the mechanics of cell motility and a unifying theory for

225
The Mechanics of Cell Motility and a Unifying Theory for Characterizing Directed Motion by Alex Loosley B. Sc, Simon Fraser University, Burnaby, BC, Canada, 2009 Sc. M., Brown University, Providence, RI, USA, 2011 A Dissertation submitted in partial fulfillment of the requirements for the Degree of Doctor of Philosophy in the Department of Physics at Brown University Providence, Rhode Island May 2015

Upload: others

Post on 27-Oct-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The Mechanics of Cell Motility and a Unifying Theory for

The Mechanics of Cell Motility

and a Unifying Theory for

Characterizing Directed Motion

by

Alex Loosley

B. Sc, Simon Fraser University, Burnaby, BC, Canada, 2009

Sc. M., Brown University, Providence, RI, USA, 2011

A Dissertation submitted in partial fulfillment of the

requirements for the Degree of Doctor of Philosophy

in the Department of Physics at Brown University

Providence, Rhode Island

May 2015

Page 2: The Mechanics of Cell Motility and a Unifying Theory for

c© Copyright 2015 by Alex Loosley

Page 3: The Mechanics of Cell Motility and a Unifying Theory for

This dissertation by Alex Loosley is accepted in its present form

by the Department of Physics as satisfying the

dissertation requirement for the degree of Doctor of Philosophy.

Date

Jay X. Tang, Ph.D., Advisor (�)

Recommended to the Graduate Council

Date

Jonathan S. Reichner, Ph.D., Reader (⊥)

Date

James M. Valles Jr., Ph.D., Reader (‖)

Date

Robert A. Pelcovits, Ph.D., Reader (∴)

Approved by the Graduate Council

Date

Peter Weber, Ph.D, Dean of the Graduate School

iii

Page 4: The Mechanics of Cell Motility and a Unifying Theory for

Vitæ

Alex J Loosley was born in Vancouver, British Columbia, Canada, on May 26,

1985 to Ellen B Loosley. Alex liked extreme sports while growing up, thrilling himself

by skateboarding, snowboarding, sailing, and playing ice hockey. At Templeton Sec-

ondary School, Alex accelerated through science and math courses (except biology),

actively participating in math competitions, and eventually graduating as the top

math student in his year.

During his college years at Simon Fraser University, Alex received a rigorous ed-

ucation in physics. In addition to learning from many fantastic professors (too many

to name), Alex benefited from being around many bright award winning classmates

including Cisco Gooding (NSERC CGS M and D, Quantum Gravity), Michael Mc-

Dermott, (NSERC CGS M and D, High Energy Theory), Joel Zylberberg (Fulbright

Scholar, Theoretical Neuroscience), and Karen Chan (NSERC CGS M and D, Theo-

retical Chemistry) to name a few. It was truly a special group.

Later, Alex moved to Providence, Rhode Island, to pursue a Ph.D. in Physics

under the guidance of Profs. Jay X. Tang and Jonathan Reichner. When not working

hard on his Ph.D. projects, Alex could be found refereeing ice hockey at the minor

and college levels, or racing sailboats in Newport, Rhode Island.

iv

Page 5: The Mechanics of Cell Motility and a Unifying Theory for

An abbreviated C.V. and other resources can be found at:

http://www.brown.edu/research/labs/tang-biophysics/lab-members/alex-loosley

v

Page 6: The Mechanics of Cell Motility and a Unifying Theory for

“It’s not poison, it’s clove.”

Xian speaking about die Clovenstoop

vi

Page 7: The Mechanics of Cell Motility and a Unifying Theory for

Dedicated to My Family

(ABCDE, R)

vii

Page 8: The Mechanics of Cell Motility and a Unifying Theory for

Acknowledgements

I would not be here writing this Ph.D. thesis without my research advisers Prof.

Jay Tang, Prof Jonathan Reichner, and Dr. Xian O’Brien. Jay, you have been there

for me from day 0. I will also be grateful that you brought me into the physics

department at Brown. Thanks also for many hours spent playing a key role in my

development as a scientist, from reading fellowship applications to supporting me

through a number of projects. Jonathan, I still have not finished that 4-D vector

analysis of god knows what that you asked me to do the first time you met me. You

are a fantastic translator of biology into layman’s terms, a subject which at one point

in my life appeared to be a forest too dense to navigate. To be able to speak to

somebody in personalized metaphors is a gift. Xian, I should have moved my office

to the division of surgical research near you long before I finally did to begin my fifth

year. Every young graduate student could benefit from being around good senior

graduate students and post-docs. You exemplify the role of a guide to a younger

researcher and I am a better researcher when I am around you. Thank you for always

being happy to heed my ideas and answer questions, no matter when I decided to

spontaneously pop into your office to present them. Thank you to all of my thesis

readers Profs. Jay Tang, Jonathan Reichner, James Valles Jr. and Robert Pelcovits.

I have been blessed with many good lab mates over my years as a graduate student.

My apologizes in advance if I miss anybody. In the Tang Lab and department of

physics, Angus McMullen, you have been an outstanding friend and lab mate with

viii

Page 9: The Mechanics of Cell Motility and a Unifying Theory for

many qualities that I have tried emulate. I am sorry that Yinda and I almost killed

your future wife the day you met her. Dr. Jingjing Wang, thanks for the many fun

science conversations and being a great desk mate. Drs. Patrick Oakes and Hyeran

Kang, I have always looked up to you both. Thank you Ryan Handoko, Wei Li, and

Cole Morrissette for all the hard work you put into my projects. Mike Morse, Jun He,

Nelson Leung, Barbara Dailey, Sabina Griffin, and Mary Ellen Woycik, thank you. In

the Reichner Lab and in the division of surgical research, Dr. Meredith Crane (Dingo

Dingo Dingo), Dr. Angel Byrd, Courtney Johnson, Estefany Flores, Dr. Dipan

Patel, Dr. Liz Lavigne, Dr. Hilary McGruder, Patty Young, Tracy Monteforte and to

anybody else that I missed, thank you all. Kyle Glass and Prof. Sunil Shaw, thanks

for letting me mess around with your microscope and flow chamber. Thanks to Prof.

Christian Franck, Eyal Bar-Kochba, and Dr. Jennet Toyjanova. Jennet, thanks for

all the reagent care packages and helping me be a better presenter. Also, thanks to

past advisers and colleagues that helped me get to Brown, especially Prof. Eldon

Emberly.

