br />

Winter reviews research on the postural control during quiet standing and normal walking<br />

[Winter 95]. A kinetic analysis <strong>of</strong> the head, trunk, and pelvis movements during<br />

walking show an increasing stabilization in the horizontal plane towards the head. A look<br />

at emg measurements explains this observation as paraspinal muscle activation travels<br />

from the upper neck downwards and reaches the lumbar levels with about 60ms <strong>of</strong>fset.<br />

This suggest a top-down anticipatory posture control <strong>of</strong> the spine.<br />

The main work in sagittal balance control <strong>of</strong> the upper body is raised by the hip extensors<br />

and flexors. Analyzing the dynamics <strong>of</strong> the trunk reveals significant unbalancing moments<br />

about the hip axis <strong>of</strong> about 40Nm. The upper body’s inertia causes a flexion during<br />

weight acceptance and an extension during push-<strong>of</strong>f. The unbalancing moment is nearly<br />

completely compensated by hip muscle action <strong>of</strong> the stance limb, which explains the small


3.2. Normal <strong>Walking</strong> in <strong>Human</strong>s 51<br />

Figure 3.22: The extrapolated center <strong>of</strong> mass (XcoM) lies just within the center <strong>of</strong> pressure<br />

during single support in walking. From [H<strong>of</strong> 07], p253.<br />

changes in trunk angle during walking. But this muscle response occurs practically at the<br />

same moment as head acceleration is measured by the vestibular organ. Thus, considering<br />

the latency <strong>of</strong> the neuromuscular system, this sensory information cannot account for the<br />

full compensation moment. Rather, it suggest that an anticipatory component is involved.<br />

Besides the control <strong>of</strong> the upper body posture in the sagittal plane, the balance system has<br />

to prevent the supporting leg from collapsing. The unbalancing torques <strong>of</strong> the trunk as just<br />

described result in a high variability <strong>of</strong> the hip muscle action, which again influences the<br />

knee and ankle joint and could result in limb instability. However, it can be observed that<br />

the knee joint reacts with additional flexion to extra hip extension and vice versa. This can<br />

partly be explained by biarticular muscles, for instance the rectus femoris muscle causes<br />

simultaneous hip flexion and knee extension. Still the nervous system must contribute to<br />

this fact, probably by corresponding muscle synergies.<br />

Considering the balance in the frontal plane, a single inverted pendulum behavior can be<br />

assumed as the hip ad- and abductors keep the body nearly rigid. These muscles have to<br />

compensate the torque induced by gravity as the CoM always lies medial <strong>of</strong> the subtalar<br />

joint acting as rotation axis in the foot. Further torques are generated by the inertia <strong>of</strong><br />

the trunk at touchdown. Similar torques act at the subtalar joint, which itself can only<br />

produce small correctional torques due to the short support leverage. Hence disturbances<br />

in the frontal plane must be compensated by lateral adaption <strong>of</strong> foot placement that has<br />

significant impact on the torques induced by gravity.<br />

H<strong>of</strong> et al. also investigate stabilization in the frontal plane during walking [H<strong>of</strong> 05, H<strong>of</strong> 07].<br />

They suggest to consider not only the center <strong>of</strong> mass but also an extended version <strong>of</strong> it.<br />

This extrapolated center <strong>of</strong> mass (XcoM) in the frontal plane is defined as<br />

XcoM = CoMz + ˙ CoMz<br />

, ω0 =<br />

ω0<br />

� g<br />

h<br />

(3.2)<br />

CoMx denotes the lateral position <strong>of</strong> the center <strong>of</strong> mass, ω0 the eigen frequency corresponding<br />

to an inverted pendulum model, g the gravity, and h the vertical height <strong>of</strong> the


52 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

center <strong>of</strong> mass. Figure 3.22 illustrates the XcoM during several steps <strong>of</strong> normal walking.<br />

It can be observed that the center <strong>of</strong> pressure which is closely related to the foot position<br />

during single support is always displaced by a small distance lateral <strong>of</strong> the XcoM. This<br />

could imply a simple strategy <strong>of</strong> balance control in walking where foot placement is chosen<br />

based on the XcoM just before heel strike [H<strong>of</strong> 08].<br />

Bauby and Kuo analyze balance control based on a model <strong>of</strong> a passive dynamic walking<br />

(see also Section 3.4.1) [Bauby 00]. They find that passive control in the sagittal plane is<br />

sufficient to achieve stability. However, in the lateral direction, active control by feedback<br />

seems necessary. This control most <strong>like</strong>ly involves visual and vestibular sensory information,<br />

and thus would be originated from higher centers <strong>of</strong> the nervous system, <strong>like</strong> the brain stem<br />

or cerebellum. This assumption is supported by experimental results, as a restriction <strong>of</strong><br />

sensory information <strong>like</strong> walking with eyes closed leads to a significant higher variability in<br />

lateral foot placement, but only small changes in the fore-aft direction. Similar indications<br />

are given by experiments comparing normal walking with an externally lateral stabilized<br />

gait [Donelan 04].<br />

The influence <strong>of</strong> vestibular and visual sensory information is examined in several research<br />

works. The role <strong>of</strong> the vestibular system is analyzed by Bent et al. by phase dependent<br />

galvanic vestibular stimulation [Bent 04]. The strongest changes in foot placement result<br />

from stimulation during the double support phase. This suggests that the decision for limb<br />

positioning is already taken when the foot leaves the ground. Furthermore the information<br />

from the vestibular system must be integrated with somatosensory information to gain an<br />

appropriate representation <strong>of</strong> the body’s pose. During double support the somatosensory<br />

information is more reliable. In contrast, the vestibular sensory information seems to be<br />

used independently during the whole walking cycle for stabilizing the head and trunk.<br />

The influence <strong>of</strong> vision on human locomotion has also been examined by confronting<br />

walking subjects with tunnel-<strong>like</strong> optical flow patterns [Prokop 97]. An effect on stride<br />

length and walking velocity can be observed. The body pose estimation is also integrating<br />

information from the vision system [Patla 97]. This integration depends more strongly on<br />

visual input but can dynamically be adapted based on available perception [Deshpande 05].<br />

Conflicting information from different sensory organs can cause instability or nausea.<br />

As already indicated above, it should again be mentioned that postural control is not<br />

only achieved by purely reactive control components. Rather, the central nervous system<br />

seems to coordinate posture and voluntary movement by anticipatory muscle patterns.<br />

Gahery [Gahery 87] classifies the postural adjustments in three strategies: postural preparations,<br />

postural accompaniments, and postural reactions. Sequences <strong>of</strong> postural muscle<br />

activity can be observed prior to activation <strong>of</strong> focal muscles. Frank et al. [Frank 90] discuss<br />

these postural accompaniments in different motor task, e.g. during elbow flexion-extension<br />

or rising to toes. The postural regulations serve to displace the body’s center <strong>of</strong> mass,<br />

to generate an opposing displacement <strong>of</strong> the center <strong>of</strong> mass, or to position the center <strong>of</strong><br />

mass over a new base <strong>of</strong> support. For example when rising to toes, healthy subjects move<br />

the center <strong>of</strong> mass to the front about 60ms before activation <strong>of</strong> the soleus and medial<br />

gastrocnemius. When pulling a stiff handle, a preceding movement <strong>of</strong> center-<strong>of</strong>-pressure<br />

(cop) and center-<strong>of</strong>-mass (com) counteracting reactive forces can be observed. As different<br />

conditions modulate timing and gain <strong>of</strong> the regulations, Frank et al. argue that the cns<br />

must have learned a simplified but state-dependent model <strong>of</strong> body dynamics to predict the<br />

expected shift <strong>of</strong> mass. The necessary control for this task could be simplified by only using


3.3. Key Aspects <strong>of</strong> Biological <strong>Walking</strong> <strong>Control</strong> 53<br />

a limited set <strong>of</strong> postural synergies that are associated with different focal motor patterns.<br />

The same postural reactions could come to use in response to external perturbations.<br />

Implications for Robotics:<br />

Postural control includes maintaining body stability in the sagittal and<br />

front plane and controlling forward velocity. This is a high-level skill<br />

requiring an estimation <strong>of</strong> the robot’s pose using information on joint angles,<br />

acceleration, and velocity information from an inertial measurement<br />

unit, and optical flow from vision, if available. Adjusting the foot<br />

placement seems to have the major influence on whole body balancing.<br />

Anticipatory torque patterns seem to be necessary to compensate<br />

segment movements resulting from mass inertia during normal walking,<br />

especially in the hip and trunk joints.<br />

3.3 Key Aspects <strong>of</strong> Biological <strong>Walking</strong> <strong>Control</strong><br />

Having reviewed research on human motion control in general and walking in particular<br />

on the preceding pages, this section will shortly summarize the key aspects considered to<br />

be relevant to robotics by the author. The following section will then go into detail on<br />

control approaches for bipedal machines applying one or more <strong>of</strong> the mentioned concepts.<br />

Regarding the structural aspects <strong>of</strong> human motion control, the following significant<br />

conceptions can be identified:<br />

Functional Morphology The human morphology is highly optimized for efficient biped<br />

locomotion regarding its geometry, mass distribution, or actuation. This“intelligent<br />

mechanics”relieves the nervous system <strong>of</strong> some <strong>of</strong> the control burden.<br />

Exploitation <strong>of</strong> Inherent Dynamics The semi-flat gait as exhibited in human walking<br />

takes advantage <strong>of</strong> this functional morphology. Inertia <strong>of</strong> body segments, compliance<br />

<strong>of</strong> the muscle-tendon system, or conversion between static and kinetic energy are<br />

exploited to reduce energy consumption and control effort, and result in natural<br />

looking motions.<br />

Self-Stabilization Suited elastic properties <strong>of</strong> muscles and tendons increase stability<br />

without active control by reacting to perturbations with purposeful joint movements.<br />

This allows for slower communication and higher control delay times.<br />

Hierarchical Network Neural control exhibits a hierarchical layout spanning from nerve<br />

fibers via the spinal cord to the brain. Muscle stimulation and perception run down<br />

respectively up along this network. Nerve cells or groups <strong>of</strong> cells can be gradually<br />

stimulated or inhibited.<br />

Distribution and Spatial Relations <strong>Control</strong> units act in a distributed way and are<br />

located spatially related to their place <strong>of</strong> activity. Like the hierarchical layout, this<br />

reduces signal density and allows to cope with high complexity.


54 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

Feed-Forward Motor Patterns Activated at kinetic or kinematic events or due to<br />

learned timing, feed-forward patterns <strong>of</strong> muscle activity create coordinated synergies<br />

<strong>of</strong> movement. Spinal pattern generators manage the correct timing and sequencing<br />

<strong>of</strong> motor patterns to create the desired cyclic motion.<br />

Feedback <strong>of</strong> Varying Complexity Feedback control occurs at various levels <strong>of</strong> complexity<br />

regarding the extend <strong>of</strong> perception and deployment <strong>of</strong> muscle action. Basic<br />

reflexes only affect a single muscle group, while postural reflexes create whole body<br />

motions and rely on a multitude <strong>of</strong> sensing organs.<br />

Phase-dependent Modulation Depending on the current task or phase <strong>of</strong> motion,<br />

reflex action can be modulated, reinforced, or suppressed. Phases <strong>of</strong> locomotion are<br />

originating from spinal pattern generators being part <strong>of</strong> the hierarchical layout <strong>of</strong><br />

control.<br />

While the structural aspects just mentioned provide valuable clues for designing an<br />

appropriate architecture, the functional analysis <strong>of</strong> human walking helps to derive the<br />

actual control units generating a walking bipedal robot. The following functional elements<br />

<strong>of</strong> human gait prove to be meaningful for developing a control for dynamic walking:<br />

<strong>Control</strong> Units from Gait Analysis Especially kinetic analysis and emg recordings <strong>of</strong><br />

human gait can provide information on motor patterns stimulated during walking.<br />

While only qualitative answers can be given, gait analysis still helps in deciding<br />

which basic patterns are necessary to generate the desired movements.<br />

Five <strong>Walking</strong> Phases Analysis <strong>of</strong> emg data shows five basic components <strong>of</strong> muscle<br />

activations <strong>of</strong> which walking is composed. As these components are related to kinetic<br />

or kinematic events, a pattern generator <strong>of</strong> the walking control system can be based<br />

on these five phases. Sensor information can be used to trigger phase transitions.<br />

Bilateral Synchronization In humans, the five phases <strong>of</strong> walking for each side <strong>of</strong> the<br />

body are synchronized. This bilateral coupling can simplify control and guarantee<br />

correct coordination <strong>of</strong> the two legs.<br />

Reflexes during <strong>Walking</strong> By investigating impaired subjects or by specific inhibition<br />

<strong>of</strong> sensing organs, reflex action and its modulation during walking can be identified.<br />

Certain reflexes can be transferred to robotic systems.<br />

Postural <strong>Control</strong> is Supraspinal Postural control requires more than local reflexes.<br />

To a certain degree, estimating the body pose and its dynamics is required for<br />

posture stability. Adjustment <strong>of</strong> foot placement has strong influence on whole body<br />

balancing.<br />

It shall be understood that the features just mentioned do not exhaustively describe<br />

human walking control. On the one hand, this is due to the fact that biomechanical and<br />

neuroscientific knowledge on this problem is still incomplete. On the other hand, some<br />

aspects were deliberately omitted as it was decided not to include them in the control<br />

system developed in this work. Most prominently, this concerns the capability <strong>of</strong> learning<br />

and long-term adaptation <strong>of</strong> the human brain. This topic will be attended to in future


3.4. Biologically Inspired <strong>Control</strong> <strong>of</strong> <strong>Bipedal</strong> <strong>Robots</strong> 55<br />

(a) (b) (c) (d)<br />

Figure 3.23: Derivation <strong>of</strong> passive walking on downwards slope following McGeer’s approach:<br />

(a) simple wheel, (b) rimless wheel, (c) synthetic wheel, (d) simple walker.<br />

work. Furthermore, this thesis will not try to design a control system on neural level but<br />

rather a system based on larger control units. Consequentially, mechanism on the level <strong>of</strong><br />

nerve cells, nerve fibers, or neural oscillators were only treated on a very basic level. As the<br />

next section will show, some researchers try to approach the problem <strong>of</strong> controlling bipeds<br />

at such a fine granularity. And finally, even the parts <strong>of</strong> neurosciences and biomechanics<br />

approached above can certainly not be covered in their vast completeness on the few pages<br />

<strong>of</strong> this introduction.<br />

3.4 Biologically Inspired <strong>Control</strong> <strong>of</strong> <strong>Bipedal</strong> <strong>Robots</strong><br />

Considering the two-legged locomotion capabilities <strong>of</strong> humans, it is not surprising that<br />

robotics researchers are trying to transfer biological concepts as those just discussed to<br />

the control <strong>of</strong> bipedal robots. The remainder <strong>of</strong> this chapter reviews research efforts being<br />

motivated in one or another way by insights on natural motion control. The section is<br />

divided in two parts, one discussing research activities concerning passive control aspects<br />

<strong>like</strong> the exploitation <strong>of</strong> inherent dynamics or elasticities, the other focusing on active<br />

control concepts <strong>like</strong> reflexes or oscillators. In case a research work includes several <strong>of</strong> the<br />

characteristics introduced above, it is listed in the section discussing its most prominent<br />

feature.<br />

3.4.1 Exploitation <strong>of</strong> Inherent Dynamics and Elasticities<br />

One <strong>of</strong> the most important aspect <strong>of</strong> locomotion in biological systems is the exploitation <strong>of</strong><br />

the natural dynamics. As mentioned before, this is done in several ways, e.g. clever mass<br />

distribution, movements that follow the inertia <strong>of</strong> body segments, or energy storage and<br />

self-stabilization by elastic elements. Unfortunately, most actuated bipeds cannot benefit<br />

from these effects due to the high friction <strong>of</strong> their joints.<br />

Beginning in the 1990s, researchers tried to build machines purely relying on inherent body<br />

dynamics. These so called passive dynamic walkers have no actuators and therefore can<br />

exhibit very low joint friction. No active control is taking place, instead these machines<br />

purely rely on well-balanced mechanics and gravity.<br />

McGeer was one <strong>of</strong> the first to investigate the motion laws behind passive dynamic walkers<br />

motivated by simple ramp-walking tinkertoys [McGeer 90, McGeer 93]. He built a machine<br />

l<br />

r<br />

mH<br />

mL


56 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

(a) (b) (c) (d)<br />

Figure 3.24: Passive dynamic walking machines: (a) McGeer’s original walker, (b) Collin’s<br />

Walker, (c) Cornell Biped, (d) Meta from TU Delft.<br />

that can walk down a slope, only using the gained potential energy to compensate for<br />

energy loss (Figure 3.24a). It can be shown that self-stabilizing cycles exist for such<br />

systems, i.e. within small margins they will stay inside stable trajectories. The models for<br />

such a simple walker can be derived starting from a simple wheel as shown in Figure 3.23.<br />

A wheel can roll along a level surface without loss <strong>of</strong> energy as far as friction is neglected.<br />

Removing the rim <strong>of</strong> the wheel leaves an equally spaced set <strong>of</strong> legs dispensing energy<br />

each time the next leg hits the ground. Assuming inelastic and impulsive impacts and<br />

conservation <strong>of</strong> angular momentum, the loss <strong>of</strong> speed can be calculated: on a level surface<br />

the rimless wheel decelerates exponentially. But on a downhill slope it can regain velocity<br />

during the rotation about the stance leg.<br />

Instead <strong>of</strong> removing the rim, the wheel could be reduced to only two spokes that can<br />

vary their angle in a joint at the former axle <strong>of</strong> the wheel (Figure 3.23c). Assuming the<br />

swinging leg would clear the ground, McGeer shows that a solution to the dynamics <strong>of</strong><br />

such a system can be found so that the next swing leg arrives just in time to continue<br />

the rolling motion as next stance leg. A generalization <strong>of</strong> this walker can be found in<br />

Figure 3.23d. McGeer found stable walking cycles for this model considering different<br />

parameters <strong>like</strong> the leg length l, the foot arc radius r, the hip and leg masses mH and mL<br />

as well as <strong>of</strong>fsets <strong>of</strong> the leg mass relative to the leg.<br />

Further research on this topic addresses the extention to three dimensions, adding a passive<br />

knee for implementing ground clearance <strong>of</strong> the swing leg, comparing the motions <strong>of</strong> passive<br />

walkers to human gait, or investigating energy consumption, step length, and stability<br />

issues [Garcia 98, Kuo 99, Kuo 01, Wisse 04]. Figure 3.24 shows some <strong>of</strong> the machines<br />

that emerged from this research.<br />

Despite the elegance <strong>of</strong> its motions and the simplicity <strong>of</strong> the control, the principle <strong>of</strong><br />

passive dynamic walking has some draw-backs and limitations. As the machines are purely<br />

passive, they rely on the potential energy gained while walking down a slope. Furthermore,<br />

the design <strong>of</strong> the robot must be carefully chosen and tuned for it to be able to walk at all.<br />

And as there is no feed-back, the machine cannot react to disturbances, but only walk on


3.4. Biologically Inspired <strong>Control</strong> <strong>of</strong> <strong>Bipedal</strong> <strong>Robots</strong> 57<br />

(a) (b)<br />

Figure 3.25: Jerry Pratt’s Spring Flamingo and the finite state machine <strong>of</strong> its control system.<br />

From [Pratt 00], p68.<br />

very level ground without external forces. It is not possible to walk outside the sagittal<br />

plane as degrees <strong>of</strong> freedom and an active steering mechanism are missing.<br />

The necessity <strong>of</strong> additional energy can be solved by substituting the energy gained while<br />

walking down a slope by simple actuators [Collins 05]. The robots Mike and Denise<br />

developed at TU Delft are powered by two pneumatic actuators in the hip [Wisse 05].<br />

Electrical actuation with series elastic elements is used for hip and ankle actuation <strong>of</strong><br />

the robot Meta shown in Figure 3.24d. Only ankle actuation is used for the Cornell<br />

Biped (Figure 3.24c) having a comparable energy consumption as human walking, whereas<br />

Honda’s Asimo uses at least ten times the energy <strong>of</strong> human walking [Collins 05].<br />

Another work propagating the exploitation <strong>of</strong> inherent control and natural dynamics <strong>of</strong><br />

walking is Jerry Pratt’s control approach for the robot Spring Flamingo (Figure 3.25a).<br />

He states that bipedal walking is a difficult problem only if viewed as a dynamical system<br />

solution but simpler if viewed as a specific mechanism [Pratt 00]. Starting from simplified<br />

models for two-legged walking, intuitive control laws are derived to make sure that five<br />

conditions are met: stabilization <strong>of</strong> height, <strong>of</strong> pitch, and <strong>of</strong> speed; swing leg motion; and<br />

properly timed transition from one support leg to the next. For each leg, a finite state<br />

machine as shown in Figure 3.25b toggles between these controllers. Some <strong>of</strong> the controllers<br />

make use <strong>of</strong> the virtual model control method as described in Section 2.2.2. The approach<br />

is shown to work on the 6 degree <strong>of</strong> freedom planar walking Spring Flamingo as well as for<br />

a 12 degree <strong>of</strong> freedom machine [Pratt 09]. Both robots feature series elastic actuators with<br />

low impedance to allow for the passive dynamics to take effect and for the implementation<br />

<strong>of</strong> a torque joint control [Pratt 95a, Pratt 02].<br />

The main achievements <strong>of</strong> Pratt’s work is to successfully combine feed-back control with<br />

exploitation <strong>of</strong> passive dynamics. A very natural and energy efficient walking gait is


58 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

produced and the system remains stable against small perturbations and changes in slope.<br />

But while the principles <strong>of</strong> exploiting inherent robustness and natural dynamics appear to<br />

beuniversal, thesuggestedalgorithmsandtheirparametershavetobesignificantlychanged<br />

for a robot <strong>of</strong> different mass distribution, size, or shape. Furthermore it is not easily<br />

possible to extend the control to different modes <strong>of</strong> locomotion or adapt characteristics <strong>like</strong><br />

the step height. And for some <strong>of</strong> the controllers, modelling is necessary to a considerable<br />

amount, <strong>like</strong> calculating the Jacobian <strong>of</strong> the direct kinematic for a virtual component or<br />

partly using a dynamic model for fast leg control.<br />

Roa et al. develop simple robots using a design methodology based on models <strong>of</strong> passive<br />

walking [Roa 05]. A dynamic simulation <strong>of</strong> the preliminary mechanical design including<br />

actuation is then used to refine the contruction. The three UNROCA robots originating<br />

from this work use servo motors to track the trajectories resulting from stimulation. While<br />

this approach makes use <strong>of</strong> the system dynamics during the design process, the actual<br />

robots do not exploit it due to their stiff actuators.<br />

A different approach combining the benefits <strong>of</strong> natural system dynamics and actuated<br />

joints is proposed by Katja Mombaur et al. [Mombaur 05b]. The main idea is to use<br />

periodical feed-forward control trajectories that will generate cyclic and stable open-loop<br />

solutions. These trajectories are generated using techniques from numerical optimization<br />

and optimal control. For this approach, a dynamic rigid multibody system model <strong>of</strong><br />

the robot and its interaction with the environment has to be established and periodic<br />

constraints must be formulated. The resulting multi-phase optimal control problem can<br />

efficiently be solved by optimization techniques based on the direct boundary value problem<br />

approach using multiple shooting [Leineweber 03a, Leineweber 03b]. This technique involves<br />

a discretization <strong>of</strong> both state and control variables transforming the optimal control<br />

problem into a nonlinear programming problem (nlp) with simultaneous simulation for<br />

the evaluation <strong>of</strong> objective function and constraints. The approach does not only allow<br />

to calculate an optimal control solution, but also to optimize mechanical parameters<br />

in respect to the target functional. Open-loop stable solutions can be shown for walking<br />

[Mombaur 01], running [Mombaur 05a], and somersaults [Mombaur 05c]. It can even<br />

be shown that human-<strong>like</strong> running with a high degree <strong>of</strong> freedom model can be open-loop<br />

stable [Mombaur 07].<br />

As stability can be included as part <strong>of</strong> the target functional during the optimization process,<br />

self-stabilizing, natural behavior can be achieved. Never-the-less the approach lacks any<br />

feed-back, so robust locomotion over unknown terrain cannot be realized. Furthermore,<br />

the approach depends on a complete mathematical model <strong>of</strong> the robot and its environment<br />

and the resulting trajectories only work for this representation. This makes it difficult<br />

to transfer it to an actual robot, as the optimization process works on a simultaneous<br />

simulation that cannot be easily replaced with a real machine.<br />

As mentioned in Section 3.1.1 exploitation <strong>of</strong> elasticity results in several benefits in<br />

natural locomotion. Alexander suggests three different applications <strong>of</strong> springs in legged<br />

robotics [Alexander 90]. Pogo-<strong>like</strong> springs can store energy during impact that can be<br />

reused for propulsion; return springs can save energy during the reverse <strong>of</strong> direction in<br />

the swing leg; and elastic foot pads can reduce impacts and provide better grip on uneven<br />

terrain. Already in the 1980s Raibert made heavy use <strong>of</strong> pogo-<strong>like</strong> springs in his running<br />

and hopping machines [Raibert 86].


3.4. Biologically Inspired <strong>Control</strong> <strong>of</strong> <strong>Bipedal</strong> <strong>Robots</strong> 59<br />

(a) (b) (c) (d) (e)<br />

Figure 3.26: Bipeds featuring elastic actuation: (a) Lucy and (b) Veronica from University <strong>of</strong><br />

Brussels, (c) the biped developed by the Shadow company, (d) elastic biped developed at fzi,<br />

Germany, (e) Pneumat-BR from Hosoda Lab.<br />

Only a few research projects cover the use <strong>of</strong> elasticity during bipedal walking. The 6 degree<br />

<strong>of</strong> freedom robot Lucy as shown in Figure 3.26a uses pleated pneumatic artificial muscles as<br />

actuation that allow to control joint stiffness [Vermeulen 04, Verrelst 05, Vanderborght 07,<br />

Vanderborght 08]. The aim <strong>of</strong> the project is to combine the versatility <strong>of</strong> robots controlled<br />

by technical methods as discussed in Chapter 2 with exploitation <strong>of</strong> natural dynamics and<br />

thus reducing energy consumption. Joint trajectories are pre-generated using a dynamic<br />

model <strong>of</strong> the robot and the pneumatic muscles. zmp calculations are applied as stability<br />

criterion. A trajectory tracker then controls the pressure within the muscles to follow the<br />

desired motion. In this way walking velocities <strong>of</strong> up to 0.15 m/s could be achieved. By<br />

changing the stiffness <strong>of</strong> the joint during the trajectory tracking depending on the natural<br />

system dynamics, energy usage can be reduced. This approach has so far only be tested<br />

on a pendulum structure actuated by the artificial muscles. While this research work tries<br />

to include passive dynamics to the control, it still depends on a dynamic model <strong>of</strong> the<br />

robot and its environment.<br />

Another project<strong>of</strong> the same groupcombiningpassivedynamic walkingandelasticactuation<br />

is the robot Veronica shown in Figure 3.26b [Ham 06a]. It is powered by six maccepa<br />

(Mechanically Adjustable Compliance and <strong>Control</strong>lable Equilibrium Position Actuator)<br />

actuators in the hip, knee, and ankle. This rotational actuator features controlable stiffness<br />

and equilibrium position by combining two servo motors and a mechanical spring. The<br />

planar robot is 1m in height, weights 5.6kg and is controlled by a concept called <strong>Control</strong>led<br />

Passive <strong>Walking</strong>: the compliant actuators change the natural frequency <strong>of</strong> the mechanical<br />

system thus enabling passive dynamic walking at different velocities. The underlying<br />

controller divides the walking cycle into six phases. For each phase, a set <strong>of</strong> joint control<br />

parameters is determined for the hip, knee, and ankle joints. For instance, the hip joint’s<br />

equilibrium position is either set to the flexed or extended postion, and the controller is<br />

configured to be stiff or compliant. Supported by a human operator, several walking steps<br />

could be achieved. While the actuator has interesting properties, it cannot produce high<br />

stiffness or high torque by design. The control concept lacks feedback mechanisms to react<br />

to disturbances or unpredicted ground conditions.


60 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

At the Locomotion Laboratory in Jena, Seyfarth and his colleagues also investigate the role<br />

<strong>of</strong> compliant actuators on bipedal walking and running [Geyer 04,Iida 07,Blum 09]. In the<br />

scope <strong>of</strong> this research, several basic planar robots with only a few active degrees <strong>of</strong> freedom<br />

are built to verify the models derived from biomechanical analysis. The underactuated<br />

robot Jena Walker II is used to examine the function <strong>of</strong> the elastic muscle-tendon system<br />

during walking [Smith 07]. It features two actuated hip joints and passive knee and ankle<br />

joints. Four major muscle-tendon groups are represented by a spring-cable system, three <strong>of</strong><br />

them spanning two joints. Stable, supported toe-walking can be demonstrated at different<br />

velocities up to 0.6 m/s showing leg compliance via the ankle joint even during stance with<br />

a stretched leg.<br />

Since the pioneer work <strong>of</strong> Kato in 1969, several projects use pneumatic artificial muscles as<br />

actuation for biped, e.g. the biped by the Shadow company in Figure 3.26c, but only a few<br />

machines are actually able to walk, let alone make use <strong>of</strong> the elastic properties. Kerscher et<br />

al. are designing a compliant 3D bipedal robot as shown in Figure 3.26d [Kerscher 07]. The<br />

approach combines electrical actuation in the hip joint with pneumatic muscles produced by<br />

the company Festo in the knee and ankle joints. The control concept combines the Virtual<br />

Model <strong>Control</strong> as describes in Section 2.2.2 with adaptation <strong>of</strong> the stiffness. A detailed<br />

model <strong>of</strong> the pneumatic muscle is developed by using e.g. a quick release experiment to<br />

derive dynamic properties [Kerscher 06]. Based on dynamic simulation <strong>of</strong> the robot and<br />

the model <strong>of</strong> the muscle, first walking steps could be achieved.<br />

Hosoda et al. have built several robots for walking and jumping actuated by McKibben<br />

pneumatic artificial muscles [Hosoda 07a, Hosoda 07b]. The robot Pneumat-BR shown in<br />

Figure 3.26e is able to jump as well as walk in the sagittal plane. The main focus <strong>of</strong> this<br />

work lies in the adjustability <strong>of</strong> the joint compliance depending on the mode <strong>of</strong> locomotion.<br />

The control system consists <strong>of</strong> a simple feed-forward mechanism opening or closing the<br />

valves in a fixed sequence for an experimentally determined time. While indeed passive<br />

walking-<strong>like</strong> locomotion could be achieved, the robot lacks any feedback system and can<br />

only walk on flat terrain without disturbances.<br />

3.4.2 Neuro-, Reflex-, and Oscillator-based <strong>Control</strong><br />

Several research projects developing control approaches for bipedal walking based on<br />

neuroscientific findings can be found in literature. As presented earlier in this chapter,<br />

possible features to be considered are the functionality <strong>of</strong> the neurons, central pattern<br />

generators, reflexes, or muscle synergies. In the following, research work involving one or<br />

several <strong>of</strong> these features will be discussed.<br />

Paul demonstrates that a purely reactive sensorimotor neural network can produce a stable<br />

walking gait for a simulated 8 degrees <strong>of</strong> freedom lower body biped [Paul 05]. Instead<br />

<strong>of</strong> relying on a central pattern generator, simple neural networks directly connect sensor<br />

inputs <strong>like</strong> waist orientation or contact information to actuator control signals. A genetic<br />

algorithm is used to evolve the controllers, resulting in a rhythmic, cyclic gait behavior. It<br />

is shown that the cutaneous sensor information is sufficient to establish control networks<br />

that produce a walking limit cycle, while joint angle and orientation sensing seems to<br />

be less important. It is assumed that a central pattern generator could mainly control<br />

upper body movement to adjust e.g. the step frequency. While the results underline<br />

the importance <strong>of</strong> reflexes and the possibility <strong>of</strong> reactive switching <strong>of</strong> walking phases,


3.4. Biologically Inspired <strong>Control</strong> <strong>of</strong> <strong>Bipedal</strong> <strong>Robots</strong> 61<br />

(a) (b) (c) (d)<br />

Figure 3.27: (a) HOAP-1 and (b) HOAP-3 by Fujitsu, (c) RunBot from bccn Göttingen, (d)<br />

small biped from Nakamura Lab.<br />

it remains to be seen whether this simple control approach would still work on a more<br />

complex machine and in a disturbed environment.<br />

Zaier et al. investigate the combination <strong>of</strong> reflexes with a cpg [Zaier 07, Zaier 08]. On<br />

the small humanoid HOAP-3 (Figure 3.27b) they implement a piecewise linear pattern<br />

generator using gyro sensor information for stabilization during single support. A recurrent<br />

neural network is triggered by sudden sensor events to recognize the disturbance and<br />

initiate a reflex motion. The motion can be adapted on the basis <strong>of</strong> sensor information: in<br />

case the robot steps on an obstacle, its position on the foot sole is calculated using four<br />

force sensors in the foot. Then a roll or pitch motion in the ankle tries to increase the<br />

size <strong>of</strong> the support polygon in order to relocate the current zmp in a stable region. This<br />

approach combines generated trajectories with reflexes. The pattern generation clearly<br />

benefits from the large foot plates and the low weight <strong>of</strong> the robot.<br />

Geng et al. implement a reflexive neural network for a small planar walker [Geng 06,<br />

Manoonpong 07]. They show that fast walking is possible without planing <strong>of</strong> trajectories<br />

but rather by using local reflexes and by exploiting the passive dynamics <strong>of</strong> the mechanical<br />

system. The robot RunBot as shown in Figure 3.27c is 23cm in height, has curved feet<br />

measuring ground contact, and is actuated in hip and knees by four modified servo motors.<br />

An additional servo motor can move a small weight at the upper body to simulate leaning<br />

forward or backwards. Further sensors include an accelerometer and one infrared sensor<br />

pointing to the floor to distinguish between even black floor and sloped white floor. Simple<br />

neural networks locally control the leg and the upper body movement, creating reflex-<strong>like</strong><br />

actions depending on joint angles, ground contact, or acceleration. For instance, when<br />

the swing leg reaches the anterior extreme position (aep), the knee joint is extended to<br />

prepare for touchdown and weight acceptance. An additional adaptive network modifies<br />

the local networks based on the ir information to achieve a predictive change <strong>of</strong> gait on<br />

slopes. As a result, stable planar walking on flat and sloped terrain could be generated<br />

that combines exploitation <strong>of</strong> passive dynamics and active neural control. Scaled by leg<br />

length, the walking velocity is one <strong>of</strong> the fastest observed in bipeds and is comparable


62 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

Figure 3.28: Oscillator control systems as applied by Taga. From [Taga 91], p150.<br />

to fast human walking. However, it still has to be shown whether the simplicity <strong>of</strong> the<br />

control approach scales to larger machines with more degrees <strong>of</strong> freedom.<br />

A combination <strong>of</strong> feed-forward patterns and reflexes accounts for the control approach<br />

developed by Huang et al. [Huang 05]. The feed-forward joint angle trajectories are<br />

pre-calculated <strong>of</strong>fline as third-order splines using zmp as stability criterion. Here, it is<br />

assumed that the dynamics <strong>of</strong> the robot and its environment are known. Three reflexes<br />

add joint angle <strong>of</strong>fsets to the rhythmic trajectories in case disturbances are detected. The<br />

zmp reflex will shift the ankle joints if the actual zmp moves too close to the border<br />

<strong>of</strong> the foot. By considering the <strong>of</strong>fset between the estimated and effective touchdown<br />

time, the landing-phase reflex changes the leg length <strong>of</strong> the stance leg, thus compensating<br />

changes in forward velocity. Finally, the body posture reflex will adapt the hip joints if<br />

the measured body inclination varies from that <strong>of</strong> the calculated trajectory. The control<br />

approach is demonstrated in simulation and using the small biped shown in Figure 3.27d.<br />

While a combination <strong>of</strong> feed-forward and feedback control seems promising, the concept <strong>of</strong><br />

Huang et al. still requires a complete dynamics model for the <strong>of</strong>fline as well as the online<br />

calculation <strong>of</strong> the zmp. Furthermore, the use <strong>of</strong> joint angles as control variable and the<br />

stiff mechanics forbid the use <strong>of</strong> passive dynamics.<br />

Taga was one <strong>of</strong> the first to control biped locomotion by neural oscillators [Taga 91,<br />

Taga 95b, Taga 95a]. In a simulated five and eight link planar biped, neural rhythm<br />

generators similar to those proposed by Matsuoka control each hip, knee, and ankle<br />

(Figure 3.28) [Matsuoka 87]. As passive dynamics are allowed to act, also mechanical<br />

oscillation <strong>like</strong> the pendulum motions <strong>of</strong> the limb is included. Inertial joint angles, joint<br />

velocities, and foot contact information are used in the feedback terms <strong>of</strong> the individual


3.4. Biologically Inspired <strong>Control</strong> <strong>of</strong> <strong>Bipedal</strong> <strong>Robots</strong> 63<br />

Figure 3.29: Hierarchical oscillator control systems as applied by Shan. From [Shan 02].<br />

oscillators. The relation between the oscillators is self-organized as they are mutually<br />

entrained and globally interact via the controlled system and the environment. Stable<br />

walking on even and sloped surfaces and under perturbations is achieved. An important<br />

result <strong>of</strong> Taga’s work is to demonstrate embodied control <strong>of</strong> a biped with rhythmic<br />

locomotion patterns emerging as result from dynamic interaction <strong>of</strong> the global entrainment<br />

between the musculoskeletal system, the neural system, and the environment. Yet it is not<br />

clear if the approach remains valid with more complex models or an actuation similar to real<br />

bipedal machines. Also it has to be decided for each oscillator which sensor systems will be<br />

used as feedback. Similar projects continued Taga’s ideas including slightly more degrees<br />

<strong>of</strong> freedom or a more sophisticated model <strong>of</strong> the neuro-muscular system [Yamazaki 96,<br />

Hase 99, Ogihara 01]. Katayama et al. study the theory <strong>of</strong> neural oscillators for the use in<br />

bipedal locomotion control and show solutions for the stability <strong>of</strong> the nonlinear differential<br />

equations and possible applications for walking pattern generation [Katayama 95].<br />

<strong>Walking</strong> control systems based on neural oscillators do not scale well with increasing<br />

degrees <strong>of</strong> freedom as their parameters do not have a physical interpretation and therefore<br />

are difficult to tune or to learn. Shan et al. approach the problem <strong>of</strong> complexity by using a<br />

hierarchical system <strong>of</strong> cpg sub circuits [Shan 02]. Each <strong>of</strong> these circuits is implemented as<br />

recurrent neural network and represent a certain motor program for walking (Figure 3.29).<br />

Represented by differential equations <strong>of</strong> different order, the oscillator sub circuits create a<br />

lateral roll movement, swing leg lifting, and forward/backward motion. The parameter<br />

space is reduced by inverse kinematics calculations. The control system is tested on the<br />

small biped robot HOAP-1 as shown in Figure 3.27a featuring 20 degrees <strong>of</strong> freedom, force<br />

measurement in the feet, and a three dimensional gyroscope. <strong>Walking</strong> and simple stair<br />

climbing could be achieved. However, as joint position trajectories <strong>of</strong> sine-<strong>like</strong> patterns are


64 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

Figure 3.30: Oscillator control system as applied by Aoi et al. From [Aoi 05], p222.<br />

forced on the robots joint, no inherent dynamics can be exploited and only highly tuned<br />

parameters allow the system to walk in an undisturbed environment.<br />

Aoi and his colleagues also control a HOAP-1 robot by a network <strong>of</strong> nonlinear oscillators<br />

[Tsuchiya 03, Aoi 05]. One oscillator for each foot, each leg, and the trunk are<br />

coordinated by an inter-oscillator (Figure 3.30). Together these units form the rhythm<br />

generator <strong>of</strong> the control system. Their phases encode the joint trajectories for the limbs.<br />

Again, inverse kinematics is used to calculate joint positions resulting in ellipsoid trajectories<br />

<strong>of</strong> hands and feet in swing and stance phase. Inverse dynamics and numerical<br />

simulation lead to a trunk motions that allows stable walking. Stability is determined by<br />

analysis <strong>of</strong> Poincaré maps <strong>of</strong> the motion equations. The ground contact measurements<br />

within the feet provide sensory feedback that is used to reset the oscillators and modify<br />

the trajectories e.g. based on the foot position at touchdown. Experiments using the robot<br />

and a simulation environment show that the cyclic motion falls back to its limit cycles at<br />

very small disturbances but cannot cope with larger perturbations. While this approach<br />

employs oscillators at the center <strong>of</strong> its control system, it still relies on many methods <strong>of</strong><br />

technically controlled bipeds and thus inherits the corresponding drawbacks as discuss at<br />

the end <strong>of</strong> Chapter 2.<br />

The application <strong>of</strong> neural oscillators is also examined by Endo, Cheng, et al. for a basic<br />

planar walker and for a small, fully actuated biped [Endo 04, Endo 05]. The control system<br />

based on Matsuoka oscillators is set up in analogy to Kimura’s work on quadrupedal<br />

robots [Kimura 04]. While the oscillators <strong>of</strong> the planar walker directly control the joint<br />

torques, those <strong>of</strong> the control for the small humanoid act on the foot positions (Figure 3.31a).<br />

While this approach reduces the number <strong>of</strong> neurons and parameters, it makes an inverse<br />

kinematics calculation necessary and excludes the possibility to exploit system dynamics.<br />

The vertical leg movement is modulated by two postural reflexes, an extensor response,<br />

and a vestibulo-spinal-<strong>like</strong> reflex that generate feedback terms for the oscillator equations.<br />

Another feedback term synchronized the horizontal leg movement with the body sway. The<br />

approach is tested in simulation and using a QRIO robot by Sony as shown in Figure 3.31b.<br />

Slow walking under small perturbations with a step length <strong>of</strong> approximately the length <strong>of</strong><br />

the feet can be achieved. In doing so, the dynamic effects during walking do play a minor<br />

role and postural stability is easier to obtain.


3.4. Biologically Inspired <strong>Control</strong> <strong>of</strong> <strong>Bipedal</strong> <strong>Robots</strong> 65<br />

(a) (b)<br />

Figure 3.31: (a) Oscillator and reflex control systems as applied by Endo, Cheng et al.,<br />

from [Endo 05], p599. (b) The QRIO robot by Sony.<br />

The forgoing literature survey presents an overview on biological inspired approaches <strong>of</strong><br />

bipedal robots. While not being complete, it still should provide a solid view on this topic’s<br />

research projects. Table 3.4 summarizes these works regarding the application <strong>of</strong> aspects<br />

<strong>of</strong> biological motion control as stated in the previous section. The following chapters will<br />

show how these aspects are integrated in the concept suggested in this work.


66 3. <strong>Human</strong> Locomotion <strong>Control</strong><br />

Approach Details ED HC DC FP RE PG PM PC<br />

[McGeer 90, McGeer 93] planar passive walking � - - - - - - -<br />

[Collins 05, Wisse 05] actuated passive walking � - - - ◦ - - -<br />

[Pratt 00] Spring Flamingo, planar, compliant � - ◦ - - ◦ ◦ ◦<br />

[Roa 05] exploit dynamics during design ◦ - - - - - - -<br />

[Mombaur 01, Mombaur 05b] optimal control, simulation ◦ - - � - - - -<br />

[Vermeulen 04, Vanderborght 08] Lucy, pneumatic muscles ◦ - - ◦ - - - -<br />

[Ham 06a] Veronica, maccepa, variable stiffness � - - ◦ - - - -<br />

[Geyer 04, Smith 07] Jena Walker II, passive compliance � - - ◦ - - - -<br />

[Kerscher 06, Kerscher 07] virtual model control, pneumatic muscles ◦ - ◦ - ◦ - ◦ -<br />

[Hosoda 07a, Hosoda 07b] Pneumat-BR, pneumatic muscles ◦ - - ◦ - - ◦ -<br />

[Paul 05] purely reflex-based, simulation, 8 DoF - - ◦ - � - - -<br />

[Zaier 07, Zaier 08] HOAP-3, recurrent neural network - - ◦ ◦ � � - ◦<br />

[Geng 06, Manoonpong 07] RunBot, planar, neural network � ◦ ◦ - � ◦ - ◦<br />

[Huang 05] small servo biped, zmp - - - � � ◦ - �<br />

[Taga 91, Taga 95b, Taga 95a] simulation, 5/8 DoF, neural oscillators ◦ ◦ � - ◦ � - -<br />

[Shan 02] HOAP-1, hierarchy <strong>of</strong> cpgs, inv. kinematics - � - ◦ ◦ � - -<br />

[Tsuchiya 03, Aoi 05] HOAP-1, neural osc., inv. kinematics, <strong>of</strong>fline - ◦ - - ◦ � - -<br />

[Endo 04, Endo 05] QRIO, neural osc., inv. kinematics - - - - � � - -<br />

this work simulation, 21 DoF, compliance � � ◦ � � � � �<br />

ED: Exploit Inh. Dynamics RE: Reflexes, Feedback <strong>Control</strong><br />

HC: Hierarchical <strong>Control</strong> PG: Central Pattern Generators<br />

DC: Distrib. <strong>Control</strong> Units PM: Phase-dep. Stim. & Modul.<br />

FP: F.-Forw. Motor Patterns PC: High-Level Posture <strong>Control</strong><br />

Table 3.4: <strong>Control</strong> approaches for bipedal robots adapting ideas from biological motion control.


4. A Biologically Supported Concept<br />

for <strong>Control</strong>ling Bipeds<br />

In the previous chapter, key findings <strong>of</strong> biological motion control have been suggested to be<br />

considered worthwhile for robotic bipeds. They concern not only the control system itself<br />

in form <strong>of</strong> the central nervous system. Rather, they affect all aspects <strong>of</strong> the human body,<br />

its morphology, physiology, muscles, and receptors. Even more importantly, these aspects<br />

are interleaved as one cannot function without the other. This leads to a holistic approach<br />

for designing bionical bipeds or other robots strongly based on a biological example. As<br />

illustrated in Figure 4.1, all facets <strong>of</strong> the robot’s design can adopt or transfer ideas from<br />

biology. However, this thesis is focused on the control system for biped locomotion. And<br />

while solutions for the other problems <strong>like</strong> the mechatronics or muscle-<strong>like</strong> actuators<br />

cannot be provided, these topics must be covered nevertheless. A biological inspired<br />

control method as the one suggested here could not function without similar biological<br />

characteristics <strong>of</strong> the mechanical system. In order to test the control concept without the<br />

need <strong>of</strong> developing a full bipedal robot, a suitable simulation is designed (see Chapter 6),<br />

while keeping in mind the requirements for a real robot.<br />

This chapter introduces a methodology for controlling the locomotion <strong>of</strong> complex bipeds<br />

based on above findings from neurosciences and biomechanics. As the approach described<br />

here partly relies on passive dynamics, the following section describes passive control<br />

aspects. These lead to several requirements for the structure <strong>of</strong> the active control system<br />

that will be introduced in the subsequent section. Classes <strong>of</strong> control units that can be<br />

derived from their natural complements will be presented. Thereupon, the hierarchical<br />

layout and the control flow within the emerging network is described. Preliminary versions<br />

<strong>of</strong> this approach have been published in [Luksch 08a, Luksch 08b].<br />

4.1 Passive <strong>Control</strong> Aspects<br />

A strong advantage <strong>of</strong> natural locomotion control is the passive exploitation <strong>of</strong> mechanical<br />

properties. This includes the functional morphology <strong>of</strong> the human body, the inherent<br />

system dynamics, and self-stabilization by elastic elements. All these effects do not require


68 4. A Biologically Supported Concept for <strong>Control</strong>ling Bipeds<br />

Morphology<br />

Kinematical<br />

Layout<br />

Muscles<br />

Actuators<br />

Biology<br />

Bionics<br />

Robotics<br />

Receptors<br />

Sensors<br />

Nervous<br />

System<br />

<strong>Control</strong><br />

System<br />

Figure 4.1: Bionics as guiding principle for the design <strong>of</strong> bipedal robots. This thesis focuses on<br />

the derivation <strong>of</strong> the control system.<br />

active control or feedback. The presented approach tries to make use <strong>of</strong> these benefits.<br />

Naturally, this mainly affects the mechanical properties <strong>of</strong> the simulated or real robot.<br />

However, it also imposes demands on various levels <strong>of</strong> the active control system.<br />

Regarding the mechanical setup, beneficial passive effects can only emerge when certain<br />

requirementsaremet. Themorphology<strong>of</strong>therobotmustbedesignedtosupportthedesired<br />

forms <strong>of</strong> locomotion. This necessitates proper segment lengths and mass distribution.<br />

Besides manually tuning these parameters, a common procedure in the development <strong>of</strong><br />

passive walkers, two further strategies come to mind: as human morphology has been<br />

functionally optimized during millions <strong>of</strong> years <strong>of</strong> evolution, mimicking the masses and<br />

segment lengths <strong>of</strong> human trunk and limbs would be intelligible. This approach is taken<br />

in this work, as a simulated biped is not restricted by constructive constraints. The<br />

second strategy would make use <strong>of</strong> numerical optimization techniques to find a suitable<br />

robot design while taking the target motions and constraints <strong>of</strong> the construction into<br />

account. This idea was also investigates by the author but is not a substantial part <strong>of</strong> this<br />

work [Luksch 05, Luksch 07].<br />

Besides the robot’s morphology, the actuation system must also be suited for passive<br />

control. Especially high joint friction and damping can prevent exploitation <strong>of</strong> inherent<br />

dynamics. In this work, a compliant joint actuator featuring some <strong>of</strong> the characteristics<br />

found in the human muscle-tendon system is assumed. It allows for direct torque control<br />

and position control with adjustable compliance as well as for parallel elastic elements.<br />

This elasticity makes self-stabilization possible. Details on this topic can be found in<br />

Chapter 6.<br />

Finally, the active control system itself must be designed in such a way it does not inhibit<br />

the positive effects <strong>of</strong> passive dynamics. At no point, full body joint position trajectories<br />

should be used. Even if generated using a dynamics model, they can never fully match<br />

actual body motions in a partly unknown environment. Rather, torque control is to be<br />

preferred to position control whenever possible, and control units should act locally and<br />

be distributed. If position control is inevitable, at least a minimum compliance should<br />

remain, and at no time position control should affect more then only a few joints. Active


4.2. <strong>Control</strong> Unit Classes 69<br />

control only occurs to steer the limbs towards the deliberate motions and during reflexive<br />

action. The next section will introduce control units implementing these ideas.<br />

4.2 <strong>Control</strong> Unit Classes<br />

In human motion control, the basic control unit is the neuron. Each neuron collects<br />

incoming activity and fires itself if a certain threshold is reached. All data processing<br />

and muscle control is based on this simple unit. Since decades, researchers are designing<br />

artificial neural network to capture certain properties <strong>of</strong> their natural examples. While<br />

the idea <strong>of</strong> building a control system <strong>of</strong> simple, uniform units is intriguing, the concept<br />

described here chooses a structural abstraction above the neural level. Following the<br />

philosophy <strong>of</strong> behavior-based robotics, each control unit already represents a self-contained<br />

functional unit. Even though the approach is probably not as elegant as artificial neural<br />

networks, it yet shows several advantages: the human nervous system has about 100 billions<br />

<strong>of</strong> neuron, some <strong>of</strong> them handling thousands <strong>of</strong> incoming signal. A robot walking control<br />

system will be much smaller, but still the complexity <strong>of</strong> the network and its communication<br />

would be very high and increasing in size with each new feature. This complexity as well<br />

as the fact that individual neurons and signal have no semantics make it hard to design,<br />

interpret, and debug such a control system. On the other hand, if each control unit has its<br />

own function, these tasks become easier to manage. Furthermore, the parameters <strong>of</strong> the<br />

units have a physical meaning, facilitating the tuning and error handling <strong>of</strong> the control<br />

system.<br />

Nevertheless, some properties <strong>of</strong> neurons can be transferred to control units. A neuron can<br />

only be active if stimulated, and its activation can be decreased or suppressed by inhibiting<br />

signals. This idea <strong>of</strong> gradual stimulation or inhibition is applied to the control units at<br />

hand. The flow <strong>of</strong> activity will indicate which parts <strong>of</strong> the network will be responsible for<br />

the control at a certain point in time. Section 5.1.3 will describe how these features are<br />

implemented within the behavior-based architecture deployed in this work.<br />

The previous chapter, in particular Section 3.1, presented the control units found in human<br />

locomotion control. In nature the distinctions between the classes <strong>of</strong> control units are not<br />

sharply defined. Some might not even exist as a self-contained group <strong>of</strong> neurons, but only<br />

emerge as a unit from consistant effects observed during gait analysis. Still it remains<br />

valid to derive control units for technical systems from these biological finding as they<br />

clearly exist as a functional entity. Thus the following section lists the unit classes <strong>of</strong> the<br />

suggested control concept, succeeded by the descriptions <strong>of</strong> the hierarchical network in<br />

which the units are arranged. For each class, a symbolic representation is introduced which<br />

will be used throughout the remaining chapters <strong>of</strong> this text.<br />

4.2.1 Locomotion Modes<br />

In humans and animals, the decision on which form <strong>of</strong> locomotion is applied in a certain<br />

situation is taken by the brain. The control system in this work leaves this choice to the<br />

operator <strong>of</strong> the bipedal robot. Thus, the highest-level control units to be considered are<br />

the locomotion modes:<br />

Locomotion<br />

Mode


70 4. A Biologically Supported Concept for <strong>Control</strong>ling Bipeds<br />

Examples for locomotion modes could be standing, walking, or running. Selected by the<br />

operator, these units stimulate the corresponding parts <strong>of</strong> the control network. Inputs for<br />

the modulation <strong>of</strong> the locomotion mode could include the desired velocity or direction <strong>of</strong><br />

movement. Locomotion modes directly stimulate the pattern generators introduced next,<br />

but can also modulate control units on a lower level.<br />

4.2.2 Spinal Pattern Generators<br />

Section 3.1.4 showed that muscle activation seems to be coordinated with the phases <strong>of</strong><br />

the walking cycle. This and other evidence brought forth in Section 3.2.2 indicate the<br />

existence <strong>of</strong> a central unit generating rhythmic patterns <strong>of</strong> muscle action. It is commonly<br />

believed that these pattern generators are located within the spinal cord. Consequently,<br />

the corresponding control unit is labeled as spinal pattern generator (spg):<br />

SPG<br />

While many research works model central pattern generators as neural oscillator, another<br />

approach is followed here. spgs are represented as finite state machines, where state<br />

transitions are triggered by sensory events. The drawback <strong>of</strong> this approach is the absence<br />

<strong>of</strong> a strong timing to create periodic movement. Here, the periodicity emerges from the<br />

robots interaction with the environment. As an advantage, the walking phases are always<br />

synchronized to the corresponding kinematic or kinetic events. This can result in higher<br />

robustness <strong>of</strong> the system. Furthermore, this approach is beneficial for the exploitation <strong>of</strong><br />

inherent dynamics as less timing is forced on the motion <strong>of</strong> the segment masses.<br />

4.2.3 Motion Phases<br />

The spgs could directly stimulate all the lower control units working in one phase <strong>of</strong><br />

locomotion. But to support a clear arrangement and to further follow the natural example<br />

<strong>of</strong> hierarchical layout, for each phase its own control unit is introduced:<br />

Motion<br />

Phase<br />

In general, each motion phase corresponds to a section <strong>of</strong> the cyclic locomotion mode<br />

between significant kinetic or kinematic events. Within each phase unit it is coded to<br />

what extend the lower level control units are stimulated or inhibited. Modulation <strong>of</strong> reflex<br />

action is also possible. Furthermore, motion phases set the stiffness and equilibrium point<br />

<strong>of</strong> passive joints.<br />

4.2.4 Motor Patterns<br />

Bizzi, Ivanenko and others showed that stimulation <strong>of</strong> certain regions in the spinal cord<br />

result in muscle action producing coordinated joint or limb motions (Section 3.1.4). These<br />

components or motion primitives seem also to be recruited in phases <strong>of</strong> locomotion. The<br />

corresponding control unit <strong>of</strong> this concept is called motor pattern:


