Development of a Hill-Type Muscle Model With Fatigue for the ...

Development of a Hill-Type Muscle Model With Fatigue for the ... Development of a Hill-Type Muscle Model With Fatigue for the ...

dspace.ist.utl.pt
from dspace.ist.utl.pt More from this publisher
12.05.2013 Views

Development of a Hill-Type Muscle Model With Fatigue for the Calculation of the Redundant Muscle Forces using Multibody Dynamics ANDRÉ FERRO PEREIRA Dissertação para obtenção do Grau de Mestre em ENGENHARIA BIOMÉDICA Júri Presidente: Prof. Helder Carriço Rodrigues Orientador: Prof. Miguel Pedro Tavares da Silva Prof. Mamede de Carvalho Vogais: Prof. Jorge Manuel Mateus Martins Prof. João Nuno Marques Parracho Guerra da Costa Outubro 2009

<strong>Development</strong> <strong>of</strong> a <strong>Hill</strong>-<strong>Type</strong> <strong>Muscle</strong> <strong>Model</strong> <strong>With</strong> <strong>Fatigue</strong> <strong>for</strong><br />

<strong>the</strong> Calculation <strong>of</strong> <strong>the</strong> Redundant <strong>Muscle</strong> Forces using<br />

Multibody Dynamics<br />

ANDRÉ FERRO PEREIRA<br />

Dissertação para obtenção do Grau de Mestre em<br />

ENGENHARIA BIOMÉDICA<br />

Júri<br />

Presidente: Pr<strong>of</strong>. Helder Carriço Rodrigues<br />

Orientador: Pr<strong>of</strong>. Miguel Pedro Tavares da Silva<br />

Pr<strong>of</strong>. Mamede de Carvalho<br />

Vogais: Pr<strong>of</strong>. Jorge Manuel Mateus Martins<br />

Pr<strong>of</strong>. João Nuno Marques Parracho Guerra da Costa<br />

Outubro 2009


Resumo<br />

O objectivo deste trabalho é o desenvolvimento de um modelo muscular<br />

versátil e a sua implementação de <strong>for</strong>ma robusta e eficiente num código de<br />

dinâmica de sistemas multicorpo com coordenadas naturais. São considera-<br />

dos dois tipos de modelos: o primeiro é um modelo muscular do tipo <strong>Hill</strong> que<br />

simula o comportamento das estruturas contrácteis, tanto para análises em<br />

dinâmica directa como inversa. O segundo é um modelo dinâmico de fadiga<br />

muscular que toma em consideração o historial de cada músculo, em termos<br />

de <strong>for</strong>ça produzida, estimando o seu nível de aptidão física usando um mod-<br />

elo multi-compartimentar triplo e a hierarquia de recrutamento muscular.<br />

A <strong>for</strong>mulação para as equação do movimento é adaptada, de <strong>for</strong>ma a incluir<br />

os modelos descritos, através do método de Newton. Isto permitirá, numa<br />

perspectiva de dinâmica directa, que se proceda ao cálculo da cinemática do<br />

sistema mecânico resultante para determinadas activações musculares previ-<br />

amente conhecidas, ou, numa perspectiva de dinâmica inversa, a computação<br />

das activações musculares necessárias para provocarem um movimento artic-<br />

ular prescrito. Para ambas estas <strong>for</strong>mulações, os sistemas são redundantes,<br />

uma propriedade que é ultrapassada no segundo caso usando um algoritmo<br />

de optimização.<br />

As metodologias e modelos são aplicados para vários casos de estudo, de<br />

<strong>for</strong>ma a avaliar a sua robustez e precisão. Um modelo da extremidade su-<br />

perior com sete músculos é considerado para evidenciar a aplicabilidade de<br />

um modelo de fadiga muscular num sistema multicorpo. Um segundo mod-<br />

elo, que inclui um aparato músculo-esquelético da extremidade inferior com<br />

doze músculos, é utilizado com único propósito de calcular as activações<br />

musculares. Os resultados associados a estes modelos são apresentados bem<br />

como as respectivas conclusões. O trabalho conclui perspectivando eventuais<br />

desenvolvimentos futuros.


Palavras-chave: Dinâmica multicorpo, dinâmica de contracção muscular,<br />

dinâmica de fadiga muscular, <strong>for</strong>ças musculares, activações musculares, op-<br />

timização.


Abstract<br />

The aim <strong>of</strong> this work is to develop a versatile muscle model and robustly<br />

implement it in an existent multibody system dynamics code with natural<br />

coordinates. Two different models are included: <strong>the</strong> first is a <strong>Hill</strong>-type muscle<br />

model that simulates <strong>the</strong> functioning <strong>of</strong> <strong>the</strong> contractile structures both in<br />

<strong>for</strong>ward and in inverse dynamic analysis; <strong>the</strong> second is a dynamic muscular<br />

fatigue model that considers <strong>the</strong> <strong>for</strong>ce production history <strong>of</strong> each muscle and<br />

estimates its fitness level using a three-compartment <strong>the</strong>ory approach and a<br />

physiological muscle recruitment hierarchy.<br />

The existent equations <strong>of</strong> motion <strong>for</strong>mulation is rearranged to include <strong>the</strong><br />

referred models using <strong>the</strong> Newton’s method approach. This allows, in a<br />

<strong>for</strong>ward dynamics perspective, <strong>for</strong> <strong>the</strong> calculation <strong>of</strong> <strong>the</strong> system’s motion<br />

that results from a pattern <strong>of</strong> given muscle activations, or, in an inverse<br />

dynamics perspective, <strong>for</strong> <strong>the</strong> computation <strong>of</strong> <strong>the</strong> muscle activations that are<br />

required to produce a prescribed articular movement. In both perspectives<br />

<strong>the</strong> system presents a redundant nature that is overcome in <strong>the</strong> latter case<br />

using an SQP optimization algorithm.<br />

The methodologies and models are applied to several case studies to eval-<br />

uate <strong>the</strong>ir robustness and accuracy. An upper extremity model with seven<br />

muscles is designed to evidence <strong>the</strong> effectiveness <strong>of</strong> <strong>the</strong> implementation <strong>of</strong><br />

a muscle fatigue model in a multibody system. A second model, encom-<br />

passing <strong>the</strong> lower extremity musculoskeletal apparatus with twelve muscles,<br />

is proposed <strong>for</strong> <strong>the</strong> exclusive calculation <strong>of</strong> muscle activations. The results<br />

are presented and conclusions are discussed. The work concludes with a<br />

perspective <strong>of</strong> possible future developments.


Keywords: Multibody dynamics, muscle contraction dynamics, muscle fa-<br />

tigue dynamics, muscle <strong>for</strong>ces, muscle activations, optimization.


Acknowledgements<br />

I would like to start by expressing my deepest gratitude to my supervisor Dr.<br />

Miguel Silva, to whom I owe <strong>for</strong> his inspiration, knowledge, encouragement<br />

and patience. This <strong>the</strong>sis would not have been possible without his wisdom<br />

and total dedication. The uncountable meetings and brainstorming sessions<br />

were <strong>the</strong> core <strong>of</strong> this project and are definitely <strong>the</strong> highlights <strong>of</strong> my academic<br />

experience, so far.<br />

To Pr<strong>of</strong>. Dr. Mamede de Carvalho and Dr. João Costa, <strong>for</strong> providing <strong>the</strong>ir<br />

highly motivational medical feedback and points-<strong>of</strong>-view.<br />

A big thanks to Dr. Jorge Martins and to my colleagues Rita Malcata,<br />

Pedro Moreira and Daniel Lopes <strong>for</strong> <strong>the</strong>ir help in <strong>the</strong> whole project process.<br />

To Dr. Marko Ackermann, Dr. Maury Hull, Dr. Maxime Raison and Dr.<br />

Ting Xia <strong>for</strong> <strong>the</strong>ir kindness in providing <strong>the</strong>ir papers and work.<br />

To <strong>the</strong> University <strong>of</strong> Washington <strong>for</strong> providing <strong>the</strong> Musculoskeletal Images<br />

from <strong>the</strong> "Musculoskeletal Atlas: A Musculoskeletal Atlas <strong>of</strong> <strong>the</strong> Human<br />

Body" by Carol Teitz, M.D. and Dan Graney, Ph.D.<br />

To all <strong>of</strong> my friends that supported me throughout <strong>the</strong>se University years.<br />

Never<strong>the</strong>less, some <strong>of</strong> <strong>the</strong>m should be mentioned, due to <strong>the</strong>ir true friendship<br />

and com<strong>for</strong>t: Nadir Abu-Samra, Diogo Almeida, Ana Barradinhas, André<br />

Bento, João Cabaça, Luís Cabecinha, Akshay Chaudry, Fábio Coelho, João<br />

Fayad, Artur Ferreira, Daniel Fitas, Diogo Geraldes, Maria João Lascas,<br />

Gonçalo Marcelo, Daniel Martins, André Medeiros, Diana Nunes, João Maia<br />

de Oliveira, Susana Palma, Manuel Rosa, Rafael Rosário, João Lála dos<br />

Santos, Ricardo Serrano, César Silveira, João Venes.<br />

To all my family, namely to my parents, my sister, my grandmo<strong>the</strong>r, cousins<br />

and Maria.


The deepest acknowledgment goes to my mo<strong>the</strong>r. I thank her <strong>for</strong> being <strong>the</strong><br />

person who always was <strong>the</strong>re <strong>for</strong> me, who motivated me all <strong>the</strong> way through,<br />

and granted that I was guided by <strong>the</strong> same values <strong>of</strong> ambition and drive that<br />

I recognise in herself. Indeed, this work is dedicated to her.


Para a minha mãe, Maria Isabel.<br />

To my mo<strong>the</strong>r, Maria Isabel.


Contents<br />

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

List <strong>of</strong> Tables xi<br />

List <strong>of</strong> Symbols xiii<br />

Glossary xv<br />

1 Introduction 1<br />

1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1<br />

1.2 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2<br />

1.3 Literature review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3<br />

1.4 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7<br />

1.5 Thesis Organization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

2 Musculoskeletal System <strong>Model</strong>ling 11<br />

2.1 Skeletal <strong>Muscle</strong> Anatomy and Physiology . . . . . . . . . . . . . . . . . 12<br />

2.1.1 Skeletal <strong>Muscle</strong> Anatomy . . . . . . . . . . . . . . . . . . . . . . 12<br />

2.1.2 Skeletal <strong>Muscle</strong> Physiology . . . . . . . . . . . . . . . . . . . . . 18<br />

2.2 Dynamics <strong>of</strong> <strong>the</strong> <strong>Muscle</strong> Tissue . . . . . . . . . . . . . . . . . . . . . . . 20<br />

2.2.1 Activation Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

2.2.2 Contraction Dynamics . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

2.2.3 <strong>Muscle</strong> <strong>Fatigue</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

2.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37<br />

v


CONTENTS<br />

3 Multibody Dynamics 39<br />

3.1 Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />

3.2 Equations <strong>of</strong> motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43<br />

3.3 Generic muscle <strong>for</strong>ces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

3.4 Inverse Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50<br />

3.5 Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54<br />

3.6 Forward Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58<br />

3.7 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60<br />

4 Biomechanical <strong>Model</strong>s 63<br />

4.1 <strong>Muscle</strong> model verification . . . . . . . . . . . . . . . . . . . . . . . . . . 63<br />

4.2 <strong>Muscle</strong> <strong>Fatigue</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

4.3 Human Gait . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74<br />

4.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

5 Conclusions and Future <strong>Development</strong>s 93<br />

5.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93<br />

5.2 Future <strong>Development</strong>s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95<br />

A Apollo – <strong>Hill</strong>-type muscles manual 99<br />

A.1 MDL File . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

A.2 Simulation file . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101<br />

B MHILL Data Visualizer manual 103<br />

References 109<br />

vi


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

2.1 Arrangement <strong>of</strong> <strong>the</strong> skeletal muscle structure, from an external level to<br />

a molecular level [53]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13<br />

2.2 Representation <strong>of</strong> <strong>the</strong> penation angle α between muscle fibers and ten-<br />

dons [42]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

2.3 Detail <strong>of</strong> a myosin molecule (A) and an actin filament (B) [53]. . . . . . 15<br />

2.4 Illustration <strong>of</strong> <strong>the</strong> cross-bridges <strong>for</strong>med by <strong>the</strong> connections <strong>of</strong> actin fila-<br />

ments with a myosin filament. Adapted from Reference [53]. . . . . . . . 15<br />

2.5 Detail <strong>of</strong> a muscle fiber [62]. . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

2.6 Neural muscular junction with three different detail levels [62]. . . . . . 17<br />

2.7 Contraction regulation mechanism by <strong>the</strong> troponin-tropomyosin complex,<br />

dependent on <strong>the</strong> concentration level <strong>of</strong> Ca 2+ [62]. . . . . . . . . . . . . 18<br />

2.8 Crossbridge <strong>the</strong>ory steps [62]. . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

2.9 Relation between twitch frequency and effective muscle contraction [1]. . 21<br />

2.10 Scheme <strong>of</strong> muscle tissue dynamics, with a series model <strong>of</strong> Activation<br />

Dynamics and Contraction Dynamics. . . . . . . . . . . . . . . . . . . . 21<br />

2.11 Activation dynamics model consistency [20]. . . . . . . . . . . . . . . . . 23<br />

2.12 <strong>Hill</strong>-type muscle models. . . . . . . . . . . . . . . . . . . . . . . . . . . . 24<br />

2.13 Discrete length-tension diagram <strong>of</strong> <strong>the</strong> contractile contribution <strong>of</strong> a single<br />

fully activated sarcomere [53] (a) and <strong>the</strong> same relationship scaled to <strong>the</strong><br />

whole muscle [1] (b). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

2.14 Force-length relationship <strong>of</strong> <strong>the</strong> passive element [1]. . . . . . . . . . . . . 27<br />

2.15 Force-velocity relationship [18]. . . . . . . . . . . . . . . . . . . . . . . . 28<br />

2.16 Activation scaling evidence: Force-length and <strong>for</strong>ce-velocity relationships<br />

<strong>for</strong> different levels <strong>of</strong> muscle activation a(t) [18]. . . . . . . . . . . . . . . 28<br />

vii


LIST OF FIGURES<br />

2.17 Block diagram <strong>of</strong> <strong>the</strong> muscle contraction dynamics model implemented<br />

in <strong>the</strong> <strong>for</strong>ward dynamics multibody system routines. . . . . . . . . . . . 30<br />

2.18 Block diagram <strong>of</strong> <strong>the</strong> muscle contraction dynamics model implemented<br />

in <strong>the</strong> inverse dynamics multibody system routines. . . . . . . . . . . . . 30<br />

2.19 Curve based on Rohmert’s relationship between %MVC (percentage <strong>of</strong><br />

maximum voluntary contraction) and endurance time (minutes) [56]. . . 32<br />

2.20 Three-compartment <strong>the</strong>ory flowchart. . . . . . . . . . . . . . . . . . . . . 33<br />

2.21 <strong>Muscle</strong> recruitment hierarchy pile chart [59]. . . . . . . . . . . . . . . . . 36<br />

3.1 <strong>Muscle</strong> representation <strong>for</strong> a biomechanical system. . . . . . . . . . . . . 45<br />

3.2 Application <strong>of</strong> a <strong>for</strong>ce F m to pointo p, belonging to <strong>the</strong> rigid body defined<br />

by points i and j and vectors u and v [1]. . . . . . . . . . . . . . . . . . 46<br />

3.3 The different representations <strong>of</strong> a generic muscle <strong>for</strong>ce F m p . . . . . . . . . 48<br />

3.4 <strong>Muscle</strong> <strong>for</strong>ce representation <strong>for</strong> <strong>the</strong> 3 via-point model in Figure 3.1. . . . 50<br />

3.5 Used optimization methodology. . . . . . . . . . . . . . . . . . . . . . . . 56<br />

3.6 Direct integration algorithm flowchart <strong>for</strong> a <strong>for</strong>ward dynamics problem [1]. 60<br />

4.1 Mechanical system with muscles m X and m Y <strong>for</strong> <strong>Hill</strong>-type muscle model<br />

verification purposes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

4.2 Activations pattern <strong>for</strong> muscles m X and m Y . . . . . . . . . . . . . . . . 65<br />

4.3 The state <strong>of</strong> <strong>the</strong> mechanical model <strong>for</strong> different time instants when mus-<br />

cles m X and m Y are activated as in Figure 4.2. . . . . . . . . . . . . . . 66<br />

4.4 <strong>Muscle</strong> m X contractile element <strong>for</strong>ce output (black line) and possible<br />

<strong>for</strong>ces <strong>for</strong> <strong>the</strong> model (coloured surface). . . . . . . . . . . . . . . . . . . . 67<br />

4.5 Results obtained <strong>for</strong> three cycles <strong>of</strong> isometric contractions <strong>of</strong> <strong>the</strong> described<br />

model. The quantities MA + MR, MF and a(t) are scaled to F0. . . . . . 69<br />

4.6 Simple upper extremity model with 7 elbow muscles and constant <strong>for</strong>ce<br />

P applied at <strong>the</strong> hand. This image was conceived using <strong>the</strong> OpenSim<br />

s<strong>of</strong>tware [71]. Local reference frames only indicated in <strong>the</strong> lateral view. . 70<br />

4.7 Resultant muscle <strong>for</strong>ces <strong>of</strong> <strong>the</strong> contractile element F m CE<br />

<strong>of</strong> <strong>the</strong> biomechan-<br />

ical model <strong>of</strong> <strong>the</strong> upper extremity, when susceptible to fatigue dynamics. 71<br />

4.8 Simple leg model with 12 muscles spanning <strong>the</strong> knee and ankle joints.<br />

This image conceived using <strong>the</strong> OpenSim s<strong>of</strong>tware [71]. Local reference<br />

frames only indicated in <strong>the</strong> lateral view. . . . . . . . . . . . . . . . . . . 75<br />

viii


LIST OF FIGURES<br />

4.9 Scheme with relative progress <strong>of</strong> <strong>the</strong> gait cycle, indicating <strong>the</strong> stance and<br />

swing phases [74]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80<br />

4.10 Part <strong>of</strong> <strong>the</strong> gait cycle with four points <strong>of</strong> reference [75]. . . . . . . . . . . 81<br />

4.11 Point numbering used in <strong>the</strong> model: Point 1 - Lower torso, point 2 - hip<br />

joint, point 3 - knee joint, point 4 - ankle joint, point 5 - metatarsopha-<br />

langeal joint, 6 - heel position reference [75]. . . . . . . . . . . . . . . . . 81<br />

4.12 Driver 1: Trajectory driver with <strong>the</strong> space coordinates in time <strong>of</strong> point 1. 82<br />

4.13 Angular directions <strong>of</strong> Drivers 3, 4 and 5. . . . . . . . . . . . . . . . . . . 82<br />

4.14 The motion obtained using <strong>the</strong> prescribed kinematic drivers <strong>for</strong> <strong>the</strong> model<br />

<strong>of</strong> <strong>the</strong> lower extremity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83<br />

4.15 Reaction <strong>for</strong>ce vector components, acquired from a <strong>for</strong>ce plat<strong>for</strong>m. . . . 84<br />

4.16 Ground reaction <strong>for</strong>ce application point coordinates, or centre <strong>of</strong> pressure<br />

curve. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84<br />

4.17 <strong>Muscle</strong> activation patterns obtained from <strong>the</strong> solution <strong>of</strong> <strong>the</strong> EOM. . . . 86<br />

4.18 Resultant total muscle <strong>for</strong>ces <strong>of</strong> each considered muscle group in <strong>the</strong> lower<br />

leg extremity model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

4.19 Comparison between <strong>the</strong> calculated total <strong>for</strong>ce <strong>of</strong> <strong>the</strong> tibialis anterior<br />

muscle considering (Physiological case) and not considering (Non-physiological<br />

case) a contraction dynamics model. . . . . . . . . . . . . . . . . . . . . 88<br />

4.20 Comparison between <strong>the</strong> calculated activation <strong>of</strong> <strong>the</strong> tibialis anterior<br />

muscle considering (Physiological case) and not considering (Non-physiological<br />

case) a contraction dynamics model. . . . . . . . . . . . . . . . . . . . . 89<br />

B.1 General aspect <strong>of</strong> <strong>the</strong> MHILL Data Visualizer. . . . . . . . . . . . . . . . 103<br />

B.2 Operation boxes <strong>of</strong> <strong>the</strong> MHILL Data Visualizer. . . . . . . . . . . . . . . 104<br />

B.3 Files addition <strong>for</strong> <strong>the</strong> folder system. . . . . . . . . . . . . . . . . . . . . . 105<br />

B.4 Example <strong>of</strong> data plots. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105<br />

B.5 Selection <strong>of</strong> muscle data to be plotted. . . . . . . . . . . . . . . . . . . . 106<br />

B.6 Plotting <strong>of</strong> muscle activations. . . . . . . . . . . . . . . . . . . . . . . . . 106<br />

B.7 Removing some <strong>of</strong> <strong>the</strong> included files. . . . . . . . . . . . . . . . . . . . . 107<br />

B.8 Time instant selection. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107<br />

ix


LIST OF FIGURES<br />

x


List <strong>of</strong> Tables<br />

4.1 <strong>Fatigue</strong> parameters used <strong>for</strong> <strong>the</strong> muscle fatigue model employed, and <strong>the</strong><br />

same used in <strong>the</strong> work by Xia and Law [59]. The values <strong>for</strong> <strong>the</strong> F , R,<br />

LD and LR are given in units <strong>of</strong> 1/s. . . . . . . . . . . . . . . . . . . . . 68<br />

4.2 Body index number and local coordinates <strong>of</strong> <strong>the</strong> points that define <strong>the</strong><br />

rigid bodies <strong>for</strong> <strong>the</strong> upper extremity model. . . . . . . . . . . . . . . . . 70<br />

4.3 Inertial data <strong>for</strong> rigid bodies considered in <strong>the</strong> upper extremity model. . 71<br />

4.4 Geometrical and physiological parameters <strong>for</strong> <strong>the</strong> considered muscle in<br />

<strong>the</strong> upper extremity model. Considered values were taken from <strong>the</strong> work<br />

by Holzbaur et. al. [72]. Pictures reproduced with permission: "Copy-<br />

right 2003-2004 University <strong>of</strong> Washington. All rights reserved including<br />

all photographs and images. No re-use, re-distribution or commercial<br />

use without prior written permission <strong>of</strong> <strong>the</strong> authors and <strong>the</strong> University<br />

<strong>of</strong> Washington." [73]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

4.5 <strong>Fatigue</strong> parameters used <strong>for</strong> all <strong>the</strong> muscles in <strong>the</strong> upper extremity with<br />

elbow muscles model. These have no correlation with experimental re-<br />

sults [59]. Values given in units <strong>of</strong> 1/s. . . . . . . . . . . . . . . . . . . . 73<br />

4.6 Body index number bn and local coordinates <strong>of</strong> <strong>the</strong> points that defined<br />

each rigid body considered on <strong>the</strong> lower extremity model. . . . . . . . . 74<br />

4.7 Geometrical and physiological properties inherent to <strong>the</strong> muscles existent<br />

in <strong>the</strong> lower extremity model. Source: Reference [1]. . . . . . . . . . . . 76<br />

4.8 Anthropometrical properties associated with <strong>the</strong> rigid bodies <strong>of</strong> <strong>the</strong> lower<br />

extremity model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79<br />

xi


LIST OF TABLES<br />

xii


List <strong>of</strong> Symbols<br />

u(t) Neural signal a(t) <strong>Muscle</strong> activation<br />

τrise Time constant <strong>of</strong> muscle activation<br />

rise <strong>of</strong> a(t)<br />

F m <strong>Muscle</strong> <strong>for</strong>ce F m CE<br />

τfall Time constant <strong>of</strong> muscle activation<br />

fall <strong>of</strong> a(t)<br />

<strong>Muscle</strong> contractile element <strong>for</strong>ce<br />

F m P E <strong>Muscle</strong> passive element <strong>for</strong>ce ˆ F m CE <strong>Muscle</strong> available contractile element<br />

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

F m 0 <strong>Muscle</strong> maximum isometric <strong>for</strong>ce F m L<br />

F m ˙ L<br />

<strong>Muscle</strong> contractile element <strong>for</strong>cevelocity<br />

relationship<br />

<strong>Muscle</strong> contractile element <strong>for</strong>celength<br />

relationship<br />

L m <strong>Muscle</strong> length<br />

L m 0 <strong>Muscle</strong> length change rate ˙ L m <strong>Muscle</strong> rate <strong>of</strong> length change<br />

˙v m <strong>Muscle</strong> speed ˙ L m 0 <strong>Muscle</strong> maximum contractile<br />

velocity<br />

F <strong>Muscle</strong> fatigue factor R <strong>Muscle</strong> recovery factor<br />

BE Brain ef<strong>for</strong>t MA <strong>Fatigue</strong> model activated MU<br />

compartment<br />

MF <strong>Fatigue</strong> model fatigued MU compartment<br />

RC <strong>Muscle</strong> residual capacity ˆ F f<br />

CE<br />

MR <strong>Fatigue</strong> model resting MU<br />

compartment<br />

<strong>Fatigue</strong>d muscle available contractile<br />

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

C(t) <strong>Fatigue</strong> model muscle activa- LD <strong>Muscle</strong> <strong>for</strong>ce development factor<br />

tion–deactivationtroller<br />

driving con-<br />

LR muscle <strong>for</strong>ce relaxation factor T L Target load<br />

xiii


LIST OF SYMBOLS<br />

q Vector <strong>of</strong> generalised coordinates ˙q Vector <strong>of</strong> generalised velocities<br />

¨q Vector <strong>of</strong> generalised accelerations Φ Kinematic constraints expressions<br />

Φq Jacobian matrix <strong>of</strong> Φ in order to q ν Right-hand-side <strong>of</strong> <strong>the</strong> velocity<br />

equation<br />

γ Right-hand-side <strong>of</strong> <strong>the</strong> acceleration<br />

equation<br />

P ∗<br />

Virtual power produced by <strong>the</strong> external<br />

<strong>for</strong>ces<br />

˙q ∗ Virtual velocity vector M Global mass matrix<br />

g Generalised <strong>for</strong>ce vector gΦ Internal constraint <strong>for</strong>ces vector<br />

λ Vector <strong>of</strong> Lagrange multipliers vp Number <strong>of</strong> via-points<br />

ud <strong>Muscle</strong> direction d F m p <strong>Muscle</strong> <strong>for</strong>ce cartesian vector<br />

representation<br />

oξηζ Rigid body local reference frame 0xyz Global reference frame<br />

rp Global coordinates <strong>of</strong> point p Cp Cartesian-generalised coordinates<br />

trans<strong>for</strong>mation matrix <strong>for</strong> point p<br />

g F e<br />

gF m<br />

CE<br />

ˆg F m<br />

CE<br />

Rigid body generalised representation<br />

<strong>of</strong> <strong>for</strong>ce F<br />

Whole system generalised representation<br />

<strong>of</strong> <strong>the</strong> contractile element<br />

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

Whole system generalised representation<br />

<strong>of</strong> <strong>the</strong> available contractile<br />

element <strong>for</strong>ce<br />

χ Set <strong>of</strong> generalised available contractile<br />

element <strong>for</strong>ces <strong>the</strong> considered<br />

muscles<br />

x Optimization problem control variables<br />

F0<br />

λ R<br />

Optimization problem cost function<br />

Lagrange multiplier without constraints<br />

in <strong>the</strong> optimization problem<br />

xiv<br />

g F Whole system generalised representation<br />

<strong>of</strong> <strong>for</strong>ce F<br />

gF m<br />

P E<br />

Whole system generalised representation<br />

<strong>of</strong> <strong>the</strong> passive element<br />

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

g ext Generalised <strong>for</strong>ces excluding muscle<br />

<strong>for</strong>ces<br />

a <strong>Muscle</strong> activations vector<br />

feq Optimization problem equality<br />

constraints<br />

λ ∗<br />

Lagrange multipliers associated<br />

with <strong>the</strong> kinematic drivers<br />

g Gravity <strong>for</strong>ce


Glossary<br />

ACh Acetylcholine<br />

AD Activation dynamics<br />

ADP Adenosine diphosphate<br />

ATP Adenosine triphosphate<br />

BIClong biceps brachii long head<br />

BICshort biceps brachii short head<br />

BRA brachialis<br />

BRD brachioradialis<br />

CE <strong>Muscle</strong> contraction model contractile element<br />

CNS Central nervous system<br />

DE <strong>Muscle</strong> contraction model damping element<br />

EMG Electromyograph<br />

EOM Equations <strong>of</strong> motion<br />

FD Forward dynamics<br />

HS Heel Strike<br />

ID Inverse dynamics<br />

MSO Modified Static optimization<br />

MU Motor units<br />

OHS Opposite Heel Strike<br />

OTO Opposite Toe Off<br />

PE <strong>Muscle</strong> contraction model passive element<br />

SE <strong>Muscle</strong> contraction model series elastic element<br />

TO Toe Off<br />

TRIlat triceps brachii lateral head<br />

TRIlong triceps brachii long head<br />

TRImed triceps brachii medial head<br />

xv


GLOSSARY<br />

xvi


Chapter 1<br />

Introduction<br />

1.1 Motivation<br />

The discipline <strong>of</strong> computational mechanics uses ma<strong>the</strong>matical models to process <strong>the</strong><br />

state <strong>of</strong> a system regulated by <strong>the</strong> mechanical laws <strong>of</strong> physics. These models are math-<br />

ematical <strong>for</strong>mulations that emulate <strong>the</strong> behaviour <strong>of</strong> a determined system, and can<br />

<strong>the</strong>re<strong>for</strong>e be rendered to simple or complex networks <strong>of</strong> computational programs.<br />

This way, a computational model simulation can predict <strong>the</strong> functioning <strong>of</strong> complex<br />

systems, only by <strong>for</strong>mulating its nature in a model. These simulations required an<br />

immense set <strong>of</strong> algebraic calculations, impracticably solved by human beings. The<br />

main advantage <strong>of</strong> an in<strong>for</strong>matic application comes from <strong>the</strong> rapid development that<br />

computer machines have exhibited in <strong>the</strong> last few decades. Nowadays, a simple personal<br />

computer is capable <strong>of</strong> calculating a series <strong>of</strong> processes regarding <strong>the</strong> phenomena <strong>of</strong> any<br />

computational science. Modern day computers not only have substantial capacity <strong>for</strong><br />

data processing, but also show a day-to-day improvement with respect to this potential.<br />

The field <strong>of</strong> biomechanics, i.e., <strong>the</strong> scientific domain that concerns to <strong>the</strong> mechanics<br />

<strong>of</strong> living systems, has additionally suffered a dramatic innovation in terms <strong>of</strong> its com-<br />

putational applications. Multibody dynamics systems is one <strong>of</strong> <strong>the</strong> mechanical <strong>for</strong>mula-<br />

tions that emerged in this domain with <strong>the</strong> recent progress <strong>of</strong> new computational tools.<br />

These systems model <strong>the</strong> dynamics <strong>of</strong> linked rigid (or flexible) bodies that experience<br />

large displacements <strong>of</strong> position and rotation, and are easily adjusted to biomechanical<br />

systems, with significance <strong>for</strong> <strong>the</strong> musculoskeletal complex <strong>of</strong> <strong>the</strong> human body.<br />

Multibody dynamics is governed by its equations <strong>of</strong> motion (EOM) that will dictate<br />

1


1. INTRODUCTION<br />

<strong>the</strong> kinematics and kinetic properties <strong>of</strong> a certain mechanical system. From a classic<br />

perspective, it is possible to solve <strong>the</strong>se equations to obtain ei<strong>the</strong>r [1]:<br />

• The dynamic response <strong>of</strong> <strong>the</strong> multibody system experiencing known external <strong>for</strong>ce<br />

solicitations;<br />

• Or <strong>the</strong> internal and external <strong>for</strong>ces that explain a given motion pattern.<br />

The multibody <strong>for</strong>mulation that is used to solve <strong>the</strong> first paradigm is called Forward<br />

Dynamics (FD), <strong>the</strong> latter is known as Inverse Dynamics (ID).<br />

In <strong>the</strong> human body, <strong>the</strong>se <strong>for</strong>mulation can be adapted to analyse <strong>the</strong> kinetics <strong>of</strong><br />

body articulations with <strong>the</strong> existence <strong>of</strong> bone stress, joint reactions or even muscle<br />

<strong>for</strong>ces. To implement such a template, some existent ma<strong>the</strong>matical models that simu-<br />

late <strong>the</strong> constitutive laws that rule <strong>the</strong> behaviour <strong>of</strong> muscle activation and contraction<br />

are practicably employable. These have <strong>the</strong> potential <strong>of</strong> calculating <strong>the</strong> <strong>for</strong>ce output<br />

<strong>of</strong> <strong>the</strong> contractile structures <strong>of</strong> a determined muscle, knowing some physiological and<br />

geometrical parameters. In <strong>the</strong> same way, o<strong>the</strong>r models predict <strong>the</strong> fatigue state <strong>of</strong> a<br />

muscle, <strong>for</strong> a given developed <strong>for</strong>ce history and can be included in a multibody system.<br />

Implementing such models, <strong>the</strong> researcher is able to use <strong>the</strong>m in all kinds <strong>of</strong> scientific<br />

areas that deal with <strong>the</strong> mechanical interactions <strong>of</strong> <strong>the</strong> human body. The applications<br />

<strong>of</strong> such a tool are undeniably appealing, regarding its competence <strong>for</strong> calculating, in a<br />

non-invasive manner, muscle ef<strong>for</strong>ts, neural motor activations or even to predict how<br />

physically exhausted is a person after an established activity. A wide range <strong>of</strong> areas<br />

benefit from <strong>the</strong>se kind <strong>of</strong> models, such as medicine, sports or ergonomics.<br />

1.2 Objectives<br />

The main goal <strong>of</strong> this <strong>the</strong>sis is to implement a muscle model that accounts <strong>for</strong> <strong>the</strong> con-<br />

traction and fatigue dynamics inherent to <strong>the</strong> muscle tissue and to adapt this model in<br />

<strong>the</strong> multibody system dynamics program APOLLO [1], in <strong>the</strong> interest <strong>of</strong> allowing <strong>the</strong><br />

inclusion <strong>of</strong> muscle actuators and incorporate its characteristics to <strong>the</strong> mechanical sys-<br />

tem. The final scheme must be capable <strong>of</strong> calculating ei<strong>the</strong>r <strong>the</strong> muscle activations <strong>for</strong> a<br />

given kinematics, or <strong>the</strong> dynamics response <strong>of</strong> <strong>the</strong> system when <strong>the</strong> muscle activations<br />

are prescribed, with <strong>the</strong> option <strong>of</strong> considering muscle fatigue. There<strong>for</strong>e, it is desired to<br />

use a versatile <strong>for</strong>mulation, that will allow <strong>the</strong> operation in both <strong>for</strong>ward and inverse<br />