The summer of 2012 was the best summer of my life, and completely transformed

my way of thinking science. That summer in Singapore at the Mechanobiology In-

stitute (MBI), I worked with many fantastic researchers doing nanopillar array force

spectroscopy. Feroz Musthafa, my mentor, worked with me from 9pm to 5am just so

I could stay in my home time zone. Thanks for fabricating all of my nanopillar arrays

and bringing me on many exciting motor cycle rides. Aneesh Sathe, you helped get

my experiments up and running, endowing me with the cell culture and microscopy

skills to succeed. You along with Feroz also brought me into your family introduc-

ing me to many people I will never forget: Sharvari Sathe, Schweta Pradip Jadhav,

Soumya Mohanty, Shelly Kaushik, Aarthi Ravichandran, Rohith Kutty, and Dipanjan

ix

Page 10: The Mechanics of Cell Motility and a Unifying Theory for

Bhattacharya. Nils Gauthier, our experiments failed miserably except for that one

fibroblast. I wish I had more time to do science with you. Cheng-han Yu, you got me

to think of the cell membrane as the 2-D fluid that it is. Prof. Michael Sheetz, I was

always amazed that you would spontaneously take time out of your busy schedule to

see how I was doing. I now find myself hearing your voice as a read your manuscripts.

Many of your words of wisdom have become permanent pillars of my way of thinking.

Thanks also to Thomas Masters, Esther Anon, Bo Yang, Prabuddha Gupta, Alvin

Guo (hook’n me up with reagents day and night), Keiko Kawauchi (your HL-60 cell

line saved my summer research plan), Andrea and Christina (you taught me the im-

portance of coffee breaks), Joseph Tarango, Dacotah Melicher, Katelyn Goetz, Tanya

Gordonov, Hau Chan, Loretta Au and to anybody that I missed, there were so many

people that made an impact on my life that summer.

Outside of academia, I want to graciously thank my entire family. I have amazing

parents that have supported me every step of the way. Cisco Gooding, you helped

shape me as a physicist I am today. Callan MacKinlay, Racan Souiedan, Wendy

and Mitchell Parks, Angus and Jenica McMullen, Christina Miles, Marc Bertucci,

Kyle Helson, Lu Lu, Mingming Jiang, Simone Konig, I could go on. Thank you

fundamental constants of the universe for enabling this beautiful existence.

x

Page 11: The Mechanics of Cell Motility and a Unifying Theory for

Preface

This dissertation covers nearly five years of cell mechanics and motility research

findings as a graduate student in the Department of Physics at Brown University.

Most of this work was done as part of a collaboration with Profs. Jay X. Tang

(Physics), Xian O’Brien (Surgical Research), and Jonathan S. Reichner (Surgical

Research). This collaboration was originally fostered by the work of previous graduate

student, Dr. Patrcik Oakes. Patrick laid the groundwork for many of the experiments

described in this dissertation and has been supportive from afar over the last five years.

Six chapters grouped into two parts are presented here to report on an analyti-

cal tool for characterizing directed motion, multiple spring models that recapitulate

the shape and movement of motile cells, and the effects of pharmacological inter-

vention on immune cell motility and mechanics. Chapter 1 features a review style

blanket overview of cell motility and mechanobiology. Part 1 incorporates Chapters

2 and 3 about analytical tools for characterizing cell motility and the mechanisms

of mechanosensitive neutrophil motility in 2D. Chapter 2 describes a novel analysis

tool for measuring the extend of honing in directed motion. This tool was derived

from random walk theory, coming together through many critical back and forth dis-

cussions between Xian, Jonathan, Jay, and myself. Both Xian and Dr. Kate Oakley

provided the experimental data that was used to test our tool. Chapter 3 shows the

application of a simplified version of this tool to probe the mechanism of substrate

stiffness mediated migration of neutrophils. Xian and Jonathan were the driving

xi

Page 12: The Mechanics of Cell Motility and a Unifying Theory for

forces behind the work and much of this chapter is made up of text and figures from

our manuscript published in the Journal of Leukocyte Biology [1] (Copyright 2014 by

the Federation of American Societies for Experimental Biology). In Part 2, the scope

of this dissertation broadens with the additional consideration of cell mechanics in

describing cell motility. In Chapter 4, several toy models of cell motility using springs

and stick-slip adhesion as ingredients, are presented. Much of this work was inspired

by a Biophysics Journal Club presentation by Dr. Jingjing Wang, who did a presen-

tation about bipedal locomotion in crawling cells. This chapter contains a modified

version of our manuscript published in Physical Review Letters E [2] (Copyright 2012

by the American Physical Society, ). Chapter 5 contains much of the experimental

data I have collected over five years in collaboration with Xian, Jonathan, and Jay.

Several research assistants, namely Mr. Ryan Handoko, Ms. Wei Li, and Mr. Cole

Morrissette, helped with experiments and data analysis. Dr. Eliza Fox and Jonathan

Reichner were instrumental in setting up an IRB for the acquisition of neutrophils

from sepsis patients in the Rhode Island Hospital Intensive Care Unit. This lab work

would also not be possible without the general assistance of Dr. Angel Byrd and

(soon to be) Dr. Courtney Johnson. The original findings in this dissertation are

tied together in a final chapter containing unpublished data for future directions, and

closing remarks section.

Alex Loosley

Providence, Rhode Island

xii

Page 13: The Mechanics of Cell Motility and a Unifying Theory for

Publications

• Alex J. Loosley and Jay X. Tang. “Stick-slip motion and elastic coupling

in crawling cells.” Phys. Rev. E 86: 031908 (2012)

• Xian M. O’Brien, Alex J. Loosley, Kate E. Oakley, Jay X. Tang, and

Jonathan S. Reichner. “Technical advance: introducing a novel metric, di-

rectionality time, to quantify human neutrophil chemotaxis as a function of

matrix composition and stiffness.” J. Leukocyte Biol. 95: 993-1004 (2014)

• Alex J. Loosley, Xian M. O’Brien, Jonathan S. Reichner, and Jay X. Tang.