4.2. <strong>Control</strong> Unit Classes 71<br />

Motor<br />

Pattern<br />

Similar to muscle activity measured with emg, these units produce uniform patterns <strong>of</strong><br />

torque for one or more joints in a feed-forward manner. Only a minimum number <strong>of</strong><br />

parameters describe the characteristics <strong>of</strong> the torque function. Motor patterns always act<br />

locally, i.e. they will not affect joints that are located far apart (e.g. a joint <strong>of</strong> an arm and<br />

a joint in one <strong>of</strong> the feet). Should a more global synergy be necessary, it must be provided<br />

at a higher level. For instance a motion phase can stimulate several motor patterns at<br />

a time. While motor patterns are designed without feedback, fortitude and duration <strong>of</strong><br />

the torque output can vary depending on the modulation by motion phases or locomotion<br />

modes. Motor patterns support or shape the passive system dynamics to create the desired<br />

motion. Still, as many joints as possible should act in a purely passive or compliant way.<br />

4.2.5 Local Reflexes<br />

Reflex action provides the basic feedback mechanism in human motion control. It can occur<br />

on various levels <strong>of</strong> complexity, ranging frommonosynaptic and propriospinalto supraspinal<br />

depending on the area <strong>of</strong> the spinal cord and the brain they cover (Section 3.1.3). <strong>Control</strong><br />

units corresponding to the first two types, i.e. those affecting only a few, spatially adjoining<br />

joint, are called local reflexes:<br />

Local<br />

Reflex<br />

Local reflexes show a tight coupling between sensor information and motor action. The<br />

control answer can have different forms: some reflexes show a loopback controller-<strong>like</strong><br />

behavior with a linear or nonlinear relation between sensor data and control output. The<br />

other type <strong>of</strong> reflexes works event-based: as soon as a certain sensor event occurs, the<br />

output state is changed.<br />

4.2.6 Postural Reflexes<br />

As in the natural model, reflexes vary in their spatial and semantic expansion. Most reflexes<br />

regulating the body’s posture during locomotion require a more complex mechanism than<br />

local loopback control. They belong to the class <strong>of</strong> supraspinal reflexes as their neural<br />

activity is not restricted to an enclosed part <strong>of</strong> the spinal cord, but also include the brain.<br />

The analogues to these units are called postural reflexes:<br />

Postural<br />

Reflex<br />

The main characteristics <strong>of</strong> this type <strong>of</strong> control units is the extend <strong>of</strong> their action and<br />

the complexity <strong>of</strong> their internal model: the reflex action at one place <strong>of</strong> the robot can<br />

be the result <strong>of</strong> a sensor event at the opposite end <strong>of</strong> the machine, possibly supported<br />

by a simplified dynamical model. The state variables most frequently used or calculated


72 4. A Biologically Supported Concept for <strong>Control</strong>ling Bipeds<br />

‘Brain’ ‘Spinal Cord’<br />

Locomotion<br />

Modes<br />

Sense <strong>of</strong><br />

Balance<br />

stimulation stimulation<br />

SPGs<br />

Mechatronics<br />

Vestibular<br />

System<br />

(IMU)<br />

modulation signals<br />

Somatosensory<br />

System<br />

Motion<br />

Phases<br />

Environment<br />

Postural<br />

Reflexes<br />

torques, positions, forces<br />

Joint<br />

<strong>Control</strong><br />

Mechanical <strong>Control</strong> (Elasticities, Inerita, ...)<br />

Joint Groups<br />

Local<br />

Reflexes<br />

Motor<br />

Patterns<br />

Sensor<br />

Handling<br />

torque, position,<br />

stiffness<br />

Figure 4.2: The hierarchical organization <strong>of</strong> control units and the interaction <strong>of</strong> the control<br />

system, the robot’s mechatronics and the environment.<br />

by postural reflexes are estimations <strong>of</strong> the upper body orientation, the rough position or<br />

velocity <strong>of</strong> the center <strong>of</strong> mass, or load distribution in the feet. While this information<br />

partly requires some kind <strong>of</strong> reduced model <strong>of</strong> the system dynamics, it can be kept simple<br />

enough to avoid the break <strong>of</strong> the control loop by intensive processing operations. This<br />

is possible because an approximation <strong>of</strong> a single or only a few variables <strong>of</strong> the dynamic<br />

system is sufficient. In collaboration with the local reflexes, the postural reflexes enhance<br />

the global stability <strong>of</strong> the biped.<br />

4.3 Hierarchical Layout<br />

Having defined the basic types <strong>of</strong> control units, the questions remains on how to organize<br />

them as a control network. Section 3.1.4 stated that natural motion control has a<br />

hierarchical layout. As suggested by Ivanenko et al. in the hypothetical scheme shown<br />

in Figure 3.9, stimulation <strong>of</strong> control units runs downwards from higher to lower levels.<br />

Perceptional information feeds back from sensory neurons to various stages <strong>of</strong> the control<br />

network. Figure 4.2 illustrates how these ideas and other conceptions from the previous<br />

chapter are translated to a control concept for bipeds. It also shows how the control system<br />

is integrated with the robot’s mechatronics.<br />

The hierarchy <strong>of</strong> control units is laid out from left to right. From the highest level, i.e. the<br />

locomotion modes, stimulation and modulation signals run along the control units until<br />

motor patterns and local reflexes are reached on the lowest level. As the last-mentioned<br />

units only act locally, they can be pooled in so called joint groups. This corresponds to the


4.3. Hierarchical Layout 73<br />

Locomotion<br />

Mode 1<br />

Locomotion<br />

Mode n<br />

SPG 1<br />

SPG n<br />

1<br />

2<br />

modulation<br />

Motion<br />

Phase 1<br />

Motion<br />

Phase 2<br />

Motion<br />

m Phase m<br />

;<br />

Motion<br />

Phases<br />

Postural<br />

Reflex 1<br />

Postural<br />

Reflex 2<br />

Postural<br />

Reflex l<br />

Joint Group 1<br />

Local<br />

Reflexes<br />

Motor<br />

Patterns<br />

Local<br />

Reflexes<br />

Motor<br />

Patterns<br />

Joint Group q<br />

Figure 4.3: Flow <strong>of</strong> stimulation and modulation signals within the hierarchy <strong>of</strong> control units.<br />

areas <strong>of</strong> the spinal cord where sensor and motor neurons are spatially related depending<br />

on their dedicated region <strong>of</strong> the body (Section 3.1.2 and Figure 3.8). For the biped in<br />

this work, the joints and sensors <strong>of</strong> the upper trunk, the lower trunk, and each leg are<br />

subsumed in such a group. This distribution <strong>of</strong> control units enhances the clearness <strong>of</strong><br />

the control system and enables a possible assignment <strong>of</strong> units to dedicated, distributed<br />

processing hardware <strong>like</strong> small dsp boards [Hillenbrand 09].<br />

Similarly to the somatosensory system <strong>of</strong> the human body, information on torques and<br />

forces acting in joints and feet as well as joint positions are measured. This sensor data<br />

can be pre-processes in the corresponding joint group and forwarded to various levels <strong>of</strong><br />

control. As in human postural control, the sense <strong>of</strong> balance needs to be derived from sensor<br />

information at a higher level. Where nature uses the vestibular system, eyes, or the hairs<br />

on the skin as source <strong>of</strong> perception, a robotic system deploys an inertial measurement unit<br />

(imu), or vision-based sensors. The resulting information is fed to the postural reflexes.<br />

Compared to human locomotion control, the sense <strong>of</strong> balance and the locomotion modes<br />

would be situated within the brain, the rest <strong>of</strong> the control units within the spinal cord.<br />

Reflexes and motor patterns issue control commands for the robot’s joint controller. The<br />

interface consists <strong>of</strong> torque, stiffness, and equilibrium position values. Driven by the joint<br />

controllers, the actuation system applies forces and torques to the robot’s mechanics. In<br />

interaction with parallel elastic elements, external disturbances, and the environment, the<br />

bipeds motions emerge.<br />

In case several control units want to access the same actuator, a fusion <strong>of</strong> the control<br />

signals takes place. Depending on the type <strong>of</strong> the control data and the activity <strong>of</strong> the<br />

unit, one <strong>of</strong> two fusion methods is applied: torque signals are fused by a weighted sum<br />

function, similar to muscle stimulation adding up from different sources. All other control<br />

signals are fused using a weighted average function. More on this topic can be found in<br />

Section 5.1.3 discussing the behavior-based architecture.


74 4. A Biologically Supported Concept for <strong>Control</strong>ling Bipeds<br />

For further clarification <strong>of</strong> the hierarchical structure <strong>of</strong> the control system and <strong>of</strong> the flow<br />

<strong>of</strong> stimulation, Figure 4.3 again shows the layout <strong>of</strong> the control units, this time omitting<br />

the mechatronic part <strong>of</strong> the robot and the environment. Each locomotion mode stimulates<br />

a spinal pattern generator which again stimulates one motion phase at a time, or possibly<br />

one phase per body side. A motion phase activates postural reflexes as well as motor<br />

patterns and local reflexes within the joint groups. Reflexes and motor patterns can be<br />

recruited by more than one phase, and even be shared between locomotion modes. This<br />

matches the fact mentioned in Section 3.1.4 that the control processes <strong>of</strong> human walking<br />

and running seem to share the same control components.<br />

Figure 4.3 also illustrates how one phase <strong>of</strong> a locomotion mode is stimulated at one point<br />

in time. The control units <strong>of</strong> lighter shade indicate inactivity, those <strong>of</strong> darker shade<br />

indicate activation. Exemplary, the dashed line explains the information flow when a<br />

locomotion mode modulates the behavior <strong>of</strong> a motor pattern, e.g. based on the desired<br />

walking velocity. It must be mentioned that while only one phase <strong>of</strong> a locomotion mode<br />

can be stimulated at once for each side <strong>of</strong> the body, more than one locomotion mode can<br />

be activated simultaneously. For instance, during walking initiation, the standing mode<br />

can still continue to work.<br />

As a further remark it can be said that the suggested control concept could also work<br />

for motions other than locomotion. It is conceivable that a spinal pattern generator does<br />

not only create cyclic limb movement, but also voluntary and purposeful motion. For<br />

instance, a grasping process could be subdivided in different phases, each consisting <strong>of</strong><br />

motor patterns and reflexes <strong>like</strong> collision avoidance. Modulation from higher levels could<br />

coordinate the movement to reach in the direction <strong>of</strong> the desired goal. However, future<br />

work is needed to show the implementation <strong>of</strong> such a different task within this control<br />

methodology. The next chapter will describe how stable standing and dynamic walking <strong>of</strong><br />

a two-legged robot can be designed using this approach.


5. <strong>Control</strong>ling Dynamic Locomotion<br />

<strong>of</strong> a Fully Articulated Biped<br />

The preceding chapter presented the structural organization <strong>of</strong> the control concept. Six<br />

classes <strong>of</strong> control units were defined and a hierarchical network structure introduced.<br />

However, to apply this concept to the task <strong>of</strong> controlling the locomotion <strong>of</strong> a robot,<br />

appropriate control units must be found and be integrated within the suggested structure,<br />

thereby creating the desired system behavior. This chapter will demonstrate the design<br />

process for a dynamically walking biped. Methods for finding suitable functional control<br />

units will be discussed, and the resulting network for standing and walking will be<br />

introduced. But first prerequisites and assumptions for the underlying system will be<br />

addressed. These concern the kinematic layout and the actuation system <strong>of</strong> the assumed<br />

robot as well as the behavior-based architecture serving as framework for the control<br />

system.<br />

5.1 System Premises<br />

To achieve human-<strong>like</strong> walking with a bipedal robot, not only the control system but<br />

also the morphology and actuation system must resemble the human body to a sufficient<br />

degree. Nowadays it is impossible to construct a robot featuring the multitude <strong>of</strong> degrees<br />

<strong>of</strong> freedom found in the human body, or an actuation system similar in capabilities to the<br />

sophisticated muscle-tendon system. Likewise, the human perceptual abilities outdo any<br />

available technical sensors. Thus, this work has to agree on a practicable trade-<strong>of</strong>f. The<br />

following paragraphs will outline the characteristics <strong>of</strong> the robotic system upon which the<br />

control system will be established.<br />

5.1.1 Kinematic Layout<br />

As just mentioned, it is not possible to mimic all kinematic degrees <strong>of</strong> freedom the human<br />

body possesses. Instead, a subset <strong>of</strong> joints must be found that is simple enough to be<br />

implemented in a robotic system, but still extensive enough to allow for flexible locomotion.<br />

With dynamic walking being the target mode <strong>of</strong> locomotion, most degrees <strong>of</strong> freedom


76 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

z<br />

x y<br />

(a) (b) (c) (d)<br />

Figure 5.1: Kinematic layout <strong>of</strong> the assumed bipedal robot’s knee and ankle (a), hip (b),<br />

spine (c), and shoulder and elbow (d). Dashed lines denote the rotation axes.<br />

should be contained within the legs and hips. During normal walking, no extensive or<br />

elaborated movement can be observed within the spine, the neck, or the arms, consequently<br />

those parts <strong>of</strong> the robot can be simplified, though not completely ignored.<br />

Section 3.2.1 discussed the human gait with its dominant kinematic characteristics. Originating<br />

from this analysis, necessary joints for the leg design can be found. Figure 5.1a<br />

illustrates the resulting leg kinematics. To make stabilization in the frontal and lateral<br />

direction possible and to allow adaptation to uneven ground, the feet need at least two<br />

rotational degrees <strong>of</strong> freedom along both horizontal axes. The rotation about the vertical<br />

axis is neglected as it is also present in the hip joint where the main rotation about this<br />

axis takes place. The same holds true for the knee joint. During push-<strong>of</strong>f, the foot as well<br />

as the toes bend upwards. This rolling motion can be achieved by an additional joint or a<br />

curved foot shape. The latter solution is chosen for this work. As the knee joint mainly<br />

bends in the frontal plane and shows no other significant motions, only this degree <strong>of</strong><br />

freedom is included.<br />

The hip joint is implemented with all three rotational degrees <strong>of</strong> freedom (Figure 5.1b). All<br />

<strong>of</strong> these are essential for leg swing, lateral postural corrections, and curve walking. Even<br />

though the pelvis rotation during human walking shows only low amplitude, rotational<br />

joints around all three axes are included in the spine (Figure 5.1c). This allows to<br />

compensate the small pelvis motions as well as potentially larger pelvis shifts while e.g.<br />

walking on sloped ground. Furthermore, adjustments <strong>of</strong> the center <strong>of</strong> mass executed by<br />

postural reflexes are possible.<br />

Figure 5.1d illustrates the design <strong>of</strong> the upper trunk. In the scope <strong>of</strong> this work, additional<br />

degrees <strong>of</strong> freedom in the neck are omitted, but could become necessary for head stabilization<br />

should a vision system be included. The shoulder joint enables the arms to swing<br />

forth and back as well as to lift them in lateral direction for posture correction. An elbow<br />

joint is included for a more natural arm swing.<br />

Altogether, this kinematic setup results in a fully articulated biped with 21 degrees <strong>of</strong><br />

freedom to control. While this number requires the control system to cope with a system<br />

<strong>of</strong> considerable complexity, the above analysis suggests that this setup is necessary to<br />

achieve human-<strong>like</strong> three-dimensional walking.


5.1. System Premises 77<br />

5.1.2 Assumed Actuation and Sensor System<br />

Similar to the kinematics, the actuation system must possess certain properties to make a<br />

human-<strong>like</strong> motion control possible. Some properties <strong>of</strong> the muscle-tendon system have<br />

been discussed in Section 3.2.1. Hill’s mechanical muscle model (Figure 3.14b) features<br />

serial and parallel elastics elements, dampers, and a contracting unit for each muscle. A<br />

human joint is actuated by at least two <strong>of</strong> those muscles in an antagonistic setup. This<br />

results in a possible torque control and a position control with a large range <strong>of</strong> stiffness.<br />

Unfortunately, no technical actuator is available that comes even close to its natural<br />

counterpart. Pneumatic muscles at least partly resemble the biological muscle, but are<br />

difficult to model and to control, and the necessity to provide compressed air complicates<br />

the development <strong>of</strong> a self-sustaining system. Electronic actuation in combination with<br />

elastic elements could provide a partial solution. In the scope <strong>of</strong> the research efforts<br />

presented here, some considerations have been put on this topic [Luksch 05, Luksch 07,<br />

Wahl 09, Blank 09b]. Chapter 6 will present a short overview, but otherwise these studies<br />

will not be discussed in depth within this text.<br />

For this work, a direct joint actuation is assumed, neglecting the antagonistic setup.<br />

This has the advantage <strong>of</strong> reducing complexity in design and control. A drawback could<br />

arise as no biarticular structures can be implemented mechanically. In future work, an<br />

antagonistic joint actuation system similar to the human one could be investigated as a<br />

possible extension.<br />

Still, the actuation system applied here features some properties <strong>of</strong> the natural example.<br />

Otherwise, the suggested control concept could not demonstrate all <strong>of</strong> its potential benefits.<br />

It is assumed that each joint can swing freely with low friction if no control commands<br />

are given. A direct torque demand is possible. Beside this, position control with variable<br />

compliance simulates the series elastic elements <strong>of</strong> the muscle. A fusion <strong>of</strong> both control<br />

methods is also provided for. Similar to the muscle-tendon system, a parallel elastic<br />

element with a fixed spring constant and equilibrium point concludes the joint actuation<br />

characteristics. The following chapter will describe how these properties were implemented<br />

in the simulated robotic system.<br />

Furthermore, some requirements have to be met regarding the sensor configuration to<br />

enable the transfer <strong>of</strong> certain biological control strategies to technical systems. As it is the<br />

case in the human muscle, each joint outputs the current angular position and the acting<br />

torque. Cutaneous and load receptor information is required by certain reflexes in human<br />

motion control. Consequently, it is assumed that force sensors in each <strong>of</strong> the robot’s feet<br />

provide data on ground contact and load distribution. Finally, postural reflexes depend<br />

on an estimation <strong>of</strong> the upper trunk’s pose and movements. For this purpose, an inertial<br />

measurement unit (imu) or equivalent sensors have to be installed.<br />

5.1.3 The Behavior-based <strong>Control</strong> Architecture iB2C<br />

Besides the morphology and the sensor-actor system, an appropriate control framework is<br />

a further requirement for implementing a human-<strong>like</strong> locomotion control system. Over the<br />

last 20 years, behavior-based architectures have been established as a preferred control<br />

approach for robotic systems acting in situations and environments that are not known a<br />

priori. Compared to other robot control architectures, they can be classified as a hybrid


78 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

s<br />

�ı<br />

�e<br />

F (�e,s,�ı )<br />

(a)<br />

�u<br />

�a<br />

r<br />

s<br />

�ı<br />

�e<br />

F (�e,s,�ı )<br />

Figure 5.2: The basic behavior module (a) and the fusion behavior module (b) <strong>of</strong> the iB2C<br />

architecture.<br />

system, being located between purely reactive and highly deliberative systems. But in<br />

contrast to normal hybrid systems, behavior-based control shows no conceptual break<br />

between a reactive and deliberative layer. Rather, it is characterized by a homogeneous<br />

layout.<br />

The subsumption architecture by R. Brooks is commonly seen as being the first behaviorbased<br />

strategy [Brooks 86]. Here, mainly reactive control units are arranged in a hierarchical<br />

layout, and suppression signals serve as arbitration method. The subsumption<br />

architecture also shows the feasibility <strong>of</strong> controlling biologically inspired walking machines<br />

by behavior-based architectures in its application to the six-legged robots Genghis<br />

and Attila [Brooks 89, Ferrell 95]. Further biologically motivated robots controlled in a<br />

behavior-based manner include the lobster-<strong>like</strong> machine by Ayers or the quadruped WARP1<br />

by Pettersson et al. [Ayers 00, Pettersson 98]. Arkin provides an extensive overview on<br />

the topic <strong>of</strong> behavior-based architectures, Pirjanian discusses possible coordination mechanisms<br />

[Arkin 98, Pirjanian 99]. Behavior-based control has also been successfully applied<br />

to wheel-driven robots [Arkin 00, Matarić 97]. So far, the control <strong>of</strong> dynamic locomotion<br />

<strong>of</strong> bipedal robots has not been in the focus <strong>of</strong> behavior-based control research.<br />

Considering the key aspects <strong>of</strong> biological locomotion control identified in Chapter 3 and<br />

the understanding <strong>of</strong> the role <strong>of</strong> control units, utilizing a behavior-based architecture<br />

as framework for the suggested control concept seems promising. The behavior-based<br />

control architecture iB2C (integrated Behavior-Based <strong>Control</strong>) is well suited for implementing<br />

the aspired hierarchical layout, the control unit classes, or the stimulation and<br />

inhibition mechanism. It is based on architectural ideas developed for the control <strong>of</strong><br />

multipedal walking machines [Luksch 02, Albiez 03a, Albiez 03b, Albiez 03c, Albiez 07].<br />

Since then the specification <strong>of</strong> its behavioral components has been further refined and<br />

formalized [Proetzsch 07, Proetzsch 10]. The architecture is also applied to the control <strong>of</strong><br />

complex wheeled robots for indoor and <strong>of</strong>f-road applications [Proetzsch 04, Kleinlützum 05,<br />

Proetzsch 05, Schäfer 05, Schmidt 06]. The following paragraphs will give a short introduction<br />

on the ideas <strong>of</strong> the iB2C and will describe how the features required by the control<br />

concept can be implemented.<br />

5.1.3.1 The Basic iB2C Behavior Module<br />

Each individual behavior <strong>of</strong> iB2C is contained in a module as shown in Figure 5.2a. That<br />

way, all behaviors feature the same, common interface. An iB2C behavior B can be<br />

formulated as a three-tuple<br />

(b)<br />

�u<br />

�a<br />

r


5.1. System Premises 79<br />

B = (fa,fr,F) (5.1)<br />

where fa denotes the activity function, fr the target rating function, and F the transfer<br />

function. These functions calculate the activity information �a, a target rating r, and an<br />

output vector �u. The behavior interface is completed by the input vector�e, the stimulation<br />

s, and the inhibition vector�ı.<br />

The stimulation s ∈ [0,1] is used by higher levels <strong>of</strong> the control system to adjust the<br />

influence <strong>of</strong> the behavior. It is complemented by the inhibition i ∈ [0,1] enabling other<br />

control units to actively suppress the behavior. The combined inhibition i <strong>of</strong> a behavior is<br />

calculated based on the inhibition vector as<br />

i = max<br />

j=0,...,k−1 (ij). (5.2)<br />

The stimulation and inhibition determine the activation ι = s·(1−i) <strong>of</strong> the behavior. An<br />

activation <strong>of</strong> ι = 0 corresponds to a fully inactive behavior, ι = 1 to maximum possible<br />

activity and output values. The signals s and�ı correspond to the stimulating and inhibiting<br />

signals <strong>of</strong> control units in neural control. Thus, the requirement stated in Section 4.2 <strong>of</strong><br />

allowing the influence on the activation <strong>of</strong> control units within the network is met.<br />

In addition to these influencing input values, two more behavior signals complete the<br />

module interface. The activity a ∈ [0,1] denotes the influence the behavior wants to exert<br />

on the network in the current state. An activity <strong>of</strong> a = 0 represents an inactive unit,<br />

whereas a = 1 denotes that the behavior intends to have maximum impact with its output<br />

values. It is calculated by the activity function<br />

fa :R m ×[0,1] → [0,1]×[0,1] l , fa(�e,ι) =�a = (a,�a) T<br />

(5.3)<br />

which can also generate an optional vector �a <strong>of</strong> derived activities a i ≤ a used for reduced<br />

stimulation <strong>of</strong> other behaviors. High activity can be interpreted as the firing <strong>of</strong> a neuron,<br />

thus again bridging to natural control concepts. As the activity <strong>of</strong> one behavior, or a<br />

value derived from it, normally serves as stimulation for a lower behavior in the control<br />

hierarchy, a ≤ ι must hold true. As a result, the total activity does not increase as it flows<br />

through the control network.<br />

Thetargetratingr ∈ [0,1]calculatedbythetargetratingfunctionfr statesthecontentment<br />

<strong>of</strong> the behavior with the current situation. A rating <strong>of</strong> r = 0 indicates that the behavior<br />

is satisfied with the state, r = 1 represents maximum discontentment. The target rating<br />

function only refers to the restricted view on the current situation as perceived by the<br />

input vector �e. As with the activity signal, the target rating can prove to be useful during<br />

network debugging, or it can serve as input signal for other control units.<br />

Finally, the transfer function formulates the relation between the input data vector �e and<br />

the output data vector �u:<br />

F :R m ×[0,1] →R n , F (�e,ι) = �u. (5.4)<br />

It further depends on the current activation ι <strong>of</strong> the behavior. In most cases <strong>of</strong> this<br />

work, the activation scales the control values <strong>of</strong> the output vector. The transfer function<br />

implements the actual intelligence <strong>of</strong> the behavior based on its view on the current situation<br />

and a possible internal model. It codes whether the behavior represents a locomotion<br />

mode, a reflex, or one <strong>of</strong> the other possible classes <strong>of</strong> control units. Actual implementations<br />

<strong>of</strong> the behavior functions just described will be found in this and the following chapters.


80 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

B1<br />

5.1.3.2 Behavior Coordination<br />

B2<br />

Fusion<br />

Behavior<br />

Figure 5.3: Example for the fusion <strong>of</strong> three behaviors.<br />

In case several behaviors want to access the same control value, e.g. a motor torque or the<br />

stimulation <strong>of</strong> another behavior, their action must be coordinated. For this purpose, fusion<br />

modules are introduced (Figure 5.2b). Being behaviors themselves, they feature the same<br />

interface, but have a special form <strong>of</strong> input and output vectors as well as transfer function.<br />

A fusion behavior always collects identical outputs <strong>of</strong> competing behaviors and generates<br />

a similar output (Figure 5.3). The input vector �e <strong>of</strong> the fusion behavior is composed <strong>of</strong><br />

the activity ai, the target rating ri, and the output vector �ui for each fused behavior. The<br />

incoming activities are used for arbitration, the target ratings are used to calculate a<br />

meaningful target rating for the fusion behavior itself.<br />

For now, iB2C <strong>of</strong>fers three methods for behavior fusion: maximum fusion, weighted average,<br />

and weighted sum. A fusion module using maximum fusion mirrors the outputs <strong>of</strong> the<br />

behavior with the highest activity as its own. In the case <strong>of</strong> weighted average, the activity,<br />

the target rating, as well as the output values are weighted by the activity <strong>of</strong> the p<br />

competing behaviors:<br />

a =<br />

p−1 �<br />

a<br />

j=0<br />

2 j<br />

p−1 �<br />

ak<br />

k=0<br />

·ι r =<br />

p−1 �<br />

aj ·rj<br />

j=0<br />

p−1 �<br />

ak<br />

k=0<br />

�u =<br />

B3<br />

p−1 �<br />

aj ·�uj<br />

j=0<br />

p−1 �<br />

ak<br />

k=0<br />

(5.5)<br />

In this way, the behavior with the highest activity has the strongest impact on the control<br />

network. The weighted sum fusion method weights the activities and the control outputs<br />

by the activity relative to the maximum activity amax <strong>of</strong> all connected behaviors:<br />

a = min<br />

� p−1<br />

�<br />

a 2 j<br />

amax<br />

j=0<br />

,1<br />

�<br />

·ι r =<br />

p−1 �<br />

aj ·rj<br />

j=0<br />

p−1 �<br />

ak<br />

k=0<br />

�u =<br />

p−1 �<br />

j=0<br />

aj ·�uj<br />

amax<br />

(5.6)<br />

This last fusion method acts similar to motor neurons collecting the incoming activity<br />

<strong>of</strong> connected neurons. Consequently, this fusion method is used when coordinating joint<br />

torque commands as will be described in Chapter 6. For most other control outputs, e.g.<br />

joint stiffness or joint position, weighted average fusion is applied. In relying on a uniform<br />

and established coordination concept for competing control units, the design <strong>of</strong> control<br />

networks is simplified as no additional effort must be invested in arbitration mechanisms.


5.2. Guidelines for Designing <strong>Control</strong> Units 81<br />

5.2 Guidelines for Designing <strong>Control</strong> Units<br />

After selecting a target locomotion mode, functional control units have to be selected<br />

and designed that will achieve the desired motions. The discussion on human walking<br />

in Section 3.2 can serve as source for this process. Gait analysis and neuroscientific<br />

research provide suitable information on selecting feed-forward and feedback control units.<br />

Examining the human gait also helps to find appropriate compliance settings for passive<br />

joints during different phases <strong>of</strong> locomotion. The following section will present some<br />

guidelines on how to design control networks.<br />

5.2.1 Feed-Forward <strong>Control</strong> Units<br />

As mentioned earlier, motor patterns provide the means for feed-forward control. They<br />

allow to shape the passive dynamics according to the target gait, or to initiate limb and<br />

body motions. Fast and promising results in designing these patterns can be achieved<br />

when starting from biomechanical gait analysis. emg data on the muscle activities and<br />

kinetic calculations during the individual phases allow to draw conclusion on the necessity<br />

<strong>of</strong> torque input.<br />

In most cases it is possible to find a semantic interpretation <strong>of</strong> the muscle activity. For<br />

instance, right before the leg swing can be observed, the muscles bending the hip joint are<br />

stimulated. The reason for this can be found in the necessary acceleration <strong>of</strong> the tibia, also<br />

resulting in the passive bending <strong>of</strong> the knee, and thus in ground clearance <strong>of</strong> the swing<br />

foot. Similar patterns can be found in the hip joints during walking initiation, or in the<br />

ankle joint during the push-<strong>of</strong>f phase.<br />

In any case it is necessary to find a suitable function to create torque patterns in a technical<br />

system. It should avoid sudden torque changes creating stress on the mechanics. Also, it<br />

should be possible to define the function with only a few parameters, and still be flexible<br />

enough to allow for a variety <strong>of</strong> torque trajectories. Equation 5.7 presents the sigmoid<br />

function that is suggested as basis for motor patterns:<br />

⎧<br />

1 1 t 1 ⎪⎨ + sin( π( − ) ) 0 ≤ t < T1<br />

2 2 T1 2<br />

ˆτ = A· 1 T1 ≤ t < T2<br />

⎪⎩ 1 1 t−T2 1 − sin( π( − 2 2 T3−T2 2 ) ) T2<br />

(5.7)<br />

≤ t ≤ T3<br />

Using the three parameters T1, T2, and T3 it is possible to define the length <strong>of</strong> the sine-<strong>like</strong><br />

rise <strong>of</strong> the function, the duration <strong>of</strong> the plateau at maximum torque output, and the<br />

time until the sine-<strong>like</strong> fall <strong>of</strong> the function reaches zero. The output value ˆτ ∈ [−1,1] is<br />

normalized to the maximum torque <strong>of</strong> the joint in question, i.e. the effective torque τ<br />

arises to τ = ˆτ ·τmax. Figure 5.4 depicts the sigmoid function and its parameters, and also<br />

gives some examples illustrating how sign and amplitude can be varied by the factor A.<br />

For technical system that do not resemble a biological example, no biomechanical data<br />

is available to be consulted for finding motor patterns. One possibility to obtain similar<br />

data can be optimization techniques. Optimal control theory can generate torque control<br />

patterns for target movements based on a simplified model <strong>of</strong> the robot and its environment.<br />

These torque trajectories can then again be used as foundation for designing motor patterns.<br />

This method has also be investigated by the author and others for a robot <strong>of</strong> reduced<br />

complexity [Luksch 07].


82 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

1.0<br />

-1.0<br />

ˆτ<br />

T1 T2 T3<br />

t<br />

ˆτ<br />

Figure 5.4: Motor patterns generate torque commands as sigmoid function. Three parameters<br />

define the end <strong>of</strong> the rise (T1), the start <strong>of</strong> the fall (T2), and the end <strong>of</strong> the fall (T3). As shown<br />

in the second example, a factor can scale the output.<br />

5.2.2 Feedback <strong>Control</strong> Units<br />

Feedback is introduced by three types <strong>of</strong> control units: spinal pattern generators, local<br />

reflexes, and postural reflexes. One spinal pattern generator exists for each mode <strong>of</strong><br />

locomotion. For designing its phases, findings from the decomposition <strong>of</strong> emg data as<br />

described in Sections 3.1.4 and 3.2.2 can be called on. Each component identified in the<br />

emg analysis can be translated to one locomotion phase. The kinematic or kinetic events<br />

associated with the components provide the conditions for phase transitions.<br />

Local reflexes can partly be adopted from those found during human locomotion (Section<br />

3.1.3). Whether this transition is feasible mainly depends on the availability <strong>of</strong><br />

corresponding sensory information and actuators. Possibly the reflex action has to be<br />

adopted to the robots capabilities. For instance, cutaneous reflexes normally rely on<br />

sensitive skin. If no artificial skin is available, these reflexes could be triggered by isolated<br />

force measurements or data from load cells. Other local reflexes can be deduced by<br />

matching kinematic events with distinctive muscle activities. One example can be found<br />

in the passive knee movement during the swing phase. As soon as the knee is stretched<br />

again before touching down with the swing leg, muscle activity stabilizing the knee can be<br />

observed.<br />

For designing postural reflexes, several ways are conceivable. Some possible reflexive<br />

actions concerning posture stabilization, e.g. the labyrinthine righting reflex, are suggested<br />

in neuroscientific literature (Section 3.2.4). Again, it might prove difficult to adapt these<br />

reflexes due to missing sensor data. Another way to find postural reflexes is based on<br />

simplifying biomechanical models. As discussed in Chapter 3, H<strong>of</strong>, Seyfarth, or Bauby<br />

among others have tried to derive simplified mechanical models explaining correcting<br />

movements or muscle activity observed during human locomotion. These models can be<br />

<strong>of</strong> low complexity and thus be predestined as possible postural reflex. Finally, empirical<br />

validation <strong>of</strong> the robot’s behavior during experiments can point out existing flaws in<br />

posture control, leading to additional reflexes. This strategy, by the way, can be applied<br />

for all other classes <strong>of</strong> control units, too. The common problem in designing postural<br />

reflexes is the requirement for information on the body pose or the behavior <strong>of</strong> the center<br />

<strong>of</strong> mass. The simplified biomechanical models just mentioned can provide a possibility to<br />

derive the necessary data.<br />

t<br />

ˆτ<br />

t<br />

ˆτ<br />

t


5.3. Stable Standing 83<br />

5.2.3 Guidelines for Implementing iB2C Features<br />

The preceding paragraphs focused on the functionality <strong>of</strong> control units. In terms <strong>of</strong> iB2C,<br />

this corresponds to the transfer function F, describing the calculation <strong>of</strong> control signals<br />

based on the behavior activation and the received sensor information. But while iB2C<br />

<strong>of</strong>fers powerful mechanisms and tools, it requires the designer to take care <strong>of</strong> additional<br />

behavior signals, mainly the stimulation, inhibition, and activity. General guidelines and<br />

principles facilitating the development <strong>of</strong> iB2C control systems are discussed by Proetzsch<br />

et al. [Proetzsch 10]. Some rules being <strong>of</strong> importance during the design process following<br />

the control concept suggested here are given in the following.<br />

Stimulation and inhibition <strong>of</strong> one behavior by another should be based on the activity <strong>of</strong><br />

the controlling behavior. In most cases it is sufficient to distinguish between no stimulation<br />

(s = 0) and full stimulation (s = 1). A locomotion mode is fully stimulated by the operator<br />

and shows full activity. Its activity a = 1 will then stimulate a spinal pattern generator,<br />

which again will stimulate motion phases only by a value <strong>of</strong> zero or one. A stimulated<br />

motion phase should show an activity <strong>of</strong> a = 1, and will normally use this activity to<br />

stimulate reflexes and motor patterns. In a few cases it might prove necessary to use<br />

a reduced activity value for stimulation, e.g. when a reflex should only show half <strong>of</strong> its<br />

normal activity during a certain motion phase.<br />

While most higher level behaviors can and should be reduced to activity <strong>of</strong> only one or<br />

zero, reflexes and motor patterns need more sophisticated implementations <strong>of</strong> their activity<br />

function. Motor patterns feature an activity matching their torque output. Reflexes<br />

should increase to full activity as soon as the situation assessment based on their restricted<br />

sensor information indicates a risk <strong>of</strong> loosing stability. Section 6.2.2 <strong>of</strong> the following<br />

chapter will describe how the activity functions <strong>of</strong> the behaviors developed in this work<br />

are implemented.<br />

With control unit classes defined and guidelines for selecting and designing functional units<br />

suggested, control networks for bipedal locomotion can now be developed. The control<br />

units will be arranged following the hierarchical layout described in the previous chapter.<br />

In the remainder <strong>of</strong> this chapter, networks for stable standing and dynamic walking are<br />

presented.<br />

5.3 Stable Standing<br />

Although this work focuses on dynamic walking, stable standing is considered, too. This<br />

allows to observe the transition from standing to walking, and validates the control concept<br />

for a second task <strong>of</strong> different nature. A short review <strong>of</strong> human standing control will<br />

motivate the control network introduced in this section. Additional information on this<br />

part regarded from a slightly different perspective and including further details on the<br />

implementation can be found in [Steiner 08].<br />

5.3.1 Balanced Standing <strong>Control</strong> in <strong>Human</strong>s<br />

<strong>Human</strong>s show a remarkable ability to stand securely on highly unstructured ground and<br />

under considerable disturbances. Generally, postural control is considered to be a complex<br />

skill rather then the result <strong>of</strong> only local reflex action [Horak 06]. Two tasks can be


84 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

Figure 5.5: The ankle and hip strategies for postural balance generate different joint movements<br />

and muscle activities during a movement <strong>of</strong> the support surface. From [Ting 07], p305.<br />

distinguished: postural orientation and postural equilibrium. Postural orientation involves<br />

active adaptation <strong>of</strong> the body segment’s pose considering gravity, structure <strong>of</strong> the support<br />

surface, or additionally body load. Head, trunk, and extremities are aligned in a way that<br />

a relaxed, stable posture emerges. Information <strong>of</strong> the body’s whole sensory apparatus is<br />

necessary to achieve this goal.<br />

Postural equilibrium however implies the reaction to self-induced or external disturbances.<br />

Three basic strategies to achieve this goal can be discerned: the ankle, the hip, and the step<br />

strategy. Which strategy is chosen does not only depend on the kind and magnitude <strong>of</strong> the<br />

disturbance. Personal experience, goals, and current expectations can result in different<br />

responses for each individual. Furthermore, the contribution <strong>of</strong> the human sensory systems<br />

varies from subject to subject [Horak 90]. Most notably this is the case in disabled patients,<br />

e.g. suffering from vestibular loss. One estimation for the distribution <strong>of</strong> the different<br />

senses for balancing in healthy subjects states a partial contribution <strong>of</strong> approximately<br />

70% for the somatosensory system, 20% for the vestibular system, and 10% for the visual<br />

system [Horak 06].<br />

The most basic response strategy being observed at small to medium disturbances is the<br />

ankle strategy. It lends its name from the fact that a correcting movement is issued by the<br />

ankle joint while the rest <strong>of</strong> the body is kept stiff. This <strong>of</strong> course implies muscle action<br />

all along one side <strong>of</strong> the body. Figure 5.5 illustrates the reaction to a backward shift <strong>of</strong><br />

the supporting platform a subject is standing on. Muscles in the back and hip keep the<br />

body rigid while the gastrocnemius and soleus muscles prevent the subject from falling<br />

to the front and create a plantarflexion <strong>of</strong> the ankle for stabilization [Ting 07]. Winter<br />

et al. suggest a simple control scheme for quiet standing where the ankle muscles act as<br />

springs [Winter 98, Winter 01]. This results in a stiffness control for balancing that can<br />

react mechanically and as such almost instantly without a delay caused by the nervous<br />

system. It also explains the body sway that can usually be observed during quiet standing.<br />

The ankle strategy is mechanically constrained by the support area <strong>of</strong> the foot. If the<br />

center <strong>of</strong> preassure moves outside this support area, the ankle strategy alone can no longer<br />

regain balance [Horak 90].<br />

In this case, the hip strategy is able to exert greater influence. In contrast to the ankle<br />

strategy, itgeneratesashift<strong>of</strong>thecenter<strong>of</strong>massintheoppositedirection<strong>of</strong>thedisturbance.<br />

This is accomplished by a coordinated movement <strong>of</strong> the ankle, knee, hip, and spinal column.


5.3. Stable Standing 85<br />

<strong>Walking</strong><br />

Stable<br />

Standing<br />

Standing<br />

States<br />

1<br />

2<br />

3<br />

4<br />

Ground<br />

Adapt.<br />

Stabilization<br />

Optimization<br />

Relaxed<br />

Posture<br />

Figure 5.6: <strong>Control</strong> network for stable standing.<br />

As shown in Figure 5.5, the upper body moves in the direction <strong>of</strong> the disturbance, the<br />

pelvis in the opposite direction. The arm support the posture correction by moving in the<br />

same direction as the trunk. Finally, the step strategy consists <strong>of</strong> an evasive step taken in<br />

the direction <strong>of</strong> the disturbing force. This response has the largest effect on the center <strong>of</strong><br />

mass and can compensate the strongest disturbances.<br />

For the standing control <strong>of</strong> a bipedal robot it is preferable to implement both the postural<br />

orientation and the postural equilibrium task. The first allows to operate not only on<br />

flat surface but also on unstructured ground and with different mass distributions <strong>of</strong> the<br />

robot. At least the ankle strategy should be included for basic functionality <strong>of</strong> the postural<br />

equilibrium task. An extended stiffness control as postulated by Winter et al. seems to<br />

<strong>of</strong>fer the necessary abilities.<br />

5.3.2 Stabilized Standing <strong>Control</strong><br />

The control network for balanced standing suggested in this work does include the postural<br />

orientation and equilibrium task as well as the ankle strategy. Furthermore the assumption<br />

<strong>of</strong> a level ground is discarded in favor <strong>of</strong> an approach for flexible ground adaptation.<br />

Figure 5.6 illustrates the upper layers <strong>of</strong> the resulting control network. As introduced<br />

in the previous chapter, a finite state machine coordinates the different phases occurring<br />

during standing. It stimulates one <strong>of</strong> four phase units depending on the current state <strong>of</strong> the<br />

robot. The Ground Adaptation phase is responsible for setting a sensible posture before<br />

the feet touch the ground and for achieving an appropriate initial posture considering the<br />

terrain geometry as the robot settles on the support surface. As its name suggests, the<br />

Stabilization phase balances the system against disturbances to gain a stable posture. This<br />

phase corresponds to the postural equilibrium task and includes a basic ankle strategy.<br />

The postural orientation task is handled during the Optimization phase. It changes the<br />

target angles to reduce the necessary torques within the joints. In case such a low energy<br />

posture can be found, the joint stiffness is reduced in the Relaxed Posture phase.<br />

5.3.3 Stable Standing State Machine<br />

The spg for stable standing is implemented as the finite state machine shown in Figure 5.7.<br />

Depending on its current state, the corresponding phase is stimulated. The initial state<br />

responsible for ground adaptation is left as soon as full ground contact is achieved. This<br />

information is received from the leg joint groups that preprocess various contact sensor


86 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

Ground<br />

Adaptation<br />

Relaxed<br />

Posture<br />

full ground contact<br />

disturbances<br />

low joint torques<br />

disturbances<br />

Stabilization<br />

Optimization<br />

quite stance<br />

Figure 5.7: Transitions <strong>of</strong> finite state machine managing phases during stable standing.<br />

measurements to a combined ground contact signal per leg. Then the stabilization state is<br />

entered in which the adopted pose is steadied mainly by high joint stiffness.<br />

If a quiet stance can be achieved, the state transition to the optimization state takes place.<br />

This decision is based on measurements from the inertial measurement unit. It provides<br />

three-dimensional acceleration and angular velocity information for the trunk. Based on<br />

the pose estimation the gravitational components <strong>of</strong> the acceleration vector is eliminated.<br />

By integrating this data a rough approximation <strong>of</strong> the trunk’s velocity is calculated. Only<br />

if this velocity and the acceleration stay below a certain threshold, the state transition is<br />

performed.<br />

During the optimization phase several postural behavior modules adapt the robot’s stance<br />

to find a posture consuming only little energy. Following the iB2C specification, these<br />

behaviors calculate a target rating to evaluate the current situation regarding their goal.<br />

Only if all <strong>of</strong> these ratings are low, i.e. all postural reflexes are content, the spg proceeds<br />

to the next state. As now a balanced stance should be adopted, the joint stiffness can be<br />

reduced during the Relaxed Posture phase.<br />

In case disturbances are perceived by the inertial measurement unit or the ground contact<br />

sensors, the state machine switches back to the stabilization state. Similarly, should ground<br />

contact be lost, or the locomotion mode for standing be inhibited, the ground adaptation<br />

state is resumed. In Figure 5.7, the corresponding transition are omitted for clarity. The<br />

following paragraphs will describe the functioning <strong>of</strong> the introduced phases.<br />

5.3.4 Ground Adaptation<br />

The basic idea <strong>of</strong> ground adaptation lies in the compliance <strong>of</strong> the joints. By setting the<br />

joints to a low stiffness, they passively adjust their position to external forces. For instance,<br />

the ankle joint can rotate in a way that the foot settles on uneven ground. As soon as full<br />

ground contact is made, the current joint angles are applied as set point and the stiffness<br />

can be raised in the stabilization phase.<br />

The described functionality is assumed by the Hold Position reflex. An instance <strong>of</strong> this<br />

reflex is included for each joint and stimulated by the standing phase as illustrated in<br />

Figure 5.8. The reflex can be parametrized by an initial angle and three stiffness values:<br />

the initial stiffness stiff init is set at first stimulation and in case there is no ground contact.<br />

The minimal and maximal stiffness parameters stiff min and stiff max define an interval <strong>of</strong>


5.3. Stable Standing 87<br />

Standing<br />

States<br />

1<br />

2<br />

3<br />

4<br />

Ground<br />

Adapt.<br />

Stabilization<br />

Optimization<br />

Relaxed<br />

Posture<br />

Upper Trunk<br />

Hold<br />

Hold<br />

Position<br />

Position<br />

Lower Trunk<br />

Hold<br />

Hold<br />

Position<br />

Position<br />

Left/Right Leg<br />

Hold<br />

Hold<br />

Position<br />

Position<br />

Figure 5.8: <strong>Control</strong> network during the Ground Adaptation phase <strong>of</strong> stable standing.<br />

compliance used during ground contact. A stiffness factor received as modulation input<br />

determines the value within this interval that is actually set as stiffness. Algorithm 5.1<br />

summarizes the behavior <strong>of</strong> the Hold Position reflex.<br />

Algorithm 5.1: Target angle α and stiffness stiff calculation <strong>of</strong> the Hold Position reflex<br />

1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

Initiation:<br />

α = αinit<br />

stiff = stiff init<br />

if a = 0 or no ground contact then<br />

α = αinit<br />

stiff = stiff init<br />

end<br />

else if initial ground contact then<br />

α = αcurrent<br />

stiff = stiff min<br />

end<br />

else if full ground contact then<br />

α = α+∆α<br />

stiff = stiff min +(stiff max −stiff min)·stiffness factor<br />

end<br />

At the start <strong>of</strong> the control system or following a reset command, the target values are set<br />

to the initial angle and stiffness. As soon as first ground contact is measured, the target<br />

angle α follows along the current joint angle αcurrent. When full ground contact is detected,<br />

the target joint angle is fixed, and the compliance is set depending on the stiffness factor.<br />

During the ground adaptation phase, this factor is set to zero resulting in the stiffness<br />

being set to the minimal stiffness parameter.<br />

Beside the stiffness factor, the Hold Position reflex receives a modulation signal ∆α for<br />

changing the set point <strong>of</strong> the joint angle. This angular velocity is integrated by the reflex


88 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

Standing<br />

States<br />

1<br />

2<br />

3<br />

4<br />

Ground<br />

Adapt.<br />

Stabilization<br />

Optimization<br />

Relaxed<br />

Posture<br />

Relax<br />

Joints<br />

Reduce<br />

Tension<br />

Upper Trunk<br />

Hold<br />

Hold<br />

Position<br />

Position<br />

Lower Trunk<br />

Hold<br />

Hold<br />

Position<br />

Position<br />

Left/Right Leg<br />

Hold<br />

Hold<br />

Position<br />

Position<br />

Figure 5.9: <strong>Control</strong> network during the Optimization phase <strong>of</strong> stable standing. The dashed<br />

arrows indicate modulation signals.<br />

and added to the current set point. In case the ground contact is lost or the reflex is no<br />

longer activated, the integral part is set to zero. Using the modulation signal, the posture<br />

optimization control units described below can influence the stance.<br />

5.3.5 Posture Stabilization<br />

The stabilization phase starts as soon as full ground contact is established or when a<br />

disturbance is measured in one <strong>of</strong> the other standing phases. Its control network is laid out<br />

the same way as the one during the Ground Adaptation phase. Only the stiffness factor is<br />

set to a high value to stabilize the stance. In doing so, the joints generate high torques in<br />

case the body is deflected from its target posture. With the trunk and extremities braced,<br />

the compensation will mainly origin from the ankle joints, thus implementing the ankle<br />

strategy. Similar reactions can be observed in humans being disturbed during a relaxed<br />

stance. Muscle activities throughout the body increase and the stiffened body is pushed<br />

against the disturbance by the ankle joint to regain a balanced pose. By only locally<br />

controlling the joints towards a determined target angle, no global feedback loop or model<br />

depending control is necessary and the delay times can be kept low. This strategy can<br />

only work if the target stance retains at least the center <strong>of</strong> mass above the support surface<br />

spanned by the feet. Otherwise, the robot will not be able to remain standing.<br />

5.3.6 Posture Optimization<br />

After the posture is stabilized and a sufficiently quiet stance is achieved, the target pose<br />

can be optimized. This phase corresponds to the postural orientation task. As shown in<br />

Figure 5.9, the Hold Position reflexes are still stimulated. Additionally, postural reflexes<br />

adapt the stance by shifting the joint set points according to the provided modulation<br />

signal. Two types <strong>of</strong> these adaptation reflexes can be distinguished: those trying to relax<br />

joints, i.e. to reduce the necessary torque, and those trying to release tension acting


5.3. Stable Standing 89<br />

<strong>Control</strong> Unit Manipulated Joints<br />

Relax Spine X x-axis <strong>of</strong> spine joint<br />

Relax Spine Y y-axis <strong>of</strong> spine joint<br />

Relax Hip Y y-axis <strong>of</strong> both hip joints<br />

Relax Knee both knee joints<br />

Balance Ankle Pitch y-axis <strong>of</strong> both ankle joints<br />

Reduce Tension Hip Y y-axis <strong>of</strong> both hip joints<br />

Reduce Tension Ankle Y y-axis <strong>of</strong> both ankle joints<br />

Table 5.1: Postural reflexes adapting the posture during the Optimization phase.<br />

because <strong>of</strong> opposing set points. Table 5.1 lists the implemented postural reflexes and notes<br />

the joint angles being adapted by each reflex.<br />

The relax joint control units receive the current joint angles and the corresponding torques<br />

as input. In case the joint limits are not exceeded, a change rate for the joints is forwarded<br />

to the Hold Position reflexes. The change rate is proportional to the torque, thus enabling<br />

a posture adaptation reducing the torques. To keep dynamic effects low during the<br />

adaptation, the change rate is relatively slow. Only the Balance Ankle Pitch follows a<br />

different optimization approach. Instead <strong>of</strong> reducing the joint torque, it tries to optimize<br />

the center <strong>of</strong> pressure acting in the foot. This information is derived from four force sensors<br />

located within the outer parts <strong>of</strong> the foot sole. Based on these measurements the center <strong>of</strong><br />

pressure can be estimated and shifted below the ankle joint by the reflex.<br />

While the relax joint reflexes move one or two joints in the same direction, the reduce<br />

tension control units adjust two paired joints in opposed direction. This strategy is useful<br />

if tension arises due to external forces and limitations by two-legged ground contact. For<br />

instance, concurrent dorsiflexion <strong>of</strong> the one foot and plantarflexion <strong>of</strong> the contralateral<br />

foot causes opposing torques that can be resolved by adjusting the ankle angles without<br />

altering the posture. The same holds true for flexion and extension <strong>of</strong> the hip joints. The<br />

control unit works by trying to minimize the torque discrepancy <strong>of</strong> the two joints being<br />

considered. The generated change rate is proportional to the difference <strong>of</strong> one joint’s<br />

torque and the average torque <strong>of</strong> both joints. Again, the velocity <strong>of</strong> adaptation is low<br />

to reduce dynamic effects. Both the relax joint and reduce tension control units can act<br />

simultaneously on the same joint as it is the case for the y-axes <strong>of</strong> the ankles and the hips.<br />

The change rates from the competeting reflexes are then combined by a fusion behavior<br />

using the weighted average method.<br />

5.3.7 Relaxed Stance<br />

Following the iB2C specification, the angle adjustment reflexes just described indicate<br />

their progress by the target rating signal. The maximum <strong>of</strong> all these ratings can serve as a<br />

reference on the balance <strong>of</strong> the current stance. If this value drops below a certain threshold,<br />

the Relaxed Posture phase <strong>of</strong> stable standing starts. Its control network again mimics the<br />

one <strong>of</strong> the Ground Adaptation or Stabilization phase, i.e. the Hold Position reflexes are<br />

stimulated. But as a balanced posture can be assumed, the stiffness factor can slowly<br />

be decreased to take up a relaxed stance with low energy consumption. If the posture is


90 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

(Running<br />

Initiation)<br />

(Running)<br />

Operator<br />

Standing <strong>Walking</strong><br />

(Cyclic<br />

Running)<br />

Cyclic<br />

<strong>Walking</strong><br />

<strong>Walking</strong><br />

Initiation<br />

Figure 5.10: Coordination <strong>of</strong> three locomotion modes standing, walking, and running.<br />

not as stable as expected, or it is disturbed by external forces, this will be indicated by<br />

raising target ratings <strong>of</strong> the angle adjustments reflexes or by increasing accelerations and<br />

velocities. In this case the stable standing spg changes back to the Stabilization phase to<br />

regain a balanced stance.<br />

5.4 Dynamic <strong>Walking</strong><br />

While stable standing presents a rather static balancing problem, walking is a highly<br />

dynamic motion. The rest <strong>of</strong> this chapter will present a control network for dynamic<br />

walking <strong>of</strong> a human-<strong>like</strong> bipedal robot. First, the transition <strong>of</strong> standing to walking will be<br />

described. In this context, a detailed example <strong>of</strong> how to derive a motor pattern from gait<br />

analysis data will be given. The succeeding main part will introduce the control <strong>of</strong> cyclic<br />

dynamic walking. Its spinal pattern generator determines the phases <strong>of</strong> walking and their<br />

transition conditions. Then the feed-forward and feedback control units <strong>of</strong> each motion<br />

phase will be presented, followed by a discussion <strong>of</strong> the postural reflexes.<br />

Throughout the following pages, the behavior <strong>of</strong> some control units is explained and<br />

illustrated by extracts <strong>of</strong> experimental data. This data is based on the results from the<br />

experiments within the simulation framework introduced in Chapter 6. Basically, the<br />

physical simulation <strong>of</strong> a complex, anthropomorphic bipedal robot following the system<br />

premises stated at the beginning <strong>of</strong> this chapter is used.<br />

5.4.1 Interrelation between Locomotion Modes<br />

Before going into the details <strong>of</strong> the dynamic walking control, the coordination <strong>of</strong> high-level<br />

locomotion modes should be described. Figure 5.10 shows three possible locomotion modes<br />

standing, walking, and running. The latter two are again divided into the cyclic motion<br />

itself, and an initiation or transition mode to switch from standing to walking, and from<br />

walking to running. Switching back to walking or standing is left out for clearer appearance.<br />

The user simply stimulates more and more locomotion modes depending on the desired<br />

form <strong>of</strong> motion. A target velocity and direction is also passed to the control units.