2


1.3 Literature review<br />

dynamics paradigms. An important point <strong>of</strong> any inverse dynamics system that deals<br />

with redundant muscle <strong>for</strong>ces is <strong>the</strong> optimization process, that will select one <strong>of</strong> <strong>the</strong> in-<br />

finite available combinations <strong>of</strong> muscle <strong>for</strong>ces that justify a certain kinematic condition.<br />

In addition, a special regard is given to <strong>the</strong> implementation <strong>of</strong> muscle fatigue dynamics,<br />

as <strong>the</strong> implementation <strong>of</strong> a muscle fatigue model in multibody dynamics constitutes one<br />

<strong>of</strong> <strong>the</strong> main novelties <strong>of</strong> this work.<br />

1.3 Literature review<br />

<strong>Muscle</strong> <strong>for</strong>ces determination is a quintessential problem in <strong>the</strong> biomechanics field. Mus-<br />

cle synchronisation and <strong>the</strong> estimation <strong>of</strong> <strong>the</strong> internal <strong>for</strong>ces <strong>of</strong> a musculoskeletal sys-<br />

tem (internal loads on both bones and joints) are <strong>the</strong> main targets in this kind <strong>of</strong><br />

calculation [2]. Direct measurement methods have been widely applied, as in Fujie et.<br />

al. [3] and Nigg and Herzog [4], however <strong>the</strong>se have <strong>the</strong> evident downside <strong>of</strong> its inva-<br />

sive nature. Despite <strong>of</strong> <strong>the</strong>ir accuracy, <strong>the</strong>se techniques take no advantage <strong>of</strong> numerical<br />

methods whatsoever, lacking versatility and applicability. There<strong>for</strong>e, multibody system<br />

modelling became <strong>the</strong> widespread choice <strong>for</strong> representation <strong>of</strong> musculoskeletal systems.<br />

Nikravesh and Attia [5] and Schiehlen [6] studied <strong>the</strong> application <strong>of</strong> multibody dy-<br />

namics in general rigid body mechanics, describing its <strong>for</strong>mulation and equations <strong>of</strong><br />

motion. The implementation <strong>of</strong> <strong>the</strong>se methodologies in natural coordinates is distinctly<br />

detailed in <strong>the</strong> work <strong>of</strong> Jalón and Bayo [7] and Silva [1]. This type <strong>of</strong> <strong>for</strong>mulation has<br />

a major importance <strong>for</strong> medical applications in rehabilitation pros<strong>the</strong>tics [8, 9], gait<br />

disorders [2], sports [10], study <strong>of</strong> muscle diseases [11] and equipment ergonomics [12]<br />

where its application can be specifically adapted to a certain subject by adjustment <strong>of</strong><br />

several anthropometrical and physiological parameters.<br />

Multibody dynamics can be categorised in two important methodologies <strong>for</strong> solving<br />

<strong>the</strong> equations <strong>of</strong> motion: inverse and <strong>for</strong>ward dynamics. The inverse dynamics approach<br />

calculates both external and internal <strong>for</strong>ces based on <strong>the</strong> anthropometrical properties<br />

and <strong>the</strong> motion <strong>of</strong> a given biomechanical system. Several inverse dynamics studies were<br />

made uniquely concerning <strong>the</strong> calculation <strong>of</strong> joint torques <strong>for</strong> prescribed kinematics [13–<br />

15]. In a musculoskeletal system, <strong>the</strong>se external <strong>for</strong>ces may include muscle ef<strong>for</strong>ts. In <strong>the</strong><br />

human body, <strong>the</strong>re is an average number <strong>of</strong> 2.6 muscles per degree-<strong>of</strong>-freedom [16] which<br />

act in synergy, in co-contraction or in overactuation around <strong>the</strong> spanned joints [17], i.e.,<br />

3


1. INTRODUCTION<br />

different muscle activations can generate <strong>the</strong> same mechanical model motion or posture,<br />

leading to what is know as muscular redundancy [1]. From a multibody dynamics<br />

perspective, this is numerically explained by <strong>the</strong> fact that <strong>the</strong> number <strong>of</strong> unknowns in<br />

<strong>the</strong> equations <strong>of</strong> motions is larger than <strong>the</strong> number <strong>of</strong> available equations, resulting in<br />

an indeterminate system with an infinite set <strong>of</strong> possible solutions. Numerically this<br />

problem is tackled by optimization techniques [1, 18]. This way, <strong>the</strong> optimal solution<br />

should minimise a specific cost function while fulfilling <strong>the</strong> equations <strong>of</strong> motion <strong>for</strong><br />

a set <strong>of</strong> prescribed kinematics and kinetics [19]. This approach is known as static<br />

optimization and several studies have produced results regarding muscular ef<strong>for</strong>ts <strong>for</strong><br />

different physiological situations [1, 11, 17, 18, 20–27].<br />

Static optimization has associated an important drawback: it only processes instant<br />

frames and <strong>the</strong>re<strong>for</strong>e only takes into account instantaneous per<strong>for</strong>mance criteria, that<br />

do not simulate in <strong>the</strong> best way <strong>the</strong> behaviour <strong>of</strong> <strong>the</strong> central nervous system (CNS) when<br />

choosing a muscle activation set [2]. In response to this aspect some authors [2, 28–<br />

30] used what is know as dynamic optimization (a solution developed using <strong>for</strong>ward<br />

dynamics). This approach operates with a nodal point parametrisation <strong>of</strong> <strong>the</strong> neural<br />

excitation, followed by several numerical integrations <strong>of</strong> <strong>the</strong> equations <strong>of</strong> motion [31].<br />

This option allows <strong>the</strong> correct implementation <strong>of</strong> activation dynamics, since it relies on<br />

multiple integrations, however it induces high computational times, which is its major<br />

drawback. Despite <strong>of</strong> modern day personal computers being able to solve simple me-<br />

chanical systems in less than 24 hours, <strong>the</strong>se times can reach several days or weeks <strong>for</strong><br />

complex systems [2, 18]. Several biomechanical studies have been made <strong>for</strong> different<br />

types <strong>of</strong> motion, such as normal gait [29], jumping [32] or cycling [33]. Anderson and<br />

Pandy [30] concluded that, <strong>for</strong> normal gait, <strong>the</strong> use <strong>of</strong> dynamic optimization in not jus-<br />

tified since <strong>the</strong> obtained results, when compared with <strong>the</strong> ones from static optimization,<br />

are practically <strong>the</strong> same. In <strong>the</strong> present work, only static optimization procedures were<br />

developed.<br />

Cost functions used in static optimization, as <strong>the</strong> ones proposed by Crowninshield<br />

and Brand [19], were widely used in classic works such as <strong>the</strong> ones by Collins [34],<br />

Glitsch and Baumann [35], Patriarco [36] and Yamaguchi [18, 23]. An excelent review<br />

regarding <strong>the</strong> various presented cost functions was made by Tsirakos et. al. [37]. Accord-<br />

ing to Ackermann [2], <strong>the</strong>se instantaneous functions are unable to accurately describe<br />

<strong>the</strong> key per<strong>for</strong>mance criterion: <strong>the</strong> total energy expenditure. There<strong>for</strong>e, two different<br />

4


1.3 Literature review<br />

approaches are introduced in reaction to <strong>the</strong> high computational times <strong>of</strong> dynamic op-<br />

timization: Extended inverse dynamics and Modified static optimization. For fur<strong>the</strong>r<br />

reading on <strong>the</strong>se methods please use reference [2].<br />

Bearing in mind <strong>the</strong> second paradigm <strong>of</strong> multibody dynamics, <strong>for</strong>ward dynamics, <strong>the</strong><br />

equations <strong>of</strong> motion are usually solved in order to obtain both <strong>the</strong> kinematic output and<br />

internal reaction <strong>for</strong>ces <strong>of</strong> a constrained biomechanical system when compelled to certain<br />

external <strong>for</strong>ces. Several authors developed studies regarding analysis <strong>of</strong> <strong>the</strong> <strong>for</strong>ward<br />

dynamics <strong>of</strong> musculoskeletal models <strong>for</strong> different motions, such as human gait [38, 39],<br />

arm control [40], cycling [20] and high-jumping [41].<br />

These types <strong>of</strong> multibody analysis require a physiological validation <strong>for</strong> <strong>the</strong> computed<br />

muscle ef<strong>for</strong>ts. This is obtained by <strong>the</strong> use <strong>of</strong> ma<strong>the</strong>matical models that mimic <strong>the</strong><br />

physical properties <strong>of</strong> excitable muscle tissue. <strong>Muscle</strong> models in this work are divided<br />

in three different types <strong>of</strong> dynamics: activation dynamics, contraction dynamics and<br />

fatigue dynamics. Activation dynamics model <strong>the</strong> time lag between <strong>the</strong> neural signal<br />

that arrives at a motor neuron and <strong>the</strong> correspondent muscle activation level, and is<br />

associated with <strong>the</strong> calcium dynamics <strong>of</strong> <strong>the</strong> muscle [1, 18, 20, 42]. Using such model, it<br />

is possible to predict <strong>the</strong> neural signal that triggered a certain muscle activation pattern.<br />

Several authors [18, 43] used bang-bang control models developed by He et al. [44],<br />

where <strong>the</strong> neural signal is regarded as a control signal that switches steeply between<br />

two states. These models fail when <strong>the</strong> control takes intermediate values between its<br />

boundaries [20], a fact that lead to new model developments to account <strong>for</strong> neural<br />

signal values amid 0 and 1 [31, 32]. In a numerical processing basis, some situations<br />

need to have a stable model <strong>for</strong> negative values, <strong>the</strong>re<strong>for</strong>e Raash et. al. [10] presented a<br />

non-linear model, used by Neptune [45] and Kaplan [20], a model that is well-behaved<br />

when handling neural signals that violate <strong>the</strong> physiological limits. Activation dynamics<br />

models deploy two physiological parameters: <strong>the</strong> time constants <strong>of</strong> rise and fall <strong>of</strong> <strong>the</strong><br />

muscle activation [45].<br />

<strong>Muscle</strong> contraction dynamics reports to <strong>the</strong> relation between muscle activation and<br />

respective <strong>for</strong>ce applied, i.e., to <strong>the</strong> contractile behaviour <strong>of</strong> muscle tissue. Regular<br />

tissue models are based on classical material models, such as <strong>the</strong> Maxwell model and<br />

<strong>the</strong> Kelvin-Voigt model, and only simulate <strong>the</strong> response <strong>of</strong> s<strong>of</strong>t tissues under compres-<br />

sive and tensile loads, not being able to mimic <strong>the</strong> active behaviour <strong>of</strong> muscle tissue<br />

at a macroscopic scale [1, 18]. Archibald <strong>Hill</strong>, in his experimental work, introduced an<br />

5


1. INTRODUCTION<br />

adaptation <strong>of</strong> <strong>the</strong> referred models, including a contractile component that explains <strong>the</strong><br />

functional nature <strong>of</strong> muscle tissue at a microscopic scale [46, 47]. Conversely, o<strong>the</strong>r com-<br />

putational muscle models are aimed to simulate <strong>the</strong> microscopic behaviour <strong>of</strong> muscle<br />

physiology, based on <strong>the</strong> cross-bridge <strong>the</strong>ory by Huxley [48]. These are very accurate,<br />

but highly complex and conditional on a considerable number <strong>of</strong> parameters that turn<br />

it computationally impracticable <strong>for</strong> elaborated musculoskeletal models [42]. This lead<br />

to <strong>the</strong> usage <strong>of</strong> <strong>the</strong> macroscopic <strong>Hill</strong>-based muscle models in <strong>the</strong> majority <strong>of</strong> studies that<br />

deal with muscle ef<strong>for</strong>ts [1, 2, 10, 11, 17, 18, 24, 29, 38, 42, 44]. The contraction dynam-<br />

ics model used in this work is composed by <strong>Hill</strong>-type model that comprises a passive<br />

element and a contractile element, each with <strong>the</strong> respective ma<strong>the</strong>matical <strong>for</strong>mulation<br />

equivalent to <strong>the</strong> one used in <strong>the</strong> work by Kaplan [20] and Silva [1]. This ma<strong>the</strong>mat-<br />

ical representation was conceived by relating <strong>the</strong> effective muscle <strong>for</strong>ce obtained <strong>for</strong><br />

variations <strong>of</strong> length, velocity and activation [18, 42].<br />

<strong>Muscle</strong>s are defined in <strong>the</strong> computational code by its geometry and <strong>for</strong>ce generating<br />

properties [1]. They are geometrically defined by <strong>the</strong> locations where <strong>the</strong>y blend with<br />

tendons, <strong>the</strong> origin and insertion, and by eventual via-points [1, 42]. Some studies<br />

consider additionally that muscles may be defined by wrapping around objects, in order<br />

to define in a much more accurate manner <strong>the</strong> way that muscle geometry is arranged [49].<br />

These can wrap around single objects [50] or multiple ones [51]. There is associated a<br />

type <strong>of</strong> path optimization when processing <strong>the</strong> more lifelike <strong>for</strong>m <strong>of</strong> joint wrapping [52].<br />

The third component <strong>of</strong> <strong>the</strong> considered muscle model takes into consideration <strong>the</strong> dy-<br />

namics <strong>of</strong> muscle fatigue. Physiologically and apart from o<strong>the</strong>r causes, peripheral fatigue<br />

increases proportionally to <strong>the</strong> consumption level <strong>of</strong> intramuscular glycogen and Adeno-<br />

sine triphosphate (ATP) [53]. Initial muscle fatigue studies by Scherrer and Monod [54]<br />

and Rohmert [55] stated <strong>the</strong> first interpretations <strong>of</strong> <strong>the</strong> relationship between ef<strong>for</strong>t levels<br />

and endurance times [56]. First models were based on empirical relationships, that lack<br />

applicability to different situations due to <strong>the</strong> absence <strong>of</strong> parameters. The first <strong>the</strong>o-<br />

retical models introduced a collection <strong>of</strong> physiological parameters and fatigue dynamic<br />

laws [57, 58], and were only practicable <strong>for</strong> single muscle analysis regarding its com-<br />

plexity and numerous parameters [59]. New models approach fatigue dynamics using<br />

simple biophysical principles, such as compartment dynamics and muscle recruitment<br />

hierarchy [59].<br />

6


1.4 Contributions<br />

Liu et al. [60] developed a muscle unit (MU) based fatigue model, where dynamic<br />

fatigue laws dictate <strong>the</strong> evolution <strong>of</strong> <strong>the</strong> muscle fitness using compartment <strong>the</strong>ory, and<br />

solely three physiological parameters [60]. This model however does not account <strong>for</strong><br />

variable muscle ef<strong>for</strong>ts. Xia and Law [59] used Liu’s model and developed a model that<br />

considers peripheral muscle fatigue <strong>for</strong> both plain and complex exercises at fluctuating<br />

muscle <strong>for</strong>ce intensities [59]. In this model, <strong>the</strong> fitness state <strong>of</strong> a specific muscle varies<br />

along a continuos scale, with each one <strong>of</strong> enclosed MUs having <strong>the</strong> possibility <strong>of</strong> being<br />

ideally activated, ideally resting or ideally fatigued [59]. In addition, it optionally<br />

takes into consideration <strong>the</strong> muscle recruitment hierarchy, described by Henneman [61].<br />

Guyton and Hall [53] identifies two main types <strong>of</strong> muscle skeletal fibers: slow and fast<br />

fibers. In his recruitment scheme, Henneman categorises muscle fibers in three different<br />

sub-types that are recruited in a specific order: slow, fast fatigue-resistant, and fast<br />

fatigable [59, 61]. In this work, a fatigue model based in <strong>the</strong> work by Xia and Law [59]<br />

was included in <strong>the</strong> implemented muscle model <strong>of</strong> <strong>the</strong> multibody dinamics routines,<br />

coupled with <strong>the</strong> muscle contraction dynamics model. It was tested <strong>for</strong> a single muscle<br />

case and <strong>for</strong> a more complex musculoskeletal model <strong>of</strong> <strong>the</strong> upper limb.<br />

1.4 Contributions<br />

The contributions <strong>of</strong> this <strong>the</strong>sis are fourfold:<br />

• To develop a flexible muscle contraction model, adapted from <strong>the</strong> work by Ka-<br />

plan [20] and Silva [1], using known <strong>for</strong>ce-length-velocity and activation scaling<br />

properties.<br />

• To develop a muscle fatigue model, adapted from <strong>the</strong> work by Xia and Law [59],<br />

using a three-compartment model and a muscle recruitment hierarchy model.<br />

• To implement <strong>the</strong> coupling <strong>of</strong> <strong>the</strong> developed contraction and fatigue dynamics<br />

muscle models in an existing FORTRAN-based multibody system dynamics pro-<br />

gram, in way that allows <strong>the</strong>ir application in both <strong>for</strong>ward and inverse dynamics<br />

perspectives.<br />

• To adapt <strong>the</strong> equations <strong>of</strong> motion <strong>of</strong> <strong>the</strong> multibody system, using <strong>the</strong> Newton<br />

method [1, 7], and solve <strong>the</strong>se equations by means <strong>of</strong> optimization procedures.<br />

7


1. INTRODUCTION<br />

1.5 Thesis Organization<br />

This <strong>the</strong>sis is divided in five chapters:<br />

Chapter 1 This first chapter explains <strong>the</strong> motivation and objectives <strong>of</strong> this work, and<br />

introduces <strong>the</strong> reader to <strong>the</strong> outline <strong>of</strong> this document.<br />

Chapter 2 The second chapter regards <strong>the</strong> underlying mechanism <strong>of</strong> muscle contrac-<br />

tion and its ma<strong>the</strong>matical model <strong>for</strong>mulation, being divided in two main sections:<br />

Skeletal <strong>Muscle</strong> Anatomy and Physiology, where <strong>the</strong> reader in introduced to <strong>the</strong><br />

anatomical and physiological properties <strong>of</strong> skeletal muscle tissue in order to un-<br />

derstand <strong>the</strong> muscle contraction mechanism, and Dynamics <strong>of</strong> <strong>the</strong> <strong>Muscle</strong> Tissue,<br />

where <strong>the</strong> muscle models are presented, with special consideration <strong>for</strong> <strong>the</strong> con-<br />

traction, activation and fatigue models that were used in <strong>the</strong> developed routines.<br />

Chapter 3 The third chapter starts by introducing <strong>the</strong> foundations <strong>of</strong> a multibody<br />

system <strong>for</strong>mulated with natural coordinates. Such preamble is followed by <strong>the</strong><br />

<strong>for</strong>mulation <strong>of</strong> concentrated <strong>for</strong>ces in our multibody approach. Subsequently, this<br />

<strong>for</strong>mulation is employed in <strong>the</strong> rearranging <strong>of</strong> <strong>the</strong> equations <strong>of</strong> motion. The next<br />

two sections explain how <strong>the</strong> equations <strong>of</strong> motion will be used in <strong>the</strong> standards<br />

<strong>of</strong> inverse and <strong>for</strong>ward dynamic analysis. Then, <strong>the</strong> optimization problem that is<br />

solved to deal with redundant muscle <strong>for</strong>ces is described. The chapter closes with<br />

a discussion regarding <strong>the</strong> used approach and methodology.<br />

Chapter 4 The fourth chapter encloses <strong>the</strong> results obtained in this work. It begins by<br />

explaining <strong>the</strong> method applied <strong>for</strong> <strong>the</strong> confirmation <strong>of</strong> <strong>the</strong> validity <strong>of</strong> <strong>the</strong> muscle<br />

contraction dynamics model that was put into practice. It follows with two ex-<br />

amples <strong>of</strong> success <strong>of</strong> <strong>the</strong> introduction <strong>of</strong> a fatigue model in multibody routines,<br />

showing <strong>the</strong> obtained results. Fur<strong>the</strong>rmore, a gait cycle is analysed from an in-<br />

verse dynamics perspective, where <strong>the</strong> muscle activations where determined <strong>for</strong> a<br />

lower extremity model considering some <strong>of</strong> <strong>the</strong> muscles that span <strong>the</strong> knee and<br />

ankle joints. The chapter terminates with a discussion contemplating <strong>the</strong> obtained<br />

results.<br />

8


1.5 Thesis Organization<br />

Chapter 5 In <strong>the</strong> last chapter <strong>the</strong> author exposes some <strong>of</strong> <strong>the</strong> most important con-<br />

clusions, suggesting some future developments related to <strong>the</strong> employed models,<br />

methodologies and optimization techniques.<br />

9


1. INTRODUCTION<br />

10


Chapter 2<br />

Musculoskeletal System <strong>Model</strong>ling<br />

The complex <strong>for</strong>med by muscles and tendons, stimulated by <strong>the</strong> central nervous<br />

system, will work as <strong>the</strong> biological actuator in order to per<strong>for</strong>m a specific motor task.<br />

As mentioned in [42]<br />

"[...] muscles and tendons are <strong>the</strong> interface between <strong>the</strong> CNS and <strong>the</strong><br />

articulated body segments."<br />

When excited, muscles apply a moment <strong>of</strong> <strong>for</strong>ce about specific joints. When ex-<br />

erted in a synchronised way <strong>the</strong>se moments will result in body balance or motion. A<br />

muscle can be considered uniarticular or biarticular, when able to span one or two<br />

joints [2] (some authors consider as biarticular <strong>the</strong> set <strong>of</strong> muscles that span more than<br />

two joints [27]).<br />

<strong>Muscle</strong> fibers can only produce contractile <strong>for</strong>ce, i.e., <strong>the</strong>y can only pull [2]. The<br />

tensile <strong>for</strong>ces produced by muscle are transmitted through <strong>the</strong> tendons to <strong>the</strong> location<br />

where tendons are attached to bone – muscle’s origin and insertion. Due to this limita-<br />

tion, we find many muscles in sets <strong>of</strong> antagonist pairs, which produce opposite segment<br />

movements. For instance, in <strong>the</strong> upper limb, we have <strong>the</strong> triceps brachii which is <strong>the</strong><br />

main muscle <strong>for</strong> elbow joint extension movements. Its antagonist is <strong>the</strong> biceps brachii,<br />

responsible <strong>for</strong> <strong>the</strong> flexion <strong>of</strong> this joint (along with <strong>the</strong> supination <strong>of</strong> <strong>the</strong> <strong>for</strong>earm). In<br />

biomechanical systems, degrees <strong>of</strong> freedom are usually overdriven, i.e., <strong>for</strong> a certain<br />

joint movement <strong>the</strong>re is more than one muscle that can deliver enough torque to pre-<br />

scribe <strong>the</strong> movement in question. In addition, even <strong>the</strong> simplest antagonist pair has a<br />

11


2. MUSCULOSKELETAL SYSTEM MODELLING<br />

countless number <strong>of</strong> muscle <strong>for</strong>ce combinations requested by <strong>the</strong> CNS to per<strong>for</strong>m joint<br />

rotation, since <strong>the</strong>re is <strong>the</strong> possibility <strong>of</strong> co-contraction. We call muscle redundancy to<br />

this situation, where <strong>the</strong> minimum number <strong>of</strong> muscles required is outnumbered. This<br />

will prompt one <strong>of</strong> <strong>the</strong> major aspects in <strong>the</strong> computation <strong>of</strong> muscle activations, since a<br />

specific motion to be studied will have an infinite set <strong>of</strong> muscle <strong>for</strong>ces combinations to<br />

generate it.<br />

There are three types <strong>of</strong> muscle tissue: cardiac, smooth and skeletal. The first is<br />

found in most <strong>of</strong> <strong>the</strong> cardiac structure. Smooth muscle is accountable <strong>for</strong> <strong>the</strong> movements<br />

<strong>of</strong> hollow inner structures. These are classified as involuntary muscles – that is, <strong>the</strong>ir<br />

contraction happens without a conscious control. Skeletal muscle produces voluntary<br />

anatomical segment displacements and promotes body balance, thus this is <strong>the</strong> relevant<br />

type <strong>of</strong> muscle <strong>for</strong> this work [62].<br />

In this chapter, an introduction to <strong>the</strong> underlying aspects <strong>of</strong> skeletal muscle anatomy<br />

and physiology will be made. We will begin with a brief reference to its main structures<br />

and design, and briefly describe <strong>the</strong> basic physical and biochemical phenomena in muscle<br />

contraction. In order to <strong>for</strong>mulate a computationally feasible model, <strong>the</strong> macroscopic<br />

properties <strong>of</strong> <strong>the</strong> behaviour <strong>of</strong> muscle tissue must be assumed. The dynamics <strong>of</strong> muscle<br />

contraction and muscle activation are <strong>for</strong>mulated. The chapter closes with a simple<br />

model <strong>for</strong> muscle fatigue.<br />

2.1 Skeletal <strong>Muscle</strong> Anatomy and Physiology<br />

Prior to any review <strong>of</strong> <strong>the</strong> employed muscle modelling methodology, a brief description <strong>of</strong><br />

<strong>the</strong> fundamentals <strong>of</strong> muscle anatomy and physiology is <strong>of</strong>fered. This section is presented<br />

to provide <strong>the</strong> essential background <strong>for</strong> <strong>the</strong> comprehension <strong>of</strong> <strong>the</strong> models presented<br />

in Section 2.2.<br />

2.1.1 Skeletal <strong>Muscle</strong> Anatomy<br />

As illustrated by Figure 2.1, muscles have a fasciculated structure. An outer layer<br />

covers <strong>the</strong> whole muscle, called epimysium. Inside it, assemblages <strong>of</strong> parallel muscle<br />

fibers that are assumed to extend to <strong>the</strong> entire length <strong>of</strong> <strong>the</strong> muscle, embedded in<br />

endomysium, constitute muscle fascicles [53]. Each one <strong>of</strong> <strong>the</strong>se fascicles are delimited<br />

by <strong>the</strong> perimysium. These three type <strong>of</strong> layers <strong>of</strong> connective tissue have elastic properties<br />

12


2.1 Skeletal <strong>Muscle</strong> Anatomy and Physiology<br />

and continue beyond <strong>the</strong> limits <strong>of</strong> muscle tissue, becoming more collagenous, to <strong>for</strong>m<br />

tendons [62].<br />

Figure 2.1: Arrangement <strong>of</strong> <strong>the</strong> skeletal muscle structure, from an external level to a<br />

molecular level [53].<br />

<strong>Muscle</strong> fibers are stacked in parallel and oriented at an acute angle to <strong>the</strong> ten-<br />

don, that can be different <strong>of</strong> 0 (zero). This angle is called pennation angle [42]. This<br />

arrangement is shown in Figure 2.2.<br />

In <strong>the</strong> same way that bundles <strong>of</strong> muscle fibers <strong>for</strong>m fascicles, muscle fibers are com-<br />

13


2. MUSCULOSKELETAL SYSTEM MODELLING<br />

Figure 2.2: Representation <strong>of</strong> <strong>the</strong> penation angle α between muscle fibers and tendons<br />

[42].<br />

posed by my<strong>of</strong>ibrils. These have a series arrangement <strong>of</strong> basic units called sarcomeres,<br />

which are <strong>the</strong> fundamental structure responsible <strong>for</strong> muscle contraction. Each sarcomere<br />

is composed by a complex <strong>of</strong> proteins comprised between two dark regions <strong>of</strong> proteins<br />

called Z discs and three bands are visible – A, H and I. The A bands are darker due to<br />

<strong>the</strong> existence <strong>of</strong> a brew <strong>of</strong> thick and thin filaments. Each A band includes a central H<br />

band which corresponds to <strong>the</strong> region where only thick filaments are present. Finally,<br />

<strong>the</strong> I band consists <strong>of</strong> a pale region where <strong>the</strong>re are only thin filaments [62].<br />

Thick filaments are <strong>for</strong>med by myosin proteins whose shape is shown in Figure 2.3-A.<br />

Anchored to Z discs, thin filaments extend to connect to <strong>the</strong> myosin complex. These thin<br />

filaments are composed by a system <strong>of</strong> actin, tropomyosin and troponin – Figure 2.3-B.<br />

Myosin and actin proteins overlap in <strong>the</strong> A band, with a projection <strong>of</strong> myosin heads into<br />

actin. These projections are known as cross-bridges – Figure 2.4 – and <strong>the</strong>ir interactions<br />

will devise <strong>the</strong> contraction machinery. Troponin and tropomyosin <strong>for</strong>m a complex that<br />

inhibit <strong>the</strong> bonding <strong>of</strong> actin with <strong>the</strong> myosin <strong>of</strong> <strong>the</strong> thick filaments when significant<br />

concentrations calcium ions (Ca 2+ ) are not present. This will impede contraction from<br />

happening when not desired. The whole overlapped structure is supported by a frame-<br />

work <strong>of</strong> titin molecules [53]. During contraction, thick and thin filaments glide without<br />

changing length. Only <strong>the</strong> sarcomere length reduces due to this gliding. There<strong>for</strong>e,<br />

throughout <strong>the</strong> whole process, A band length is constant, and H and I lengths shortens.<br />

Additionally, it should be referred that, an important thin membrane covers each<br />

muscle fiber, vital to sustain electrical membrane potential – <strong>the</strong> sarcolemma. This<br />

14


2.1 Skeletal <strong>Muscle</strong> Anatomy and Physiology<br />

Figure 2.3: Detail <strong>of</strong> a myosin molecule (A) and an actin filament (B) [53].<br />

Figure 2.4: Illustration <strong>of</strong> <strong>the</strong> cross-bridges <strong>for</strong>med by <strong>the</strong> connections <strong>of</strong> actin filaments<br />

with a myosin filament. Adapted from Reference [53].<br />

15


2. MUSCULOSKELETAL SYSTEM MODELLING<br />

layer has extensions that passes through <strong>the</strong> fiber inner structures called T tubules. A<br />

muscle fiber cytoplasm is known as sarcoplasm. Apart from several my<strong>of</strong>ibrils, it also<br />

contains a nucleus, various mitochondria associated with high concentrations <strong>of</strong> ATP,<br />

and a substantial network <strong>of</strong> smooth sarcoplasmic reticulum, that will store Ca 2+ .<br />

These structures are illustrated in Figure 2.5.<br />

Figure 2.5: Detail <strong>of</strong> a muscle fiber [62].<br />

Skeletal muscle contraction is voluntary, i.e., it happens under <strong>the</strong> conscious control<br />

<strong>of</strong> <strong>the</strong> brain. <strong>Muscle</strong> fibers must <strong>the</strong>n be innervated by a neuron – <strong>the</strong> motor neuron.<br />

The set <strong>of</strong> fibers innervated by a motor neuron plus <strong>the</strong> neuron itself is called a motor<br />

unit (MU ). The number <strong>of</strong> fibers in a MU may vary considerably. <strong>Muscle</strong>s that are<br />

responsible <strong>for</strong> precise movements will have few muscle fibers per MU, but a high number<br />

<strong>of</strong> motor units. This example contrasts with <strong>the</strong> one <strong>of</strong> muscles that control gross<br />

movements. The number <strong>of</strong> MU will be inferior, but each one will have a higher number<br />

<strong>of</strong> muscle fibers [62].<br />

Motor neurons ramify into several terminal axons that get close to a region <strong>of</strong> <strong>the</strong><br />

sarcolemma called motor end plate, keeping a small distance – <strong>the</strong> synaptic cleft [62].<br />

This region is known as neuromuscular junction – Figure 2.6.<br />

All <strong>the</strong> described structures will play a major role in <strong>the</strong> mechanisms <strong>of</strong> contraction,<br />

that will be described in <strong>the</strong> following subsection.<br />

16


2.1 Skeletal <strong>Muscle</strong> Anatomy and Physiology<br />

Figure 2.6: Neural muscular junction with three different detail levels [62].<br />

17


2. MUSCULOSKELETAL SYSTEM MODELLING<br />

2.1.2 Skeletal <strong>Muscle</strong> Physiology<br />

The general mechanism <strong>for</strong> muscle contraction – known as excitation-contraction cou-<br />

pling – undergoes several stages, commencing with <strong>the</strong> progress <strong>of</strong> <strong>the</strong> action potential<br />

from <strong>the</strong> nervous system to <strong>the</strong> sarcolemma – Figure 2.6 (a). When <strong>the</strong> action poten-<br />

tial travelling across <strong>the</strong> myelinated motor nerve reaches <strong>the</strong> synaptic bulbs (<strong>the</strong> motor<br />

axon end projections), a small quantity <strong>of</strong> acetylcholine or ACh (neurotransmitter) is<br />

secreted to <strong>the</strong> synaptic cleft. The ACh combines with its receptors in channels located<br />

in <strong>the</strong> motor end plate, allowing Na + ions to diffuse into <strong>the</strong> sarcoplasm, firing up a<br />

action potential in <strong>the</strong> sarcolemma [53].<br />

Once <strong>the</strong> action potential occurs, <strong>the</strong> contractile machinery must be activated –<br />

Figure 2.7. The discharged action potential will cross through <strong>the</strong> sarcolemma and<br />

widespread across <strong>the</strong> core <strong>of</strong> <strong>the</strong> fiber by <strong>the</strong> T tubules. This will lead to a release <strong>of</strong><br />

sizeable quantities <strong>of</strong> Ca 2+ from <strong>the</strong> sarcoplasmic reticulum. Calcium ions will bond to<br />

troponin, changing <strong>the</strong> shape <strong>of</strong> <strong>the</strong> troponin-tropomyosin complex and exposing actin<br />

double helix bonding sites to myosin cross-bridges [53].<br />

Figure 2.7: Contraction regulation mechanism by <strong>the</strong> troponin-tropomyosin complex,<br />

dependent on <strong>the</strong> concentration level <strong>of</strong> Ca 2+ [62].<br />

18


2.1 Skeletal <strong>Muscle</strong> Anatomy and Physiology<br />

Figure 2.8: Crossbridge <strong>the</strong>ory steps [62].<br />

<strong>Muscle</strong> contraction will happen solely in <strong>the</strong> presence <strong>of</strong> ATP and Ca 2+ . The sar-<br />

comere shortening process, that depends fundamentally on <strong>the</strong>se two components, can<br />

be explained by <strong>the</strong> crossbridge <strong>the</strong>ory and has its steps illustrated in Figure 2.8. Step<br />