“Describing Directionally Biased Paths by a Characteristic Directionality

Time.” submitted (2014)

xiii

Page 14: The Mechanics of Cell Motility and a Unifying Theory for

Contents

Vitæ iv

Dedication vii

Acknowledgements viii

Preface xi

Publications xiii

List of Tables xx

List of Figures xxi

Chapter 1. Introduction 1

1. A General Overview of Cell Motility 2

2. Dissertation Outline 7

Part 1. Cell Kinematics 9

Chapter 2. Directionality Time: A New Theory for Characterizing Directed

Motion 10

Forward 11

1. Introduction 11

1.1. Commonly Used Analytical Tools for Characterizing Migration 11

1.2. Sampling Interval Dependent Metrics 12

xiv

Page 15: The Mechanics of Cell Motility and a Unifying Theory for

1.3. Sampling Interval Independent Metrics 13

2. Results 15

2.1. Mean Squared Displacement Analysis 15

2.2. Deriving the Directionality Time Model 16

2.3. Computational Modeling to Test Robustness 22

2.4. Application to Real Data 27

2.5. Neutrophil Chemotaxis 30

3. Discussion and Future Direcitons 33

4. Supplementary Material 36

4.1. Methods 36

4.2. Appendix A: Analytical Modeling 36

4.3. Appendix B: Deviations Caused by Variances and Ergodicity 46

Chapter 3. Directionality Time Analysis Identifies Rigidity Sensing Pathway

Through αMβ2 Integrins 51

Forward 52

1. Introduction 52

1.1. Neutrophils 52

1.2. Neutrophil Mechanosensing 52

1.3. Integrins and Mechanosensing Mechanisms 53

1.4. Integrin Ligand Engagement 54

2. Materials and Methods 56

2.1. Reagents 56

2.2. Neutrophil Isolation 56

2.3. Substrate Preparation 56

xv

Page 16: The Mechanics of Cell Motility and a Unifying Theory for

2.4. Neutrophil Chemotaxis 58

2.5. Microscopy 58

2.6. Cell Tracking Tools and Analytics 58

2.7. Statistics 59

2.8. Online Supplemental Material 59

3. Results and Discussion 60

3.1. Neutrophil morphology is dependent on substrate stiffness and

independent of ligand coating 61

3.2. The MSD mechanosensitivity of the neutrophil chemotaxis toward fMLP

is ligand-dependent 63

3.3. The TAD mechanosensitivity of neutrophil chemotaxis toward fMLP is

ligand-dependent 64

3.4. The tortuosity of neutrophil chemotaxis toward fMLP is ligand-dependent

and independent of substrate stiffness 65

3.5. Neutrophils migrating on Fgn-coated substrates toward fMLP show

mechanosensitive changes in td and β− 67

3.6. Neutrophils migrating on Col-coated substrates toward fMLP show a td

independent of substrate stiffness 67

3.7. Neutrophils migrating on Fn-coated substrates toward fMLP show

mechanosensitive differences in td and β− 69

3.8. The mechanosensitive component of directionality for neutrophils

migrating on Fn-coated substrates toward fMLP is dependent on

β2 integrins 70

Part 2. Integrating Cell Kinematics with Cell Mechanics 78

xvi

Page 17: The Mechanics of Cell Motility and a Unifying Theory for

Chapter 4. Cell Mechanics and Motility: Springs and Stick-Slip Adhesion 79

Forward 80

1. Introduction 80

2. Methods 82

2.1. Model overview 82

2.2. Simulation methods and criteria for characterizing dynamics 86

2.3. Choice of model parameters 88

2.4. Simulation benchmarking 89

3. Results 90

3.1. Viable spring configurations 90

3.2. Crawling dynamics depend on cell aspect ratio 95

4. Discussion 96

5. Supplementary Material 101

5.1. Analysis of fish keratocyte movies 101

5.2. Shape dynamics of fast crawling keratocytes 101

5.3. The two dimensional model 103

5.4. Comparing 2-D Elastic Configurations 107

Movie Legends (Movies Online) 112

Chapter 5. Immune System Modifier β-Glucan Regulates Motility Through the

Actin Cytoskeleton 114

Abstract and Forward 115

1. Introduction 115

1.1. Neutrophils 115

1.2. Traction Forces 116

xvii

Page 18: The Mechanics of Cell Motility and a Unifying Theory for

1.3. Force Motility Relationships 116

1.4. β-Glucan 117

2. Materials and Methods 118

2.1. Reagents 118

2.2. Substrate Preparation 118

2.3. Neutrophil Preparation 119

2.4. Microscopy 120

2.5. Cell Tracking and Migration Analysis 120

2.6. Traction Force Microscopy 121

2.7. Force and Energy Measurements 121

2.8. Dipole Moment Analysis 123

2.9. Statistics 123

3. Results 123

3.1. Spread Area and Motility Biphasically Dependent on BG Concentration 123

3.2. Mechanotactic Conservation of Energy and Contractility Time 124

3.3. Optimal Motility Correlated Inversely With Mechanical Output 128

3.4. Dipole Case Study: Forces Perpendicular to the Direction of Motion 129

3.5. Estimating Active Contractile Force and Cytoskeletal Stiffness 132

4. Discussion 134

5. Supplementary Material 138

5.1. Supplementary Movie Legends 138

5.2. Supplementary Tables and Figures 140

Chapter 6. Future Directions and Closing Remarks 143

Forward 144

xviii

Page 19: The Mechanics of Cell Motility and a Unifying Theory for

1. Future Project 1: Signaling During Early Stage Chemotaxis 145

1.1. Introduction 145

1.2. Computational Model 147

1.3. Results 153

1.4. Discussion 155

1.5. Pseudocode 157

2. Future Project 2: Neutrophils From Patients With Sepsis Show Novel

Stress Distributions During Chemotaxis 159

2.1. Introduction 159

2.2. Results 159

2.3. Discussion 162

3. Closing Remarks 164

Chapter 7. Extras 170

1. Cell Trajectory Simulator Class (MATLAB) 170

2. How to Save $150 On A New Mercury Lamp 179

Bibliography 180

xix

Page 20: The Mechanics of Cell Motility and a Unifying Theory for

List of Tables

2.1 Mathematical notation used to derive directionality time and inter-

pret cell motility .................................................................................... 37

3.1 Mechanosensitive parameters of human neutrophil chemotaxis to-

ward fMLP ............................................................................................ 74

4.1 List of parameters and physiologically viable values for the 2D stick-

slip model .............................................................................................. 89