5.4. Dynamic <strong>Walking</strong> 91<br />

(a) (b)<br />

Figure 5.11: (a) Joint angles <strong>of</strong> the stance and swing leg, and (b) movement <strong>of</strong> the center <strong>of</strong> mass<br />

(dotted line) and the center <strong>of</strong> pressure (solid line) during initiation <strong>of</strong> walking. From [Elble 94],<br />

p142f.<br />

During the transition phases, both locomotion modes are activated. As soon as the<br />

transition is finished, the cyclic part <strong>of</strong> the next mode will inhibit the previous one. The<br />

next section will describe this process for the transition from standing to walking in more<br />

detail.<br />

5.4.2 <strong>Walking</strong> Initiation<br />

In many research works on bipedal dynamic walking, the initiation is not considered.<br />

Rather it is assumed that the robot is already moving within the cyclic part <strong>of</strong> the motion.<br />

Here, this assumption is not made. Taking a short look at the initiation <strong>of</strong> human walking,<br />

simple motor patterns are derived to achieve the transition from standing to walking.<br />

5.4.2.1 Gait Initiation in <strong>Human</strong> <strong>Walking</strong><br />

Elble et al. have found a strong stereotypy in the process <strong>of</strong> gait initiation [Elble 94].<br />

Essentially, the center <strong>of</strong> mass is shifted in lateral and anterior direction (Figure 5.11b).<br />

That way, weight is taken from the swing leg, and the body moves forward in preparation<br />

<strong>of</strong> the first step. The forward motion is mainly achieved by the activity <strong>of</strong> the tibialis<br />

anterior, resulting in a dorsiflexion <strong>of</strong> both ankles (Figure 5.11a). A slight flexion <strong>of</strong> the<br />

knee and hip <strong>of</strong> the stance leg can also be observed. This induces a small lateral movement<br />

<strong>of</strong> the trunk in direction <strong>of</strong> the stance leg.<br />

However, the main part <strong>of</strong> lateral weight transfer is initiated by the activity <strong>of</strong> the aband<br />

adductor muscles in the hip joints. This hip muscle function has been investigated<br />

by Kirker et al. during stepping as well as balancing [Kirker 00]. Figure 5.12 plots the<br />

activities <strong>of</strong> the muscle involved when starting to walk with the left leg. Action <strong>of</strong> the<br />

right adductor and the left abductor (gluteus medius) can be observed. This results in the<br />

shearing motion <strong>of</strong> the quadrangle formed by the feet and the two hip joints, which again


92 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

Figure 5.12: Hip muscles activities during gait initiation first stepping with the left leg. The<br />

area marked by the rectangle indicates the initial weight transfer resulting in a lateral torque<br />

and movement <strong>of</strong> the com. From [Kirker 00], p461.<br />

moves the trunk sideways in direction <strong>of</strong> the right stance leg. During this process, muscular<br />

activity in the trunk stabilizes the spine, thus reducing swaying and head movement.<br />

While the center <strong>of</strong> mass only moves anteriorly and in direction <strong>of</strong> the stance leg, the<br />

center <strong>of</strong> pressure follows a more complex trajectory (Figure 5.11b). This is mainly due<br />

to the lateral hip action just described. As the torque in the hip increases, the trunk is<br />

pushed to the side <strong>of</strong> the stance leg. The force that is necessary for this motion connects<br />

to the ground via the swing leg as it pushes downwards during the shearing movement.<br />

Only when the hip torque falls <strong>of</strong>, the weight shifts to the stance leg as the upper body<br />

mass has moved over this leg.<br />

Interestingly enough, in its first half this behavior is counterproductive to the goal <strong>of</strong><br />

transferring the weight onto the stance leg. Only during the second half <strong>of</strong> the initiation<br />

motion, the swing leg is relieved. Consequently, it is not possible to design a simple<br />

controller working on the foot load.<br />

As soon as the walking initiation has finished to shift the weight towards the stance leg,<br />

the activity <strong>of</strong> the swing leg’s adductor and the stance leg’s abductor decreases. In fact,<br />

the direction <strong>of</strong> the hip torque changes to stabilize the pelvis. This is necessary as the foot<br />

leaves the ground and the pelvis can no longer be supported by the swing leg. However,<br />

this behavior is already part <strong>of</strong> the normal cyclic walking phases described later in this<br />

chapter.


5.4. Dynamic <strong>Walking</strong> 93<br />

<strong>Walking</strong><br />

Stable<br />

Standing<br />

Cyclic<br />

<strong>Walking</strong><br />

<strong>Walking</strong><br />

Initiation<br />

<strong>Walking</strong><br />

SPG<br />

Standing<br />

States<br />

...<br />

...<br />

Initiation<br />

Phase L<br />

Initiation<br />

Phase R<br />

Lower Trunk Group<br />

Lateral<br />

Shift R<br />

Lateral<br />

Shift L<br />

Leg Groups<br />

Lean<br />

Forward<br />

Figure 5.13: <strong>Control</strong> network for walking initiation, first step with the left leg.<br />

5.4.2.2 Motor Patterns for <strong>Walking</strong> Initiation<br />

The muscle activity observed during human walking initiation can be translated to motor<br />

patterns. Following the principle <strong>of</strong> control unit distribution and spatially related action,<br />

the muscle activity is allocated to several motor pattern units. As mentioned before, the<br />

local units are distributed in four joint groups representing the upper trunk, the lower<br />

trunk, and each leg. The movement in anterior direction origins from the ankle, hence a<br />

motor pattern is added to each leg group and stimulated during walking initiation. As<br />

lateral weight shift results from muscle action in the hips and the spine, a motor pattern<br />

is added to the lower trunk group, also stimulated during walking initiation. Below, this<br />

control unit will be discussed in more detail.<br />

Figure 5.13 presents the resulting control network for the initiation <strong>of</strong> walking. During<br />

initiation, the standing mode is still active and is combined with the control commands<br />

from the initiation units. The walking locomotion mode signals the initiation mode with<br />

which leg to take the first step. This results in the stimulation <strong>of</strong> one <strong>of</strong> the two initiation<br />

phases. As the initiation process consists <strong>of</strong> only one phase per starting leg, no spg is<br />

necessary. The figure illustrates the activation during initiation with the first step done by<br />

the left leg. The Initiation Phase Left unit is stimulated by the <strong>Walking</strong> Initiation mode,<br />

and itself stimulates the motor patterns for leaning forward and for shifting the weight to<br />

the stance leg’s side (Lateral Shift Right). Similarly, walking initiation with the right leg<br />

as first swing leg would also stimulate the Lean Forward units in both leg groups, and the<br />

Lateral Shift Left to transfer the weight to the contralateral side. The <strong>Walking</strong> mode will<br />

stimulate cyclic walking and gradually deactivate the initiation as soon as the weight is<br />

removed from the swing leg. Then, <strong>Walking</strong> Cyclic will inhibit the standing mode, and<br />

normal walking as described below will be in progress.<br />

A closer look at the control unit for lateral weight shift reveals its similarity to the muscle<br />

activities observed in human walking initiation. Analysis <strong>of</strong> the emg data during this<br />

phase (Figure 5.12) reveals activity <strong>of</strong> the gluteus medius on the swing leg’s side, and<br />

<strong>of</strong> the adductor muscle group on the contralateral side. Knowing the function <strong>of</strong> these<br />

muscles allows to map this muscle action to the corresponding joint action. Appendix A<br />

includes a list <strong>of</strong> the muscles relevant during human walking along with their function and<br />

abbreviation. The main action <strong>of</strong> the gluteus medius is abduction <strong>of</strong> the thigh, meaning an


94 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

Norm. Torque ˆτ<br />

0.00<br />

−0.05<br />

−0.10<br />

−0.15<br />

Left Hip X<br />

0.0 0.25 0.5 0.75 1.0<br />

Time [s]<br />

0.15<br />

0.10<br />

0.05<br />

Right Hip X<br />

0.00<br />

0.0 0.25 0.5 0.75 1.0<br />

Time [s]<br />

0.00<br />

−0.02<br />

−0.04<br />

−0.06<br />

−0.08<br />

Spine X<br />

0.0 0.25 0.5 0.75 1.0<br />

Time [s]<br />

Figure 5.14: Torque output generated by the Lateral Shift Right motor patterns.<br />

outward rotation <strong>of</strong> the hip joint around the x-axis pointing in walking direction. Similarly,<br />

action <strong>of</strong> the adductor muscles result in an inward rotation <strong>of</strong> the hip joint around the<br />

same axis.<br />

Using this information, the muscle action can then be translated to joint torque patterns.<br />

Figure 5.14 illustrates the torques generated by the muscle pattern for lateral weight shift.<br />

The sigmoid function formulated in Equation 5.7 is parametrized and applied to three<br />

joints with individual factors. The torques generated are normalized torque values ranging<br />

from -1 to 1. A positive sign <strong>of</strong> the normalized torque signifies a flexion or adduction <strong>of</strong> the<br />

joint, a negative sign amounts to extension or abduction. Within a hardware abstraction<br />

layer, these values are multiplied by the maximum torque <strong>of</strong> the joint and the sign is<br />

possibly swapped to obtain the mathematically correct rotation direction. The figure<br />

illustrates a negative torque pattern <strong>of</strong> the left hip resulting in an thigh abduction, and a<br />

simultaneous positive torque pattern <strong>of</strong> the right hip leading to adduction <strong>of</strong> the other<br />

thigh. Additionally, a torque about the x-axis <strong>of</strong> the spine joint is generated to compensate<br />

a sideward bending <strong>of</strong> the upper trunk that would occur due to its mass inertia.<br />

It is important to note that all other joints remain passive during the initiation process.<br />

As the motor patterns just mentioned are superimposed on the control <strong>of</strong> stable standing,<br />

a relatively high stiffness and a fixed equilibrium point is set as command for these joints.<br />

Only a few actively controlled joints and the passive dynamics <strong>of</strong> the system result in the<br />

movement preparing for the first step.<br />

Figure5.15comparesthegroundreactionforcesduringwalkinginitiation<strong>of</strong>humansubjects<br />

with results from simulation measured by simulated force plates located under the left and<br />

right body side. As described above, the center <strong>of</strong> pressure first moves to the swing leg<br />

before the weight is shifted fully to the stance leg. When regarding the individual vertical<br />

forces for each leg, a progress as plotted in Figure 5.15a can be observed [Elble 94]. The<br />

vertical forces measured by the force plates during walking initiation <strong>of</strong> the simulated biped<br />

show the same characteristics (Figure 5.15b). The progress <strong>of</strong> the ground reaction forces<br />

is less smooth compared to human measurements as the hip movements <strong>of</strong> the simulated<br />

robot at the beginning <strong>of</strong> the swing phase are slightly jerky. The whole process <strong>of</strong> walking<br />

initiation takes only about 0.1sec longer than in human subjects.<br />

As soon as the load on the swing leg is sufficiently smaller than the load on the stance leg,<br />

the walking locomotion mode switches from walking initiation to cyclic walking. The spinal<br />

pattern generator described in the next section is initialized at those phases responsible


5.4. Dynamic <strong>Walking</strong> 95<br />

Vertical GRF [N]<br />

Vertical GRF [N]<br />

800<br />

600<br />

400<br />

200<br />

0<br />

800<br />

600<br />

400<br />

200<br />

0<br />

Stance Leg<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

Swing Leg<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

Time [s]<br />

(a)<br />

Vertical GRF [N]<br />

Vertical GRF [N]<br />

800<br />

600<br />

400<br />

200<br />

0<br />

800<br />

600<br />

400<br />

200<br />

0<br />

Stance Leg<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

Swing Leg<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

Time [s]<br />

Figure 5.15: Comparing vertical ground reaction forces <strong>of</strong> human walking initiation (a) with<br />

those produced by the Lateral Shift Right motor pattern in simulated biped (b). <strong>Human</strong> data<br />

reproduced from [Elble 94], p143.<br />

for generating the swing <strong>of</strong> the designated leg and ensuring support <strong>of</strong> the body weight<br />

upon the other leg.<br />

5.4.3 Spinal Pattern Generator for <strong>Walking</strong><br />

After initiation <strong>of</strong> walking, the locomotion mode for cyclic walking will be activated. It<br />

again stimulates the spinal pattern generator for walking which is responsible for the<br />

coordination <strong>of</strong> the different phases during walking. This pattern generator and the<br />

walking phases are derived from the work <strong>of</strong> Vaughan, Ivanenko and others as described in<br />

Section 3.1.4 and Section 3.2.2. Their research results based on emg analysis suggest a<br />

central control unit generating the signals for synergistic muscle activity and reflex action.<br />

The transition from one phase to the next could be triggered in two ways: time-based or<br />

event-based. Due to the high complexity <strong>of</strong> the human nervous system, it cannot clearly<br />

be said how this process works in humans. As presented in Section 3.4 some approaches<br />

utilize artificial neural oscillators to create cyclic motions, thus are based on a fixed timing.<br />

In this work, the phase transition is triggered by sensor events. This complies with the<br />

statement <strong>of</strong> Ivanenko et al. that the components <strong>of</strong> muscle activity during walking are<br />

related to kinematic or kinetic events. The duration <strong>of</strong> each walking phase emerges from<br />

the interaction <strong>of</strong> the dynamic system with the environment.<br />

One advantage <strong>of</strong> this approach is a higher contribution <strong>of</strong> the inherent system dynamics.<br />

By not forcing the robot’s motions into a predefined timing, passive dynamics can be<br />

(b)


96 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

right<br />

leg<br />

left<br />

leg<br />

1 weight acceptance<br />

3<br />

4<br />

2 propulsion<br />

5<br />

3 stab.<br />

4 leg swing<br />

5 heel strike<br />

full contact<br />

toe-<strong>of</strong>f locked knee full contact<br />

toe-<strong>of</strong>f locked knee full contact<br />

Figure 5.16: Phases <strong>of</strong> walking for left and right side and the sensor events triggering the phase<br />

transitions.<br />

exploited more easily. Another benefit can be seen in a potentially higher robustness. As<br />

phase transitions are based on sensor events, less additional mechanisms must be provided<br />

to adapt the control to unexpected disturbances. Rather, the system reacts automatically,<br />

e.g. in case the foot contact after the leg swing occurs a bit earlier or later. A possible<br />

drawback <strong>of</strong> this approach could be that the determining events must be perceived in a<br />

reliable way. For this reason, transition conditions are formulated in a fuzzy and when<br />

possible redundant way. Furthermore the number <strong>of</strong> relevant conditions can be reduced by<br />

bilateral synchronization.<br />

According to the discussed emg analysis, human walking seems to be composed <strong>of</strong> five<br />

phases that can be be associated to kinematic or kinetic events: weight acceptance, loading<br />

or propulsion, trunk stabilization during double support, toe lift-<strong>of</strong>f, and heel strike. Based<br />

on these finding, the spinal pattern generator for walking also stimulates five phases for<br />

both the left and the right side <strong>of</strong> the robot. Figure 5.16 illustrates these phases and the<br />

events triggering the phase transitions. As a reference, snapshots <strong>of</strong> the gait cycle are<br />

depicted on the upper part <strong>of</strong> the figure.<br />

The first phase <strong>of</strong> walking starts after full contact <strong>of</strong> the swing foot, i.e. after heel and toes<br />

touch the ground. The weight <strong>of</strong> the trunk is accepted and then supported by the former<br />

swing leg. During the next phase additionally forward velocity can be inserted into the<br />

system in form <strong>of</strong> propulsion by the stance foot. This process starts as soon as the other<br />

leg finishes its swing and is ready for ground impact. The third phase is characterized by<br />

double support during which stabilization <strong>of</strong> the body posture is facilitated. After toe-<strong>of</strong>f,<br />

the swing phase starts, transferring the leg in front <strong>of</strong> the body. As soon as the swing leg’s<br />

knee is locked or its foot touches the ground, the last phase begins. It is responsible for<br />

managing the foot impact during heel strike and for providing control towards the full<br />

contact <strong>of</strong> the foot.<br />

Figure 5.16 also illustrates the bilateral synchronization <strong>of</strong> the walking phases. As observed<br />

in human walking, phase 1 <strong>of</strong> the one leg and phase 3 <strong>of</strong> the contralateral leg are temporally<br />

coordinated. Thesameholdstruefortheonset<strong>of</strong>phases2and5thattakesplacehalfacycle<br />

1<br />

2


5.4. Dynamic <strong>Walking</strong> 97<br />

<strong>Walking</strong><br />

Stable<br />

Standing<br />

<strong>Walking</strong><br />

Initiation<br />

Cyclic<br />

<strong>Walking</strong><br />

Standing<br />

States<br />

...<br />

...<br />

<strong>Walking</strong><br />

SPG<br />

1<br />

2<br />

Weight<br />

Weight<br />

Accept.<br />

1 Accept.<br />

ProPropulsion<br />

2 pulsion<br />

Trunk<br />

3<br />

Trunk<br />

Stabiliz. 3 Stabiliz.<br />

Leg<br />

Leg<br />

4<br />

4<br />

Swing<br />

Swing<br />

5<br />

left right<br />

Heel<br />

Heel<br />

5Strike<br />

Strike<br />

Figure 5.17: Stimulation flow <strong>of</strong> locomotion modes, spinal pattern generator, and walking<br />

phases during cyclic walking.<br />

later. Only the fourth phase associated with the leg swing begins without a corresponding<br />

phase <strong>of</strong> the other leg. This bilateral synchronization guarantees coordinated movement <strong>of</strong><br />

both legs. It further reduces the number <strong>of</strong> conditions relevant for phase transitions to<br />

three, thus increasing the reliability <strong>of</strong> the pattern generator.<br />

Considering this, the spinal pattern generator can be represented by a state machine <strong>of</strong><br />

three states for the walking phases and a second state machine <strong>of</strong> two states switching<br />

the assignment <strong>of</strong> the phases to the body sides. State one corresponds to phase 1 and the<br />

contralateral phase 3, the next state to phase 1 and phase 4, and the third state to phase<br />

2 and phase 5. Each time one cycle <strong>of</strong> this state machine is completed, the assignment <strong>of</strong><br />

the phases to the left and the right side <strong>of</strong> the body is toggled.<br />

The stimulation flow <strong>of</strong> the locomotion modes, the spinal pattern generator, and the<br />

walking phases during cyclic walking is visualized in Figure 5.17. The locomotion mode for<br />

cyclic walking is inhibiting the standing mode and stimulating the spinal pattern generator.<br />

Depending on its state, the pattern generator activates a matching pair <strong>of</strong> walking phases,<br />

one for the left and one for the right body side, each being equipped with its own set <strong>of</strong><br />

the five phases. The figure illustrates the stimulation <strong>of</strong> phase 1 for the left and phase 3<br />

for the right side <strong>of</strong> the body. This pairwise activation <strong>of</strong> phases progresses as the relevant<br />

sensor events described above are perceived.<br />

Each active walking phase is responsible for stimulating the appropriate control units and<br />

for managing the joints being passive during the phase. The passive joints have a constant<br />

stiffness and equilibrium point set throughout the phase. In case a joint or the action <strong>of</strong> a<br />

control unit cannot unambiguously be associated to one body side, it will be controlled by<br />

the phase that is semantically closer related to it. The next sections will name the passive<br />

joints and describe the active control units for each walking phase in detail.<br />

5.4.4 <strong>Walking</strong> Phase 1: Weight Acceptance<br />

Phase 1 <strong>of</strong> cyclic walking begins as soon as the foot <strong>of</strong> the swing leg fully touches the<br />

ground. As in human walking, this phase is responsible for accepting the body weight<br />

with the new stance leg, generating the semi-flat walking gait and shifting the task <strong>of</strong>


98 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

<strong>Walking</strong><br />

SPG<br />

1<br />

2<br />

3<br />

4<br />

5<br />

left<br />

Weight<br />

Accept.<br />

Propulsion<br />

Trunk<br />

Stabiliz.<br />

Leg<br />

Swing<br />

Heel<br />

Strike<br />

Forward<br />

Velocit.L<br />

Upright<br />

Trunk L<br />

Lat. Bal.<br />

Ankle L<br />

Lower Trunk<br />

Stabilize<br />

Pelvis L<br />

Left Leg<br />

Weight<br />

Accept.<br />

Figure 5.18: <strong>Control</strong> units stimulated during phase 1 <strong>of</strong> cyclic walking.<br />

stabilizing the pelvis and the trunk from one leg to the other. Figure 5.18 illustrates the<br />

control units being stimulated in this phase for the left body side. The locomotion modes<br />

and the control units <strong>of</strong> the right body side are omitted for sake <strong>of</strong> clarity and readability.<br />

A trailing ‘L’ or ‘R’ in the unit’s name denotes that it is responsible for action on the left<br />

or right body side. The postural reflexes that aim for suitable forward velocity, an upright<br />

trunk, and lateral stabilization will be discussed separately in Section 5.4.9. Thus, the<br />

passive joints and two control units called Stabilize Pelvis and Weight Acceptance remain<br />

to be described.<br />

The motor pattern Weight Acceptance controls the knee joint to guarantee a smooth weight<br />

acceptance and a semi-flat gait during and after heel strike. This is achieved by actively<br />

bending the knee, already starting in walking phase 5, i.e. just before the foot touches the<br />

ground (Figure 5.19a). Similar to what can be observed in human gait analysis, the leg is<br />

bent preventively even before ground contact, thus reducing impact energy. The motor<br />

pattern receives the desired walking velocity as modification signal. The faster the walking<br />

speed, the stronger the knee is bent. The duration <strong>of</strong> the pattern is set to last until the<br />

trunk has traveled beyond the foot <strong>of</strong> the stance leg in a normal stride. After the pattern<br />

has ended, the knee joint remains in a mostly extended position with high stiffness. As a<br />

result, only little energy must be invested despite <strong>of</strong> the knee having to support the whole<br />

body weight. Figure 5.19b illustrates the resulting semi-flat gait by plotting the z-position<br />

<strong>of</strong> the head’s center. Close to what is found in human walking, the height <strong>of</strong> the head<br />

oscillates only a few centimeters at twice the frequency <strong>of</strong> the gait cycle.<br />

To stabilize the pelvis via the stance leg, a second control unit is stimulated during walking<br />

phase 1. As the pelvis would rotate about the x-axis <strong>of</strong> the hip joint as soon as the swing<br />

leg leaves the ground, this joint must be kept stiff. This corresponds to the activity <strong>of</strong> the<br />

abductor muscles <strong>of</strong> the stance leg observed in human walking where the pelvis is kept<br />

almost horizontal throughout the walking cycle.<br />

Unfortunately, it is not possible to stabilize the pelvis by simply setting a constant joint<br />

angle during the whole stance phase. As the position <strong>of</strong> the stance leg’s foot in lateral<br />

direction will shift depending on disturbances or during walking sidewards, another control<br />

strategy has to be found. Two possible ways come into mind and have been examined


5.4. Dynamic <strong>Walking</strong> 99<br />

Left Knee Angle [ ◦ ]<br />

60<br />

40<br />

20<br />

Heel Strike<br />

0<br />

0.0 0.5 1.0 1.5 2.0 2.5<br />

Time [s]<br />

(a)<br />

Head Height [m]<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

0.0 1.0 2.0<br />

Time [s]<br />

3.0 4.0<br />

Figure 5.19: (a) The Weight Acceptance motor pattern starts to bend the knee before heel<br />

strike to reduce impact. (b) The height <strong>of</strong> the head’s center illustrates the semi-flat walking gait.<br />

in the scope <strong>of</strong> this work. The first possibility assumes that the hip angle in the x-axis<br />

does not have to change during stance phase. Then the Stabilize Pelvis reflex can stiffen<br />

the joint at its current angle as soon as the leg touches the ground. To gain additionally<br />

ground clearance for the swing leg, a small <strong>of</strong>fset can be added to the hip angle, raising<br />

the contralateral side <strong>of</strong> the pelvis.<br />

A second strategy to stabilize the pelvis can be implemented by actually controlling the roll<br />

angle <strong>of</strong> the pelvis. Assuming the pose <strong>of</strong> the trunk is known, this roll angle can be derived<br />

and adjusted by changing the x-angle <strong>of</strong> the hip joint. This method has the advantage <strong>of</strong><br />

being more flexible, as it keeps an appropriate pelvis orientation no matter what situation<br />

occurs. But as this reflex needs to hold a stiff position during strong disturbances and<br />

relatively high cycle time, it easily inclines to oscillate.<br />

In addition to the joints controlled by the two units just described, the postural reflexes<br />

actively influence the ankle pitch and roll angles, and the hip roll angle. The remaining<br />

joints, namely the hip joint about the vertical axis, the spine joints, and the arm joints<br />

are passive and set to a medium, low, and high compliance, respectively. The equilibrium<br />

points <strong>of</strong> these joints are set to their neutral position, correlating to an upright, relaxed<br />

standing posture.<br />

5.4.5 <strong>Walking</strong> Phase 2: Propulsion<br />

The next walking phase starts as soon as the contralateral leg has finished the swing phase<br />

and has stretched the knee in preparation to heel strike. The main task <strong>of</strong> this phase is to<br />

insert additional energy to the system so the forward velocity can be maintained. As the<br />

body mass has already traveled beyond the stance leg, this propulsion can be achieved by<br />

plantarflexion <strong>of</strong> the ankle. Furthermore, the leg swing is already initialized by active leg<br />

flexion in the hip joint. The Stabilize Pelvis control unit as well as the postural reflexes<br />

from phase 1 are kept stimulated. Figure 5.20 illustrates the resulting control network<br />

during phase 2.<br />

The motor pattern Leg Propel creates a plantarflexion <strong>of</strong> the ankle. Thus it corresponds to<br />

the strong muscular activity <strong>of</strong> the gastrocnemius and soleus before toe-<strong>of</strong>f. As in human<br />

walking, the amplitude <strong>of</strong> the motor pattern increases with the walking velocity that it<br />

receives as modulation signal. Additionally, the motor pattern output is scaled by a factor<br />

(b)


100 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

<strong>Walking</strong><br />

SPG<br />

1<br />

2<br />

3<br />

4<br />

5<br />

left<br />

Weight<br />

Accept.<br />

Propulsion<br />

Trunk<br />

Stabiliz.<br />

Leg<br />

Swing<br />

Heel<br />

Strike<br />

Forward<br />

Velocit.L<br />

Upright<br />

Trunk L<br />

Lat. Bal.<br />

Ankle L<br />

Lower Trunk<br />

Stabilize<br />

Pelvis L<br />

Initialize<br />

Swing L<br />

Left Leg<br />

Leg<br />

Propel<br />

Figure 5.20: <strong>Control</strong> units stimulated during phase 2 <strong>of</strong> cyclic walking.<br />

determined by the ankle angle at the beginning <strong>of</strong> the pattern. Advancing dorsiflexion <strong>of</strong><br />

the ankle results in an amplification <strong>of</strong> the motor pattern torque. As a result, a longer step<br />

size automatically leads to stronger propulsion by the stance leg. To prevent excessive<br />

plantarflexion, the output <strong>of</strong> the motor pattern is reduced by the increasing pitch angle <strong>of</strong><br />

the ankle.<br />

Gait analysis <strong>of</strong> human walking shows that leg swing not only emerges passively by the<br />

weight <strong>of</strong> the leg segments after toe-<strong>of</strong>f. Rather, energy is added by the hip joint to actively<br />

start the leg swing and to increase the leg’s velocity even before the foot leaves the ground.<br />

The more energy is added, the more the swing duration can be shortened, thus allowing<br />

for higher walking speeds. The motor pattern Initialize Swing corresponds to this muscle<br />

activity generated by the iliopsoas, sartorius, and tensor fascia latae. As in the Leg Propel<br />

control unit, its amplitude is modulated by the desired walking velocity. The duration <strong>of</strong><br />

the pattern is chosen to last well into walking phase 4.<br />

The torque induced in the hip joint does not only cause a flexion <strong>of</strong> the leg, but also bends<br />

the trunk forward. This trunk movement would be corrected by the postural reflexes, but<br />

only after a deflection has occurred which is big enough to trigger a reflexive response. By<br />

then, the trunk already has started to sway. To prevent this outcome, the motor pattern<br />

additional generates a torque in the contralateral hip. Properly dimensioned, this torque<br />

compensates the unwanted trunk deflection. A corresponding muscle activity can also be<br />

observed in human walking.<br />

The leg flexion generated by Initialize Swing induces a flexion <strong>of</strong> the knee that is set to high<br />

compliance during walking phase 2. This would lead to contra-productive shortening <strong>of</strong> the<br />

stance leg which still is responsible for supporting the body weight. Only in combination<br />

with the plantarflexion <strong>of</strong> the ankle produced by the Leg Propel motor pattern, body<br />

support and propulsion is possible. This synergistic effect is created by the synchronous<br />

stimulation <strong>of</strong> the two control units, coordinated by the motion phase unit.<br />

As already mentioned, the knee joint acts passively with high compliance during this phase.<br />

The other passive joints are similar to those <strong>of</strong> the previous phase.


5.4. Dynamic <strong>Walking</strong> 101<br />

<strong>Walking</strong><br />

SPG<br />

1<br />

2<br />

3<br />

4<br />

5<br />

left<br />

Weight<br />

Accept.<br />

Propulsion<br />

Trunk<br />

Stabiliz.<br />

Leg<br />

Swing<br />

Heel<br />

Strike<br />

Upright<br />

Trunk L<br />

Lat. Bal.<br />

Ankle L<br />

Upper Trunk<br />

Arm<br />

Swing L<br />

Lower Trunk<br />

Stabilize<br />

Pelvis L<br />

Initialize<br />

Swing L<br />

Lateral<br />

Shift R<br />

Left Leg<br />

Leg<br />

Propel<br />

Figure 5.21: <strong>Control</strong> units stimulated during phase 3 <strong>of</strong> cyclic walking.<br />

5.4.6 <strong>Walking</strong> Phase 3: Stabilization<br />

During the third phase <strong>of</strong> walking the robot’s posture can be stabilized as both feet touch<br />

the ground. It begins as soon as the foot <strong>of</strong> the contralateral leg has full contact with the<br />

ground after its swing phase. As the stance phase is nearly finished, the postural reflexes<br />

are gradually diminished in their stimulation. The forward velocity is now controlled by<br />

the former swing leg, and lateral balancing by the ankle joint becomes more difficult as the<br />

ground contact lessens. For similar reasons, the reflex stabilizing the pelvis is only partly<br />

stimulated. Still, the Leg Propel motor pattern pushes the body forward until the foot<br />

leaves the ground. The Initialize Swing pattern started in the previous phase continuous<br />

its action, too.<br />

As illustrated in Figure 5.21, two additional control units are stimulated in phase 3: the<br />

Lateral Shift Right pattern already introduced during walking initiation is stimulated<br />

depending on the body posture. This allows to direct the movement <strong>of</strong> the center <strong>of</strong> mass<br />

to the next stance leg. Additionally, the Arm Swing motor pattern controls the rotation<br />

about the y-axis <strong>of</strong> the ipsi- and contralateral shoulder joint. By adding contrary torque<br />

impulses, a synchronized arm flexion respectively extension is generated. As in human<br />

walking, this arm swing compensates possible trunk rotation disturbing the walking cycle.<br />

The remaining arm joints stay passive.<br />

Figure 5.22 visualizes the torques generated by the Arm Swing motor pattern and the<br />

resulting trajectories. The upper plots represent the y-axis <strong>of</strong> left, the lower plots the<br />

y-axis <strong>of</strong> the right shoulder joint. On the left side, the effective motor torques illustrate<br />

the synchronous but opposite action <strong>of</strong> both shoulder joints. This results in sine-<strong>like</strong> angle<br />

trajectories <strong>of</strong> the swinging arms despite the fact that only short torque impulses are<br />

created. This behavior is possible as the shoulder joints can move passively throughout<br />

the rest <strong>of</strong> the walking cycle.


102 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

Left Shoulder<br />

Right Shoulder<br />

Torque [Nm]<br />

Torque [Nm]<br />

10<br />

0<br />

−10<br />

10<br />

0<br />

−10<br />

0.0 0.5 1.0 1.5 2.0 2.5<br />

0.0 0.5 1.0 1.5 2.0 2.5<br />

Time [s]<br />

Angle [ ◦ ]<br />

Angle [ ◦ ]<br />

40<br />

20<br />

0<br />

−20<br />

−40<br />

0.0 0.5 1.0 1.5 2.0 2.5<br />

40<br />

20<br />

0<br />

−20<br />

−40<br />

0.0 0.5 1.0 1.5 2.0 2.5<br />

Time [s]<br />

Figure 5.22: Angle and torque progress <strong>of</strong> the arm swing generated by the Arm Swing motor<br />

pattern.<br />

5.4.7 <strong>Walking</strong> Phase 4: Leg Swing<br />

The fourth phase <strong>of</strong> walking covers the actual swing <strong>of</strong> the leg. It begins as soon as the<br />

foot has left the ground and ends when the knee <strong>of</strong> the swing leg is stretched again. The<br />

task <strong>of</strong> this phase is tw<strong>of</strong>old: the swing leg has to be brought forth to accept the body<br />

weight while ensuring ground clearance <strong>of</strong> the foot; and the foot must be positioned in a<br />

way that postural stability can by retained. The latter function is performed by the local<br />

reflex Lock Hip and the postural reflex Lateral Foot Placement which will be described<br />

later on. Active control <strong>of</strong> the upper body’s posture is not possible as the leg does not<br />

touch the ground and no forces can be induced into the trunk.<br />

As shown in Figure 5.23, the Arm Swing and Initialize Swing motor patterns continue<br />

their action and three new reflexes are introduced. The Lock Knee reflex is responsible for<br />

ensuring the knee is stretched at the end <strong>of</strong> the leg swing. This is necessary as only an<br />

extended leg can create the pendulum-<strong>like</strong> walking gait aimed for. A bended leg would<br />

allow the center <strong>of</strong> mass to fall too low, and additional joint work would be required to lift<br />

the trunk again.<br />

Before the knee can be stretched to prepare it for weight acceptance, it must bend to<br />

guarantee ground clearance <strong>of</strong> the foot while it travels forward. In human walking, no<br />

muscle action affecting the knee joint can be observed during the main part <strong>of</strong> the leg<br />

swing. Rather, the knee bends passively due to mass inertia <strong>of</strong> the lower leg and the foot.<br />

The same holds true for this approach: during phase 4, the knee joint is allowed to move<br />

freely until the end <strong>of</strong> the swing when the Lock Knee reflex reacts. It bends passively as<br />

an effect <strong>of</strong> the inherent dynamics.<br />

In case the momentum <strong>of</strong> the lower leg is high enough, the knee would extend by itself and<br />

the Lock Knee reflex would not be needed. But if the momentum is too low, active control


5.4. Dynamic <strong>Walking</strong> 103<br />

<strong>Walking</strong><br />

SPG<br />

1<br />

2<br />

3<br />

4<br />

5<br />

left<br />

Weight<br />

Accept.<br />

Propulsion<br />

Trunk<br />

Stabiliz.<br />

Leg<br />

Swing<br />

Heel<br />

Strike<br />

Lat. Foot<br />

Placem.L<br />

Upper Trunk<br />

Arm<br />

Swing L<br />

Lower Trunk<br />

Lock<br />

Hip L<br />

Initialize<br />

Swing L<br />

Left Leg<br />

Lock<br />

Knee<br />

Cutan.<br />

Reflex<br />

Figure 5.23: <strong>Control</strong> units stimulated during phase 4 <strong>of</strong> cyclic walking.<br />

<strong>of</strong> knee joint will be necessary. This control takes place as soon as the angle <strong>of</strong> the knee<br />

joint draws near to the stretched state. The reflex then starts to increase the stiffness <strong>of</strong><br />

the knee joint with the equilibrium point set to the stretched angle. If the leg extends by<br />

itself, no additional energy will be expended. Otherwise, a torque will be generated by the<br />

knee motor to support the leg extension.<br />

Figure 5.24 shows the knee angle emerging during a step <strong>of</strong> the simulated biped and<br />

compares it to human walking. It can be seen that similar trajectories arise. The knee<br />

passively bends to an angle <strong>of</strong> about 60 ◦ during the swing phase.<br />

The Lock Hip reflex controls the flexion <strong>of</strong> the hip joint, i.e. the rotation about the y-axis<br />

by applying a damping torque. It aims at stopping the swing <strong>of</strong> the leg to achieve a<br />

suitable angle <strong>of</strong> attack at heel strike. This follows the ideas <strong>of</strong> Blum, Seyfarth et al. who<br />

show a correlation <strong>of</strong> the swing leg’s angle <strong>of</strong> attack and a stable forward velocity in human<br />

walking and running [Blum 07]. Varying step length at the same kinetic forward energy <strong>of</strong><br />

the trunk influences the forward velocity. Longer steps result in slowing down, shorter<br />

steps in accelerating the upper body.<br />

To set a suitable angle <strong>of</strong> attack, the Lock Hip (lh) reflex has to know the pitch angle φ <strong>of</strong><br />

the trunk. It also receives the desired walking velocity. The resulting normalized target<br />

hip angle ˆ φhip y, lh amounts to<br />

ˆαhip y, lh = ˆαhip y, max ·vtarget +Kpitch ·φ (5.8)<br />

where the maximum hip angle ˆαhip y, max and the body pitch factor Kpitch are set experimentally.<br />

This calculation results in shorter steps when the body leans backwards and<br />

longer steps when the body leans forward. The more the hip flexion <strong>of</strong> the swing leg<br />

approaches the target angle, the higher the applied damping. The damping torque is<br />

proportional to the rotational velocity <strong>of</strong> the hip joint, thus falling to zero as soon as the


104 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

Knee Angle [ ◦ ]<br />

60<br />

40<br />

20<br />

<strong>Human</strong> <strong>Walking</strong><br />

0<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

(a)<br />

Knee Angle [ ◦ ]<br />

60<br />

40<br />

20<br />

Simulated Biped<br />

0<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

Figure 5.24: Comparison <strong>of</strong> the knee angle trajectory in human walking (a) and in the biped<br />

simulation, averaged over 20 steps (b). <strong>Human</strong> data reproduced from [Hamill 03].<br />

leg swing is stopped. Then the reflex stimulates the position control to fix the hip joint at<br />

the desired angle <strong>of</strong> attack.<br />

The Cutaneous Reflex is motivated by its namesake discussed in Chapter 3. Its task is to<br />

re-establish ground clearance in case the foot touches down during the swing phase. To<br />

this end, it receives the force measurements from the toe and heel sensors. In case the<br />

toe sensors indicate ground contact, ankle dorsiflexion is generated by an ankle torque<br />

proportional to the measured force. Similar, ankle plantarflexion is initiated as soon as heel<br />

contact is conceived. In both cases, an accompanying torque to bend the knee supports<br />

the necessary ground clearing reaction.<br />

As long as the cutaneous reflex does not become active, the knee joint acts purely passively<br />

during this motion phase. Furthermore, both ankle joints are set to a fixed, neutral<br />

equilibrium point with medium compliance. The spring-<strong>like</strong> behavior <strong>of</strong> the ankle pitch<br />

angle is sufficient to lift the toes <strong>of</strong>f the ground after the Leg Propel motor pattern <strong>of</strong> the<br />

previous phases has finished the plantarflexion <strong>of</strong> the ankle. Hip ab- and adduction is not<br />

passive but controlled by the postural reflex Lateral Foot Placement. The flexion <strong>of</strong> the hip<br />

is passive between the end <strong>of</strong> the Initialize Swing pattern and the action <strong>of</strong> the Lock Hip<br />

reflex. As in the phases before, hip rotation and the spine joints are set to low compliance.<br />

5.4.8 <strong>Walking</strong> Phase 5: Heel Strike<br />

The last phase <strong>of</strong> walking takes place around the heel strike event. As soon as the swing<br />

leg is extended, it must prepare for touching the ground. This phase is responsible for<br />

reducing the ground impact and for generating a controlled lowering <strong>of</strong> the toes after heel<br />

strike. Figure 5.25 illustrates the involved control units. Compared to the previous phase,<br />

the knee joint is no more controlled by the Lock Knee reflex but by the Weight Acceptance<br />

motor pattern already described above. Additionally, the Heel Strike reflex controls the<br />

ankle dorsiflexion. <strong>Control</strong> <strong>of</strong> the posture and <strong>of</strong> the leg’s angle <strong>of</strong> attack remain the same.<br />

Due to the body weight resting on the foot after touch-down, the ankle joint would<br />

plantarflex and the foot would rotate about the heel contact point. In human walking,<br />

activity <strong>of</strong> the tibialis anterior muscle prevents the toes from hitting the ground too<br />

(b)


5.4. Dynamic <strong>Walking</strong> 105<br />

<strong>Walking</strong><br />

SPG<br />

1<br />

2<br />

3<br />

4<br />

5<br />

left<br />

Weight<br />

Accept.<br />

Propulsion<br />

Trunk<br />

Stabiliz.<br />

Leg<br />

Swing<br />

Heel<br />

Strike<br />

Lat. Foot<br />

Placem.L<br />

Lower Trunk<br />

Lock<br />

Hip L<br />

Left Leg<br />

Heel<br />

Strike<br />

Weight<br />

Accept.<br />

Figure 5.25: <strong>Control</strong> units stimulated during phase 5 <strong>of</strong> cyclic walking.<br />

forcefully. Here, the Heel Strike reflex assumes this task. As a force is measured by the heel<br />

force sensors <strong>of</strong> the foot, a damping torque is produced to s<strong>of</strong>ten plantarflexion. The torque<br />

is proportional to the rotational velocity <strong>of</strong> the ankle joint. In contrast to the biological<br />

ankle joint, technical systems already introduce joint damping, e.g. by gear friction. The<br />

simulation used in this work also includes damping within the joints. Depending on the<br />

amount <strong>of</strong> damping that acts within the ankle joint, the Heel Strike reflex can be omitted.<br />

Otherwise, the ankle joint remains in a passive, relaxed state as no other control unit<br />

affects its action. Again, hip rotation and the spine joints are set to low compliance at a<br />

neural equilibrium position.<br />

5.4.9 Posture <strong>Control</strong><br />

Postural reflexes provide additional body stabilization based on global sensor information.<br />

The main source <strong>of</strong> information on the current body pose is provided by the inertial<br />

measurement unit. It is able to measure accelerations in all three directions as well as<br />

three rotational velocities. In this work it is assumed that the upper body orientation can<br />

be derived from this data. While this assumption might not be easily implemented, it<br />

would go beyond the scope <strong>of</strong> this thesis. In literature, a few approaches can be found on<br />

deriving the orientation based on an inertial measurement unit and certain heuristics or<br />

additional model knowledge [Koch 04, Gienger 02].<br />

The main challenges <strong>of</strong> postural control are threefold: the upper body should be kept erect<br />

or slightly bend forward depending on the walking velocity; the forward velocity must be<br />

controlled to keep the robot from stumbling or falling forward or backwards; finally, lateral<br />

stability must be ensured.<br />

To a certain degree, passive dynamics and compensating motor patterns work towards these<br />

challenges. For instance, parallel elasticities in the ankle joint help in lateral stabilization<br />

and in controlling the forward velocity: larger steps result in stronger dorsiflexion during<br />

push-<strong>of</strong>f, thus stretching the ankle elasticities and adding more energy to the system.<br />

The motor pattern for initiating the leg swing also inserts an opposing torque in the<br />

contralateral hip to stabilize the trunk. However, postural reflexes still need to generate<br />

correcting motions to compensate self-induced or external disturbances.


106 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

φtarget<br />

−<br />

eφ<br />

˙φtarget e ˙ φ ˆτhip<br />

P PI<br />

φ<br />

−<br />

˙φ<br />

Process<br />

Figure 5.26: <strong>Control</strong> block diagram <strong>of</strong> the postural reflex Upright Trunk.<br />

5.4.9.1 Upper Body Balance<br />

In human walking, the upper body is kept remarkably stable. Despite significant accelerations<br />

caused by hip movements, ground impact, or foot push-<strong>of</strong>f, the trunk remains upright.<br />

Especially the head shows only small deflections in the horizontal plane, facilitating the<br />

processing <strong>of</strong> visual information, assisted by the vestibulo-ocular reflex [Ito 82].<br />

The postural reflex Upright Trunk assumes the task <strong>of</strong> upper body stabilization in the<br />

frontal plane. It receives the current estimation on the body pitch as sensor information.<br />

The walking velocity is passed as modulation signal to achieve a forward inclination <strong>of</strong> the<br />

trunk when walking faster.<br />

As described above, there is the need to have one instance <strong>of</strong> the reflex for each body<br />

side to enable specific stimulation during different walking phases. Each <strong>of</strong> the two reflex<br />

instances will generate torque values acting on the hip joint <strong>of</strong> its particular body half.<br />

The activity <strong>of</strong> the reflex is further scaled by the ground contact derived from the force<br />

sensors <strong>of</strong> the respective foot. This guarantees that the torque applied at the hip joint<br />

only affects the orientation <strong>of</strong> the trunk as the leg is securely rooted to the ground.<br />

TheUpright Trunk reflexisdesignedasacascadedposition-velocityloop-backcontroller. As<br />

illustrated in Figure 5.26, the position control is implemented as P-control, the underlying<br />

velocity part as PI-control. The target pitch angle φtarget, i.e. the rotation <strong>of</strong> the trunk<br />

about the y-axis, is slightly increased with the walking velocity to boost the forward<br />

momentum. Depending on the resulting error <strong>of</strong> body pitch, a target change <strong>of</strong> pitch<br />

˙φtarget is calculated and passed to the PI loop. The controller then generates a normalized<br />

torque ˆτhip being applied to the hip joint <strong>of</strong> the corresponding body side.<br />

The action <strong>of</strong> the left Upright Trunk reflex during six consecutive steps and the resulting<br />

posture <strong>of</strong> the upper body is depicted in Figure 5.27. The activity a <strong>of</strong> the reflex illustrates<br />

the dependency <strong>of</strong> its reaction on the ground contact information. During the swing phase,<br />

the reflex shows no activity as it is not stimulated by the corresponding walking phase.<br />

The plotted normalized torque ˆτhip will only be applied to the hip joint if the reflex is<br />

active.<br />

As indicated by the head’s pitch angles, swaying <strong>of</strong> the upper body in the frontal plane<br />

can be kept within a few degrees. The head shows slightly more rotation than the inertial<br />

measurement unit being attached to the pelvis. This is caused by the elastic action <strong>of</strong> the<br />

spine joint. The plotted angles also illustrate that the upper body is leaning forward by<br />

a few degrees during walking. Depending on the aspired walking velocity, this <strong>of</strong>fset is<br />

raised or decreased.


5.4. Dynamic <strong>Walking</strong> 107<br />

Head Pitch Angle [ ◦ ]<br />

IMU Pitch Angle [ ◦ ]<br />

10<br />

0<br />

−10<br />

10<br />

0<br />

−10<br />

0.0 1.0 2.0 3.0<br />

0.0 1.0 2.0<br />

Time [s]<br />

3.0<br />

Reflex Activity a<br />

Norm. Hip Torque ˆτhip<br />

1.0<br />

0.5<br />

0.0<br />

1.0<br />

0.0<br />

−1.0<br />

0.0 1.0 2.0 3.0<br />

0.0 1.0 2.0<br />

Time [s]<br />

3.0<br />

Figure 5.27: The postural reflex Upright Trunk regulates the pitch angle <strong>of</strong> the upper body.<br />

The imu is mounted within the pelvis.<br />

5.4.9.2 Estimation <strong>of</strong> Center <strong>of</strong> Mass Position<br />

Regulation <strong>of</strong> the stability in lateral direction and adjustments <strong>of</strong> the forward velocity<br />

require additional information processing. As discussed in Section 3.2.4, postural control<br />

in human walking could be based on the movement <strong>of</strong> the center <strong>of</strong> mass (com). H<strong>of</strong> et al.<br />

suggest a control strategy for foot placement relying on the extrapolated center <strong>of</strong> mass<br />

(xcom) [H<strong>of</strong> 08]. To adopt a similar approach for the functioning <strong>of</strong> postural reflexes, an<br />

estimation <strong>of</strong> the current position <strong>of</strong> the com becomes necessary.<br />

The com calculations used in this work only provide a coarse estimation <strong>of</strong> its position. It is<br />

assumed that there is no elaborate dynamic model coded within the human cerebellum, but<br />

rather a simplified model that is sufficient for postural regulation. Figure 5.28 illustrates<br />

the approach <strong>of</strong> a reduced model <strong>of</strong> the com for one vertical plane. It disregards all<br />

motions <strong>of</strong> the upper body <strong>like</strong> e.g. arm swing that are changing the mass distribution<br />

and are introducing additional dynamic effects. Variation <strong>of</strong> the stance leg’s length is also<br />

neglected. The calculations are equivalent for the frontal and the sagittal plane, so the<br />

following considerations can be limited to two dimensions.<br />

The center <strong>of</strong> mass is assumed to be located at a constant distance l1 from the hip joint,<br />

aligned by an angle δ. These variables can be derived from the horizontal and vertical<br />

shift ∆com <strong>of</strong> the com from the hip joint as:<br />

l1 =<br />

�<br />

∆comhoriz 2 +∆comvert 2 , (5.9)<br />

tanδ = ∆comvert<br />

∆comhoriz<br />

. (5.10)<br />

The rotational deflection φ <strong>of</strong> the pelvis from the neutral vertical position is provided by<br />

the pose estimation <strong>of</strong> the imu. The stance leg <strong>of</strong> length l2 is rotated by the known angle


108 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

φ<br />

CoM l1<br />

l3<br />

δ<br />

d<br />

β<br />

α<br />

β<br />

γ<br />

l2<br />

β +δ + π<br />

2<br />

Figure 5.28: Simplified model for the estimation <strong>of</strong> the center <strong>of</strong> mass position.<br />

β about the hip joint. From these values, the distance l3 <strong>of</strong> the com from the foot contact<br />

point can be calculated as<br />

�<br />

l3 = l1 2 +l2 2 �<br />

−2l1l2cos β +δ + π<br />

�<br />

. (5.11)<br />

2<br />

To derive the wanted distance d <strong>of</strong> the projection <strong>of</strong> the com from the foot contact point,<br />

the angle α = β +γ +φ is necessary. With γ and φ already known, the complementing<br />

angle γ can be found as<br />

tanγ = l1sin � β +δ + π<br />

�<br />

2<br />

l2 −l1cos � β +δ + π<br />

�. (5.12)<br />

2<br />

This results in the distance d being<br />

d = l3sinα. (5.13)<br />

These calculations are done for both the left and right leg and for the frontal and sagittal<br />

plane, resulting in four values comx,left, comx,right, comy,left, and comy,right. They provide<br />

an estimation <strong>of</strong> the projection <strong>of</strong> the static com position to the ground plane relative to<br />

the position <strong>of</strong> the left and right ankle joint. Following H<strong>of</strong>’s suggestions, the extrapolated<br />

center <strong>of</strong> mass (xcom) can be derived from these values by including the velocity ˙ d <strong>of</strong> the<br />

respective com projection normalized by the eigen frequency ω0 <strong>of</strong> the assumed inverted<br />

pendulum model:<br />

xcom = d+ ˙ d<br />

(5.14)<br />

ω0<br />

with the eigen frequency ω0 for the pendulum length l3 <strong>of</strong> the com above the the foot<br />

point given as<br />

�<br />

g<br />

ω0 = . (5.15)<br />

l3


5.4. Dynamic <strong>Walking</strong> 109<br />

comx [m]<br />

comy [m]<br />

comy,fusion [m]<br />

1.0<br />

0.0<br />

−1.0<br />

0.2<br />

0.0<br />

−0.2<br />

0.2<br />

0.0<br />

−0.2<br />

0.0 1.0 2.0 3.0 4.0 5.0<br />

0.0 1.0 2.0 3.0 4.0 5.0<br />

0.0 1.0 2.0 3.0 4.0 5.0<br />

Time [s]<br />

xcomx [m]<br />

xcomy [m]<br />

xcomy,fusion [m]<br />

1.0<br />

0.0<br />

−1.0<br />

0.2<br />

0.0<br />

−0.2<br />

0.2<br />

0.0<br />

−0.2<br />

0.0 1.0 2.0 3.0 4.0 5.0<br />

0.0 1.0 2.0 3.0 4.0 5.0<br />

0.0 1.0 2.0 3.0 4.0 5.0<br />

Time [s]<br />

Figure 5.29: Estimation <strong>of</strong> the com movement. Solid lines represent the left side, dashed line<br />

the right side. The lower two plots show merged (x)com values in y-direction weighted by ground<br />

contact.<br />

Again, equation 5.14 can then be applied for both legs and directions resulting in the four<br />

values xcomx,left, xcomx,right, xcomy,left, and xcomy,right. The extrapolated interpretation<br />

<strong>of</strong> the com projection can be used in case the static calculations are insufficient for<br />

maintaining stability.<br />

Figure 5.29 illustrates the progression <strong>of</strong> the estimated center <strong>of</strong> mass for undisturbed<br />

walking. The solid lines in the upper two rows represent the com and xcom values relative<br />

to the left leg, the dashed lines those corresponding to the right leg. The phase shift <strong>of</strong><br />

half a cycle between the left and the right side is clearly visible. During stance phase, the<br />

comx value increases as the center <strong>of</strong> mass swings over the supporting leg. It decreases<br />

again as the leg swings forward. The absolute value <strong>of</strong> comy has its minimum during the<br />

stance phase as the body mass sways in the direction <strong>of</strong> the support leg. During leg swing,<br />

the value is <strong>of</strong> only minor significance as it depends on the lateral foot position controlled<br />

by the related postural reflex.<br />

The extrapolated com values on the right side <strong>of</strong> the plot reflect the integration <strong>of</strong> the<br />

com’s velocity. Both directions exhibit larger amplitudes. The additional noise in both<br />

signals origins from the oscillation <strong>of</strong> the upper body that inflicts the measurements <strong>of</strong> the<br />

imu. Both the xcomx and the xcomy values show a negative phase shift compared to their<br />

com counterparts due to their velocity component. Thus their interpretation by a postural


110 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

xcomx [m]<br />

1.0<br />

0.0<br />

−1.0<br />

0.0 1.0 2.0 3.0 4.0 5.0<br />

Time [s]<br />

Correction Value<br />

0.5<br />

0.0<br />

−0.5<br />

0.0 1.0 2.0 3.0 4.0 5.0<br />

Time [s]<br />

Figure 5.30: <strong>Control</strong> <strong>of</strong> forward velocity by approximating the xcomx trajectory during normal<br />

walking and calculating the derivation.<br />

reflex not only allows for integrating dynamic properties but also gives a temporal benefit.<br />

This advantages must be weighted against the drawback <strong>of</strong> additional noise.<br />

The lower two plots <strong>of</strong> Figure 5.29 show merged values <strong>of</strong> the comy and xcomy <strong>of</strong> both<br />

the left and right leg. The fusion is calculated using the normalized ground contact force<br />

ˆFz ∈ [0,1] measured by the load cells as weights:<br />

comy,fusion = comy,left · ˆ Fz,left +comy,right · ˆ Fz,right<br />

ˆFz,left + ˆ Fz,right<br />

(5.16)<br />

This combined value illustrates the lateral sway <strong>of</strong> the robot during normal walking. During<br />

single support the center <strong>of</strong> mass travels to the side <strong>of</strong> the supporting leg. With both legs<br />

on the ground, the merged com value levels out at about zero, as the left and right com<br />

estimations show similar absolute values and both legs have full ground contact. As soon<br />

as the swing leg lifts <strong>of</strong>f, the merged com value again moves to the side <strong>of</strong> the stance leg.<br />

5.4.9.3 <strong>Control</strong>ling the Forward Velocity<br />

The control <strong>of</strong> forward velocity relies on three control units: the local reflex Lock Hip, the<br />

motor pattern Leg Propel, and the postural reflex Forward Velocity, each <strong>of</strong> those existing<br />

for both body sides. The postural adaptation <strong>of</strong> the former control unit has already been<br />

discussed above: the step length is varied depending on the pitch <strong>of</strong> the upper body. The<br />

latter <strong>of</strong> the three control units calculates a correction value for the forward velocity that<br />

is used to generate an suitable ankle torque. It is also passed to the leg propulsion control<br />

unit as modulation value scaling the amplitude <strong>of</strong> the motor pattern.<br />

The computation <strong>of</strong> the correction value relies on the xcomx estimation <strong>of</strong> the respective<br />

leg as shown on the upper right plot in Figure 5.29. The plot depicts the xcomx during<br />

undisturbed, normal walking on level ground. Assuming that a deviation from this<br />

trajectory stems from external or self-induced disturbances, the correction factor is based<br />

on this deviation. Therefore, the xcomx trajectory over the time interval during which<br />

the corresponding leg has ground contact is approximated by the first quarter <strong>of</strong> a sine<br />

function:<br />

xcomx,target = vtarget<br />

�<br />

xcomx,start +(xcomx,end −xcomx,start)sin<br />

�<br />

π<br />

2 t<br />

��<br />

(5.17)<br />

where xcomx,start and xcomx,end denote the interval defining the amplitude, vtarget the<br />

desired walking velocity passed as modulation signal, and t ∈ [0,1] indicating the time


5.4. Dynamic <strong>Walking</strong> 111<br />

comy [m]<br />

0.2<br />

0.1<br />

0.0<br />

0.0 2.0 4.0 6.0<br />

Time [s]<br />

Correction Value<br />

0.1<br />

0.0<br />

−0.1<br />

0.0 2.0 4.0 6.0<br />

Time [s]<br />

Figure 5.31: <strong>Control</strong> <strong>of</strong> lateral stability by approximating the comy trajectory during normal<br />

walking and calculating the derivation.<br />

progression over the average duration <strong>of</strong> the reflex’s activity. Figure 5.30 illustrates the<br />

approximated trajectory and the calculated correction value. Both values are only relevant<br />

during full ground contact <strong>of</strong> the foot as only then the leg can influence the robot’s posture.<br />

Hence the calculations are limited to walking phases one and two, during the other phases<br />

the values are set to zero.<br />

The correction value is then applied as torque to the ankle joint to achieve an appropriate<br />

dorsi- or plantarflexion. The maximum absolute torque is limited to prevent lifting <strong>of</strong> the<br />

heel or toes <strong>of</strong> the foot. In case the forward velocity is too high, i.e. the xcomx value is<br />

greater than the approximated trajectory, plantarflexion <strong>of</strong> the foot will slow down the<br />

pendulum movement <strong>of</strong> the upper body over the stance leg. Likewise a small xcomx value<br />

will result in an ankle dorsiflexion and increase the velocity <strong>of</strong> the upper body.<br />

5.4.9.4 Lateral Stability<br />

Similar to the control <strong>of</strong> forward velocity, lateral stability is improved by foot placement<br />

and by correcting ankle torque. As the xcom estimation in y-direction is strongly influenced<br />

by the lateral sway <strong>of</strong> the upper body and the pelvis rotation during hip stabilization, the<br />

com estimation is utilized to calculate reflex action. Again, the trajectory <strong>of</strong> comy during<br />

normal walking is approximated, in this case by a forth order polynomial:<br />

�<br />

comy,target = vtarget comy,start +(comy,end −comy,start) � 1−(2t−1) 4��<br />

(5.18)<br />

The variables are defined similar to the trajectory approximation for forward velocity<br />

control. Figure 5.31 illustrates the resulting correction value. The dashed line in the left<br />

plot denotes the estimated comy values, the solid line the approximated trajectory.<br />

The postural reflex Lateral Balance Ankle uses the correction value to generate an ankle<br />

torque about the x-axis. Again, the absolute torque values are limited to maintain full<br />

ground contact <strong>of</strong> the foot. The resulting inversion or eversion <strong>of</strong> the foot can only produce<br />

a small correction motion as the leverage <strong>of</strong> toes and heel in y-direction is relatively short.<br />

Nevertheless the reflex provides a small amount <strong>of</strong> lateral balance.<br />

Foot placement has much stronger influence on the movement <strong>of</strong> the center <strong>of</strong> mass in the<br />

sagittal plane. Similar to what is suggested by H<strong>of</strong> et al., the postural reflex Lateral Foot<br />

Placement (lfp) uses the correction value just described to adapt the hip angle during leg<br />

swing in motion phases four and five. As the leg influenced by the reflex is in its swing


112 5. <strong>Control</strong>ling Dynamic Locomotion <strong>of</strong> a Fully Articulated Biped<br />

phase, the com estimation relative to the opposite leg is considered for foot placement<br />

adaptation. The normalized hip angle ˆαhip, lfp passed to the joint controller amounts to<br />

ˆαhip x, lfp = ˆαhip x, normal +g ·σ ·Kcom ·(comy,target −comy) (5.19)<br />

where ˆαhip, normal denotes the hip angle for the undisturbed case, σ the change <strong>of</strong> sign due<br />

to the different directions <strong>of</strong> normalized hip angles <strong>of</strong> the left and right leg, Kcom a factor<br />

to translate the com error to a normalized angle change, and g a gradual activation <strong>of</strong> the<br />

angle correction rising from 0 to 1 to prevent a jerky movement <strong>of</strong> the leg at the start <strong>of</strong><br />

its swing phase. The reflex activity indicating both the behavior influence and desired<br />

joint stiffness quickly rises with an increasing correction value to enforce the adjustment<br />

<strong>of</strong> the foot placement.