1 – In <strong>the</strong> relaxed <strong>for</strong>m, myosin heads catalyse <strong>the</strong> decomposition <strong>of</strong> ATP (whenever<br />

present), increasing its energy and becoming active. Adenosine diphosphate (ADP) and<br />

phosphate ion resultant from this cleavage remain attached to <strong>the</strong> myosin head. If Ca 2+<br />

is present (Step 2), by <strong>the</strong> action potential mechanism already described, cross-briges<br />

will <strong>the</strong>n be capable <strong>of</strong> bond to actin filaments with a perpendicular configuration,<br />

loosing <strong>the</strong> phosphate ion (Step 3). Step 4 –Reaching this step, <strong>the</strong> energy stored will<br />

provoke what is known as power stroke, i.e., <strong>the</strong> head <strong>of</strong> myosin molecules will suffer<br />

a con<strong>for</strong>mational modification slopping its orientation towards <strong>the</strong> centre <strong>of</strong> <strong>the</strong> sar-<br />

comere. This ultimately results in <strong>the</strong> sliding <strong>of</strong> filaments and at this point <strong>the</strong> ADP<br />

molecule is released. Step 5 – A new ATP binds causing detaching <strong>the</strong> thick from <strong>the</strong><br />

thin filament. Step 6 – This molecule is cleaved such as in Step 1, making <strong>the</strong> head<br />

recover its perpendicular position to <strong>the</strong> actin filament and prompting <strong>the</strong> fiber <strong>for</strong> a<br />

new power stroke. For as long as calcium ions are available (not to mention ATP),<br />

19


2. MUSCULOSKELETAL SYSTEM MODELLING<br />

fiber contraction will proceed – Step 7 – per<strong>for</strong>ming <strong>the</strong> previously mentioned steps,<br />

providing that <strong>the</strong> physiological conditions are satisfied (presence <strong>of</strong> ATP and Ca 2+ ,<br />

physical limits <strong>of</strong> <strong>the</strong> sarcomere and bearable muscle load [53]) [62].<br />

In <strong>the</strong> absence <strong>of</strong> an action potential, <strong>the</strong> sarcoplasm tends to have low levels <strong>of</strong><br />

calcium ions. This is due to <strong>the</strong> existence <strong>of</strong> active transport membrane pumps that<br />

constantly store Ca 2+ ions back to <strong>the</strong> sarcoplasmic reticulum hindering <strong>the</strong> bonding<br />

<strong>of</strong> actin with myosin and <strong>the</strong>re<strong>for</strong>e inhibiting muscle contraction [53].<br />

The <strong>for</strong>ce applied by a certain muscle depends on various parameters and physical<br />

conditions. Its regulation is defined by <strong>the</strong> number <strong>of</strong> recruited muscle fibers and <strong>the</strong><br />

stimulation frequency <strong>of</strong> <strong>the</strong> neural signal. If a muscle fiber is stimulated by <strong>the</strong> CNS<br />

above a certain magnitude, <strong>the</strong>n <strong>the</strong> sarcolemma will become depolarised and develop<br />

a brief contraction, designated as twitch. An important aspect becomes apparent in<br />

terms <strong>of</strong> <strong>the</strong> frequency <strong>of</strong> stimulation, shown in Figure 2.9. When <strong>the</strong> time interval<br />

between successive twitches becomes smaller than <strong>the</strong> duration <strong>of</strong> a single twitch, <strong>the</strong>n<br />

contraction level is amplified by <strong>the</strong> sum <strong>of</strong> <strong>the</strong> twitches superposition. For increasing<br />

twitching frequency, <strong>the</strong> produced muscle <strong>for</strong>ce will also increase until it reaches a<br />

frequency threshold, where it is observed that <strong>the</strong> muscle contraction does not intensify<br />

any more and holds its value in a plateau – this level is known as tetanic contraction.<br />

2.2 Dynamics <strong>of</strong> <strong>the</strong> <strong>Muscle</strong> Tissue<br />

In order to implement a computational model describing muscle behaviour, <strong>the</strong> dynam-<br />

ics <strong>of</strong> muscle tissue play a vital role in <strong>the</strong> acquisition <strong>of</strong> physiologically valid muscle <strong>for</strong>ce<br />

values, and <strong>the</strong>re<strong>for</strong>e in <strong>the</strong> construction <strong>of</strong> such model. There are some reductionist<br />

models [48, 63] that describe <strong>the</strong> microscopic properties introduced in Section 2.1. These<br />

models are usually based on <strong>the</strong> crossbridge <strong>the</strong>ory and regardless <strong>of</strong> <strong>the</strong>ir accuracy in<br />

describing <strong>the</strong> mechanics and chemistry involved in contraction <strong>of</strong> muscle fibers, <strong>the</strong>ir<br />

high complexity and numerous parameters lead to computationally inefficient routines<br />

that prove to be disadvantageous. These drawbacks are specially evident when complex<br />

mechanical structures, with many muscle structures, are considered. For a robust and<br />

computationally efficient modelling, several authors [1, 2, 20, 42] employed a macro-<br />

scopic approach <strong>for</strong> <strong>the</strong> muscle model, where some <strong>of</strong> its most important properties are<br />

considered in order to identify <strong>the</strong> inherent parameters.<br />

20


2.2 Dynamics <strong>of</strong> <strong>the</strong> <strong>Muscle</strong> Tissue<br />

Figure 2.9: Relation between twitch frequency and effective muscle contraction [1].<br />

In what muscle tissue modelling is concerned, two major types <strong>of</strong> dynamics are usu-<br />

ally identified: Activation Dynamics (Section 2.2.1) and Contraction dynamics (Sec-<br />

tion 2.2.2) – Figure 2.10. The <strong>for</strong>mer describes <strong>the</strong> conversion <strong>of</strong> a CNS neural signal<br />

u(t) into a muscle tissue activation state a(t). The latter correlates a(t) with muscle<br />

<strong>for</strong>ce development. A novel model to multibody methodologies is introduced: a muscle<br />

fatigue dynamics model (Section 2.2.3).<br />

u(t) a m (t) F m (t)<br />

Activation<br />

Contraction<br />

✲ ✲ ✲<br />

Dynamics<br />

Dynamics<br />

Figure 2.10: Scheme <strong>of</strong> muscle tissue dynamics, with a series model <strong>of</strong> Activation Dynamics<br />

and Contraction Dynamics.<br />

2.2.1 Activation Dynamics<br />

Also known as excitation-contraction dynamics, activation dynamics (AD) models <strong>the</strong><br />

relation between <strong>the</strong> neural signal that arrives at a motor neuron u(t) and <strong>the</strong> muscle<br />

21


2. MUSCULOSKELETAL SYSTEM MODELLING<br />

activation level a(t), i.e., <strong>the</strong> dimensionless proportion <strong>of</strong> muscle fibers that are instan-<br />

taneously active at time t [18]. By including an AD model, it is possible to predict <strong>the</strong><br />

neural signal that triggered a certain muscle activation pattern and <strong>the</strong>re<strong>for</strong>e <strong>the</strong> motion<br />

<strong>of</strong> <strong>the</strong> biomechanical system at issue. Despite <strong>the</strong> fact <strong>the</strong> developed routines in work<br />

do not count with this sort <strong>of</strong> model, it is important to comprehend its characteristics<br />

in order to implement it in a future work.<br />

Physiologically <strong>the</strong>re is a time lag between <strong>the</strong> neural signal u(t) and <strong>the</strong> corre-<br />

sponding muscle activation a(t). The physical meaning <strong>of</strong> this lag is associated with<br />

<strong>the</strong> muscle’s calcium dynamics [20, 42], described in Section 2.1.2. In <strong>the</strong> same manner<br />

that <strong>the</strong> contraction dynamics model was developed, it is desired to simplify our AD<br />

model by representing <strong>the</strong> behaviour observed from a macroscopic point-<strong>of</strong>-view. At<br />

this level, AD is represented by first-order ordinary differential equations [1].<br />

Several authors like Yamaguchi [18] and Pandy et al. [43] used a model developed by<br />

He et al. [44], where <strong>the</strong> neural signal u(t) was regarded as bang-bang control, i.e., as a<br />

control signal that switches steeply between two states (in this case taking <strong>the</strong> boundary<br />

values – 0 or 1). These models report with precision <strong>the</strong> AD <strong>of</strong> a signal u(t) with a<br />

binary nature, however failing when <strong>the</strong> control takes intermediate values between its<br />

boundaries [20]. <strong>Development</strong>s from this model, such as <strong>the</strong> one in <strong>the</strong> work <strong>of</strong> Anderson<br />

and Pandy [32], or <strong>the</strong> one developed in Pandy and Hull [31] account <strong>for</strong> values <strong>of</strong> u(t)<br />

amid 0 and 1, but are unstable <strong>for</strong> negative values <strong>of</strong> <strong>the</strong> neural signal. These negative<br />

values may occur, despite <strong>the</strong> fact that u(t) is bounded between 0 and 1, since some<br />

iterative methods require its calculation below <strong>the</strong> lower bound. In response to that,<br />

Raash et. al. [10] presented a non-linear model, used by Neptune [45] and Kaplan [20],<br />

that is stable when handling neural signals that violate <strong>the</strong> physiological limits. This<br />

model states that <strong>the</strong> change rate <strong>of</strong> muscle activation ˙a(t) is related to a(t) and u(t)<br />

in <strong>the</strong> following way:<br />

where<br />

˙a(t) =<br />

<br />

(u(t) − a(t))(c1u(t) + c2) u(t) ≥ a(t)<br />

(u(t) − a(t))c2<br />

c1 =<br />

c2 =<br />

1<br />

τrise<br />

1<br />

τfall<br />

22<br />

− 1<br />

τfall<br />

u(t) < a(t)<br />

(2.1)<br />

(2.2)<br />

(2.3)


2.2 Dynamics <strong>of</strong> <strong>the</strong> <strong>Muscle</strong> Tissue<br />

The parameters τrise and τfall describe respectively <strong>the</strong> time constants <strong>of</strong> rise and fall<br />

<strong>of</strong> <strong>the</strong> a(t), i.e., <strong>the</strong> time <strong>of</strong> activation and relaxation [45]. The latter term is generally<br />

bigger than <strong>the</strong> first [64]. The consistency <strong>of</strong> this model is illustrated in Figure 2.11.<br />

(a) Response to bang-bang control signal.<br />

(b) Response to an intermediate value <strong>of</strong> neu- (c) <strong>Model</strong> consistency <strong>for</strong> negative values <strong>of</strong><br />

ral signal control u(t).<br />

u(t).<br />

Figure 2.11: Activation dynamics model consistency [20].<br />

2.2.2 Contraction Dynamics<br />

In order to simulate <strong>the</strong> active behaviour <strong>of</strong> muscles in body motion, an explicit model<br />

<strong>of</strong> <strong>the</strong> contractile structures must be implemented to estimate <strong>the</strong> muscle <strong>for</strong>ces. The<br />

dynamics <strong>of</strong> muscle contraction are described in this subsection. These will bridge<br />

values <strong>of</strong> muscle activation a(t) with muscle <strong>for</strong>ce F m .<br />

The primary ma<strong>the</strong>matical models used in generic tissue modelling include passive<br />

ones (such as <strong>the</strong> Maxwell and <strong>the</strong> Kelvin-Voigt models), that are able to precisely mimic<br />

<strong>the</strong> behaviour <strong>of</strong> s<strong>of</strong>t tissues under compressive and tensile loads [1, 18]. However, due<br />

to <strong>the</strong> absence <strong>of</strong> an active element, <strong>the</strong>se models are incapable <strong>of</strong> representing <strong>the</strong><br />

dynamics <strong>of</strong> muscle contractile structures. Archibald <strong>Hill</strong> introduced an adaptation <strong>of</strong><br />

23


2. MUSCULOSKELETAL SYSTEM MODELLING<br />

<strong>the</strong> Kelvin model, including an additional contractile element [46] that simulates <strong>the</strong><br />

active macroscopic action <strong>of</strong> <strong>the</strong> cross-brige cycle explained in Subsection 2.1.2. Due<br />

to this active characteristic <strong>Hill</strong>-type models are widely used in <strong>the</strong> biomechanics field<br />

to reproduce both contractile and passive behaviour <strong>of</strong> muscle tissue, since <strong>the</strong>se have<br />

accessible parameters and <strong>the</strong>y are computationally tractable <strong>for</strong> systems with several<br />

muscles.<br />

The model used in this work is illustrated in Figure 2.12(a). It is composed by a<br />

passive element (PE) that takes <strong>the</strong> non-linear passive elastic properties <strong>of</strong> <strong>the</strong> muscle<br />

tissue and a contractile element (CE) that accounts <strong>for</strong> both <strong>the</strong> contractile structures<br />

and <strong>the</strong> viscous <strong>for</strong>ce produced by intracellular and intercellular fluids enclosed in <strong>the</strong><br />

muscle [1]. This work, similarly to <strong>the</strong> works by Silva [1] and Kaplan [20], does not<br />

consider <strong>the</strong> standard <strong>Hill</strong>-model – a Kelvin model with <strong>the</strong> CE inserted parallel to <strong>the</strong><br />

damping element (DE) as representated in Figure 2.12(b). It neglects <strong>the</strong> series elastic<br />

element (SE) related to cross-bridge stiffness and includes <strong>the</strong> behaviour <strong>of</strong> <strong>the</strong> DE in<br />

<strong>the</strong> expressions <strong>of</strong> <strong>the</strong> CE. The expression <strong>of</strong> <strong>the</strong> exerted muscle <strong>for</strong>ce <strong>for</strong> <strong>the</strong> model in<br />

consideration <strong>for</strong> this work, is straight<strong>for</strong>wardly described by:<br />

F m = F m CE + F m P E<br />

(2.4)<br />

(a) The used ma<strong>the</strong>matical <strong>Hill</strong>-type muscle model [1] (b) Generic <strong>Hill</strong>-type muscle model [1].<br />

Figure 2.12: <strong>Hill</strong>-type muscle models.<br />

According to Zajac [42] and Yamaguchi [18], <strong>the</strong>re are three key properties in <strong>the</strong><br />

muscle tissue. The first one, already mentioned in this chapter, states that it can only<br />

produce tensile <strong>for</strong>ces. Even in <strong>the</strong> possibility <strong>of</strong> muscles being able to push, tendons<br />

would buckle, cancelling that effect [18]. The following two properties describe <strong>the</strong> <strong>for</strong>ce<br />

24


2.2 Dynamics <strong>of</strong> <strong>the</strong> <strong>Muscle</strong> Tissue<br />

output <strong>for</strong> a certain state <strong>of</strong> muscle length and muscle velocity. <strong>Muscle</strong> activation is<br />

evidently also taken into consideration <strong>for</strong> <strong>the</strong> ma<strong>the</strong>matical expressions correspondent<br />

to <strong>the</strong> calculation <strong>of</strong> <strong>the</strong> muscle <strong>for</strong>ces, since <strong>the</strong> amount <strong>of</strong> <strong>for</strong>ce produced by a muscle<br />

depends on <strong>the</strong> number <strong>of</strong> muscle unites recruited.<br />

Force-length relationship<br />

Several authors, such as <strong>Hill</strong> [47] or more recently Zajac [42], exposed <strong>the</strong> evolution<br />

<strong>of</strong> <strong>the</strong> <strong>for</strong>ces produced by a fully activated muscle fiber (a m (t) = 1) with its length.<br />

There were some evolutions concerning <strong>the</strong> bounding values <strong>for</strong> <strong>the</strong> region where active<br />

muscle <strong>for</strong>ce is generated. The more recent studies state that this region <strong>for</strong> a muscle<br />

length L m is comprised between 0.5L m 0 < Lm < 1.5L m 0 , where Lm 0<br />

corresponds to <strong>the</strong><br />

muscle fiber resting length(or relaxed length). The muscle resting length L m 0 usually<br />

verifies to have a close value to <strong>the</strong> length at which <strong>the</strong> active muscle <strong>for</strong>ce reaches its<br />

maximum value when isometrically contracted [42], <strong>the</strong> muscle’s optimal length [18, 42].<br />

The <strong>for</strong>ce developed at total muscle activation and fiber length L m = L m 0<br />

is designated<br />

maximum isometric <strong>for</strong>ce or optimal <strong>for</strong>ce F m 0 [1, 18]. This value does not correspond<br />

to <strong>the</strong> maximum <strong>for</strong>ce that <strong>the</strong> muscle can exert, but to <strong>the</strong> maximum active <strong>for</strong>ce in an<br />

isometric contraction. As we will see, <strong>the</strong>re is a passive <strong>for</strong>ce that can sum to <strong>the</strong> active<br />

muscle contribution, overtaking <strong>the</strong> value <strong>of</strong> F m 0<br />

. The value <strong>of</strong> <strong>the</strong> maximum isometric<br />

<strong>for</strong>ce may be estimated by multiplying <strong>the</strong> muscle’s average cross sectional area by <strong>the</strong><br />

maximal muscle specific tension (roughly 31.39N/cm 2 ) [18].<br />

Figure 2.13(a) relates <strong>the</strong> overlap level between <strong>the</strong> actin and myosin fibers with<br />

<strong>the</strong> contractile <strong>for</strong>ce <strong>of</strong> a fully activated muscle fiber. This overlapping varies along <strong>the</strong><br />

sarcomere length and it will have influence in <strong>the</strong> muscle’s capacity <strong>of</strong> generating con-<br />

tractile <strong>for</strong>ce. Points A and D represent states where <strong>the</strong> muscle is unable <strong>of</strong> per<strong>for</strong>ming<br />

its maximum <strong>for</strong>ce, respectively due to an adjacent positioning <strong>of</strong> <strong>the</strong> myosin with <strong>the</strong><br />

Z discs and no actin-myosin overlap [53]. The contractile <strong>for</strong>ce peaks when <strong>the</strong> highest<br />

number <strong>of</strong> cross-bridges are <strong>for</strong>med. When reporting <strong>the</strong>se facts to <strong>the</strong> whole muscle,<br />

a curve similar to <strong>the</strong> one in Figure 2.13(b) is obtained. Its shape arises due to <strong>the</strong><br />

evolution <strong>of</strong> <strong>the</strong> actin-myosin overlapping conditions already mentioned.<br />

A passive <strong>for</strong>ce also emerges from <strong>the</strong> length evolution <strong>of</strong> muscle fibers. It is<br />

thought [42] that this <strong>for</strong>ce is due to intrafiber elasticity. For muscle lengths larger<br />

25


2. MUSCULOSKELETAL SYSTEM MODELLING<br />

(a) (b)<br />

Figure 2.13: Discrete length-tension diagram <strong>of</strong> <strong>the</strong> contractile contribution <strong>of</strong> a single<br />

fully activated sarcomere [53] (a) and <strong>the</strong> same relationship scaled to <strong>the</strong> whole muscle [1]<br />

(b).<br />

than <strong>the</strong> resting length, <strong>the</strong>se tensions become apparent, reaching values larger than<br />

F m 0<br />

as observed in Figure 2.14.<br />

Force-velocity relationship<br />

The rate at which <strong>the</strong> length <strong>of</strong> muscle changes affects <strong>the</strong> magnitude <strong>of</strong> <strong>the</strong> obtained<br />

muscle <strong>for</strong>ce. As illustrated in Figure 2.15, <strong>for</strong> concentric contraction, that is <strong>the</strong> short-<br />

ening <strong>of</strong> activated muscle, <strong>the</strong> contractile <strong>for</strong>ce is less that <strong>the</strong> one observed at isometric<br />

contractions with <strong>the</strong> same level <strong>of</strong> muscle activation [1, 42]. The opposite happens in<br />

eccentric contractions when <strong>the</strong> activated muscle is leng<strong>the</strong>ning. Empirically <strong>the</strong> muscle<br />

apparatus shows a per<strong>for</strong>mance similar to a fluid "damper" [18] and this behaviour was<br />

described by <strong>Hill</strong> [46] as<br />

(F m + a) (b + v m ) = (F m 0 + a) b (2.5)<br />

where v m is <strong>the</strong> muscle speed, that corresponds to <strong>the</strong> opposite <strong>of</strong> <strong>the</strong> <strong>the</strong> rate <strong>of</strong> length<br />

change ˙ L m (or <strong>the</strong> muscle contraction velocity). A special attention should be taken<br />

to this relation ( ˙ L m = −v m ), in order to relate <strong>the</strong> presented model graphs with <strong>the</strong><br />

used equations. ˙ L m is positive <strong>for</strong> eccentric contractions and negative <strong>for</strong> concentric<br />

26


2.2 Dynamics <strong>of</strong> <strong>the</strong> <strong>Muscle</strong> Tissue<br />

Figure 2.14: Force-length relationship <strong>of</strong> <strong>the</strong> passive element [1].<br />

ones. Coefficients a and b are <strong>the</strong> values that define <strong>the</strong> asymptotes that define <strong>the</strong><br />

hyperbole curve in Figure 2.15. The observed range <strong>for</strong> <strong>the</strong> muscle rate <strong>of</strong> length<br />

change is [− ˙ L m 0 , 0.5 ˙ L m 0 ], where ˙ L m 0<br />

˙L m 0 ≈ 10Lm 0 [1].<br />

is <strong>the</strong> maximum contractile velocity, normalised as<br />

Only <strong>the</strong> active contribution <strong>of</strong> muscle will be affected by its velocity. The passive<br />

<strong>for</strong>ces mentioned previously will be considered to be <strong>the</strong> same <strong>for</strong> all muscle velocities,<br />

i.e., <strong>the</strong>y will only depend on <strong>the</strong> length <strong>of</strong> <strong>the</strong> muscle.<br />

Activation scaling<br />

In <strong>the</strong> model considered <strong>for</strong> this work, <strong>the</strong> <strong>for</strong>ces developed by a muscle are classified<br />

as passive element <strong>for</strong>ces and contractile element <strong>for</strong>ces. Both depend on <strong>the</strong> length <strong>of</strong><br />

muscle fibers L m , but <strong>the</strong> latter is additionally conditioned by <strong>the</strong> muscle’s contractile<br />

velocity ˙ L m and activation level a(t). The assumption that all fibers respond to neural<br />

activation in <strong>the</strong> same manner is made (which is not completely true, since it is known<br />

that <strong>the</strong>re are several types <strong>of</strong> muscle fibers with different contraction patterns) and<br />

<strong>the</strong>re<strong>for</strong>e it is considered that <strong>the</strong> <strong>for</strong>ce-length and <strong>for</strong>ce-velocity curves are linearly<br />

scaled by <strong>the</strong> muscle activation a(t) as it can be observed in Figure 2.16 [42].<br />

The <strong>for</strong>ce exerted by <strong>the</strong> contractile element is <strong>the</strong>n composed by a factor dependent<br />

on its relationship with <strong>the</strong> muscle length F m L (Lm (t)) and velocity F m ˙ L (L m (t)). This<br />

27


2. MUSCULOSKELETAL SYSTEM MODELLING<br />

Figure 2.15: Force-velocity relationship [18].<br />

Figure 2.16: Activation scaling evidence: Force-length and <strong>for</strong>ce-velocity relationships<br />

<strong>for</strong> different levels <strong>of</strong> muscle activation a(t) [18].<br />

28


factor is <strong>the</strong>n multiplied by a(t) as indicated in Equation 2.6.<br />

2.2 Dynamics <strong>of</strong> <strong>the</strong> <strong>Muscle</strong> Tissue<br />

F m CE(a(t), L m (t), ˙ L m (t)) = F m L (Lm (t))F m L ˙ ( ˙ Lm (t))<br />

F m a(t) (2.6)<br />

0<br />

The factor to be multiplied by <strong>the</strong> muscle activation a m (t) corresponds to <strong>the</strong> avail-<br />

able muscle contractile <strong>for</strong>ce which is referred hereafter as ˆ F m CE<br />

tion 2.7.<br />

and indicated in Equa-<br />

ˆF m CE = F m L (Lm (t))F m L ˙ ( ˙ Lm (t))<br />

F m . (2.7)<br />

0<br />

This term plays an important role in <strong>the</strong> following description <strong>of</strong> <strong>the</strong> fatigue model<br />

(Section 2.2.3) and in <strong>the</strong> integration <strong>of</strong> <strong>the</strong> muscle constitutive model with <strong>the</strong> multi-<br />

body methodology (Chapter 3). Equation 2.4 can be rewritten considering <strong>the</strong> activa-<br />

tion term as an independent one:<br />

Analytical expressions<br />

F m = F m P E + ˆ F m CEa m . (2.8)<br />

In order to analytically express <strong>the</strong> major characteristics <strong>of</strong> <strong>the</strong> contraction dynamics<br />

model described be<strong>for</strong>e in its computational counterpart, <strong>the</strong> ma<strong>the</strong>matical expressions<br />

developed in <strong>the</strong> works <strong>of</strong> Kaplan [20] and Silva [1] are used. These take into account<br />

<strong>the</strong> properties previously described.<br />

The <strong>for</strong>ce-length and <strong>for</strong>ce-velocity relationships <strong>of</strong> <strong>the</strong> contractile element, referred<br />

in Equation 2.6, are described in Equation 2.9 and Equation 2.10.<br />

F m ˙ L ( ˙ L m (t)) =<br />

F m L (L m (t)) = F m −<br />

0 e<br />

⎪⎩ − πF m 0<br />

" »<br />

− 9<br />

4<br />

„ L m (t)<br />

L m 0<br />

− 19<br />

«– 4<br />

− 20<br />

1<br />

»<br />

− 4<br />

9<br />

„<br />

L<br />

m<br />

(t)<br />

4 Lm −<br />

0<br />

19<br />

«– #<br />

2<br />

20<br />

(2.9)<br />

⎧<br />

0<br />

L˙ m (t) < −L˙ m<br />

0<br />

⎪⎨<br />

− F m 0<br />

arctan (5) arctan<br />

<br />

−5 ˙ Lm (t)<br />

˙L m <br />

+ F<br />

0<br />

m 0 − ˙ Lm 0 ≤ ˙ Lm (t) ≤ 0.2 ˙ Lm 0<br />

4 arctan (5) + F m 0 0.2 ˙ L m 0 < ˙ L m (t)<br />

(2.10)<br />

Equation 2.10 is <strong>the</strong> used expression that corresponds to <strong>the</strong> relation <strong>of</strong> Equation 2.5.<br />

The passive element <strong>for</strong>ce will only depend on <strong>the</strong> muscle’s length L m (t) and <strong>the</strong> math-<br />

ematical approximation used is<br />

29


2. MUSCULOSKELETAL SYSTEM MODELLING<br />

⎧<br />

0 L m (t) < L m 0<br />

F m P E(L m ⎪⎨<br />

(t)) = 8<br />

⎪⎩<br />

F m 0<br />

Lm 0 3 (Lm (t) − L m 0 ) 3 Lm 0 ≤ Lm (t) ≤ 1.63Lm 0<br />

2F m 0 1.63L m 0 < Lm (t)<br />

. (2.11)<br />

This contractile model is introduced in <strong>the</strong> multibody system dynamics routines <strong>of</strong><br />

<strong>the</strong> s<strong>of</strong>tware APOLLO, as schemed in <strong>the</strong> block diagrams <strong>of</strong> Figure 2.17 (on a <strong>for</strong>ward<br />

dynamics perspective) and Figure 2.18 (on an inverse dynamics perspective).<br />

Figure 2.17: Block diagram <strong>of</strong> <strong>the</strong> muscle contraction dynamics model implemented in<br />

<strong>the</strong> <strong>for</strong>ward dynamics multibody system routines.<br />

Figure 2.18: Block diagram <strong>of</strong> <strong>the</strong> muscle contraction dynamics model implemented in<br />

<strong>the</strong> inverse dynamics multibody system routines.<br />

30


2.2.3 <strong>Muscle</strong> <strong>Fatigue</strong><br />

2.2 Dynamics <strong>of</strong> <strong>the</strong> <strong>Muscle</strong> Tissue<br />

<strong>Muscle</strong> fatigue is a common condition and empirically understood by <strong>the</strong> majority <strong>of</strong><br />

people as <strong>the</strong> debilitation <strong>of</strong> <strong>the</strong> muscle <strong>for</strong>ce production per<strong>for</strong>mance. The metabolic<br />

conditions required <strong>for</strong> proper muscle contraction (described in Subsection 2.1.2) are<br />

altered during sustained contractions and this will lead to an incapacity <strong>of</strong> <strong>the</strong> muscle<br />

to deliver <strong>the</strong> same fitness levels. From a physiological point-<strong>of</strong>-view, this weakness<br />

increases proportionally with <strong>the</strong> consumption level <strong>of</strong> intramuscular glycogen (where<br />

glucose is stored) and ATP [53]. In addition, <strong>the</strong> conditions <strong>of</strong> neural signal transmis-<br />

sion are also altered after extended muscle action, narrowing its output [53]. There are<br />

several o<strong>the</strong>r justifications <strong>for</strong> muscle fatigue not related to <strong>the</strong> effects <strong>of</strong> prolonged mus-<br />

cle activity, such as <strong>the</strong> restriction <strong>of</strong> blood supply (and consequent nutrient privation),<br />

however <strong>the</strong>se are not related to <strong>the</strong> scope <strong>of</strong> this work.<br />

Initial empirical models were developed by Rohmert [55] who gave name to what<br />

became known as <strong>the</strong> Rohmert curves, <strong>for</strong> which an example is depicted in Figure 2.19.<br />

However, due to versatility limitations, <strong>the</strong>se first models were overpowered by new<br />

<strong>the</strong>oretical ones. Hawkins and Hull [57] studied <strong>the</strong> effects <strong>of</strong> fatigue in long-lasting<br />

ef<strong>for</strong>ts by considering a set <strong>of</strong> empirical fatigue parameters and a model that calculates<br />

muscle <strong>for</strong>ces as <strong>the</strong> sum <strong>of</strong> individual fiber <strong>for</strong>ces. The dependence <strong>of</strong> <strong>for</strong>ce with<br />

time is <strong>the</strong>n based on empirical input that requires an experimental consistence, which<br />

leads to a <strong>for</strong>eseeable inaccuracy. The first <strong>the</strong>oretical models introduced a collection<br />

<strong>of</strong> physiological parameters and fatigue dynamic laws, such as <strong>the</strong> ones presented in<br />

Ding el al. [58]. Despite <strong>the</strong> precision <strong>of</strong> <strong>the</strong>se models, <strong>the</strong>ir complexity and numerous<br />

parameters lead <strong>the</strong>m to be computationally disadvantageous when analysing multiple<br />

muscles. In order to tackle <strong>the</strong>se drawbacks, new <strong>the</strong>oretical models derive <strong>the</strong> muscle<br />

<strong>for</strong>ce function from simple biophysical principles. This work uses an example <strong>of</strong> <strong>the</strong>se<br />

new <strong>the</strong>oretical models, adapted from <strong>the</strong> work by Xia and Law [59].<br />

The three-compartment <strong>the</strong>ory<br />

The principle <strong>of</strong> <strong>the</strong> model used in this work derives from <strong>the</strong> MU-based fatigue model<br />

proposed by Liu et al. [60]. In <strong>the</strong>ir work, a fatigue model where dynamic fatigue<br />

laws prescribe <strong>the</strong> evolution <strong>of</strong> fiber conditions in three ideal states (resting, activated<br />

and fatigued), over a certain period <strong>of</strong> time, is considered. This concept is known as<br />

31


2. MUSCULOSKELETAL SYSTEM MODELLING<br />

Figure 2.19: Curve based on Rohmert’s relationship between %MVC (percentage <strong>of</strong><br />

maximum voluntary contraction) and endurance time (minutes) [56].<br />

compartment <strong>the</strong>ory, which as been applied in a diverse set <strong>of</strong> scientific areas, namely<br />

substance transport and chemical reaction phenomena modelling [59, 65, 66].<br />

Liu’s model is based in simple biophysical principles and solely considers three pa-<br />

rameters: a fatigue factor (F ), a recovery factor (R) and <strong>the</strong> total number <strong>of</strong> motor<br />

units in <strong>the</strong> muscle [60]; and an input factor describing <strong>the</strong> brain ef<strong>for</strong>t (BE). The<br />

BE variable plays an important role in this model, since it is analogous to <strong>the</strong> muscle<br />

activation a m mentioned in <strong>the</strong> previous subsections. Note that this model becomes<br />

appropriate <strong>for</strong> fitting experimental data, due to <strong>the</strong> few parameters that are involved,<br />

and <strong>for</strong> a far more general use than <strong>the</strong> o<strong>the</strong>r mentioned models. However, this model<br />

fails to consider conditions <strong>of</strong> non-constant muscle ef<strong>for</strong>ts, which is a major disadvan-<br />

tage in our work, as it involves a non-linear application <strong>of</strong> muscle <strong>for</strong>ces to a multibody<br />

system.<br />

Considering <strong>the</strong> consistency and computational efficiency <strong>of</strong> Liu’s model, and in<br />

response to <strong>the</strong> fact that this model is not ready to deal with variable muscle ef<strong>for</strong>ts, Xia<br />

and Law [59] developed a model that considers peripheral muscle fatigue <strong>for</strong> both plain<br />

and complex exercises <strong>for</strong> fluctuating muscle <strong>for</strong>ce intensities. The same compartment<br />

32


2.2 Dynamics <strong>of</strong> <strong>the</strong> <strong>Muscle</strong> Tissue<br />

<strong>the</strong>ory model <strong>of</strong> Liu’s was taken into consideration. To apply it, and despite <strong>the</strong> fact that<br />

<strong>the</strong> fitness state <strong>of</strong> a specific MU varies along a continuous scale, each one <strong>of</strong> <strong>the</strong>se MUs<br />

are minded as being ideally activated, ideally resting <strong>of</strong> ideally fatigued. Physiologically<br />

it is never expectable to have a MU ei<strong>the</strong>r fully fatigued or fully comprised in one <strong>of</strong><br />

<strong>the</strong> o<strong>the</strong>r states. However, this model gives us <strong>the</strong> whole muscle fitness state by <strong>the</strong><br />

mixture <strong>of</strong> <strong>the</strong> set <strong>of</strong> MUs that <strong>for</strong>m it and <strong>the</strong>re<strong>for</strong>e ma<strong>the</strong>matically it will be <strong>the</strong> sum<br />

<strong>of</strong> <strong>the</strong> compartment proportions. <strong>Muscle</strong> units exerting maximum <strong>for</strong>ce will fall in <strong>the</strong><br />

activated compartment MA and <strong>the</strong> ones with zero available contractile <strong>for</strong>ce will be<br />

consigned to <strong>the</strong> fatigued compartment MF . The compartment <strong>of</strong> <strong>the</strong> resting MUs is<br />

symbolised as MR as schematically indicated in Figure 2.20.<br />

Figure 2.20: Three-compartment <strong>the</strong>ory flowchart.<br />

For model flexibility purposes [59], <strong>the</strong> quantity <strong>of</strong> muscle fibers set in a particular<br />

compartment are given in percent <strong>of</strong> maximum voluntary contraction (%MVC). In <strong>the</strong><br />

beginning <strong>of</strong> <strong>the</strong> exercise, it is considered that every muscle fiber is resting, i.e., it is<br />

assumed that MR = 100%. The available contractile <strong>for</strong>ce will <strong>the</strong>n be limited to <strong>the</strong><br />

fitness state <strong>of</strong> <strong>the</strong> whole muscle. The MUs that are available <strong>for</strong> <strong>for</strong>ce production will be<br />