5.1 Table of Alternating Conditional Expectations values .......................... 140

xx

Page 21: The Mechanics of Cell Motility and a Unifying Theory for

List of Figures

2.1 The effects of sampling interval on measurements and characteriza-

tion of migration trajectories................................................................. 14

2.2 Log-log EASD slopes of the 1D persistent biased random walk and

its correspondence to the directionality time model.............................. 21

2.3 The universal fit for measuring directionality time ............................... 23

2.4 Tangent-bias correlation plotted against the von Mises distribution

bias factor.............................................................................................. 24

2.5 Time scales at which βPBRW(t) converges to the directionality time

model when measurement errors are nonzero (σm > 0)......................... 25

2.6 Additional data plots investigating the robustness of the directional-

ity time model ....................................................................................... 27

2.7 Sampling interval dependent metrics applied to simulated 2D persis-

tent biased random walk trajectories .................................................... 28

2.8 Step-by-step flow-chart for processing experimental migration data

to measure directionality time............................................................... 31

2.9 Measuring directionality time from experimental data.......................... 32

2.10 Sampling interval dependent metrics applied to chemotactic poly-

morphonuclear neutrophils trajectories ................................................. 34

2.11 Explicit demonstration of the measurement error correction on ex-

perimentally measured log-log MSD slope............................................. 35

xxi

Page 22: The Mechanics of Cell Motility and a Unifying Theory for

3.1 Introduction to Integrins ....................................................................... 55

3.2 Mobility and TAD of neutrophil chemotaxis toward fMLP on surfaces

of varying stiffness depend on ligand coating ........................................ 62

3.3 Tortuosity of neutrophil chemotaxis toward fMLP is ligand-dependent

but independent of substrate stiffness ................................................... 66

3.4 Cells migrating on Fgn-coated gels toward fMLP have similar MSD

and RMS speed but show a mechanosensitive change in td and β− ...... 68

3.5 Neutrophils migrating on Col IV-coated substrates toward fMLP

show td, β−, and β+ independent of substrate stiffness ......................... 69

3.6 Neutrophils migrating on Fn-coated substrates toward fMLP show

mechanosensitive differences in MSD, td, and β−. ................................. 71

3.7 The mechanosensitive component of directionality for neutrophils mi-

grating on Fn-coated substrates toward fMLP is dependent on β2

integrins................................................................................................. 73

4.1 1-D model schematic and stick-slip adhesion definitions ....................... 85

4.2 2-D model schematic ............................................................................. 87

4.3 Overview and dynamics of viable 2-D stick-slip toy models (Configs.

1-4) ........................................................................................................ 92

4.4 Characterized dynamical response of Config. 4 with respect to pa-

rameters KD, KL and g ......................................................................... 94

4.5 Phase diagrams of Config. 4 dynamics: cell shape matters ................... 97

4.6 Cell aspect ratio, speed, and amplitude of cell length oscillation distri-

butions measured from previously published movies of fish keratocytes102

xxii

Page 23: The Mechanics of Cell Motility and a Unifying Theory for

4.7 Bipedal locomotion and lateral oscillations of the nucleus observed

in fish keratocytes.................................................................................. 104

4.8 Representative dynamical responses of four select configurations of

the 2-D cell crawling model ................................................................... 109

4.9 Drag force sensitivity analysis of Config. 4 ........................................... 113

5.1 Traction force microscopy cartoon with representative DIC and trac-

tion field frames..................................................................................... 121

5.2 Neutrophil motility and mechanical output depend on BG concen-

tration and substrate stiffness ............................................................... 125

5.3 Two graphs showing the Pearson product-moment correlation coeffi-

cients (PCCs) between variables describing the motility and mechan-

ical output of neutrophils ...................................................................... 129

5.4 Slower moving cells exert more energy contracting their substrate

than faster moving cells......................................................................... 130

5.5 Dipole moment analysis of a motile BG0 cell migrating on a 5 kPa

substrate................................................................................................ 133

5.6 Estimating total myosin mediated contractile force and cell stiffness ... 134

5.7 The perceived cell boundary may differ from the actual cell boundary 141

5.8 Control experiments showing that the BG effect cannot be explained

by changes in osmolarity ....................................................................... 141

5.9 Neutrophil adhesion rate, activity rate, and contractility rate .............. 142

6.1 Schematic showing the chemical signaling of chemotaxis and the

downstream cellular components that drive cell migration ................... 145

6.2 Classification of neutrophil traction stress fields during 2D chemotaxis148

xxiii

Page 24: The Mechanics of Cell Motility and a Unifying Theory for

6.3 The LEGI model ................................................................................... 150

6.4 LEGI model results using Cartesian coordinate system ........................ 154

6.5 LEGI model results using polar coordinate system............................... 156

6.6 Neutrophil activation, endothelium transmigration, and chemotaxis

through the extracellular matrix ........................................................... 160

6.7 Differential Traction Force Mapping of representative neutrophils

from healthy (H) and septic (S) donors................................................. 161

6.8 Neutrophils from healthy (H) and septic (S) donors differ in adhesion

and Mean Square Displacement ............................................................ 163

6.9 Schematics summarizing all the force modes of motile and non motile

neutrophils............................................................................................. 165

7.1 The $165 kymograph............................................................................. 179

xxiv

Page 25: The Mechanics of Cell Motility and a Unifying Theory for

Abstract of “The Mechanics of Cell Motility and a Unifying Theory for

Characterizing Directed Motion”

by Alex Loosley, Ph.D., Brown University, May 2015

This dissertation is split into two parts with the first focusing on characterizing cell

motility, and the second focusing on linking cell motility with cell mechanics. Cellular

motility has been routinely characterized by parameters such as migration speed,

tortuosity, and persistence time. However, many of these parameters are generally

not reproducible because their numerical values depend on technical parameters such

as the experimental sampling interval and measurement error. In Part 1, we use

random walk theory, simulations, and experimental data to address the need for

a metric that quantifies the directionality of a migration path in a manner that

decouples from technical parameters. We call this novel metric directionality time

because it can be interpreted as the minimum observation time required to determine

that the migratory motion is directed. Along with measures of migration speed and

persistence, we used the directionality time metric to determine the directedness

of chemotactic neutrophil migration paths as a function of physiologically relevant

substrate stiffnesses and compositions. We find that engagement of the β2-integrin,

CR3, is required for the substrate stiffness dependent regulation of neutrophil honing

in 2D on fibronectin.