6. Implementation <strong>of</strong> <strong>Bipedal</strong><br />

<strong>Walking</strong> <strong>Control</strong><br />

To validate the suggested control concept and its application to dynamic walking and<br />

standing as described in the previous chapters, it is implemented as control system for<br />

a simulated biped. The simulation environment must support the dynamics calculation<br />

<strong>of</strong> the robot and its interaction with the environment. As already mentioned before, the<br />

bipedal robot has to show some properties <strong>of</strong> human morphology, muscles, and perception.<br />

This chapter will introduce the test environment as well as the biped model and its features.<br />

Some aspects <strong>of</strong> the implementation <strong>of</strong> the control system are discussed.<br />

6.1 Simulation Environment<br />

As this thesis focuses on the control <strong>of</strong> bipedal walking, it cannot also provide a concept<br />

for a suitable mechanical construction and actuation system. But while the design <strong>of</strong> a<br />

biped with human-<strong>like</strong> properties is subject to current research activities, there still is<br />

no satisfactory solution for this problem. To nevertheless validate the suggested control<br />

approach, it is implemented and tested within a simulation environment. In doing so, the<br />

essential features for the functioning <strong>of</strong> the biologically inspired control can be provided<br />

without the need <strong>of</strong> a mechanical implementation. Still, attention is paid to the fact that<br />

the simulated biped should potentially be realizable as a prototype. Additional information<br />

on the simulation environment can be found in Appendix B, preliminary studies to this<br />

topic can be found in [Chiderova 05, Lamp 05, Gilsdorf 06].<br />

6.1.1 Embedding the Physics Engine<br />

The control system and simulation <strong>of</strong> this work is implemented using the framework<br />

mca2 (modular controller architecture) 1 [Scholl 02]. Its basic control entities are modules<br />

featuring an interface <strong>of</strong> control and sensor inputs and outputs. Modules are connected<br />

by edges transferring processed data between their interfaces. Additionally, a blackboard<br />

1 The RRLab version <strong>of</strong> mca2 can be downloaded at http://rrlib.cs.uni-kl.de


114 6. Implementation <strong>of</strong> <strong>Bipedal</strong> <strong>Walking</strong> <strong>Control</strong><br />

mechanism allows the exchange <strong>of</strong> larger data objects without the need <strong>of</strong> edges. Modules<br />

can be subsumed in groups that feature the same interface as modules. The interconnection<br />

<strong>of</strong> modules and groups defines a control hierarchy <strong>of</strong> layers and the order in which their<br />

processing methods are called during the control cycle running with a fixed timing.<br />

The behavior-based framework iB2C is also implemented within mca2. Each behavior<br />

is represented by a module, the behavior signals and the input and output vectors are<br />

transferred via edges. mca2 comes with an extensive tool set to support the developing<br />

process. In particular, the behavior-based design is assisted by visualization and debugging<br />

tools, e.g. the mcabrowser allows to show the flow <strong>of</strong> stimulation and activity between<br />

behaviors during runtime.<br />

mca2 already provides a library called SimVis3D for the simulation and visualization <strong>of</strong><br />

robots and their environment [Braun 07]. Its visualization part is based on Coin3D 2 , a free<br />

implementation <strong>of</strong> the OpenInventor standard. SimVis3D features a comfortable way to<br />

define the simulation scene via a description file, dynamic object management, separation<br />

<strong>of</strong> simulation and visualization, and <strong>of</strong>fers support for various sensor systems <strong>like</strong> cameras<br />

or laser range finders.<br />

In the scope <strong>of</strong> this work, SimVis3D has been supplemented with dynamical simulation<br />

capabilities. To achieve this, two possible methods are possible: either the dynamics<br />

equation can be derived manually for each robot, or a physics engine can be used that<br />

automates this process. While the first approach allows to describe more specialized<br />

models, the second approach has been chosen due to its versatility. Furthermore, changes<br />

<strong>of</strong> the model can be transferred straightforward to the simulation without the need to<br />

re-derive the equations <strong>of</strong> motion. The free Newton physics engine 3 used in this work is<br />

developed by Julio Jerez. Its main advantages are easy adaptability by a function callback<br />

mechanism, the possibility to extract various information <strong>like</strong> acting forces or torques,<br />

and a flexible universal joint implementation system. Newton’s penalty-based contact<br />

calculation features adjustable elasticity, s<strong>of</strong>tness, and coulomb-<strong>like</strong> static and kinetic<br />

frictions. These parameters can be defined for each pair <strong>of</strong> materials.<br />

To embed the physics engine into the simulation environment, the so-called elements <strong>of</strong><br />

SimVis3D are used. In contrast to parts that define the visualization <strong>of</strong> objects, elements<br />

provide the means to insert sensors or other specialized entries into the scene description<br />

graph. In the context <strong>of</strong> the physics engine, three new types <strong>of</strong> elements are defined:<br />

static physical objects, dynamic objects, and joints. Physical objects define the association<br />

between the visualization parts and colliding objects within the physical simulation. Static<br />

objects describe the environment, they only need to provide a collision geometry and a<br />

material type, and are assumed to be unmovable and to have infinite mass. Dynamic<br />

objects will move within the environment, and therefore a transformation matrix node is<br />

added to the scene graph. During run-time, this matrix is modified by the physics engine<br />

to set the pose <strong>of</strong> the visual object. Besides their collision geometry, dynamic objects<br />

additionally are defined by a mass and a material changing the contact interaction between<br />

objects. It is assumed that the mass is distributed evenly within the collision geometry<br />

and an inertia tensor is calculated accordingly.<br />

2 http://www.coin3d.org<br />

3 http://www.newtondynamics.com


6.1. Simulation Environment 115<br />

axis <strong>of</strong> rotation<br />

motor<br />

series spring<br />

segment i segment i+1<br />

parallel spring<br />

Figure 6.1: The layout <strong>of</strong> the simulated joint is assumed as a combination <strong>of</strong> a motor, a series<br />

spring, and a parallel spring.<br />

A third element type is used to add joints to the simulation model. Each joint defines<br />

constraints acting between two objects. Besides several standard joint types, Newton <strong>of</strong>fers<br />

so called universal joints to allow a high flexibility in joint design. Using the universal<br />

joint type, six constraints for the three linear and the three rotational motions between<br />

two objects can be specified. These constraints can either detain the movement along<br />

respectively about the corresponding axis, or induce an acceleration or torque on the<br />

two connected objects. This enables the implementation <strong>of</strong> loopback control mechanisms.<br />

In addition it is possible to read the force or torque that is necessary to maintain each<br />

constraint. Appendix B gives further details on the implementation <strong>of</strong> the simulation<br />

environment and presents an example <strong>of</strong> the scene description file.<br />

6.1.2 Actuators and Joint <strong>Control</strong><br />

In accordance to the system premises discussed in Section 5.1, a suitable model for joint<br />

actuation needs to be simulated. Its main characteristics are adjustable compliance,<br />

ranging from freely moving to high stiffness, an additional parallel elastic element, and<br />

direct torque control as well as position control. All joints used in the biped model <strong>of</strong> this<br />

work are revolute joints, i.e. rotational joints with five <strong>of</strong> their six degrees <strong>of</strong> freedom fixed.<br />

The ball joints <strong>of</strong> hip and spine and the ankle joints are simulated by a concatenation <strong>of</strong><br />

these revolute joints.<br />

The principle <strong>of</strong> the joint design suggested for the simulation is illustrated in Figure 6.1.<br />

The joint connects two robot segments i and i+1. In series with a revolving motor an<br />

elastic element with adjustable stiffness connects to the actuated axis. Another spring<br />

with fixed stiffness and constant equilibrium point is attached in parallel.<br />

Figure 6.2 presents the torque calculation used for implementing the constraint <strong>of</strong> the<br />

joint’s rotational degree <strong>of</strong> freedom. The controller receives a target torque τtarget and a<br />

target angle αtarget as inputs. Following the ideas <strong>of</strong> the behavior-based architecture iB2C,<br />

each input value is complemented by a stimulation value. The stimulation sτ is used to<br />

scale the target torque, thus only control units with full activity can generate the maximum<br />

torque output. For the position control part, sα is interpreted as stiffness. If the position<br />

controller is stimulated less, the compliance <strong>of</strong> the joint will be higher. This behavior is


116 6. Implementation <strong>of</strong> <strong>Bipedal</strong> <strong>Walking</strong> <strong>Control</strong><br />

sτ<br />

τtarget<br />

sα<br />

αtarget<br />

−<br />

Limit P P<br />

α<br />

α<br />

“compliance”<br />

˙αtarget<br />

−<br />

˙α<br />

τpos<br />

×<br />

×<br />

α0 = const.<br />

� wi<br />

sτ, sα<br />

Nonlinear<br />

Limit<br />

parallel spring<br />

damping<br />

P<br />

τspring<br />

τdamping<br />

τcontrol<br />

Figure 6.2: <strong>Control</strong> block diagram for the constraint <strong>of</strong> the rotational degree <strong>of</strong> freedom used<br />

for the simulated revolute joint.<br />

implemented by scaling the output τpos <strong>of</strong> the position control by the stimulation value.<br />

The controller itself is designed as a cascaded P-P control, with the outer position loop<br />

generating the target value ˙αtarget for the inner velocity loop. Resembling the iB2C fusion<br />

behaviors, the two torques resulting from the stimulation multiplication are combined by<br />

weighted summation, using the stimulation values as weights:<br />

τcontrol = sτ(τtargetsτ)+sα(τpossα)<br />

sτ +sα<br />

−<br />

−<br />

τ<br />

(6.1)<br />

This weighted sum is symbolized by � wi in Figure 6.2. The output <strong>of</strong> the loopback<br />

control part is given as τcontrol and limited by the maximum torque <strong>of</strong> the simulated motor.<br />

The parallel elasticity as it is found in the biological muscle is modeled as nonlinear spring.<br />

Given the constant equilibrium position α0 <strong>of</strong> the spring and the current joint angle α, the<br />

resulting torque is calculated as<br />

τspring = sgn(α0 −α)·Kspring ·(α0 −α) 2 . (6.2)<br />

with Kspring being the spring constant. Despite the fact that damping is nearly absent<br />

in biological joints, it is by all means present for technical systems. Thus, an additional<br />

torque τdamping proportional to the joint’s rotational velocity is calculated. All three torque<br />

values τcontrol, τspring, and τdamping are then combined and used as the acting torque τ for<br />

the rotational constraint <strong>of</strong> the simulated joint.<br />

This joint setup allows to create various joint behaviors as needed by the biologically<br />

motivated control concept. If both stimulation values sτ and sα are set to zero, the joint<br />

can rotate freely except for the parallel spring and the damping. Setting the position<br />

stimulation to zero, a direct torque command can be implemented. With sα > 0, an elastic,


6.1. Simulation Environment 117<br />

(a) (b)<br />

Figure 6.3: (a) Joint construction with (from top to bottom) the motor, the gearbox, and the<br />

gas spring highlightened. (b) Monopod prototype attached to a vertical slider.<br />

passive joint action with the given equilibrium point αtarget is achieved. As described<br />

below, precedence <strong>of</strong> the torque command over the position command is guaranteed by an<br />

adequate layer <strong>of</strong> fusion behaviors.<br />

In the antagonistic setup <strong>of</strong> two muscle group acting on a human joint, information on<br />

the current status <strong>of</strong> the joint is perceived by the muscle spindles, the golgi tendon organ,<br />

the pacinian corpuscle, the Ruffini endings, and the surrounding skin. The joint used in<br />

this simulation emits the current joint angle and the torque acting in the joint as sensor<br />

information. Both values can be extracted from the physics engine. In a real robotic joint,<br />

the current angle can be measured by e.g. encoders. The torque can be captured by using<br />

a load cell. Given an appropriate loopback control, the torque can also be derived by<br />

measuring the motor current, but then a low ration gear is necessary. Otherwise, external<br />

forces would not be transferred to the motor.<br />

In the scope <strong>of</strong> this work, some consideration has also been put towards the development <strong>of</strong><br />

an actuator implementing the characteristics just described [Luksch 05, Luksch 07]. The<br />

preliminary prototype follows the joint layout presented in Figure 6.1. A dc motor with<br />

disc-shaped rotor is used as main actuator. The serial elastic element is omitted in the<br />

first prototype. Instead, active compliance is implemented by a suitable controller similar<br />

to the one presented above [Wahl 09, Blank 09b]. A low friction gear box with a ratio <strong>of</strong>


118 6. Implementation <strong>of</strong> <strong>Bipedal</strong> <strong>Walking</strong> <strong>Control</strong><br />

(a) (b)<br />

Figure 6.4: (a) Inertial measurement unit, and (b) three-axis load cell developed at the Robotics<br />

Research Lab.<br />

1:32 still allows the joint to move freely. In total, the setup can produce a peak torque <strong>of</strong><br />

more than 150Nm. Figure 6.3a depicts the resulting joint construction, with the main<br />

components marked in red.<br />

In the first setup <strong>of</strong> the prototype joint, the parallel spring is designed as a novel rotatory<br />

gas spring [Flörchinger 05]. By changing the volume being compressed by the rotating<br />

piston, the stiffness <strong>of</strong> the spring would be adaptable. Also, the elasticity would be<br />

nonlinear as in biological muscles. Unfortunately, the construction turned out to be very<br />

difficult to manufacture. Remaining air leaks permit to produce the desired torque. Thus<br />

the gas spring is complemented by a conventional linear coil spring that is stretched during<br />

the flexion <strong>of</strong> the joint.<br />

Using this actuator, the monopod prototype shown in Figure 6.3b is developed [Roth 08,<br />

Schumann 08]. The active hip and knee joints allow the robot to jump on the spot. The<br />

upper body is guided by a passive linear slider to neglect the problem <strong>of</strong> stabilization. The<br />

setup is intended for the evaluation <strong>of</strong> the joint construction and the controller in a highly<br />

dynamic task. For more details on this ongoing research, please refer to the cited papers.<br />

While the presented actuator is still under development and not implementing all features<br />

preferable for the suggested control concept, it at least shows that the design <strong>of</strong> such an<br />

actuator should be possible.<br />

6.1.3 Simulation <strong>of</strong> Sensors<br />

Similar to the joint actuation, a suitable sensor setup must be provided by the simulation<br />

framework. Senses necessary for the suggested control system include a sense <strong>of</strong> balance,<br />

load perception, and cutaneous sensor information <strong>of</strong> the foot soles. These senses are<br />

implemented as inertial measurement unit (imu), load cells, and force sensors in the feet.<br />

To stay as close as possible to real sensor systems, the imu and the load cell developed at<br />

the Robotics Research Laboratory were modeled (see Figure 6.4).<br />

The imu is able to measure three linear accelerations and three rotational velocities. To<br />

model the functionality <strong>of</strong> the imu within the simulation, an object <strong>of</strong> known mass is


6.1. Simulation Environment 119<br />

Joint min. angle [rad] max. angle [rad] max. torque [Nm]<br />

Spine X -1.0 1.0 180<br />

Spine Y -1.0 1.0 180<br />

Spine Z -1.0 1.0 100<br />

Shoulder X left 0.0 2.0 80<br />

Shoulder X right -2.0 0.0 80<br />

Shoulder Y -1.0 1.0 80<br />

Elbow Y -2.0 0.0 30<br />

Hip X -1.0 1.0 220<br />

Hip Y -0.85 0.15 220<br />

Hip Z -1.0 1.0 150<br />

Knee Y 0.0 2.0 150<br />

Ankle X -0.5 0.5 80<br />

Ankle Y -0.5 0.5 150<br />

Table 6.1: Minimum and maximum angles and the maximum torque for all degrees <strong>of</strong> freedom<br />

<strong>of</strong> the simulated biped.<br />

attached to the mount point <strong>of</strong> the imu by a joint fixing all six degrees <strong>of</strong> freedom. The accelerations<br />

are calculated by dividing the forces necessary to maintain the linear constraints<br />

by the objects mass. This method is similar to the working <strong>of</strong> real microelectromechanical<br />

acceleration sensors using a small suspended mass and measuring its deflection via a change<br />

<strong>of</strong> capacitance. The rotational velocities can be calculated by transforming the velocities<br />

<strong>of</strong> the imu object provided by the physics engine to the current coordinate system <strong>of</strong> the<br />

imu.<br />

Similar to the imu, the load cell is implemented by a fixed joint attached between two<br />

robot segments. The load cell shown in Figure 6.4b measures the force in z-direction and<br />

the torque about the x- and y-axis. These values <strong>of</strong> interest can directly be read from<br />

the physics engine as it provides the forces and torques produced to preserve the joint<br />

constraints. The force sensors placed within the foot soles are <strong>of</strong> equal design, only they<br />

just return the force in one direction and no torque values.<br />

6.1.4 Model <strong>of</strong> Simulated Biped<br />

Deploying the joint actuation and sensors just described, the model for the simulated biped<br />

can be designed. The kinematic layout corresponds to the compromise <strong>of</strong> joint selection<br />

described in Section 5.1. The ball joints mentioned there are implemented as chained<br />

revolute joints. Table 6.1 summarizes the range <strong>of</strong> movement and the maximum torques<br />

for each joint. Segment lengths and masses are similar to what can be found in human<br />

subjects. Masses were taken from Table 3.2, but in contrast to human segments made <strong>of</strong><br />

bone, muscle, and tissue, the mass distribution is assumed to be uniform for reasons <strong>of</strong><br />

simplicity. In total, the number <strong>of</strong> degrees <strong>of</strong> freedom amounts to 21, the robot’s weight<br />

adds up to 76kg, and its overall height is 1.8m. Details on the exact masses and the<br />

placement <strong>of</strong> all segments can be found in Appendix B where the scene description file for<br />

the robot setup and the environment is given.


120 6. Implementation <strong>of</strong> <strong>Bipedal</strong> <strong>Walking</strong> <strong>Control</strong><br />

(a)<br />

z<br />

x y<br />

Figure 6.5: (a) Simulated biped as visualized by the simulation framework. (b) Joint layout <strong>of</strong><br />

the biped model. The hip and spine joints are modeled by three, the shoulder and ankle joints as<br />

two subsequent revolute joints.<br />

Figure 6.5 illustrates the visualization and the joint layout <strong>of</strong> the biped model. The<br />

coordinate system depicts the direction <strong>of</strong> the axes, namely the x-axis pointing in walking<br />

direction, the y-axis pointing sidewards to the left, and the z-axis pointing upwards,<br />

respectively. The visible shapes correspond to the collision geometries used by the physics<br />

engine. The trunk and pelvis are modeled in such a way that the center <strong>of</strong> mass is located<br />

slightly in front <strong>of</strong> the hip joints. Similarly, the head position is shifted in x-direction. This<br />

position <strong>of</strong> the center <strong>of</strong> mass facilitates the forward motion while walking, in particular<br />

during the stance phase as the trunk travels over the supporting foot.<br />

In contrast to the vestibular system located in the human head, the inertial measurement<br />

unit is attached to the pelvis <strong>of</strong> the biped model. This has the advantage <strong>of</strong> circumventing<br />

the oscillation <strong>of</strong> the upper body, facilitating the stabilization <strong>of</strong> the pelvis. Knowing the<br />

angles <strong>of</strong> the spine joint, the trunk can still be balanced at an upright pose.<br />

The foot geometry is shown as a close-up in Figure 6.6. The white-colored part between<br />

the ankle joint and the actual foot represents the load cell. It measures the force acting in<br />

z-direction and the torques about the x- and y-axis. The elliptical shape <strong>of</strong> the toe objects<br />

allows the foot to roll over the ground during push-<strong>of</strong>f, creating a behavior similar to the<br />

bending <strong>of</strong> the forefoot and the toes. This geometry can save an additional joint in the<br />

foot. Each <strong>of</strong> the round toe and heel objects is attached to the central foot object by a<br />

fixed joint measuring the force acting in z-direction. By knowing these four load values for<br />

each foot, the center <strong>of</strong> pressure can be estimated. The force sensors in the feet correspond<br />

to the cutaneous perception found in the human foot.<br />

(b)


6.2. Notes on the Implementation 121<br />

Figure 6.6: Shape <strong>of</strong> the foot model allowing to roll over the toes.<br />

6.2 Notes on the Implementation<br />

As mentioned above, the control concept postulated in this work is implemented within<br />

the behavior-based architecture iB2C and the robot control framework mca2. Both <strong>of</strong><br />

these systems <strong>of</strong>fer special features, but also pose certain restrictions on the design. Some<br />

<strong>of</strong> the resulting decision necessary during the development process will be discussed in the<br />

following.<br />

6.2.1 Group Layout and Phase Representatives<br />

As illustrated in Figure 6.7, the control and simulation system is arranged within three<br />

main groups. The mca2 framework executes all processing and data transportation<br />

routines <strong>of</strong> these groups and the contained modules once for each control cycle. For the<br />

experiments done in this work, this cycle runs at 40Hz. This results in a delay time for<br />

reflex action <strong>of</strong> at least 25ms, thus the control system has to manage similar dead times<br />

as found in neural pathways. The group gSimulation contains the modules responsible<br />

for the dynamical simulation and the three-dimensional visualization <strong>of</strong> the current scene.<br />

The iteration step size for the dynamics calculation is set to 1ms, resulting in 25 iterations<br />

per control cycle. As the low level joint controller described above is embedded within the<br />

physics simulation cycle, it also runs at 1kHz.<br />

The interface <strong>of</strong> the simulation group is designed in a way to enable its replacement by<br />

a real robot. The control interface is composed <strong>of</strong> the inputs for the joint controllers,<br />

i.e. the target torque and position as well as the two corresponding stimulation values.<br />

On its sensor side, the simulation provides the current joint positions and torques, the<br />

accelerations and velocities measured by the inertial measurement unit, and the forces and<br />

torques measured by the load cells. To abstract from actual limits <strong>of</strong> joint angles, motor<br />

torques, or sensor information, the gAbstractionLayer group introduces a normalization<br />

layer between the simulation and the control system.<br />

The group gBiped<strong>Control</strong> contains the control system as described in the previous chapter.<br />

Its control units are distributed to six groups: the locomotion modes are located within<br />

the brain group, the spgs and the postural reflexes within the spinal cord group. As local<br />

reflexes and motor patterns only act on a few joints and do not receive sensor information<br />

from other parts <strong>of</strong> the robot, they can be separated in four joint groups for the upper<br />

and lower trunk and the legs.


122 6. Implementation <strong>of</strong> <strong>Bipedal</strong> <strong>Walking</strong> <strong>Control</strong><br />

gBiped<strong>Control</strong><br />

gAbstractionLayer<br />

gSimulation<br />

(a)<br />

gBiped<strong>Control</strong><br />

gBrain<br />

gSpinalCord<br />

gUpperTrunk gLowerTrunk gLeftLeg gRightLeg<br />

Figure 6.7: (a) Main layout <strong>of</strong> the control and simulation system. (b) The control units are<br />

arranged within six groups. The four joint groups at the lower end contain the motor interface.<br />

The motion phases are located in the spinal cord group. But to improve clarity, to reduce<br />

communication bandwidth and interface complexity, and to increase local coherence, a<br />

representative <strong>of</strong> each motion phase is introduced in each joint group. These units are<br />

stimulated by their counterpart in the spinal cord and again manage the stimulation <strong>of</strong><br />

local reflexes and motor patterns <strong>of</strong> their joint group. They also set the compliance and<br />

equilibrium point for the passive joints.<br />

6.2.2 Application <strong>of</strong> iB2C Features<br />

The joint groups also contain the interface to the joint controllers in form <strong>of</strong> iB2C fusion<br />

behaviors. They provide the means to coordinate concurrent access to the same joint<br />

by several control units. Only units <strong>of</strong> the classes motion phase, postural reflex, motor<br />

pattern, and local reflex generate control commands for the joints, with motion phases<br />

only setting a fixed equilibrium point and compliance. As a behavior’s activity signal is<br />

used for calculating its influence during the fusion, only control units stimulated by the<br />

active motion phase will be considered.<br />

Two different fusion methods are applied for the joint interface: position commands are<br />

fused by the weighted average method, thereby finding a compromise between the joint<br />

angle demands <strong>of</strong> active control units. Torque commands are combined by the weighted<br />

sum function. As the target torque for a joint roughly corresponds to muscle activity,<br />

several possible torque demands are added up as a motor neuron would add up neural<br />

activity. The resulting torque command must be limited as it can extend the maximum<br />

value.<br />

Figure 6.8 illustrates the behavior-based joint control interface using the example <strong>of</strong> the<br />

leg groups. The left and right leg group each manage the knee and the two ankle joints.<br />

For each <strong>of</strong> the group’s joints, two fusion behaviors are inserted, one being responsible<br />

for position, the other for torque commands. The torque fusion behavior inhibits the<br />

position fusion behavior using its activity. This design decision is based on the assumption<br />

that control units generating torque commands have higher priority than those providing<br />

(b)


6.2. Notes on the Implementation 123<br />

Motor Pattern<br />

(F) Knee<br />

Torque<br />

sτ,knee<br />

ˆτknee<br />

(F) Knee<br />

Position<br />

sα,knee<br />

ˆαknee<br />

Local Reflex<br />

(F) Ankle X<br />

Torque<br />

sτ,ankle x<br />

ˆτankle x<br />

(F) Ankle X<br />

Position<br />

sα,ankle x<br />

ˆαankle x<br />

(F) Ankle Y<br />

Torque<br />

sτ,ankle y<br />

ˆτankle y<br />

(F) Ankle Y<br />

Position<br />

sα,ankle y<br />

ˆαankle y<br />

Figure 6.8: Interface to joint controllers implemented as layer <strong>of</strong> fusion behaviors. Precedence<br />

<strong>of</strong> torque fusion is achieved by inhibiting edges.<br />

position and stiffness commands. This also agrees to the suggested design guideline that<br />

torque control should be preferred to position control.<br />

Each fusion behavior forwards the resulting control value and its activity to the joint<br />

controllers described above. The normalized torque and angle commands ˆτi and ˆαi are<br />

scaled to effective values while passing the abstraction layer. The activity <strong>of</strong> the torque<br />

fusion behavior serves as stimulation sτi for the torque part <strong>of</strong> the controller, <strong>like</strong>wise the<br />

activity <strong>of</strong> the position fusion behavior is used as stimulation sαi<br />

for the position part. As<br />

the fusion behavior’s activity results from the activity <strong>of</strong> the connected control units, these<br />

units can use their own activity signal to influence the desired compliance or to scale the<br />

effective torque value.<br />

The activity <strong>of</strong> motor patterns is based on the sigmoid function used to calculate the<br />

torque output. Similar to Equation 5.7, the signal is depending on the three parameters<br />

describing the temporal progress <strong>of</strong> the torque value, only a factor for scaling the amplitude<br />

is missing as the activity will always rise up to the reflex activation ι and fall back to zero:<br />

⎧<br />

1 1 t 1 ⎪⎨ + sin( π( − ) ) 0 ≤ t < T1<br />

2 2 T1 2<br />

a = ι· 1 T1 ≤ t < T2<br />

⎪⎩ 1 1 t−T2 1 − sin( π( − 2 2 T3−T2 2 ) ) T2<br />

(6.3)<br />

≤ t ≤ T3<br />

In contrast to the uniform activity <strong>of</strong> motor patterns, the activity <strong>of</strong> postural and local<br />

reflexes depends on sensory information. For instance, the postural reflex Upright Trunk<br />

calculates a normalized hip torque ˆτhip to balance the upper body based on the trunk<br />

pitch angle and its velocity. The activity signal <strong>of</strong> this reflex is derived from the output<br />

torque, but is then scaled by the normalized ground contact value ˆ Fz ∈ [0,1] measured by<br />

the force sensors:<br />

a = ι·|ˆτhip|· ˆ Fz<br />

(6.4)


124 6. Implementation <strong>of</strong> <strong>Bipedal</strong> <strong>Walking</strong> <strong>Control</strong><br />

This ensures that the activity decreases with the load on the stance leg as torque applied<br />

to the hip joint can only act on the upper body if the leg is firmly connected to the ground.<br />

The multiplication by the reflex activation ι corresponds to the iB2C principle stating that<br />

the activity <strong>of</strong> a behavior must never exceed it activation. As activity signals are used to<br />

stimulate other behaviors, this principle guarantees a limitation <strong>of</strong> the activity throughout<br />

the behavior network.<br />

This principle is also obeyed by the activity calculation <strong>of</strong> the motion phase control<br />

units. As described in Chapter 4, these units are responsible for stimulating the motor<br />

patterns and reflexes being active during the respective phase <strong>of</strong> motion, and for setting<br />

the compliance and equilibrium points <strong>of</strong> the passive joints. As just described, setting<br />

the stiffness <strong>of</strong> a joint corresponds to the appropriate stimulation <strong>of</strong> the associated fusion<br />

behavior. Thus stimulating a reflex or motor pattern and setting a joint compliance<br />

functions by the same mechanism. Consequentely it is possible to implement all motion<br />

phases as instance <strong>of</strong> the same behavior class called mbbStimulationBehavior. Each<br />

behavior or joint k controlled by the motion phase is registered by passing a stimulation<br />

factor fs,k along with a possible modulation value or equilibrium point. The activity <strong>of</strong><br />

the motion phase itself is simply set to its activation (a = ι), but the stimulation signal a k<br />

used for the controlled units is calculated as<br />

a k = a·fs,k. (6.5)<br />

By setting fs,k <strong>of</strong> a passive joint to zero, it is allowed to move freely during the motion<br />

phase, only the passive elastic element and external forces will influence its action. A<br />

factor fs,k > 0 results in compliant joint behavior around its equilibrium point. Setting<br />

fs,k <strong>of</strong> a motor pattern or reflex to a value smaller than one will restrict its influence in the<br />

control network during this motion phase. The actual compliance for each joint during all<br />

motion phases <strong>of</strong> walking will be given in Chapter 8 when discussing passive joint action<br />

during normal walking.<br />

The target ratings <strong>of</strong> the control units do not play as crucial a role as their activity. Still,<br />

an analysis <strong>of</strong> the target rating can be used during implementation and testing to better<br />

understand the working <strong>of</strong> the behaviors in a given situation. For this purpose, the target<br />

rating should have a meaningful value representing the contentment <strong>of</strong> the behavior. To<br />

continue the above example <strong>of</strong> the postural reflex Upright Trunk, its target rating r is<br />

given as<br />

r = min(1,|φ−φtarget|) (6.6)<br />

with φ being the current body pitch angle in radians and φtarget the target pitch angle<br />

depending on the desired walking velocity. This kind <strong>of</strong> target rating calculation allows to<br />

interpret the state <strong>of</strong> the robot regarding the goal <strong>of</strong> an individual control unit.<br />

A further use <strong>of</strong> the target rating is its application as pre-processed sensor value. This<br />

method can reduce the extend <strong>of</strong> sensor information that must be passed to higher control<br />

levels. An example for this kind <strong>of</strong> usage will be given below, where the target rating<br />

<strong>of</strong> the Lock Knee reflex serves as indication for switching the state <strong>of</strong> the cyclic walking<br />

spinal pattern generator.


7. Balanced Standing Experiments<br />

Before discussing the performance <strong>of</strong> the implemented control system during dynamic<br />

walking, this chapter will present its capability for stable standing [Steiner 08]. Several<br />

experiments will demonstrate different aspects <strong>of</strong> the control introduced in Section 5.3.<br />

The adaptation to different ground geometries becomes apparent when letting the robot<br />

fall on inclined terrain or steps. The control <strong>of</strong> joint stiffness can be observed by applying<br />

external forces to body segments during standing or by moving the platform the robot<br />

is standing on. Posture optimization takes place after ground contact, when tilting the<br />

ground platform, or in case the robots mass distribution changes, e.g. by adding weight to<br />

a backpack mounted on the robot. Some <strong>of</strong> these experiments are inspired by studies on<br />

human postural control.<br />

The setup used for the following experiments is composed <strong>of</strong> the robot standing on or<br />

falling on a square plate with an edge length <strong>of</strong> 1m. The plate can be moved along and<br />

rotated about all three axis to apply disturbances. Additionally, arbitrary forces can be<br />

deployed to individual body segments. Sequences <strong>of</strong> platform motions or external forces<br />

can be programmed and executed automatically to generate repeatable runs.<br />

7.1 Adaptation to the Ground Geometry<br />

In order to test the ground adaptation capabilities <strong>of</strong> the standing control, the robot<br />

is dropped from a height <strong>of</strong> about 20cm on different ground geometries as exemplified<br />

in Figure 7.1. The actual shape <strong>of</strong> the terrain is unknown to the control system. The<br />

adaptation relies on the initially moderate compliance in most joints and on the high<br />

compliance <strong>of</strong> the ankle joints. This strategy enables a passive matching to the ground<br />

geometry. Then, high joint stiffness is used to stabilize the posture.<br />

Figure 7.2 illustrates the control behavior during a drop on sloped ground. The platform<br />

the robot is landing on is inclined by 6 ◦ about the y-axis. In fact, the robot remains stable<br />

at inclinations <strong>of</strong> up to 30 ◦ , but the parallel springs <strong>of</strong> the ankle joints act with increasing<br />

torque and the joint relaxing phase cannot be observed anymore.<br />

While still in the air, the spinal pattern generator for stable standing stimulates the first<br />

phase called Ground Adaptation. The local Hold Position reflexes are stimulated, but the


126 7. Balanced Standing Experiments<br />

(a) (b) (c) (d)<br />

Figure 7.1: Experiment on ground geometry adaptation. The robot is dropped from about<br />

20cm height on sloped ground with both feet (a,b) and on a step with just the toes <strong>of</strong> one foot<br />

(c,d), then the posture is stabilized.<br />

stiffness modulation factor is zero. The activity <strong>of</strong> the reflexes is medium for most joints<br />

and near zero for the ankle joints as shown on the lower right side <strong>of</strong> the figure. As the<br />

reflexes connect to the position fusion nodes, their activity is interpreted as joint stiffness,<br />

resulting in compliant ankles.<br />

At point in time t1 the robot touches the ground, easily recognizable by the force peak in<br />

z-direction measured by the ankle load cells. In the short time interval until full contact,<br />

the joints passively adapt to the ground shape, with the major adaptation taking place<br />

in the ankle joint. As seen in the lower left plot, the ankle compensates for most <strong>of</strong> the<br />

6 ◦ <strong>of</strong> ground inclination. With full foot contact at t2, the standing spg switches to the<br />

stabilization phase. The stiffness factor increases to one in order to brace the joints at<br />

their current angles. With the raising activity <strong>of</strong> the Hold Position reflexes, the position<br />

control stabilizes the joint angles.<br />

As the taken stance is relatively steady, i.e. no strong accelerations are measured, the<br />

spg directly moves on to the next phase trying to optimize the posture. This phase<br />

will be described in detail below, but Figure 7.2 exemplary shows the activity <strong>of</strong> the<br />

postural reflex Relax Spine Y to clarify the processes during stable standing. The posture<br />

optimization causes minor joint movements resulting in accelerations recorded by the<br />

inertial measurement unit (imu). As soon as the joint torques have reached a lower<br />

threshold at t3, the Relaxed Posture phase is entered. The stiffness factor is reduced,<br />

resulting in lower activity <strong>of</strong> the Hold Position reflexes and thus in higher joint compliance.<br />

7.2 External Forces and Platform Movements<br />

The stabilization phase <strong>of</strong> the standing control mainly relies on passing favorable joint<br />

position and high stiffness values to the joint controllers. This results in a stance stabilization<br />

similar to the human ankle strategy. Two kind <strong>of</strong> experiments will validate the


7.2. External Forces and Platform Movements 127<br />

Loadcell Force Z [kN]<br />

Acceleration [ m/s 2 ]<br />

Spine Y Angle [ ◦ ]<br />

Hip Y Angle [ ◦ ]<br />

Knee Angle [ ◦ ]<br />

Ankle Y Angle [ ◦ ]<br />

3.0<br />

2.0<br />

1.0<br />

0.0<br />

0.4<br />

0.2<br />

0.0<br />

2.0<br />

1.0<br />

0.0<br />

−1.0<br />

5.0<br />

4.0<br />

3.0<br />

t1t2 t3 t1t2 t3<br />

0.0 1.0 2.0 3.0 4.0<br />

0.0 1.0 2.0 3.0 4.0<br />

0.0 1.0 2.0 3.0 4.0<br />

2.0<br />

0.0 1.0 2.0 3.0 4.0<br />

7.0<br />

6.0<br />

5.0<br />

4.0<br />

0.0 1.0 2.0 3.0 4.0<br />

0.0<br />

−2.0<br />

−4.0<br />

−6.0<br />

0.0 1.0 2.0<br />

Time [s]<br />

3.0 4.0<br />

Stiffness Factor<br />

Relax Spine Y a<br />

Hold Pos. Spine Y a<br />

Hold Pos. Hip Y a<br />

Hold Pos. Knee a<br />

Hold Pos. Ankle Y a<br />

1.0<br />

0.5<br />

0.0<br />

0.0 1.0 2.0 3.0 4.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 1.0 2.0 3.0 4.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 1.0 2.0 3.0 4.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 1.0 2.0 3.0 4.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 1.0 2.0 3.0 4.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 1.0 2.0<br />

Time [s]<br />

3.0 4.0<br />

Figure 7.2: Experiment on ground geometry adaptation. t1: ground contact, t2: full contact,<br />

t3: joint relaxation.


128 7. Balanced Standing Experiments<br />

Figure 7.3: Experiment on stabilization against external forces. A force <strong>of</strong> 100N in direction <strong>of</strong><br />

the x-axis is applied to the head for 300ms. The arrow above the head indicates the acceleration<br />

measured by the imu.<br />

performance <strong>of</strong> this part <strong>of</strong> the control system. External forces are applied to different<br />

segments <strong>of</strong> the robot, and the platform it is standing on is moved horizontally. Both<br />

<strong>of</strong> these setups can frequently be found in studies on human balancing capabilities. The<br />

following analysis is limited to movements in the frontal plane. Lateral disturbances can<br />

be compensated more easily as the feet are placed side by side in normal stance.<br />

Figure 7.3 shows the setup during the application <strong>of</strong> an external force to the head segment.<br />

Beginning with a quiet stance, the force is pushing the robot forward. The resulting<br />

acceleration is visualized by the arrow above the robot’s head. The low compliance <strong>of</strong> the<br />

joints in the stabilization phase only allows for small joint movements before the robot is<br />

moving back to a stable posture.<br />

The results <strong>of</strong> a run applying an external force <strong>of</strong> 100N in x-direction to the head segment<br />

for 300ms are shown in Figure 7.4. The gray bar starting at t = 0.65sec is indicating<br />

the time span during which the force is effective. The impact <strong>of</strong> the force is clearly<br />

demonstrated by the acceleration in x-direction measured by the imu, or by the change <strong>of</strong><br />

the pitch angle <strong>of</strong> the torso as it is leaning forward.<br />

Before the external force is applied, the robot is standing in a relaxed posture and the<br />

spg is stimulating the fourth motion phase. This results in a low stiffness factor and<br />

moderately compliant joints. As soon as the disturbance is measured by the imu, the spg<br />

switches to the stabilization phase. The stiffness factor is pulled up and the compliance<br />

<strong>of</strong> the joint position control is reduced. The center <strong>of</strong> pressure moves to the front <strong>of</strong> the<br />

feet as the maximum leverage is applied to move the body back into an upright position.<br />

The plot marked“CoP Position”in Figure 7.4 indicates the movement <strong>of</strong> the center <strong>of</strong><br />

pressure. Being based on the four force measurement in each foot, a value <strong>of</strong> 1 denotes<br />

that all vertical force is shifted to the front <strong>of</strong> the foot, at a value <strong>of</strong> −1 it is located at<br />

the heels. A cop position <strong>of</strong> zero will occur if the center <strong>of</strong> pressure lies below the ankle<br />

joint, i.e. in case the system is statically balanced.<br />

Analysis <strong>of</strong> the arising torques <strong>of</strong> the spine, hip, knee, and ankle joints reveals the reaction<br />

<strong>of</strong> the joint controllers to the disturbance. As soon as the external force deflects the trunk,<br />

an opposite torque can be observed. The reaction is propagated with a small delay to the


7.2. External Forces and Platform Movements 129<br />

Accel. X [ m/s 2 ]<br />

Torso Pitch [ ◦ ]<br />

Spine Y Torque [Nm]<br />

Hip Y Torque [Nm]<br />

5<br />

0<br />

−5<br />

10<br />

0<br />

0.0 2.0 4.0 6.0<br />

−10<br />

0.0 2.0 4.0 6.0<br />

50<br />

0<br />

−50<br />

50<br />

0<br />

−50<br />

0.0 2.0 4.0 6.0<br />

0.0 2.0 4.0 6.0<br />

Time [s]<br />

Stiffness Factor<br />

CoP Position<br />

Knee Torque [Nm]<br />

Ankle Y Torque [Nm]<br />

1.0<br />

0.5<br />

0.0<br />

0.0 2.0 4.0 6.0<br />

1.0<br />

0.0<br />

−1.0<br />

0.0 2.0 4.0 6.0<br />

50<br />

0<br />

−50<br />

50<br />

0<br />

−50<br />

0.0 2.0 4.0 6.0<br />

0.0 2.0 4.0 6.0<br />

Time [s]<br />

Figure 7.4: Experiment on stabilization against external forces. The gray bar marks the time<br />

interval <strong>of</strong> 300ms during which an external force <strong>of</strong> 100N is pulling the head segment forwards.<br />

Experiment Torque [Nm] Acceleration<br />

Force [N] Time [ms] spine y hip y knee ankle y x-axis [ m/s2] 1000 25 160 82 68 64 75.0<br />

100 300 57 30 37 50 7.6<br />

50 600 38 22 30 46 4.1<br />

25 ∞ 24 14 23 40 2.0<br />

−750 25 −130 −62 −61 −45 −57.0<br />

−70 300 −39 −22 −32 −35 −5.0<br />

−40 600 −30 −17 −30 −34 −2.9<br />

−25 ∞ −23 −14 −28 −34 −2.0<br />

Table 7.1: Experiment on stabilization against external forces. The columns give the maximum<br />

forces and the resulting accelerations at which the robot does still keep standing.


130 7. Balanced Standing Experiments<br />

Figure 7.5: Major muscle activity <strong>of</strong> human subjects during movements <strong>of</strong> the supporting<br />

platform. From [Ting 07], p302.<br />

joints more distant to the disturbance. The external force pulling at the head segment<br />

results in the robot standing solely on its toes. During this timespan, the lower joints have<br />

to support a larger amount <strong>of</strong> the machine’s weight, consequently the torque in the ankle<br />

joints is higher than in the other joints. The robot then shifts back and is again supported<br />

by the whole foot. After only a short oscillation, the posture is stabilized and the joint<br />

torques fall back to a low value.<br />

The control system can only handle disturbances up to a certain threshold. If the intensity<br />

<strong>of</strong> the external force is increased, the timespan for which this force may act without<br />

the robot falling down decreases, and vice versa. Table 7.1 lists some <strong>of</strong> these limits.<br />

The first two columns give the force value and the interval during which the force was<br />

applied. A negative force indicates that the head segment was pulled backwards. Small<br />

disturbances can be compensated continuously. The remaining columns list the resulting<br />

maximum torques and the maximum acceleration in x-direction measured during the<br />

corresponding experiment. As expected, stronger forces result in higher torques and<br />

accelerations. Disturbances pushing the robot backwards must not be as intense due to<br />

the shorter leverage <strong>of</strong> the foot on the heel side. Applying the force to other segments<br />

yields similar results. When pulling or pushing the pelvis segment, the sign <strong>of</strong> the spine<br />

torque changes because <strong>of</strong> the mass inertia <strong>of</strong> the upper body.<br />

A further series <strong>of</strong> experiments analyzes the control behavior at linear movements <strong>of</strong> the<br />

platform the robot is standing on. Similar to the studies on human subjects by Ting et al.<br />

(Figure 7.5), the platform is moved by about 12cm with a gradual acceleration in forward<br />

and backward direction [Ting 07].<br />

Figure 7.6 illustrates the reaction <strong>of</strong> the control system to a 12cm platform movement in<br />

forward direction accelerating to a maximum velocity <strong>of</strong> 0.25 m/s. The upper row outlines<br />

the platform motion, the gray bar denotes the time interval during which the platform<br />

is moving. The robot’s behavior basically resembles the one <strong>of</strong> the previously described<br />

experiments. As soon as the disturbance is perceived by the acceleration sensor, the joint


7.2. External Forces and Platform Movements 131<br />

Platform Vel. [ m/s]<br />

Accel. X [ m/s 2 ]<br />

Torso Pitch [ ◦ ]<br />

Spine Y Torque [Nm]<br />

Hip Y Torque [Nm]<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0.0 1.0 2.0 3.0 4.0<br />

5<br />

0<br />

−5<br />

5<br />

0<br />

−5<br />

50<br />

0<br />

−50<br />

50<br />

0<br />

−50<br />

0.0 1.0 2.0 3.0 4.0<br />

0.0 1.0 2.0 3.0 4.0<br />

0.0 1.0 2.0 3.0 4.0<br />

0.0 1.0 2.0<br />

Time [s]<br />

3.0 4.0<br />

Platform Pos. [m]<br />

Stiffness Factor<br />

CoP Position<br />

Knee Torque [Nm]<br />

Ankle Y Torque [Nm]<br />

0.15<br />

0.10<br />

0.05<br />

0.00<br />

0.0 1.0 2.0 3.0 4.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 1.0 2.0 3.0 4.0<br />

1.0<br />

0.0<br />

−1.0<br />

0.0 1.0 2.0 3.0 4.0<br />

50<br />

0<br />

−50<br />

50<br />

0<br />

−50<br />

0.0 1.0 2.0 3.0 4.0<br />

0.0 1.0 2.0<br />

Time [s]<br />

3.0 4.0<br />

Figure 7.6: Experiment on stabilization against platform movement. The gray bar marks the<br />

time interval during which the platform is moving for 12cm at a maximum velocity <strong>of</strong> 0.25m/s.<br />

stiffness is raised to stabilize the posture. While the upper body sways backwards due<br />

to the platform acceleration, the center <strong>of</strong> pressure completely moves to the heel side<br />

<strong>of</strong> the feet. As the body swings forwards again during platform deceleration, the center<br />

<strong>of</strong> pressure moves to the toe and then levels out at its equilibrium point as the robot<br />

stabilizes.<br />

The joint torques correspond to the muscle activities observed by Ting in her experiments<br />

with human subjects (Figure 7.5). The main work is done by the tibialis anterior which


132 7. Balanced Standing Experiments<br />

Max. platform Torque [Nm] Acceleration<br />

velocity [ m/s] spine y hip y knee ankle y x-axis [ m/s2] 2.0 −110 −60 −65 −46 16.0<br />

1.0 −75 −45 −54 −44 11.0<br />

0.5 −24 −15 −31 −35 3.1<br />

0.25 −19 −12 −30 −35 2.6<br />

−2.0 115 62 62 60 −16.0<br />

−1.0 80 45 50 55 −11.0<br />

−0.5 33 20 30 45 −4.2<br />

−0.25 22 14 24 42 −3.0<br />

Table 7.2: Experiment on stabilization against forward and backward platform movement. The<br />

first column states the maximum platform velocity during a displacement <strong>of</strong> 12cm. The following<br />

columns list the maximum torques and x-acceleration during the respective experiment.<br />

is active throughout the disturbances. Similarly, Figure 7.6 shows the prominent torque<br />

<strong>of</strong> the ankle joint lasting until the robot sways forward due to platform deceleration.<br />

As in human subjects, the reaction <strong>of</strong> the distal muscles respectively joint actuators is<br />

increasingly delayed with the distance to the ankle joint. The knee, hip, and spine joints<br />

apply a torque to prevent the body from bending backwards. A <strong>like</strong>wise reaction can be<br />

observed in the activity <strong>of</strong> the quadriceps and abdominal muscles.<br />

<strong>Human</strong> subjects manage to compensate the disturbance with less sway <strong>of</strong> the upper<br />

trunk. As the center <strong>of</strong> mass displacement shown in Figure 7.5 illustrates, the body stays<br />

in a backward position and slowly moves back to an upright pose after the platform<br />

movement stops. The pitch angle <strong>of</strong> the robot’s torso visualizes the slight overreaction<br />

<strong>of</strong> the joint controllers due to strong movements <strong>of</strong> body masses. This also leads to the<br />

change <strong>of</strong> direction in joint torques after the platform motion, a reaction not visible in<br />

the corresponding muscles <strong>of</strong> human subjects. A less aggressively tuned controller would<br />

reduce the body sway, but would also decrease the maximum platform velocities that can<br />

still be stabilized. Furthermore, the experiment just described pushes the control systems<br />

to its limits. Slower platform movements result in considerably less overshooting <strong>of</strong> the<br />

upper body.<br />

After the robot has stabilized its posture again, the joint torques and the measured<br />

accelerations decline. As the remaining torques rest within the limits being accepted as<br />

optimized posture, the standing control system switches back to the relaxed posture phase<br />

and slowly decreases the joint stiffness.<br />

Asinthepreviousexperiment, aseries<strong>of</strong>runswithdifferentdisturbanceintensitieshasbeen<br />

performed. Table 7.2 lists the torques and accelerations occurring at forward and backward<br />

platform movement <strong>of</strong> different velocities. At the evaluated platform displacement <strong>of</strong><br />

12cm, the maximum forward velocity the robot can still compensate is about 2.0 m/s, the<br />

maximum backward velocity about 2.5 m/s. The difference stems from the longer foot<br />

leverage in toe direction, as most stabilization origins from ankle torque. With decreasing<br />

platform velocity, the observed ankle torque is only slightly reduced, whereas the reaction<br />

<strong>of</strong> the more distal joints strongly declines. This behavior can be explained by the fact<br />

that stronger platform motions induce greater displacement <strong>of</strong> the center <strong>of</strong> mass. This


7.3. Posture Optimization 133<br />

again results in greater body pitch angles and thereby in higher torque requirements for<br />

the joints closer to the center <strong>of</strong> mass.<br />

7.3 Posture Optimization<br />

A last series <strong>of</strong> experiments shall illustrate the behavior <strong>of</strong> the posture optimization control<br />

units. To this end, two different types <strong>of</strong> disturbances are applied to the robot: on the one<br />

hand, the platform the robot is standing on is rotated about the y-axis, on the other hand<br />

the robot is equipped with a backpack <strong>of</strong> changing weight. In both cases, the robot starts<br />

with an already optimized posture and has to adapt to the changing situation.<br />

Figure 7.7 shows the control behavior during a modification <strong>of</strong> the support surface. The<br />

platform the robot is standing on rotates at an angular velocity <strong>of</strong> 1.15 ◦ /s about the y-axis,<br />

i.e. it tilts forward. The rotation is stopped after three seconds at an angle <strong>of</strong> about 3 ◦ ;<br />

the time interval <strong>of</strong> the platform motion is marked by the gray bar. As the accelerations<br />

remain below the threshold for switching to the stabilization motion phase, the angle<br />

adjustments control units are continuously stimulated. The units only show activity if the<br />

robot is standing still, stronger motions induced by their own adjustments result in short<br />

intervals <strong>of</strong> inactivity. The two Relax Joint reflexes plotted in the figure try to reduce the<br />

torque acting in the corresponding joint, whereas the Balance Ankle Pitch reflex tries to<br />

keep the relative cop position shown in the upper right at its equilibrium point below the<br />

ankle joint. All three control units adjust the posture by passing an angle change rate to<br />

the appropriate Hold Position reflex.<br />

It can be observed that most <strong>of</strong> the change in angle and the highest reflex activity is<br />

accomplished in the ankle joint, compensating most <strong>of</strong> the rotation <strong>of</strong> the supporting<br />

platform. Additional adaptation is done in the hip and spine joints, the angle <strong>of</strong> the<br />

knee joint remains constant. In contrast to what could be expected, the sum <strong>of</strong> angle<br />

adjustments exceeds the amount <strong>of</strong> platform rotation. The posture assumed after the<br />

optimization is slightly shifted to the back. But actually, this posture is superior to the<br />

one with an upright trunk, as the rotation <strong>of</strong> the platform and thereby the rotation <strong>of</strong> the<br />

ankle joint shifts the cop in direction <strong>of</strong> the toes. This reduces the stability margin in<br />

forward direction and can be compensated by the observed backward recline.<br />

The second experiment on posture optimization deals with different weight distributions <strong>of</strong><br />

the robot itself. For this purpose, a virtual backpack is mounted to the torso <strong>of</strong> the robot.<br />

The weight <strong>of</strong> the backpack can be changed at random. Figure 7.8 depicts the setup <strong>of</strong><br />

the experiments and shows the final postures after increasing backpack weight to 5, 10,<br />

and 15kg. Figure 7.9 illustrates the control behavior with increasing backpack weight up<br />

to 20kg in step <strong>of</strong> 5kg.<br />

The postural reflexes are parametrized identically to the previously described experiment.<br />

Nonetheless the adapted posture differs significantly. Instead <strong>of</strong> compensating the disturbance<br />

mainly by change <strong>of</strong> angle in the ankle, the spine and hip joint show the strongest<br />

adjustments. The reason for this can be found in the optimization objective to minimize<br />

joint torques. Compensating the additional backpack weight purely by changing the ankle<br />

joint to regain equilibrium while keeping the body rigid would maintain considerable<br />

torques in spine and hip. The final posture at a backpack weight <strong>of</strong> 20kg results in a torso<br />

pitch <strong>of</strong> approximately 13 ◦ , a fact also visible in the raw acceleration values in x-direction<br />

which increasingly includes a gravitational component.