<strong>the</strong> ones laying in <strong>the</strong> activated and resting compartments (<strong>the</strong> fatigued compartment<br />

is not considered). For that effect, a new term is introduced to describe <strong>the</strong> muscle<br />

strength that still can be developed considering fatigue, <strong>the</strong> residual capacity (RC).<br />

33


2. MUSCULOSKELETAL SYSTEM MODELLING<br />

RC(t) = MA + MR = 100% − MF . (2.12)<br />

When ma<strong>the</strong>matically adapted to this work, <strong>the</strong> fatigued muscle available contractile<br />

<strong>for</strong>ce ˆ F f<br />

CE will be <strong>the</strong> product <strong>of</strong> <strong>the</strong> reposed muscle available contractile <strong>for</strong>ce ˆ F m CE with<br />

RC and <strong>the</strong> muscle activation a(t). Hence:<br />

ˆF f<br />

CE = RC(t) × ˆ F m CE × a(t) (2.13)<br />

The flow <strong>of</strong> MUs between compartments used by Xia and Law, i.e., <strong>the</strong> dynamic<br />

process <strong>of</strong> muscle fatigue, depends on <strong>the</strong> same paramenters from Liu’s model: <strong>the</strong><br />

fatigue coefficient F and <strong>the</strong> recovery coefficient R, as indicated in Equations 2.14<br />

to 2.16. These differential equations dictate <strong>the</strong> percentage <strong>of</strong> muscle fibers that are<br />

transferred between compartments and used when a <strong>for</strong>ce solicitation is made, updating<br />

<strong>the</strong> fitness state <strong>of</strong> <strong>the</strong> muscle structure. Accordingly:<br />

dMR<br />

dt<br />

dMA<br />

dt<br />

dMF<br />

dt<br />

= −C(t) + R × MF (t) (2.14)<br />

= C(t) − F × MA(t) (2.15)<br />

= F × MA(t) − R × MF (t) (2.16)<br />

where <strong>the</strong> muscle activation–deactivation driving controller C(t) (described in <strong>the</strong> next<br />

paragraph) is <strong>the</strong> term that gives <strong>the</strong> competence <strong>of</strong> this model to process <strong>the</strong> dynamics<br />

<strong>of</strong> muscle fatigue with variable ef<strong>for</strong>ts. The rate <strong>of</strong> resting MUs, in Equation 2.14, is<br />

reduced by C(t) and increased by a multiplication factor between <strong>the</strong> recovery coefficient<br />

R and <strong>the</strong> amount <strong>of</strong> fatigued MUs MF , a term that expresses <strong>the</strong> amount <strong>of</strong> fatigued<br />

fibers that recovered in <strong>the</strong> course <strong>of</strong> <strong>the</strong> muscle ef<strong>for</strong>t history. Equation 2.15 dictates<br />

<strong>the</strong> evolution <strong>of</strong> active muscle fibers and has a similar behaviour to <strong>the</strong> previous one.<br />

This term will increase with <strong>the</strong> driving controller C(t) and decrease with <strong>the</strong> number<br />

<strong>of</strong> freshly fatigued fibers (<strong>the</strong> multiplication between <strong>the</strong> fatigue coefficient F and <strong>the</strong><br />

active fibers MA). The rate <strong>of</strong> fatigued units, as indicated in Equation 2.16, is prescribed<br />

by <strong>the</strong> difference between <strong>the</strong> amount <strong>of</strong> recently fatigued fibers and <strong>the</strong> <strong>the</strong> amount <strong>of</strong><br />

fatigued fibers that just recovered.<br />

To define <strong>the</strong> driving controller C(t), stated in Equation 2.17, Xia and Law added<br />

two additional parameters LD and LR, respectively <strong>the</strong> muscle <strong>for</strong>ce development and<br />

34


2.2 Dynamics <strong>of</strong> <strong>the</strong> <strong>Muscle</strong> Tissue<br />

relaxation factors, whose values are not <strong>of</strong> major relevance since <strong>the</strong>se were added<br />

merely to ensure a good system behaviour [59] (despite <strong>of</strong> <strong>the</strong> name resemblance <strong>of</strong><br />

<strong>the</strong>se parameters with <strong>the</strong> activation and relaxation times in <strong>the</strong> Activation Dynamics<br />

model <strong>of</strong> Section 2.2.1, <strong>the</strong>re is no evidence <strong>of</strong> a correlation between <strong>the</strong>se terms).<br />

In Xia and Law’s model, C(t) is a bounded controller that depends on <strong>the</strong> relation<br />

between <strong>the</strong> compartments’ state and a target load T L, i.e., <strong>the</strong> <strong>for</strong>ce that <strong>the</strong> muscle is<br />

required to exert. In this work <strong>the</strong> target load is referred as <strong>the</strong> instantaneous solicited<br />

contractile muscle <strong>for</strong>ce F m CE , since Xia and Law assume that <strong>the</strong> neuromuscular system<br />

can produce T L [59], i.e., this term, just like F m CE , is limited by <strong>the</strong> available contractile<br />

<strong>for</strong>ce ˆ F m CE . Moreover, <strong>the</strong> model used in this work assumes that <strong>the</strong> passive element <strong>for</strong>ce<br />

contribution does not play a part in <strong>the</strong> fatigue process, <strong>the</strong>re<strong>for</strong>e it is not considered<br />

<strong>for</strong> <strong>the</strong> calculation <strong>of</strong> C(t).<br />

⎧<br />

⎪⎨ LR × (F<br />

C(t) =<br />

⎪⎩<br />

m CE − MA × ˆ F m CE ) F m ≤ MA ˆ F m CE<br />

LD × (F m CE − MA × ˆ F m CE ) MA ˆ F m CE < F m ≤ (MA + MR) ˆ F m CE<br />

LD × MR × ˆ F m CE (MA + MR) ˆ F m CE < F m<br />

(2.17)<br />

Since F , R, LD and LR are <strong>the</strong> only parameters under consideration, <strong>the</strong>n it becomes<br />

accessible to infer expressions that express <strong>the</strong>se factors by experimental data fitting,<br />

granting a relevant flexibility and applicability to this model.<br />

<strong>Muscle</strong> recruitment hierarchy<br />

All muscles are composed by a brew <strong>of</strong> different fiber types that span from slow-twitch<br />

to fast-twitch fibers. <strong>Muscle</strong>s with a high percentage <strong>of</strong> fast fibers are able to rapidly<br />

develop strong contractions, while slow fiber muscles have larger reaction times but<br />

substantial endurance. There are several reasons [53] <strong>for</strong> this contrasts: fast fibers<br />

have a bigger diameter and a higher activity <strong>of</strong> enzymes that catalyse a rapid release <strong>of</strong><br />

energy, leading to faster contractions (but in shorter periods); slow fibers contain a large<br />

amount <strong>of</strong> mitochondria and myoglobin in <strong>the</strong> sarcoplasm, inducing higher amounts <strong>of</strong><br />

available AT P and higher rate <strong>of</strong> oxygen diffusion. In <strong>the</strong>ir work, Xia and Law [59]<br />

identified three principal types <strong>of</strong> muscle fibers: slow, fast fatigue-resistant and fast<br />

fatigable.<br />

The recruitment <strong>of</strong> muscle fibers was described by Henneman [61] that stated that<br />

slow MUs are <strong>the</strong> first to be recruited, followed by fast-resistant and fast-fatigable at<br />

35


2. MUSCULOSKELETAL SYSTEM MODELLING<br />

last. This mechanism is explained by <strong>the</strong> stack structure in Figure 2.21 where every level<br />

includes an independent three-compartment fatigue sub-system indicated in Figure 2.20,<br />

since each type <strong>of</strong> muscle fiber will have a characteristic fatigue development process.<br />

Note that fast fatigue-resistant fibers are only activated when <strong>the</strong> RC <strong>of</strong> all slow fibers<br />

is fully solicited. The same happens with fast-fatigable fibers in relation to fast-fatigable<br />

ones.<br />

Figure 2.21: <strong>Muscle</strong> recruitment hierarchy pile chart [59].<br />

The importance <strong>of</strong> implementing such recruitment hierarchy model is that it allows<br />

<strong>the</strong> same model to define, in a more specific way, muscles with different characteristics<br />

and fiber constitution. Considering this hierarchy, it is possible to indicate <strong>the</strong> per-<br />

centage <strong>of</strong> each type <strong>of</strong> fiber, improving <strong>the</strong> way <strong>the</strong> model recreates real-life fatigue<br />

behaviour and approaching it towards a smaller (microscopic) scale. The muscle fatigue<br />

model becomes characterised by a non-linear behaviour, even <strong>for</strong> constant target loads.<br />

36


2.3 Discussion<br />

2.3 Discussion<br />

In this chapter, <strong>the</strong> fundamentals <strong>of</strong> muscle anatomy and physiology were introduced.<br />

Skeletal muscle has a well defined fasciculated structure and known mechanical and<br />

biochemical organisation, which voluntary contraction delivers tension that arises from<br />

action <strong>of</strong> <strong>the</strong> myosin-actin cross-bridge cycling. The contraction process, starting with<br />

action potential from <strong>the</strong> CNS, was described in order to acknowledge its microscopic<br />

functioning when introducing <strong>the</strong> ma<strong>the</strong>matical models that will simulate muscle con-<br />

traction behaviour.<br />

Three types <strong>of</strong> skeletal muscle tissue contraction dynamics were described: activa-<br />

tion, contraction and fatigue dynamics. Activation dynamics models, that describe <strong>the</strong><br />

time lag between <strong>the</strong> neural signal u(t) and corresponding muscle activation a(t), have<br />

evolved from bang-bang control models to more versatile and stable models, such as<br />

<strong>the</strong> one described. In this kind <strong>of</strong> model, it is required to solve <strong>the</strong> activation dynamics<br />

differential equations in order to obtain <strong>the</strong> neural signal u(t). These type <strong>of</strong> models<br />

are usually included when using dynamic optimization, since integration is required.<br />

The majority <strong>of</strong> <strong>the</strong> developed contraction dynamics models are based in a <strong>Hill</strong>-type<br />

tissue model. Archibald <strong>Hill</strong> [46] introduced, to <strong>the</strong> classic tissue models, a contractile<br />

element capable <strong>of</strong> developing active tension. Three important properties <strong>of</strong> muscle<br />

tissue were reported: <strong>the</strong> <strong>for</strong>ce-length relationship, <strong>the</strong> <strong>for</strong>ce-velocity relationship and<br />

activation scaling. The described contraction model, included in this work, is <strong>the</strong> one<br />

proposed by Kaplan [20] and Silva [1].<br />

<strong>Fatigue</strong> dynamics are a novelty to multibody dynamics, <strong>the</strong>re<strong>for</strong>e a simple model<br />

was considered, adapted from <strong>the</strong> work by Xia and Law [59]. This model is based in<br />

a three-compartment <strong>the</strong>ory, considering that muscle units are in one <strong>of</strong> three ideal<br />

states: activated (exerting <strong>for</strong>ce), resting (recovering <strong>the</strong> contractile capabilities) and<br />

fatigued (no muscle <strong>for</strong>ce can be delivered). The dynamics <strong>of</strong> <strong>the</strong> "transfer" <strong>of</strong> muscle<br />

units between compartments are made with basis in differential equations ruled by a<br />

recovery coefficient R, a fatigue coefficient F , a driving controller C(t) and <strong>the</strong> instan-<br />

taneous state <strong>of</strong> <strong>the</strong> muscle (MR, MA and MF ). The driving controller C(t) allows <strong>the</strong><br />

model to consider a non-constant muscle <strong>for</strong>ce history, an aspect <strong>of</strong> major importance<br />

<strong>for</strong> musculoskeletal multibody routines, which are highly non-linear. The model is com-<br />

plemented by a muscle recruitment hierarchy model that enables <strong>the</strong> same model to be<br />

37


2. MUSCULOSKELETAL SYSTEM MODELLING<br />

more descriptive in terms <strong>of</strong> muscle fiber type composition.<br />

38


Chapter 3<br />

Multibody Dynamics<br />

According to Jalón and Bayo [7], a multibody system is defined as an assembly <strong>of</strong><br />

two or more rigid bodies imperfectly connected (by what is known as kinematic pair or<br />

joint), being able to move relatively to each o<strong>the</strong>rs. Multibody routines are developed to<br />

numerically analyse three-dimensional mechanical systems that undergo large displace-<br />

ments and rotations, and are acted upon by external <strong>for</strong>ces. These routines assemble<br />

and solve <strong>the</strong> system’s equations <strong>of</strong> motion in a methodical manner [7]. This technique,<br />

along with o<strong>the</strong>r computational mechanics methodologies, multiplied its relevance in<br />

<strong>the</strong> last few decades due to <strong>the</strong> increasing capacities <strong>of</strong> modern computers, able to lead<br />

with elaborate systems.<br />

The <strong>for</strong>mulation presented in this work employs fully Cartesian (or natural) coor-<br />

dinates as introduced by Jalon and Bayo [7] and applied in Silva [1]. The system is<br />

defined as <strong>the</strong> set <strong>of</strong> Cartesian coordinates <strong>of</strong> <strong>the</strong> points and vectors that define <strong>the</strong><br />

body elements, without making use <strong>of</strong> angular variables. Natural coordinates allows<br />

<strong>the</strong> use <strong>of</strong> shared points and vectors when modelling kinematic joints, which leads to<br />

a reduction <strong>of</strong> <strong>the</strong> system’s equations. This loss <strong>of</strong> in<strong>for</strong>mation however conditions <strong>the</strong><br />

calculation <strong>of</strong> reaction <strong>for</strong>ces in joints. Silva [1] overcomes this drawback by creating an<br />

expanded system from <strong>the</strong> original, with no shared-point joints, when <strong>the</strong> calculation<br />

<strong>of</strong> <strong>the</strong>se reactions is needed.<br />

This work adapts an already existent FORTRAN multibody code, developed by<br />

Silva [1], to <strong>the</strong> models described in Chapter 2, in order to analyse biomechanical systems<br />

where skeletal muscles are considered. This chapter starts with a first introduction to<br />

<strong>the</strong> kinematics and equations <strong>of</strong> motion <strong>of</strong> multibody systems. The used <strong>for</strong>mulation<br />

39


3. MULTIBODY DYNAMICS<br />

<strong>for</strong> muscle <strong>for</strong>ces is described and <strong>the</strong> process <strong>of</strong> introduction <strong>of</strong> muscle equations in<br />

<strong>the</strong> equations <strong>of</strong> motion is explained. The implication <strong>of</strong> such action is analysed in<br />

both Inverse and Forward Dynamics paradigms. The vital optimization process that<br />

is utilised <strong>for</strong> solving <strong>the</strong> equations <strong>of</strong> motion, is described. The chapter closes with a<br />

brief discussion and conclusions about this methodology and <strong>the</strong> chosen approach <strong>for</strong><br />

muscle implementation.<br />

3.1 Kinematics<br />

The kinematic analysis <strong>of</strong> a multibody system comprises solely <strong>the</strong> geometrical as-<br />

pects <strong>of</strong> position and orientation <strong>of</strong> each body, disregarding <strong>the</strong> <strong>for</strong>ces and torques that<br />

produce <strong>the</strong> observed movement [67].<br />

Since we are working with natural coordinates, <strong>the</strong> configuration <strong>of</strong> <strong>the</strong> whole sys-<br />

tem is characterised by <strong>the</strong> Cartesian coordinates <strong>of</strong> every point and vector use in <strong>the</strong><br />

description <strong>of</strong> <strong>the</strong> model. These are arranged in <strong>the</strong> column vector <strong>of</strong> generalised coor-<br />

dinates q. For a system with n existing points and m existing vectors, <strong>the</strong> generalised<br />

coordinates are described as shown below:<br />

q = { xP1 yP1 zP1 ... xPn yPn zPn xV1 yV1 zV1 ... xVn yVn zVn } T<br />

(3.1)<br />

where x, y and z refer to <strong>the</strong> coordinates in <strong>the</strong> three Cartesian directions, and P and<br />

V symbolises a point or vector, respectively. Vector q has a total size <strong>of</strong> nc = 3(n + m)<br />

coordinates.<br />

Taking into consideration that <strong>the</strong>se coordinates show some dependencies that arise<br />

from <strong>the</strong> anatomical and dynamic properties, supplementary algebraic equations must<br />

be considered, in order to uphold <strong>the</strong> topology <strong>of</strong> <strong>the</strong> mechanical system. These equa-<br />

tions are known as kinematic constraint equations and, to ensure <strong>the</strong> validation <strong>of</strong> <strong>the</strong><br />

system status, <strong>the</strong>y must be fulfilled <strong>for</strong> every time instant in <strong>the</strong> analysis. In a multi-<br />

body system with ns scleronomic constraints (equations with no explicit dependence<br />

on <strong>the</strong> time variable) and nr rheonomic constraints (where time dependency is explicit,<br />

usually related to driver actuators), <strong>the</strong> kinematic constraint equations are held by <strong>the</strong><br />

equality <strong>of</strong> <strong>the</strong> column vector Φ to <strong>the</strong> null vector:<br />

Φ(q, t) = { Φ1(q) ... Φns(q) Φns+1(q, t) ... Φns+nr(q, t) } T = 0 T<br />

40<br />

(3.2)


3.1 Kinematics<br />

The total number <strong>of</strong> equations will <strong>the</strong>n be nh = ns + nr, where nh stands <strong>for</strong> <strong>the</strong><br />

number <strong>of</strong> holonomic constraints. Holonomic constraints are defined as constraints<br />

with an exact differential and <strong>the</strong>re<strong>for</strong>e integrable from <strong>the</strong> velocities equations. This<br />

is important <strong>for</strong> <strong>the</strong> evaluation <strong>of</strong> <strong>the</strong> respective Jacobian entries (described in <strong>the</strong> next<br />

subsection). For a detailed understanding <strong>of</strong> <strong>the</strong> characteristics <strong>of</strong> holonomic and non-<br />

holonomic constraints, please refer to Jalón and Bayo [7] or Silva [1].<br />

There are several types <strong>of</strong> constraints to be considered in <strong>the</strong> construction <strong>of</strong> Φ.<br />

To model rigid bodies, constant lengths between points that belong to <strong>the</strong> same body<br />

must be held constant, using rigid body constraints. For more complex structures,<br />

linear combination constraints are used. Joint constraints are employed to describe<br />

relative positions between components and driver constraints <strong>for</strong> motion prescription<br />

purposes. These four types <strong>of</strong> constraints are <strong>the</strong> most relevant <strong>for</strong> <strong>the</strong> developed work.<br />

For a detailed description <strong>of</strong> kinematic constraint equations <strong>the</strong> reader should refer to<br />

References [7] and [1].<br />

However, a special note should be taken regarding driver constraints. This is <strong>the</strong> type<br />

<strong>of</strong> constraints that will correlate certain equations, numerically imposing <strong>the</strong> prescribed<br />

motion. Driver constraints will play a major role in Section 3.5, where <strong>the</strong> proposed<br />

optimization methods is described, and <strong>the</strong> calculation <strong>of</strong> muscle <strong>for</strong>ces and activations<br />

will be per<strong>for</strong>med. As it will be described in that section, its importance comes as a<br />

consequence <strong>of</strong> our muscle <strong>for</strong>ces implementation in <strong>the</strong> equations <strong>of</strong> motion: it will be<br />

required to "transfer" <strong>the</strong> numerical contribution that justifies <strong>the</strong> prescribed motion<br />

from <strong>the</strong> kinematic drivers to <strong>the</strong> muscle equations.<br />

Kinematic analysis<br />

In our work, it is necessary on occasions to corroborate <strong>the</strong> consistency <strong>of</strong> prescribed<br />

kinematics making use <strong>of</strong> a kinematic analysis routine. The position, velocity and<br />

acceleration <strong>of</strong> <strong>the</strong> system’s driven elements are given in order to obtain <strong>the</strong> same<br />

physical quantities <strong>of</strong> <strong>the</strong> remaining constituents <strong>of</strong> <strong>the</strong> system. This analysis is made<br />

by solving Equation 3.2 in order to q using <strong>the</strong> first two terms <strong>of</strong> its Taylor series<br />

expansion:<br />

Φ(q, t) ∼ = Φ(qi, t) + Φq(qi)(q − qi) = 0 (3.3)<br />

41


3. MULTIBODY DYNAMICS<br />

where q corresponds to an initial approximation <strong>of</strong> <strong>the</strong> generalised coordinates <strong>for</strong> iter-<br />

ation i. Matrix Φq is <strong>the</strong> Jacobian <strong>of</strong> Φ in order to q, defined as<br />

Φq(q) = ∂Φi<br />

⎡<br />

∂Φ1<br />

⎢ ∂q1 ⎢<br />

= ⎢<br />

∂qj<br />

⎢ .<br />

⎣∂Φnh<br />

∂q1<br />

...<br />

. ..<br />

...<br />

⎤<br />

∂Φ1<br />

∂qnc ⎥<br />

. ⎥ .<br />

∂Φnh<br />

⎦<br />

∂qnc<br />

(3.4)<br />

where nc and nh are <strong>the</strong> already referred number <strong>of</strong> coordinates and number <strong>of</strong> holo-<br />

nomic constraints, respectively. The Taylor expansion in Equation 3.3 will be employed<br />

to implement <strong>the</strong> Newton-Raphson method. Since this method is iterative, <strong>the</strong> following<br />

adaptations are required, where subscripts are <strong>the</strong> number <strong>of</strong> <strong>the</strong> iteration:<br />

q = qi+1 (3.5)<br />

∆qi = qi+1 − qi (3.6)<br />

Adapting Equation 3.3 to <strong>the</strong> arrangement <strong>of</strong> Newton-Raphson method results in<br />

Φq(qi)∆qi = −Φ(qi, t) (3.7)<br />

being this equations used iteratively until ∆qi reaches a value considered negligible. At<br />

this point, <strong>the</strong> algorithm is ready to increase <strong>for</strong> <strong>the</strong> next time step.<br />

The automatic generation <strong>of</strong> <strong>the</strong> vector <strong>of</strong> constraints Φ(q, t) may occasionally lead<br />

to an overconstrained system, due to <strong>the</strong> presence <strong>of</strong> redundant constraint equations<br />

(equations that describe identical topological properties). In this case, <strong>the</strong> Jacobian<br />

matrix will have linearly dependent lines, i.e., it is rank-defficient. To overcome this<br />

aspect [15] and [7] apply <strong>the</strong> least-squares method below:<br />

Φq(qi) T Φq(qi)∆qi = −Φq(qi) T Φ(qi). (3.8)<br />

To work out <strong>the</strong> velocities <strong>of</strong> <strong>the</strong> bodies that constitute <strong>the</strong> multibody system, Equa-<br />

tion 3.2 is differentiated in time, yielding:<br />

˙Φ(q,<br />

dΦ(q, t)<br />

˙q, t) = =<br />

dt<br />

∂Φ(q, t)<br />

+<br />

∂t<br />

∂Φ(q, t) ∂q<br />

= 0 (3.9)<br />

∂q ∂t<br />

where <strong>the</strong> vector ∂Φ(q, t)/∂t corresponds to <strong>the</strong> partial derivatives <strong>of</strong> <strong>the</strong> constraints<br />

with respect to time. Knowing that <strong>the</strong> term ∂Φ(q, t)/∂q = Φq and ∂q/∂t = ˙q,<br />

Equation 3.9 can be <strong>the</strong>n rewritten as<br />

Φq ˙q = ν (3.10)<br />

42


where<br />

∂Φ(q, t)<br />

ν(t) = −<br />

∂t<br />

3.2 Equations <strong>of</strong> motion<br />

(3.11)<br />

A similar procedure is taken <strong>for</strong> <strong>the</strong> calculation <strong>of</strong> <strong>the</strong> accelerations ¨q. The derivation<br />

<strong>of</strong> Equation 3.9 in time yields:<br />

that can be rewritten as<br />

where<br />

¨Φ(q, ˙q, ¨q, t) = d ˙Φ(q, ˙q, t)<br />

dt<br />

= Φq¨q + (Φq ˙q)q ˙q − νt = 0, (3.12)<br />

Φq¨q = γ (3.13)<br />

γ(q, ˙q, t) = νt − (Φq ˙q)q ˙q (3.14)<br />

Note that Equations 3.10 and 3.13 will be used <strong>for</strong> verification <strong>of</strong> consistency <strong>of</strong> <strong>the</strong><br />

initial values in a <strong>for</strong>ward dynamics analysis as it will be described in Section 3.6.<br />

3.2 Equations <strong>of</strong> motion<br />

Since <strong>the</strong> calculation <strong>of</strong> <strong>the</strong> muscle <strong>for</strong>ces is <strong>the</strong> major goal <strong>of</strong> this work, <strong>the</strong> kinetic<br />

laws used in multobody dynamics with natural coordinates must be examined. When<br />

per<strong>for</strong>ming ei<strong>the</strong>r <strong>for</strong>ward and inverse dynamic calculations, <strong>the</strong> equations <strong>of</strong> motion <strong>of</strong><br />

a generic system must be worked out and modified. This section makes an overview <strong>of</strong><br />

some <strong>of</strong> <strong>the</strong> algebraic principles implicated in <strong>the</strong> physical laws <strong>of</strong> multibody systems.<br />

The principle <strong>of</strong> <strong>the</strong> virtual power is one <strong>of</strong> <strong>the</strong> possible approaches that can be<br />

used to obtain <strong>the</strong> equations <strong>of</strong> motion in constrained multibody systems [1]. It states<br />

that, <strong>for</strong> a determined mechanical system, <strong>the</strong> sum <strong>of</strong> <strong>the</strong> virtual power produced by<br />

<strong>the</strong> external <strong>for</strong>ces must be zero <strong>for</strong> all instants <strong>of</strong> time, i.e.,<br />

P ∗ = ˙q ∗T (M¨q − g) = 0 (3.15)<br />

where ˙q ∗ is a virtual velocity vector, M¨q are <strong>the</strong> inertial <strong>for</strong>ces (M is <strong>the</strong> global mass<br />

matrix and ¨q <strong>the</strong> vector <strong>of</strong> generalised accelerations), and g <strong>the</strong> generalised <strong>for</strong>ce vector<br />

(containing external and velocity-dependent <strong>for</strong>ces) [1].<br />

43


3. MULTIBODY DYNAMICS<br />

Vector ˙q ∗ must contain <strong>for</strong> all time steps a set <strong>of</strong> fictional velocities consistent with<br />

Equation 3.10 [1]. It belongs to <strong>the</strong> null space <strong>of</strong> <strong>the</strong> Jacobian matrix and <strong>the</strong>re<strong>for</strong>e this<br />

means that it must be orthogonal to <strong>the</strong> constraint manifold.<br />

Equation 3.15 does not take into consideration <strong>the</strong> internal reaction <strong>for</strong>ces related<br />

to <strong>the</strong> kinematic constraints since <strong>the</strong>se produce null virtual power. These <strong>for</strong>ces are<br />

developed by <strong>the</strong> system in order to assure that <strong>the</strong> system fulfils <strong>the</strong> imposed kine-<br />

matic constraints [1]. Internal <strong>for</strong>ces may be added to <strong>the</strong> equations <strong>of</strong> motion using<br />

<strong>the</strong> Lagrange Multipliers method. This method correlates each internal <strong>for</strong>ce to its as-<br />

sociated kinematic constraint, yielding nh equations to <strong>the</strong> equations <strong>of</strong> motion <strong>of</strong> <strong>the</strong><br />

mechanical system [1]. The internal constraint <strong>for</strong>ces vector g Φ is given by:<br />

gΦ = Φ T q<br />

λ (3.16)<br />

where λ is known as <strong>the</strong> vector <strong>of</strong> Lagrange multipliers that, from a physical perspec-<br />

tive, provides <strong>the</strong> magnitude <strong>of</strong> <strong>the</strong> constraint <strong>for</strong>ces, whereas <strong>the</strong> rows <strong>of</strong> <strong>the</strong> Jacobian<br />

matrix Φq correspond to <strong>the</strong> directions <strong>of</strong> <strong>the</strong>se <strong>for</strong>ces. Combining Equation 3.15 with<br />

Equation 3.16, <strong>the</strong> virtual power comes<br />

P ∗ = ˙q ∗T (M¨q − g + Φ T q<br />

λ) = 0. (3.17)<br />

According to Silva [1], it is always possible to find nc − nh arbitrarily vitual velocities<br />

and nh Lagrange multipliers (<strong>for</strong> <strong>the</strong> mentioned system with nc generalised coordinates<br />

and nh holonomic constraints) that, using 3.17, will ensure that<br />

M¨q − g + Φ T q<br />

λ = 0. (3.18)<br />

This equation embodies <strong>the</strong> equation <strong>of</strong> motion (EOM) to be solved in order to compute<br />

<strong>the</strong> system unknowns.<br />

3.3 Generic muscle <strong>for</strong>ces<br />

In order to insert a muscle model into <strong>the</strong> existing multibody routines, its data man-<br />

agement must be carefully defined and implemented. As stated in Section 2.1.1, muscle<br />

structures are spatially defined by <strong>the</strong> coordinates <strong>of</strong> <strong>the</strong>ir origin (subscript o), insertion<br />

(subscript i) and eventual via-points (subscripts v1, v2, ... , vvp, with vp as <strong>the</strong> number<br />

<strong>of</strong> via-ponts). This representation, <strong>for</strong> two muscle structures with 0 and 3 via-points,<br />

44


3.3 Generic muscle <strong>for</strong>ces<br />

both applying a <strong>for</strong>ce F m , is shown in Figure 3.1 (a) and (b) respectivelly. Note that<br />

some important geometrical assumptions are made, that do not correspond to <strong>the</strong> mus-<br />

cle anatomy <strong>of</strong> humans: muscles have rectilinear orientations, constant cross-sectional<br />

area and no wrap around structures in its via-points.<br />

(a) <strong>Model</strong> with no via-points (b) <strong>Model</strong> with 3 via-points<br />

Figure 3.1: <strong>Muscle</strong> representation <strong>for</strong> a biomechanical system.<br />

Generically, a muscle exerts in <strong>the</strong> respective application points a number <strong>of</strong> <strong>for</strong>ces<br />

nf = 2 × vp + 2. The simplest case <strong>of</strong> a muscle with no via-points (vp = 0) will exert<br />

a pair <strong>of</strong> <strong>for</strong>ces F m at both <strong>the</strong> origin and insertion points, while <strong>the</strong> case with vp = 3<br />

will have 4 pairs <strong>of</strong> opposite <strong>for</strong>ces. A muscle with via-points will be defined by more<br />

than one direction vectors that will be used to define <strong>the</strong> orientation <strong>of</strong> muscle <strong>for</strong>ces<br />

(u1, u2, ... , ud), where <strong>the</strong> number <strong>of</strong> direction vectors is d = vp + 1. The application<br />

point <strong>for</strong> each applied <strong>for</strong>ce is associated to a rigid body, where <strong>the</strong> <strong>for</strong>ce is actually<br />

being exerted. From a vectorial point-<strong>of</strong>-view, a muscle <strong>for</strong>ce with direction ud ′ and<br />

applied in a point p will be expressed as<br />

F m p = ud ′F m . (3.19)<br />

This work has a <strong>for</strong>mulation that uses <strong>for</strong>ces <strong>for</strong> <strong>the</strong> whole muscle structures implemen-<br />

tation, i.e., numerically muscles will be defined as <strong>for</strong>ces. Differently from this, Silva [1]<br />

45


3. MULTIBODY DYNAMICS<br />

developed a methodology that uses Lagrange multipliers λmd associated with muscle<br />

actuators. These multipliers were related to a <strong>for</strong>ce per unit <strong>of</strong> length (N/m, in SI<br />

units). This approach considers a constant <strong>for</strong>ce per unit length in a muscle which is a<br />

physiological assumption that requires additional pro<strong>of</strong>. The advantage <strong>of</strong> working with<br />

<strong>for</strong>ces, ra<strong>the</strong>r that <strong>for</strong>ce per unit length, is inherent to <strong>the</strong> fact that <strong>the</strong> exerted muscle<br />

<strong>for</strong>ce magnitude will be <strong>the</strong> same <strong>for</strong> different length locations in <strong>the</strong> same muscle, and<br />

will correspond to <strong>the</strong> <strong>for</strong>ce exerted by <strong>the</strong> whole muscle F m .<br />

Concentrated <strong>for</strong>ces<br />

<strong>Muscle</strong> <strong>for</strong>ces must be adapted to <strong>the</strong> multibody system. There<strong>for</strong>e, we must consider<br />

an algebraic method to bridge vectorial with generalised <strong>for</strong>ces. The way external <strong>for</strong>ces<br />

are processed in this work is <strong>the</strong> one described by Jalón and Bayo [7] and Silva [1]. Let<br />

us consider <strong>the</strong> rigid body in Figure 3.2 with orthogonal local reference frame oξηζ<br />

defined by two points i and j, and two non-coplanar vectors u and v, in an inertial<br />

reference frame 0xyz. In this rigid body, a generic <strong>for</strong>ce F m is applied in a point p, as<br />

illustrated.<br />

Figure 3.2: Application <strong>of</strong> a <strong>for</strong>ce F m to pointo p, belonging to <strong>the</strong> rigid body defined<br />

by points i and j and vectors u and v [1].<br />

46


3.3 Generic muscle <strong>for</strong>ces<br />

Now, <strong>the</strong> global coordinates <strong>of</strong> point p, represented by vector rp, are related to <strong>the</strong><br />

Cartesian coordinates <strong>of</strong> ri, rj, u and v as follows<br />

rp − ri = c1(rj − ri) + c2u + c3v (3.20)<br />

where <strong>the</strong> coefficients c1, c2 and c3 correspond to <strong>the</strong> coordinates <strong>of</strong> vector rip in <strong>the</strong><br />

three-dimensional basis <strong>for</strong>med by vectors rij, u and v. Arranging Equation 3.20 to<br />

obtain rp comes<br />

or simply:<br />

rp 3×1 = (1 − c1)I3 c1I3 c2I3 c3I3<br />

<br />

3×12<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

ri<br />

rj<br />

u<br />

v<br />

⎫<br />

⎪⎬<br />

⎪⎭<br />

12×1<br />

(3.21)<br />

rp = Cpqe. (3.22)<br />

Matrix Cp plays an important role, since it is responsible to express <strong>the</strong> Cartesian<br />

coordinates <strong>of</strong> any given point p, belonging to a rigid body e, as a linear combination<br />