In Part 2, we investigate the role of cell mechanics in cell motility. As cells migrate,

they exert cytoskeletal driven contractile forces on their environment. Using spring

models with stick-slip adhesion, we are able to predict how the mechanical properties

and shape of the cell affect its dynamics and traction forces. We then switch the

focus back to neutrophils and explore the relationship between neutrophil contractile

xxv

Page 26: The Mechanics of Cell Motility and a Unifying Theory for

force, motility, and the biological response modifier β-glucan, a substance that has

shown promise as a clinic grade therapeutic. Using live cell imaging and traction force

microscopy to measure neutrophil traction forces as a function of substrate stiffness

and β-glucan concentration, we report that increasing the concentration of β-glucan

leads to diminished traction forces. Our findings indicate that biological response

modifiers may act through the modulation of cell mechanics and motility. Further,

we report an overall inverse correlation between migration speed and mechanical

output that is affected by β-glucan. These data should be useful for modeling cell

motility.

Page 27: The Mechanics of Cell Motility and a Unifying Theory for

CHAPTER 1

Introduction

.

1

Page 28: The Mechanics of Cell Motility and a Unifying Theory for

1. A General Overview of Cell Motility

Cell motility is the ability of a cell to actively locomote from one place to another. Such

locomotion can arise in the form of swimming [3, 4] or crawling [5–7] with the latter the

focus of this dissertation. Crawling motility is required for many physiological processes that

sustain life. For example, in wound healing, an influx of inflammatory cells clear harmful

pathogens and dead material [8]. Following the inflammatory response, motile fibroblasts

must arrive and proliferate at the scene of the injury where they facilitate extracellular

matrix synthesis and repair [9, 10]. Finally, epithelial cells collectively migrate in a motion

called plithotaxis to close tissue lesions [11–13].

Research aiming to describe and elucidate the mechanisms of cell motility has been

under way since the 1680s when the first study of bacterial motion was completed by van

Leeuwenhoek [14]. This was also one of the first studies of micro-organisms using a micro-

scope, which had been invented nearly 100 years prior. In the twentieth, ground breaking

optical techniques, such as differential interference contrast (DIC), confocal microscopy, and

the discovery of green fluorescent protein (GFP) [15], have paved the way for the acquisition

of massive amounts of data in real time at spatial resolutions down to 10−7 m (0.1µm), two

orders of magnitude smaller than the size of most cells. These technologies, aided by a

sharp upswing in computational power, have led to an influx of quantitative data describ-

ing cell migration and a new frontier of opportunities for interdisciplinary collaborations

between biologists and physicists who wish to understand the physics behind the question,

how do crawling cells locomote? In the chapters that follow, this question is answered by

combining the principles of classical mechanics, fluid dynamics, and random walk theory

with the underlying biomolecular mechanisms that generate and sustain active matter.

Biomolecular mechanisms are often initiated by a physical interaction, and often lead

to a subsequent physical interaction. For this reason, biomolecular mechanisms and their

corresponding physical interactions are indubitably connected. Active matter can be defined

2

Page 29: The Mechanics of Cell Motility and a Unifying Theory for

as matter that spontaneously converts a fuel source (energy source) into active forces.

Without active matter, cellular processes such as morphogenesis [16], mitosis [17], and cell

motility [18], would not be possible.

At least two sets of components within the cell are necessary for active matter driven

cell crawling motility. The first is the actin cytoskeleton, a viscoelastic network that spans

the cell and is made up of actin filaments called F-actin, cross-linking components, and

molecular motors. F-actin is highly rigid at the length scale of cells, with a filament per-

sistence length of 17.7µm [19] (cell lengths ∼ 10 to 15µm). Thus, the actin cytoskeleton

acts as a strong scaffolding medium maintaining cell shape [20, 21] and providing a base

for molecular cargo transport [22]. The actin cytoskeletal network is highly dynamic and

can be disassembled into globular actin monomers called G-actin, which can be reused to

polymerize new F-actin elsewhere. As part of the motility process, actin polymerization at

the leading edge of the cell generates localized protrusions of the cellular plasma membrane,

while molecular machines such as ATP driven myosin motors cause contraction of the actin

cytoskeleton. As the actin cytoskeleton contracts, depolymerization mediated by cofilin

[23, 24] occurs near the center of the cell, and material is recycled for F-actin polymeriza-

tion and protrusion at the edges of the cell. As a result of both depolymerization at the

center and polymerization at periphery, the actin cytoskeleton flows toward the center of

the cell in what is called retrograde flow.

The second set of components required for crawling motility is adhesion between the cell

and its substrate. Without an adhesive clutch to anchor the cell, it would not be possible

for protrusion at the leading edge of the cell to propel the cell forward. For many cells, this

anchorage is achieved through transmembrane proteins called integrins [25]. Integrins form

bonds with the substrate and a host of intracellular components, ultimately coupling the

substrate to the actin cytoskeleton [26]. Retrograde flow of the actin cytoskeleton leads to

3

Page 30: The Mechanics of Cell Motility and a Unifying Theory for

traction forces that are inward directed (contractile) and highly correlated across the cell

body [27–29].

In past decades, techniques have been invented to detect cellular traction forces, provid-

ing a new spotlight for illuminating features of the cytoskeleton/adhesion driven cell motility

process. In 1980, Harris et al. were the first to see traction forces by culturing fibroblasts on

very thin silicon substrates that wrinkled under cellular traction stress [30]. Later in 1994,

Lee et al. improved the spatiotemporal resolution of this procedure by implanting small

latex beads into the silicon substratum to measure traction force generated displacements

[31]. In 1999, Dembo and Wang were the first to report traction force measurements using

bead embedded polyacrylamide substrates, an improvement over silicon substrates because

of their well characterized rheological properties [32]. Dembo and Wang assumed that the

size of their substrates was effectively infinite compared to the 3T3 fibroblasts they studied

in order to apply the Boussinesq equation to map micrometer sized bead displacements to

traction stresses on the order of 0.1 to 10 nN/µm2. This study marked the beginning of

modern traction force microscopy measurements. In the last 15 years, numerous improve-

ments have been made to resolution, accuracy, and throughput of calculating traction forces.