134 7. Balanced Standing Experiments<br />

Platform Rot. [ ◦ ]<br />

Torso Pitch [ ◦ ]<br />

Spine Y Angle [ ◦ ]<br />

Hip Y Angle [ ◦ ]<br />

Ankle Y Angle [ ◦ ]<br />

Hip Y Torque [Nm]<br />

3.0<br />

2.0<br />

1.0<br />

0.0<br />

2.0<br />

0.0<br />

−2.0<br />

4.0<br />

2.0<br />

0.0<br />

0.0<br />

−2.0<br />

−4.0<br />

2.0<br />

0.0<br />

−2.0<br />

10.0<br />

5.0<br />

0.0<br />

−5.0<br />

0.0 4.0 8.0 12.0<br />

0.0 4.0 8.0 12.0<br />

0.0 4.0 8.0 12.0<br />

0.0 4.0 8.0 12.0<br />

0.0 4.0 8.0 12.0<br />

0.0 4.0 8.0 12.0<br />

Time [s]<br />

CoP Position<br />

Accel. X [ m/s 2 ]<br />

Relax Spine Y a<br />

Relax Hip Y a<br />

Bal. Ankle Pitch a<br />

Spine Y Tor. [Nm]<br />

1.0<br />

0.0<br />

−1.0<br />

0.0 4.0 8.0 12.0<br />

1.0<br />

0.0<br />

−1.0<br />

0.0 4.0 8.0 12.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 4.0 8.0 12.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 4.0 8.0 12.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 4.0 8.0 12.0<br />

10.0<br />

5.0<br />

0.0<br />

−5.0<br />

0.0 4.0 8.0 12.0<br />

Time [s]<br />

Figure 7.7: Experiment on posture optimization during platform rotation. The gray bar marks<br />

the duration <strong>of</strong> the platform motion. The angular velocity <strong>of</strong> the platform rotation is 1.15 ◦ /s.


7.3. Posture Optimization 135<br />

Figure 7.8: Experiment on posture optimization while changing weight distribution by different<br />

weights <strong>of</strong> a backpack.<br />

As illustrated in Figure 7.9, the posture adaptation is generated by reflex activity appearing<br />

as soon as the backpack weight is increased. The activity abates as the robot approaches<br />

a balanced posture. Indeed, the necessary torque in the spine and the hip cannot be<br />

completely eliminated, but only a small amount <strong>of</strong> a few Nm remains. The cop position is<br />

shifted back to zero by respective adjustment <strong>of</strong> the ankle joint, thus bringing the robot<br />

back to equilibrium.<br />

The experiment has been continued with increasing backpack weight. Starting at a weight<br />

<strong>of</strong> about 30kg the postural reflexes cannot fully stabilize the posture any longer. The<br />

robot starts to sway. At a backpack load <strong>of</strong> about 70kg, i.e. nearly the weight <strong>of</strong> the<br />

whole robot, standing cannot be sustain and the robot falls over.<br />

It should again be noted that the posture adaptation relies purely on local reflexes acting<br />

at a balanced rate <strong>of</strong> change. The optimization process does neither require a whole body<br />

model nor a global assessment <strong>of</strong> the posture. As a drawback, the suggested method might<br />

stabilize in a local minimum and takes relatively long to reach it as a static posture is<br />

assumed. Too fast changes would induce dynamics distorting the measured joint torques<br />

that are evaluated for postural corrections. The method also depends on equal rates <strong>of</strong><br />

adaptation <strong>of</strong> all reflexes involved to reduce oscillation and to facilitate the stabilization <strong>of</strong><br />

the process. In return, it allows to cope with different and unforeseen kinds <strong>of</strong> disturbances<br />

in a robust and intuitive way.<br />

Comparing the robot’s postures after optimization with those assumed by human subjects<br />

in similar situations, a strong resemblance can be observed. This might justify the<br />

conclusion that humans use analog, local strategies focused on necessary muscle work while<br />

optimizing their posture against static disturbances <strong>like</strong> ground variations or additional<br />

weight attached to the body.


136 7. Balanced Standing Experiments<br />

Backp. Weight [kg]<br />

Torso Pitch [ ◦ ]<br />

Spine Y Angle [ ◦ ]<br />

Hip Y Angle [ ◦ ]<br />

Ankle Y Angle [ ◦ ]<br />

Hip Y Torque [Nm]<br />

20<br />

10<br />

0<br />

15<br />

10<br />

5<br />

0<br />

0<br />

−5<br />

−10<br />

0<br />

−5<br />

−10<br />

0<br />

−5<br />

−10<br />

0<br />

−10<br />

−20<br />

0.0 5.0 10.0 15.0<br />

0.0 5.0 10.0 15.0<br />

0.0 5.0 10.0 15.0<br />

0.0 5.0 10.0 15.0<br />

0.0 5.0 10.0 15.0<br />

0.0 5.0 10.0 15.0<br />

Time [s]<br />

CoP Position<br />

Accel. X [ m/s 2 ]<br />

Relax Spine Y a<br />

Relax Hip Y a<br />

Bal. Ankle Pitch a<br />

Spine Y Tor. [Nm]<br />

1.0<br />

0.0<br />

−1.0<br />

0.0 5.0 10.0 15.0<br />

0.0<br />

−1.0<br />

−2.0<br />

1.0<br />

0.5<br />

0.0 5.0 10.0 15.0<br />

0.0<br />

0.0 5.0 10.0 15.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 5.0 10.0 15.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 5.0 10.0 15.0<br />

0<br />

−10<br />

−20<br />

0.0 5.0 10.0 15.0<br />

Time [s]<br />

Figure 7.9: Experiment on posture optimization while changing weight distribution by different<br />

weights <strong>of</strong> a backpack. The dashed lines mark the changes <strong>of</strong> weight ranging from 0 to 20kg.


8. Dynamic <strong>Walking</strong> Experiments<br />

To evaluate the performance <strong>of</strong> the suggested control concept for dynamic walking, experiments<br />

have been performed in the test environment as presented in Chapter 6 and<br />

Appendix B. Normal walking on flat ground is analyzed in depth, and the robustness <strong>of</strong> the<br />

control system is examined by changing the ground geometry or by applying other external<br />

disturbances. Throughout all experimental runs, the same set <strong>of</strong> control parameters is<br />

used, and no adaptations are made to handle any particular situation.<br />

8.1 Normal <strong>Walking</strong><br />

Thefirstpart<strong>of</strong>thischapterfocusesonnormalwalkingoneventerrainwithoutthepresence<br />

<strong>of</strong> external disturbances. It gives a detailed evaluation <strong>of</strong> the internal mechanisms <strong>of</strong> the<br />

control system and the interaction <strong>of</strong> the control units. A kinematic analysis reveals the<br />

emerging joint trajectories and the absolute movements <strong>of</strong> the robot’s segments, allowing<br />

to conclude on gait parameters <strong>like</strong> the walking velocity or the step length. Studying the<br />

kinetic results gives insights into the necessary joint torques or the ground reaction forces<br />

occurring during a gait cycle, and allows to estimate locomotion efficiency. Comparison<br />

to both kinematic and kinetic data <strong>of</strong> human walking can expose differences as well as<br />

similarities suggesting common control strategies. Appendix B includes an image series <strong>of</strong><br />

one gait cycle <strong>of</strong> normal walking on level ground recorded in the simulation framework.<br />

8.1.1 Locomotion Modes<br />

The highest level <strong>of</strong> the biped control system is composed <strong>of</strong> the locomotion modes. During<br />

walking initiation, the interplay <strong>of</strong> these behaviors is essential to allow for a fast and<br />

smooth start <strong>of</strong> walking. This task is managed by the <strong>Walking</strong> control unit.<br />

Figure 8.1 illustrates this process by plotting the flow <strong>of</strong> activity between the involved<br />

behaviors. The Standing mode is stimulated and active by default. The stimulation <strong>of</strong> the<br />

<strong>Walking</strong> locomotion mode as initiated by the operator is marked by the dotted, vertical<br />

line. The walking mode directly stimulates the <strong>Walking</strong> Initiation mode. During this first<br />

part <strong>of</strong> walking, both the standing and the initiation behaviors are active. The compliant


138 8. Dynamic <strong>Walking</strong> Experiments<br />

Standing a<br />

Cyclic <strong>Walking</strong> a<br />

Left Foot Load<br />

1.0<br />

0.5<br />

0.0<br />

0.0 1.0 2.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 1.0 2.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 1.0<br />

Time [s]<br />

2.0<br />

<strong>Walking</strong> a<br />

<strong>Walking</strong> Init. a<br />

Right Foot Load<br />

1.0<br />

0.5<br />

0.0<br />

0.0 1.0 2.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 1.0 2.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 1.0<br />

Time [s]<br />

2.0<br />

Figure 8.1: Activity <strong>of</strong> the locomotion modes and normalized foot load. The dashed line marks<br />

the beginning <strong>of</strong> walking stimulation initiated by the operator.<br />

joint settings <strong>of</strong> the current standing posture are superimposed with the torque commands<br />

<strong>of</strong> the motor patterns stimulated by the initiation control unit.<br />

In a next step, the cyclic walking behavior is stimulated. Its activity is passed as inhibiting<br />

signal to the standing behavior, thus terminating its activity and giving full control to<br />

the walking behavior. At a time <strong>of</strong> 1.0sec in Figure 8.1, the flow <strong>of</strong> activity between<br />

standing and cyclic walking can be observed. It takes only about 0.6sec from stimulating<br />

the walking behavior to the activity <strong>of</strong> cyclic walking. During this first moments <strong>of</strong> cyclic<br />

walking, walking initiation is still active as some joints require different compliance during<br />

the first step compared to later steps <strong>of</strong> normal walking. Only when the first swing leg<br />

takes <strong>of</strong>f, the stimulation <strong>of</strong> the walking initiation behavior is pulled down.<br />

The actual points in time when the walking behavior changes the stimulation <strong>of</strong> its two<br />

sub-modes are determined by evaluating the load <strong>of</strong> the prospective stance and swing<br />

leg as depicted by the lower two plots <strong>of</strong> Figure 8.1. Algorithm 8.1 details this decision<br />

making. The target rating <strong>of</strong> the initiation behavior indicates its contentment as the load<br />

<strong>of</strong> the swing leg falls below the load <strong>of</strong> the stance leg and is passed as sensor information<br />

to the walking locomotion mode. The latter behavior uses this load comparison to decide<br />

when to stimulate cyclic walking and when to reduce the stimulation <strong>of</strong> walking initiation.<br />

The first switch is done as soon as the load <strong>of</strong> the swing leg is sufficiently below the one<br />

<strong>of</strong> the stance leg (t = 1.0sec in Figure 8.1). <strong>Walking</strong> initiation is fully terminated when<br />

there is no more load at all on the swing leg (t = 1.15sec).


8.1. Normal <strong>Walking</strong> 139<br />

Algorithm 8.1: Stimulation calculation <strong>of</strong> <strong>Walking</strong> Initiation and Cyclic <strong>Walking</strong><br />

1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

calculate value for load comparison (target rating <strong>of</strong> <strong>Walking</strong> Initiation):<br />

if load swing leg ≤ 0 then<br />

load comparison = 0<br />

end<br />

else if load stance leg ≤ 0 or load swing leg > load stance leg then<br />

load comparison = 1<br />

end<br />

else<br />

load comparison = load swing leg/load stance leg<br />

end<br />

calculate walking mode stimulation within locomotion mode <strong>Walking</strong>:<br />

if load comparison ≤ 0.8 then<br />

stimulate Cyclic <strong>Walking</strong><br />

end<br />

if Cyclic <strong>Walking</strong> stimulated and load comparison = 0 then<br />

reduce stimulation <strong>of</strong> <strong>Walking</strong> Initiation<br />

end<br />

8.1.2 Switching <strong>Walking</strong> Phases<br />

Following the walking initiation, the control system switches to cyclic locomotion. As<br />

described in Section 5.4.3, cyclic walking is composed <strong>of</strong> five consecutive motion phases.<br />

The responsible spinal pattern generator (spg) manages the phase sequence and stimulates<br />

the respective motion phase control units for the left and right body side.<br />

Figure 8.2 illustrates the sequence <strong>of</strong> phases for the first 18 seconds <strong>of</strong> normal walking<br />

by plotting the left and right states <strong>of</strong> the walking spg. The dashed, vertical line marks<br />

the beginning <strong>of</strong> activity <strong>of</strong> the Cyclic <strong>Walking</strong> control unit. As the left leg is to take the<br />

first step, the spg starts by stimulating the fourth phase <strong>of</strong> this body side. The opposite<br />

side begins with the corresponding phase one. From then on, the sequence <strong>of</strong> walking<br />

phases complies with the order prescribed by the spg. Exemplary, the figure also plots the<br />

activity <strong>of</strong> the second walking phase <strong>of</strong> the right body side. These individual behaviors<br />

are stimulated by the spg according to its current state.<br />

To further explain the mechanisms during phase transitions, Figure 8.3 enlarges seconds<br />

two and three <strong>of</strong> the previous figure. The upper plot now contains the walking phases <strong>of</strong><br />

both the left and right body side. The phase transitions are marked by the thin, vertical<br />

lines. In this representation, the bilateral synchronization <strong>of</strong> both sides can clearly be<br />

identified. This corresponds to the observations by Ivanenko et al. during their emg<br />

analysis <strong>of</strong> normal walking <strong>of</strong> human subjects.<br />

In the control system described here, this bilateral synchronization is achieved by switching<br />

the left and right motion phases on the basis <strong>of</strong> the same sensor events. Those events<br />

responsible for the phase transitions are tagged by the thick vertical lines in the lower four<br />

plots in Figure 8.3. The second and third row show the measurements <strong>of</strong> the four force<br />

sensors mounted in each foot. The solid lines represent the inner and outer toe sensors,<br />

the dashed lines those two sensors mounted in the heel. Plots four and five <strong>of</strong> the figure<br />

depict the knee angles <strong>of</strong> both legs.


140 8. Dynamic <strong>Walking</strong> Experiments<br />

Left Phase<br />

Right Phase<br />

Right Phase 2 a<br />

5<br />

4<br />

3<br />

2<br />

1<br />

5<br />

4<br />

3<br />

2<br />

1<br />

1.0<br />

0.5<br />

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0<br />

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0<br />

0.0<br />

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0<br />

Time [s]<br />

Figure 8.2: Motion phases <strong>of</strong> the left and the right body side during normal walking. The<br />

lower plot exemplifies the activity <strong>of</strong> an individual motion phase stimulated by the spg, in this<br />

case the right <strong>Walking</strong> Phase 2.<br />

The transition from phase one to phase two <strong>of</strong> the stance leg and from phase four to five<br />

<strong>of</strong> the swing leg is triggered by the lock knee event. It is based on the knee angle <strong>of</strong> the<br />

swing leg. During the swing motion, the knee passively bends due to mass inertia and then<br />

extends again as the hip joint decelerates. As soon as the leg is stretched, the heel strike<br />

is expected, and the spg switches to the next phase. Three <strong>of</strong> these events are marked in<br />

the knee angle plots <strong>of</strong> the figure.<br />

The next phase transition takes place after the heel strike is finished and the foot has made<br />

full contact with the ground. This fact can be extracted from the force measurements<br />

by the four sensors attached in the corners <strong>of</strong> each foot. As shown in the figure, first the<br />

heel sensors represented by the dashed lines detect contact with the ground. As soon as<br />

the toe sensors depicted as solid lines also touch down, the phase transition is triggered.<br />

The small gap between the actual full contact shown in the plot and the phase switch can<br />

be explained by two facts: First, a certain threshold must be exceeded before contact is<br />

confirmed. Secondly, the control system’s cycle time <strong>of</strong> 25ms results in a sampling rate <strong>of</strong><br />

40Hz clearly being visible in the zoom level <strong>of</strong> this figure.<br />

The final phase transition event is also based on the force sensors <strong>of</strong> the feet. Phase four<br />

starts as soon as the load on the prospective swing leg falls below a threshold <strong>of</strong> about<br />

150N. The force plots in the figure illustrate the fact that first the heel <strong>of</strong> the foot leaves<br />

the ground. The Leg Propel motor pattern generates the plantarflexion <strong>of</strong> the foot, thus<br />

pushing the toe sensors towards the ground to support the forward motion <strong>of</strong> the robot.


8.1. Normal <strong>Walking</strong> 141<br />

<strong>Walking</strong> Phases<br />

Left Foot Forces [N]<br />

Right Foot Forces [N]<br />

Left Knee Angle [ ◦ ]<br />

Right Knee Angle [ ◦ ]<br />

5<br />

4<br />

3<br />

2<br />

1<br />

400<br />

200<br />

0<br />

400<br />

200<br />

0<br />

60<br />

40<br />

20<br />

right<br />

left<br />

2.0 2.5 3.0 3.5 4.0<br />

2.0 2.5 3.0 3.5 4.0<br />

2.0 2.5 3.0 3.5 4.0<br />

0<br />

2.0 2.5 3.0 3.5 4.0<br />

60<br />

40<br />

20<br />

0<br />

2.0 2.5 3.0<br />

Time [s]<br />

3.5 4.0<br />

Figure 8.3: Transitions <strong>of</strong> motion phases are triggered by three events toe <strong>of</strong>f, full contact, and<br />

locked knee. Thick vertical lines mark the decisive events in the respective plots. Solid lines in<br />

foot force plots represent the inner and outer toe forces, dashed lines those <strong>of</strong> the heel sensors.


142 8. Dynamic <strong>Walking</strong> Experiments<br />

When this torque pattern subsides and the Active Hip Swing motor pattern indirectly<br />

initiates the knee bending, the load is taken from the toe sensors and the phase transition<br />

will take place.<br />

8.1.3 Behavioral Activity<br />

With each state transition <strong>of</strong> the walking spg, the stimulation proceeds from one motion<br />

phase to the next. Each motion phase in then responsible for stimulating those reflexes<br />

and motor patterns that should be active during this part <strong>of</strong> the walking cycle. This<br />

section illustrates the flow <strong>of</strong> stimulation and activity to clarify the behavioral interaction.<br />

8.1.3.1 Motor Pattern Activity<br />

Motor patterns act as the feed-forward control units <strong>of</strong> the control system. Therefore their<br />

activity only depends on their activation ι and not on any sensor information. Figure 8.4<br />

shows the activation and the activity <strong>of</strong> all participating motor patterns during the same<br />

run <strong>of</strong> normal walking as used in the previous plots. The first row <strong>of</strong> the figure depicts<br />

the walking phases <strong>of</strong> the selected 2.5sec interval, the thin vertical lines mark the phase<br />

transitions. The shaded plots in the remaining rows represent the behavior activation ι, the<br />

solid lines stand for their activity a. Motion phases always fully stimulate the respective<br />

motor patterns by a value <strong>of</strong> s = 1.<br />

The Arm Swing motor patterns are stimulated in phases three and four to correspond to<br />

the leg swing. Since the pattern generates a backward torque impulse <strong>of</strong> the ipsilateral<br />

and a forward impulse for the contralateral arm, the left pattern is activated during the<br />

left leg swing and vice versa. As soon as the control unit is activated, it starts its activity<br />

corresponding to equation 6.3 as discussed in Section 6.2.2. The behavior is stimulated<br />

long enough for the pattern to finish, and no inhibition interrupts is progress.<br />

The Initiate Swing motor pattern is stimulated from phase three onwards. Hence its<br />

pattern generation is started as soon as the foot load is low enough. Otherwise the high<br />

friction between the foot and the ground would not allow the foot to lift, and the hip<br />

torque would induce a rotation <strong>of</strong> the robot about the z-axis as well as a forward bending<br />

<strong>of</strong> the hip and the upper body. Due to the inhibition <strong>of</strong> the motor pattern by the activity<br />

<strong>of</strong> the Lock Hip reflex, the activation ι = s·(1−i) <strong>of</strong> the behavior can be reduced despite<br />

the fact that it is still stimulated by the motion phase. The step-<strong>like</strong> progress <strong>of</strong> the<br />

activation plots in the figure must be accounted for by the control cycle frequency <strong>of</strong> 40Hz.<br />

As the activity <strong>of</strong> a behavior must not exceed its activation, the motor pattern generation<br />

is suppressed.<br />

During motion phases two and three, the Leg Propel motor pattern is stimulated. As no<br />

inhibition signal is connected to the behavior, its activation lasts until the stimulation is<br />

terminated. Similar to the Initiate Swing control unit, the pattern is suppressed if the<br />

activation decreases before the pattern is completed. This behavior makes perfect sense as<br />

the control unit should not continue the plantarflexion after lift-<strong>of</strong>f, but should still propel<br />

the body forward in case the stance phase is prolonged.<br />

Finally, the Weight Acceptance motor pattern shows an exception in its activity calculation.<br />

As it needs full influence on the knee angle just before and during heel strike, the activity<br />

is set to the full value <strong>of</strong> activation. In doing so, the knee can be bent right before heel


8.1. Normal <strong>Walking</strong> 143<br />

<strong>Walking</strong><br />

Phases<br />

Left Arm<br />

Swing ι, a<br />

Right Arm<br />

Swing ι, a<br />

Left Init.<br />

Swing ι, a<br />

Right Init.<br />

Swing ι, a<br />

Left Leg<br />

Propel ι, a<br />

Right Leg<br />

Propel ι, a<br />

Left Weight<br />

Accept. ι, a<br />

Right Weight<br />

Accept. ι, a<br />

5<br />

4<br />

3<br />

2<br />

1<br />

right<br />

left<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0 2.5 3.0 3.5 4.0 4.5<br />

Time [s]<br />

Figure 8.4: Activation ι and activity a <strong>of</strong> the motor patterns during normal walking. The<br />

shaded plots represent the activation <strong>of</strong> the behaviors.


144 8. Dynamic <strong>Walking</strong> Experiments<br />

strike to reduce the impact, and its angle can be controlled during the first part <strong>of</strong> the<br />

stance phase to achieve the semi-flat walking gait. The actual torques generated during<br />

this phase will be discussed below.<br />

8.1.3.2 Reflex Activity<br />

Un<strong>like</strong> motor patterns, reflexes are feedback control units. Consequently, their activity<br />

raises due to sensor information. But <strong>like</strong> all iB2C behaviors, they must be stimulated to<br />

enable any activity. Figure 8.5 shows the activation and activity <strong>of</strong> relevant local reflexes<br />

during normal walking. Again, the figure is based on the same data as the previous plots.<br />

The Lock Knee reflex has been omitted as it only shows minor activity at the end <strong>of</strong> phase<br />

four to dampen the knee swing. Otherwise, it mainly serves as sensor pre-processor for<br />

the walking spg as its target rating indicates when the knee is stretched again at the end<br />

<strong>of</strong> the swing phase.<br />

The Lock Hip reflex is responsible for controlling the hip joint at the end <strong>of</strong> the leg swing.<br />

Its activity rises as the hip angle about the y-axis approaches its target position. With<br />

increasing activity, the dampening force is applied to the joint to slow down the leg swing.<br />

Then the hip position is retained by passing the target angle to the position control. The<br />

high activity value provides for the necessary joint stiffness. During heel strike, the reflex<br />

is still stimulated and continues to be active. Thereby the upper body is stabilized against<br />

its forward momentum during foot impact.<br />

To keep the pelvis from rotating about the hip joint’s x-axis, the Stabilize Pelvis reflex<br />

is stimulated during phases one and two, i.e. during the part <strong>of</strong> the stance phase when<br />

most <strong>of</strong> the body weight is supported. The reflex’s activity is based on the foot load <strong>of</strong> the<br />

supporting leg. As can be seen in Figure 8.5, the activity mostly parallels the activation <strong>of</strong><br />

the behavior. Only towards the end <strong>of</strong> phase two, the activity declines with the load <strong>of</strong> the<br />

support foot due to the touchdown <strong>of</strong> the contralateral foot. Small notches in the activity<br />

can appear when some amount <strong>of</strong> the body weight is supported by the contralateral leg,<br />

e.g. caused by a dragging <strong>of</strong> the foot or excessive leg propulsion. The figure also illustrates<br />

how the left and right sided instances <strong>of</strong> the reflex alternate in stabilizing the pelvis.<br />

The activity <strong>of</strong> the local reflex Heel Strike is based in the force information <strong>of</strong> the two<br />

load sensors mounted at the back <strong>of</strong> each foot. As soon as heel contact is measured, the<br />

reflex activity increases and a dampening force is applied to the ankle joint based on its<br />

rotational velocity. This action prevents the foot from slapping down at impact in case the<br />

mechanical damping is not sufficient. The activity declines together with the stimulation<br />

<strong>of</strong> the reflex at the end <strong>of</strong> phase five.<br />

The Cutaneous Reflex is stimulated during the swing phase <strong>of</strong> the leg. In normal walking<br />

on level terrain it should show only minor activity. While the left sided reflex remains<br />

nearly fully inactive during the interval shown in the figure, the right sided reflex displays<br />

at least a small peak <strong>of</strong> activity. This peak is caused by a remaining ground contact <strong>of</strong> the<br />

toes during the begin <strong>of</strong> the swing phase. The reflex makes sure that the foot completely<br />

loses contact to the ground by a short torque impulse in the ankle and the knee joint.<br />

It is interesting to note that relevant activity <strong>of</strong> some reflexes is necessary during normal<br />

walking while other reflexes mainly handle disturbances. Similar observations can be<br />

made on the reflex action in human walking. This phenomenon can also be noticed in the<br />

activity <strong>of</strong> postural reflexes as illustrated by Figure 8.6. In contrast to the local reflexes,


8.1. Normal <strong>Walking</strong> 145<br />

<strong>Walking</strong><br />

Phases<br />

Left Lock<br />

Hip ι, a<br />

Right Lock<br />

Hip ι, a<br />

L. Stabilize<br />

Pelvis ι, a<br />

R. Stabilize<br />

Pelvis ι, a<br />

Left Heel<br />

Strike ι, a<br />

Right Heel<br />

Strike ι, a<br />

L. Cutaneous<br />

Reflex ι, a<br />

R. Cutaneous<br />

Reflex ι, a<br />

5<br />

4<br />

3<br />

2<br />

1<br />

right<br />

left<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0 2.5 3.0 3.5 4.0 4.5<br />

Time [s]<br />

Figure 8.5: Activation ι and activity a <strong>of</strong> the local reflexes during normal walking. The shaded<br />

plots represent the activation <strong>of</strong> the behaviors. The Lock Knee reflex has been omitted.


146 8. Dynamic <strong>Walking</strong> Experiments<br />

here those control units responsible for compensating disturbances predominate, but also<br />

show substantial activity during normal walking.<br />

The postural reflex Upright Trunk is stimulated throughout all phases <strong>of</strong> cyclic walking<br />

to achieve maximum stability <strong>of</strong> the upper body. But as the reflex cannot and should<br />

not influence the hip joint in case the leg <strong>of</strong> the considered body side has no ground<br />

contact, the activity function includes the normalized foot load as factor. Thereby the<br />

reflex suspends its activity during the leg swing in phase four and the beginning <strong>of</strong> phase<br />

five. As soon as the leg has regained ground contact, the reflex resumes its task. Figure 8.6<br />

also illustrates the fact that during double support around phase three both the left and<br />

the right Upright Trunk control units are active. This allows an even better stabilization<br />

<strong>of</strong> the upper body as the action <strong>of</strong> both reflexes work in concert. Furthermore, it can<br />

be observed that the behavior shows significant activity even during normal walking in<br />

even terrain. This is caused by the considerable mass inertia <strong>of</strong> the upper body keeping it<br />

moving forward while the pelvis decelerates at heel strike, and vice versa during propulsion.<br />

Likewise, the torque generated by the Initiate Swing motor pattern not only causes the<br />

leg to swing but also effects in a bending <strong>of</strong> the trunk. These self-induced disturbances are<br />

partly compensated by anticipatory torque impulses produced by the responsible motor<br />

patterns, but must additionally be balanced by the Upright Trunk reflexes. Accordingly,<br />

the highest activity <strong>of</strong> these behaviors can be observed right after heel strike.<br />

The Forward Velocity control unit is stimulated by walking phase one. During this part <strong>of</strong><br />

the walking cycle, the foot has full ground contact and the postural reflex can influence<br />

the forward velocity by applying a correcting torque to the ankle joint. Again, relevant<br />

activity can be observed even during normal walking, however not as strong as in the<br />

previously discussed reflex. Movements <strong>of</strong> the upper body and the swinging extremities<br />

can cause variations in the forward velocity, or more precisely in the estimated position<br />

<strong>of</strong> the extrapolated center <strong>of</strong> mass (xcom). These disturbances are then compensated by<br />

the reflex. Also, it cannot be avoided that the xcom estimation is inaccurate, resulting in<br />

slight reflex action that would not have been necessary otherwise. Nevertheless, the xcom<br />

calculations are sufficient for stabilizing the forward velocity.<br />

The same considerations hold true for the Lateral Balance Ankle behavior. It is also<br />

stimulated during walking phase one as it balances the robot by exerting torques via the<br />

ankle joint. Even on level ground as in the run presented in Figure 8.6, considerable<br />

activity can be observed. This behavior can be explained by the fact that the reflex<br />

can only fine-tune the posture due to the short sidewards leverage <strong>of</strong> the feet, so the<br />

maximum activity is already reached at minor disturbances. Combined with appropriate<br />

foot placement, this reflex action contributes to the lateral stability.<br />

Concluding the discussion on behavioral activity, the Lateral Foot Placement control<br />

unit shows a different activity signature than the other postural reflexes. As it controls<br />

the adduction and abduction <strong>of</strong> the leg by passing a target angle and a low compliance<br />

value to the position control <strong>of</strong> the hip joint about the x-axis, the behavior must be<br />

active throughout the swing phase. Its activity is again based on the normalized foot<br />

load because influencing the foot position is only possible during ground clearance. The<br />

activity value is used directly as stiffness demand for the position controller as described<br />

in Section 6.1.2. Thus, the behavior activity cannot serve as indication on the amount <strong>of</strong><br />

lateral foot placement correction. Rather, the behavior’s target rating should be consulted<br />

for that information. Being stimulated during phases four and five, the reflex can start its


8.1. Normal <strong>Walking</strong> 147<br />

<strong>Walking</strong><br />

Phases<br />

L. Upright<br />

Trunk ι, a<br />

R. Upright<br />

Trunk ι, a<br />

L. Forward<br />

Velocity ι, a<br />

R. Forward<br />

Velocity ι, a<br />

L. Lat. Bal.<br />

Ankle ι, a<br />

R. Lat. Bal.<br />

Ankle ι, a<br />

L. Lat. Foot<br />

Placem. ι, a<br />

R. Lat. Foot<br />

Placem. ι, a<br />

5<br />

4<br />

3<br />

2<br />

1<br />

right<br />

left<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0<br />

1<br />

2.5 3.0 3.5 4.0 4.5<br />

0<br />

2.0 2.5 3.0 3.5 4.0 4.5<br />

Figure 8.6: Activation ι and activity a <strong>of</strong> the postural reflexes during normal walking. The<br />

shaded plots represent the activation <strong>of</strong> the behaviors.


148 8. Dynamic <strong>Walking</strong> Experiments<br />

task as soon as the foot leaves the ground and can continue to correct the foot position<br />

right up to the point when the foot touches down.<br />

In summary, three different types <strong>of</strong> behavioral activity can be observed in the low-level<br />

control units: motor patterns show a predefined progression <strong>of</strong> activity, only depending on<br />

the starting point when the pattern is triggered. Some reflexes <strong>like</strong> the Lock Hip reflex<br />

are an essential part <strong>of</strong> the walking cycle, showing high activity at certain kinematic or<br />

kinetic events to contribute to normal walking. Finally, the remainder <strong>of</strong> reflexes work<br />

towards stabilizing the robot. They display more or less activity depending on internal or<br />

external disturbances.<br />

8.1.4 Kinematic Analysis<br />

The kinematic analysis <strong>of</strong> gait handles joint angles trajectories and the translation and<br />

rotation <strong>of</strong> body segments relative to a fixed coordinate system. This kind <strong>of</strong> evaluation<br />

allows to gain a better understanding <strong>of</strong> the resulting motions by considering angle ranges,<br />

variations in trajectories, or locomotion velocity. Furthermore, the results can be compared<br />

to human walking. Other than in clinical gait studies, in the experiments presented here<br />

the relevant data can directly be extracted from the simulation environment.<br />

8.1.4.1 Joint Angles<br />

Figure 8.7 presents the angle trajectories <strong>of</strong> all joints over the course <strong>of</strong> one gait cycle. To<br />

this end, 30 consecutive steps <strong>of</strong> walking on level ground are averaged by manually tagging<br />

the intervals from one heel strike to the next. Each interval is then normalized in length,<br />

and the mean, minimum, and maximum values are calculated. The 15 left and the 15 right<br />

data sets are combined while taking sign conversions into account due to the mirroring<br />

along the sagittal plane. Positive values express a joint flexion, abduction, or dorsiflexion,<br />

while negative values signify extension, adduction, or plantarflexion, respectively. This<br />

complies to the standards in biomechanical analysis and facilitates the comparison <strong>of</strong> the<br />

trajectories to human data. The average duration <strong>of</strong> the gait cycles is 1.27sec, with the<br />

shortest duration being 1.20sec and the longest duration amounting to 1.30sec. In the<br />

figure, the mean values are represented by the solid lines, the minimum and maximum<br />

values are indicated by the dashed lines. The vertical dashed line at 69% marks the mean<br />

location <strong>of</strong> the transition from stance to swing. The left column <strong>of</strong> plots depicts the joints<br />

in the sagittal plane, i.e. revolving about the y-axis, while the right column presents the<br />

remaining joints.<br />

The spine joint remains within a small angle range throughout the walking cycle since it is<br />

set to a fixed position at relatively low compliance. The joint trajectory about the y-axis<br />

exhibits a pattern that is repeated after half the cycle, once for each leg. For the x- and<br />

z-axis, a similar behavior can be observed, but the pattern swaps its sign from one side to<br />

the next. This can be explained by the fact that the forward motion <strong>of</strong> the upper body is<br />

symmetrical for each leg and runs at twice the frequency <strong>of</strong> the gait cycle. The lateral<br />

movement <strong>of</strong> the trunk is in phase with the cycle and is mirrored for each leg. Especially<br />

the rotation about the x-axis illustrates the robot’s sway from left to right during one gait<br />

cycle, as the inertia <strong>of</strong> the upper body causes a deflection about the spine joint.<br />

The sine-<strong>like</strong> motion <strong>of</strong> the shoulder joint about the y-axis was already discussed in<br />

Chapter 5. It results from torque impulses generated by the Arm Swing motor patterns


8.1. Normal <strong>Walking</strong> 149<br />

Spine Y [ ◦ ]<br />

Shoulder Y [ ◦ ]<br />

Elbow Y [ ◦ ]<br />

Hip Y [ ◦ ]<br />

Knee Y [ ◦ ]<br />

Ankle Y [ ◦ ]<br />

2<br />

0<br />

−2<br />

40<br />

20<br />

0<br />

−20<br />

0 25 50 75 100<br />

−40<br />

0 25 50 75 100<br />

15<br />

10<br />

5<br />

0<br />

40<br />

20<br />

0<br />

−20<br />

60<br />

40<br />

20<br />

−20<br />

0 25 50 75 100<br />

0 25 50 75 100<br />

0<br />

0 25 50 75 100<br />

20<br />

0<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

Spine X [ ◦ ]<br />

Spine Z [ ◦ ]<br />

Shoulder X [ ◦ ]<br />

Hip X [ ◦ ]<br />

Hip Z [ ◦ ]<br />

Ankle X [ ◦ ]<br />

2<br />

0<br />

−2<br />

5<br />

0<br />

0 25 50 75 100<br />

−5<br />

0 25 50 75 100<br />

2<br />

0<br />

−2<br />

5<br />

0<br />

0 25 50 75 100<br />

−5<br />

0 25 50 75 100<br />

4<br />

2<br />

0<br />

−2<br />

2<br />

0<br />

0 25 50 75 100<br />

−2<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

Figure 8.7: Joint angles over the course <strong>of</strong> one gait cycle. The solid lines represent the mean<br />

values over 30 steps <strong>of</strong> normal walking on even ground. The dashed lines mark the minimum<br />

and maximum values. The vertical dashed line indicates the transition from stance to swing.


150 8. Dynamic <strong>Walking</strong> Experiments<br />

and is out-<strong>of</strong>-phase with the hip trajectory. Shoulder rotation in the x-axis is restricted<br />

by a high stiffness value keeping the arm aligned to the trunk. Still, some compliance<br />

remains and the arm mass inertia combined with the body sway from left to right create a<br />

deflection <strong>of</strong> about ±1 ◦ .<br />

The elbow joint is set to low compliance throughout the walking cycle. It rests at its<br />

extension limit while the arm swings forward due to mass inertia. As the shoulder joints<br />

change direction, the elbow passively bends up to an angle <strong>of</strong> about 15 ◦ . Due to the<br />

relatively low mass <strong>of</strong> the lower arm segment, its movement has little influence on the<br />

walking process.<br />

The action <strong>of</strong> the hip, knee, and ankle joint about the y-axis has the most significant<br />

influence on forward walking. The hip joints mainly reflects the leg swing. At heel strike,<br />

the hip is flexed to its maximum and then extends during the stance phase as the body<br />

travels over the stance leg. This trajectory is not predefined, rather it emerges from the<br />

pendulum motion <strong>of</strong> the body and the work <strong>of</strong> the postural reflex Upright Trunk keeping<br />

the pelvis and the upper body in an upright orientation. As the motor pattern initiating<br />

the leg swing starts its activity, the hip joint moves forward until its position is controlled<br />

by the Lock Hip reflex. During this last part <strong>of</strong> the gait cycle, the variation in angle is<br />

nearly disappearing as the swing leg’s angle <strong>of</strong> attack is set by a position control with low<br />

compliance.<br />

Bending <strong>of</strong> the knee joint starts even before heel strike as the responsible motor pattern<br />

initiates this action at the beginning <strong>of</strong> walking phase five. The knee then flexes up to<br />

the defined angle during weight acceptance before stretching again in anticipation <strong>of</strong> the<br />

swing phase. As the hip initiates the leg swing, the knee begins to bend passively due to<br />

the mass inertia <strong>of</strong> the lower leg. Still without active control, it then extends again as<br />

the hip decelerates towards the end <strong>of</strong> the walking cycle. The trajectory <strong>of</strong> the passive<br />

knee flexion shows considerable variation caused by different starting points, varying robot<br />

velocity, and diverging stride lengths. Similar to the hip joint, the bending <strong>of</strong> the knee<br />

starts well before the swing foot leaves the ground.<br />

The onset <strong>of</strong> the ankle trajectory about the y-axis reflects the fact that the foot planarflexes<br />

as it moves from only heel contact to full ground contact. During the stance phase, the<br />

body passively swings over the foot, only influenced by the passive elasticity <strong>of</strong> the joint<br />

and the Forward Velocity reflex. As soon as the leg propulsion starts at the end <strong>of</strong> the<br />

stance phase, the ankle dorsiflexes to push the robot forward. At lift-<strong>of</strong>f, the foot travels<br />

back to minor plantarflexion set by the respective motion phase with low compliance to<br />

facilitate ground clearance during the swing phase.<br />

Similar to the spine joint, the hip joint’s z-axis reveals the small rotation <strong>of</strong> the pelvis<br />

during walking. The hip joint’s x-axis is used for stabilizing the pelvis during stance phase<br />

and adjusting the lateral foot placement during swing phase. At the beginning <strong>of</strong> the cycle<br />

the weight <strong>of</strong> the robot shifts from one leg to another, resulting in a small indulge <strong>of</strong> the<br />

pelvis before it is steady again. The sway <strong>of</strong> the upper body to the new stance leg at the<br />

middle <strong>of</strong> the gait cycle results in adduction <strong>of</strong> the hip joint by a few degrees. During<br />

the swing phase, the postural reflex Lateral Foot Placement moves the leg to the left or<br />

right depending on the com estimation. This explains the relatively large variation in<br />

joint angle. The robot sheers in the lateral direction, resulting in an <strong>of</strong>fset <strong>of</strong> the hip angle<br />

throughout the gait cycle.


8.1. Normal <strong>Walking</strong> 151<br />

Joint Minimum [ ◦ ] Maximum [ ◦ ] Joint Limits [ ◦ ]<br />

Spine X -1.1 1.2 -57 – 57<br />

Spine Y -1.9 0.9 -57 – 57<br />

Spine Z -3.2 3.2 -57 – 57<br />

Shoulder X 0.7 1.5 0 – 115<br />

Shoulder Y -33.8 33.9 -57 – 57<br />

Elbow Y -0.1 15.1 0 – 115<br />

Hip X -0.5 3.4 -57 – 57<br />

Hip Y -18.9 35.8 -10 – 49<br />

Hip Z -1.9 2.7 -57 – 57<br />

Knee Y 4.3 60.2 0 – 115<br />

Ankle X -0.3 2.4 -29 – 29<br />

Ankle Y -24.6 23.8 -29 – 29<br />

Table 8.1: Minimum and maximum values <strong>of</strong> the mean joint angle trajectories over 20 steps<br />

during normal walking on level ground.<br />

For equal reasons, a similar <strong>of</strong>fset can be observed in the ankle joint’s x-axis. During the<br />

stance phase, the robot leans to the side <strong>of</strong> the stance leg, resulting in an increasing angle.<br />

Before the swing phase, the robot sways back towards the sagittal plane. During leg swing,<br />

the ankle remains passive except for the parallel spring pulling it back to neutral position.<br />

At touchdown, the ankle adapts to the ground surface, thus, given a level ground plane,<br />

the angle only depends on the robot orientation.<br />

Itshouldbenotedagainthattheangletrajectoriesjustpresentedresultfromtheinteraction<br />

<strong>of</strong> passive dynamics including parallel and series springs, torque impulses by the motor<br />

patterns, and local reflex action. Except for a small part <strong>of</strong> the knee trajectory during<br />

weight acceptance, no joint position trajectories are predefined. The position control is<br />

only used with constant equilibrium points and compliance settings throughout a motion<br />

phase.<br />

The minimum and maximum values <strong>of</strong> the mean joint angle trajectories occurring during<br />

the presented experiment are summarized in Table 8.1. Furthermore, the joint limits are<br />

listed for comparison. It can be seen that the joints stay well below their limits that have<br />

been chosen while keeping a development <strong>of</strong> a real robot in mind. Only the hip joint<br />

shows an over-extension in its y-axis. This demonstrates that the joint limits are not<br />

implemented as fixed mechanical stops, but rather as elastic limits generating increasing<br />

torque as the joint exceeds its boundaries.<br />

8.1.4.2 Segment Translation and Rotation<br />

While the above discussion on joint angles considers relative coordinates, kinematic analysis<br />

also handles absolute poses <strong>of</strong> segments. Figure 8.8 illustrates the trajectories <strong>of</strong> the robot’s<br />

head and pelvis relative to a fixed coordinate system. The head is plotted as solid line,<br />

the pelvis is represented by the dashed line. The figure shows the first 22 seconds <strong>of</strong> the<br />

same run <strong>of</strong> normal walking used in the previous plots.<br />

The progression <strong>of</strong> both the head and the pelvis in x-direction displays a very steady<br />

velocity after the initial acceleration during walking initiation. The resulting walking


152 8. Dynamic <strong>Walking</strong> Experiments<br />

Head/Pelvis<br />

Position X [m]<br />

Head/Pelvis<br />

Position Y [m]<br />

Head/Pelvis<br />

Position Z [m]<br />

30<br />

20<br />

10<br />

0<br />

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0 20.0 22.0<br />

1<br />

0<br />

−1<br />

2<br />

1<br />

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0 20.0 22.0<br />

0<br />

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0 20.0 22.0<br />

Time [s]<br />

Figure 8.8: Head (solid line) and pelvis (dashed line) position during normal walking. Caused<br />

by slip and joint compliance, the robot shows slight derivation from straightforward direction.<br />

Left Foot<br />

Position Z [m]<br />

Right Foot<br />

Position Z [m]<br />

0.4<br />

0.2<br />

0.0<br />

0.4<br />

0.2<br />

0.0<br />

toe<br />

heel<br />

ankle<br />

walking dir.<br />

3.0 3.2 3.4 3.6 3.8 4.0 4.2 4.4 4.6 4.8 5.0<br />

0.76m 0.75m<br />

3.0 3.2 3.4 3.6 3.8 4.0 4.2 4.4 4.6 4.8 5.0<br />

Position X [m]<br />

Figure 8.9: Movement <strong>of</strong> the foot in the sagittal plane during a stride. The plots mark the<br />

trajectories <strong>of</strong> the center <strong>of</strong> the left toe, the ankle, and left heel segment the foot is composed <strong>of</strong>.


8.1. Normal <strong>Walking</strong> 153<br />

velocity in this run amounts to 1.2 m/s or 4.3 km/h. This velocity corresponds to relaxed<br />

walking <strong>of</strong> human subjects.<br />

Two characteristics can be observed in the trajectories in the lateral direction. On the one<br />

hand, head and pelvis sway slightly to the left and to the right during a walking cycle.<br />

The oscillation <strong>of</strong> the head amounts to about ±4cm, the one <strong>of</strong> the pelvis to ±3cm. On<br />

the other hand, the robot does not stay on a straight line while walking forward. Rather,<br />

it tends to show minor rotations which vary from run to run. These rotations are caused<br />

by a slip <strong>of</strong> the feet on the ground and by the compliance <strong>of</strong> the joints about the z-axis.<br />

The head and pelvis trajectories in z-direction illustrate the semi-flat walking gait already<br />

mentioned in previous chapters. During each stance phase, the body travels over the<br />

stance leg and is first lifted and than lowered again during this process. The amplitude <strong>of</strong><br />

this oscillation is reduced by the bending <strong>of</strong> the leg initiated by the Weight Acceptance<br />

motor pattern. Due to additional sway <strong>of</strong> the upper body in the sagittal plane, the head’s<br />

amplitude is slightly higher than the one <strong>of</strong> the pelvis, coming to approximately 5cm.<br />

The movement <strong>of</strong> the feet is detailed in Figure 8.9. Three trajectories in the sagittal<br />

plane for both feet are given, namely those <strong>of</strong> the inner toe, the ankle, and the inner heel<br />

segments. As the plots represent the position <strong>of</strong> the segment centers, the z-position <strong>of</strong> toe<br />

and heel does not reach zero but stays 3cm above the ground, corresponding to the radius<br />

<strong>of</strong> the two segments. To provide a reference to the progress <strong>of</strong> time, the plots are drawn<br />

with markers at a time interval <strong>of</strong> 25ms.<br />

The foot location during the stance phase can easily be identified by the clustering <strong>of</strong> plot<br />

markers. During leg propulsion, the heel lifts from the ground and the toes are pushed<br />

towards the ground by plantarflexion <strong>of</strong> the foot to increase the robots forward velocity.<br />

The toes lift <strong>of</strong>f the ground as soon as the swing phase begins and the foot rotates to a<br />

neutral ankle position. During this first part <strong>of</strong> the swing phase, the toes are the foot<br />

segment closest to the ground and determine the ground clearance <strong>of</strong> 5–6cm. In the second<br />

half <strong>of</strong> the swing phase, the knee joint extends. This causes the foot to rotate as the ankle<br />

joint stays passive and does not significantly change its angle. Thus, the toe segment<br />

moves upwards and the height <strong>of</strong> the heel segment is reduced. During this part <strong>of</strong> the<br />

swing phase, the heel is located closest to the floor and the ground clearance is reduced to<br />

only 2cm at the lowest point. Just before heel strike, this distance increases again as the<br />

knee reaches its full extension.<br />

Figure 8.9 also illustrates the step length during normal walking. Measuring the distance<br />

between the successive heel strikes <strong>of</strong> the left and the right leg amounts to 0.76m and<br />

0.75m. This length varies only slightly among the steps <strong>of</strong> the presented run. The mean<br />

step length results in 0.78m, with a minimum value <strong>of</strong> 0.73m, and a maximum value <strong>of</strong><br />

0.81m. The standard deviation <strong>of</strong> the step length amounts to 0.02m.<br />

The foot placement is further visualized by Figure 8.10, showing 28 foot points <strong>of</strong> the<br />

considered run. The consistent step length is illustrated as well as the fact that the robot<br />

does not manage to remain on a straight line, but instead displays small displacements<br />

and rotation due to slip and joint compliance. As in human walking, additional external<br />

sensor systems <strong>like</strong> vision would be necessary to correct the walking direction.<br />

The figure also allows to identify the step width. Analysis <strong>of</strong> the underlying data reveals a<br />

mean step width <strong>of</strong> 0.16m at a standard deviation <strong>of</strong> 0.03m. The minimum value <strong>of</strong> the<br />

run amounts to 0.10m, the maximum value comes to 0.22m. The higher variation <strong>of</strong> the


154 8. Dynamic <strong>Walking</strong> Experiments<br />

Foot Points<br />

Position Y [m]<br />

2.0 walking dir.<br />

0.0<br />

−2.0<br />

2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0 20.0 22.0 24.0<br />

Position X [m]<br />

Figure 8.10: Foot points on the horizontal plane during 28 steps <strong>of</strong> normal walking. The dotted<br />

line marks the trajectory <strong>of</strong> the pelvis.<br />

step width compared to the step length can be explained by the difficulty to stabilize the<br />

robot in the lateral direction due to its narrow foot placement and the high center <strong>of</strong> mass.<br />

This results in considerable activity <strong>of</strong> the postural reflex Lateral Foot Placement and thus<br />

in stronger scattering <strong>of</strong> the step width.<br />

Figure 8.11 illustrates the orientation <strong>of</strong> the head and the pelvis segment during the<br />

walking cycle. The rotation about the x-axis indicates the raise <strong>of</strong> the pelvis during the<br />

stance phase by the Stabilize Pelvis reflex. The trunk and head show even stronger sway<br />

from left to right due to the rotation about the spine joint and the considerable mass <strong>of</strong><br />

the upper body.<br />

The upper body oscillation in the sagittal plane explains the orientation <strong>of</strong> head and pelvis<br />

about the y-axis. It appears with twice the frequency <strong>of</strong> the stride, as the trunk leans<br />

forward after each heel strike. The Upright Trunk postural reflex cannot fully compensate<br />

for the mass inertia pushing the upper body forward after impact. A pelvis rotation <strong>of</strong> up<br />

to 8 ◦ remains. Again, the head displays the higher amplitude as the spine flexion is added<br />

to the pelvis rotation.<br />

The rotation <strong>of</strong> the pelvis about the z-axis mainly emerges from the compliance <strong>of</strong> the<br />

hip joint about the same axis. As the swing leg moves forward, a torque is induced to the<br />

pelvis that increases with the step length. Here, the rotation <strong>of</strong> the head and trunk is<br />

reduced by the flexibility <strong>of</strong> the spine and the compensating swing <strong>of</strong> the arms.<br />

8.1.4.3 Variation <strong>of</strong> <strong>Walking</strong> Velocity<br />

The figures so far have been based on a run <strong>of</strong> normal walking with the velocity modulation<br />

factor set to 0.8. This modulation signal influences several reflexes and motor patterns by<br />

adapting target values or scaling the amplitude <strong>of</strong> torque impulses. All these adjustments<br />

are obtained by simply multiplying the relevant values by the velocity factor. No other<br />

elaboratefunctionsareused. Bychangingthismodulationvalue, differentwalkingvelocities<br />

can be achieved. Figure 8.12 illustrates this relation by plotting the forward progression <strong>of</strong><br />

the pelvis for five different velocity factors.<br />

By varying the velocity modulation factor, walking velocities ranging from 1.1 m/s to<br />

1.4 m/s can be reached. At higher velocities, the steps lengthen and the oscillation <strong>of</strong> the<br />

upper body increases due to the stronger impact at heel strike. Slow walking reduces<br />

this oscillation in the sagittal plane, but lateral sway is still prominent. The lower torque<br />

impulse <strong>of</strong> the hip joint initiating the leg swing lessens the passive flexion angle <strong>of</strong> the


8.1. Normal <strong>Walking</strong> 155<br />

Pelvis X [ ◦ ]<br />

Pelvis Y [ ◦ ]<br />

Pelvis Z [ ◦ ]<br />

2<br />

0<br />

−2<br />

10<br />

5<br />

0<br />

10<br />

0<br />

−10<br />

0 25 50 75 100<br />

0 25 50 75 100<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

Head X [ ◦ ]<br />

Head Y [ ◦ ]<br />

Head Z [ ◦ ]<br />

2<br />

0<br />

−2<br />

10<br />

5<br />

0<br />

10<br />

0<br />

−10<br />

0 25 50 75 100<br />

0 25 50 75 100<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

Figure 8.11: Orientation <strong>of</strong> head and pelvis over the course <strong>of</strong> one gait cycle. The solid lines<br />

represent the mean values over 30 steps <strong>of</strong> normal walking, the dashed lines mark the minimum<br />

and maximum values. The vertical dashed line indicates the transition from stance to swing.<br />

Pelvis Position X [m]<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0.0 1.0 2.0 3.0 4.0<br />

Time [s]<br />

5.0 6.0 7.0 8.0<br />

velocity resulting<br />

factor: velocity:<br />

0.75 1.13m/s 0.80 1.21m/s 0.85 1.27m/s 0.90 1.34m/s 0.95 1.38m/s Figure 8.12: Pelvis progression at different walking velocities. Only the walking velocity<br />

modulation value is changed, no other parameters are adapted.