<strong>of</strong> <strong>the</strong> generalised coordinates used to describe that element 1 . Since Cp depends exclu-<br />

sively <strong>of</strong> locally defined vectors, it will be independent <strong>of</strong> body motion, hence constant<br />

during <strong>the</strong> whole analysis. It also should be noticed that, since this matrix trans<strong>for</strong>ms<br />

generalised coordinates <strong>of</strong> a rigid body (arranged as 3 vectors in a column) in global<br />

coordinates <strong>of</strong> a point, its dimensions are (3 × 12) and remain constant in time.<br />

For every multibody dynamic analysis that involves any kind <strong>of</strong> external <strong>for</strong>ces,<br />

Cp must be assembled <strong>for</strong> every <strong>for</strong>ce application point. The c coefficients must be<br />

calculated and it can be done by adapting Equation 3.21 to <strong>the</strong> local frame oξηζ:<br />

(r ′ p − r ′ i) = r ′ ij u′ v<br />

⎧<br />

⎨<br />

′<br />

⎩<br />

c1<br />

c2<br />

c3<br />

⎫<br />

⎬<br />

⎭ = X′ c (3.23)<br />

with X ′ = r ′ ij u′ v ′ . This matrix always has an inverse [1] and <strong>the</strong>re<strong>for</strong>e Equa-<br />

tion 3.23 can be solved in terms <strong>of</strong> c<br />

c = X ′−1 (r ′ p − r ′ i) (3.24)<br />

1 It should be noted that when rigid bodies with complex geometry are considered, a second trans-<br />

<strong>for</strong>mation matrix V must be used. However its analysis is out <strong>of</strong> <strong>the</strong> scope <strong>of</strong> <strong>the</strong> work. Please refer to<br />

Silva [1] and Jalón and Bayo [7] <strong>for</strong> fur<strong>the</strong>r study in this subject.<br />

47


3. MULTIBODY DYNAMICS<br />

Reaching this point, we are able to express <strong>the</strong> generic muscle <strong>for</strong>ce F m p in terms <strong>of</strong><br />

an equivalent term g F m p<br />

e<br />

expressed in terms <strong>of</strong> <strong>the</strong> generalised coordinates <strong>of</strong> <strong>the</strong> rigid<br />

body in matter. It is considered that <strong>the</strong> virtual work carried out by concentrated <strong>for</strong>ce<br />

F m p and its generalised counterpart g F m p<br />

e<br />

should be <strong>the</strong> same:<br />

δW = δr T p F m p = δq T e g F m p<br />

e<br />

Using Equation 3.22 in <strong>the</strong> latter, yields that<br />

δW = δq T e C T p F m p = δq T e g F m p<br />

e<br />

(3.25)<br />

(3.26)<br />

and relating with Equation 3.19, <strong>the</strong> expression <strong>of</strong> <strong>the</strong> concentrated muscle <strong>for</strong>ce ex-<br />

pressed in <strong>the</strong> generalised coordinates that define <strong>the</strong> rigid body will be:<br />

F m<br />

ge = C T p F m p = C T p ud ′F m . (3.27)<br />

This expression plays a vital role in this muscle modelling work, since it allows us to<br />

trans<strong>for</strong>m a Cartesian representation <strong>of</strong> <strong>for</strong>ces F m p in <strong>the</strong> generalised configuration g F m p<br />

e .<br />

Never<strong>the</strong>less <strong>the</strong> latter corresponds to a physical quantity that requires additional ma-<br />

nipulation, since F m p needs to be expressed in terms <strong>of</strong> <strong>the</strong> generalised muscle <strong>for</strong>ce<br />

vector <strong>of</strong> <strong>the</strong> whole system, <strong>the</strong> column vector g F m p <strong>of</strong> size nc. In his work, Silva [1]<br />

developed natural-coordinates-specific computational routines whose function is to as-<br />

semble g F m p<br />

e into g F m p , expressing generic concentrated muscle <strong>for</strong>ces in <strong>the</strong> scope <strong>of</strong><br />

multibody equations <strong>of</strong> motion. The three different representations <strong>of</strong> muscle <strong>for</strong>ce F m<br />

are schemed in Figure 3.3.<br />

Vectorial<br />

F m p<br />

Rigid body<br />

generalised coordinates<br />

g F m p<br />

e<br />

Whole system<br />

generalised coordinates<br />

Figure 3.3: The different representations <strong>of</strong> a generic muscle <strong>for</strong>ce F m p .<br />

Since each generalised <strong>for</strong>ce vector g F m p<br />

e<br />

g F m p<br />

is specific <strong>for</strong> a certain body (<strong>the</strong> body<br />

containing <strong>the</strong> point where <strong>the</strong> <strong>for</strong>ce is applied), <strong>the</strong>n a particular muscle framed in nb<br />

bodies will have nb different g F m p<br />

e<br />

that need to be added to <strong>the</strong> equations <strong>of</strong> motion.<br />

48


3.3 Generic muscle <strong>for</strong>ces<br />

Taking <strong>the</strong> case <strong>of</strong> <strong>the</strong> muscle in Figure 3.1 (b) as example, it is arranged by three dif-<br />

ferent bodies I, II and III. There<strong>for</strong>e three independent gF m<br />

e <strong>for</strong>ce vectors (subscript<br />

p will be dropped from this point) must be considered: gI e, gII e and gIII e . Each one<br />

<strong>of</strong> <strong>the</strong>se vectors will enclose in<strong>for</strong>mation regarding <strong>the</strong> <strong>for</strong>ces that are exerted in <strong>the</strong><br />

respectively attached via-points as depicted in Figure 3.4. For instance, gF m<br />

I will be defined<br />

by <strong>for</strong>ces Fm 1 , Fm2 and Fm m<br />

3 , while gF<br />

III will only be determined by <strong>the</strong> contribution<br />

<strong>of</strong> F m 8 . In ma<strong>the</strong>matical terms and assuming that <strong>the</strong> muscle <strong>for</strong>ce F m has a constant<br />

magnitude in all muscle extension, it comes that <strong>for</strong> body I<br />

g I e = g F m 1<br />

e<br />

+ g F m 2<br />

e<br />

+ g F m 3<br />

e<br />

= C T o uov1F m + C T v1uv1oF m + C T m<br />

v1uv1v2 F<br />

(3.28)<br />

(3.29)<br />

and knowing that <strong>the</strong> direction vectors defined by any two points i and j will have<br />

opposite directions<br />

uij = −uji<br />

<strong>the</strong>n we can simplify Equation 3.29 and state that, <strong>for</strong> all ge:<br />

g I e = (C T o − C T v1 )uov1 + CTv1 uv1v2<br />

m<br />

F<br />

g II<br />

e = −C T v2uv1v2 + (CTv2 − CTv3 )uv2v3 + CTv3 uv3i<br />

m<br />

F<br />

g III<br />

e = −C T m<br />

i uv3i F<br />

(3.30)<br />

(3.31)<br />

(3.32)<br />

(3.33)<br />

Having a numerical representation <strong>of</strong> all <strong>the</strong> <strong>for</strong>ces <strong>of</strong> <strong>the</strong> muscle in Figure 3.4, it is<br />

now possible to assemble gI e, gII e and gIII e<br />

– <strong>the</strong> <strong>for</strong>ces expressed in general coordinates<br />

<strong>of</strong> <strong>the</strong> involved bodies – in <strong>the</strong> vectors <strong>of</strong> generalised muscle <strong>for</strong>ces expressed in <strong>the</strong><br />

general coordinates <strong>of</strong> <strong>the</strong> whole system g I , g II and g III . Since <strong>the</strong>se are not specific<br />

to <strong>the</strong> bodies, but to <strong>the</strong> multibody system, <strong>the</strong>y are summable, and <strong>the</strong>re<strong>for</strong>e <strong>the</strong><br />

assembling <strong>of</strong> all muscle <strong>for</strong>ces in terms <strong>of</strong> <strong>the</strong> system general coordinates gF m<br />

as<br />

F m<br />

g = gI + gII + gIII<br />

is given<br />

(3.34)<br />

Note that, as <strong>the</strong> term F m is present in all expressions as a multiplication factor, this<br />

term can be divided in <strong>the</strong> <strong>Hill</strong>-type muscle model components. Using Equation 2.4 we<br />

get<br />

49


3. MULTIBODY DYNAMICS<br />

Figure 3.4: <strong>Muscle</strong> <strong>for</strong>ce representation <strong>for</strong> <strong>the</strong> 3 via-point model in Figure 3.1.<br />

F m<br />

g<br />

F m<br />

= g<br />

F m<br />

P E + gCE (3.35)<br />

where gF m<br />

m<br />

CE and gF<br />

P E are respectively <strong>the</strong> generalised <strong>for</strong>ce vectors <strong>for</strong> <strong>the</strong> contractile and<br />

passive elements contributions. In addition, and regarding <strong>the</strong> terminology <strong>of</strong> Equa-<br />

tion 2.8, gF m<br />

can be expressed in terms <strong>of</strong> <strong>the</strong> muscle activation a:<br />

F m<br />

g<br />

F m<br />

= g<br />

F m<br />

P E + ˆg<br />

CE a (3.36)<br />

where ˆg F m<br />

CE corresponds to <strong>the</strong> generalised <strong>for</strong>ce vector <strong>of</strong> <strong>the</strong> maximum available contractile<br />

<strong>for</strong>ce, i.e., <strong>the</strong> generalised representation <strong>of</strong> ˆ F m CE . These expressions will have<br />

<strong>the</strong> utmost importance in <strong>the</strong> following sections, where <strong>the</strong> equations <strong>of</strong> motion will be<br />

reshaped in order to include our muscle <strong>for</strong>ces and calculate muscle activations.<br />

3.4 Inverse Dynamics<br />

Considering that <strong>the</strong> in<strong>for</strong>mation about a mechanical system’s motion and anthro-<br />

pometry (topology and kinematic restrictions) are available, <strong>the</strong>n it will be possible to<br />

calculate both internal and external <strong>for</strong>ces by means <strong>of</strong> inverse dynamics based routines.<br />

50


3.4 Inverse Dynamics<br />

In this terms, it is possible to infer reaction <strong>for</strong>ces and net-moments in articular joints,<br />

by non-invasive procedures [1]. This is a major asset <strong>for</strong> biomechanics, specially when<br />

a living mechanical system is being analysed. In this work, <strong>the</strong> methods considered<br />

<strong>for</strong> solving our EOM are <strong>the</strong> Lagrange multipliers and <strong>the</strong> Newton method (<strong>for</strong> more<br />

developments on o<strong>the</strong>r solutions refer to Jalón and Bayo [7]).<br />

Taking <strong>the</strong> EOM (Equation 3.18) into <strong>the</strong> paradigm <strong>of</strong> inverse dynamics, <strong>the</strong> analysis<br />

will be per<strong>for</strong>med knowing <strong>the</strong> anthropometric data (specified by <strong>the</strong> mass matrix M and<br />

<strong>the</strong> Jacobian Φq), <strong>the</strong> system motion (given by <strong>the</strong> vector <strong>of</strong> generalised accelerations<br />

¨q), and both <strong>the</strong> external <strong>for</strong>ces and velocity-dependent inertial <strong>for</strong>ces (available in g).<br />

The only unknown in this analysis, and <strong>the</strong>re<strong>for</strong>e our output, is <strong>the</strong> Lagrange multipliers<br />

column vector λ, that will provide <strong>the</strong> <strong>for</strong>ces associated with each degree-<strong>of</strong>-freedom <strong>of</strong><br />

<strong>the</strong> system. Rewriting Equation 3.18 we get<br />

Φ T q λ = g − M¨q. (3.37)<br />

This equation corresponds to a system with nc equations with nh unknowns. If <strong>the</strong><br />

system is over-constrained (nh > nc), <strong>the</strong>n Equation 3.37 will have an infinite set <strong>of</strong><br />

solutions. In his work, Silva [1] employed <strong>the</strong> minimum norm condition as a means to<br />

acquire an unique solution. In this way, <strong>the</strong> best solution will be considered to be <strong>the</strong> one<br />

orthogonal to <strong>the</strong> null-space <strong>of</strong> <strong>the</strong> matrix Φ T q [1], which ma<strong>the</strong>matically corresponds<br />

to<br />

λ = Φ T q λ ∗<br />

(3.38)<br />

where λ ∗ will enclose <strong>the</strong> unique solution <strong>for</strong> λ, when replacing Equation 3.38 in Equa-<br />

tion 3.37, since Φ T q Φq is always invertible [1].<br />

There are several methods <strong>for</strong> calculation <strong>of</strong> muscle <strong>for</strong>ces: some authors, such as<br />

Raison [17] and Ackermann [2] use inverse dynamics to determine joint torques and<br />

in a subsequent post-processing step, muscle <strong>for</strong>ces are calculated using an optimizer<br />

that relates <strong>the</strong>se joint moments with muscles’ moment arm. When <strong>the</strong> goal is to<br />

calculate muscle <strong>for</strong>ces exclusively, this method assures a simple, efficient and flexible<br />

implementation, along with excellent results. However, <strong>the</strong> values obtained <strong>for</strong> joint<br />

reactions and inner tensions will not be accurate, since muscle <strong>for</strong>ces are not taken into<br />

account in <strong>the</strong> EOM.<br />

51


3. MULTIBODY DYNAMICS<br />

Silva [1] in his work, modelled muscle actuators as constraint equations describing<br />

<strong>the</strong> muscle length development history. This method is quite similar to <strong>the</strong> one devel-<br />

oped here in terms <strong>of</strong> muscle <strong>for</strong>ces representation described in Section 3.2, but differs<br />

in <strong>the</strong> way this in<strong>for</strong>mation is added to <strong>the</strong> EOM. In Silva’s work, muscle actuators are<br />

simply considered by adding <strong>the</strong> respective constraint equations to Φ. This method-<br />

ology proved [1] to rectify <strong>the</strong> drawback <strong>of</strong> <strong>the</strong> previously described procedure, where<br />

muscle <strong>for</strong>ces are not considered <strong>for</strong> <strong>the</strong> calculation <strong>of</strong> joint reactions and inner <strong>for</strong>ces,<br />

since <strong>the</strong> system is solved taking into account <strong>the</strong> system mechanical constraints and<br />

muscle contraction dynamics. This in<strong>for</strong>mation is figured as a whole and <strong>the</strong> solution is<br />

calculated in an integrated way. Never<strong>the</strong>less this representation lacks some versatility<br />

when integrating <strong>the</strong> muscle model in both inverse and <strong>for</strong>ward dynamic (Section 3.6)<br />

approaches. In this work, muscles are represented by <strong>for</strong>ces solely as it will be described.<br />

Integrating muscle <strong>for</strong>ces<br />

The goal in this Section, is to include <strong>the</strong> referred muscle <strong>for</strong>ces in <strong>the</strong> equations <strong>of</strong> mo-<br />

tion and add muscle activations in <strong>the</strong> solutions <strong>of</strong> our problem. To do so, <strong>the</strong> Newton<br />

method from Jalón and Bayo [7] is used. Employing this approach, muscle actuators<br />

must be regarded as sets <strong>of</strong> concentrated external <strong>for</strong>ces, avoiding <strong>the</strong> constraint repre-<br />

sentation. Considering <strong>the</strong> existence <strong>of</strong> nm muscles in our model, g can be expressed<br />

as<br />

g = g ext + g M1 + g M2 + . . . + g Mnm (3.39)<br />

where g ext is <strong>the</strong> remaining external <strong>for</strong>ces (excluding muscle <strong>for</strong>ces). Making use <strong>of</strong><br />

Equation 3.36, it is possible to explicitly express passive and active contributions <strong>of</strong><br />

muscle in <strong>the</strong> generalised <strong>for</strong>ce vectors, resulting in<br />

g = g ext + ˆg M1<br />

CEaM1 M1 + gP E + . . . + ˆgMnm<br />

CE aMnm + g Mnm<br />

P E<br />

Replacing this equation in Equation 3.18, it comes that<br />

. (3.40)<br />

M¨q − (g ext + ˆg M1<br />

CEaM1 M1 + gP E + . . . + ˆgMnm<br />

CE aMnm + g Mnm<br />

P E ) + ΦTq λ = 0. (3.41)<br />

Now, <strong>the</strong> aim here is to compute <strong>the</strong> magnitude <strong>of</strong> <strong>the</strong> internal constraint <strong>for</strong>ces λ<br />

and muscle activations a M1 ...a Mnm . Reshaping our system to a suitable configuration,<br />

52


Equation 3.41 suffers a rearrangement:<br />

3.4 Inverse Dynamics<br />

Φ T q λ − (ˆg M1<br />

CEaM1 Mnm<br />

+ . . . + ˆg CE aMnm ) = g ext + g M1<br />

P E + . . . + gMnm<br />

P E − M¨q. (3.42)<br />

Since <strong>the</strong> inertial term M¨q, <strong>the</strong> passive contribution in muscle <strong>for</strong>ces g m P E<br />

and <strong>the</strong><br />

remaining external <strong>for</strong>ces g ext are independent from both λ and <strong>the</strong> muscle activations,<br />

<strong>the</strong>y can be passed to <strong>the</strong> right hand side <strong>of</strong> <strong>the</strong> EOM. It is now possible to express <strong>the</strong><br />

equation in <strong>the</strong> matrix <strong>for</strong>m as:<br />

<br />

Φ T q −ˆg M1<br />

CE<br />

. . . −ˆgMnm<br />

CE<br />

or in a more compact fashion<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

λ<br />

a M1<br />

· · ·<br />

a Mnm<br />

Φ T q −χ T λ<br />

a<br />

⎫<br />

⎪⎬<br />

⎪⎭<br />

= g ext + g M1<br />

P E + . . . + gMnm<br />

P E − M¨q (3.43)<br />

<br />

= g ext + gP E − M¨q (3.44)<br />

where χ is a matrix holding all <strong>the</strong> generalised maximum available contractile <strong>for</strong>ce<br />

vectors <strong>of</strong> all <strong>the</strong> muscles included (Equation 3.45), and a is <strong>the</strong> column vector <strong>of</strong> all<br />

muscle activations (Equation 3.46):<br />

χ =<br />

a =<br />

⎡<br />

⎣<br />

⎡<br />

⎣<br />

ˆg M1<br />

CE<br />

· · ·<br />

ˆg Mnm<br />

CE<br />

aM1 · · ·<br />

aMnm ⎤<br />

⎦ (3.45)<br />

⎤<br />

⎦ (3.46)<br />

The final configuration <strong>of</strong> <strong>the</strong> EOM with muscle <strong>for</strong>ces consists in a system <strong>of</strong> nc<br />

equations and nh + nm unknowns. This will lead to a system with infinite solutions,<br />

a fact that can be physiologically understood by <strong>the</strong> existence <strong>of</strong> muscular redundancy<br />

(already mentioned in <strong>the</strong> previous chapters) understood by <strong>the</strong> infinite number <strong>of</strong> mus-<br />

cle <strong>for</strong>ce combinations that will result in a specific motion. Optimization procedures are<br />

used to solve this system, by minimising a function that will describe <strong>the</strong> muscle system<br />

energy depletion. Fur<strong>the</strong>r details about <strong>the</strong> optimization can be found in Section 3.5.<br />

53


3. MULTIBODY DYNAMICS<br />

3.5 Optimization<br />

A musculoskeletal system is typically redundant. Even if a straight<strong>for</strong>ward adductor-<br />

abductor muscle pair is modelled, any movement <strong>of</strong> <strong>the</strong> joint that <strong>the</strong>se muscles span<br />

will have an infinite set <strong>of</strong> muscle <strong>for</strong>ce combinations that will produce it, i.e., <strong>the</strong><br />

CNS will pick one from a set <strong>of</strong> infinite muscle activation combinations to prescribe <strong>the</strong><br />

movement in question.<br />

In our inverse dynamics implementation, due to <strong>the</strong> use <strong>of</strong> <strong>the</strong> Newton approach,<br />

a redundant system will be created only by adding one muscle to span a certain joint.<br />

This is due to <strong>the</strong> fact that <strong>the</strong> EOM <strong>of</strong> <strong>the</strong> system will hold in<strong>for</strong>mation related to<br />

<strong>the</strong> kinematic driver constraint that prescribes <strong>the</strong> joint motion, as well as in<strong>for</strong>mation<br />

related to <strong>the</strong> generalised <strong>for</strong>ce term ˆgCE, associated to <strong>the</strong> included muscle. This means<br />

that <strong>the</strong>re will be two unknowns related to each one <strong>of</strong> <strong>the</strong>se terms that will constitute<br />

our control variables, respectively: a Lagrange multiplier associated with <strong>the</strong> kinematic<br />

driver λ ∗ and a muscle activation a m . There<strong>for</strong>e, in any situation that considers muscles,<br />

<strong>the</strong> number <strong>of</strong> equations will always be less than <strong>the</strong> number <strong>of</strong> unknowns in <strong>the</strong> EOM,<br />

leading to an indeterminate system <strong>of</strong> equations with an infinite set <strong>of</strong> solutions. This<br />

is a major issue in muscle modelling with multibody dynamics, i.e., to choose <strong>the</strong> best<br />

possible solution.<br />

The general optimization problem<br />

The challenge in optimization is to make <strong>the</strong> mentioned solution choice. There are sev-<br />

eral ma<strong>the</strong>matical approaches to solve this problem. A typical one is trying to even out<br />

<strong>the</strong> number <strong>of</strong> equations and unknowns ei<strong>the</strong>r by decreasing <strong>the</strong> number <strong>of</strong> unknowns<br />

(situation that will lead to loss <strong>of</strong> <strong>the</strong> system’s in<strong>for</strong>mation) or by increasing <strong>the</strong> number<br />

<strong>of</strong> equations, however it is difficult to obtain such equations). <strong>With</strong> optimization meth-<br />

ods <strong>the</strong> objective is to find, from all <strong>the</strong> solutions, <strong>the</strong> one that, regarding a particular<br />

set <strong>of</strong> constraints, minimises a stipulated objective or cost function.<br />

The control variables <strong>of</strong> our optimization problem x are <strong>the</strong> terms that are meant<br />

to be worked out, i.e., <strong>the</strong> unknowns <strong>of</strong> <strong>the</strong> inverse dynamics system, which are:<br />

x =<br />

λ<br />

a<br />

54<br />

<br />

(3.47)


3.5 Optimization<br />

The solution <strong>of</strong> x must respect <strong>the</strong> limitations imposed by <strong>the</strong> constraints <strong>of</strong> our system.<br />

The first type <strong>of</strong> constraints are <strong>the</strong> equality constraints feq, that arise from <strong>the</strong> EOM<br />

<strong>of</strong> <strong>the</strong> musculoskeletal system. Equation 3.37 defines this type <strong>of</strong> constraints.<br />

⎧<br />

⎪⎨<br />

feq =<br />

⎪⎩<br />

f1<br />

.<br />

⎫<br />

⎪⎬<br />

⎪⎭ = ΦT q −χT <br />

λ<br />

+ M¨q − (g<br />

a<br />

ext + g M1<br />

P E<br />

fnc<br />

+ . . . + gMnm<br />

P E ) = 0 (3.48)<br />

The gradient <strong>of</strong> feq will be required <strong>for</strong> <strong>the</strong> optimization routines, described in <strong>the</strong><br />

development <strong>of</strong> this section:<br />

∇f n λ a<br />

<br />

o T = Φq −χT (3.49)<br />

An important point <strong>of</strong> our implementation, and a novelty in <strong>the</strong> usual procedure<br />

<strong>of</strong> <strong>the</strong> calculation <strong>of</strong> muscle <strong>for</strong>ces, is associated to <strong>the</strong> boundaries <strong>of</strong> control variables.<br />

When muscles are not implemented and a motion is prescribed, <strong>the</strong> Lagrange multi-<br />

pliers assigned to each kinematic driver hold <strong>the</strong> <strong>for</strong>ces contribution that induces <strong>the</strong><br />

movement. Let us identify <strong>the</strong> Lagrange multipliers associated with <strong>the</strong> joint drivers<br />

that prescribe <strong>the</strong> motion <strong>of</strong> <strong>the</strong> joints spanned by <strong>the</strong> muscles by λ ∗ . When muscles<br />

are considered in <strong>the</strong> model, it is desired to eliminate that contribution from <strong>the</strong>se La-<br />

grange multipliers, shifting that value to <strong>the</strong> activations <strong>of</strong> <strong>the</strong> considered muscles. To<br />

restrict <strong>the</strong>ir contribution, <strong>the</strong>se must be kept within a bound <strong>of</strong> ε. This will be <strong>the</strong><br />

first inequality type constraints.<br />

|λ ∗ | ≤ ε (3.50)<br />

The optimizer, constrained by Equation 3.48 will ma<strong>the</strong>matically justify <strong>the</strong> motion<br />

prescription by assigning positive values <strong>for</strong> <strong>the</strong> muscle activations a, but limiting <strong>the</strong><br />

values between 0 and 1 as<br />

0 ≤ a m ≤ 1 , <strong>for</strong> m = 1, . . . , nm (3.51)<br />

The remaining existent Lagrange multipliers in <strong>the</strong> control variables, identified as λ R ,<br />

will not have any special limitations and <strong>the</strong>re<strong>for</strong>e can have any value:<br />

− ∞ ≤ λ R ≤ +∞ (3.52)<br />

This process <strong>of</strong> transferring <strong>the</strong> contributions <strong>of</strong> motion prescription from <strong>the</strong> La-<br />

grange multipliers to <strong>the</strong> muscle activations is illustrated in Figure 3.5. Considering<br />

55


3. MULTIBODY DYNAMICS<br />

<strong>the</strong> system in Figure 3.5(a), where a musculoskeletal structure with two rigid bodies<br />

and muscle m bears a load P [N] in <strong>the</strong> extremity <strong>of</strong> a vertical rigid body <strong>of</strong> length<br />

l [m]. When this model is fed to an inverse dynamics solver that calculates <strong>the</strong> muscle<br />

activation a m <strong>for</strong> a constant prescribed angle <strong>of</strong> 90 o between <strong>the</strong> two rigid bodies, <strong>the</strong><br />

correspondent kinematic driver will be assigned with <strong>the</strong> value <strong>of</strong> <strong>the</strong> angular moment<br />

needed to sustain <strong>the</strong> desired position. The value <strong>of</strong> that correspondent torque in <strong>the</strong><br />

mentioned model will be P l ey[Nm], as indicated in Figure 3.5(b). Be<strong>for</strong>e <strong>the</strong> opti-<br />

mization process, all muscle activations are 0 (zero). As mentioned, <strong>the</strong> optimizer is<br />

directed to convey <strong>the</strong> numerical influence <strong>of</strong> λ ∗ to <strong>the</strong> respective muscle activations,<br />

as depicted in Figure 3.5(c). Since all λ ∗ are kept in <strong>the</strong> interval in Equation 3.50, this<br />

allows <strong>the</strong> optimizer to become more stable, since <strong>the</strong>se terms will act moreover as a<br />

scape <strong>for</strong> <strong>the</strong> optimizer in <strong>the</strong> accomplishment <strong>of</strong> <strong>the</strong> EOM.<br />

(a) Simple mechanical system with mus-<br />

cle m and exertion <strong>of</strong> load P in an ex-<br />

tremity <strong>of</strong> a body with length l.<br />

(b) Control variables x<br />

be<strong>for</strong>e <strong>the</strong> optimization<br />

procedure.<br />

Figure 3.5: Used optimization methodology.<br />

56<br />

(c) Control variables x<br />

after <strong>the</strong> optimization<br />

procedure.


Ultimately, <strong>the</strong> general optimization problem can be <strong>the</strong>n stated as<br />

Given : x<br />

Minimise : F0(x)<br />

Subject to :<br />

⎧<br />

feq(x) = 0<br />

⎪⎨ 0 ≤ am ≤ 1<br />

−ε ≤ λ<br />

⎪⎩<br />

∗ ≤ +ε<br />

−∞ ≤ λ R ≤ +∞<br />

3.5 Optimization<br />

(3.53)<br />

The only aspect that is missing a description in <strong>the</strong> <strong>for</strong>mulated optimization problem<br />

is <strong>the</strong> cost function F0.<br />

Cost functions<br />

The cost function purpose is to reproduce <strong>the</strong> response <strong>of</strong> <strong>the</strong> CNS in terms <strong>of</strong> mus-<br />

cle recruitment and muscle activation distribution, when required to provoke a certain<br />

posture and articular motion. There is an interminable number <strong>of</strong> situations and param-<br />

eters with significance depending in <strong>the</strong> type <strong>of</strong> movement and <strong>the</strong> physical conditions<br />

inherent to <strong>the</strong> model. For instance, <strong>the</strong> gait pattern <strong>for</strong> an individual with a muscle<br />

injury will differ from <strong>the</strong> gait pattern <strong>of</strong> an healthy one. The CNS <strong>of</strong> <strong>the</strong> <strong>for</strong>mer will<br />

aim to minimise <strong>the</strong> <strong>for</strong>ce exerted in <strong>the</strong> damaged muscle, and <strong>the</strong> CNS <strong>of</strong> <strong>the</strong> latter<br />

will probably try to minimise <strong>the</strong> ef<strong>for</strong>t deployed. In addition <strong>the</strong>se functions should<br />

assure a stable and efficient computational behaviour.<br />

There are several works in this area regarding cost functions. Tsirakos [37] made an<br />

extensive work regarding cost functions in linearly constrained optimization techniques.<br />

Raison [17] in his work included muscle EMG function shape <strong>for</strong> <strong>the</strong> calculation <strong>of</strong> F0.<br />

For fur<strong>the</strong>r reading about <strong>the</strong> use <strong>of</strong> different objective functions, please refer to <strong>the</strong><br />

works by Crowninshield and Brand [19], Collins [34], Tsirakos [37] and Silva [1]. In <strong>the</strong><br />

present work, <strong>the</strong> considered cost functions are <strong>the</strong> following:<br />

1. The sum <strong>of</strong> <strong>the</strong> square <strong>of</strong> <strong>the</strong> activations:<br />

F0 =<br />

nm<br />

(a m ) 2<br />

m=1<br />

(3.54)<br />

The usage <strong>of</strong> this functions aims to minimise <strong>the</strong> muscle activations, i.e., <strong>the</strong><br />

neural ef<strong>for</strong>t required <strong>for</strong> execution <strong>of</strong> a motor task.<br />

57


3. MULTIBODY DYNAMICS<br />

2. Sum <strong>of</strong> <strong>the</strong> square <strong>of</strong> <strong>the</strong> individual contractile element <strong>for</strong>ces:<br />

F0 =<br />

nm<br />

(F m CE) 2<br />

m=1<br />

This function usage aims to minimise energy consumption amounts [1, 37].<br />

3. Sum <strong>of</strong> <strong>the</strong> cube <strong>of</strong> <strong>the</strong> individual average muscle stresses:<br />

F0 =<br />

nm<br />

(σ m CE) 3<br />

m=1<br />

(3.55)<br />

(3.56)<br />

Crowninshield and Brand [19] proposed this function, by relating muscle <strong>for</strong>ce<br />

with muscle endurance, as well as results from experimental procedures [1]. In<br />

addition, physiologically accordant results were obtained regarding <strong>the</strong> prediction<br />

muscle groups co-activation [1, 37].<br />

Computational optimization routine<br />

The optimizer used in our program is <strong>the</strong> DNCONG routine, <strong>the</strong> double precision ver-<br />

sion <strong>of</strong> <strong>the</strong> NCONG routine. It is a FORTRAN routine available in <strong>the</strong> IMSL Library<br />

developed by Visual Numerics, Inc. [68], based on a subroutine developed by Schit-<br />

tkowski [69]. The DNCONG routine was developed in order to solve a general nonlin-<br />

ear programming problem, by means <strong>of</strong> a successive quadratic programming algorithm<br />

and a user-supplied gradient [68], in this case Equation 3.49. For detailed in<strong>for</strong>mation<br />

about this method, please refer to <strong>the</strong> IMSL Library Documentation [68] and <strong>the</strong> work<br />

by Schittkowski [69].<br />

3.6 Forward Dynamics<br />

From a different perspective from <strong>the</strong> last sections, it is possible to compute <strong>the</strong><br />

dynamic reaction <strong>of</strong> a constrained biomechanical system when compelled by external<br />

<strong>for</strong>ces. This is <strong>the</strong> goal <strong>of</strong> standard <strong>for</strong>ward dynamics: to calculate <strong>the</strong> system’s motion<br />

and internal reaction <strong>for</strong>ces, given <strong>the</strong> external <strong>for</strong>ces that include, in our case, <strong>the</strong><br />

muscle <strong>for</strong>ces. The way that <strong>for</strong>ward dynamics (FD) methodology is implemented in this<br />

work involves <strong>the</strong> prescription <strong>of</strong> <strong>the</strong> vector <strong>of</strong> <strong>the</strong> muscle activations a m . Never<strong>the</strong>less<br />

58


3.6 Forward Dynamics<br />

it is possible to use this kind <strong>of</strong> analysis prescribing <strong>the</strong> system’s motion in order to<br />

withdraw both internal and external <strong>for</strong>ces. The work by Anderson and Pandy [30]<br />

explores this alternative approach.<br />

The paradigm <strong>of</strong> <strong>for</strong>ward dynamics implies different available in<strong>for</strong>mation about<br />

<strong>the</strong> system. As mentioned, <strong>the</strong> muscle activations a m are prescribed and <strong>the</strong> motion is<br />

unknown, <strong>the</strong>re<strong>for</strong>e ¨q is to be determined. This means that <strong>the</strong> generalised <strong>for</strong>ces vector<br />

g is fully known, i.e., all terms in Equation 3.39 are known, and Equation 3.18 becomes<br />

a system <strong>of</strong> nc 2 nd order ordinary differential equations (ODE) with nc + nh unknowns<br />

(that correspond to <strong>the</strong> accelerations vector ¨q and <strong>the</strong> Lagrange multipliers λ), and<br />

<strong>the</strong>re<strong>for</strong>e indeterminate. To make it determinate, <strong>the</strong> nh equations <strong>of</strong> Equation 3.13<br />

are added, resulting in <strong>the</strong> following system<br />

M Φ T q<br />

Φq 0<br />

¨q<br />

λ<br />

<br />

=<br />

g<br />

γ<br />

<br />

. (3.57)<br />

Once <strong>the</strong> accelerations are known, <strong>the</strong>se can be integrated in time in order to obtain<br />

<strong>the</strong> generalised coordinates <strong>of</strong> <strong>the</strong> system q. To do so, <strong>the</strong> initial conditions <strong>for</strong> position<br />

q0 and velocity ˙q0 must be given and are required to be consistent with <strong>the</strong> kinematic<br />

constraints <strong>of</strong> <strong>the</strong> system in analysis. This requirement is assured when Equations 3.58<br />

and 3.59 are fulfilled <strong>for</strong> <strong>the</strong> initial value <strong>of</strong> <strong>the</strong> problem.<br />