This includes greatly reducing the computational power required to measure traction forces

by solving the inverse Boussinesq equation in Fourier space [33], improving resolution us-

ing multiple colors of fluorescent displacement markers [34], assuming finite traction stress

regions instead of point sources [35], and properly accounting for large nonlinear substrate

displacements and rotations [36]. Traction force microscopy has now also been adapted

for measurements in 3D [37–39], and has even been combined with total internal reflection

fluorescence (TIRF) microscopy to precisely co-localize cell-substrate contact regions with

measurements of traction stress. [40].

In this dissertation, traction force microscopy is used to address two questions: How

do cells coordinate actin cytoskeletal mechanics and adhesions to achieve locomotion in a

4

Page 31: The Mechanics of Cell Motility and a Unifying Theory for

specific direction? And, how much force (or energy) does the cellular machinery apply to

(or use in contracting) its environment during locomotion? The latter is important because

it is part of an even bigger question currently under intense investigation: What is the

relationship between cellular traction forces and motility [41, 42]?

The ability for cells to detect and adapt their function based on the mechanical prop-

erties of their environment is another important factor that plays a significant role in cell

motility. This phenomenon, sometimes called mechanosensing, has been the topic of nu-

merous studies over the last 15 years [1, 29, 32, 43–55].

Mechanosensing is one of the central themes of this dissertation, specifically the role

of mechanosensing in neutrophils. Neutrophils play a primary role in the inflammatory

response as part of the innate host defense against pathogenic bodies. Vital to their role

is the ability to migrate from blood vessels, through the extracellular matrix, to sites of

inflammation. During the course of migration, neutrophils encounter a variety of tissues

characterized by differing physical topologies and stiffnesses ranging in Young’s moduli from

10−1 kPa for brain tissue, to more than 106 kPa for bone [29]. That cell motility depends on

the mechanical properties of a cell’s surroundings is now widely accepted [41, 48, 49]. How

this mechanosensitivity comes about is another matter. We ask the following questions:

Does mechanotactic regulation depend on tissue type? Which integrins are involved in the

mechanotactic regulation of cell motility? And, what specific aspect of motility do specific

integrins regulate? Note above that the units kPa were used to describe substrate stiffness.

This is to be consistent with the bulk of the recent literature. However, most other quantities

in this dissertation, including applied stress, have been written using the fundamental size,

time, and force scales of cell motility: µm, s, and nN, respectively (1 nN/µm2 = 1 kPa).

These scales are convenient for making quick and dirty estimates of quantities based on the

data presented in this dissertation.

5

Page 32: The Mechanics of Cell Motility and a Unifying Theory for

In order to answer the second question, precise analytical tools are required to charac-

terize the trajectories of cell migration. Common has been the usage of persistent random

walk theory to fit the mean squared displacement (MSD) of non directed migration trajecto-

ries thereby characterizing them in terms of parameters for migration speed and persistence

time [56–62]. When cell migration is directed (i.e. biased towards an external cue), one

more parameter is necessary to fully characterize the motion. One parameter that has been

used is the fraction of all turning angles between ±φ, where φ is an arbitrary angle typ-

ically chosen to be ≤ 90o [1, 48, 63, 64]. Another extra parameter that has been used is

tortuosity, the end to end distance divided by the arc length of the migration trajectory

(also known as chemotactic index [1, 49, 50], directionality fraction [65], and straightness

index [60, 66]). Both of these two parameters have shortcomings. Mainly, they are sampling

interval dependent. The sampling interval used to collect data biases the measurement of

both turning angle fraction and tortuosity (more about this in Chapter 2). Some rules for

characterizing MSD scaling exponent crossover (currently used in physics [67] and contin-

uum mechanics [68]) could potentially be developed into a directed cell migration analysis

tool that is sampling interval independent. But, these theories are not designed to analyze

noisy and potentially nonergodic migratory motion. To address the need for an analytical

tool that measures the honing portion of directed motion, we used random walk theory to

derive a tool called directionality time. Armed with the proper analytical tools to character-

ize directional cell migration, we identified the integrins responsible for the mechanotactic

regulation of 2D neutrophil honing towards a source of chemoattractant.

Finally, towards the end of this dissertation, we combine the study of motility and

traction microscopy to investigate the effects of biological response modifiers on the cell

mechanics of immune cells. As a model biological response modifier, we use is β-glucan

(BG) from yeast that has been purified into a clinical grade, water soluble form [69]. BG

of this form has been acts as a biological response modifier by improving wound healing

6

Page 33: The Mechanics of Cell Motility and a Unifying Theory for

in mice [70], by showing promise as an anti-cancer agent in in vivo tumor models [71–

73] and clinical trials [74], and by enhancing chemotactic honing of neutrophils on glass

[75, 76]. To promote an effect in neutrophils, BG is known to bind and allosterically

regulate the affinity of Complement Receptor 3 (CR3, also known as αMβ2, Mac-1, and

CD11b/CD18) [76–79], an integrin that plays a significant role in mechanosensitive behavior

[1, 48]. Therefore, it is rational to suggest that BG affects biological response modification

through the regulation of cell mechanics. Altogether, BG has the potential to be a positive

pharmacological intervention and we use live cell tracking and traction force microscopy to

investigate the role of cell mechanics as part of this intervention.

2. Dissertation Outline

Although the primary cell directly studied in this dissertation is the human neutrophil,

many of the analytical techniques developed herein are generally applicable to a variety of

other motile organisms, and our findings provide clues towards a broad understanding of

crawling motility in all cell types. Below is the story of cell motility beginning in Chapter

2 by first using random walk theory to develop the analytical tools necessary to analyze

motility kinematics. Next, in Chapter 3, these tools are applied to mathematically charac-

terize directed cell migration in terms of speed, persistence, and honing. This analysis is

applied to determine the specific integrins on neutrophils that mechanotactically regulate

each of these components of cell migration. In Chapter 4, we switch gears ever so slightly

to explain cell motility and traction forces using minimalist toy models using springs and

stick-slip adhesion as ingredients. Next, in Chapter 5, we bring together the studies of

traction forces, mechanosensing, and cell motility to determine which factors of mechanical

output correlate to optimal neutrophil motility. As part of this study, we show for the first

time the relationship between immune system priming with β-glucan, and traction forces,

demonstrating that the immunological response is tied to the mechanical output of immune

7

Page 34: The Mechanics of Cell Motility and a Unifying Theory for

cells. The dissertation finishes with a future directions Chapter that contrasts the motil-

ity and traction forces of BG treated neutrophils with data from neutrophils treated with

lipopolysaccharide, or harvested from human patients with sepsis.