156 8. Dynamic <strong>Walking</strong> Experiments<br />

knee. This decreases the ground clearance distance and raises the probability <strong>of</strong> the robot<br />

to stumble. Table 8.2 summarizes the resulting mean velocities and step lengths during<br />

walking with different velocity modulation factors.<br />

Velocity Factor <strong>Walking</strong> Velocity Step Length Max. Ankle Torque<br />

[ m/s] [m] [N]<br />

0.75 1.13 0.68 52<br />

0.80 1.21 0.76 56<br />

0.85 1.27 0.84 60<br />

0.90 1.34 0.88 65<br />

0.95 1.38 0.92 69<br />

Table 8.2: Variation <strong>of</strong> walking velocity, stride length, and maximum ankle torque during leg<br />

propulsion <strong>of</strong> walking with different velocity modulation values.<br />

The described variations <strong>of</strong> velocity are achieved without adapting any parameter <strong>of</strong> the<br />

control units involved. Further reducing or increasing the factor causes instabilities mainly<br />

due to stumbling during leg swing or tilting in the lateral direction. By manually tuning<br />

the parameters <strong>of</strong> some control units, faster or slower walking is possible, but no extensive<br />

gain <strong>of</strong> velocity can be obtained. For very slow and deliberate walking, the fundamental<br />

form <strong>of</strong> locomotion would have to be adapted. This gait transition can also be observed in<br />

human subjects. Likewise, to further increase the speed, the locomotion form should be<br />

changed to a running gait.<br />

8.1.4.4 Comparison to human data<br />

As human morphology and motion control serves as basis for this thesis, it stands to reason<br />

to compare the results from simulation to those from human gait analysis. In 1998, the<br />

Vicon company 1 , a manufacturer <strong>of</strong> systems for biomechanical gait analysis, provided a<br />

kinematic and kinetc data set <strong>of</strong> human walking that is used as basis <strong>of</strong> the following<br />

comparisons.<br />

Figure 8.13 presents the differences and similarities <strong>of</strong> four joint angle trajectories. The<br />

solid lines represent the results from simulation, the dashed lines are taken from the Vicon<br />

data set. The first plot illustrates that humans show stronger pelvis lift during the stance<br />

phase by additional hip abduction. Still, both curves have two elevations <strong>of</strong> hip abduction<br />

in common, one occurring during stance and one during the swing phase. However, humans<br />

keep the swing leg closer to the stance leg during the swing phase.<br />

More similarities can be found in the trajectories <strong>of</strong> the hip joint’s y-axis. Both curves<br />

have the same characteristic progress. The main difference is the additional hip extension<br />

<strong>of</strong> the simulated biped towards the end <strong>of</strong> the stance phase.<br />

The knee trajectories also show a similar behavior. Here, it must be admitted that the<br />

knee angle during weight acceptance is loosely modeled after the human example, as a<br />

semi-flat walking gait is aimed for. As such, the resemblance is to be expected. However,<br />

both in human and in the bipedal robot’s walking, the knee angle during the swing phase<br />

1 http://www.vicon.com/


8.1. Normal <strong>Walking</strong> 157<br />

Hip X [ ◦ ]<br />

Knee Y [ ◦ ]<br />

10<br />

0<br />

−10<br />

60<br />

40<br />

20<br />

0<br />

0 25 50 75 100<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

Hip Y [ ◦ ]<br />

Ankle Y [ ◦ ]<br />

40<br />

20<br />

0<br />

−20<br />

20<br />

0<br />

−20<br />

0 25 50 75 100<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

Figure 8.13: Comparing joint angles in human walking (dashed lines) to those <strong>of</strong> the simulated<br />

biped (solid lines). <strong>Human</strong> data is taken from a data set provided by the Vicon company in 1998.<br />

emerges by passive dynamics and comparable trajectories arise. Generally, the peak angle<br />

<strong>of</strong> the simulated robot during this phase is slightly smaller and the duration <strong>of</strong> the knee<br />

bending is shorter. Furthermore, a small delay just before the leg swing can be observed.<br />

This is caused by the fact that, compared to human subjects, the robot lacks a small<br />

amount <strong>of</strong> forward velocity giving it more time between leg extension and touchdown <strong>of</strong><br />

the contralateral leg. In this point, the human gait clearly shows a superior optimization.<br />

As last example, the comparison <strong>of</strong> the ankle movements again reveals strong similarities.<br />

The robot tends to stronger plantarflexion after heel strike, whereas humans longer role<br />

over the heel. During this phase <strong>of</strong> point contact, no torques can be induced by ankle<br />

flexion. Again, the optimization <strong>of</strong> human walking allows this sacrifice <strong>of</strong> potential control<br />

in favor <strong>of</strong> a more efficient gait. Another difference can be found in additional plantarflexion<br />

during leg propulsion and in stronger dorsiflexion during the swing phase.<br />

Similar to the joint angle trajectories, resemblance <strong>of</strong> the absolute movements <strong>of</strong> body<br />

segments in humans and in the simulated robot can be observed. Figure 8.14 illustrates<br />

this fact using the position <strong>of</strong> the toe segment as an example. The vertical displacement <strong>of</strong><br />

the toe is plotted over the progress <strong>of</strong> one stride. The experimental data <strong>of</strong> the biped robot<br />

is averaged over 30 steps <strong>of</strong> normal walking on level terrain. As the position <strong>of</strong> the center<br />

<strong>of</strong> the inner toe segment is shown, there remains an <strong>of</strong>fset <strong>of</strong> 3cm to the ground plane.<br />

Both curves show the same distinctive course and similar amplitude. After heel strike, the<br />

toes are lowered to the ground where they stay throughout the stance phase and during<br />

leg propulsion. At the onset <strong>of</strong> the swing phase, the toes lift from the ground as the foot<br />

dorsiflexes to a neutral angle. During the leg swing, the toes first come closer to the ground<br />

as the leg traverses under the trunk. As soon as the knee extends towards the end <strong>of</strong> the<br />

swing, the foot rotates and the toes point upwards, resulting in the maximum distance to<br />

the ground.<br />

In summary, the joint angle and segment position trajectories <strong>of</strong> human subjects and the<br />

simulated robot show remarkable resemblance. The reason for this can be found in several


158 8. Dynamic <strong>Walking</strong> Experiments<br />

(a)<br />

Toe Position Z [m]<br />

0.2<br />

0.1<br />

0.0<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

Figure 8.14: Comparing the vertical toe displacement <strong>of</strong> human subjects (a) to those <strong>of</strong> the<br />

simulated biped (b), averaged over 30 steps. <strong>Human</strong> data plot taken from [Winter 92], p47.<br />

facts. First, both control strategies rely on the exploitation <strong>of</strong> passive dynamics. As the<br />

masses and lengths <strong>of</strong> the robot’s body segments have been modeled based on human<br />

data, one expects that similar motions emerge. Furthermore, the active part <strong>of</strong> the control<br />

system is designed following insights from neuroscience and biomechanics. Several <strong>of</strong> the<br />

relevant control concepts have been transferred, and indeed <strong>like</strong>wise locomotion patterns<br />

originate. Finally, many parameters <strong>of</strong> the control units might be unconsciously tuned<br />

to achieve a human-<strong>like</strong> gait <strong>of</strong> the simulated biped because <strong>of</strong> repeated consultation <strong>of</strong><br />

biomechanical data and observation <strong>of</strong> the supremacy <strong>of</strong> human walking during the design<br />

and development <strong>of</strong> the suggested control system.<br />

8.1.5 Kinetic Analysis<br />

Kinetic analysis considers the acting <strong>of</strong> internal as well as external forces and torques.<br />

This gives a deeper insight into the inner workings <strong>of</strong> the control system and allows to<br />

draw conclusions regarding the feasibility <strong>of</strong> the approach on a real robot. Furthermore,<br />

preliminary statements on the efficiency <strong>of</strong> the emerging gait are possible.<br />

8.1.5.1 Ground Reaction Forces<br />

During undisturbed walking, external forces are only induced by ground contact. The<br />

robot is equipped with several sensors to measure these forces: A three axis load cell<br />

connects the foot with the ankle joint. In addition, four one-dimensional force sensors<br />

are mounted in the corners <strong>of</strong> each feet, and referred to as inner toe, outer toe, inner<br />

heel, and outer heel sensor. Figure 8.15 visualizes the measurements <strong>of</strong> these sensors.<br />

Again, the data is normalized to one gait cycle and averaged over 30 successive steps <strong>of</strong><br />

walking on level ground. The solid lines represent the mean values, the dashed lines show<br />

the minimum and maximum values. The vertical dashed line marks the stance-to-swing<br />

transition. To ease orientation, the angle trajectory <strong>of</strong> the ankle joint is included in the<br />

upper right corner.<br />

The force in z-direction perceived by the load cell summarizes the load the foot has to bear.<br />

Adding up the forces measured by the individual foot sensors results in the same trajectory.<br />

During initial weight acceptance, the foot load surpasses the amount <strong>of</strong> gravitational force<br />

generated by the body weight due to the impact. However, this overshoot is only marginal<br />

(b)


8.1. Normal <strong>Walking</strong> 159<br />

LC Force Z [N]<br />

LC Torque X [Nm]<br />

Inner Toe [N]<br />

Outer Toe [N]<br />

900<br />

600<br />

300<br />

0<br />

20<br />

0<br />

0 25 50 75 100<br />

−20<br />

0 25 50 75 100<br />

400<br />

200<br />

0<br />

400<br />

200<br />

0<br />

0 25 50 75 100<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

Ankle Y [ ◦ ]<br />

LC Torque Y [Nm]<br />

Inner Heel [N]<br />

Outer Heel [N]<br />

20<br />

0<br />

−20<br />

60<br />

30<br />

0<br />

0 25 50 75 100<br />

−30<br />

0 25 50 75 100<br />

400<br />

200<br />

0<br />

400<br />

200<br />

0<br />

0 25 50 75 100<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

Figure 8.15: Forces and torques measured by the foot sensors over one gait cycle. The solid<br />

lines represent the mean values over 30 steps <strong>of</strong> normal walking on even ground. The dashed<br />

lines mark the minimum and maximum values. The vertical dashed line indicates the transition<br />

from stance to swing.<br />

thanks to the preventive leg bending before heel strike by the Weight Acceptance motor<br />

pattern. Also, some <strong>of</strong> the impact is absorbed by damping in the ankle joint as the foot<br />

rotates from heel strike to full contact during the first ten percent <strong>of</strong> the gait cycle. The<br />

short decline in foot load right after full contact can be explained by the activity <strong>of</strong> the<br />

leg propulsion motor pattern that pushes the robot forward via the contralateral ankle<br />

joint. This action shifts some <strong>of</strong> the body weight back to the other leg. As the swing <strong>of</strong><br />

the contralateral leg sets in after about 20 percent <strong>of</strong> the cycle, the body weight is fully<br />

supported by the stance leg. A short peak in vertical force can be observed during leg<br />

propulsion, then the force drops to zero at lift-<strong>of</strong>f. During the swing phase, only small<br />

deviations from zero appear due to the weight <strong>of</strong> the foot pulling at the load cell.<br />

The torques measured by the load cell give a first indication on the progress <strong>of</strong> the center<br />

<strong>of</strong> pressure (cop). During the stance phase, the torque about the x-axis oscillates about


160 8. Dynamic <strong>Walking</strong> Experiments<br />

Medial Lateral<br />

Center <strong>of</strong> Pressure<br />

outer heel outer toe<br />

inner heel inner toe<br />

Posterior Anterior<br />

Ant.<br />

CoP X<br />

Post.<br />

Lat.<br />

CoP Y<br />

Med.<br />

0 25 50 75 100<br />

0 25 50 75 100<br />

% <strong>of</strong> Gait Cycle<br />

Figure 8.16: Center <strong>of</strong> pressure trajectories calculated from the four force sensors in each foot.<br />

The values present an average over 30 steps <strong>of</strong> normal walking on even terrain.<br />

zero due to the action <strong>of</strong> the Lateral Balance Ankle reflex. The dashed minimum and<br />

maximum values show considerable variation in the trajectory, underlining the reactive<br />

character <strong>of</strong> the responsible motor pattern. The torque value remains within ±10Nm,<br />

considering the short leverage <strong>of</strong> the foot in the lateral direction. Higher torques would<br />

cause the foot to turn onto its edge.<br />

The torque about the y-axis illustrates that the cop travels from heel to toe. At heel-strike,<br />

the foot pushes the robot forward by the torque <strong>of</strong> the passive ankle spring and the activity<br />

<strong>of</strong> the Forward Velocity reflex. Due to the action <strong>of</strong> the leg propulsion motor pattern, a<br />

strong torque peak <strong>of</strong> more than 60Nm can be observed. Both torque values disappear<br />

during the swing phase since no more external forces act on the foot.<br />

The measurements <strong>of</strong> the individual toe and heel sensors provide further details on the<br />

ground reaction forces. At touchdown, only the heel sensors detect the impact, with an<br />

emphasis on the outer side <strong>of</strong> the foot. The ankle joint then rotates to full foot contact<br />

and the toe sensors also measure a force, with the peak value showing strong variations<br />

due to deviations in the robot’s forward velocity. Already after 50 percent <strong>of</strong> the gait<br />

cycle, i.e. well before the swing phase, the heel loses contact with the ground and leg<br />

propulsion pushes the toes downwards. Here, a maximum force <strong>of</strong> over 400N can occur.<br />

Not surprisingly, the foot sensors do not measure relevant forces during the swing phase.<br />

The four force sensors mounted in the foot corners also allow to estimate the trajectory <strong>of</strong><br />

the cop. Computing the ratio between the inner and outer sensors, or the toe and heel<br />

sensors, respectively, yields<br />

CoPX =<br />

CoPY =<br />

Finner toe +Fouter toe<br />

Finner toe +Fouter toe +Finner heel +Fouter heel<br />

Fouter toe +Fouter heel<br />

Finner toe +Fouter toe +Finner heel +Fouter heel<br />

(8.1)<br />

(8.2)


8.1. Normal <strong>Walking</strong> 161<br />

Figure 8.16 shows the progress <strong>of</strong> the cop values over the progress <strong>of</strong> one gait cycle on<br />

the right side, and the trajectory <strong>of</strong> the cop in the horizontal plane on the left side. Both<br />

visualizations illustrate the characteristics <strong>of</strong> the cop trajectory. At heel strike, the cop<br />

is located at the posterior end <strong>of</strong> the foot and quickly moves towards the center. During<br />

the middle part <strong>of</strong> the stance phase, the cop stays around the center <strong>of</strong> the foot while the<br />

postural reflexes generate correcting ankle torques to stabilize the robot. Towards the end<br />

<strong>of</strong> the stance phase, the cop travels to the anterior end and to the medial side <strong>of</strong> the foot<br />

as the leg propulsion accelerates the robot forward by plantarflexion <strong>of</strong> the foot. Since<br />

the upper body moves in the direction <strong>of</strong> the next stance leg, the main propulsion force is<br />

raised by the inner toe segment.<br />

Disregarding the influence <strong>of</strong> the postural reflexes, the cop trajectory travels smoothly<br />

from heel to toes during the stance phase. In y-direction, it remains well within the center<br />

<strong>of</strong> the foot with a small bias to the lateral side since the robot leans towards the side <strong>of</strong><br />

the stance leg. The distance to the stability margin stays large throughout the walking<br />

cycle, indicating that there is still ample reserve concerning the machine’s stability and<br />

there is no immediate risk for the robot to tumble.<br />

8.1.5.2 Joint Torques<br />

Analysis <strong>of</strong> the torques occurring in the joints during walking provides additional information<br />

on the working <strong>of</strong> the control system and allows to assess its application on a real<br />

robot. Figure 8.17 displays the effective torques as solid lines. The torque value <strong>of</strong> the<br />

simulated torque controller are given as dotted lines. These values do not include the<br />

effects <strong>of</strong> the parallel spring and the joint damping. All plots are again based on the same<br />

run as the previous figures and represent the mean values <strong>of</strong> 30 consecutive steps. The<br />

dashed lines illustrate the minimum and maximum values, the vertical dashed line marks<br />

the transition from stance to swing.<br />

The torque acting on the y-axis <strong>of</strong> the spine joint is mainly generated by the position control<br />

working at low compliance and constant target angle. High torque values <strong>of</strong> −100Nm can<br />

be observed after each heel strike. They compensate the mass inertia <strong>of</strong> the trunk swaying<br />

forward. The further progress <strong>of</strong> the torque trajectory reflects the continuous stabilization<br />

<strong>of</strong> the upper body. Similar behavior can be seen in the other axis <strong>of</strong> the spine joint. The<br />

z-axis in particular shows lower joint stiffness in the smooth motor torque values.<br />

The shoulder joint exhibits the motor torque impulses generated by the Arm Swing motor<br />

pattern. Depending on the current joint angle and velocity, damping and a weak parallel<br />

elasticity partly work against the requested arm swing. The ad- and abduction <strong>of</strong> the arm<br />

are restricted by the torque produced by the compliant position controller. The elbow<br />

joint is set to zero stiffness and the motor only shows activity when the joint reaches its<br />

limits in the fully extended position.<br />

The y-axis <strong>of</strong> the hip joint exhibits the highest torque peak <strong>of</strong> all joints. As the corresponding<br />

spine joint, it has to decelerate the upper body after heel strike, but due to the longer<br />

leverage and the additional weight <strong>of</strong> the pelvis segment, the torque values are even higher<br />

and reach nearly 200Nm. A second peak occurs during the acceleration <strong>of</strong> the hip joint<br />

for initiating the leg swing. At higher walking velocities, a short swing phase is necessary<br />

to move the leg to its anterior position in time for the next heel strike. As already realized<br />

by J. Pratt, the maximum velocity and torque <strong>of</strong> the hip joint limits the robot’s walking


162 8. Dynamic <strong>Walking</strong> Experiments<br />

Spine Y [Nm]<br />

Shoulder Y [Nm]<br />

Elbow Y [Nm]<br />

Hip Y [Nm]<br />

Knee Y [Nm]<br />

Ankle Y [Nm]<br />

100<br />

0<br />

−100<br />

10<br />

0<br />

−10<br />

5<br />

0<br />

0 25 50 75 100<br />

0 25 50 75 100<br />

−5<br />

0 25 50 75 100<br />

100<br />

0<br />

−100<br />

−200<br />

0 25 50 75 100<br />

50<br />

0<br />

−50<br />

−100<br />

0 25 50 75 100<br />

30<br />

0<br />

−30<br />

−60<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

Spine X [Nm]<br />

Spine Z [Nm]<br />

Shoulder X [Nm]<br />

Hip X [Nm]<br />

Hip Z [Nm]<br />

Ankle X [Nm]<br />

50<br />

0<br />

−50<br />

0 25 50 75 100<br />

20<br />

0<br />

−20<br />

10<br />

0<br />

−10<br />

50<br />

0<br />

−50<br />

−100<br />

20<br />

0<br />

−20<br />

20<br />

0<br />

−20<br />

0 25 50 75 100<br />

0 25 50 75 100<br />

0 25 50 75 100<br />

0 25 50 75 100<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

Figure 8.17: Average joint torques acting during one gait cycle. The solid line represents the<br />

mean values over 30 steps <strong>of</strong> normal walking on level ground. The dashed lines mark the minimum<br />

and maximum values. Torques generated by the motor without the influence <strong>of</strong> damping and<br />

parallel springs are shown by the dotted line.


8.1. Normal <strong>Walking</strong> 163<br />

velocity [Pratt 00]. At the end <strong>of</strong> the swing phase, the torque is reduced to the amount<br />

necessary to lock the hip angle and to control the legs angle <strong>of</strong> attack.<br />

During the stance phase, the hip joint also shows high torque values about its x-axis.<br />

This torque is used to stabilize the pelvis against the weight <strong>of</strong> the upper body while the<br />

contralateral leg is swinging. As soon as the double support phase starts after 50 percent<br />

<strong>of</strong> the gait cycle, the load is distributed and the torque is reduced. During the legs swing<br />

phase, some torque is necessary to adjust the lateral foot placement. The torque about<br />

the hip joint’s z-axis is produced due to the compliant joint action and is proportional to<br />

the current joint angle.<br />

The main strain on the knee joint also occurs during the stance phase. The minor flexion<br />

<strong>of</strong> the knee implicates that the joint has to support the body weight while it is not fully<br />

stretched. With the onset <strong>of</strong> the swing phase, the knee joint is set to passive action at<br />

maximum compliance, and the motor does not have to produce any torque.<br />

At heel strike, the ankle joint slows down the plantarflexion <strong>of</strong> the foot by a passive torque<br />

resulting from damping and the parallel spring. The joint’s motor remains mainly inactive<br />

except for the corrections <strong>of</strong> the Forward Velocity reflex until the leg propulsion motor<br />

pattern sets in. Then considerable torque is produced to accelerate the robot forward.<br />

During the swing phase, the foot is kept in minor dorsiflexion and the motor has to<br />

compensate for the torque generated by the parallel spring due to the deflection from its<br />

equilibrium position. The torques produced in the x-axis <strong>of</strong> the ankle joint result from the<br />

action <strong>of</strong> the Lateral Balance Ankle postural reflex.<br />

While the peak torque values, in particular those <strong>of</strong> the hip joint, appear to be relatively<br />

large, they are still far below the capabilities <strong>of</strong> humans or state-<strong>of</strong>-the-art bipedal robots.<br />

For instance, the robot LOLA <strong>of</strong> the University <strong>of</strong> Munich can generate a peak torque <strong>of</strong><br />

370Nm in the y-axis <strong>of</strong> its hip joint, at a total robot weight <strong>of</strong> 55kg and a similar number<br />

<strong>of</strong> degrees <strong>of</strong> freedom. Thus it should be possible to develop joint actuators meeting at least<br />

the torque demands. There still remains the problem <strong>of</strong> achieving adjustable compliance<br />

and sufficiently low joint friction.<br />

8.1.5.3 Joint Power<br />

In knowing the joint angles and torques, it is possible to calculate the mechanical work<br />

done by the joints as well as their mechanical power. Rotational work W is defined as<br />

W = τθ, (8.3)<br />

with τ being the torque and θ the joint revolution this torque is applied through. Power<br />

is the rate at which this work is performed. Using the data set <strong>of</strong> the run presented<br />

throughout this chapter, the mechanical power Pj for each joint j is calculated as<br />

Pj(t) =<br />

1<br />

2 [τj(t+∆T)−τj(t−∆T)]·θj(t+∆T)−θj(t−∆T)<br />

, (8.4)<br />

2·∆T<br />

with ∆T = 25msec. The resulting power trajectories are again averaged over 30 steps <strong>of</strong><br />

normal walking, and the mean, minimum, and maximum values are shown in Figure 8.18.<br />

The solid lines represent the power based on the effective torque values, while the dotted<br />

lines depict the mechanical motor power.


164 8. Dynamic <strong>Walking</strong> Experiments<br />

Spine Y [W]<br />

Shoulder Y [W]<br />

Elbow Y [W]<br />

Hip Y [W]<br />

Knee Y [W]<br />

Ankle Y [W]<br />

50<br />

0<br />

−50<br />

0 25 50 75 100<br />

20<br />

10<br />

0<br />

−10<br />

0 25 50 75 100<br />

0<br />

−1<br />

−2<br />

−3<br />

800<br />

400<br />

0<br />

100<br />

0<br />

−100<br />

0 25 50 75 100<br />

0 25 50 75 100<br />

−200<br />

0 25 50 75 100<br />

500<br />

250<br />

0<br />

−250<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

Spine X [W]<br />

Spine Z [W]<br />

Shoulder X [W]<br />

Hip X [W]<br />

Hip Z [W]<br />

Ankle X [W]<br />

0<br />

−5<br />

−10<br />

5<br />

0<br />

−5<br />

0 25 50 75 100<br />

−10<br />

0 25 50 75 100<br />

1<br />

0<br />

−1<br />

−2<br />

0 25 50 75 100<br />

60<br />

30<br />

0<br />

5<br />

0<br />

−5<br />

−10<br />

5<br />

0<br />

−5<br />

0 25 50 75 100<br />

0 25 50 75 100<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

Figure 8.18: Average approximated joint power arising during one gait cycle. The solid line<br />

represents the mean values over 30 steps <strong>of</strong> normal walking on level ground. The dashed lines<br />

mark the minimum and maximum values. Power generated by the motor without the influence<br />

<strong>of</strong> damping and parallel springs is shown by the dotted line.


8.1. Normal <strong>Walking</strong> 165<br />

Peak Peak Motor Maximum Max. Motor<br />

�<br />

Motor<br />

Joint Torque [Nm] Torque [Nm] Power [W] Power [W] Power [W]<br />

Spine X 29.0 13.4 0.5 1.5 0.3<br />

Spine Y 111.1 103.8 20.3 21.4 3.2<br />

Spine Z 16.3 9.5 1.6 1.9 0.5<br />

Shoulder X 7.4 6.9 0.2 0.3


166 8. Dynamic <strong>Walking</strong> Experiments<br />

�<br />

ΣPj =<br />

max(0,Pj(t))dt (8.5)<br />

Totalizing these values gives a mechanical power consumption <strong>of</strong> the biped during walking<br />

<strong>of</strong> 318.8W, with the two hip joints alone contributing nearly two thirds <strong>of</strong> this sum. As the<br />

control units were designed without any concern about energy consumption, it is highly<br />

probable that this number could be reduced after appropriate optimization.<br />

In an article on passive walkers, Collins et al. propose to use the dimensionless mechanical<br />

cost <strong>of</strong> transport cmt = (energy used)/(weight×distance traveled) to compare the walking<br />

efficiency <strong>of</strong> different robots and humans [Collins 05]. They give an estimate <strong>of</strong> cmt = 0.05<br />

for human walking and <strong>of</strong> cmt = 1.6 for the robot Asimo by Honda. With a weight <strong>of</strong><br />

76kg and an approximate walking velocity <strong>of</strong> 1.2 m/s during the analyzed run, the cost <strong>of</strong><br />

transport for the presented biped amounts to<br />

cmt = 318.8W<br />

760N·1.2 m/s<br />

= 0.35 (8.6)<br />

While this value is still well above the one <strong>of</strong> human walking, it attributes the presented<br />

bipedal robot to be nearly five times more efficient than Asimo using a technical control<br />

approachbasedonjointanglecontrol. Andasnotedbefore,thesystemstillhasthepotential<br />

to be optimized regarding its energy consumption. Of course, the simulation environment<br />

does not model all aspects <strong>of</strong> a real robot. Therefore, this result can only indicate the<br />

walking efficiency <strong>of</strong> the suggested methodology by exploiting inherent dynamics and<br />

making use <strong>of</strong> elastic energy storage.<br />

8.1.5.4 Active and Passive Joint Action<br />

To conclude the kinetic analysis, Table 8.4 lists which joints are actively controlled and<br />

which joints are passive during each walking phase. Only the spine joints and the joints <strong>of</strong><br />

the left body half are given. The number states the stiffness factor defining the compliance<br />

<strong>of</strong> each joint for the constant equilibrium point. A stiffness factor <strong>of</strong> 1.0 results in minimum<br />

compliance, a factor <strong>of</strong> 0.0 in a freely rotating joint with only the damping acting. The<br />

letters ‘M’ and ‘R’ indicate that a motor pattern or a reflex is influencing the joint during<br />

that phase. Reflexes put in parentheses only act due to disturbances and remain mostly<br />

inactive during normal walking.<br />

Including both the left and the right sided joints in the calculation results in the following<br />

totals: during stimulation <strong>of</strong> phases one and three, 8 joint are actively controlled and 13<br />

joints act passively. In phase combination one and four, 10 joints are active and 11 joints<br />

are passive. And during phases two and five, again 8 joints are active and 13 joints are<br />

passive. In total, no more than half <strong>of</strong> the joints are simultaneously actively controlled<br />

throughout the walking cycle, reducing the overall complexity <strong>of</strong> the control system. In<br />

fact, only 1–4 <strong>of</strong> the 21 degrees <strong>of</strong> freedom must be controlled by motor patterns for<br />

directing the passive system dynamics towards a walking gait.<br />

Furthermore, some passive joints remain at the same compliance setting during all phases <strong>of</strong><br />

walking. These joints could be replaced by fixed springs and would not need an additional<br />

actuator. However, a closer look reveals that this situation would change when starting to<br />

walk on intense slopes, or when including curved walking. Then, the spine joints would be


8.1. Normal <strong>Walking</strong> 167<br />

Joint Phase 1 Phase 2 Phase 3 Phase 4 Phase 5<br />

Spine X 0.9 0.9 0.9 0.9 0.9<br />

Spine Y 1.0 1.0 1.0 1.0 1.0<br />

Spine Z 0.2 0.2 0.2 0.2 0.2<br />

Shoulder X 0.3 0.3 0.3 0.3 0.3<br />

Shoulder Y 0.0 0.0 0.0 M 0.0 M 0.0<br />

Elbow Y 0.0 0.0 0.0 0.0 0.0<br />

Hip X 0.1 R 0.1 R 0.1 0.1 R 0.1 R<br />

Hip Y 0.0 R 0.0 R 0.0 M,R 0.0 M,R 0.0 R<br />

Hip Z 0.5 0.5 0.5 0.5 0.5<br />

Knee Y 0.0 M 0.3 0.0 0.0 (R) 0.0 M<br />

Ankle X 0.4 R 0.0 R 0.1 0.1 0.1<br />

Ankle Y 0.0 R 0.0 R 0.0 M 0.1 (R) 0.1 (R)<br />

Active 5 4 3 5 4<br />

Passive 7 8 9 7 8<br />

Table 8.4: Passive and active joints during each walking phase for the left sided joints. The<br />

numbers give the joint stiffness, M and R mark the active control by a motor pattern or reflex.<br />

necessary to compensate the inclination, and the hip’s z-axis would be used to control the<br />

walking direction. Still, the passive joint action suffices during walking on minor slopes or<br />

small steps as will be shown below.<br />

8.1.5.5 Comparison to human data<br />

As done for the kinematic data, some results from the kinetic analysis are compared to<br />

human walking. Figure 8.19 shows the effective torque <strong>of</strong> four joints for both human<br />

walking and the presented biped. <strong>Human</strong> data is again taken from the data set provided by<br />

Vicon and is plotted as dashed line. The torque values are given in Nm/kg, normalizing the<br />

torque by the body weight as commonly done in biomechanics to make results comparable.<br />

The torque trajectories <strong>of</strong> the hip joint show strong similarities. Besides the additional<br />

oscillation occurring in the robot data, the shape <strong>of</strong> the curves match. The torque about<br />

the hip’s x-axis even resembles in amplitude. In contrast, the robot’s torque values <strong>of</strong> the<br />

y-axis are up to three times higher than in human subjects. During the stance phase, this<br />

can be explained by stronger sway <strong>of</strong> the upper body that must be compensated. Here,<br />

suitable anticipatory motor patterns could reduce the necessary torque.<br />

The knee joint shows a different characteristic during weight acceptance in the first part<br />

<strong>of</strong> the stance phase. <strong>Human</strong> subjects actively flex the knee before and during heel strike.<br />

Then the body weight is decelerated by a torque working against excessive bending <strong>of</strong> the<br />

knee. The robot also actively bends the knee. But then the plantarflexion <strong>of</strong> the foot on<br />

the contralateral side accepts some <strong>of</strong> the body weight, and the flexion <strong>of</strong> the knee must<br />

actually be supported. During leg swing, in both human and robot, minor torques appear<br />

due to elasticities and damping.<br />

The ankle torque trajectories again show notable resemblance. The main difference is the<br />

shift <strong>of</strong> the onset <strong>of</strong> the push-<strong>of</strong>f motor pattern that occurs earlier in human subjects and<br />

is <strong>of</strong> higher amplitude.


168 8. Dynamic <strong>Walking</strong> Experiments<br />

Ext 2<br />

Hip X<br />

[ Nm/kg]<br />

1<br />

0<br />

Flex −1<br />

Ext 2<br />

Knee Y<br />

[ Nm/kg]<br />

1<br />

0<br />

Flex −1<br />

0 25 50 75 100<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

Ext 2<br />

Hip Y<br />

[ Nm/kg]<br />

1<br />

0<br />

Flex −1<br />

Ext 2<br />

Ankle Y<br />

[ Nm/kg]<br />

1<br />

0<br />

Flex −1<br />

0 25 50 75 100<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

Figure 8.19: Comparing joint moments in human walking (dashed lines) to those <strong>of</strong> the<br />

simulated biped (solid lines). Torque values are normalized by body weight. <strong>Human</strong> data is<br />

again taken from the data set provided by Vicon.<br />

Figure 8.20 presents a similar comparison <strong>of</strong> joint power trajectories. Unfortunately, the<br />

Vicon data set does not include any information on the power <strong>of</strong> hip ab- and adduction,<br />

so only three joint axes are compared. As expected after the analysis <strong>of</strong> the torque output,<br />

the hip joint’s power consumption is considerable higher in the presented robot than in<br />

human subjects. In particular during the initiation <strong>of</strong> the leg swing, a strong power peak<br />

can be observed. As already argued above, this might originate from the relatively strong<br />

damping.<br />

The knee joint again produces a very similar power progression. Both human subjects and<br />

the robot exhibit two small peaks <strong>of</strong> power absorption during the leg swing. Those <strong>of</strong> the<br />

robot are located closer to one another as the knee flexion is <strong>of</strong> shorter duration. The<br />

power trajectories <strong>of</strong> the ankle joint also show strong resemblance. However, the robot<br />

absorbs more energy during the ankle rotation at heel strike. During leg propulsion, the<br />

robot produces slightly more power output.<br />

In summary, the kinetic comparison renders a similar picture as the kinematic comparison.<br />

The torque as well as the power trajectories during human and robot walking mostly<br />

resemble each other. The main differences can be found in the flexion and extension <strong>of</strong> the<br />

hip joint. Here, the robot produces considerably more torque and power output. Thus,<br />

any optimization towards reducing the energy consumption should begin by considering<br />

the control units affecting this joint.<br />

8.2 <strong>Walking</strong> under Disturbances<br />

The first part <strong>of</strong> this chapter illustrated that the proposed control approach is well able<br />

to achieve efficient, dynamic walking <strong>of</strong> a bipedal robot. Already in the normal walking


8.2. <strong>Walking</strong> under Disturbances 169<br />

Prod 4<br />

Knee Y<br />

[ W/kg]<br />

2<br />

0<br />

Abs −2<br />

Prod 4<br />

Hip Y<br />

[ W/kg]<br />

Abs −2<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

2<br />

0<br />

0 25 50 75 100<br />

Prod 4<br />

Ankle Y<br />

[ W/kg]<br />

2<br />

0<br />

Abs −2<br />

0 25 50<br />

% <strong>of</strong> Gait Cycle<br />

75 100<br />

Figure 8.20: Comparing joint power in human walking (dashed lines) to those <strong>of</strong> the simulated<br />

biped (solid lines). Power values are normalized by body weight. <strong>Human</strong> data is taken from the<br />

Vicon data set.<br />

described there, local and postural reflexes help to stabilize the robot against self-induced<br />

disturbances and against the impacts occurring at each foot contact. It is one conjecture <strong>of</strong><br />

this thesis that the distributed and purely reactive action <strong>of</strong> these independent control units<br />

enables walking under the influence <strong>of</strong> additional external and unforeseen disturbances.<br />

This section will evaluate the performance <strong>of</strong> the reflex network during perturbed walking<br />

by examining three different kinds <strong>of</strong> disturbances as shown in Figure 8.21: walking on<br />

sloped terrain, walking over steps, and walking while external forces are applied to the<br />

trunk <strong>of</strong> the robot. It should be noted again that all experiments are performed with<br />

the same set <strong>of</strong> model and control unit parameters. Only a different walking velocity<br />

modulation factor is used in some runs to vary the amount <strong>of</strong> energy put into the system<br />

by the motor patterns.<br />

8.2.1 Sloped Terrain<br />

In a first experiment, the robot is placed on sloped terrain <strong>of</strong> different angles. The robot<br />

is dropped from a small height on the ground and the standing control is balancing the<br />

robot. Then walking is initiated and the biped has to start and to continue to walk on the<br />

sloped surface. Downhill and uphill walking is examined as well as locomotion on terrain<br />

inclined to the side.<br />

Figure 8.22 compares the walking on level ground with walking on a surface inclined<br />

upwards by one degree and inclined downward by two and four degrees. The pelvis height<br />

is included to illustrate the progression <strong>of</strong> the slope and to show that the robot manages<br />

to maintain a smooth and steady gait. On the left, the figure plots both the effective as<br />

well as the pure motor torques <strong>of</strong> the ankle joint about the y-axis. The trajectories are


170 8. Dynamic <strong>Walking</strong> Experiments<br />

(a) (b) (c)<br />

Figure 8.21: Experiments on walking under disturbances: (a) walking on a downhill slope, (b)<br />

stepping up a platform, (c) walking with an external force acting at the torso.<br />

averaged over the steps <strong>of</strong> the respective run. The first eight seconds <strong>of</strong> the activity <strong>of</strong> the<br />

left (green) and right (red) Forward Velocity postural reflexes are depicted on the lower<br />

right. During the runs on level and downhill ground, the walking velocity factor is set to<br />

0.8, during the uphill run the factor is set to 0.95 as will be explained below. The resulting<br />

effective walking velocity amount to approximately 1.2 m/s, differing slightly depending on<br />

the ground inclination.<br />

Regarding the action <strong>of</strong> the postural reflex, it can be noted that only when walking downhill<br />

at 4 ◦ considerably more activity can be observed. In the other terrains, the required<br />

postural correction is indeed different in its characteristic but similar in strength. Only at<br />

steep inclinations, higher reflex intensity becomes necessary.<br />

The results <strong>of</strong> the reflex action can best be observed in the ankle torque during the first<br />

half <strong>of</strong> the gait cycle. After about 5% <strong>of</strong> the cycle, the foot has full ground contact and the<br />

Forward Velocity control units influence the ankle torque until the heel lifts <strong>of</strong>f at about<br />

45%. In Figure 8.22, this interval is marked by the light red box. During this phase, foot<br />

dorsiflexion, i.e. positive ankle torque values, push the heel to the ground and accelerate<br />

the robot. In contrast, plantarflexion represented by negative torque values, decelerate the<br />

robot. During uphill walking, dorsiflexion <strong>of</strong> the foot slightly predominates, but no distinct<br />

difference to walking on level ground can be observed. The situation is more obvious as<br />

the ground inclines forward. Already at two degrees, prominent foot plantarflexion can be<br />

noted. This effect intensifies with increasing slope. At four degrees, a peak motor torque<br />

<strong>of</strong> nearly −40Nm occurs.<br />

Another distinction in ankle torque can be noticed during the subsequent leg propulsion.<br />

As the body mass has traveled in front <strong>of</strong> the stance leg, the situation reverses and foot<br />

plantarflexion now results in accelerating the robot. From downhill to uphill walking, a<br />

considerable increase in leg propulsion from −40Nm to nearly −80Nm can be observed.<br />

This behavior stabilizes the robot since the gain or reduction in potential energy caused by<br />

the slope must be compensated. The reason for the change in ankle torque can be found<br />

in two facts: during uphill walking, the parallel spring mounted in the ankle is extended


8.2. <strong>Walking</strong> under Disturbances 171<br />

Uphill Slope 1 ◦<br />

Level Ground<br />

Downhill Slope 2 ◦<br />

Downhill Slope 4 ◦<br />

Ankle Torque Y [Nm]<br />

Ankle Torque Y [Nm]<br />

Ankle Torque Y [Nm]<br />

Ankle Torque Y [Nm]<br />

40<br />

0<br />

−40<br />

−80<br />

40<br />

0<br />

−40<br />

−80<br />

40<br />

0<br />

−40<br />

−80<br />

40<br />

0<br />

−40<br />

−80<br />

0 25 50 75 100<br />

% <strong>of</strong> Gait Cycle<br />

0 25 50 75 100<br />

% <strong>of</strong> Gait Cycle<br />

0 25 50 75 100<br />

% <strong>of</strong> Gait Cycle<br />

0 25 50 75 100<br />

% <strong>of</strong> Gait Cycle<br />

Pelvis Z [m]<br />

Forw. Vel. a<br />

Pelvis Z [m]<br />

Forw. Vel. a<br />

Pelvis Z [m]<br />

Forw. Vel. a<br />

Pelvis Z [m]<br />

Forw. Vel. a<br />

1.0<br />

0.5<br />

0.0<br />

0.0 2.0 4.0 6.0 8.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 2.0 4.0 6.0 8.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 2.0 4.0 6.0 8.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 2.0 4.0 6.0 8.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 2.0 4.0 6.0 8.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 2.0 4.0 6.0 8.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 2.0 4.0 6.0 8.0<br />

1.0<br />

0.5<br />

0.0<br />

0.0 2.0 4.0<br />

Time [s]<br />

6.0 8.0<br />

Figure 8.22: Comparison <strong>of</strong> the ankle torque, the pelvis height, and the activity <strong>of</strong> the Forward<br />

Velocity postural reflex at different ground slopes.


172 8. Dynamic <strong>Walking</strong> Experiments<br />

further towards the end <strong>of</strong> the stance phase and thus induces more torque during leg<br />

propulsion. Similarly, ankle torque is reduced while walking downhill. This effect occurs<br />

without any control and illustrates the self-stabilizing properties <strong>of</strong> elastic elements. The<br />

second reason for the change in ankle torque lies in the Leg Propulsion motor pattern.<br />

The amplitude <strong>of</strong> its torque impulse is influenced by two modulation factors. On the one<br />

hand, the ankle’s angle at the onset <strong>of</strong> the pattern increases or decreases the amplitude,<br />

resulting in a behavior similar to the one <strong>of</strong> the parallel spring. On the other hand, the<br />

control unit receives the same modulation factor as the Forward Velocity reflex based on<br />

the trajectory <strong>of</strong> the estimated com position. By this means, the leg propulsion effect is<br />

reinforced in case the estimated velocity is too low, and vice versa.<br />

By repeating the experiment on different slopes, the limits <strong>of</strong> the postural control system<br />

can be determined. At downhill inclinations <strong>of</strong> more than 5 ◦ , the robot does not manage<br />

to maintain stability for a long time. The step length increases too much, resulting in<br />

stronger oscillation <strong>of</strong> the trunk, but also in reduced lateral stability as changes in foot<br />

placement and ankle torque have less influence. Uphill walking shows to be more difficult.<br />

Already at minor slopes, two problems can be observed: the robot’s forward velocity<br />

decreases at each step as additional potential energy must be raised; and due to the small<br />

ground clearance during foot swing, the robot tends to stumble. To a certain degree, both<br />

difficulties can be solved by increasing the walking velocity modulation factor. That way,<br />

more energy is added to the system, and the higher torque in the hip joint during swing<br />

initiation results in a slightly larger ground clearance. Still, starting at uphill slopes <strong>of</strong><br />

about two 2 ◦ , the robot cannot maintain walking for more than a few steps.<br />

Stability during up- and downhill walking could be increased by adapting the parameters<br />

<strong>of</strong> some control units and by introducing additional postural reflexes. Tuning <strong>of</strong> the step<br />

length, the propulsion forces, or the correcting ankle torques would slightly improve the<br />

walking abilities. To achieve a more substantial enhancement <strong>of</strong> stability, a postural reflex<br />

adapting the orientation <strong>of</strong> the upper body seems more promising. As it can be observed<br />

in humans, leaning backwards during downhill and bending forward during uphill walking<br />

generates the desired effect. Such a control unit would either require an estimation <strong>of</strong> the<br />

ground inclination based on the inertial measurement unit, or could observe the activity<br />

and the target rating <strong>of</strong> the postural behaviors as indication on the need to change the<br />

upper body pose. As already validated in preliminary test, the ground clearance can be<br />

increased by introducing an additional motor pattern being stimulated during leg swing<br />

and creating auxiliary torque impulses in the knee and hip joint.<br />

<strong>Walking</strong> on a slope directed orthogonally to the walking direction serves as evaluation <strong>of</strong><br />

the lateral stabilization capabilities. The setup is the same as in the previously described<br />

experimental runs, only the inclination <strong>of</strong> the ground points to the side <strong>of</strong> the robot.<br />

Figure 8.23 illustrates the differences between walking on level terrain and walking on a<br />

surface inclined to the right by 1.5 ◦ .<br />

The most obvious distinction can be observed in the action <strong>of</strong> the Lateral Foot Placement<br />

postural reflex. The inclination pushes the robot to the right with every step it takes,<br />

resulting also in a rotation <strong>of</strong> the whole machine around the x-axis. The reflex reacts to this<br />

disturbances by placing the feet further to the right than in normal walking. As illustrated<br />

by the foot point diagrams, the left foot shown as green circle is placed practically in front<br />

<strong>of</strong> the last stance foot, whereas the right foot is positioned up to 60cm to the right. As a


8.2. <strong>Walking</strong> under Disturbances 173<br />

Slope Down to Right 1.5 ◦<br />

Level Ground<br />

Foot Points<br />

Position Y [m]<br />

Lat. Balance<br />

Ankle a<br />

Foot Points<br />

Position Y [m]<br />

Lat. Balance<br />

Ankle a<br />

−1<br />

−2<br />

0 walking dir.<br />

−3<br />

3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0 12.0<br />

Position X [m]<br />

1.0<br />

0.5<br />

0.0<br />

0.0 2.0 4.0 6.0<br />

Time [s]<br />

8.0 10.0 12.0<br />

1 walking dir.<br />

0<br />

−1<br />

3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0 12.0<br />

Position X [m]<br />

1.0<br />

0.5<br />

0.0<br />

0.0 2.0 4.0 6.0<br />

Time [s]<br />

8.0 10.0 12.0<br />

Figure 8.23: Comparison <strong>of</strong> lateral foot placement and the activity <strong>of</strong> the Lateral Balance<br />

Ankle postural reflex on level terrain and ground sloped to the right.<br />

result, the robot is walking sidewards down the slope, whereas on level ground, the minor<br />

foot placement corrections do not deflect the robot distinctly from its path.<br />

In addition, the Lateral Balance Ankle reflex also exhibits plainly increased activity. It<br />

works against the robot tilting in the direction <strong>of</strong> the inclination by raising the ankle<br />

torque about the x-axis. Since only a small leverage is available in the foot, the reflex<br />

already reaches its saturation at some points. Due to this reflex action, the cop is located<br />

further to the margin <strong>of</strong> the foot, thus increasing the risk <strong>of</strong> rolling over the edge <strong>of</strong> the<br />

foot while at the same time pushing the robot to a more stable, upright position.<br />

It shows that the robot cannot maintain stability during walking at sidewards slopes <strong>of</strong> 2 ◦<br />

or more. The main reason for this is that the robot is expecting level ground at heel strike<br />

but instead the foot must be lowered even further. Since the hip is kept rigid during this<br />

phase, the robot has to tilt to the side to reach the ground. Again, an additional control


174 8. Dynamic <strong>Walking</strong> Experiments<br />

Pelvis<br />

Position Z [m]<br />

Foot Points<br />

Position Y [m]<br />

Forward<br />

Velocity a<br />

1.2<br />

1.0<br />

0.8<br />

1.0<br />

0.0<br />

−1.0<br />

1.0<br />

0.5<br />

0.0<br />

walking dir.<br />

1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0<br />

1.0 2.0 3.0 4.0 5.0<br />

Position X [m]<br />

6.0 7.0 8.0 9.0<br />

2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 7.5<br />

Time [s]<br />

Figure 8.24: Pelvis height, foot points, and activity <strong>of</strong> the Forward Velocity postural reflex<br />

during walking up and down a platform <strong>of</strong> 3cm in height. The gray box marks the position <strong>of</strong><br />

the platform.<br />

unit could solve this problem. By rotation the hip towards the ground inclination, and<br />

at the same time bending the spine joint to lean the trunk in the opposite direction, the<br />

disturbance just described could be reduced.<br />

8.2.2 <strong>Walking</strong> over Steps<br />

Similar to the experiments on walking up- and downhill, traversing steps mainly poses<br />

demands on the postural reflexes controlling the forward velocity. But in contrast to slope<br />

walking, steps produce short peaks <strong>of</strong> disturbances and present a discontinuity to the<br />

control system. The main difficulties lie in sudden accelerations and stumbling over the<br />

steps.<br />

Figure 8.24 presents the results <strong>of</strong> walking over a platform <strong>of</strong> 3cm in height and 4m in<br />

length. The robot first has to step upon the platform, walk along it, and then step down<br />

again. The platform is visualized by the gray box. The time scale <strong>of</strong> the reflex activity in<br />

the lower plot is scaled to approximately fit to the translational progression shown in the<br />

upper two plots.<br />

The side view <strong>of</strong> the pelvis trajectory illustrates how the robot gains height during walking<br />

on the platform. Additionally, it reveals two variations from its normal trajectory, as


8.2. <strong>Walking</strong> under Disturbances 175<br />

marked by the arrows in the plot. Directly after stepping on the platform, the pelvis<br />

stays higher at its lowest point, whereas after stepping down, it descends a bit more. The<br />

lowest point <strong>of</strong> this trajectory is reached after heel strike before the upper body again is<br />

raised by the stance leg, thus indicating a change in step length. More precisely, the step<br />

length is reduced when stepping up and lengthened when stepping down. This behavior<br />

compensates the loss respectively the gain <strong>of</strong> potential energy and results from the change<br />

in walking velocity. The variance in step length can also be observed in the foot placement<br />

plot. Furthermore it shows that the robot slightly walks to the left after stepping down the<br />

platform. This arises from the lateral acceleration while stepping down with the left foot.<br />

The lowest plot <strong>of</strong> Figure 8.24 depicts the activity <strong>of</strong> the Forward Velocity postural reflexes.<br />

The green plot represents the left reflex and the red plot the right reflex. They correspond<br />

to the left and right foot points shown in the plot above. It can be observed that the reflex<br />

activity increases right after stepping on and stepping <strong>of</strong>f the platform. After these events,<br />

the activity drops back to normal level. The postural reflex tries to correct the deviations<br />

in walking velocity by foot dorsi- or plantarflexion in a similar way as it was described<br />

during walking on slopes.<br />

The maximum platform height the robot can step upon is mainly restricted by the feet’s<br />

ground clearance during the swing phase. As shown in Figure 8.9, for the major part<br />

<strong>of</strong> the swing phase, the distance <strong>of</strong> the heel or toe segment to the ground is below 5cm.<br />

Already for the experiment just described, the platform had to be placed at the appropriate<br />

distance to the robot so it would manage to step upon it. Again, an additional motor<br />

pattern adding torque impulses in the knee and the hip would increase ground clearance<br />

and would allow to walk over higher steps.<br />

<strong>Walking</strong> down the platform works up to a height <strong>of</strong> 5cm. Above this height, the robot<br />

gains too much velocity and does not manage to finish the subsequent step in time, i.e.<br />

the robot stumbles during the swing phase. Additional hip torque would be needed to<br />

accelerate the swing phase. Furthermore, higher steps cause additional lateral disturbances<br />

that the robot cannot always compensate by correcting the foot placement.<br />

8.2.3 Constant External Forces<br />

In a final series <strong>of</strong> experiments, a constant external force is applied to the robot’s trunk<br />

pushing it forward or pulling it backwards. Other than during walking on a slope, the<br />

robot does not risk to stumble on a rising inclination or to excessively lengthening its<br />

steps. But the external force constrains the trunk movements and thus disturbs the passive<br />

dynamics <strong>of</strong> the system.<br />

Figure 8.25 illustrates the ankle’s power output and the activity <strong>of</strong> the Forward Velocity<br />

postural reflex at no external force and during both a pushing and a pulling force <strong>of</strong> 10N.<br />

As expected, additional reflex activity can be observed in both cases <strong>of</strong> external forces.<br />

The ankle power trajectories show only small but distinctive variations. Right after heel<br />

strike, marked by the left arrow in the plots, negative work is done while the foot is rotated<br />

to full ground contact. The amount <strong>of</strong> work increases while pushing the robot, resulting<br />

in a dissipation <strong>of</strong> the additional energy. In case the robot is restricted by the force, the<br />

negative work is reduced.<br />

The opposite situation can be seen during leg propulsion as marked by the right arrow in<br />

the ankle power plots. Here additional energy is put into the system when the robot is


176 8. Dynamic <strong>Walking</strong> Experiments<br />

Push Torso 10N<br />

Undisturbed<br />

Pull Torso 10N<br />

Ankle Power Y [W]<br />

Ankle Power Y [W]<br />

Ankle Power Y [W]<br />

400<br />

200<br />

0<br />

−200<br />

0 25 50 75 100<br />

% <strong>of</strong> Gait Cycle<br />

400<br />

200<br />

0<br />

−200<br />

0 25 50 75 100<br />

% <strong>of</strong> Gait Cycle<br />

400<br />

200<br />

0<br />

−200<br />

0 25 50 75 100<br />

% <strong>of</strong> Gait Cycle<br />

Forw. Vel. a<br />

Forw. Vel. a<br />

Forw. Vel. a<br />

1.0<br />

0.5<br />

0.0<br />

2.0 4.0 6.0<br />

Time [s]<br />

8.0 10.0<br />

1.0<br />

0.5<br />

0.0<br />

2.0 4.0 6.0<br />

Time [s]<br />

8.0 10.0<br />

1.0<br />

0.5<br />

0.0<br />

2.0 4.0 6.0<br />

Time [s]<br />

8.0 10.0<br />

Figure 8.25: Comparison <strong>of</strong> the ankle power and the activity <strong>of</strong> the Forward Velocity postural<br />

reflex at different constant external forces.<br />

pulled backwards, and less work is performed in case the robot is pushed forwards. Since<br />

the power output <strong>of</strong> the motor remains more or less the same during the indicated phases,<br />

the variations in effective power must result from the parallel spring and from damping.<br />

Again, passive action by an elastic element shows a self-stabilizing effect.<br />

Similar results are achieved by forces acting from different directions. For instance, a torso<br />

force directed to the front and the left generates activity <strong>of</strong> the reflexes acting both in<br />

the frontal and the sagittal plane, and the robot takes sidewards steps to the left. Forces<br />

acting in walking direction should not exceed 15N, lateral forces should remain below 8N,<br />

otherwise the robot will stumble.<br />

In summary, the postural control system works satisfactorily when confronted with different<br />

and unforeseen disturbances and manages to maintain stability at small to medium<br />

perturbations. The suggested combination <strong>of</strong> local reflexes, postural reflexes, and elastic<br />

elements can achieve robust dynamic walking. This is the case despite the fact that the<br />

control system has not yet being optimized for locomotion stability. Additional control<br />

units targeted at certain classes <strong>of</strong> disturbances could increase the robustness even further.


9. Conclusion and Outlook<br />

This thesis approached the task <strong>of</strong> designing a locomotion control methodology for twolegged<br />

robots. The presented concept is based on findings from the study <strong>of</strong> human motion<br />

control and can achieve dynamic, efficient, and robust walking <strong>of</strong> three-dimensional and<br />

fully articulated bipeds. The key characteristics <strong>of</strong> the suggested control system are:<br />

� Designed above the neuron level, the control is structured as a hierarchical network <strong>of</strong><br />

feedback and feed-forward control units, allowing gradual stimulation and inhibition.<br />

Neither a monolithic dynamic model nor pre-calculated trajectories are used.<br />

� Stimulation and synchronization <strong>of</strong> control units originates from spinal pattern<br />

generators; state transitions are triggered by kinetic and kinematic events.<br />

� Passive dynamics and self-stabilizing effects <strong>of</strong> elasticities are exploited by relying<br />

mainly on local torque commands instead <strong>of</strong> whole body joint angle control.<br />

� Postural control is based on reflexes <strong>of</strong> varying complexity. High-level postural<br />

reflexes apply simplified local dynamic models. Reflex action is phase-dependent.<br />

� The computational demands <strong>of</strong> the control system are comparatively low since<br />

cooperating local control units are used instead <strong>of</strong> a full-featured dynamic model.<br />

With the aid <strong>of</strong> a dynamic simulation environment, the control approach has been applied<br />

to an anthropomorphic biped model with 21 degrees <strong>of</strong> freedom and human-<strong>like</strong> morphology<br />

and actuation. This validation yielded the following main results:<br />

� The control system can achieve both dynamic walking and balanced standing; it can<br />

cope with the high complexity and the mechanical elasticities <strong>of</strong> the modeled biped.<br />

� The kinematic and kinetic analysis <strong>of</strong> the emerging, naturally looking walking gait<br />

reveals remarkable similarities to human walking.<br />

� Only a subset <strong>of</strong> the available degrees <strong>of</strong> freedom must be actively controlled in each<br />

phase <strong>of</strong> walking to direct the passive system dynamics towards a walking gait.