Φ(q0) = 0 (3.58)<br />

Φq ˙q0 = ν(t0) (3.59)<br />

Verifying <strong>the</strong> consistency <strong>of</strong> <strong>the</strong> initial conditions is <strong>the</strong> first step in a <strong>for</strong>ward dynam-<br />

ics analysis <strong>of</strong> a multibody system, illustrated in <strong>the</strong> flowchart <strong>of</strong> Figure 3.6. In his<br />

work, Silva [1] uses an iterative method from <strong>the</strong> work <strong>of</strong> Jalón and Bayo [7] called<br />

<strong>the</strong> Augmented Lagrange Formulation (ALF). This method consists in a penalty-type<br />

<strong>for</strong>mulation whose task is to stabilise <strong>the</strong> EOM when <strong>the</strong> Lagrange multipliers λ are<br />

removed from Equation 3.57, so it becomes a 2 nd order ODE with nc equations, instead<br />

<strong>of</strong> nc + nh. This way only <strong>the</strong> generalised accelerations vector ¨q is worked out, and <strong>the</strong><br />

Lagrange multipliers are only calculated when required.<br />

Having <strong>the</strong> accelerations <strong>of</strong> <strong>the</strong> elements <strong>of</strong> <strong>the</strong> system, <strong>the</strong> process continues by <strong>the</strong><br />

integration in time <strong>of</strong> this in<strong>for</strong>mation, so that <strong>the</strong> motion is computed. The generalised<br />

velocities and accelerations vectors ( ˙q and ¨q) are assembled in a vector ˙yt<br />

<br />

˙q<br />

˙yt =<br />

¨q<br />

59<br />

(3.60)


3. MULTIBODY DYNAMICS<br />

Figure 3.6: Direct integration algorithm flowchart <strong>for</strong> a <strong>for</strong>ward dynamics problem [1].<br />

and integrated using <strong>the</strong> direct integration method [70], a numerical process that <strong>for</strong> a<br />

time t will integrate ˙yt, obtaining yt+∆t, i.e., <strong>the</strong> vector that contains <strong>the</strong> generalised<br />

coordinates q and velocities ˙q <strong>for</strong> <strong>the</strong> following time step:<br />

<br />

q<br />

yt+∆t =<br />

˙q<br />

(3.61)<br />

<strong>With</strong> this in<strong>for</strong>mation, <strong>the</strong> cycle ends by updating <strong>the</strong> positions and velocities <strong>for</strong><br />

<strong>the</strong> next iteration, as well as <strong>the</strong> time t = t + ∆t. As elucidated in Figure 3.6, <strong>the</strong><br />

algorithm feeds <strong>the</strong> new cycle with <strong>the</strong> calculated data whose initial conditions, due to<br />

<strong>the</strong>ir nature, are not checked <strong>for</strong> consistency anymore.<br />

3.7 Discussion<br />

Some important points with respect to this work’s <strong>for</strong>mulations must be mentioned.<br />

The chosen approach incorporates <strong>the</strong> described muscle models in <strong>the</strong> kernel <strong>of</strong> multi-<br />

body <strong>for</strong>mulation. This way, <strong>the</strong> <strong>for</strong>ce network inherent to <strong>the</strong> kinetics <strong>of</strong> <strong>the</strong> mechanical<br />

system in analysis will take in consideration muscle ef<strong>for</strong>ts toge<strong>the</strong>r with <strong>the</strong> o<strong>the</strong>r ex-<br />

ternal and internal <strong>for</strong>ces.<br />

As previously mentioned in this chapter, <strong>the</strong> majority <strong>of</strong> <strong>the</strong> studies concerning <strong>the</strong><br />

calculation <strong>of</strong> redundant muscle ef<strong>for</strong>ts do not carry out this inclusion, but separate<br />

60


3.7 Discussion<br />

<strong>the</strong> analysis in two parts. A first one where, by means <strong>of</strong> inverse dynamics routines,<br />

<strong>the</strong> joint moments required to carry out an given motion are computed. A second part<br />

uses optimization techniques to predict <strong>the</strong> shought-after muscle ef<strong>for</strong>ts. This approach<br />

has <strong>the</strong> plus side <strong>of</strong> its simpler implementation and a broader range <strong>of</strong> optimization<br />

methods adapted to this situation. Never<strong>the</strong>less, <strong>the</strong> calculated reaction <strong>for</strong>ces and<br />

structure stress values in <strong>the</strong> first part will not take into consideration eventual muscle<br />

<strong>for</strong>ces. So this approach is only applicable when <strong>the</strong> objective <strong>of</strong> <strong>the</strong> whole analysis is<br />

to infer muscle <strong>for</strong>ces exclusively.<br />

The developed method on <strong>the</strong> o<strong>the</strong>r side, increases its complexity in terms <strong>of</strong> multi-<br />

body implementation, but allows <strong>the</strong> usage <strong>of</strong> a <strong>for</strong>mulation that carries <strong>the</strong> whole batch<br />

<strong>of</strong> existent <strong>for</strong>ces. Including muscle ef<strong>for</strong>ts in <strong>the</strong> system’s equations has <strong>the</strong> natural<br />

asset <strong>of</strong> correctly calculating joint reactions and rigid body stresses, since it takes into<br />

consideration <strong>the</strong> muscle <strong>for</strong>ces that may alter <strong>the</strong>ir values. This is a very important<br />

point, <strong>for</strong> instance, in cases where muscle pair co-contraction happens, a situation that<br />

leads to an intensification <strong>of</strong> joint stiffening and bone stress.<br />

Regarding <strong>the</strong> defined optimization problem, it will be proved in <strong>the</strong> next chapter<br />

that this is a well defined problem and will lead in <strong>the</strong> desired way with <strong>the</strong> computation<br />

<strong>of</strong> <strong>the</strong> <strong>the</strong> equation <strong>of</strong> motion’s solutions. Never<strong>the</strong>less, <strong>the</strong> used optimization algorithm,<br />

DNCONG, is a very sensible one. For such an intricate and highly non-linar system,<br />

it is expected to verify a hard convergence to <strong>the</strong> solution. In addition, this algorithm<br />

only guarantees <strong>the</strong> acquisition <strong>of</strong> local minima solutions. This means that it is possible<br />

to obtain solutions that do not correspond to <strong>the</strong> lowest energy combination <strong>of</strong> applied<br />

muscle <strong>for</strong>ces. However, this allows <strong>the</strong> prediction <strong>of</strong> muscle co-contraction <strong>of</strong> antagonist<br />

pair muscle groups, a situation that is actually observed in human gait, <strong>for</strong> instance.<br />

61


3. MULTIBODY DYNAMICS<br />

62


Chapter 4<br />

Biomechanical <strong>Model</strong>s<br />

The models and <strong>for</strong>mulations described in <strong>the</strong> previous chapters were implemented in <strong>the</strong><br />

Apollo program. This inclusion involved <strong>the</strong> conception <strong>of</strong> new routines incorporating<br />

a muscle <strong>Hill</strong> model, a muscle fatigue model, EOM manipulation and <strong>the</strong> usage <strong>of</strong><br />

optimization routines. In this chapter, case test examples are considered to evaluate<br />

<strong>the</strong> robustness, accuracy and efficiency <strong>of</strong> <strong>the</strong> proposed methods. The first model is a<br />

minimal upper limb musculoskeletal model with no physiological relevance, that is used<br />

to test <strong>the</strong> validation <strong>of</strong> <strong>the</strong> muscle contractile model described in Section 2.2.2. The<br />

second one is a more complex upper limb model with 7 muscles used to test <strong>the</strong> muscle<br />

fatigue model described in Section 2.2.3, with an initial test <strong>for</strong> a single immobilised<br />

muscle. The third model is a 12 muscles lower limb musculoskeletal model, used to<br />

calculate <strong>the</strong> muscle activations and <strong>for</strong>ces <strong>for</strong> a prescribed gait motion, testing <strong>the</strong><br />

validation <strong>of</strong> <strong>the</strong> muscle <strong>for</strong>ce calculation method introduced in Chapter 3.<br />

4.1 <strong>Muscle</strong> model verification<br />

After <strong>the</strong> complete implementation <strong>of</strong> <strong>the</strong> <strong>Hill</strong>-type muscle model described in Sec-<br />

tion 2.2, <strong>the</strong> accuracy <strong>of</strong> <strong>the</strong> conceived routines is evaluated. To do so, a specific<br />

mechanical model with muscles is conceived in order to evidence <strong>the</strong> relationship <strong>of</strong><br />

muscle <strong>for</strong>ce F m with muscle length L m and velocity ˙ L m , and confirm if <strong>the</strong> output<br />

values fit <strong>the</strong> model’s equations. The conceived mechanical model has no physiologi-<br />

cal relevance in both <strong>the</strong> used muscle parameters and anthropometric properties, apart<br />

63


4. BIOMECHANICAL MODELS<br />

from some slight physical resemblance to <strong>the</strong> musculoskeletal system <strong>of</strong> <strong>the</strong> elbow joint<br />

– Figure 4.1.<br />

It was required to have a situation where a certain muscle m X was fully activated<br />

while <strong>the</strong> rigid bodies vary <strong>the</strong>ir relative positions in time. This way m X spans a cer-<br />

tain range <strong>of</strong> fiber lengths and velocities while employing its full contractile capacity <strong>for</strong><br />

each state. This model is solved <strong>for</strong> a <strong>for</strong>ward dynamics analysis, where muscle activa-<br />

tions are prescribed. The relation between muscle m X ’s length and <strong>the</strong> correspondent<br />

output contractile <strong>for</strong>ce, <strong>for</strong> a constant activation a X (t) = 1, is examined. In order<br />

to control <strong>the</strong> system’s movement and <strong>the</strong>re<strong>for</strong>e <strong>the</strong> muscle’s length, a second muscle<br />

m Y was included and its variable activation a Y (t) pattern directs angle α while body<br />

1 keeps its vertical orientation still. Only <strong>the</strong> active contribution <strong>of</strong> <strong>the</strong> muscle model<br />

F X CE (Lm (t), ˙ L m (t), a(t) = 1) is tested, since it plays <strong>the</strong> most complex role in <strong>the</strong> model.<br />

The activation patterns <strong>for</strong> both <strong>the</strong> modelled muscles is illustrated in Figure 4.2 and<br />

<strong>the</strong> correspondent motion is depicted in Figure 4.3.<br />

Figure 4.1: Mechanical system with muscles m X and m Y <strong>for</strong> <strong>Hill</strong>-type muscle model<br />

verification purposes.<br />

Analysing Figure 4.3, it should be noted that <strong>the</strong> muscle in question, muscle m X ,<br />

starts per<strong>for</strong>ming an eccentric contraction increasing its length and velocity. This first<br />

phase is followed by a deceleration that will lead to a concentric contraction with a<br />

maximum <strong>of</strong> its velocity. The system halts its concentric contraction and returns to<br />

<strong>the</strong> initial position where <strong>the</strong> analysis ends with an immobilised system. <strong>With</strong> this<br />

64


4.1 <strong>Muscle</strong> model verification<br />

Figure 4.2: Activations pattern <strong>for</strong> muscles m X and m Y .<br />

description, <strong>the</strong> system should go through a set <strong>of</strong> muscle <strong>for</strong>ce values that travel in <strong>the</strong><br />

length-velocity (l ˙ l) space in <strong>the</strong> shape <strong>of</strong> a cardioid, and should lay in <strong>the</strong> surface that<br />

define <strong>the</strong> muscle <strong>for</strong>ce model, as described in Section 2.2. By plotting this surface and<br />

overlaying <strong>the</strong> muscle contractile <strong>for</strong>ce values calculated, this prediction proved to be<br />

true, as illustrated in Figure 4.4.<br />

The contractile <strong>for</strong>ce F X CE (Lm (t), ˙ L m (t), 1) in <strong>the</strong> l ˙ l space, appears as a cardioid<br />

laying in <strong>the</strong> muscle model active <strong>for</strong>ce surface <strong>for</strong> a fully activated model with <strong>the</strong> con-<br />

sidered parameters. This proved that our model is correctly implemented and responds<br />

with accuracy to <strong>the</strong> parameters imposed.<br />

65


4. BIOMECHANICAL MODELS<br />

Figure 4.3: The state <strong>of</strong> <strong>the</strong> mechanical model <strong>for</strong> different time instants when muscles<br />

m X and m Y are activated as in Figure 4.2.<br />

66


4.2 <strong>Muscle</strong> <strong>Fatigue</strong><br />

Figure 4.4: <strong>Muscle</strong> m X contractile element <strong>for</strong>ce output (black line) and possible <strong>for</strong>ces<br />

<strong>for</strong> <strong>the</strong> model (coloured surface).<br />

4.2 <strong>Muscle</strong> <strong>Fatigue</strong><br />

One <strong>of</strong> <strong>the</strong> novel aspects <strong>of</strong> this work is <strong>the</strong> couple <strong>of</strong> a fatigue model with <strong>the</strong> devel-<br />

oped <strong>Hill</strong>-type muscle contraction model and its consequent integration in a multibody<br />

dynamics system. The used fatigue model requires as input <strong>the</strong> parameters that will de-<br />

fine <strong>the</strong> dynamics associated with <strong>the</strong> three-compartment <strong>the</strong>ory described in Chapter 2,<br />

and <strong>the</strong> response in terms <strong>of</strong> muscle <strong>for</strong>ce delivery by <strong>the</strong> musculoskeletal system.<br />

At first, <strong>the</strong> muscle fatigue model is examined apart from <strong>the</strong> rest <strong>of</strong> <strong>the</strong> <strong>for</strong>mulation,<br />

without coupling it to <strong>the</strong> <strong>Hill</strong>-type muscle model. For this case, an immobilised muscle<br />

with F0 = 300N and <strong>the</strong> fatigue parameters displayed in Table 4.1, is subjected to<br />

cycles <strong>of</strong> isometric contractions to which is associated <strong>the</strong> fatigue model. It is assumed<br />

that initially all muscle units are in <strong>the</strong> resting state. The values <strong>of</strong> <strong>the</strong> <strong>for</strong>ces developed<br />

by <strong>the</strong> isometric contractions <strong>of</strong> <strong>the</strong> muscle, are obtained by imposing a target muscle<br />

<strong>for</strong>ce P to <strong>the</strong> muscle. This target <strong>for</strong>ce values are given, in Newtons, by <strong>the</strong> function<br />

proposed in Equation 4.1<br />

67


4. BIOMECHANICAL MODELS<br />

P = F0<br />

<br />

0.25 + 0.15g<br />

where g(x) is a square wave function, given by:<br />

<br />

1 sin(x) ≥ 0<br />

g(x) =<br />

−1 sin(x) < 0<br />

<br />

2πt<br />

[N] (4.1)<br />

50<br />

(4.2)<br />

The values <strong>of</strong> <strong>the</strong> <strong>for</strong>ce P <strong>for</strong> an analysis time ta = 150 are plotted in Figure 4.5 (red<br />

line).<br />

Table 4.1: <strong>Fatigue</strong> parameters used <strong>for</strong> <strong>the</strong> muscle fatigue model employed, and <strong>the</strong> same<br />

used in <strong>the</strong> work by Xia and Law [59]. The values <strong>for</strong> <strong>the</strong> F , R, LD and LR are given in<br />

units <strong>of</strong> 1/s.<br />

Composition F R LD LR<br />

S 50 % 0.01 0.002 10 10<br />

FR 25 % 0.05 0.01 10 10<br />

FF 25 % 0.1 0.02 10 10<br />

An analysis is per<strong>for</strong>med and <strong>the</strong> obtained results are shown in Figure 4.5. Let<br />

us first mind <strong>the</strong> available muscle units <strong>for</strong> <strong>for</strong>ce delivery, i.e., <strong>the</strong> residual capacity<br />

RC = MA + MR (dark blue line). It is observable that, <strong>for</strong> <strong>the</strong> time intervals where<br />

P = 0.4F0 = 120N that <strong>the</strong> available muscle <strong>for</strong>ce drops considerably and <strong>for</strong> <strong>the</strong> times<br />

steps where P = 0.1F0 = 30N <strong>the</strong> recovery process is verifiable. The opposite happens<br />

in terms <strong>of</strong> <strong>the</strong> evolution <strong>of</strong> fatigued fibers. In <strong>the</strong> periods where P is larger, it is<br />

observable that MF has a steep increase, in opposition to <strong>the</strong> remaining time intervals,<br />

where muscle recovery is observable. In response to <strong>the</strong> diminishing <strong>of</strong> RC, <strong>the</strong> muscle<br />

activation a m must increase. This has a valid correspondence to what happens in <strong>the</strong><br />

human body: since muscle units can not deliver <strong>the</strong> same levels <strong>of</strong> contraction <strong>for</strong>ces,<br />

<strong>the</strong>n <strong>the</strong> CNS must increase its level <strong>of</strong> fiber recruiting. Never<strong>the</strong>less, in <strong>the</strong> second<br />

cycle <strong>of</strong> this analysis, slightly be<strong>for</strong>e <strong>the</strong> first 60s, <strong>the</strong> available muscle contractile <strong>for</strong>ce<br />

is unable to keep <strong>the</strong> desired tonus, and <strong>the</strong> muscle <strong>for</strong>ce F m will be unfit to keep <strong>the</strong><br />

desired values <strong>of</strong> <strong>for</strong>ce. The activation a m becomes <strong>the</strong>n saturated <strong>for</strong> <strong>the</strong> periods where<br />

P = 120N and <strong>the</strong> fatigue state <strong>of</strong> <strong>the</strong> muscle will unfit it to provide <strong>the</strong> desired muscle<br />

<strong>for</strong>ce: note in Figure 4.5 that F m (red line) displays occasionally values smaller than<br />

<strong>the</strong> target <strong>for</strong>ce P (black line).<br />

68


4.2 <strong>Muscle</strong> <strong>Fatigue</strong><br />

Figure 4.5: Results obtained <strong>for</strong> three cycles <strong>of</strong> isometric contractions <strong>of</strong> <strong>the</strong> described<br />

model. The quantities MA + MR, MF and a(t) are scaled to F0.<br />

The validated fatigue model block is incorporated into <strong>the</strong> multibody routines and<br />

additional tests are carried out. To test its functionality, a simple right upper extremity<br />

model was designed with an immobilised vertically standing humerus and considering<br />

<strong>the</strong> most important muscles <strong>of</strong> <strong>the</strong> elbow joint. A graphic representation <strong>of</strong> <strong>the</strong> model is<br />

shown in Figure 4.6. Notice that a concentrated <strong>for</strong>ce P = 150N was applied at 30cm<br />

from <strong>the</strong> elbow joint.<br />

The model is defined by three rigid bodies: torso, arm and a <strong>for</strong>earm-hand complex.<br />

Table 4.2 holds <strong>the</strong> local coordinates <strong>of</strong> <strong>the</strong> points that define <strong>the</strong> referred rigid bodies<br />

<strong>of</strong> <strong>the</strong> model, while Table 4.3 shows <strong>the</strong>ir mass and inertial properties. The torso rigid-<br />

body is not considered in Table 4.3, since it was only considered to allow <strong>the</strong> insertion<br />

<strong>of</strong> <strong>the</strong> origins <strong>of</strong> some <strong>of</strong> <strong>the</strong> muscles and is considered as <strong>the</strong> inertial (fixed) ground<br />

body. The whole biomechanical system is also subjected to a constant gravitational<br />

<strong>for</strong>ce g = −9.81ez [m/s].<br />

The geometrical and physiological parameters <strong>of</strong> <strong>the</strong> considered muscle are displayed<br />

in Table 4.4. The remaining muscles that cross <strong>the</strong> elbow joint, such as <strong>the</strong> extensor<br />

carpi radialis longus or <strong>the</strong> pronator teres, are not considered in <strong>the</strong> model, since <strong>the</strong>se<br />

muscle’s functions are not related to <strong>the</strong> action in analysis.<br />

This model is tested in a Inverse Dynamics analysis perspective where a kinematic<br />

driver <strong>for</strong> <strong>the</strong> elbow joint is prescribed, imposing a constant angle <strong>of</strong> 90 o , i.e., <strong>the</strong><br />

69


4. BIOMECHANICAL MODELS<br />

Figure 4.6: Simple upper extremity model with 7 elbow muscles and constant <strong>for</strong>ce P<br />

applied at <strong>the</strong> hand. This image was conceived using <strong>the</strong> OpenSim s<strong>of</strong>tware [71]. Local<br />

reference frames only indicated in <strong>the</strong> lateral view.<br />

Table 4.2: Body index number and local coordinates <strong>of</strong> <strong>the</strong> points that define <strong>the</strong> rigid<br />

bodies <strong>for</strong> <strong>the</strong> upper extremity model.<br />

Proximal point coordinates Distal point coordinates<br />

Body bn ξ[m] η[m] ζ[m] ξ[m] η[m] ζ[m]<br />

Torso 1 0.00 0.00 0.00 0.00 0.00 0.00<br />

Arm 2 0.00 0.00 0.180496 0.00 0.00 -0.109904<br />

Forearm<br />

and hand<br />

3 0.00 0.00 0.181479 0.00 0.00 -0.118521<br />

70


4.2 <strong>Muscle</strong> <strong>Fatigue</strong><br />

Table 4.3: Inertial data <strong>for</strong> rigid bodies considered in <strong>the</strong> upper extremity model.<br />

Inertial Moments [Kg.m 2 ]<br />

bn Iξ Iη Iζ mass [Kg]<br />

2 0.014810 0.004551 0.013193 1.864572<br />

3 0.019281 0.001571 0.020062 1.534315<br />

humerus is immobilised in a vertical position and <strong>the</strong> muscles in <strong>the</strong> biomechanical<br />

model maintain <strong>the</strong> body that represents <strong>the</strong> complex radius-ulna-hand horizontal. The<br />

fatigue parameters used in this case are <strong>the</strong> ones in Table 4.5. Similarly to <strong>the</strong> previous<br />

example, <strong>the</strong>se have no correlation with any experimental results or literature values.<br />

They are only <strong>for</strong> example purposes. The analysis is per<strong>for</strong>med till <strong>the</strong> time instant<br />

when <strong>the</strong> fatigue state <strong>of</strong> <strong>the</strong> muscles caused <strong>the</strong> system to be incapable to hold <strong>the</strong><br />

load P and sustain a horizontal orientation <strong>of</strong> <strong>the</strong> <strong>for</strong>earm. The calculated muscle <strong>for</strong>ces<br />

<strong>of</strong> <strong>the</strong> contractile element F m CE<br />

<strong>for</strong> <strong>the</strong> considered muscles are illustrated in Figure 4.7.<br />

Figure 4.7: Resultant muscle <strong>for</strong>ces <strong>of</strong> <strong>the</strong> contractile element F m CE <strong>of</strong> <strong>the</strong> biomechanical<br />

model <strong>of</strong> <strong>the</strong> upper extremity, when susceptible to fatigue dynamics.<br />

The optimizer computed that, in <strong>the</strong> first instants, <strong>the</strong> position is sustained by full<br />

activation <strong>of</strong> <strong>the</strong> brachialis and brachioradialis muscles and a partial activation <strong>of</strong> <strong>the</strong><br />

long head <strong>of</strong> <strong>the</strong> biceps brachii. The first two loose <strong>the</strong> competence <strong>of</strong> maintaining<br />

<strong>the</strong> tonus, since <strong>the</strong>y are using <strong>the</strong> full capacity <strong>of</strong> <strong>the</strong> muscle, becoming importantly<br />

71


4. BIOMECHANICAL MODELS<br />

Table 4.4: Geometrical and physiological parameters <strong>for</strong> <strong>the</strong> considered muscle in <strong>the</strong><br />

upper extremity model. Considered values were taken from <strong>the</strong> work by Holzbaur et.<br />

al. [72]. Pictures reproduced with permission: "Copyright 2003-2004 University <strong>of</strong> Washington.<br />

All rights reserved including all photographs and images. No re-use, re-distribution<br />

or commercial use without prior written permission <strong>of</strong> <strong>the</strong> authors and <strong>the</strong> University <strong>of</strong><br />

Washington." [73].<br />

biceps brachii long head (BIClong)<br />

F0[N] L0[N] LT [N] α[ o ] bn ξ[m] η[m] ζ[m]<br />

624.3 0.1157 0.2723 0<br />

1<br />

1<br />

2<br />

2<br />

2<br />

2<br />

2<br />

2<br />

3<br />

-0.0392<br />

-0.0289<br />

0.0213<br />

0.0238<br />

0.0135<br />

0.0107<br />

0.0170<br />

0.0228<br />

0.0075<br />

0.0221<br />

0.0136<br />

-0.0103<br />

-0.0120<br />

-0.0014<br />

0.0017<br />

-0.0002<br />

0.0063<br />

-0.0218<br />

0.0035<br />

0.0139<br />

0.1984<br />

0.1754<br />

0.1522<br />

0.1031<br />

0.0593<br />

0.0051<br />

0.1331<br />

biceps brachii short head (BICshort)<br />

F0[N] L0[N] LT [N] α[ o ] bn ξ[m] η[m] ζ[m]<br />

435.56 0.1321 0.1923 0<br />

1<br />

1<br />

2<br />

2<br />

2<br />

3<br />

0.0047<br />

-0.0070<br />

0.0112<br />

0.0170<br />

0.0228<br />

0.0075<br />

0.0353<br />

0.0249<br />

0.0110<br />

0.0108<br />

0.0063<br />

-0.0218<br />

-0.0123<br />

-0.0400<br />

0.1047<br />

0.0593<br />

0.0051<br />

0.1331<br />

brachialis (BRA)<br />

F0[N] L0[N] LT [N] α[ o ] bn ξ[m] η[m] ζ[m]<br />

987.26 0.0858 0.0535 0<br />

72<br />

2<br />

3<br />

0.0068<br />

-0.0032<br />

0.0036<br />

-0.0009<br />

0.0066<br />

0.1576


4.2 <strong>Muscle</strong> <strong>Fatigue</strong><br />

(Table 4.4 continuation)<br />

brachioradialis (BRD)<br />

F0[N] L0[N] LT [N] α[ o ] bn ξ[m] η[m] ζ[m]<br />

261.33 0.1726 0.133 0<br />

2<br />

3<br />

3<br />

-0.0098<br />

0.03577<br />

0.0419<br />

-0.00223<br />

-0.02315<br />

-0.0224<br />

-0.019134<br />

0.054059<br />

-0.039521<br />

triceps brachii lateral head (TRIlat)<br />

F0[N] L0[N] LT [N] α[ o ] bn ξ[m] η[m] ζ[m]<br />

261.33 0.1726 0.133 0<br />

2<br />

2<br />

2<br />

2<br />

3<br />

-0.00599<br />

-0.02344<br />

-0.03184<br />

-0.01743<br />

-0.0219<br />

-0.00428<br />

-0.00928<br />

0.01217<br />

0.01208<br />

0.00078<br />

0.054036<br />

0.035216<br />

-0.045874<br />

-0.087074<br />

0.191939<br />

triceps brachii long head (TRIlong)<br />

F0[N] L0[N] LT [N] α[ o ] bn ξ[m] η[m] ζ[m]<br />

798.52 0.134 0.143 0.21<br />

1<br />

2<br />

2<br />

2<br />

3<br />

-0.05365<br />

-0.02714<br />

-0.03184<br />

-0.01743<br />

-0.0219<br />

0.02277<br />

0.00664<br />

0.01217<br />

0.01208<br />

0.00078<br />

-0.01373<br />

0.066086<br />

-0.045874<br />

-0.087074<br />

0.191939<br />

triceps brachii medial head (TRImed)<br />

F0[N] L0[N] LT [N] α[ o ] bn ξ[m] η[m] ζ[m]<br />

624.3 0.1138 0.098 0.157<br />

2<br />

2<br />

2<br />

2<br />

3<br />

-0.00599<br />

-0.02344<br />

-0.03184<br />

-0.01743<br />

-0.0219<br />

-0.00428<br />

-0.00928<br />

0.01217<br />

0.01208<br />

0.00078<br />

0.054036<br />

0.035216<br />

-0.045874<br />

-0.087074<br />

0.191939<br />

Table 4.5: <strong>Fatigue</strong> parameters used <strong>for</strong> all <strong>the</strong> muscles in <strong>the</strong> upper extremity with elbow<br />

muscles model. These have no correlation with experimental results [59]. Values given in<br />

units <strong>of</strong> 1/s.<br />

F R LD LR<br />

0.01 0.002 10 10<br />

73


4. BIOMECHANICAL MODELS<br />

fatigued. To compensate <strong>the</strong> loss in <strong>the</strong> potential <strong>for</strong> creating articular angular momen-<br />

tum <strong>of</strong> <strong>the</strong>se muscles, both long and short heads <strong>of</strong> <strong>the</strong> biceps brachii increase <strong>the</strong> <strong>for</strong>ce<br />

output, and <strong>the</strong>re<strong>for</strong>e <strong>the</strong>ir muscle activations. Note that, around <strong>the</strong> 14s <strong>of</strong> analysis,<br />

<strong>the</strong> long head reaches its maximum level <strong>of</strong> activation, leading to an increase <strong>of</strong> <strong>the</strong> rate<br />

<strong>of</strong> short head recruitment. At time = 22.4s <strong>the</strong> analysis terminates, since <strong>the</strong> horizontal<br />

position <strong>of</strong> <strong>the</strong> <strong>for</strong>earm cannot be supported by <strong>the</strong> muscle framework. In addition, it<br />

should be noted that <strong>the</strong> activation <strong>of</strong> <strong>the</strong> three heads <strong>of</strong> <strong>the</strong> triceps brachii muscle is 0<br />

(zero) <strong>for</strong> <strong>the</strong> whole analysis period, since <strong>the</strong>se work as antagonist muscles, not being<br />

able to per<strong>for</strong>m a practical action towards <strong>the</strong> bearing <strong>of</strong> load P .<br />

4.3 Human Gait<br />

A simple model <strong>of</strong> <strong>the</strong> right lower extremity with 12 muscles is conceived with <strong>the</strong><br />

purpose <strong>of</strong> testing both <strong>the</strong> practicability and accuracy <strong>of</strong> all <strong>the</strong> <strong>for</strong>mulation described<br />

in Chapter 3. The model consists in four rigid bodies: torso, thigh, shank and foot,<br />

however <strong>the</strong> torso inertial properties are again not considered <strong>for</strong> this analysis, since it is<br />

only included <strong>for</strong> <strong>the</strong> insertion <strong>of</strong> muscle origins. The model is illustrated in Figure 4.8.<br />

Tables 4.6 to 4.7 display <strong>the</strong> anthropometrical data <strong>of</strong> <strong>the</strong> alluded rigid bodies, and <strong>the</strong><br />

physiological and geometrical parameters <strong>of</strong> <strong>the</strong> considered muscles.<br />

Table 4.6: Body index number bn and local coordinates <strong>of</strong> <strong>the</strong> points that defined each<br />

rigid body considered on <strong>the</strong> lower extremity model.<br />

Proximal point coordinates Distal point coordinates<br />

Body bn ξ[m] η[m] ζ[m] ξ[m] η[m] ζ[m]<br />

Torso 1 - - - - - -<br />

Thigh 2 0.00 0.00 0.215 0.00 0.00 -0.219<br />

Shank 3 0.00 0.00 0.151 0.00 0.00 -0.288<br />

Foot 4 -0.04980 0.00 0.06882 0.07529 0.00 -0.03042<br />

74


(a) Anterior view. (b) Lateral view.<br />

4.3 Human Gait<br />

(c) Posterior view.<br />

Figure 4.8: Simple leg model with 12 muscles spanning <strong>the</strong> knee and ankle joints. This<br />

image conceived using <strong>the</strong> OpenSim s<strong>of</strong>tware [71]. Local reference frames only indicated<br />

in <strong>the</strong> lateral view.<br />

75


4. BIOMECHANICAL MODELS<br />

Table 4.7: Geometrical and physiological properties inherent to <strong>the</strong> muscles existent in<br />

<strong>the</strong> lower extremity model. Source: Reference [1].<br />

76


(continuation <strong>of</strong> Table 4.7)<br />

77<br />

4.3 Human Gait


4. BIOMECHANICAL MODELS<br />

(continuation <strong>of</strong> Table 4.7)<br />

78


(continuation <strong>of</strong> Table 4.7)<br />

4.3 Human Gait<br />

Table 4.8: Anthropometrical properties associated with <strong>the</strong> rigid bodies <strong>of</strong> <strong>the</strong> lower<br />

extremity model.<br />

bn mass [Kg]<br />

Inertial Moments [Kg.m 2 ]<br />

Iξ Iη Iζ<br />

2 4.922 0.718 7.970 4.934<br />

3 1.813 0.543 1.915 1.570<br />

4 1.182 0.129 0.128 2.569<br />

79


4. BIOMECHANICAL MODELS<br />

The gait cycle is <strong>the</strong> movement output <strong>of</strong> <strong>the</strong> lower limbs <strong>of</strong> a walking subject, and<br />

is mainly divided in two phases: <strong>the</strong> stance phase (period where <strong>the</strong> limb, in this case<br />

<strong>the</strong> right leg in this case, is in contact with <strong>the</strong> ground) and <strong>the</strong> swing phase (<strong>the</strong> period<br />

where <strong>the</strong> foot is <strong>of</strong>f <strong>the</strong> ground surface), as depicted in Figure 4.9.<br />

Figure 4.9: Scheme with relative progress <strong>of</strong> <strong>the</strong> gait cycle, indicating <strong>the</strong> stance and<br />

swing phases [74].<br />

The stance phase commences with <strong>the</strong> landing <strong>of</strong> <strong>the</strong> heel, <strong>the</strong> initial contact known<br />

as Heel Strike (HS). Starting a period <strong>of</strong> double limb support that results in a loading<br />

response stage. Around 20% <strong>of</strong> <strong>the</strong> gait cycle, <strong>the</strong> opposite limb toe leaves <strong>the</strong> ground,<br />

an instant called Opposite Toe Off (OTO). The body advances while <strong>the</strong> right lower<br />

extremity holds <strong>the</strong> weight, passing through a mid stance stage at 30% <strong>of</strong> <strong>the</strong> gait tread.<br />

The terminal stance happens when <strong>the</strong> limb arranges itself to start its swing phase, just<br />

be<strong>for</strong>e <strong>the</strong> Opposite Heel Strike (OHS). When OHS occurs (approximately at 50% <strong>of</strong><br />

<strong>the</strong> stride), a new double support period starts, and <strong>the</strong> pre swing phase takes place.<br />