8

Page 35: The Mechanics of Cell Motility and a Unifying Theory for

Part 1

Cell Kinematics

Page 36: The Mechanics of Cell Motility and a Unifying Theory for

CHAPTER 2

Directionality Time: A New Theory for Characterizing

Directed Motion

10

Page 37: The Mechanics of Cell Motility and a Unifying Theory for

Forward. In this chapter, we review and develop the analytical tools necessary to

characterizing directed cell migration. These tools are applied in Chapter 3 to study the

role of integrins in regulating directed cell motility as a function of substrate stiffness.

1. Introduction

Directed cell migration is the process where a single cell or a group of cells bias their

direction of locomotion by coupling to an external cue. One example of an external cue is

a chemical gradient. Migration in the direction of (or opposite to) a particular chemical

gradient is called chemotaxis. Besides chemotaxis, there are many types of directed cell

migration named according to the external cue, including gravitaxis [80], aerotaxis [81],

durotaxis [82], haptotaxis [83], and plithotaxis [13]. These processes are ubiquitous in

nature, facilitating the innate and adaptive immune systems [84, 85], sexual reproduction

[86], embryonic development [87], cancer metastasis [88, 89], and more. The efficacy to

which cells are able to carry out these functions is often tied to the characteristics of their

migration, including migration speed, persistence, and tortuosity. These characteristics

can be quantified to determine which biochemical and biomechanical factors affect cell

migration, and by how much.

1.1. Commonly Used Analytical Tools for Characterizing Migration. Mean

squared displacement (MSD) is one of the most common metrics for measuring migration

speed and distance traveled because it is easily interpretable and readily derived from math-

ematical models of motion. Numerous studies that characterize directed migration use MSD

in conjunction with at least one other metric for quantifying path persistence or tortuosity

[1, 41, 48, 49, 90]. Three examples of such metrics used are: the distribution of turning an-

gles between discrete measurements of centroid displacements (turning angle distribution,

TAD); tortuosity (also known as straightness index [60, 66], chemotactic index [49, 50], or

directionality ratio [65]) defined as the end-to-end distance traveled divided by the total

11

Page 38: The Mechanics of Cell Motility and a Unifying Theory for

migration path length; and tangent-tangent correlation, which describes the correlation in

migration path orientation over a specific length or time interval.

1.2. Sampling Interval Dependent Metrics. In order to gain insight from quanti-

tative characterizations of the migration path, the numerical values of the metrics applied

need to be reproducible from one set of experiments to another. Such values should also

reflect the true kinematic properties of migration by decoupling from pseudo random kine-

matics induced by measurement error along the migration path. The shortcoming of TAD,

tortuosity, and tangent-tangent correlation is that they each implicitly depend on sampling

interval, ∆t, which is the time interval between position measurements. Sampling interval

can be chosen arbitrarily, implying that TAD, tortuosity, and tangent-tangent correlation

curves are not generally reproducible without applying equivalent sampling intervals across

all experiments. Even when sampling intervals are accounted for, a sampling interval de-

pendent metric only characterizes migration at an arbitrarily chosen time scale at which

the metric may or may not decouple from measurement error.

To visualize sampling interval dependence, consider two experimental measurements of

a migration path, one using a “long” sampling interval, ∆t = ∆t>, the other using a “short”

sampling interval, ∆t = ∆t< (Fig. 2.1 a, top and bottom, respectively). Circles are centroid

positions and the corresponding perceived migration paths (blue line segments connecting

circles) are juxtaposed against the true migration path (thick grey curve). The deviation

between centroid positions (rn = r(tn), where tn = n∆t, n = 1, 2, ..., N , ∆t = ∆t< or ∆t>)

and the true migration path represents centroid measurement error, σm, which depends

on factors such as image resolution and cell boundary detection accuracy. Angles between

successive blue line segments, φn, are turning angles (−π < φ ≤ π). Taking into account

all turning angles, normalized TADs, ρφn(φ; ∆t), are calculated for both sampling intervals

(Fig. 2.1 b). As the sampling interval increases towards the total duration of observation,

the TAD curve becomes sharply peaked at φ = 0. Conversely, as the sampling interval

12

Page 39: The Mechanics of Cell Motility and a Unifying Theory for

decreases towards zero, the effects of diffusive kinematics and centroid measurement error

flatten the TAD curve. Hence, TAD depends notably on sampling interval. One measure of

persistence is the so-called turning angle persistence (TAD persistence), the fraction of all

turning angles between ±π2 (shaded area under TAD curves in Fig. 2.1 b). TAD persistence

depends on the sampling interval just as TAD does.

The dependence of tortuosity on sampling interval is apparent when considering the

limit that the sampling interval approaches the total duration of the migration path. In

this limit, total path length approaches the end-to-end length resulting in a tortuosity

of 1. When sampling interval decreases, the total path length increases due to centroid

measurement error and the underlying fractal nature of the migration path itself [91, 92].

1.3. Sampling Interval Independent Metrics. Tangent-tangent correlation, v(t+ τ) · v(t),

is the time averaged cosine of the angle between tangent vectors v(t+ τ) and v(t) that are

separated by a time interval τ . The overline denotes an average over all time t. When ap-

plied to discrete experimental data captured at a specific sampling interval, v(t) is replaced

by vn (Fig. 2.1 a). Tangent-tangent curves are sampling interval dependent because the

uncertainty of measuring tangent vector direction increases with increasing sampling inter-

val. However, there is a sampling interval independent measure called persistence time that

derives from the tangent-tangent correlation curve. Persistence time, tp, is the time scale

over which directional orientation of the migration path remains correlated. Measurements

of persistence time are sampling interval independent as long as the sampling interval is

small enough for it to be resolved. In general, persistence time would be measured as a fit

parameter in a model used to fit the tangent-tangent correlation curve over all time scales.