178 9. Conclusion and Outlook<br />

� The achieved velocity <strong>of</strong> up to 5 km/h with a robot <strong>of</strong> human-<strong>like</strong> size and weight<br />

surpasses the walking speed <strong>of</strong> most state-<strong>of</strong>-the-art bipeds. At the same time,<br />

locomotion is considerably more energy efficient than in joint angle controlled robots.<br />

� <strong>Walking</strong> and standing are robust against unknown and unexpected disturbances <strong>of</strong><br />

small to medium intensity <strong>like</strong> inclined ground, steps, or external forces.<br />

� The presented results and the analysis <strong>of</strong> biological motion control suggest that<br />

human-<strong>like</strong> morphology and actuation is a necessary prerequisite to achieve human<strong>like</strong><br />

walking skills.<br />

In conclusion, this thesis has shown that a control system based on concepts <strong>of</strong> human<br />

locomotion control applied to an anthropomorphic biped can in fact result in walking<br />

properties comparable to those observed in humans. The main drawback <strong>of</strong> the suggested<br />

approach remains the necessity to adjust a multitude <strong>of</strong> control unit parameters. Some<br />

possible solutions to this problem are discussed in the outlook succeeding the detailed<br />

summary presented in the next section.<br />

9.1 Summary<br />

The thesis started with an assessment <strong>of</strong> state-<strong>of</strong>-the-art biped control systems based on an<br />

extensive analysis <strong>of</strong> related work. It was argued that despite several decades <strong>of</strong> research,<br />

today’s robots are still far behind the locomotion capabilities <strong>of</strong> humans. The majority <strong>of</strong><br />

biped control systems rely on analytical approaches based on multibody dynamic modelling<br />

and pre-calculated joint angle trajectories. <strong>Walking</strong> stability is ensured using Zero-Moment<br />

Point constraints or similar considerations. While these methods are mathematically sound<br />

and allow for exact and purposeful motions, they show several drawbacks <strong>like</strong> strong model<br />

dependency, high energetic and computational costs, or low robustness to unforeseen<br />

disturbances. In contrast, the elegant motions observed in humans are highly robust, fast,<br />

energy efficient, and do not rely on a central model but rather emerge from a distributed<br />

control network and the inherent system dynamics in interaction with the environment.<br />

These facts gave rise to the main hypothesis <strong>of</strong> this thesis, namely that a control system<br />

based on concepts <strong>of</strong> human motion control can yield human-<strong>like</strong> locomotion capabilities<br />

in bipedal robots.<br />

Features <strong>of</strong> the <strong>Control</strong> Concept<br />

To verify the hypothesis mentioned above, literature on biomechanics and neuroscience<br />

was reviewed to find structural and functional features <strong>of</strong> human motion control that can<br />

be transferred to technical control systems. The following key aspects were identified:<br />

human walking relies on functional morphology and leaves some <strong>of</strong> the control burden<br />

to its“intelligent mechanics”. Inherent dynamics and self-stabilizing properties <strong>of</strong> elastic<br />

elements are exploited to further reduce the control effort and the energy consumption.<br />

The active control system is organized as a hierarchical network. Spinal control units issue<br />

coordinated feed-forward patterns <strong>of</strong> muscle activity. Feedback exists on various levels <strong>of</strong><br />

complexity, ranging from monosynaptic reflexes to postural control. Reflex action differs<br />

depending on the current phase <strong>of</strong> motion. emg analysis suggests that walking is divided


9.1. Summary 179<br />

in five phases being stimulated in bilateral synchronization. Postural control is supraspinal,<br />

i.e. it requires more than local sensor information.<br />

Based on these insights, the suggested control concept was designed. As passive dynamics<br />

should be exploited, no whole body joint trajectories are calculated. Rather, as in biological<br />

systems, the motions emerge from a hierarchical network <strong>of</strong> control units. Being based<br />

on ideas from behavioral robot control architectures, these units already have a semantic<br />

interpretation and thus are located above the level <strong>of</strong> individual neurons found in biological<br />

control systems. Six classes <strong>of</strong> control units were derived: locomotion modes are located at<br />

the highest level <strong>of</strong> control and represent a form <strong>of</strong> locomotion <strong>like</strong> walking or standing.<br />

They stimulate spinal pattern generators, state machine-<strong>like</strong> units that stimulate motion<br />

phases based on kinetic and kinematic events instead <strong>of</strong> oscillators with fixed timing.<br />

The motion phases then provide for synchronized stimulation <strong>of</strong> feed-forward control<br />

commands and activate the appropriate feedback units. Feed-forward control is issued by<br />

motor patterns in the form <strong>of</strong> local torque impulses directed at only one or a few adjacent<br />

joints. They shape the passive system dynamics to create the desired motion. Feedback is<br />

implemented by local and postural reflexes: local reflexes only affect spatially related joints<br />

based on data <strong>of</strong> adjacent sensors and introduce a tight sensor-actor coupling. Postural<br />

reflexes require whole body sensor information and can use simplified dynamic models<br />

to calculate their reaction. Arranged in a hierarchical network based on stimulation and<br />

inhibition, instances <strong>of</strong> these control unit classes generate the system behavior. Modulation<br />

signals <strong>like</strong> the desired walking velocity can influence the output <strong>of</strong> control units.<br />

Results from gait analysis and neuroscientific research helped to identify the necessary<br />

feed-forward and feedback control units. To simplify the design <strong>of</strong> motor patterns, a<br />

common sigmoid function has been suggested to represent all feed-forward joint commands.<br />

The spinal pattern generator (spg) for walking and the corresponding walking phases have<br />

been derived from findings based on emg analysis <strong>of</strong> human walking. The spg manages<br />

five motion phases, namely weight acceptance, leg propulsion, trunk stabilization, leg<br />

swing, and heel strike. On the basis <strong>of</strong> biomechanical gait analysis, motor patterns have<br />

been identified for each <strong>of</strong> these phases to achieve a walking gait. Reflexes similar to those<br />

working in human walking have been inserted.<br />

<strong>Control</strong> <strong>of</strong> stability during walking is based on several postural reflexes. They control<br />

the forward velocity, lateral stability, and the pose <strong>of</strong> the upper body by adjusting joint<br />

torque or changing foot placement. Most <strong>of</strong> these reflexes derive their reaction from an<br />

estimation <strong>of</strong> the robot’s center <strong>of</strong> mass based on a simplified dynamical model. The<br />

normal trajectories <strong>of</strong> the center <strong>of</strong> mass projection or an extrapolated version including<br />

its velocity are approximated. Deviation from these trajectories serve as indication for the<br />

reaction <strong>of</strong> the postural reflexes.<br />

The suggested control concept requires certain characteristics <strong>of</strong> the underlying mechatronic<br />

system. For instance, beneficial effects <strong>of</strong> the passive system dynamics cannot be exploited<br />

if the robot’s joints are self-locking or have overly high friction. Similarly, self-stabilizing<br />

properties <strong>of</strong> elastic elements can only be used if such elastic elements are present. This<br />

supports the statement <strong>of</strong> some biomechanics researchers claiming that biomimetic robots<br />

should be based on functional morphology. Consequentially, the presented control was<br />

applied to a suitable robotic system featuring an actuator with parallel springs, low friction,<br />

and adjustable series compliance.


180 9. Conclusion and Outlook<br />

Experimental Validation<br />

Since no real robot fullfilling the requirements mentioned above was available, a validation<br />

framework based on dynamical simulation has been developed. It features rigid body<br />

mechanics, customizable inertia tensors for each segment, a penalty-based contact model<br />

including coulomb-<strong>like</strong> friction calculation, and access to internal variables <strong>like</strong> forces and<br />

torques acting on bodies and joints. Thus, the framework is well suited to assess the<br />

feasibility <strong>of</strong> the suggested control concept and to analyze its performance. On top <strong>of</strong> the<br />

simulation’s constraint mechanism for defining joints, a model <strong>of</strong> the joint itself and its<br />

control system has been designed. It includes a parallel spring with constant rate and<br />

equilibrium point, damping, and a combined position and torque control allowing to adjust<br />

the series compliance. Within this framework, a bipedal robot with 6 degrees <strong>of</strong> freedom<br />

in each leg, a 3 d<strong>of</strong> spine joint, and 3 d<strong>of</strong> arms has been modeled. The mass distribution,<br />

the total weight, and the segment lengths were chosen to resemble typical human subjects.<br />

Flexible layout <strong>of</strong> the simulated environment allows to easily design different experimental<br />

setups.<br />

Experiments in the validation framework demonstrated that the presented control concept<br />

is indeed capable <strong>of</strong> achieving three-dimensional dynamic walking <strong>of</strong> an anthropomorphic<br />

bipedal robot, including the transition from balanced standing to walking. The naturally<br />

looking gait emerges from the combined control output <strong>of</strong> the phase-dependent stimulated<br />

reflexes and motor patterns, the inherent system dynamics, and the interaction <strong>of</strong> the<br />

robot with the environment. In fact, it shows that only about half <strong>of</strong> the 21 degrees <strong>of</strong><br />

freedom need to be actively controlled simultaneously during each phase <strong>of</strong> walking. The<br />

remaining joints act passively at different degrees <strong>of</strong> compliance. Even more, the control<br />

<strong>of</strong> only 1–4 d<strong>of</strong> by motor patterns is sufficient to direct the passive dynamics towards a<br />

walking gait.<br />

The presented approach is also capable to achieve stabilized standing. The control for<br />

balanced standing is mainly based on joint compliance to adapt to the ground, and on<br />

high joint stiffness to stabilize against external disturbances. Similar strategies can be<br />

observed in balancing human subjects. To evaluate this so called ankle strategy, the robot<br />

was confronted with different ground geometries and disturbances. The robot can stabilize<br />

its posture on ground inclination <strong>of</strong> up to 30 ◦ and withstand short force impulses <strong>of</strong> 1000N<br />

or movements <strong>of</strong> the supporting platform <strong>of</strong> up to 2.5 m/s. After stabilizing, the control<br />

system optimizes the robot’s posture by locally minimizing joint torques. This has been<br />

demonstrated by gradually inclining the surface the robot is standing on, or by changing<br />

the weight <strong>of</strong> a virtual backpack up to the weight <strong>of</strong> the robot itself.<br />

Analysis <strong>of</strong> the Resulting Gait<br />

An extensive behavioral, kinematic, and kinetic analysis <strong>of</strong> normal walking on level ground<br />

provided detailed information on the emerging gait and the interaction <strong>of</strong> the control<br />

units. It was illustrated that only a few kinetc and kinematic events are sufficient to<br />

trigger the state transitions <strong>of</strong> the walking spg in bilateral synchronization. It was shown<br />

that the behavior-based architecture iB2C with its stimulation and fusion mechanisms<br />

is an appropriate foundation <strong>of</strong> a biologically inspired control system. No elaborated<br />

stimulation strategies are necessary, thus simplifying the network design. The behavior<br />

signals allow to inspect the control unit activity throughout the walking phases to gain


9.1. Summary 181<br />

a solid understanding <strong>of</strong> the internal processes <strong>of</strong> the control network. Three different<br />

activity characteristics can be observed in the low-level control units: motor patterns<br />

follow a predefined activity progression; some reflexes always contribute an essential part<br />

to normal walking and show high activity at certain events <strong>of</strong> the gait cycle; finally, other<br />

reflexes working towards stabilization are more or less active depending on internal or<br />

external disturbances.<br />

Both kinematic and kinetic results show remarkable resemblance to human walking. Joint<br />

angle trajectories show the same characteristics and even similar amplitude. The same is<br />

true for the absolute position <strong>of</strong> body segments, joint torques, and power values. This can<br />

be observed despite the fact that no joint angle control based on human data has been<br />

applied. Reasons for this similarity can be found in the <strong>like</strong>wise exploitation <strong>of</strong> passive<br />

system dynamics at comparable segment masses and lengths, and in the fact that indeed a<br />

control system close to the human motion control has been found. These results might even<br />

serve as indication that some assumptions <strong>of</strong> neuroscientists on the function <strong>of</strong> locomotion<br />

control in humans that were used in this work are pointing in the right direction.<br />

<strong>Walking</strong> Velocity, Efficiency, and Robustness<br />

Inspection <strong>of</strong> the segment translation reveals the walking velocity <strong>of</strong> the robot. Varying<br />

the velocity modulation factor allows to change the walking velocity between 1.1 m/s and<br />

1.4 m/s. This shows that the mechanism <strong>of</strong> passing modulation signals to control units to<br />

scale their control output is working. Similarly, modulation <strong>of</strong> the walking direction or<br />

the step height could be implemented. The maximum velocity <strong>of</strong> about 5 km/h for a robot<br />

<strong>of</strong> human-<strong>like</strong> weight and height surpasses the walking speed <strong>of</strong> most <strong>of</strong> today’s bipeds.<br />

To compare the walking velocity <strong>of</strong> legged animals and robots, Alexander suggested to<br />

use the dimensionless Froude number defined as Fr = v2 /gl, where v is the velocity, g<br />

the gravitational acceleration, and l the leg length [Alexander 84]. For the presented<br />

biped, this number amounts to Fr = 0.21. At the time <strong>of</strong> writing, the fastest walking<br />

fully articulated biped is the robot Asimo, walking at a speed <strong>of</strong> 4 km/h, and running even<br />

faster [Takenaka 09a]. This results in a similar Froude number <strong>of</strong> Fr = 0.19, whereas<br />

KHR-4 comes to Fr = 0.02, Johnnie to 0.05, or HRP-2 to Fr = 0.07. <strong>Robots</strong> with a<br />

biologically motivated control or actuation system generally achieve high values, e.g. the<br />

robots Spring Flamingo and Rabbid both have a Froude number <strong>of</strong> Fr = 0.18, and the<br />

23cm small RunBot even reaches Fr = 0.28. The last three robots are supported by a<br />

rotating beam and only walk in the sagittal plane, so it still has to be shown whether<br />

their control approaches scale to three-dimensional walking. For comparison, fast human<br />

walking yields a Froude number <strong>of</strong> about Fr = 0.5.<br />

To assess the efficiency <strong>of</strong> the resulting walking gait, an estimation <strong>of</strong> the joint work<br />

and power output based on the torque and rotational displacement has been done. The<br />

overall mechanical power consumption during walking adds up to a bit more than 300W.<br />

Normalizingthisvaluetotherobot’sweightandthewalkingvelocitygivesanapproximation<br />

<strong>of</strong>thecosts<strong>of</strong>transport. Comparingthisresulttoasimilarestimation<strong>of</strong>therobotAsimo as<br />

a typical representative for joint angle controlled machines yields an efficiency improvement<br />

by a factor <strong>of</strong> five. Still, the energy consumption is not as low as in humans or actuated<br />

passive walkers. It should be noted that the control system has not yet been optimized<br />

towards low energy consumption at all. Of course, simulation can only provide a rough<br />

guess on the necessary power when applying the control concept to a real robot, but


182 9. Conclusion and Outlook<br />

nevertheless a considerable energetic benefit is to be expected. The necessary requirements<br />

for the peak values <strong>of</strong> joint angles, torques, or power can easily be met by today’s hardware.<br />

Still, the design <strong>of</strong> an actuator featuring the characteristics assumed in this work is a<br />

matter <strong>of</strong> ongoing research. First efforts in this direction made in the scope <strong>of</strong> this thesis<br />

have been introduced.<br />

Finally, the experiments on walking under external disturbances attribute a considerable<br />

robustness to the control system. <strong>Walking</strong> on uphill and downhill slopes as well as over<br />

steps <strong>of</strong> up to 5cm demonstated the capabilities <strong>of</strong> postural reflexes to regulate the forward<br />

velocity by varying step length and ankle torque. With the ground plane inclined to the<br />

side, lateral stabilization by foot placement adaptation and correcting ankle torque could<br />

be illustrated. Also, constant external forces acting on the trunk could be compensated by<br />

the postural reflexes. In all cases, self-stabilizing effects <strong>of</strong> the elastic elements could be<br />

observed to some degree. The analyzed disturbances were <strong>of</strong> small to medium intensity<br />

and a priori unknown to the control system. No external sensor systems <strong>like</strong> cameras or<br />

range finders captured the ground geometry. Considering that the design <strong>of</strong> a postural<br />

control system was not the main focus <strong>of</strong> this thesis, the resulting walking robustness is<br />

more than satisfactory. It shows that a combination <strong>of</strong> local reflexes, independent postural<br />

reflexes using simplified dynamical models, and parallel mechanical elasticity can yield<br />

a capable stability control. It was pointed out that tuning <strong>of</strong> the existing control units<br />

and adding further postural reflexes could increase the robustness <strong>of</strong> the robot against<br />

disturbances.<br />

9.2 Future Work<br />

Several refinements and extensions <strong>of</strong> the presented control system are conceivable. One<br />

potential improvement <strong>of</strong> the existing control units is the reduction <strong>of</strong> the energy consumption.<br />

As already mentioned in the previous chapter, the major part <strong>of</strong> the mechanical<br />

work is currently done by the hip joint during trunk stabilization and leg swing initiation.<br />

Improved timing and slower accelerations could already reduce the necessary energy. Furthermore,<br />

introducing an anticipatory motor pattern for deceleration the upper body at<br />

heel strike could lead to further enhancements.<br />

Possible ways to increase postural stability have been proposed at the end <strong>of</strong> the previous<br />

chapter. Optimizing the existing postural reflexes and the estimation <strong>of</strong> the extrapolated<br />

center <strong>of</strong> mass trajectory would already enhance stability. To further increase walking<br />

robustness, additional control units would have to be introduced. For instance, based on<br />

an estimation <strong>of</strong> the inclination <strong>of</strong> the ground plane, adjustments <strong>of</strong> the upper trunk and<br />

pelvis orientation would facilitate walking on sloped ground. Observation <strong>of</strong> the activity as<br />

well as the target rating <strong>of</strong> reflexes can indicate the necessity to change modulation factors<br />

<strong>like</strong> the walking velocity, or to stimulate additional control units, e.g. for generating higher<br />

leg swings to increase ground clearance.<br />

Such a control unit responsible for raising the foot during the leg swing could be implemented<br />

by an additional motor pattern being superimposed on the action <strong>of</strong> the already<br />

existing control units during the swing phase. Ivanenko et al. have shown that the emg<br />

activity observed during lifting a leg while standing can again be found when human<br />

subjects are ask to lift their legs higher during walking. This suggests a reuse <strong>of</strong> a muscle<br />

stimulation pattern that could be transferred as motor pattern unit acting on the knee


9.2. Future Work 183<br />

learning<br />

mode<br />

global<br />

rating<br />

s<br />

�ı<br />

�e<br />

F (�e,s,�ı)<br />

�u<br />

F (l,g,a,r,�e)<br />

�u<br />

�a<br />

r<br />

�a<br />

r<br />

�esensors<br />

Figure 9.1: Connecting a reinforcement learning module to the control signals <strong>of</strong> an iB2C<br />

behavior. From [Steiner 09], p41.<br />

and hip joint. Making use <strong>of</strong> the behavior-based weighted sum fusion method applied<br />

during the fusion <strong>of</strong> torque commands allows a simple insertion <strong>of</strong> such a control unit to<br />

the existing network. Preliminary tests in this direction have already yielded successful<br />

results.<br />

In combination with an additional modulation signal, a similar approach could be used to<br />

include curve walking. A motor pattern adding a torque impulse to the z-axis <strong>of</strong> the hip<br />

joint during the stance phase would initiate the necessary body rotation. Depending on<br />

the resulting self-induced disturbances, the postural reflexes would have to be modulated<br />

by the walking direction signal. With additional external sensors <strong>like</strong> a vision system,<br />

walking towards a target or directional corrections during straight walking would become<br />

possible.<br />

Of course it is desirable to add further locomotion skills to the control system. A first step<br />

would be the inclusion <strong>of</strong> walking termination. In preliminary experiments, it appeared<br />

to be sufficient to reduce the walking velocity, shorten the step length, and switch back<br />

to standing mode after the heel contact <strong>of</strong> the last step. This procedure will then be<br />

implemented by a gait transition control unit similar to the locomotion mode for walking<br />

initiation. Neuroscientists assume that the same five emg components that explain muscle<br />

activity during walking are also deployed during running. They only differ in timing and<br />

amplitude. It would be interesting to see whether a running gait could also be achieved by<br />

a similar reuse <strong>of</strong> motor patterns for the presented biped.<br />

Up to this point, the suggested future improvements do not require any change <strong>of</strong> the<br />

control concept presented in this thesis. However, the existing drawback <strong>of</strong> having to<br />

manually tune the parameters <strong>of</strong> the control units and <strong>of</strong> the biped model persists. While<br />

it helps that the units have a semantic interpretation and some parameters can be deduced<br />

from their physical meaning, there still remains a set <strong>of</strong> parameters that are difficult to<br />

optimize empirically. One possibility to support the refinement <strong>of</strong> control parameters<br />

could be the inclusion <strong>of</strong> machine learning algorithms. In a first study, the integration<br />

<strong>of</strong> reinforcement learning into the behavior-based robot control architecture iB2C was<br />

investigated [Steiner 09]. As shown in Figure 9.1, a generalized learning module was


184 9. Conclusion and Outlook<br />

inserted behind each behavior module that should be tuned. Based on Q- or Q(λ)-learning<br />

and the iB2C target rating signal used as reward, the activity signal and the control<br />

signals <strong>of</strong> the behavior can be modified by a learned factor or <strong>of</strong>fset. It was shown that<br />

the determination <strong>of</strong> a purely local reward will not be sufficient for the application <strong>of</strong><br />

Q-learning to a complex and interleaved problem <strong>like</strong> bipedal walking. However, the<br />

definition <strong>of</strong> a global reward suitable for reinforcement learning algorithms is far from<br />

being straight-forward.<br />

Another possible approach would be the integration <strong>of</strong> optimization techniques. As<br />

already mentioned, first steps in this direction have been done during the development<br />

on the monopod prototype [Luksch 07]. The applied method based on multiple shooting<br />

optimization techniques required a special representation <strong>of</strong> a complete dynamics model<br />

<strong>of</strong> the robot and its environment. It allows to optimize mechanical parameters as well<br />

as to find an optimal control based on joint torque commands. One idea would be to<br />

include the biologically motivated control system as additional layer and find optimal<br />

parameters instead <strong>of</strong> directly searching for torque trajectories. To this purpose, a suitable<br />

representation <strong>of</strong> the control system that can be used by the optimization tools or an<br />

integration <strong>of</strong> the optimization algorithms in the presented framework would have to<br />

be designed. Alternative optimization algorithms could be considered. In any case, the<br />

problem <strong>of</strong> finding an appropriate target functional remains.<br />

Finally, still no solution exists to the problem <strong>of</strong> finding a joint actuator featuring the<br />

characteristics required by some parts <strong>of</strong> the presented control concept. Several groups are<br />

working on a design for a robotic actuation system with adjustable compliance. Examples<br />

are the series elastic actuators used in the robot Spring Flamingo, the maccepa concept<br />

based on a spring and two actuators for changing the stiffness and the equilibrium point,<br />

amasc working with large leaf springs, and pneumatic muscles [Pratt 95a, Ham 06b,<br />

Hurst 08, Vanderborght 06]. All these approaches have individual drawbacks and do not<br />

fulfill all requirements stated in Chapter 8. First efforts <strong>of</strong> the author and others towards<br />

the development <strong>of</strong> an actuator have been briefly introduced in this work. Initial results<br />

on the combination <strong>of</strong> electrical motors, parallel elastic elements, and compliance control<br />

seem promising [Wahl 09, Blank 09a, Blank 09b]. Further ideas deal with the introduction<br />

<strong>of</strong> an additional series elasticity and the use <strong>of</strong> a different motor and gearbox.<br />

The application <strong>of</strong> the suggested control concept to a real robot should be possible without<br />

major changes except for the adaptation <strong>of</strong> parameters. The dynamics simulation might<br />

not include all effects found in a real machine, but it features at least all major physical<br />

characteristics relevant during walking, thus giving the insights <strong>of</strong> this thesis significant<br />

validity. Since the control concept is based on reactive units and does not depend on<br />

an exact model <strong>of</strong> the robot and its environment, additional discrepancy introduced by<br />

the mechanical or electronic components will be inherently compensated up to a certain<br />

degree.<br />

In conclusion, this thesis has shown that a control methodology based on the transfer <strong>of</strong><br />

concepts found in human motion control can yield remarkable locomotion skills in bipedal<br />

robot. Already now, the presented robot can compete with today’s most advanced bipeds<br />

regarding walking velocity, efficiency, and robustness. Thanks to the flexible, modular<br />

design and the behavior-based architecture, the control network remains easily extensible,<br />

thus allowing to smoothly implement the suggested future work and to advance the control<br />

system’s capabilities even further.


A. Biomechanical and Anatomical<br />

Terms<br />

Throughout this thesis, biomechanical terms are used as they unambiguous describe<br />

processes during bipedal walking. While most terms are be explained at first appearance,<br />

this appendix again summarizes those expressions that can be more easily apprehended<br />

figuratively. Therefore, the upper part <strong>of</strong> Figure A.1 illustrates anatomical directions.<br />

Commonly, these are given relative to the center <strong>of</strong> the body located in the abdomen. The<br />

lower part <strong>of</strong> the figure depicts the horizontal and vertical planes as well as the axes used<br />

to define a coordinate system <strong>of</strong> the human body. Again, these planes and lines all go<br />

through the center <strong>of</strong> the body.<br />

Figure A.2 explains the terms used in biomechanics to describe the direction <strong>of</strong> joint<br />

rotations. These expressions are extensively used in Chapter 3 during the discussion <strong>of</strong><br />

biomechanical gait studies, but are also adopted during the presentation <strong>of</strong> control unit<br />

functions and the analysis <strong>of</strong> the robot’s walking gait. The figure is confined to the flexions,<br />

extensions, ab- and adductions, and rotations <strong>of</strong> those joints that have major influence<br />

during the emergence <strong>of</strong> walking.<br />

At several points <strong>of</strong> this text, the anatomical names <strong>of</strong> human bones are used to avoid<br />

ambiguity. Figure A.3 thus presents a labeled diagram <strong>of</strong> the human skeleton in anterior<br />

and posterior view to indicate the location <strong>of</strong> these bones.<br />

During the design <strong>of</strong> motor patterns and reflexes, data on muscle activity was consulted<br />

repeatedly in form <strong>of</strong> emg analysis. To transfer these biomechanical and neuroscientific<br />

results to technical control units, the location and the function <strong>of</strong> the respective muscles<br />

must be known. To this end, Table A.1 lists the muscles involved in human walking along<br />

with their established abbreviation and the joint action resulting from their contraction.<br />

Biarticular muscle span two joint and thus yield the movement <strong>of</strong> more than only one joint.<br />

Finally, Figure A.4 shows a diagram <strong>of</strong> most <strong>of</strong> the major skeletal muscles to indicate their<br />

location and insertion point.


186 A. Biomechanical and Anatomical Terms<br />

Figure A.1: Anatomical directions and body planes. From [Hamill 03], p15,p22.


Figure A.2: Directions <strong>of</strong> joint flexion, extension, abduction, adduction, and rotation.<br />

From [Hamill 03], pp16–19.<br />

187


188 A. Biomechanical and Anatomical Terms<br />

Abbrev. Muscle Action<br />

Shoulder Complex<br />

BIC Biceps brachii Arm abduction<br />

DELT Deltoid Arm abduction, flexion, horiz. flexion & extension<br />

LD Latissimus dorsi Arm internal rotation, adduction, extension<br />

TRAP Trapezius Shoulder girdle upw. rot., elevation, retract.; arm abd.<br />

Forearm<br />

BIC Biceps brachii Forearm flexion, supination<br />

TRIC Triceps brachii Forearm extension<br />

Hip Joint<br />

ADDL Adductor longus Thigh adduction, internal rotation<br />

ADDM Adductor magnus Thigh adduction, internal rotation<br />

BF Biceps femoris Thigh extension; shank flexion, external rotation<br />

GM Gluteus maximus Thigh extension, external rotation<br />

GMed Gluteus medius Thigh abduction, internal rotation<br />

RF Rectus femoris Thigh flexion; leg extension<br />

SART Sartorius Thigh flexion, external rot.; leg flexion, internal rot.<br />

ST Semitendinosus Thigh extension, internal rot.; leg flexion, intern. rot.<br />

TFL Tensor fascia latae Thigh flexion, abduction, internal rotation<br />

Knee Joint<br />

LG Gastrocnemius lateralis Leg flexion; foot plantarflexion<br />

MG Gastrocnemius medialis Leg flexion; foot plantarflexion<br />

RF Rectus femoris Thigh flexion; leg extension<br />

SART Sartorius Thigh flexion, external rot.; leg flexion, internal rot.<br />

VLat Vastus lateralis Leg extension<br />

VMed Vastus medialis Leg extension<br />

Ankle and Foot<br />

EDL Extensor digitorium brevis Toe extension<br />

FDB Flexor digitorium brevis Toe flexion<br />

LG Gastrocnemius lateralis Leg flexion; foot plantarflexion<br />

MG Gastrocnemius medialis Leg flexion; foot plantarflexion<br />

PERB Peroneus brevis Foot eversion, plantarflexion<br />

PERL Peroneus longus Foot eversion, plantarflexion; forefoot abduction<br />

SOL Soleus Foot plantarflexion<br />

TA Tibialis anterior Foot dorsiflexion, inversion<br />

TP Tibialis posterior Foot inversion, plantarflexion<br />

Vertebral Column<br />

ES Erector spinae Trunk extension<br />

ILIO Iliopsoas Thigh flexion; trunk flexion<br />

OE External oblique Trunk flexion, lateral flexion, rotation to opposite side<br />

OI Internal oblique Trunk flexion, lateral flexion, rotation to same side<br />

RAM Rectus abdominis middle Trunk flexion<br />

RAS Rectus abdominis superior Trunk flexion<br />

STER Sternocleidomastoid Head, cervical flexion, lateral flex., rot. to same side<br />

Table A.1: Skeletal muscles involved during human walking. Muscle actions from [Hamill 03].


�������������<br />

��������<br />

�����������������<br />

��������<br />

�����������������������<br />

������<br />

���������������������<br />

������<br />

������<br />

189x� ����<br />

�������<br />

��������<br />

�������<br />

��������<br />

���������<br />

�������<br />

�������<br />

����<br />

�����������<br />

���������<br />

�����<br />

�������<br />

�����<br />

������<br />

�������<br />

�����������<br />

���������<br />

�������<br />

����<br />

������<br />

�������������<br />

Figure A.3: Diagram <strong>of</strong> major human body bones (Source: Wikimedia Commons, public<br />

domain).


190 A. Biomechanical and Anatomical Terms<br />

Figure A.4: Diagram <strong>of</strong> the skeletal muscles <strong>of</strong> the human body (Source: Wikimedia Commons,<br />

public domain).


B. Dynamics Simulation Framework<br />

As already motivated in the main part <strong>of</strong> this thesis, a suitable simulation environment<br />

is inevitable for the validation <strong>of</strong> the suggested control concept. Building a real robot<br />

possessing all biologically motivated features necessary for the control system unfortunately<br />

is out <strong>of</strong> question in the scope <strong>of</strong> this work. Consequently, the existing simulation and<br />

visualization framework SimVis3D has been extended by the capability to simulate the<br />

dynamics <strong>of</strong> rigid bodies by embedding a physics engine. The selected physics simulation<br />

tool features rigid body mechanics, customizable inertia tensors for each segment, a<br />

penalty-based contact model including coulomb-<strong>like</strong> friction calculation, and access to<br />

internal variables <strong>like</strong> forces and torques acting on bodies and and joint Thus, it provides<br />

a sufficiently details simulation frameworm to evaluate the feasibility <strong>of</strong> the suggested<br />

control concept and to analyze its performance. Flexible mechanisms to design different<br />

environment allow to test the control system in various setups. This appendix gives<br />

additional information on these topics and provide the scene description file used for the<br />

experiments <strong>of</strong> this thesis. A few details on the graphical tools are provided and the<br />

walking gait is visualized in form <strong>of</strong> an image sequence extracted from the simulation<br />

framework.<br />

B.1 MCA2 and SimVis3D<br />

SimVis3D 1 has been and is being developed at the Robotics Research Lab for simulating<br />

robots and their environments [Braun 07]. Being part <strong>of</strong> robotics control framework mca2 2 ,<br />

it can directly be connected to the robot control system substituting an actual robot.<br />

SimVis3D is built on top <strong>of</strong> the rendering library Coin3D 3 , an implementation <strong>of</strong> the<br />

OpenInventor standard by sgi. It manages a tree-<strong>like</strong> scene graph containing the graphical<br />

elements, geometrical transformations, and additional annotation. This graph can then be<br />

rendered by OpenGL for visualization within the mca2 user interface.<br />

Theinitialscenedescriptionisprovidedinform<strong>of</strong>axmlfileastheoneshowninListingB.1.<br />

Three types <strong>of</strong> entries can be added to the scene: parts, elements, and sensors.<br />

1 http://rrlib.cs.uni-kl.de/simvis3d/<br />

2 http://rrlib.cs.uni-kl.de/mca2-kl/<br />

3 http://www.coin3d.org


192 B. Dynamics Simulation Framework<br />

Listing B.1: Sample scene description file <strong>of</strong> SimVis3D.<br />

1 <br />

3 <br />

5


B.2. Embedding <strong>of</strong> the Physics Engine 193<br />

1 <br />

Listing B.2: Sample <strong>of</strong> a collision geometry file.<br />

2 <br />

5 <br />

B.2 Embedding <strong>of</strong> the Physics Engine<br />

The mechanism <strong>of</strong> adding new element types is used to embed the physics engine into<br />

SimVis3D. As rigid body dynamics calculation package, the Newton Game Dynamics<br />

library 4 is applied. Three new element types are defined to modify the scene graph and pass<br />

information to the physics library: physics static, physics object, and physics joint.<br />

Elements <strong>of</strong> the type physics static describe the stationary environment. As illustrated<br />

in line 8 <strong>of</strong> Listing B.1, only a collision geometry needs to be provided. Additionally, the<br />

object’s material can be passed, otherwise it falls back to the default material.<br />

In contrast to static objects, physics object entries as those added in line 9 and line 10<br />

define the movable rigid bodies <strong>of</strong> the simulation scene. Besides the collision geometry<br />

and the material, a mass value is assigned to these objects. The inertia tensor <strong>of</strong> the<br />

object is calculated with the assumption that this mass is distributed evenly within the<br />

given collision body. In addition, an initial linear and rotatory velocity <strong>of</strong> the object can<br />

be set by the velocity and omega parameters. If omitted, these values are set to zero.<br />

Each physics object entry inserts a transformation matrix node to the scene graph at<br />

the given insertion point and exports its values as SimVis3D parameters. After every<br />

simulation cycle, the physics engine will set the current pose <strong>of</strong> each object by accessing<br />

the corresponding matrix entries within the mca2 blackboard containing the SimVis3D<br />

parameters.<br />

Entries <strong>of</strong> the type physics joint define constraints acting between two physical objects.<br />

Chapter 6 already introduced the joint mechanism <strong>of</strong> the Newton library. The necessary<br />

parameters are set within the xml entries as shown in line 11 <strong>of</strong> the listing. The parent<br />

and child parameters indicate which objects are affected by the joint. Leaving the parent<br />

object empty attaches the child object to the static world. The joint_type parameter<br />

determines which kind <strong>of</strong> constraints are added to the objects. Possible joint types include<br />

fixed joints, hinges, sliders, ball joints, or force and torque sensors. Some joint types<br />

<strong>of</strong>fer different actuation methods, e.g. mechanical stops, position control, springs, or the<br />

biological motivated actuator used in this thesis. The actuator can be specified by passing<br />

further parameters such as min, max, motor, or spring. The joint location is defined by<br />

giving the pivot point about which the joint revolves. Its orientation respectively its<br />

translational movement axis is set by the direction parameter. Joint elements do not<br />

modify the scene graph but only pass information to the physics engine.<br />

4 http://www.newtondynamics.com


194 B. Dynamics Simulation Framework<br />

1 <br />

Listing B.3: Sample <strong>of</strong> a material interaction file.<br />

2 <br />

3 <br />

4 <br />

5 <br />

7 <br />

8 <br />

Joint entries as well as physical object entries can extend the interface <strong>of</strong> the mca2 module<br />

encapsulating the physics engine. This is done by adding controller input and sensor<br />

output edges to the module that can then be accessed by the control system. A joint will<br />

normally insert controller input edges to set target values <strong>of</strong> the joint controller. Sensor<br />

information passed to the control system can include data on the current joint position or<br />

the acting torque. An inertial measurement unit provides the forces and torques necessary<br />

to sustain the object constraints. By adding the sensors or controls parameters to the<br />

xml entry <strong>of</strong> physical objects, the current pose <strong>of</strong> the object can be provided or an external<br />

force can be applied to it via controller input edges.<br />

The collision geometry necessary for static and dynamic physical objects is described by an<br />

xml file as the one shown in Listing B.2. Collision geometries can either be composed <strong>of</strong> a<br />

list <strong>of</strong> object primitives or <strong>of</strong> a mesh <strong>of</strong> triangles. The meshes are passed as an additional<br />

file containing a list <strong>of</strong> triangles defined by three points each. Triangle meshes are only<br />

allowed for static objects to reduce the complexity <strong>of</strong> collision checks. Possible geometrical<br />

primitives available to approximate the object shape include spheres, cubes, and cylinders.<br />

Each shape is positioned by six pose values in object local coordinates and parameters<br />

defining the expansions <strong>of</strong> the primitive. The optional color parameters can be evaluated<br />

by a script which can convert collision geometry files to the OpenInventor format. These<br />

visualizations can directly be used in SimVis3D part entries or can help to compare the<br />

collision geometry to an existing more elaborated visualization object.<br />

In case <strong>of</strong> two objects colliding, their materials are evaluated to determine the collision<br />

behavior. Listing B.3 shows a sample file containing the pairwise definitions <strong>of</strong> material<br />

interaction. For every two materials a s<strong>of</strong>tness and elasticity values can be set specifying the<br />

behavior <strong>of</strong> the joints that are created in case <strong>of</strong> a collision. The static and kinetic friction<br />

can also be defined. Additionally, collisions can be discarded by setting the collidable<br />

option to zero. Should an even more sophisticated collision behavior be necessary, a<br />

different callback routine can be specified.


B.3. Interface to <strong>Control</strong> System and Simulation 195<br />

Figure B.1: Graphical user interfacemcagui for accessing the control system and the simulation.<br />

B.3 Interface to <strong>Control</strong> System and Simulation<br />

The robotics control framework mca2 comes with a set <strong>of</strong> graphical tools to access the<br />

control system and the simulation framework during runtime. Figure B.1 shows a section<br />

<strong>of</strong> the graphical user interface mcagui with the configuration used during the development<br />

and validation <strong>of</strong> the presented control system.<br />

A 3D visualization <strong>of</strong> the simulated scene allows to directly observe the effects <strong>of</strong> the control<br />

system (1). Several cameras can be defined and moved freely. The physics simulation can<br />

be resetted, interrupted, or executed stepwise by a control panel (2). As the simulation runs<br />

only a few times slower than realtime, it is possible to interactively control disturbances<br />

or start predefined experiments (3). Arbitrary forces can be applied to robot segments<br />

and experimental results can be recorded. The walking velocity modulation factor can<br />

also be set in the user interface (4). Several graphical widgets can be connected to sensor<br />

information provided by the simulation framework or the control system (5). Finally, the<br />

iB2C behavior signals activation, activity, and target rating <strong>of</strong> all units <strong>of</strong> the control<br />

network can be inspected (6). This provides a quick and clear glimpse on the current<br />

control processes.<br />

A more detailed view on the activity and interaction <strong>of</strong> control units is provided by the<br />

mcabrowser. This graphical tool connects to the control system at runtime and allows to<br />

navigate through the module hierarchy as well as to adapt the parameters <strong>of</strong> all control<br />

units. Figure B.2 depicts screenshots <strong>of</strong> three control groups. The first image shows the


196 B. Dynamics Simulation Framework<br />

(a) (b)<br />

(c)<br />

Figure B.2: Visualization <strong>of</strong> the control system using mcabrowser. (a) Connection <strong>of</strong> the brain,<br />

the spinal cord, and the joint groups. (b) Locomotion modes inside the brain group. (c) <strong>Control</strong><br />

units <strong>of</strong> a leg group during walking phase 1.<br />

highest level <strong>of</strong> the control system. It visualizes the main hierarchy composed <strong>of</strong> the brain<br />

group, the spinal cord group, and the four joint groups representing the upper and lower<br />

body and the two legs. Figure B.2b and c show the brain and one <strong>of</strong> the leg groups in<br />

iB2C visualization mode. The yellow, green, and red horizontal bars in the module boxes<br />

indicate the behavior’s current activation, activity, and target rating. Green and blue<br />

connecting edges mark the transfer <strong>of</strong> stimulation or derived signals. Red connections<br />

signalize inhibition. This view allows to easily grasp the activity <strong>of</strong> control units during<br />

individual motion phases and to debug possible implementation errors.<br />

Finally, Figure B.3 illustrates one cycle <strong>of</strong> the normal walking gait on even ground by<br />

giving a sequence <strong>of</strong> images recorded by the simulation visualization plug-in in the mcagui.<br />

The time interval between two pictures amounts to 75ms.


B.3. Interface to <strong>Control</strong> System and Simulation 197<br />

Figure B.3: Image sequence <strong>of</strong> one gait cycle <strong>of</strong> normal walking on even ground.


198 B. Dynamics Simulation Framework<br />

B.4 Simulation Scenario for Biped Experiments<br />

The scenario used for most experiments described in this thesis is given by the scene<br />

description file printed in the listing below. The right-hand occurrences <strong>of</strong> all segments<br />

existing for both the left and the right side <strong>of</strong> the robot have been omitted for brevity.<br />

The environment features a large plane for the robot to walk on, a movable and rotatable<br />

plate used for the stable standing experiments, a small movable ball to apply arbitrary<br />

disturbances, and to force plates to measure ground reaction forces. The masses <strong>of</strong> all<br />

robot segments can be found in the section on the physical objects starting in line 35. All<br />

joints employed in the robot model are built <strong>of</strong> the type torquepos spring implementing<br />

the actuation described in Chapter 6. The three parameters passed by the motor keyword<br />

represent the maximum motor torque and the proportional factors for the position and<br />

velocity control loop. The parallel spring is described by the three spring parameters,<br />

namely the equilibrium position, the spring rate, and the damper constant.<br />

1 <br />

2 <br />

3<br />

Listing B.4: Scene description file used for most <strong>of</strong> the experiments.<br />

4 <br />

5 <br />

6 <br />

8 <br />

9 <br />

10 <br />

12<br />

13 <br />

15 <br />

17 <br />

19 <br />

21 <br />

23 <br />

25


B.4. Simulation Scenario for Biped Experiments 199<br />

27 <br />

29 <br />

31 <br />

33 <br />

35<br />

36 <br />

37 <br />

39 <br />

40 <br />

41 <br />

42 <br />

43 <br />

44<br />

45 <br />

46 <br />

47 <br />

48 <br />

51 <br />

52 <br />

53 <br />

54 <br />

55


200 B. Dynamics Simulation Framework<br />

auto_freeze="0" attached_to="RADIUS_LEFT" /><br />

56 <br />

57 <br />

58 <br />

59 <br />

60 <br />

61 <br />

62 <br />

63 <br />

65 <br />

67<br />

68 <br />

69 <br />

71


B.4. Simulation Scenario for Biped Experiments 201<br />

max="0.0" pivot="0 0 0" direction="0 1 0" motor="40.0 100.0 30.0" spring="-0.1 '<br />

5.0 2.0" attached_to="RADIUS_LEFT" /><br />

79 <br />

80 <br />

81 <br />

82 <br />

83 <br />

84 <br />

85 <br />

98<br />

99 <br />

101 <br />

103


202 B. Dynamics Simulation Framework


Bibliography<br />

[Akachi 05] K. Akachi, K. Kaneko, N. Kanehira, S. Ota, G. Miyamori, M. Hirata, S. Kajita,<br />

F. Kanehiro,“Development <strong>of</strong> humanoid robot HRP-3P”, in Proceedings <strong>of</strong> the IEEE-<br />

RAS International Conference on <strong>Human</strong>oid <strong>Robots</strong>. 2005, pp. 50–55.<br />

[Albert 01] A. Albert, W. Gerth,“New Path Planning Algorithms for Higher Gait Stability<br />

<strong>of</strong> a <strong>Bipedal</strong> Robot”, in Proceedings <strong>of</strong> 4th International Conference on Climbing and<br />

<strong>Walking</strong> <strong>Robots</strong> (CLAWAR). Karlsruhe, Germany, 2001, pp. 521 – 528.<br />

[Albiez 03a] J. Albiez, T. Luksch, K. Berns, R. Dillmann,“An Activation-Based Behavior<br />

<strong>Control</strong> Architecture for <strong>Walking</strong> Machines”, The International Journal on Robotics<br />

Research, Sage Publications, vol. 22, pp. 203–211, 2003.<br />

[Albiez 03b] J. Albiez, T. Luksch, K. Berns, R. Dillmann,“A Behaviour Network Concept<br />

for <strong>Control</strong>ling <strong>Walking</strong> Machines”, in 2nd International Symposium on Adaptive Motion<br />

<strong>of</strong> Animals and Machines (AMAM). Kyoto, Japan, March 4–8 2003.<br />

[Albiez 03c] J. Albiez, T. Luksch, K. Berns, R. Dillmann,“Reactive reflex-based control<br />

for a four-legged walking machine”, Robotics and Autonomous Systems, Elsevier, 2003.<br />

[Albiez 07] J. Albiez, Verhaltensnetzwerke zur adaptiven Steuerung biologisch motivierter<br />

Laufmaschinen, GCA Verlag, 2007.<br />

[Alexander 84] R. M. Alexander, “The Gaits <strong>of</strong> <strong>Bipedal</strong> and Quadrupedal Animals”,<br />

International Journal <strong>of</strong> Robotics Research, vol. 3-2, pp. 49–59, 1984.<br />

[Alexander 90] R. Alexander,“Three Uses <strong>of</strong> Springs in Legged Locomotion”, The International<br />

Journal <strong>of</strong> Robotics Research, vol. 9, no. 2, pp. 53–61, 1990. Special issue on<br />

legged locomotion.<br />

[Aoi 05] S. Aoi, K. Tsuchiya, “Locomotion <strong>Control</strong> <strong>of</strong> a Biped Robot Using Nonlinear<br />

Oscillators”, Autonomous <strong>Robots</strong>, vol. 19, no. 3, pp. 219–232, 2005.<br />

[Arbulu 07] M. Arbulu, C. Balaguer,“Real -Time gait planning for Rh-1 humanoid robot,<br />

using Local Axis Gait algorithm”, in Proceedings <strong>of</strong> IEEE-RAS International Conference<br />

on <strong>Human</strong>oid <strong>Robots</strong> (<strong>Human</strong>oids). 2007.<br />

[Arisumi 09] H. Arisumi, J. R. Chardonnet, K. Yokoi,“Whole-Body Motion <strong>of</strong> a <strong>Human</strong>oid<br />

Robot for Passing through a Door - Opening a Door by Impulsive Force”, in Proceedings<br />

<strong>of</strong> IEEE/RSJ International Conference on Intelligent <strong>Robots</strong> and Systems (IROS). 2009,<br />

pp. 428–434.


204 Bibliography<br />

[Arkin 00] R. Arkin, A. Kahled, A. Weitzenfeld, F. Cervantes-Prez,“Behavioral models <strong>of</strong><br />

the Praying Mantis as a basis for robotic behavior”, Journal <strong>of</strong> Autonomous Systems,<br />

2000.<br />

[Arkin 98] R. Arkin, Behaviour-Based Robotics, MIT Press, 1998.<br />

[Ayers 00] J. Ayers, J. Witting, N. McGruer, C. Olcott, D. Massa,“Lobster <strong>Robots</strong>”, in<br />

Proceedings <strong>of</strong> the International Symposium on Aqua Biomechanisms, T. Wu, N. Kato,<br />

Eds. Tokai, 2000.<br />

[Baessler 98] U. Baessler, A. Bueschges, “Pattern Generation for Stick Insect <strong>Walking</strong><br />

Movements - Multisensory <strong>Control</strong> <strong>of</strong> a Locomoter Program”, Brain Research Reviews,<br />

no. 27, pp. 65–88, Januar 1998.<br />

[Bauby 00] C. E. Bauby, A. D. Kuo,“Active control <strong>of</strong> lateral balance in human walking”,<br />

Journal <strong>of</strong> Biomechanics, vol. 33, no. 11, pp. 1433–1440, 2000.<br />

[Bent 04] L. R. Bent, J. T. Inglis, B. J. McFadyen, “When is Vestibular Information<br />

Important During <strong>Walking</strong>?”, Journal <strong>of</strong> Neurophysiology, vol. 92, no. 3, pp. 1269–1275,<br />

2004.<br />

[Bizzi 07] E. Bizzi, V. C. K. Cheung, A. d´Avella, P. Saltiel, M. C. Tresch,“Combining<br />

modules for movement”, Brain Research Reviews, vol. 57, pp. 125–133, 2007.<br />

[Bizzi 91] E. Bizzi, F. A. Mussa-Ivaldi, S. Giszter,“Computations underlying the execution<br />

<strong>of</strong> movement: a biological perspective”, Science, vol. 253, no. 5017, pp. 287–291, 1991.<br />

[Blank 09a] S. Blank,“Distributed Compliant <strong>Control</strong> for Biologically Inspired Behavior-<br />

Based <strong>Robots</strong>”, Diploma thesis, University <strong>of</strong> Kaiserslautern, January 2009. unpublished.<br />

[Blank 09b] S. Blank, T. Wahl, T. Luksch, K. Berns,“Biologically Inspired Compliant<br />

<strong>Control</strong> <strong>of</strong> a Monopod Designed for Highly Dynamic Applications”, in IEEE/RSJ<br />

International Conference on Intelligent <strong>Robots</strong> and Systems. St. Louis (MO), USA,<br />

October 2009, pp. pp. 148–153.<br />

[Blickhan 07] R. Blickhan, A. Seyfarth, H. Geyer, S. Grimmer, H. Wagner, M. Günther,<br />

“Intelligence by Mechanics”, Philosophical Transactions <strong>of</strong> the Royal Society <strong>of</strong> London,<br />

Series A, vol. 365, pp. 199–220, 2007.<br />

[Blum 07] Y. Blum, J. Rummel, A. Seyfarth, “Advanced swing leg control for stable<br />

locomotion”, in Autonome Mobile Systeme (AMS). 2007, pp. 301–307.<br />

[Blum 09] Y. Blum, S. W. Lipfert, A. Seyfarth,“Effective leg stiffness in running”, Journal<br />

<strong>of</strong> Biomechanics, vol. 42, no. 14, pp. 2400–2405, 2009.<br />

[Bolder 07] B. Bolder, M. Dunn, M. Gienger, H. Janssen, H. Sugiura, C. Goerick,“Visually<br />

Guided Whole Body Interaction”, in Proceedings <strong>of</strong> the IEEE-RAS International<br />

Conference on <strong>Human</strong>oid <strong>Robots</strong> (<strong>Human</strong>oids). Rome, Italy, April 10–14 2007, pp.<br />

3054–3061.


Bibliography 205<br />

[Bourgeot 02] J. M. Bourgeot, N. Cislo, B. Espiau,“Path-Planning and Tracking in a 3D<br />

Complex Environment for an Anthropomorphic Biped Robot”, in Proceedings <strong>of</strong> the<br />

IEEE International Conference on Intelligent <strong>Robots</strong> and Systems (IROS). Lausanne,<br />

Suisse, 2002, pp. 2509–2514.<br />

[Braun 07] T. Braun, J. Wettach, K. Berns,“A Customizable, Multi-Host Simulation and<br />

Visualization Framework for Robot Applications”, in Recent Progress in Robotics: Viable<br />

Robotic Service to <strong>Human</strong>, Selected papers from ICAR07, S. Lee, I. H. Suh, M. S. Kim,<br />

Eds., LNCIS Series, Springer-Verlag, 2007, vol. 370, pp. 357–369.<br />

[Brooks 86] R. A. Brooks,“A Robust Layered <strong>Control</strong> System for a Mobile Robot”, IEEE<br />

Journal <strong>of</strong> Robotics and Automation, vol. RA-2, no. 1, pp. 14–23, April 1986.<br />

[Brooks 89] R. A. Brooks,“A Robot that Walks - Emergent Behaviors from a Carefully<br />

Evolved Network”, Neural Computation, vol. 1, pp. 355–363, 1989.<br />

[Bueschges 95] A. Bueschges, J. Schmitz, U. Baessler,“Rhythmic patterns in the thoratic<br />

nerve cord <strong>of</strong> the stick insect induced by pilocarpine”, The Journal <strong>of</strong> Experimental<br />

Biology, vol. 198, no. 2, pp. 435–456, September 1995.<br />

[Buschmann 09] T.Buschmann, S.Lohmeier,H.Ulbrich,“Biped<strong>Walking</strong><strong>Control</strong>Basedon<br />

Hybrid Position/Force <strong>Control</strong>”, in Proceedings <strong>of</strong> IEEE/RSJ International Conference<br />

on Intelligent <strong>Robots</strong> and Systems (IROS). 2009, pp. 3019–3024.<br />

[Cappellini 06] G. Cappellini, Y. P. Ivanenko, R. E. Poppele, F. Lacquaniti, “Motor<br />

patterns in human walking and running”, Journal <strong>of</strong> Neurophysiology, vol. 95, 2006.<br />

[Chestnutt 03] J. Chestnutt, J. J. Kuffner, K. Nishiwaki, S. Kagami, “Planning biped<br />

navigationstrategiesincomplexenvironments”, inProceedings <strong>of</strong> the IEEE Intternational<br />

Conference on <strong>Human</strong>oid Robotics (<strong>Human</strong>oids). Munich, Germany, October 2003.<br />

[Chestnutt 05] J. Chestnutt, M. Lau, G. Cheung, J. Kuffner, J. Hodgins, T. Kanade,<br />

“Footstep Planning for the Honda ASIMO <strong>Human</strong>oid”, in Proceedings <strong>of</strong> the IEEE<br />

International Conference on Robotics and Automation (ICRA). 2005, pp. 629– 634.<br />

[Cheung 05] V. C. K. Cheung, A. d´Avella, M. C. Tresch, E. Bizzi,“Central and Sensory<br />

Contributions to the Activation and Organization <strong>of</strong> Muscle Synergies during Natural<br />

Motor Behaviors”, Journal <strong>of</strong> Neuroscience, vol. 25, no. 27, pp. 6419–6434, 2005.<br />

[Chiderova 05] Z. Chiderova, “Dynamische Simulation und Steuerung eines schnelllaufenden<br />

zweibeinigen Roboters”, Diploma thesis, University <strong>of</strong> Kaiserslautern, 2005.<br />

unpublished.<br />

[Collins 05] S. H. Collins, A. L. Ruina, R. Tedrake, M. Wisse,“Efficient <strong>Bipedal</strong> <strong>Robots</strong><br />

Based on Passive-Dynamic Walkers”, Science, vol. 307, pp. 1082–1085, 2005.<br />

[Deshpande 05] N. Deshpande, A. E. Patla,“Dynamic visual–vestibular integration during<br />

goal directed human locomotion”, Experimental Brain Research, vol. 166, no. 2, pp.<br />

237–247, 2005.