The right limb starts its swing period when Toe Off (TO) happens, and <strong>the</strong> right lower<br />

extremity moves <strong>for</strong>ward passing trough a mid swing instant (that corresponds to <strong>the</strong><br />

opposite mid stance, at 80% <strong>of</strong> <strong>the</strong> cycle) and ends when <strong>the</strong> succeeding HS comes<br />

about. This completes <strong>the</strong> gait cycles.<br />

Considering an Inverse Dynamics paradigm, <strong>the</strong> positions <strong>of</strong> <strong>the</strong> mechanical system<br />

are provided. There<strong>for</strong>e, it is mandatory to feed <strong>the</strong> routines with kinematic drivers<br />

that will describe <strong>the</strong> positions <strong>of</strong> <strong>the</strong> anatomical segments in time. These positions<br />

were taken from <strong>the</strong> work by Silva [1], and a 2 nd order Butterworth (low-pass) filter<br />

was employed to eliminate <strong>the</strong> noise associated with <strong>the</strong> kinematic data. The move-<br />

80


4.3 Human Gait<br />

ment is considered to happen in <strong>the</strong> sagittal plane, similarly to <strong>the</strong> situation illustrated<br />

in Figure 4.10. Figure 4.11 shows <strong>the</strong> point numbering employed.<br />

Figure 4.10: Part <strong>of</strong> <strong>the</strong> gait cycle with four points <strong>of</strong> reference [75].<br />

Figure 4.11: Point numbering used in <strong>the</strong> model: Point 1 - Lower torso, point 2 - hip<br />

joint, point 3 - knee joint, point 4 - ankle joint, point 5 - metatarsophalangeal joint, 6 -<br />

heel position reference [75].<br />

Four kinematic different drivers are taken into account, that result in <strong>the</strong> motion<br />

<strong>of</strong> Figure 4.14:<br />

Driver 1 Trajectory driver describing <strong>the</strong> trajectory x and z <strong>of</strong> point 1 – Fig-<br />

ure 4.12.<br />

Driver 2 Rotational driver describing <strong>the</strong> angular trajectory <strong>of</strong> vector v21 with<br />

<strong>the</strong> horizontal direction. This driver has a constant value <strong>of</strong> 90 o .<br />

81


4. BIOMECHANICAL MODELS<br />

Driver 3 Rotational driver describing <strong>the</strong> angular trajectory <strong>of</strong> vector v21 with<br />

v23 – Figure 4.13(a).<br />

Driver 4 Rotational driver describing <strong>the</strong> angular trajectory <strong>of</strong> vector v32 with<br />

v34 – Figure 4.13(b).<br />

Driver 5 Rotational driver describing <strong>the</strong> angular trajectory <strong>of</strong> vector v43 with<br />

v45 – Figure 4.13(c).<br />

(a) X-axis coordinates. (b) Z-axis coordinates.<br />

Figure 4.12: Driver 1: Trajectory driver with <strong>the</strong> space coordinates in time <strong>of</strong> point 1.<br />

(a) Driver 3. (b) Driver 4. (c) Driver 5.<br />

Figure 4.13: Angular directions <strong>of</strong> Drivers 3, 4 and 5.<br />

82


4.3 Human Gait<br />

Figure 4.14: The motion obtained using <strong>the</strong> prescribed kinematic drivers <strong>for</strong> <strong>the</strong> model<br />

<strong>of</strong> <strong>the</strong> lower extremity.<br />

83


4. BIOMECHANICAL MODELS<br />

Since this gait model involves an interaction with <strong>the</strong> ground, <strong>the</strong> reaction <strong>for</strong>ces<br />

resultant from foot-ground contact are included in our biomechanical system as external<br />

<strong>for</strong>ces, i.e., added to vector gext. The evolution <strong>of</strong> <strong>the</strong> reaction <strong>for</strong>ce vector in time is<br />

illustrated in Figure 4.15 and its application point coordinates (or centre <strong>of</strong> pressure)<br />

curve in Figure 4.16. These values are obtained using a <strong>for</strong>ce plat<strong>for</strong>m synchronised<br />

with <strong>the</strong> camera acquisition settings [1] and provided to <strong>the</strong> Inverse Dynamic routines.<br />

Figure 4.15: Reaction <strong>for</strong>ce vector components, acquired from a <strong>for</strong>ce plat<strong>for</strong>m.<br />

(a) X-axis coordinate. (b) Y-axis coordinate. (c) Z-axis coordinate.<br />

Figure 4.16: Ground reaction <strong>for</strong>ce application point coordinates, or centre <strong>of</strong> pressure<br />

curve.<br />

The developed routines processed <strong>the</strong> model in<strong>for</strong>mation and <strong>the</strong> prescribed data,<br />

resulting in <strong>the</strong> muscle activation patters <strong>of</strong> Figure 4.17. In this figure, muscles are set<br />

84


4.3 Human Gait<br />

toge<strong>the</strong>r in functional groups. The total muscle <strong>for</strong>ce <strong>of</strong> each muscle in every functional<br />

group is summed and plotted in Figure 4.18.<br />

Now, examining Figure 4.18 and bearing in mind <strong>the</strong> gait phases <strong>of</strong> Figure 4.9, let us<br />

identify <strong>the</strong> <strong>for</strong>ce contributions <strong>of</strong> muscles <strong>for</strong> <strong>the</strong> resulting dynamic patterns <strong>of</strong> <strong>the</strong> right<br />

leg. The analysis start with <strong>the</strong> HS (all gait terms refer to <strong>the</strong> right leg), <strong>the</strong> beginning<br />

<strong>of</strong> <strong>the</strong> stance phase, where three groups <strong>of</strong> muscles are active. The Hamstrings are<br />

solicited to contribute <strong>for</strong> <strong>the</strong> weight transfer that happens on passage from <strong>the</strong> single<br />

support <strong>of</strong> <strong>the</strong> left leg to <strong>the</strong> double support period, supplying stability <strong>for</strong> <strong>the</strong> body<br />

foundations.<br />

During <strong>the</strong> same period, <strong>the</strong> ankle dorsiflexors are contracted to control <strong>the</strong> position<br />

<strong>of</strong> <strong>the</strong> heel <strong>for</strong> <strong>the</strong> initial contact, in order to allow <strong>the</strong> inferior surface <strong>of</strong> <strong>the</strong> calcaneus to<br />

sustain <strong>the</strong> weight acceptance at <strong>the</strong> HS. In addition, a co-contraction <strong>of</strong> <strong>the</strong> antagonist<br />

pair <strong>for</strong>med by <strong>the</strong> flexors group and <strong>the</strong> dorsiflexors group <strong>of</strong> <strong>the</strong> ankle joint is verified<br />

<strong>for</strong> joint stabilisation. From an optimization point-<strong>of</strong>-view, this co-contraction is clearly<br />

a local minima <strong>of</strong> <strong>the</strong> cost function, since <strong>the</strong>re is an excess <strong>of</strong> energy expenditure in<br />

terms <strong>of</strong> torque output. Never<strong>the</strong>less, this is something that resembles what happens in<br />

fact on human gait [1]. In order to decrease <strong>the</strong> steep variation <strong>of</strong> ankle joint reaction<br />

in <strong>the</strong> instant <strong>of</strong> HS, <strong>the</strong> CNS commands a co-contraction <strong>of</strong> <strong>the</strong>se muscles, increasing<br />

<strong>the</strong> stiffness <strong>of</strong> <strong>the</strong> ankle joint be<strong>for</strong>e and during ground contact.<br />

This first weight acceptance and loading response phases are verified till somewhat<br />

after time = 0.1s (8.3%) <strong>of</strong> <strong>the</strong> analysis, by <strong>the</strong> time that OTO occurs and <strong>the</strong> weight<br />

<strong>of</strong> <strong>the</strong> whole body structure is transferred to <strong>the</strong> right limb. This means that <strong>the</strong> body<br />

structure is in single support. There<strong>for</strong>e, <strong>the</strong> Triceps surae group comes into picture,<br />

to per<strong>for</strong>m a slight extension <strong>of</strong> <strong>the</strong> knee joint, while <strong>the</strong> leg bears most <strong>of</strong> <strong>the</strong> body<br />

weight, precipitating an initial period <strong>of</strong> <strong>for</strong>ward body propulsion, by <strong>the</strong> time <strong>of</strong> <strong>the</strong><br />

mid stance.<br />

After <strong>the</strong> mid stance, <strong>the</strong> ankle flexors start to develop a considerable muscle <strong>for</strong>ce<br />

magnitude, swivelling <strong>the</strong> lower leg around <strong>the</strong> ankle joint and inclining <strong>the</strong> body.<br />

Simultaneously, <strong>the</strong> area <strong>of</strong> <strong>the</strong> foot contacting with <strong>the</strong> floor diminishes and <strong>the</strong> body<br />

weight is gradually shifted to <strong>the</strong> <strong>for</strong>efoot region. At this time, around t = 0.5s (41.5%),<br />

<strong>the</strong> ankle flexors muscle group <strong>for</strong>ces peak controlling <strong>the</strong> body impulse that is provided<br />

in <strong>the</strong> stance phase. This stage corresponds to <strong>the</strong> terminal stance, when <strong>the</strong> OHS<br />

happens, by t = 0.6s (50%).<br />

85


4. BIOMECHANICAL MODELS<br />

(a) Knee joint flexor muscles activations.<br />

(b) Knee joint extensor muscles activations.<br />

(c) Ankle joint flexor and extensor (Tibialis anterior) muscles activa-<br />

tions.<br />

Figure 4.17: <strong>Muscle</strong> activation patterns obtained from <strong>the</strong> solution <strong>of</strong> <strong>the</strong> EOM.<br />

86


4.3 Human Gait<br />

Figure 4.18: Resultant total muscle <strong>for</strong>ces <strong>of</strong> each considered muscle group in <strong>the</strong> lower<br />

leg extremity model.<br />

The stance phase ends with a pre-swing stage, when double support comes about<br />

again, corresponding to <strong>the</strong> drop in <strong>the</strong> angle flexor group <strong>for</strong>ces that coincides with<br />

<strong>the</strong> TO. Now, <strong>the</strong> right limb starts its swing phase by progressively recruiting <strong>the</strong><br />

fibers <strong>of</strong> <strong>the</strong> Triceps surae group, throughout <strong>the</strong> initial and mid swing stages. At <strong>the</strong><br />

same time, <strong>the</strong> Tibialis anterior is slightly contracted, providing a dorsiflexion <strong>of</strong> <strong>the</strong><br />

foot that will prepare it <strong>for</strong> <strong>the</strong> following HS. In addition, by <strong>the</strong> time <strong>of</strong> <strong>the</strong> terminal<br />

swing and similarly to <strong>the</strong> beginning <strong>of</strong> <strong>the</strong> gait analysis, <strong>the</strong> ankle flexors are recruited<br />

resulting in <strong>the</strong> mentioned co-contraction. The knee flexors also contract, preparing <strong>the</strong><br />

musculoskeletal framework <strong>of</strong> <strong>the</strong> right side <strong>for</strong> a new double support period.<br />

87


4. BIOMECHANICAL MODELS<br />

4.4 Discussion<br />

Contraction dynamics model<br />

An important point <strong>of</strong> our implementation is to determine we<strong>the</strong>r <strong>the</strong> inclusion <strong>of</strong> <strong>the</strong><br />

contraction dynamics models is justified or not. A static optimization analysis <strong>for</strong> <strong>the</strong><br />

human gait musculoskeletal system where such model is not included, was pre<strong>for</strong>med.<br />

This means that <strong>the</strong> <strong>for</strong>ce-length-velocity properties were ignored, i.e., each muscle m ′<br />

available contractile <strong>for</strong>ce ˆ F m CE was equal to <strong>the</strong> maximum isometrical <strong>for</strong>ce F m 0<br />

passive element <strong>for</strong>ce was considered <strong>for</strong> <strong>the</strong> whole analysis.<br />

ˆF m′<br />

CE = F m′<br />

0<br />

and no<br />

(4.3)<br />

F m′<br />

P E = 0 (4.4)<br />

The obtained muscle <strong>for</strong>ce results <strong>for</strong> <strong>the</strong> non-physiological situation proved to be<br />

fairly similar to <strong>the</strong> ones acquired in Section 4.3. This is evidenced in <strong>the</strong> example de-<br />

picted in Figure 4.19, where <strong>the</strong> total muscle <strong>for</strong>ce <strong>of</strong> <strong>the</strong> <strong>the</strong> tibialis anterior muscle was<br />

calculated and compared with <strong>the</strong> results obtained <strong>for</strong> <strong>the</strong> same muscle in Section 4.3.<br />

Figure 4.19: Comparison between <strong>the</strong> calculated total <strong>for</strong>ce <strong>of</strong> <strong>the</strong> tibialis anterior muscle<br />

considering (Physiological case) and not considering (Non-physiological case) a contraction<br />

dynamics model.<br />

88


4.4 Discussion<br />

The similarity in muscle <strong>for</strong>ce results was discussed in <strong>the</strong> work by Anderson and<br />

Pandy [30], where it is stated that using <strong>the</strong> time-independent per<strong>for</strong>mance criterions<br />

described in Section 3.5 in a normal gait situation, <strong>the</strong> contraction model can be ne-<br />

gleted, since muscles pre<strong>for</strong>m below <strong>the</strong> intrinsic physiological limits [30]. In terms <strong>of</strong><br />

muscle activations, <strong>the</strong> results differ in more relevant manner. The value <strong>of</strong> <strong>the</strong> available<br />

contractile <strong>for</strong>ce ˆ FCE <strong>for</strong> <strong>the</strong> physiological case will vary throughout <strong>the</strong> analysis, hav-<br />

ing values different from F0 depending in <strong>the</strong> length and velocity <strong>of</strong> <strong>the</strong> muscle. Apart<br />

from <strong>the</strong> effect <strong>of</strong> <strong>the</strong> passive element <strong>for</strong>ce in <strong>the</strong> results, <strong>the</strong>re is an additional scale<br />

factor in <strong>the</strong> activations <strong>of</strong> <strong>the</strong> tibialis anterior <strong>for</strong> <strong>the</strong> two cases, due to <strong>the</strong> mentioned<br />

differences in <strong>the</strong> values <strong>of</strong> ˆ FCE. When <strong>the</strong> contraction dynamics model is considered,<br />

<strong>the</strong> system becomes more conservative, since it accounts <strong>for</strong> a muscular system which<br />

competence to develop muscle <strong>for</strong>ces is not <strong>the</strong> same <strong>for</strong> different muscle lengths and<br />

velocities, leading to situations were larger activations may be needed to pre<strong>for</strong>m a given<br />

muscle <strong>for</strong>ce.<br />

Figure 4.20: Comparison between <strong>the</strong> calculated activation <strong>of</strong> <strong>the</strong> tibialis anterior muscle<br />

considering (Physiological case) and not considering (Non-physiological case) a contraction<br />

dynamics model.<br />

The non-inclusion <strong>of</strong> muscle contraction dynamics may not lead to significant result<br />

differences when predicting muscle <strong>for</strong>ces [30], but <strong>the</strong> model’s importance becomes<br />

apparent when it is intended to acquire muscle activations or when <strong>the</strong> prescribed motion<br />

gives rise to situations where muscles work near <strong>the</strong> physiological limits (ei<strong>the</strong>r when<br />

89


4. BIOMECHANICAL MODELS<br />

large muscle <strong>for</strong>ces are produced, or muscles work at lengths/velocities considerably<br />

distant from L0/ ˙ L0).<br />

<strong>Muscle</strong> fatigue model<br />

The muscle fatigue model was successfully implemented in <strong>the</strong> multibody dynamics<br />

routines. The whole <strong>for</strong>mulation <strong>of</strong> this model proved to be well suited and easily<br />

integrated in <strong>the</strong> latter, and compatible with <strong>the</strong> considered contraction model. In<br />

addition, such a fatigue model gains an uplift in its applicability when included in a<br />

multibody dynamics system, since its usage becomes more accessible <strong>for</strong> situation where<br />

complex musculoskeletal systems with intricate kinematic patterns are to be analysed.<br />

Never<strong>the</strong>less, it should be noted that, in <strong>the</strong> employed fatigue model, <strong>the</strong>re is an<br />

inaccuracy in <strong>the</strong> used fatigue parameters F , R, LD and LR. These were random values<br />

with no correlation with any experimental results, leading to <strong>the</strong> acquisition <strong>of</strong> mus-<br />

cle fatigue-recuperation patterns that does not correspond necessarily to physiological<br />

values. In <strong>the</strong> analysis <strong>of</strong> <strong>the</strong> upper extremity model, <strong>the</strong> subject with those muscle<br />

physiological and fatigue parameters could only bear <strong>the</strong> 150N load <strong>for</strong> less than 25<br />

seconds. The analysis can not be associated to any specific real-life situation, however<br />

by carrying out experimental procedures, in order to estimate factual values <strong>the</strong> muscle<br />

contraction and fatigue parameters, a good prediction system should be <strong>for</strong>mulated.<br />

Ano<strong>the</strong>r very important point regarding <strong>the</strong> implementation <strong>of</strong> <strong>the</strong> fatigue model is<br />

<strong>the</strong> way <strong>the</strong> fatigue components state is calculated. The most accurate procedure <strong>for</strong><br />

<strong>the</strong> computation <strong>of</strong> <strong>the</strong> compartments progress from a time instant t to t + ∆t would be<br />

to integrate <strong>of</strong> <strong>the</strong> differential equations described by Equations 2.14 to 2.16 in order to<br />

correctly update those values. However, this approach conditions <strong>the</strong> efficiency <strong>of</strong> <strong>the</strong><br />

programming routines, since it would imply a significant increase <strong>of</strong> <strong>the</strong> analysis time<br />

and <strong>the</strong> programming code intricacy. The used solution was <strong>the</strong> difference quotient,<br />

i.e., <strong>for</strong> a state M with derivative (dM(t)/dt) <strong>for</strong> time instant t, its updated value <strong>for</strong><br />

t + ∆t is given as<br />

M(t + ∆t) = M(t) + dM<br />

(t) × ∆t. (4.5)<br />

dt<br />

This means that <strong>the</strong>re is an error factor associated with <strong>the</strong> loss <strong>of</strong> precision <strong>for</strong> <strong>the</strong><br />

calculation <strong>of</strong> M in <strong>the</strong> time intervals between t and t+∆t. This error is fairly noticeable<br />

90


4.4 Discussion<br />

in <strong>the</strong> situations when <strong>the</strong> term dM/dt has a large value, eventually leading to a bigger<br />

variation <strong>of</strong> M than it was supposed.<br />

When operating <strong>the</strong> implemented multibody routines associated with <strong>the</strong> fatigue<br />

model, <strong>the</strong> user is entitled to provide a value <strong>for</strong> a time variation thresholds ∆ttresh.<br />

If <strong>the</strong> time step frequency <strong>of</strong> <strong>the</strong> multibody analysis is larger than 1/∆tthresh, <strong>the</strong>n a<br />

state updating pre-processing is made, running Equation 4.5 with ∆t = ∆ttresh through<br />

consequent time intervals, in order to reduce <strong>the</strong> error factor.<br />

Human gait muscle activations<br />

The used computational language, FORTRAN, is characterised as a high-level language,<br />

never<strong>the</strong>less having a low level when compared with o<strong>the</strong>r more modern computing lan-<br />

guages. This leads to a difficult implementation <strong>of</strong> <strong>the</strong> routines and code comprehension,<br />

but substantially efficient in terms <strong>of</strong> algebraic calculations and matrix handling. All<br />

<strong>the</strong> executed analysis per<strong>for</strong>med, even in reasonably complicated musculoskeletal sys-<br />

tems, were computed in intervals <strong>of</strong> few seconds. In addition, <strong>the</strong> possibility <strong>of</strong> working<br />

with double precision variables supplies an extra precision on <strong>the</strong> numerical procedures.<br />

However, some erratic per<strong>for</strong>mances from <strong>the</strong> optimizer DNCONG were verified.<br />

This lead to a difficult interaction and result acquisition, since this optimization routine<br />

is very sensible and only allows five iterations to search <strong>for</strong> a minimum. Fur<strong>the</strong>rmore,<br />

<strong>the</strong> algorithm searches <strong>for</strong> local minima. This means that a particular configuration <strong>of</strong><br />

muscle <strong>for</strong>ces, is not necessarily <strong>the</strong> one that minimises globally <strong>the</strong> cost function, i.e.,<br />

it might not be <strong>the</strong> muscle <strong>for</strong>ce combination that reduces <strong>the</strong> most <strong>the</strong> physiological<br />

cost. This situation is evident in <strong>the</strong> results <strong>of</strong> Figure 4.7 where <strong>the</strong>re is a co-contraction<br />

<strong>of</strong> <strong>the</strong> antagonist pair <strong>for</strong>med by <strong>the</strong> ankle joint flexors and dorsiflexors. As mentioned<br />

in Section 4.3, this co-contraction situation is intelligible in a physiological point-<strong>of</strong>-<br />

view. Never<strong>the</strong>less, <strong>the</strong> optimizer’s objective is to minimise <strong>the</strong> cost function and a<br />

antagonist pair co-contraction is certainly not <strong>the</strong> lowest energetic configuration.<br />

Ano<strong>the</strong>r verified drawback from <strong>the</strong> optimizer is related to <strong>the</strong> method response to<br />

<strong>the</strong> cost function applied. There are some basic cost functions, such as Equation 3.54,<br />

that lead to situations where <strong>the</strong> optimizer could not find any solution <strong>for</strong> <strong>the</strong> system,<br />

terminating <strong>the</strong> analysis be<strong>for</strong>e <strong>the</strong> final time step. This may suggest that some cost<br />

functions exhibit a bolder physiological nature, when compared to o<strong>the</strong>rs. The cost<br />

91


4. BIOMECHANICAL MODELS<br />

function in Equation 3.56 is indicated as <strong>the</strong> one that better relates to what is cogitated<br />

by <strong>the</strong> CNS.<br />

Considering <strong>the</strong> pattern <strong>of</strong> muscle activations obtained in Figure 4.8, <strong>the</strong> author<br />

concludes that ano<strong>the</strong>r type <strong>of</strong> values should be obtained. Empirically, <strong>for</strong> a simple<br />

gait cycle, it is not expected to have some <strong>of</strong> <strong>the</strong> main muscles in our musculoskeletal<br />

structure developing a full activation condition. Some studies [17] indicate that, even<br />

<strong>for</strong> actions were maximum <strong>for</strong>ce is required, <strong>the</strong> muscle activations <strong>of</strong> <strong>the</strong> used subjects<br />

in those studies never came close to <strong>the</strong> maximum value. This means that, in voluntary<br />

contractions, muscle <strong>for</strong>ces are never close to <strong>the</strong> maximum isometric <strong>for</strong>ce F0. Two pos-<br />

sible reasons <strong>for</strong> <strong>the</strong> calculation <strong>of</strong> <strong>the</strong>se high values include <strong>the</strong> usage <strong>of</strong> an out-<strong>of</strong>-date<br />

F0 database and <strong>the</strong> misplace <strong>of</strong> <strong>the</strong> muscles’ points, leading to respective diminutive<br />

moments arm (with a smaller moment arm, a muscle need to develop larger <strong>for</strong>ces in<br />

order to produce a certain joint torque). These points may invalidate <strong>the</strong> relationship<br />

that exists between <strong>the</strong> contractile <strong>for</strong>ce F m CE and <strong>the</strong> muscle activation am terms <strong>of</strong> our<br />

muscle model in Section 2.2.2.<br />

Ano<strong>the</strong>r relevant point, is <strong>the</strong> fact that not all muscles existent in <strong>the</strong> lower limb were<br />

included. The absence <strong>of</strong> those muscles can make all <strong>the</strong> difference, even if those muscle<br />

are not recruited. A muscle that has 0 (zero) activation throughout <strong>the</strong> whole analysis,<br />

can still make a difference by exerting a <strong>for</strong>ces related to <strong>the</strong> passive contribution <strong>of</strong> our<br />

model. Over that, <strong>the</strong> hip muscles where not considered. These muscles play a major<br />

role in <strong>the</strong> gait cycle, and its addition to <strong>the</strong> model is likely to change to some extent<br />

<strong>the</strong> obtained activation patterns and muscle <strong>for</strong>ces.<br />

In addition to all <strong>the</strong> mentioned topics that may justify unfeasible values is <strong>the</strong> fact<br />

that no tendon dynamics were considered. The muscle physiological parameters used in<br />

Tables 4.7 were taken from a database available at Reference [76] that were developed<br />

<strong>for</strong> multibody models that consider <strong>the</strong> dynamics <strong>of</strong> tendon structures, with variations<br />

<strong>of</strong> length and an associated tendon tension. For instance, <strong>the</strong> upper arm model muscle<br />

parameters, taken from <strong>the</strong> work <strong>of</strong> Holzbaur et. al. [72] and used in routines developed<br />

by Delp et. al. [71], where <strong>the</strong> tendon dynamics consider length variation in <strong>the</strong> triceps<br />

long head ranging from 147.59mm to 147.71mm that result in a tendon <strong>for</strong>ce variation<br />

from 765N to 798N. The biceps brachii short head has larger tendon <strong>for</strong>ce variations<br />

(around 295N) <strong>for</strong> length variations <strong>of</strong> 2mm. This means that an important accuracy<br />

level is lost from <strong>the</strong> contribution <strong>of</strong> tendon dynamics.<br />

92


Chapter 5<br />

Conclusions and Future<br />

<strong>Development</strong>s<br />

5.1 Conclusions<br />

The <strong>for</strong>mulation <strong>of</strong> multibody system dynamics constitutes undoubtedly a powerful<br />

method to numerically evaluate three-dimensional mechanical systems that undergo<br />

large displacements and rotations, and are acted upon by external <strong>for</strong>ces. Every single<br />

tested biomechanical model involved muscle tissue modelling and was processed in a<br />

personal computer laptop, with each execution never passing 10s <strong>of</strong> analysis time.<br />

The developed ma<strong>the</strong>matical tissue models regarding <strong>the</strong> muscle structure, suc-<br />

cessfully simulated its physiological behaviour, accounting <strong>for</strong> <strong>the</strong> inherent voluntary<br />

contractile properties and <strong>the</strong> dynamics <strong>of</strong> muscle fatigue. The first model, concerning<br />

contraction dynamics, based on <strong>the</strong> macroscopic <strong>for</strong>ce-length-velocity relationships, has<br />

proved by its considerable applicability that simulates with precision <strong>the</strong> process <strong>of</strong> <strong>for</strong>ce<br />

production <strong>of</strong> muscles. Never<strong>the</strong>less, it was proven in Section 4.4 and in <strong>the</strong> work by<br />

Anderson and Pandy [30] that <strong>the</strong> non-inclusion <strong>of</strong> muscle contraction dynamics, when<br />

time-independent per<strong>for</strong>mance criterions are used, may not lead to significant result<br />

differences when predicting muscle <strong>for</strong>ces, when muscles are required to work below <strong>the</strong><br />

physiological limits. Such type <strong>of</strong> model turns to be fundamental when it is intended<br />

to acquire muscle activations or when <strong>the</strong> physiological restrictions <strong>of</strong> muscles are to be<br />

considered.<br />

93


5. CONCLUSIONS AND FUTURE DEVELOPMENTS<br />

The implementation <strong>of</strong> muscle fatigue models in multibody dynamics routines hap-<br />

pen to be successful. The whole <strong>for</strong>mulation <strong>of</strong> this model proved to be well suited and<br />

easily integrated in <strong>the</strong> multibody system routines, and compatible with <strong>the</strong> considered<br />

contraction model. The study <strong>of</strong> muscle fatigue using versatile ma<strong>the</strong>matical models<br />

in multibody system dynamics methodologies, gains a special motivation due to <strong>the</strong><br />

fact that <strong>the</strong>se can be easily applied in complex musculoskeletal systems with intricate<br />

kinematic patterns are to be analysed. In this work, <strong>the</strong> usage <strong>of</strong> <strong>the</strong> difference quotient<br />

<strong>for</strong> <strong>the</strong> calculation <strong>of</strong> <strong>the</strong> MR, MA and MF compartments state induced <strong>the</strong> presence<br />

<strong>of</strong> an error factor. This is not relevant when <strong>the</strong> rate <strong>of</strong> change <strong>of</strong> <strong>the</strong> compartments’<br />

state dM/dt is relatively small compared with <strong>the</strong> analysis time step. In addition, <strong>the</strong><br />

small number <strong>of</strong> parameters <strong>of</strong> <strong>the</strong> model facilitates <strong>the</strong> fitting <strong>of</strong> experimental data to<br />

<strong>the</strong> fatigue model. This same model may consider muscle fiber recruitment hierarchy,<br />

allowing <strong>the</strong> same model to precisely define muscles with different characteristics and<br />

fiber constitution.<br />

Our inverse dynamics approach consisted in <strong>the</strong> adaptation <strong>of</strong> <strong>the</strong> equations <strong>of</strong> mo-<br />

tion to <strong>the</strong> muscle <strong>for</strong>mulation using <strong>the</strong> Newton’s method [1, 7]. <strong>With</strong> this methodol-<br />

ogy, it is possible to calculate in <strong>the</strong> same analysis rigid body internal reactions, joint<br />

reactions, and muscle and activations, with <strong>the</strong> nuisance <strong>of</strong> increasing <strong>the</strong> complexity<br />

<strong>of</strong> its implementation and solving process. The developed <strong>for</strong>mulation was successfully<br />

included in <strong>the</strong> multibody routines, has observed in <strong>the</strong> results obtained in Chapter 4.<br />

The optimization procedure, described in Section 3.5, proved to solve <strong>the</strong> developed<br />

models in an accurate and efficient manner. The used optimizer DNCONG revealed<br />

to be very sensible, searching solely <strong>for</strong> local minima. The detection <strong>of</strong> local minima<br />

can lead to solutions that do not correspond to <strong>the</strong> configuration <strong>of</strong> muscle contractions<br />

with minimum cost.<br />

Static optimization itself only accounts <strong>for</strong> instantaneous per<strong>for</strong>mance criteria, that<br />

may not simulate in <strong>the</strong> best way <strong>the</strong> behaviour <strong>of</strong> <strong>the</strong> CNS when choosing a muscle<br />

activation set. According to Ackermann [2], <strong>the</strong>se instantaneous functions are unable<br />

to accurately describe <strong>the</strong> key per<strong>for</strong>mance criterion: <strong>the</strong> total energy expenditure. In<br />

addition, static optimization seams to operate with a single muscle even when, <strong>for</strong> a<br />

specific motion, has multiple muscles available able to exert <strong>the</strong> desired torque. These<br />

points suggest that o<strong>the</strong>r approaches, such as dynamic optimization [32], EMG-driven<br />

94


5.2 Future <strong>Development</strong>s<br />

cost functions [17] or <strong>the</strong> methodologies proposed by Ackermann [2], should be consid-<br />

ered <strong>for</strong> <strong>the</strong> calculation <strong>of</strong> individual muscle <strong>for</strong>ces and activations. Static optimization<br />

should be <strong>the</strong> favoured approach when it is intended to calculate <strong>the</strong> <strong>for</strong>ce exerted by<br />

muscle groups.<br />

5.2 Future <strong>Development</strong>s<br />

Bearing in mind <strong>the</strong> discussed points in <strong>the</strong> previous chapters, <strong>the</strong>re are some modules<br />

in this work that can be refined and several o<strong>the</strong>rs that are viable to be added. <strong>Muscle</strong><br />

modelling has a wide range <strong>of</strong> applications and <strong>for</strong>mulations, but due to <strong>the</strong> limited<br />

time frame available <strong>for</strong> this <strong>the</strong>sis, some <strong>of</strong> <strong>the</strong>se were left aside.<br />

<strong>Fatigue</strong> model and fatigue parameters<br />

Firstly, <strong>the</strong> used fatigue model is relatively straight<strong>for</strong>ward one, and such characteristic<br />

was essential <strong>for</strong> its choice. No studies were found to date that included muscle fatigue<br />

models in a multibody dynamics system, so <strong>the</strong> intention <strong>of</strong> keeping it simple is un-<br />

derstandable. O<strong>the</strong>r similar fatigue models coupled with experimental trials, such as<br />

<strong>the</strong> one in Ma et. al. [77], should be considered and <strong>the</strong> parameters values analysed.<br />

An engrossing procedure in <strong>the</strong> following <strong>of</strong> this fatigue model, would be to do some<br />

experimental trials, in order to test <strong>the</strong> correlation <strong>of</strong> <strong>the</strong> fatigue parameters with actual<br />

physiological situations, and infer <strong>the</strong> model’s validation.<br />

Tendon dynamics model<br />

One <strong>of</strong> <strong>the</strong> models lacking in this work, is <strong>the</strong> tendon dynamics model. <strong>Muscle</strong>s and<br />

tendons exist combined in nature, and tendons play a vital role in <strong>the</strong> foundations<br />

<strong>of</strong> musculoskeletal biomechanics. Tendon models convey an additional computational<br />

charge, and is far more complex that muscle modeling [42]. Never<strong>the</strong>less, as mentioned<br />

in Section 4.4 tendons play a major role in musculotendon <strong>for</strong>ces, since <strong>the</strong>se have<br />

associated considerable tensile <strong>for</strong>ce variations <strong>for</strong> small length differences. It would be<br />

important to consider such a component, in order to accurately obtain proper muscle<br />

fiber <strong>for</strong>ces and activations.<br />

95


5. CONCLUSIONS AND FUTURE DEVELOPMENTS<br />

Activation dynamics model<br />

An important additional functionality would be <strong>the</strong> determination <strong>of</strong> a neural signal u(t)<br />

by means <strong>of</strong> an activation dynamics model. As introduced in Section 2.2.1, <strong>the</strong> time<br />

lag between a neural signal u(t) and its corresponding muscle activation a(t) is ruled<br />

by a first order ordinary diferential equation. The activation dynamics must <strong>the</strong>n be<br />

inverted [2], with <strong>the</strong> purpose <strong>of</strong> infering <strong>the</strong> value <strong>of</strong> <strong>the</strong> control signal, by calculating<br />

˙a m (t), and solve Equation 2.1 in order to u(t). Additionally, <strong>the</strong> optimization problem<br />

must consider <strong>the</strong> bounds <strong>of</strong> this signal, hence <strong>the</strong> following restriction<br />

0 ≤ u m ≤ 1 , <strong>for</strong> m = 1, . . . , nm. (5.1)<br />

Ackermann [2] refers to this <strong>for</strong>mulation as Modified Static optimization (MSO). In<br />

scope <strong>of</strong> this work, this method could be implemented by adding nonlinear constraints<br />

to <strong>the</strong> individual activations a m , as additional upper and lower bounds <strong>for</strong> a specific<br />

time instant, to ensure that <strong>the</strong> calculated muscle activations are compatible with <strong>the</strong><br />

activation and contraction dynamics. There are some drawbacks however. Similarly to<br />

what happened in <strong>the</strong> fatigue model, <strong>the</strong> way <strong>the</strong> differential equation is worked out<br />

plays a vital role to <strong>the</strong> accuracy <strong>of</strong> <strong>the</strong> model. A numerical integration procedure could<br />

increase <strong>the</strong> order <strong>of</strong> <strong>the</strong> computational times considerably. On top <strong>of</strong> that, Equation 5.1<br />

represent an additional constraint to our optimization problem, increasing <strong>the</strong> sensitivity<br />