206 Bibliography<br />

[Donelan 04] J. Donelan, D. Shipman, R. Kram, A. Kuo, “Mechanical and metabolic<br />

requirements for active lateral stabilization in human walking”, Journal <strong>of</strong> Biomechanics,<br />

vol. 37, no. 6, pp. 827–835, 2004.<br />

[Elble 94] R. J. Elble, C. Moody, K. Leffler, R. Sinha,“The Initiation <strong>of</strong> Normal <strong>Walking</strong>”,<br />

Movement Disorders, pp. 139–146, 1994.<br />

[Endo 04] G. Endo, J. Morimoto, J. Nakanishi, G. Cheng,“An empirical exploration <strong>of</strong> a<br />

neural oscillator for biped locomotion control”, in Proceedings <strong>of</strong> IEEE International<br />

Conference on Robotics and Automation (ICRA), vol. 3. 2004.<br />

[Endo 05] G. Endo, J. Nakanishi, J. Morimoto, G. Cheng, “Experimental Studies <strong>of</strong><br />

a Neural Oscillator for Biped Locomotion with QRIO”, in Proceedings <strong>of</strong> the IEEE<br />

International Conference on Robotics and Automation. Barcelona, Spain, April 2005.<br />

[Escande 09] A. Escande, A. Kheddar,“Contact Planning for Acyclic Motion with Tasks<br />

Constraints”, in Proceedings <strong>of</strong> IEEE/RSJ International Conference on Intelligent <strong>Robots</strong><br />

and Systems (IROS). 2009, pp. 435–440.<br />

[Ferrell 95] C. Ferrell,“Global Behavior via Cooperative Local <strong>Control</strong>”, in Autonomous<br />

<strong>Robots</strong>, vol. 2, no. 2. Springer, 1995, pp. 105 – 125.<br />

[Fischer 06] M. S. Fischer, R. Blickhan,“The tri-segmented limbs <strong>of</strong> therian mammals:<br />

kinematics, dynamics, and self-stabilization - a review”, Journal <strong>of</strong> Experimental Zoology<br />

Part A: Comparative Experimental Biology, vol. 305A, no. 11, pp. 935–952, 2006.<br />

[Flörchinger 05] F. Flörchinger, “Konzipierung eines elastischen Roboterkniegelenks”,<br />

Project thesis, University <strong>of</strong> Kaiserslautern, 2005. unpublished.<br />

[Frank 90] J. S. Frank, M. Earl, “Coordination <strong>of</strong> Posture and Movement”, Physical<br />

Therapy, vol. 70, no. 12, December 1990.<br />

[Frigon 06] A. Frigon, S. Rossignol,“Experiments and models <strong>of</strong> sensorimotor interactions<br />

during locomotion”, Biological Cybernetics, vol. 95, no. 6, pp. 607–627, 2006.<br />

[Fujita 03] M. Fujita, Y. Kuroki, T. Ishida, T. T. Doi,“A Small <strong>Human</strong>oid Robot SDR-<br />

4X for Entertainment Applications”, in Proceedings <strong>of</strong> the IEEE/ASME International<br />

Conference on Advanced Intelligent Mechatronics (AIM). Kobe, Japan, July 20–24 2003,<br />

pp. 938–943.<br />

[Gahery 87] Y. Gahery,“Associated movements, postural adjustments and synergies: some<br />

comments about the history and significance <strong>of</strong> three motor concepts”, Arch Ital Biol.,<br />

vol. 125, pp. 345–360, 1987.<br />

[Garcia 98] M. Garcia,“Stability, scaling, and chaos in passive-dynamic gait models”, Phd<br />

thesis, Cornell University, 1998.<br />

[Geng 06] T. Geng, B. Porr, F. Wörgötter, “A Reflexive Neural Network for Dynamic<br />

Biped <strong>Walking</strong> <strong>Control</strong>”, Neural Computation, vol. 18, no. 5, pp. 1156–1196, 2006.


Bibliography 207<br />

[Geyer 04] H. Geyer, A. Seyfarth, R. Blickhan,“Spring-mass running: simple approximate<br />

solution and application to gait stability”, Journal <strong>of</strong> Theoretical Biology, vol. 232(3),<br />

pp. 315–328, 2004.<br />

[Gienger 00] M. Gienger, K. Löffler, F. Pfeiffer,“A Biped Robot that Joggs”, in Proceedings<br />

<strong>of</strong> the IEEE International Conference on Robotics and Automation (ICRA). 2000, pp.<br />

3334–3339.<br />

[Gienger 02] M. Gienger, K. Löffler, F. Pfeiffer,“<strong>Walking</strong> control <strong>of</strong> a biped robot based<br />

on inertial measurement”, in Proceedings <strong>of</strong> the International Workshop on <strong>Human</strong>oid<br />

and <strong>Human</strong> Friendly Robotics. 2002, pp. 22–30.<br />

[Gilsdorf 06] B. Gilsdorf,“Entwicklung einer Dynamiksimulationsumgebung zum Testen<br />

vonRobotersteuerungen”, Projectthesis, University<strong>of</strong>Kaiserslautern, 2006. unpublished.<br />

[Goswami 99] A. Goswami,“Foot Rotation Indicator (FRI) Point: A New Gait Planning<br />

Tool to Evaluate Postural Stability <strong>of</strong> Biped <strong>Robots</strong>”, in Proceedings <strong>of</strong> the IEEE<br />

International Conference on Robotics and Automation. Detroit, Mai 1999, pp. 47–53.<br />

[Grillner 79] S. Grillner, P. Zangger,“On the central generation <strong>of</strong> locomotion in the low<br />

spinal cat”, Experimental Brain Research, vol. 34, no. 2, pp. 241–261, 1979.<br />

[Günther 04] M. Günther, V. Keppler, A. Seyfarth, R. Blickhan, “<strong>Human</strong> leg design:<br />

optimal axial alignment under constraints”, Journal <strong>of</strong> Mathematical Biology, vol. 48,<br />

no. 6, pp. 623–646, 2004.<br />

[Ham 06a] R. V. Ham, “Compliant Actuation for Biological Inspired <strong>Bipedal</strong> <strong>Walking</strong><br />

<strong>Robots</strong>”, Dissertation, Vrije University Brussels, Brussels, Belgium, July 2006.<br />

[Ham 06b] R. V. Ham, M. V. Damme, B. Vanderborght, B. Verrelst, D. Lefeber,“MAC-<br />

CEPA, the Mechanically Adjustable Compliance and <strong>Control</strong>lable Equilibrium Position<br />

Actuator”, in Proceedings <strong>of</strong> the 10th International Conference on New Actuators (AC-<br />

TUATOR). Bremen, Germany, June 14–16 2006.<br />

[Hamill 03] J. Hamill, K. M. Knutzen, Biomechanical Basis <strong>of</strong> <strong>Human</strong> Movement, 2nd<br />

ed., Lippincott Williams & Wilkins, 2003.<br />

[Hase 99] K. Hase, N. Yamazaki,“Computational evolution <strong>of</strong> human bipedal walking by<br />

a neuro-musculo-skeletal model”, Artificial Life and Robotics, vol. 3, no. 3, pp. 133–138,<br />

1999.<br />

[Hillenbrand 09] C. Hillenbrand, “Sicheres Klettern eines radgetriebenen Roboters mit<br />

Unterdruckkammern an porösen Fächen”, Dissertation, Technische Universität Kaiserslautern,<br />

2009.<br />

[Hirai 98] K. Hirai, M. Hirose, Y. Haikawa, T. Takenaka, “The development <strong>of</strong> Honda<br />

humanoid robot”, in Proceedings <strong>of</strong> the International Conference on Robotics and<br />

Automation. 1998.<br />

[Hirai 99] K. Hirai,“The Honda humanoid robot: development and future perspective”,<br />

Industrial Robot: An International Journal, vol. 26, no. 4, pp. 260–266, 1999.


208 Bibliography<br />

[H<strong>of</strong> 05] A. L. H<strong>of</strong>, M. G. J. Gazendam, W. E. Sinke,“The condition for dynamic stability”,<br />

Journal <strong>of</strong> Biomechanics, vol. 38, no. 1, pp. 1–8, 2005.<br />

[H<strong>of</strong> 07] A. L. H<strong>of</strong>, R. M. van Bockel, T. Schoppen, K. Postema,“<strong>Control</strong> <strong>of</strong> lateral balance<br />

in walking - Experimental findings in normal subjects and above-knee amputees”, Gait<br />

& Posture, vol. 25, no. 2, pp. 250–258, 2007.<br />

[H<strong>of</strong> 08] A. L. H<strong>of</strong>,“The ”extrapolated center <strong>of</strong> mass”concept suggests a simple control<br />

<strong>of</strong> balance in walking”, <strong>Human</strong> Movement Science, vol. 27, no. 1, pp. 112–125, 2008.<br />

[H<strong>of</strong>schulte 04] J. H<strong>of</strong>schulte, M. Seebode, W. Gerth,“Parallel Manipulator Hip Joint for<br />

a <strong>Bipedal</strong> Robot”, in Proceedings <strong>of</strong> the 7th International Conference on Climbing and<br />

<strong>Walking</strong> <strong>Robots</strong> (CLAWAR). Madrid, Spain, 2004.<br />

[Horak 06] F. B. Horak,“Postural orientation and equilibrium: what do we need to know<br />

about neural control <strong>of</strong> balance to prevent falls?”, Age and Ageing, vol. 35-S2, pp. ii7–ii11,<br />

2006.<br />

[Horak 90] F. B. Horak, L. M. Nashner, H. C. Diener,“Postural strategies associated with<br />

somatosensory and vestibular loss”, Experimental Brain Research, vol. 82, pp. 167–177,<br />

1990.<br />

[Hosoda 07a] K. Hosoda, K. Narioka,“Synergistic 3D limit cycle walking <strong>of</strong> an anthropomorphic<br />

biped robot”, in Proceedings <strong>of</strong> the IEEE/RSJ International Conference on<br />

Intelligent <strong>Robots</strong> and Systems (IROS). 2007, pp. 470–475.<br />

[Hosoda 07b] K. Hosoda, T. Takuma, A. Nakamoto, S. Hayashi, “Biped robot design<br />

powered by antagonistic pneumatic actuators for multi-modal locomotion”, Robotics<br />

and Autonomous Systems, vol. 56, no. 1, pp. 46–53, 2007.<br />

[Huang 05] Q. Huang, Y. Nakamura, “Sensory reflex control for humanoid walking”,<br />

Robotics, IEEE Transactions on [see also Robotics and Automation, IEEE Transactions<br />

on], vol. 21, no. 5, pp. 977–984, 2005.<br />

[Hunt 94] K. D. Hunt,“The evolution <strong>of</strong> human bipedality: ecology and functional morphology”,<br />

Journal <strong>of</strong> <strong>Human</strong> Evolution, vol. 26, pp. 183–202, 1994.<br />

[Hurst 08] J. W. Hurst, A. A. Rizzi,“Series Compliance for an Efficient Running Gait:<br />

Lessons Learned from the Electric Cable Differential Leg”, IEEE Robotics and Automation<br />

Magazine, vol. 15, no. 3, pp. 42–51, September 2008.<br />

[Iida 07] F. Iida, J. Rummel, A. Seyfarth,“<strong>Bipedal</strong> <strong>Walking</strong> and Running with Compliant<br />

Legs”, in Proceedings <strong>of</strong> the IEEE International Conference on Robotics and Automation<br />

(ICRA). Roma, Italy, April 10–14 2007, pp. 3970–3975.<br />

[Ito 82] M. Ito, “Cerebellar control <strong>of</strong> the vestibulo-ocular reflex–around the flocculus<br />

hypothesis”, Annual Review <strong>of</strong> Neuroscience, vol. 5, no. 1, pp. 275–297, 1982.<br />

[Ivanenko 04] Y. P. Ivanenko, R. E. Poppele, F. Lacquaniti,“Five basic muscle activation<br />

patterns account for muscle activity during human locomotion”, Journal <strong>of</strong> Physiology,<br />

vol. 556, 2004.


Bibliography 209<br />

[Ivanenko 06] Y. P. Ivanenko, R. E. Poppele, F. Lacquaniti,“Motor <strong>Control</strong> Programs<br />

and <strong>Walking</strong>”, The Neuroscientist, vol. 12, no. 4, pp. 339–348, 2006.<br />

[Kagami 03] S. Kagami, M. Mochimaru, Y. Ehara, N. Miyata, K. Nishiwaki, T. Kanade,<br />

H. Inoue,“Measurement and comparison <strong>of</strong> human and humanoid walking”, in Proceedings<br />

<strong>of</strong> IEEE International Symposium on Computational Intelligence in Robotics and<br />

Automation, vol. 2. 2003, pp. 918–922.<br />

[Kajita 02] S. Kajita, T. Nagasaki, K. Yokoi, K. Kaneko, K. Tanie, “Running Pattern<br />

Generation for a <strong>Human</strong>oid Robot”, in Proceeding <strong>of</strong> IEEE International Conference on<br />

Robotics and Automation (ICRA). 2002.<br />

[Kajita 03] S. Kajita, F. Kanehiro, K. Fujiwara, K. Harada, K. Yokoi, H. Hirukawa,<br />

“Biped <strong>Walking</strong> Pattern Generation by using Preview <strong>Control</strong> <strong>of</strong> Zero-Moment Point”,<br />

in Proceedings <strong>of</strong> IEEE International Conference on Robotics and Automation (ICRA).<br />

2003.<br />

[Kajita 05] S. Kajita, T. Nagasaki, K. Kaneko, K. Yokoi, K. Tanie,“A Running <strong>Control</strong>ler<br />

<strong>of</strong> <strong>Human</strong>oid Biped HRP-2LR”, in Proceeding <strong>of</strong> IEEE International Conference on<br />

Robotics and Automation (ICRA). 2005.<br />

[Kajita 06] S. Kajita, M. Morisawa, K. Harada, K. Kaneko, F. Kanehiro, K. Fujiwara,<br />

H. Hirukawa,“Biped <strong>Walking</strong> Pattern Generator allowing Auxiliary ZMP <strong>Control</strong>”, in<br />

Proceedings <strong>of</strong> IEEE/RSJ International Conference on Intelligent <strong>Robots</strong> and Systems<br />

(IROS). 2006.<br />

[Kajita 07] S. Kajita, K. Kaneko, M. Morisawa, S. Nakaoka, H. Hirukawa,“ZMP-based<br />

Biped Running Enhanced by Toe Springs”, in Proceeding <strong>of</strong> IEEE International Conference<br />

on Robotics and Automation (ICRA). 2007.<br />

[Kaneko 02a] K.Kaneko,S.Kajita,F.Kanehiro,K.Y.K.Fujiwara,H.Hirukawa,T.Kawai,<br />

M. Hirata, T. Isozumi,“Design <strong>of</strong> Advanced Leg Module for <strong>Human</strong>oid Robotics Project<br />

<strong>of</strong> METI”, in Proceedings <strong>of</strong> the International Conference on Robotics and Automation<br />

(ICRA). 2002.<br />

[Kaneko 02b] K. Kaneko, F. Kanehiro, S. Kajita, K. Yokoyama, K. Akachi, T. Kawasaki,<br />

S. Ota, T. Isozumi, “Design <strong>of</strong> Prototype <strong>Human</strong>oid Robotics Platform for HRP”,<br />

in Proceedings <strong>of</strong> the IEEE/RSJ International Conference on Intelligent <strong>Robots</strong> and<br />

Systems (IROS). 2002.<br />

[Kaneko 04] K. Kaneko, F. Kanehiro, S. K. adn H Hirukawa, T. Kawasaki, M. Hirata,<br />

K.Akachi,T.Isozumi,“<strong>Human</strong>oidrobotHRP-2”,inIEEE-RAS International Conference<br />

on Robotics and Automation (ICRA). New Orleans, USA, April 26 – May 1 2004, pp.<br />

1083–1090.<br />

[Kaneko 08] K. Kaneko, K. Harada, F. Kanehiro, G. Miyamori, K. Akachi,“<strong>Human</strong>oid<br />

robot HRP-3”, in Proceedings <strong>of</strong> IEEE/RSJ International Conference on Intelligent<br />

<strong>Robots</strong> and Systems (IROS). 2008, pp. 2471–2478.


210 Bibliography<br />

[Katayama 85] T.Katayama, T.Ohki, T.Inoue, T.Kato,“Design<strong>of</strong>anOptimal<strong>Control</strong>ler<br />

for a Discrete Time System Subject to Previewable Demand”, International Journal <strong>of</strong><br />

<strong>Control</strong>, vol. 41, no. 3, pp. 677–699, 1985.<br />

[Katayama 95] O. Katayama, Y. Kurematsu, S. Kitamura,“Theoretical studies on neuro<br />

oscillator for application <strong>of</strong> biped locomotion”, in Proceedings <strong>of</strong> IEEE International<br />

Conference on Robotics and Automation (ICRA). 1995.<br />

[Kato 73] I. Kato,“The WABOT-1”, Bulletin <strong>of</strong> Science and engineering Research Laboratory,<br />

Waseda University, no. 62, 1973.<br />

[Kerscher 06] T. Kerscher, J. Albiez, J. M. Zöllner, R. Dillmann, “Evaluation <strong>of</strong> the<br />

Dynamic Model <strong>of</strong> Fluidic Muscles using Quick-Release”, in Proceedings <strong>of</strong> the IEEE<br />

/ RAS-EMBS International Conference on Biomedical Robotics and Biomechatronics.<br />

2006.<br />

[Kerscher 07] T. Kerscher, M. Göller, J. M. Zöllner, R. Dillmann, “Regelungsstrategie<br />

für Zweibeiniges Elastisches Laufen Mittels ”Virtual Model <strong>Control</strong>””, in Autonome<br />

Mobile Systeme 2007, ser. Informatik aktuell, K. Berns, T. Luksch, Eds. Kaiserslautern,<br />

Germany: Springer Verlag, October 18–19 2007, pp. 268 –274. ISBN: 978-3-540-74763-5.<br />

[Kim 05] J.-Y. Kim, I.-W. Park, J. Lee, M.-S. Kim, B. kyu Cho, J.-H. Oh, “System<br />

Design and Dynamic <strong>Walking</strong> <strong>of</strong> <strong>Human</strong>oid Robot KHR-2”, in Proceedings <strong>of</strong> the IEEE<br />

International Conference on Robotics and Automation (ICRA). 2005, pp. 1431–1436.<br />

[Kimura 04] H. Kimura, Y. Fukuoka, A. H. Cohen,“Biologically Inspired Adaptive Dynamic<br />

<strong>Walking</strong> <strong>of</strong> a Quadruped Robot”, in Proceedings <strong>of</strong> the 8th International Conference<br />

on the Simulation <strong>of</strong> Adaptive Behavior (SAB). Los Angles, USA, 2004, pp.<br />

201–210.<br />

[Kirker 00] S. G. B. Kirker, D. S. Simpson, J. R. Jenner, A. M. Wing,“Stepping before<br />

standing: hip muscle function in stepping and standing balance after stroke”, J. Neurol.<br />

Neurosurg. Psychiatry, vol. 68, pp. 458–464, 2000.<br />

[Kleinlützum 05] K. Kleinlützum, T. Luksch, D. Schmidt, K. Berns, “Selbstständige<br />

Erstellung einer abstrakten topologiebasierten Karte für die autonome Exploration”, in<br />

Autonome Mobile Systeme (AMS) 2005. December 2005, pp. 149–155.<br />

[Koch 04] J. Koch, “Lage- und Positionsbestimmung mobiler Roboter auf Basis eines<br />

intertialen Messsystems und ergänzender Sensorik”, Diploma thesis, University <strong>of</strong> Kaiserslautern,<br />

October 2004. unpublished.<br />

[Kuffner 01] J. J. Kuffner, K. Nishiwaki, S. Kagami, M. Inaba, H. Inoue,“Motion Planning<br />

for <strong>Human</strong>oid <strong>Robots</strong> Under Obstacle and Dynamic Balance Constraints”, in Proceedings<br />

<strong>of</strong> the IEEE International Conference on Robotics and Automation (ICRA). 2001, pp.<br />

692–698.<br />

[Kuo 01] A. D. Kuo,“A simple model <strong>of</strong> bipedal walking predicts the preferred speed-step<br />

length relationship”, Journal <strong>of</strong> Biomechanical Engineering, vol. 123, pp. 264–269, 2001.


Bibliography 211<br />

[Kuo 99] A. D. Kuo,“Stabilization <strong>of</strong> lateral motion in passive dynamic walking”, International<br />

Journal <strong>of</strong> Robotics Research, vol. 18, no. 9, pp. 917–930, 1999.<br />

[Lamp 05] D. Lamp,“Kinematische Modellierung und Simulation eines humanoiden Beins”,<br />

Project thesis, University <strong>of</strong> Kaiserslautern, 2005. unpublished.<br />

[Leineweber 03a] D. B. Leineweber, I. Bauer, H. G. Bock, J. P. Schlöder,“An efficient<br />

multiple shooting based reduced SQP strategy for large-scale dynamic process optimization<br />

- part I: theoretical aspects”, Computers and Chemical Engineering, vol. 27, pp.<br />

157–166, 2003.<br />

[Leineweber 03b] D. B. Leineweber, A. Schäfer, H. G. Bock, J. P. Schlöder,“An efficient<br />

multiple shooting based reduced SQP strategy for large-scale dynamic process optimization<br />

- part II: s<strong>of</strong>tware aspects and applications”, Computers and chemical engineering,<br />

vol. 27, pp. 167–174, 2003.<br />

[Lemay 01] M. A. Lemay, J. E. Galagan, N. Hogan, E. Bizzi,“Modulation and vectorial<br />

summation <strong>of</strong> the spinalized frog´s hindlimb end-point force produced by intraspinal electrical<br />

stimulation <strong>of</strong> the cord”, IEEE Transactions on Neural Systems and Rehabilitation<br />

Engineering, vol. 9, March 2001.<br />

[Lim 06] H. Lim, A. Takanishi,“Biped walking robots created at Waseda University: WL<br />

and WABIAN family”, Philosophical Transactions <strong>of</strong> the Royal Society A: Mathematical,<br />

Physical and Engineering Sciences, vol. 365, no. 1850, pp. 49–64, 2006.<br />

[Löffler 03] K. Löffler, M. Gienger, F. Pfeiffer,“Sensors and <strong>Control</strong> Concept <strong>of</strong> <strong>Walking</strong>”,<br />

International Journal <strong>of</strong> Robotics Research, vol. 22, pp. 229–239, 2003.<br />

[Lohmeier 04] S. Lohmeier, K. Löffler, M. Gienger, H. Ulbrich, F. Pfeiffer, “Computer<br />

System and <strong>Control</strong> <strong>of</strong> Biped”, in Proceedings <strong>of</strong> the 2004 IEEE International Conference<br />

on Robotics and Automation (ICRA). New Orleans, LA, USA, April 2004, pp. 4222 –<br />

4227.<br />

[Lohmeier 06] S. Lohmeier, T. Buschmann, H. Ulbrich, F. Pfeiffer,“Modular Joint Design<br />

for a Performance Enhanced <strong>Human</strong>oid Robot”, in Proceedings <strong>of</strong> the IEEE International<br />

Conference on Robotics and Automation (ICRA). 2006.<br />

[Luksch 02] T. Luksch,“Verhaltensbasierte freie Gangart für eine vierbeinige Laufmaschine”,<br />

Diploma thesis, University <strong>of</strong> Karlsruhe, 2002. unpublished.<br />

[Luksch 05] T. Luksch, K.Berns, F.Flörchinger,“ActuationSystemand<strong>Control</strong>Concept<br />

for a Running Biped”, in Fast Motions in Biomechanics and Robotics, ser. Lecture Notes<br />

in<strong>Control</strong>andInformationSciences,M.Diehl,K.Mombauer,Eds.,Heidelberg,Germany:<br />

Springer Verlag, September 7–9 2005.<br />

[Luksch 07] T. Luksch, K. Berns, K. Mombaur, G. Schultz,“Using Optimization TechniquesfortheDesignand<strong>Control</strong><strong>of</strong>FastBipeds”,<br />

inProceedings <strong>of</strong> the 10th International<br />

Conference on Climbing and <strong>Walking</strong> <strong>Robots</strong> (CLAWAR). Singapore, July 16–18 2007.


212 Bibliography<br />

[Luksch 08a] T. Luksch, K. Berns,“<strong>Control</strong>ling Dynamic Motions <strong>of</strong> Biped <strong>Robots</strong> with<br />

Reflexes and Motor Patterns”, in Fourth International Symposium on Adaptive Motion<br />

<strong>of</strong> Animals and Machines (AMAM). Cleveland, USA, June 1–6 2008, pp. 115–116.<br />

[Luksch 08b] T. Luksch, K. Berns,“Initiating Normal <strong>Walking</strong> <strong>of</strong> a Dynamic Biped with<br />

a Biologically Motivated <strong>Control</strong>”, in Proceedings <strong>of</strong> the 11th International Conference<br />

on Climbing and <strong>Walking</strong> <strong>Robots</strong> (CLAWAR). Coimbra, Portugal, September 8–10 2008.<br />

[Manoonpong 07] P. Manoonpong, T. Geng, B. Porr, F. Wörgötter,“The RunBot architecture<br />

for adaptive, fast, dynamic walking.”, IEEE Symposium on Circuits and Systems<br />

(ISCAS), 2007.<br />

[Massaad 07] F. Massaad, T. M. Lejeune, C. Detrembleur,“The up and down bobbing<br />

<strong>of</strong> human walking: a compromise between muscle work and efficiency”, Journal <strong>of</strong><br />

Physiology, vol. 582, no. 2, pp. 789–799, 2007.<br />

[Matarić 97] M. J. Matarić,“Behavior-Based <strong>Control</strong>: Examples from Navigation, Learning,<br />

andGroupBehavior”, Journal <strong>of</strong> Experimental and Theoretical Artificial Intelligence,<br />

vol. Special issue on S<strong>of</strong>tware Architectures for Physical Agents, 9(2-3), pp. 323–336,<br />

1997.<br />

[Matsuoka 87] K. Matsuoka,“Mechanisms <strong>of</strong> Frequency and Pattern <strong>Control</strong> in the Neural<br />

Rhythm Generators”, Biological Cybernetics, vol. 56, pp. 345–353, 1987.<br />

[McGeer 90] T. McGeer,“Passive Dynamic <strong>Walking</strong>”, International Journal <strong>of</strong> Robotics<br />

Research, vol. 1, no. 9, pp. 62–82, 1990.<br />

[McGeer 93] T. McGeer, “Dynamics and <strong>Control</strong> <strong>of</strong> <strong>Bipedal</strong> Locomotion”, Journal <strong>of</strong><br />

Theoretical Biology, vol. 163, pp. 277–314, 1993.<br />

[Mombaur 01] K.D.Mombaur, H.G.Bock, R.W.Longman,“<strong>Human</strong>-LikeActuated<strong>Walking</strong><br />

that is Asymptotically Stable Without Feedback”, in Proceedings <strong>of</strong> the International<br />

Conference on Robotics and Automation (ICRA). 2001, pp. 4128–4133.<br />

[Mombaur 05a] K. D. Mombaur, R. W. Longman, H. G. Bock, J. P. Schlöder,“Open-loop<br />

stable running”, Robotica, vol. 23, no. 1, pp. 21–33, 2005.<br />

[Mombaur 05b] K. D. Mombaur, H. G. Bock, J. P. Schlöder, R. W. Longman,“Open-loop<br />

stable solution <strong>of</strong> periodic optimal control problems”, ZAMM, vol. 85, no. No. 7, 2005.<br />

[Mombaur 05c] K. D. Mombaur, H. G. Bock, J. P. Schlöder, R. W. Longman, “Selfstabilizing<br />

somersaults”, IEEE Transactions on Robotics, vol. 21, no. 6, 2005.<br />

[Mombaur 07] K. Mombaur, “<strong>Human</strong>-<strong>like</strong> Running can be Open-Loop Stable”, in Autonome<br />

Mobile Systeme 2007, 2007, pp. 282–286.<br />

[Morris 06] B. Morris, E. R. Westervelt, C. Chevallereau, G. Buche, J. W. Grizzle,“Achieving<br />

<strong>Bipedal</strong> Running with RABBIT: Six Steps toward Infinity”, in Fast Motions in<br />

Robotics and Biomechanics - Optimization and Feedback <strong>Control</strong>, ser. Lecture Notes in<br />

<strong>Control</strong> and Information Science, K. M. M Diehl, Ed. Springer, 2006.


Bibliography 213<br />

[Murray 64] M. Murray, A. B. Drought, R. C. Kory,“<strong>Walking</strong> Patterns <strong>of</strong> Normal Men”,<br />

The Journal <strong>of</strong> Bone and Joint Surgery, vol. 46, no. 2, pp. 335–360, 1964.<br />

[Mutlu 06] B. Mutlu, S. Osman, J. Forlizzi, J. Hodgins, S. Kiesler,“Perceptions <strong>of</strong> ASIMO:<br />

an exploration on co-operation and competition with humans and humanoid robots”, in<br />

Proceedings <strong>of</strong> the 1st ACM SIGCHI/SIGART conference on <strong>Human</strong>-robot interaction<br />

(HRI). New York, USA, 2006, pp. 351–352.<br />

[Nagasaki 03] T. Nagasaki, S. Kajita, K. Yokoi, K. Kaneko, K. Tanie,“Running pattern<br />

generation andits evaluation using a realistic humanoid model”, in Proceedings <strong>of</strong> the<br />

IEEE International Conference on Robotics and Automation (ICRA). 2003, pp. 1336–<br />

1342.<br />

[Nicholls 92] J. G. Nicholls, A. R. Martin, B. G. Wallace, From Neuron to Brain, Massachusettes:<br />

Sinauer Associates Inc., 1992.<br />

[Nielsen 03] J. B. Nielsen, “How we Walk: Central <strong>Control</strong> <strong>of</strong> Muscle Activity during<br />

<strong>Human</strong> <strong>Walking</strong>”, The Neuroscientist, vol. 9, no. 3, pp. 195–204, 2003.<br />

[Nishiwaki 06] K.Nishiwaki, J.J.Kuffner, S.Kagami, M.Inaba, H.Inoue,“Theexperimental<br />

humanoid robot H7 : A research platform for autonomous behavior”, Philosophical<br />

Transactions <strong>of</strong> the Royal Society A: Mathematical, Physical and Engineering Sciences,<br />

vol. special issue, pp. 1–28, 2006.<br />

[Ogihara 01] N. Ogihara, N. Yamazaki,“Generation <strong>of</strong> human bipedal locomotion by a<br />

bio-mimetic neuro-musculo-skeletal model”, Biological Cybernetics, vol. 84, no. 1, pp.<br />

1–11, 2001.<br />

[Ogura 06a] Y. Ogura, H. Aikawa, K. Shimomura, H. Kondo, A. Morishima, H. ok Lim,<br />

A. Takanishi,“Development <strong>of</strong> A <strong>Human</strong>oid Robot WABIAN-2”, in Proceedings <strong>of</strong> the<br />

IEEE International Conference on Robotics and Automation (ICRA). 2006.<br />

[Ogura 06b] Y. Ogura, K. Shimomura, H. Kondo, A. Morishima, T. Okubo, S. Momoki,<br />

H. ok Lim, A. Takanishi,“<strong>Human</strong>-<strong>like</strong> <strong>Walking</strong> with Knee Stretched, Heel-contact and<br />

Toe-<strong>of</strong>f Motion by a <strong>Human</strong>oid Robot”, in Proceedings <strong>of</strong> the IEEE/RSJ International<br />

Conference on Intelligent <strong>Robots</strong> and Systems (IROS). 2006.<br />

[Olree 95] K. S. Olree, C. L. Vaughan,“Fundamental patterns <strong>of</strong> bilateral muscle activity<br />

in human locomotion”, Biological Cybernetics, vol. 73, no. 5, pp. 409–414, 1995.<br />

[Park 05] I. W. Park, J. H. Oh,“Mechanical Design <strong>of</strong> <strong>Human</strong>oid Robot platfrom KHR-3<br />

(KAIST <strong>Human</strong>oid Robot-3: HUBO)”, in Proceedings <strong>of</strong> the IEEE/RAS International<br />

Conference on <strong>Human</strong>oid <strong>Robots</strong> (<strong>Human</strong>oids). 2005, pp. 321–326.<br />

[Patla 85] A. E. Patla,“Some characteristics <strong>of</strong> EMG patterns during locomotion: implications<br />

for the locomotor control process”, Journal <strong>of</strong> Motor Behavior, vol. 17, no. 4,<br />

pp. 443–61, 1985.<br />

[Patla 97] A. E. Patla,“Understanding the roles <strong>of</strong> vision in the control <strong>of</strong> human locomotion”,<br />

Gait & Posture, vol. 5, no. 1, pp. 54–69, 1997.


214 Bibliography<br />

[Paul 05] C. Paul, “Sensorimotor <strong>Control</strong> <strong>of</strong> Biped Locomotion”, Adaptive Behavior -<br />

Animals, Animats, S<strong>of</strong>tware Agents, <strong>Robots</strong>, Adaptive Systems, vol. 13, no. 1, pp. 67–80,<br />

2005.<br />

[Pearson 93] K. G. Pearson,“Common Principles <strong>of</strong> Motor <strong>Control</strong> in Vertebrates and<br />

Invertebrates”, Annual Reviews on Neuroscience, vol. 16, pp. 265–297, 1993.<br />

[Peng 06] Z. Peng, Y. Fu, Z. Tang, Q. Huang, Taoxiao,“Online <strong>Walking</strong> Pattern Generation<br />

and System S<strong>of</strong>tware <strong>of</strong> <strong>Human</strong>oid BHR-2”, in Proceedings <strong>of</strong> the IEEE/RSJ<br />

International Conference on Intelligent <strong>Robots</strong> and Systems (IROS). Beijing, China,<br />

2006, pp. 5471–5476.<br />

[Pettersson 98] L. Pettersson, K. Jansson, H. Rehbinder, J. Wikander,“Behavior-based<br />

control <strong>of</strong> a four legged walking robot”, in Proceedings <strong>of</strong> Mechatronics. 1998, pp. 361–366.<br />

[Pirjanian 99] P. Pirjanian, “Behaviour Coordination Mechanisms — State-<strong>of</strong>-the-art”,<br />

Institute for Robotics and Intelligent Systems, School <strong>of</strong> Engineering, University <strong>of</strong><br />

Southern California, Tech. Rep. IRIS-99-375, October 1999.<br />

[Plagenhoef 83] S. Plagenhoef,“Anatomical Data for Analyzing <strong>Human</strong> Motion”, Research<br />

Quarterly for Exercise and Sport, vol. 54, no. 2, pp. 169–78, 1983.<br />

[Pratt 00] J. Pratt,“Exploiting Inherent Robustness and Natural Dynamics in the <strong>Control</strong><br />

<strong>of</strong> <strong>Bipedal</strong> <strong>Walking</strong> <strong>Robots</strong>.”, in PhD thesis. MIT Press, Cambridge, 2000.<br />

[Pratt 01] J. Pratt, C. M. Chew, A. Torres, P. Dilworth, G. Pratt,“Virtual Model <strong>Control</strong>:<br />

An Intuitive Approach for <strong>Bipedal</strong> Locomotion”, International Journal <strong>of</strong> Robotics<br />

Research, vol. 20, no. 2, pp. 129–143, 2001.<br />

[Pratt 02] J. Pratt, B. Krupp, C. Morse,“Series Elastic Actuators for High Fidelity Force<br />

<strong>Control</strong>”, Industrial Robot Journal, vol. 29, no. 3, pp. 234–241, 2002.<br />

[Pratt 09] J. E. Pratt, B. Krupp, V. Ragusila, J. Rebula, T. Koolen, N. van Nieuwenhuizen,<br />

C. Shake, T. Craig, J. Taylor, G. Watkins, P. Neuhaus, M. Johnson, S. Shooter,<br />

K. Buffinton,“The Yobotics-IHMC Lower Body <strong>Human</strong>oid Robot”, in Proceedings <strong>of</strong><br />

IEEE/RSJ International Conference on Intelligent <strong>Robots</strong> and Systems (IROS). 2009,<br />

pp. 410–411.<br />

[Pratt 95a] G. A. Pratt, M. M. Williamson,“Series Elastic Actuators”, in Proceedings <strong>of</strong><br />

the IEEE/RSJ International Conference on Intelligent <strong>Robots</strong> and Systems. Pittsburgh,<br />

November 1995, pp. 399–406.<br />

[Pratt 95b] J. E. Pratt, “Virtual Model <strong>Control</strong> <strong>of</strong> a Biped <strong>Walking</strong> Robot”, Diploma<br />

thesis, Massachusetts Institute <strong>of</strong> Technology, August 1995.<br />

[Pratt 97] J. Pratt, P. Dilworth, G. Pratt, “Virtual Model <strong>Control</strong> <strong>of</strong> a <strong>Bipedal</strong> <strong>Walking</strong><br />

Robot”, Proceedings <strong>of</strong> the International Conference on Robotics and Automation<br />

(ICRA), pp. 193–198, 1997. Albuquerque, NM.<br />

[Proetzsch 04] M. Proetzsch, T. Luksch, K. Berns,“Behaviour-based Motion <strong>Control</strong> for<br />

Offroad Navigation”, in International Workshop on <strong>Human</strong>itarian Demining (HUDEM).<br />

Brussels, Belgium, June 16–18 2004.


Bibliography 215<br />

[Proetzsch 05] M. Proetzsch, T. Luksch, K. Berns,“Fault-Tolerant Behavior-Based Motion<br />

<strong>Control</strong> for Offroad Navigation”, in Proceedings <strong>of</strong> the 20th IEEE International<br />

Conference on Robotics and Automation (ICRA). Barcelona, Spain, April 18–22 2005,<br />

pp. 4697–4702.<br />

[Proetzsch 07] M. Proetzsch, T. Luksch, K. Berns,“The Behaviour-Based <strong>Control</strong> Architecture<br />

iB2C for Complex Robotic Systems”, in Proceedings <strong>of</strong> the 30th Annual German<br />

Conference on Artificial Intelligence (KI). Osnabrück, Germany, September 10–13 2007,<br />

pp. 494–497.<br />

[Proetzsch 10] M.Proetzsch,T. Luksch, K.Berns,“Development<strong>of</strong>ComplexRoboticSystems<br />

Using the Behavior-Based <strong>Control</strong> Architecture iB2C”, Robotics and Autonomous<br />

Systems, vol. 58, no. 1, pp. 46–67, 2010.<br />

[Prokop 97] T. Prokop, M. Schubert, W. Berger,“Visual influence on human locomotion<br />

Modulation to changes in optic flow”, Experimental Brain Research, vol. 114, no. 1, pp.<br />

63–70, 1997.<br />

[Raibert 86] M. H. Raibert, Legged <strong>Robots</strong> That Balance, Cambridge, MA: MIT Press,<br />

1986.<br />

[Richmond 02] B. G. Richmond, D. R. Begun, D. S. Strait,“Origin <strong>of</strong> human bipedalism:<br />

The knuckle-walking hypothesis revisited”, American Journal <strong>of</strong> Physical Anthropology,<br />

vol. 116, no. S33, pp. 70–105, 2002.<br />

[Roa 05] M. Roa, R. Ramírez, D. Garzón,“Development <strong>of</strong> Biped <strong>Robots</strong> at the National<br />

University <strong>of</strong> Colombia”, in Proceedings <strong>of</strong> the 8th International Conference on Climbing<br />

and <strong>Walking</strong> <strong>Robots</strong> (CLAWAR), Springer. 2005.<br />

[Rodgers 88] M. M. Rodgers,“Dynamic biomechanics <strong>of</strong> the normal foot and ankle during<br />

walking and running”, Physical Therapy, vol. 68, no. 12, pp. 1822–1830, 1988.<br />

[Rossi 88] A. Rossi, R. Mazzocchino, C. Scarpini, “Changes in Ia reciprocal inhibition<br />

from peroneal nerve to the soleus α-motoneurons with different static body positions in<br />

man”, Neuroscience Letters, vol. 84, pp. 283–286, 1988.<br />

[Rossignol 06] S. Rossignol, R. Dubuc, J. P. Gossard,“Dynamic Sensorimotor Interactions<br />

in Locomotion”, Physiological Reviews, vol. 86, pp. 89–154, 2006.<br />

[Roth 08] M. Roth,“Entwicklung und Konstruktion einer zweibeinigen Laufmaschine -<br />

Teilbereich: Antrieb”, Project thesis, University <strong>of</strong> Kaiserslautern, 2008. unpublished.<br />

[Sabourin 05] C. Sabourin, O. Bruneau, G. Buche,“Experimental validation <strong>of</strong> a robust<br />

control strategy for the robot RABBIT”, in Proceedings <strong>of</strong> the IEEE Conference on<br />

Robotics and Automation (ICRA). 2005.<br />

[Sakagami 02] Y. Sakagami, R. Watanabe, C. Aoyama, S. Matsunaga, N. Higaki, K. Fujimura,“The<br />

intelligent ASIMO: system overview and integration”, in Proceedings <strong>of</strong> the<br />

IEEE/RSJ International Conference on Intelligent <strong>Robots</strong> and System (IROS). 2002,<br />

pp. 2478–2483.


216 Bibliography<br />

[Sardain 04] P. Sardain, G. Bessonnet,“Forces acting on a biped robot. Center <strong>of</strong> pressure<br />

- zero moment point”, IEEE Transactions on Systems, Man and Cybernetics, vol. 34, no.<br />

5, pp. 630–637, 2004.<br />

[Schäfer 05] H. Schäfer, M. Proetzsch, K. Berns,“Extension Approach for the Behaviour-<br />

Based <strong>Control</strong> System <strong>of</strong> the Outdoor Robot RAVON”, in Autonome Mobile Systeme.<br />

Stuttgart, Germany, December 8–9 2005, pp. 123–129.<br />

[Schmidt 06] D. Schmidt, T. Luksch, J. Wettach, K. Berns, “Autonomous Behavior-<br />

Based Exploration <strong>of</strong> Office Environments”, in Proceedings <strong>of</strong> the 3rd International<br />

Conference on Informatics in <strong>Control</strong>, Automation and Robotics (ICINCO). Setubal,<br />

Portugal, August 1–5 2006, pp. 235–240.<br />

[Schmitt 03] D. Schmitt,“Insights into the evolution <strong>of</strong> human bipedalism from experimental<br />

studies <strong>of</strong> humans and other primates”, The Journal <strong>of</strong> Experimental Biology,<br />

vol. 206, pp. 1437–1448, 2003.<br />

[Scholl 02] K. U. Scholl, J. Albiez, B. Gassmann, J. M. Zöllner,“MCA - Modular <strong>Control</strong>ler<br />

Architecture”, in Robotik 2002, VDI-Bericht 1679. 2002.<br />

[Schumann 08] J. Schumann,“Entwicklung und Konstruktion einer zweibeinigen Laufmaschine<br />

- Teilbereich: Struktur und Prüfstand”, Project thesis, University <strong>of</strong> Kaiserslautern,<br />

2008. unpublished.<br />

[Sellaouti 06] R. Sellaouti, O. Stasse, S. Kajita, K. Yokoi, A. Kheddar, “Faster and<br />

Smoother <strong>Walking</strong> <strong>of</strong> <strong>Human</strong>oid HRP-2 with Passive Toe Joints”, in Proceedings <strong>of</strong><br />

IEEE/RSJ International Conference on Intelligent <strong>Robots</strong> and Systems (IROS). 2006.<br />

[Seyfarth 01] A. Seyfarth, M. Guenther, R. Blickhan, “Stable Operation <strong>of</strong> an Elastic<br />

Three-Segmented Leg”, Biological Cybernetics, vol. 84, pp. 364–382, 2001.<br />

[Shan 02] J.Shan, F.Nagashima,“Neurallocomotioncontrollerdesignandimplementation<br />

for humanoid robot HOAP-1”, in Proceedings <strong>of</strong> 20th Annual Conference <strong>of</strong> the Robotics<br />

Society <strong>of</strong> Japan, 2002.<br />

[Sherrington 06] C. S. Sherrington, The Integrative Action <strong>of</strong> the Nervous System, Oxford<br />

University Press, 1906.<br />

[Smith 07] J. A. Smith, A. Seyfarth, “Exploring Toe <strong>Walking</strong> in a <strong>Bipedal</strong> Robot”, in<br />

Autonome Mobile Systeme (AMS). 2007, pp. 287–293.<br />

[Soya 06] T. Soya, “Stage <strong>of</strong> Partner Robot Development in Toyota and Feasibility <strong>of</strong><br />

Applications in the Medical Field”, Circulation Journal, vol. 70, 2006.<br />

[Steiner 08] M. Steiner,“Reactive Balance and Posture <strong>Control</strong> for a Biped during Standing”,<br />

Project thesis, University <strong>of</strong> Kaiserslautern, September 29 2008. unpublished.<br />

[Steiner 09] M.Steiner,“IntegratingReinforcementLearningintoBehaviour-Based<strong>Control</strong><br />

<strong>of</strong> <strong>Bipedal</strong> <strong>Robots</strong>”, Diploma thesis, University <strong>of</strong> Kaiserslautern, 2009. unpublished.


Bibliography 217<br />

[Taga 91] G. Taga, Y. Yamaguchi, H. Shimizu,“Self-Organized <strong>Control</strong> <strong>of</strong> <strong>Bipedal</strong> Locomotion<br />

by Neural Oscillators in Unpredictable Environment”, Biological Cybernetics,<br />

vol. 65, pp. 147–159, 1991.<br />

[Taga 95a] G. Taga,“A Model <strong>of</strong> the Neuro-Musculo-Skeletal System for <strong>Human</strong> Locomotion<br />

I: Emergence <strong>of</strong> Basic Gait”, Biological Cybernetics, vol. 73, no. 2, pp. 97–111,<br />

1995.<br />

[Taga 95b] G. Taga,“A Model <strong>of</strong> the Neuro-Musculo-Skeletal System for <strong>Human</strong> Locomotion<br />

II: Real-Time Adaptability under Various Constraints”, Biological Cybernetics, vol.<br />

73, no. 2, pp. 113–121, 1995.<br />

[Takenaka 09a] T. Takenaka, T. Matsumoto, T. Yoshiike,“Real Time Motion Generation<br />

and <strong>Control</strong> for Biped Robot - 1st Report: <strong>Walking</strong> Gait Pattern Generation”, in<br />

Proceedings <strong>of</strong> IEEE/RSJ International Conference on Intelligent <strong>Robots</strong> and Systems<br />

(IROS). 2009, pp. 1084–1091.<br />

[Takenaka 09b] T. Takenaka, T. Matsumoto, T. Yoshiike, S. Shirokura,“Real Time Motion<br />

Generation and <strong>Control</strong> for Biped Robot - 2nd Report: Running Gait Pattern<br />

Generation”, in Proceedings <strong>of</strong> IEEE/RSJ International Conference on Intelligent <strong>Robots</strong><br />

and Systems (IROS). IEEE Press, 2009, pp. 1092–1099.<br />

[Ting 07] L. H. Ting, Progress in Brain Research, Elsevier B.V., 2007, vol. 165, ch. 19.<br />

Dimensional reduction in sensorimotor systems: a framework for understanding muscle<br />

coordination <strong>of</strong> posture, pp. 1211–1221.<br />

[Torres 96] A.L.Torres.“Implementation<strong>of</strong>VirtualModel<strong>Control</strong>ona<strong>Walking</strong>Hexapod”.<br />

Undergraduate Thesis, Massachusetts Institute <strong>of</strong> Technology, May 1996.<br />

[Tsuchiya 03] K. Tsuchiya, S. Aoi, K. Tsujita,“Locomotion control <strong>of</strong> a biped locomotion<br />

robot using nonlinear oscillators”, in Proceedings <strong>of</strong> the IEEE/RSJ International<br />

Conference on Intelligent <strong>Robots</strong> and Systems (IROS). 2003.<br />

[Vanderborght 06] B. Vanderborght, B. Verrelst, R. V. Ham, M. V. Damme, P. Beyl,<br />

D. Lefeber,“Torque and Compliance <strong>Control</strong> <strong>of</strong> the Pneumatic Artificial Muscles in the<br />

Biped ”Lucy””, in Proceedings <strong>of</strong> the 2006 IEEE International Conference on Robotics<br />

and Automation (ICRA). Orland, FL, USA, May 2006, pp. 842–847.<br />

[Vanderborght 07] B. Vanderborght,“Dynamic stabilisation <strong>of</strong> the biped Lucy powered by<br />

actuators with controllable stiffness”, Dissertation, Vrije Universiteit Brussel, May 2007.<br />

[Vanderborght 08] B. Vanderborght, B. Verrelst, R. V. Ham, M. V. Damme, P. Beyl,<br />

D. Lefeber,“Development <strong>of</strong> a compliance controller to reduce energy consumption for<br />

bipedal robots”, Autonomous <strong>Robots</strong>, vol. 24, no. 4, pp. 419–434, 2008.<br />

[Vaughan 92] C. L. Vaughan, B. L. Davis, J. C. O´Connor, Dynamics <strong>of</strong> human gait,<br />

Champaign, III.: <strong>Human</strong> Kinetics Publishers, 1992.<br />

[Vermeulen 04] J. Vermeulen, D. Lefeber, B. Verrelst, B. Vanderborght,“Trajectory Planning<br />

for the <strong>Walking</strong> Biped LUCY”, in Proceedings <strong>of</strong> the 7th International Conference<br />

on Climbing and <strong>Walking</strong> <strong>Robots</strong> (CLAWAR). 2004.


218 Bibliography<br />

[Verrelst 05] B. Verrelst, R. Van Ham, B. Vanderborght, F. Daerden, D. Lefeber,“The<br />

Pneumatic Biped ”LUCY”Actuated with Pleated Pneumatic Artificial Muscles”, Autonomous<br />

<strong>Robots</strong>, vol. 18, pp. 201–213, 2005.<br />

[Vukobratovic 04] M. Vukobratovic, B. Borovac,“Zero-Moment Point - Thirty Five Years<br />

<strong>of</strong> its Life”, International Journal <strong>of</strong> <strong>Human</strong>oid Robotics, vol. 1, pp. 157–173, March<br />

2004.<br />

[Vukobratovic 72] M. Vukobratovic, Y. Stepanenko,“On the Stability <strong>of</strong> Anthropomorphic<br />

Systems”, Mathematical Biosciences, vol. 15, pp. 1–37, 1972.<br />

[Wahl 09] T. Wahl, S. Blank, T. Luksch, K. Berns,“Biological Motivated <strong>Control</strong> Architecture<br />

and Mechatronics for a <strong>Human</strong><strong>like</strong> Robot”, in Advances in Robotics Research<br />

- Theory, Implementation, Application, T. Kroeger, F. M. Wahl, Eds. Braunschweig<br />

Germany: Springer, June 2009, pp. pp. 179–190.<br />

[Wannier 01] T. Wannier, C. Bastiaanse, G. Colombo, V. Dietz,“Arm to leg coordination<br />

in humans during walking, creeping and swimming activities”, Experimental Brain<br />

Research, vol. 141, no. 3, pp. 375–379, 2001.<br />

[Wieber 06] P. B. Wieber, C. Chevallereau,“Online adaptation <strong>of</strong> reference trajectories<br />

for the control <strong>of</strong> walking systems”, Robotics and Autonomous Systems, vol. 54, no. 7,<br />

pp. 559–566, 2006.<br />

[Winter 01] D. A. Winter, A. E. Patla, S. Rietdyk, M. G. Ishac,“Ankle muscle stiffness in<br />

the control <strong>of</strong> balance during quiet standing”, Journal <strong>of</strong> Neurophysiology, vol. 85, no. 6,<br />

pp. 2630–2633, 2001.<br />

[Winter 91] D. A. Winter, Biomechanics and Motor <strong>Control</strong> <strong>of</strong> <strong>Human</strong> Gait: Normal,<br />

Elderly and Pathological, Waterloo Biomechanics, 1991.<br />

[Winter 92] D. A. Winter,“Foot trajectory in human gait: a precise and multifactorial<br />

motor control task”, Physical Therapy, vol. 72, no. 1, pp. 45–53, 1992.<br />

[Winter 95] D. A. Winter, “<strong>Human</strong> balance and posture control during standing and<br />

walking”, Gait & Posture, vol. 3, no. 4, pp. 193–214, 1995.<br />

[Winter 98] D. A. Winter, A. E. Patla, F. Prince, M. Ishac, K. Gielo-Perczak,“Stiffness<br />

control <strong>of</strong> balance in quiet standing”, Journal <strong>of</strong> Neurophysiology, vol. 80, no. 3, pp.<br />

1211–1221, 1998.<br />

[Wisse 04] M. Wisse,“Essentials <strong>of</strong> Dynamic <strong>Walking</strong>, Analysis and Design <strong>of</strong> Two-Legged<br />

<strong>Robots</strong>”, Dissertation, TU Delft, 2004.<br />

[Wisse 05] M. Wisse,“Three additions to passive dynamic walking: actuation, an upper<br />

body, and 3D stability”, International Journal <strong>of</strong> <strong>Human</strong>oid Robotics, vol. 2, no. 4, pp.<br />

459–478, 2005.<br />

[Witte 04] H. Witte, H. H<strong>of</strong>fmann, R. Hackert, C. Schilling, M. S. Fischer, H. Preusch<strong>of</strong>t,<br />

“Biomimetic Robotics should be based on Functional Morphology”, Journal <strong>of</strong> Anatomy,<br />

vol. 204, no. 5, pp. 331–342, 2004.


Bibliography 219<br />

[Witte 91] H. Witte, H. Preusch<strong>of</strong>t, S. Recknagel,“<strong>Human</strong> body proportions explained on<br />

the basis <strong>of</strong> biomechanical principles”, Zeitschrift für Morphologie und Anthropologie,<br />

vol. 78, no. 3, pp. 407–423, 1991.<br />

[Witte 97] H. Witte, S. Recknagel, J. G. Rao, M. Wüthrich, C. Lesch, “Is Elastic Energy<br />

Storage <strong>of</strong> Quantitative Relevance for the Functional Morphology <strong>of</strong> the <strong>Human</strong><br />

Locomotor Apparatus?”, Acta Anatomica, vol. 158, pp. 106–111, 1997.<br />

[Yamazaki 96] N. Yamazaki, K. Hase, N. Ogihara, N. Hayamizu,“Biomechanical Analysis<br />

<strong>of</strong> the Development <strong>of</strong> <strong>Human</strong> <strong>Bipedal</strong> <strong>Walking</strong> by a Neuro-Mosculo-Skeletal Model”,<br />

Folia Primatol, no. 66, pp. 253–271, 1996.<br />

[Zaier 07] R. Zaier, S. Kanda,“Piecewise-Linear Pattern Generator and Reflex System for<br />

<strong>Human</strong>oid <strong>Robots</strong>”, in Proceedings <strong>of</strong> the IEEE International Conference on Robotics<br />

and Automation (ICRA). Roma, Italy, April 10–14 2007, pp. 2188–2195.<br />

[Zaier 08] R. Zaier, S. Kanda, “Adaptive locomotion controller and reflex system for<br />

humanoid robots”, in Proceedings <strong>of</strong> IEEE/RSJ International Conference on Intelligent<br />

<strong>Robots</strong> and Systems (IROS). 2008, pp. 2492–2497.<br />

[Zehr 98] E. P. Zehr, R. B. Stein, T. Komiyama,“Function <strong>of</strong> sural nerve reflexes during<br />

human walking”, The Journal <strong>of</strong> Physiology, vol. 507, no. 1, pp. 305–314, 1998.<br />

[Zehr 99] E. P. Zehr, R. B. Stein,“What function do reflexes serve during human locomotion?”,<br />

Progess in Neurobiology, vol. 58, pp. 185–205, 1999.<br />

[Zhao 02] M. Zhao, L. Liu, J. Wang, K. Chen, J. Zhao, K. Xu,“<strong>Control</strong> system design <strong>of</strong><br />

THBIP-I humanoid robot”, in Proceedings <strong>of</strong> IEEE International Conference on Robotics<br />

and Automation (ICRA). 2002, pp. 2253–2258.


220 Bibliography

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

Saved successfully!

Ooh no, something went wrong!