<strong>of</strong> <strong>the</strong> optimizer.<br />

A solid global optimizer<br />

The prime fragility <strong>of</strong> <strong>the</strong> developed FORTRAN routines is most probably <strong>the</strong> used op-<br />

timizer. DNCONG is a very fast, however very sensible algorithm. When DNCONG is<br />

close to <strong>the</strong> global solution, it reaches it rapidly and with sharp precision. Never<strong>the</strong>less,<br />

it is quite demanding to find <strong>the</strong> region <strong>of</strong> <strong>the</strong> global minima. DNCONG calculates a<br />

local-minima and, when far from any solution, <strong>the</strong> algorithm simply cannot converge.<br />

A possible future development, regarding <strong>the</strong> precision <strong>of</strong> <strong>the</strong> optimizer, would be to<br />

use a genetic algorithm <strong>for</strong> <strong>the</strong> computation <strong>of</strong> an initial guess and give this guess as<br />

an input to DNCONG, letting DNCONG converge to <strong>the</strong> global minima. Ano<strong>the</strong>r ap-<br />

proach, and this one to a greater extent similar to physiological criteria, would be to<br />

use electromyography signals to drive <strong>the</strong> obtained <strong>the</strong> muscle activation results, i.e.,<br />

96


5.2 Future <strong>Development</strong>s<br />

to use and EMG-driven models to estimate muscle <strong>for</strong>ces. Several authors used this<br />

approach [17, 78–80].<br />

97


5. CONCLUSIONS AND FUTURE DEVELOPMENTS<br />

98


Appendix A<br />

Apollo – <strong>Hill</strong>-type muscles manual<br />

In <strong>the</strong> Apollo s<strong>of</strong>tware [1], <strong>Hill</strong>-type muscles are defined in both <strong>the</strong> .mdl and .sml<br />

files. In <strong>the</strong> .mdl file, <strong>the</strong> user specifies <strong>the</strong> geometrical and physiological aspects <strong>of</strong> <strong>the</strong><br />

different muscles. In <strong>the</strong> .sml file, <strong>the</strong> user provides in<strong>for</strong>mation regarding <strong>the</strong> type <strong>of</strong><br />

analysis that is intended to be done. In this appendix, templates <strong>of</strong> <strong>the</strong> relevant parts<br />

<strong>of</strong> <strong>the</strong>se files will be made inside frame boxes. The compulsory keywords are written<br />

in monospace Courier New and <strong>the</strong> numeral parameters to be provided are shown in<br />

emphasised text.<br />

A.1 MDL File<br />

The first step to include <strong>Hill</strong>-type muscles in an Apollo analysis, is to state <strong>the</strong> number<br />

<strong>of</strong> muscles that will be modelled N total<br />

muscle , in <strong>the</strong> 20th parameter <strong>of</strong> <strong>the</strong> generic *MAIN<br />

PARAMETERS keyword.<br />

A positive value <strong>of</strong> N total<br />

muscle<br />

*MAIN PARAMETERS<br />

· · · N total<br />

muscle<br />

will in<strong>for</strong>m <strong>the</strong> s<strong>of</strong>tware that muscle in<strong>for</strong>mation data<br />

will be provided subsequently in <strong>the</strong> same file. The designated keyword that encloses<br />

<strong>the</strong> in<strong>for</strong>mation about <strong>the</strong> intrinsic properties <strong>of</strong> muscle will be *MUSCLE PARAMETERS.<br />

99


A. APOLLO – HILL-TYPE MUSCLES MANUAL<br />

*MUSCLE PARAMETERS<br />

FATIGUE<br />

OptF T<br />

MUSCLE=Nmuscle<br />

L0<br />

˙L0<br />

˙Ltendon<br />

F0<br />

αpenn<br />

amin<br />

amax<br />

Nmp<br />

P 1 number P 1 ξ P 1 η P 1 z P 1 body<br />

P 2 number P 2 ξ P 2 η P 2 ζ P 2 body<br />

· · ·<br />

P Nmp<br />

number<br />

Nmp<br />

Pξ P Nmp<br />

η<br />

P Nmp<br />

ζ<br />

[F R LD LR]<br />

A brief description <strong>of</strong> each parameter is made:<br />

P Nmp<br />

body<br />

OptF T The user starts by indicating whe<strong>the</strong>r muscle fatigue is to be considered or not.<br />

OptF T = 0 <strong>Muscle</strong> fatigue is not considererd in <strong>the</strong> analysis.<br />

OptF T = 1 <strong>Muscle</strong> fatigue is considererd in <strong>the</strong> analysis. In this case, <strong>the</strong> fatigue<br />

parameters F , R, LD and LR must be provided in <strong>the</strong> last line <strong>of</strong> <strong>the</strong> *MUSCLE<br />

PARAMETERS section.<br />

L0 The resting length.<br />

˙L0 The maximum contractile velocity.<br />

˙Ltendon The tendon length, assumed to be constant.<br />

F0 Maximum contractile <strong>for</strong>ce.<br />

αpenn Pennation angle.<br />

100


amin <strong>Muscle</strong> activation lower bound (equal to 0, in <strong>the</strong>ory)<br />

amax <strong>Muscle</strong> activation upper bound (equal to 1, in <strong>the</strong>ory)<br />

A.2 Simulation file<br />

Nmp Number <strong>of</strong> points that define <strong>the</strong> muscle’s geometry. For every point n, <strong>the</strong> fol-<br />

lowing five parameters must be provided:<br />

P n number<br />

P n body<br />

Index number <strong>of</strong> point n.<br />

Number <strong>of</strong> <strong>the</strong> body where point n is attached.<br />

P n x Coordinate ξ in <strong>the</strong> local reference frame oξηζ.<br />

P n η Coordinate η in <strong>the</strong> local reference frame oξηζ.<br />

P n ζ<br />

Coordinate ζ in <strong>the</strong> local reference frame oξηζ.<br />

A.2 Simulation file<br />

In <strong>the</strong> simulation file, different entries are expected <strong>for</strong> <strong>the</strong> type <strong>of</strong> analysis to be made.<br />

In <strong>the</strong> keyword *MUSCLE ANALYSIS TYPE, <strong>the</strong> user defines what is meant to be calcu-<br />

lated.<br />

*MUSCLE ANALYSIS TYPE<br />

type<br />

The parameter type has two possible regular expressions:<br />

MOTION When it is intended to calculate <strong>the</strong> motion resultant from <strong>the</strong> prescribed acti-<br />

vations.<br />

ACTIVATIONS When <strong>the</strong> user aims to calculate <strong>the</strong> muscle activations that account <strong>for</strong><br />

<strong>the</strong> prescribed motion.<br />

If <strong>the</strong> indicated analysis type is MOTION, <strong>the</strong> s<strong>of</strong>tware solves <strong>the</strong> equations <strong>of</strong> motion<br />

in order to provide <strong>the</strong> kinematic response to a patter <strong>of</strong> muscle activation in time.<br />

The user must <strong>the</strong>n provide <strong>the</strong> file name with <strong>the</strong> in<strong>for</strong>mation <strong>for</strong> <strong>the</strong> activations <strong>of</strong><br />

<strong>the</strong> different muscles in <strong>the</strong> model. This is made employing <strong>the</strong> *ACTIVATIONS FILE<br />

keyword.<br />

*ACTIVATIONS FILE<br />

[filename].act<br />

101


A. APOLLO – HILL-TYPE MUSCLES MANUAL<br />

The indicated file must be a tab-separated text file, where <strong>the</strong> in<strong>for</strong>mation is disposed<br />

in <strong>the</strong> following manner:<br />

nt<br />

t0<br />

t1<br />

· · ·<br />

tn<br />

N total<br />

muscle<br />

actm1 t0<br />

actm1 t1<br />

act m1<br />

tn<br />

actm2 t0 · · · actm t0<br />

actm2 t1 · · · actm t1<br />

act m2<br />

tn · · · act m tn<br />

Where <strong>the</strong> value act m tn corresponds to <strong>the</strong> activation <strong>of</strong> muscle m <strong>for</strong> time instant tn. nt<br />

is <strong>the</strong> number <strong>of</strong> provided time instants and N total<br />

muscle is <strong>the</strong> number <strong>of</strong> muscles that are<br />

displayed in <strong>the</strong> activation file.<br />

If <strong>the</strong> user selects ACTIVATIONS as <strong>the</strong> analysis type, <strong>the</strong> muscle activations are<br />

calculated, regarding <strong>the</strong> kinematic data available in <strong>the</strong> driver files [1]. The user is<br />

entitled to indicate as follows <strong>the</strong> kinematic drivers to be labeled as Lagrange multipliers<br />

λ ∗ , described in Section 3.5:<br />

*BOUNDED DRIVERS<br />

Nbd<br />

εbd<br />

dr1 dr2 · · · drn<br />

where Nbd drivers, identified with <strong>the</strong> numbers indicated in <strong>the</strong> list <strong>of</strong> <strong>the</strong> third line, are<br />

bounded by <strong>the</strong> optimizer, restricting <strong>the</strong>ir values to be inside <strong>the</strong> interval [−εbd; εbd].<br />

102


Appendix B<br />

MHILL Data Visualizer manual<br />

An Apollo analysis where muscles are considered produces an output file with exten-<br />

sion .mh <strong>for</strong> each modelled muscle in a folder MHILL_BIN. This output files store data<br />

related to each muscle <strong>for</strong> different time instants <strong>of</strong> <strong>the</strong> analysis, data such as Length,<br />

Velocity, Activation, Contractile Force, etc. To better analyse this data, a small appli-<br />

cation developed in <strong>the</strong> GUI (Graphical User Interface) <strong>of</strong> MATLAB R○ . The general<br />

aspect <strong>of</strong> this s<strong>of</strong>tware is shown in Figure B.1.<br />

Figure B.1: General aspect <strong>of</strong> <strong>the</strong> MHILL Data Visualizer.<br />

103


B. MHILL DATA VISUALIZER MANUAL<br />

The main window has two chart areas (identified in Figure B.1 as A and B) and<br />

three operation boxes. Chart area A displays a 3D representation <strong>of</strong> <strong>the</strong> considered<br />

muscles, and B is <strong>the</strong> chart area where <strong>the</strong> evolution <strong>of</strong> muscle data in time is shown.<br />

The different operation sections are depicted in Figure B.2. The file box, where <strong>the</strong><br />

user is able to manage <strong>the</strong> muscle data files to be processed, is displayed in Figure B.2(a).<br />

Two buttons are available, in order to include and remove <strong>the</strong> muscle data files to be<br />

displayed, respectively <strong>the</strong> Add Files and <strong>the</strong> Delete Files buttons. The included files<br />

are displayed in <strong>the</strong> list box bellow.<br />

(a) File box. (b) Data box.<br />

(c) Time box.<br />

Figure B.2: Operation boxes <strong>of</strong> <strong>the</strong> MHILL Data Visualizer.<br />

The data box, where <strong>the</strong> options regarding <strong>the</strong> manipulation <strong>of</strong> <strong>the</strong> data may be<br />

found, is illustrated in Figure B.2(b). A pop-up menu is available to chose which data<br />

is to be displayed from a set <strong>of</strong> options: Length, Velocity, Activation, Total Force, Force<br />

CE/Activation (identified in this <strong>the</strong>sis as <strong>the</strong> available contractile <strong>for</strong>ce ˆ FCE), Force<br />

CE, Force PE). In order to plot <strong>the</strong> data, <strong>the</strong> user uses <strong>the</strong> Plot! <strong>the</strong> button. To save<br />

<strong>the</strong> plot or to export all <strong>the</strong> data to an Micros<strong>of</strong>t R○ Excel file, two options are available,<br />

respectively <strong>the</strong> Save Plot and <strong>the</strong> Export to Excel buttons. The main window may be<br />

cleared, to its initial <strong>for</strong>m (Figure B.1), by pressing <strong>the</strong> Reset button.<br />

In <strong>the</strong> time box, it is possible to select <strong>the</strong> time instant to be analysed, ei<strong>the</strong>r by set-<br />

ting its value in seconds in a text box or by means <strong>of</strong> a slider, as shown in Figure B.2(c).<br />

104


Exemple <strong>of</strong> operation<br />

The first step in <strong>the</strong> usage <strong>of</strong> this tool, is <strong>the</strong> inclusion <strong>of</strong> <strong>the</strong> muscle data files, by<br />

selecting <strong>the</strong> Add Files button <strong>of</strong> Figure B.2(a). The user selects <strong>the</strong> muscle files to be<br />

included in <strong>the</strong> analysis <strong>for</strong> <strong>the</strong> folder system, as in Figure B.3.<br />

Figure B.3: Files addition <strong>for</strong> <strong>the</strong> folder system.<br />

By clicking Push!, <strong>the</strong> results are plotted in both <strong>the</strong> chart areas. In Figure B.4,<br />

<strong>the</strong> spatial representation and <strong>the</strong> chart <strong>of</strong> <strong>the</strong> Total Force curves <strong>for</strong> <strong>the</strong> muscles in <strong>the</strong><br />

lower limb example <strong>of</strong> Section 4.3 is shown.<br />

Figure B.4: Example <strong>of</strong> data plots.<br />

105


B. MHILL DATA VISUALIZER MANUAL<br />

Using <strong>the</strong> pop-up menu <strong>of</strong> Figure B.2(b), it is possible to chose a different data to<br />

be plotter. In <strong>the</strong> example <strong>of</strong> Figure B.5, <strong>the</strong> activations <strong>of</strong> <strong>the</strong> muscles are selected. By<br />

pressing Plot! after this selection, <strong>the</strong> muscle activations are exhibited (Figure B.6).<br />

Figure B.5: Selection <strong>of</strong> muscle data to be plotted.<br />

Figure B.6: Plotting <strong>of</strong> muscle activations.<br />

If <strong>the</strong> data chart are looks messy, <strong>the</strong> user is able to remove some <strong>of</strong> <strong>the</strong> represented<br />

muscles, by deleting <strong>the</strong>m from <strong>the</strong> file box and pressing Plot!. It is possible to make a<br />

selection <strong>of</strong> <strong>the</strong> files to be excluded when using <strong>the</strong> Delete Files option (Figure B.7).<br />

106


Figure B.7: Removing some <strong>of</strong> <strong>the</strong> included files.<br />

To use <strong>the</strong> time box <strong>of</strong> Figure B.2(c), <strong>the</strong> user may opt to write <strong>the</strong> desired time<br />

value in <strong>the</strong> text box or by using <strong>the</strong> slider. In Figure B.8(a) is noticeable that <strong>the</strong> new<br />

time instant will lead to an update in <strong>the</strong> geometrical representation. This time instant<br />

will be identified in <strong>the</strong> data chart by a coloured plot <strong>of</strong> <strong>the</strong> instants between t = 0 and<br />

<strong>the</strong> selected time instant, as shown in Figure B.8(b).<br />

(a) Using <strong>the</strong> time slider. (b) The plotting colors identify in <strong>the</strong> chart, <strong>the</strong> selected time in-<br />

stant.<br />

Figure B.8: Time instant selection.<br />

107


B. MHILL DATA VISUALIZER MANUAL<br />

108


References<br />

[1] Miguel P T Silva. Human motion analysis using multibody dynamics and opti-<br />

mization tools. PhD Thesis, Departamento de Engenharia Mecânica - Instituto<br />

Superior Técnico, pages 1–347, Feb 2004.<br />

[2] Marko Ackermann. Dynamics and energetics <strong>of</strong> walking with pros<strong>the</strong>ses. PhD<br />

Thesis, Department <strong>of</strong> Mechanical Engineering, University <strong>of</strong> Stuttgart, 2007.<br />

[3] H Fujie, G Livesay, and S Woo. The use <strong>of</strong> a universal <strong>for</strong>ce-moment sensor to de-<br />

termine in-situ <strong>for</strong>ces in ligaments: A new methodology. Journal <strong>of</strong> Biomechanical<br />

Engineering, 117(1):7, 1995.<br />

[4] Benno M Nigg and Walter Herzog. Biomechanics <strong>of</strong> <strong>the</strong> musculo-skeletal systems.<br />

John Wiley & Sons Inc, Dec 1994.<br />

[5] P E Nikravesh and H A Attia. Construction <strong>of</strong> <strong>the</strong> equations <strong>of</strong> motion <strong>for</strong> multi-<br />

body dynamics using point and joint coordinates. Computer-Aided Analysis <strong>of</strong><br />

Rigid and Flexible Mechanical Systems, NATO ASI, Series E: Applied Sciences,<br />

268:31–60, 1994.<br />

[6] Werner Schiehlen. Multibody system dynamics: Roots and perspectives. Multibody<br />

Syst Dyn, 1(2):149–188, 1997.<br />

[7] Javier Garcia Jalón and Eduardo Bayo. Kinematic and dynamic simulation <strong>of</strong><br />

multibody systems: The real time challenge. Springer-Verlag, 1994.<br />

[8] J Michael. Modern pros<strong>the</strong>tic knee mechanisms. Clinical orthopaedics and related<br />

research, 361:39–47, 1999.<br />

109


REFERENCES<br />

[9] T Schmalz, S Blumentritt, and R Jarasch. Energy expenditure and biomechanical<br />

characteristics <strong>of</strong> lower limb amputee gait: The influence <strong>of</strong> pros<strong>the</strong>tic alignment<br />

and different pros<strong>the</strong>tic components. Gait & Posture, 16(2):255–263, 2002.<br />

[10] C Raasch, Felix E Zajac, B Ma, and W Levine. <strong>Muscle</strong> coordination <strong>of</strong> maximum-<br />

speed pedaling. Journal <strong>of</strong> Biomechanics, 30(6):595–602, 1997.<br />

[11] G Venture, K Yamane, and Y Nakamura. Identifying musculo-tendon parameters <strong>of</strong><br />

human body based on <strong>the</strong> musculo-skeletal dynamics computation and hill-stroeve<br />

muscle model. IEEE-RAS Int. Conf. on Humanoid Robots, pages 351–356, 2005.<br />

[12] J Rasmussen, M Damsgaard, ST Christensen, and E Surma. Design optimization<br />

with respect to ergonomic properties. Structural and Multidisciplinary Optimiza-<br />

tion, 24:89–97, 2002.<br />

[13] AJ van den Bogert, L Read, and BM Nigg. A method <strong>for</strong> inverse dynamic analysis<br />

using accelerometry. Journal <strong>of</strong> Biomechanics, 29(7):949–954, 1996.<br />

[14] I Kingma, M de Looze, and H Toussaint. Validation <strong>of</strong> a full body 3-d dynamic<br />

linked segment model. Human Movement Science, 15:833–860, 1996.<br />

[15] Miguel P T Silva and Jorge Ambrosio. Kinematic data consistency in <strong>the</strong> inverse<br />

dynamic analysis <strong>of</strong> biomechanical systems. Multibody Syst Dyn, 8(2):219–239,<br />

2002.<br />

[16] Marcelo Epstein and Walter Herzog. Theoretical models <strong>of</strong> skeletal muscle: Bio-<br />

logical and ma<strong>the</strong>matical considerations. Wiley, 1998.<br />

[17] Maxime Raison. On <strong>the</strong> quantification <strong>of</strong> joint and muscle ef<strong>for</strong>ts in <strong>the</strong> human<br />

body during motion. PhD Thesis, École Polytechnique de Louvain, Université<br />

catholique de Louvain, 2009.<br />

[18] G Tad Yamaguchi. Dynamic modeling <strong>of</strong> musculoskeletal motion: A vectorized<br />

approach <strong>for</strong> biomechanical analysis in three dimensions. Springer, pages 1–257,<br />

2005.<br />

[19] D Crowninshield and Richard A Brand. A physiologically based criterion <strong>for</strong> muscle<br />

<strong>for</strong>ce predictions on locomotion. Bulletin- Hospital <strong>for</strong> Joint Diseases, 43(2):164–<br />

170, 1983.<br />

110


REFERENCES<br />

[20] Mat<strong>the</strong>w L Kaplan. Efficient optimal control <strong>of</strong> large-scale biomechanical systems.<br />

PhD dissertation, Stan<strong>for</strong>d University, 2000.<br />

[21] D A Winter. Biomechanics and motor control <strong>of</strong> human movement. John Wiley &<br />

Sons, pages 1–384, 1990.<br />

[22] D A Winter. Biomechanics and motor control <strong>of</strong> human gait: Normal, elderly and<br />

pathological. Waterloo Biomechanics, 1991.<br />

[23] G T Yamaguchi, DW Moran, and J Si. A computationally efficient method <strong>for</strong> solv-<br />

ing <strong>the</strong> redundant problem in biomechanics. Journal <strong>of</strong> Biomechanics, 28(8):999–<br />

1005, 1995.<br />

[24] Thomas S Buchanan, DG Lloyd, Kurt T Manal, and TF Besier. Neuromuscu-<br />

loskeletal modeling: Estimation <strong>of</strong> muscle <strong>for</strong>ces and joint moments and move-<br />

ments from measurements <strong>of</strong> neural command. Journal <strong>of</strong> applied biomechanics,<br />

20(4):367–395, 2004.<br />

[25] LL Menegaldo, AT Fleury, and HI Weber. A ‘cheap’ optimal control approach<br />

to estimate muscle <strong>for</strong>ces in musculoskeletal systems. Journal <strong>of</strong> Biomechanics,<br />

39(10):1787–1795, 2006.<br />

[26] BR Umberger, KGM Gerritsen, and PE Martin. A model <strong>of</strong> human muscle en-<br />

ergy expenditure. Computer methods in biomechanics and biomedical engineering,<br />

6(2):99–112, 2003.<br />

[27] H He, K Kiguchi, and E Horikawa. A study on lower-limb muscle activities during<br />

daily lower-limb motions. International Journal <strong>of</strong> Bioelectromagnetism, 9(2):79–<br />

84, 2007.<br />

[28] D Davy and M Audu. A dynamic optimization technique <strong>for</strong> predicting muscle<br />

<strong>for</strong>ces in <strong>the</strong> swing phase <strong>of</strong> gait. Journal <strong>of</strong> Biomechanics, 20(2):187–201, 1987.<br />

[29] FC Anderson and Marcus G Pandy. Dynamic optimization <strong>of</strong> human walking.<br />

Journal <strong>of</strong> Biomechanical Engineering, 123(5):381–390, 2001.<br />

[30] FC Anderson and Marcus G Pandy. Static and dynamic optimization solutions <strong>for</strong><br />

gait are practically equivalent. Journal <strong>of</strong> Biomechanics, 34(2):153–161, 2001.<br />

111


REFERENCES<br />

[31] Marcus G Pandy, FC Anderson, and D Hull. A parameter optimization approach<br />

<strong>for</strong> <strong>the</strong> optimal control <strong>of</strong> large-scale musculoskeletal systems. Journal <strong>of</strong> Biome-<br />

chanical Engineering, 114(4):450–460, 1992.<br />

[32] FC Anderson and Marcus G Pandy. A dynamic optimization solution <strong>for</strong> jumping in<br />

three dimensions. Computer Methods in Biomechanics and Biomedical Engineering,<br />

2(3):201–231, 1999.<br />

[33] R Neptune and A Van den Bogert. Standard mechanical energy analyses do not<br />

correlate with muscle work in cycling. Journal <strong>of</strong> Biomechanics, 31(3):239–245,<br />

1998.<br />

[34] J Collins. The redundant nature <strong>of</strong> locomotor optimization laws. Journal <strong>of</strong> Biome-<br />

chanics, 28(3):251–267, 1995.<br />

[35] U Glitsch and W Baumann. The three-dimensional determination <strong>of</strong> internal loads<br />

in <strong>the</strong> lower extremity. Journal <strong>of</strong> Biomechanics, 30(11/12):1123–1131, Feb 1997.<br />

[36] A Patriarco, R Mann, and S Simon. An evaluation <strong>of</strong> <strong>the</strong> approaches <strong>of</strong> opti-<br />

mization models in <strong>the</strong> prediction <strong>of</strong> muscle <strong>for</strong>ces during human gait. Journal <strong>of</strong><br />

Biomechanics, 14(8):513–525, 1981.<br />

[37] D Tsirakos, V Baltzopoulos, and R Bartlett. Inverse optimization: functional and<br />

physiological considerations related to <strong>the</strong> <strong>for</strong>ce-sharing problem. Critical reviews<br />

in biomedical engineering, 24(4/5):371–407, 1997.<br />

[38] DG Thelen and FC Anderson. Using computed muscle control to generate <strong>for</strong>-<br />

ward dynamic simulations <strong>of</strong> human walking from experimental data. Journal <strong>of</strong><br />

Biomechanics, 39(6):1107–1115, 2006.<br />

[39] RR Neptune, CP McGowan, and SA Kautz. Forward dynamics simulations provide<br />

insight into muscle mechanical work during human locomotion. Exercise and Sport<br />

Sciences Reviews, 37(4):203–210, 2009.<br />

[40] K Min and Y Koike. Muscular skeleton <strong>for</strong>ward dynamics simulation model <strong>of</strong> hu-<br />

man arm model based on rein<strong>for</strong>cement learning. IEIC Technical Report (Institute<br />

<strong>of</strong> Electronics, 103(470):11–16, 2003.<br />

112


REFERENCES<br />

[41] M Kaphle and A Eriksson. Optimality in <strong>for</strong>ward dynamics simulations. Journal<br />

<strong>of</strong> Biomechanics, 41(6):1213–21, 2008.<br />

[42] Felix E Zajac. <strong>Muscle</strong> and tendon: properties, models, scaling, and application<br />

to biomechanics and motor control. Critical reviews in biomedical engineering,<br />

17(4):359–411, 1989.<br />

[43] Marcus G Pandy, Felix E Zajac, E Sim, and W Levine. An optimal control model<br />

<strong>for</strong> maximum-height human jumping. Journal <strong>of</strong> Biomechanics, 25(2):207–209,<br />

1990.<br />

[44] J He, W Levine, and GE Loeb. Feedback gains <strong>for</strong> correcting small perturbations<br />

to standing posture. IEEE Transactions on Automatic Control, 36(3):322–332, Jul<br />

2001.<br />

[45] R Neptune and ML Hull. A <strong>the</strong>oretical analysis <strong>of</strong> preferred pedaling rate selection<br />

in endurance cycling. Journal <strong>of</strong> Biomechanics, 32(4):409–415, 1999.<br />

[46] AV <strong>Hill</strong>. The heat <strong>of</strong> shortening and <strong>the</strong> dynamic constants <strong>of</strong> muscle. Proceedings<br />

<strong>of</strong> <strong>the</strong> Royal Society <strong>of</strong> London. Series B, Biological Sciences, 126(843):136–195,<br />

1938.<br />

[47] AV <strong>Hill</strong>. The mechanics <strong>of</strong> active muscle. Proceedings <strong>of</strong> <strong>the</strong> Royal Society <strong>of</strong><br />

London. Series B, Biological Sciences, 141(902):104–117, 1953.<br />

[48] A Huxley. <strong>Muscle</strong> structure and <strong>the</strong>ories <strong>of</strong> contraction. Progress in Biophysics<br />

and Biophysical Chemistry, pages 255–318, 1957.<br />

[49] BA Garner and MG Pandy. The obstacle-set method <strong>for</strong> representing muscle paths<br />

in musculoskeletal models. Computer methods in biomechanics and biomedical<br />

engineering, 3(1):1–30, 2000.<br />

[50] IW Charlton and GR Johnson. Application <strong>of</strong> spherical and cylindrical wrapping<br />

algorithms in a musculoskeletal model <strong>of</strong> <strong>the</strong> upper limb. Journal <strong>of</strong> Biomechanics,<br />

34(9):1209–1216, 2001.<br />

[51] S P Marsden, D C Swailes, and G R Johnson. Algorithms <strong>for</strong> exact multi-<br />

object muscle wrapping and application to <strong>the</strong> deltoid muscle wrapping around <strong>the</strong><br />

113


REFERENCES<br />

humerus. Proceedings <strong>of</strong> <strong>the</strong> Institution <strong>of</strong> Mechanical Engineers, Part H: Journal<br />

<strong>of</strong> Engineering in Medicine, 222(7):1081–1095, Oct 2008.<br />

[52] Christopher J Gatti and Richard E Hughes. Optimization <strong>of</strong> muscle wrapping<br />

objects using simulated annealing. Ann Biomed Eng, 37(7):1342–1347, Jul 2009.<br />

[53] Arthur C Guyton and J Edward Hall. Textbook <strong>of</strong> medical physiology. Elsevier<br />

Saunders, 2006.<br />

[54] J Scherrer and H Monod. Local muscle work and fatigue in man. Journal de<br />

physiologie, 52:419–501, 1960.<br />

[55] W Rohmert. Determination <strong>of</strong> <strong>the</strong> recovery pause <strong>for</strong> static work <strong>of</strong> man. Interna-<br />

tionale Zeitschrift für angewandte Physiologie, 18:123–164, 1960.<br />

[56] Eastman Kodak Company. Kodak’s ergonomic design <strong>for</strong> people at work. Wiley,<br />

page 704, 2004.<br />

[57] D Hawkins and ML Hull. <strong>Muscle</strong> <strong>for</strong>ce as affected by fatigue ma<strong>the</strong>matical model<br />

and experimental verification. Journal <strong>of</strong> Biomechanics, 26(9):1117–1128, 1993.<br />

[58] J Ding, AS Wexler, and SA Binder-Macleod. A predictive model <strong>of</strong> fatigue in<br />

human skeletal muscles. Journal <strong>of</strong> Applied Physiology, 89(4):1322–1332, 2000.<br />

[59] T Xia and L Frey Law. A <strong>the</strong>oretical approach <strong>for</strong> modeling peripheral muscle<br />

fatigue and recovery. Journal <strong>of</strong> Biomechanics, 41(14):3046–3052, 2008.<br />

[60] JZ Liu, RW Brown, and GH Yue. A dynamical model <strong>of</strong> muscle activation, fatigue,<br />

and recovery. Biophysical journal, 82(5):2344–2359, 2002.<br />

[61] E Henneman, G Somjen, and DO Carpenter. Functional significance <strong>of</strong> cell size in<br />

spinal motoneurons. Journal <strong>of</strong> Neurophysiology, 28(3):560–580, 1965.<br />

[62] G J Tortora and Sandra Reynolds Grabowski. Introduction to <strong>the</strong> human body:<br />

<strong>the</strong> essentials <strong>of</strong> anatomy and physiology. J. Wiley & Sons, pages 1–720, 2003.<br />

[63] L Chin, P Yue, J Feng, and C Seow. Ma<strong>the</strong>matical simulation <strong>of</strong> muscle cross-<br />

bridge cycle and <strong>for</strong>ce-velocity relationship. Biophysical journal, 91(10):3653–3663,<br />

2006.<br />

114


REFERENCES<br />

[64] Jack M Winters and L Stark. <strong>Muscle</strong> models: what is gained and what is lost by<br />

varying model complexity. Biological Cybernetics, 55:403–420, 1987.<br />

[65] C W Sheppard. The <strong>the</strong>ory <strong>of</strong> <strong>the</strong> study <strong>of</strong> transfers within a multi-compartment<br />

system using isotopic tracers. Journal <strong>of</strong> Applied Physics, 19(1):70–76, 1948.<br />

[66] E Conte, A Federici, and J Zbilut. On a simple case <strong>of</strong> possible non-<br />

deterministic chaotic behavior in compartment <strong>the</strong>ory <strong>of</strong> biological observables.<br />

Chaos, 22(2):277–284, 2004.<br />

[67] Jens Wittenburg. Dynamics <strong>of</strong> multibody systems. Springer, pages 1–223, 2007.<br />

[68] Imsl <strong>for</strong>tran subroutines <strong>for</strong> ma<strong>the</strong>matical applications. Visual Numerics, Inc.,<br />

1997.<br />

[69] K Schittkowski. Nlpql: A <strong>for</strong>tran subroutine solving constrained nonlinear pro-<br />

gramming problems. Annals <strong>of</strong> Operations Research, 5(1-4):485–500, 1986.<br />

[70] Parviz E. Nikravesh. Computer-aided analysis <strong>of</strong> mechanical systems. Prentice-<br />

Hall, page 370, 1988.<br />

[71] Scott L Delp, FC Anderson, AS Arnold, P Loan, A Habib, CT John, E Guen-<br />

delman, and DG Thelen. Opensim: Open-source s<strong>of</strong>tware to create and analyze<br />

dynamic simulations <strong>of</strong> movement. IEEE Transactions on Biomedical Engineering<br />

BME, 54(11):1940–50, 2007.<br />

[72] KRS Holzbaur, WM Murray, and Scott L Delp. A model <strong>of</strong> <strong>the</strong> upper extremity<br />

<strong>for</strong> simulating musculoskeletal surgery and analyzing neuromuscular control. Ann<br />

Biomed Eng, 33(6):829–840, 2005.<br />

[73] Michael L Richardson. Upper extremity muscle atlas. <strong>Muscle</strong> Atlas, Musculoskeletal<br />

Radiology - http://www.rad.washington.edu/atlas/.<br />

[74] Biomechanics. http://www.sportspodiatry.co.uk/biomechanics.htm.<br />

[75] Gait cycle. http://www.orthopaedicsurgeries.co.uk.<br />

[76] Scott L Delp. Surgery simulation: A computer-graphics system to analyze and<br />

design musculoskeletal reconstructions <strong>of</strong> <strong>the</strong> lower limb. PhD Thesis, Stan<strong>for</strong>d<br />

University. Dept. <strong>of</strong> Mechanical Engineering, 1990.<br />

115


REFERENCES<br />

[77] Liang Ma, Damien Chablat, Fouad Bennis, and W Zhang. Dynamic muscle fatigue<br />

evaluation in virtual working environment. International Journal <strong>of</strong> Industrial<br />

Ergonomics, 39(1):211–220, 2009.<br />

[78] SC White and DA Winter. Predicting muscle <strong>for</strong>ces in gait from emg signals<br />

and musculotendon kinematics. Journal <strong>of</strong> Electromyography and Kinesiology,<br />

4(1):217–231, 1993.<br />

[79] DG Lloyd and TF Besier. An emg-driven musculoskeletal model to estimate muscle<br />

<strong>for</strong>ces and knee joint moments in vivo. Journal <strong>of</strong> Biomechanics, 36(6):765–776,<br />

2003.<br />

[80] Thomas S Buchanan, Scott L Delp, and JA Solbeck. Muscular resistance to varus<br />

and valgus loads at <strong>the</strong> elbow. Journal <strong>of</strong> Biomechanical Engineering, 120:634–639,<br />

1998.<br />

116

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

Saved successfully!

Ooh no, something went wrong!