Persistence time and migration speed together can characterize non-directional motion,

but an additional metric is needed to characterize directional motion. We recently discussed

the need for such a metric in Ref. [1]. The purpose of this work is to derive a sampling

interval independent metric that measures directionality. One approach could entail fitting

13

Page 40: The Mechanics of Cell Motility and a Unifying Theory for

a b

c

∆t = ∆t<

r1

r2

rn

^

v1

^

v2

^

vnφ1

φ2

φn

∆t = ∆t>

φ

ρφn(φ;∆t<)

−π −π2

π π02

−π −π2

π π02

ρφn(φ;∆t>)

φ

Θn+1Ln

Θn

Rn

Rn+1

êy

êx

Figure 2.1. The effects of sampling interval on measurements and char-acterization of migration trajectories. (a) A migration path sampled witha long sampling interval, ∆t = ∆t> (top, outlined yellow circles) and a shortsampling interval, ∆t = ∆t< (bottom, outlined white circles). Using the longsampling interval diagram, the observed migration path is formed by connectingmeasurements of centroid positions, r1, r2, ..., with lines (in red). Unit tangentvectors are shown as v1, v2, ... while turning angles are defined as the angle be-tween successive tangent vectors, φ1, φ2, .... Differences in centroid position fromthe true migration path (thick grey curve) represent measurement error. (b)Turning angle distributions (TAD), ρφn(φ; ∆t), based on both the long (top) andshort (bottom) sampling intervals. A measure of migration persistence knownas TAD persistence is the area under the TAD curve between ±π

2(shaded).

TAD persistence depends notably on sampling interval. Similar diagrams canbe used to show the sampling interval dependence of metrics such as tortuosityand tangent-tangent correlation. (c) A diagram visually defining terms nec-essary to analytically derive the directionality time model. A cell is depictedas a random walker that steps (R1,R2, ...,Rn, ...) located in a 2-D coordinatesystem defined by the unit vectors ex and ey. Capital letters correspond torandom variables. For each step, there is a corresponding step length Ln andpolar angle Θn. The latter is defined with respect to unit vector ex and is notto be confused with the corresponding turning angle, Φn = |Θn −Θn−1|

data such as TAD persistence or tortuosity to a model with a set of fit parameters. However,

TAD persistence and tortuosity models are difficult to calculate and interpret. Mean squared

displacement (MSD) models are easier to calculate and interpret and several models have

already been analytically derived [60, 93, 94]. Nevertheless, this approach of fitting to a

model only works if the underlying kinematics of the migrating cell are understood a priori

such that a model can be chosen. While one can attempt to fit more than one model to

14

Page 41: The Mechanics of Cell Motility and a Unifying Theory for

determine which fits best, changes to MSD from one model to the next can be small with

respect to the error bars on an experimental MSD curve. Hence, there is the possibility

of a causality loop - one cannot accurately understand a set of migration paths without

knowledge of the underlying process and corresponding random walk model, but one is

unsure of the corresponding random walk model without understanding the migration paths.

To circumvent this causality loop, we take a bottom up approach to derive a novel sampling

interval independent metric called directionality time based on the slope of MSD in log-log

coordinates. We show that the directionality time concept is broadly applicable to many

types of directional motion and demonstrate its implementation on data of directionally

migrating chemotactic neutrophils.

2. Results

2.1. Mean Squared Displacement Analysis. Leading to the definition of direc-

tionality time, we begin with the assertion that ensemble averaged squared displacement

follows a power law ⟨r2(t)

⟩∼ tα (2.1)

where square brackets 〈〉 denote the ensemble average over squared displacements measured

at time t. To be precise about the type of averaging, we call this quantity the ensemble

averaged squared displacement (EASD) instead of MSD. The exponent α characterizes the

motion. A constant value of α = 1 indicates diffusive (random) motion whereas α = 2

indicates ballistic (directed) motion. Other values represent subdiffusive motion (0 < α <

1), superdiffusive motion (1 < α < 2), or no motion at all (α = 0).

When few sets of trajectories are available for the ensemble average, time averaged

squared displacement (TASD) can be calculated to reduce statistical noise. The TASD of

the ith migration path is given by

r2i (τ) = |ri(t+ τ)− ri(t)|2 (2.2)

15

Page 42: The Mechanics of Cell Motility and a Unifying Theory for

where the overline denotes an average over time t. Squared displacement that is first time

averaged and then ensemble averaged,⟨r2(τ)

⟩, is hereby referred to as MSD as this is how

it is often defined in many studies [1, 48, 95].

Ergodicity, ξi, is the conversion factor that maps EASD as a function of time t, to TASD

(and MSD) as a function of time interval τ :

r2i (τ) = ξi(τ)〈r2(t = τ)〉 (2.3)

where subscript i is the migration path index. The time averaging that goes into calcu-

lating MSD smooths out the effects of variables that may be changing over time, such as

instantaneous speed. When such factors change significantly, 〈ξ〉 6= 1, and the migration

paths are said to be nonergodic. Otherwise, the migration paths are ergodic and EASD

and MSD are interchangeable (i.e. 〈r2(τ)〉 = 〈r2(t = τ)〉).

The slope of EASD plotted in log-log coordinates is an approximate measure of the

EASD exponent α (Eq. 2.1) and therefore a measure that characterizes trajectory diffusivity

and/or directedness. Using Eq. 2.1, and noting that α can change with time t, one can define

the log-log EASD slope

β(t) =dα

dtt ln t+ α. (2.4)

When EASD exponent α is constant, β = α. Otherwise, β(t) is an estimator of α(t), and

therefore an estimator of how diffusive or ballistic the motion is at a particular time, t. Log-

log MSD slope, β(τ), can also be calculated to characterize motion as a function of time

interval τ . Two questions arise: What is the mathematical form of these β curves? Can a

β curve be used to determine a sampling interval independent quantity that characterizes

directionality?

2.2. Deriving the Directionality Time Model. When observing a directionally

biased random walk, there exists a sufficiently large sampling interval such that the motion

16

Page 43: The Mechanics of Cell Motility and a Unifying Theory for

will appear to be ballistic. Put in terms of log-log MSD slope, β(τ) → 2 as τ → ∞ for

a directionally biased random walk. The idea of using log-log MSD slope to measure this

particular time interval was recently hypothesized in our recent article about neutrophil

chemotaxis [1]. We suggested an empirical fit function for β(t) to estimate the location of

the directionality time transition. Here in this article, we rigorously develop the concept of

directionality time from the bottom up by analytically deriving a universal β(t) fit model

and using biased and persistent random walk modeling to characterize its robustness. Direc-

tionality time is defined as the time scale above which motion appears ballistic (directional)

and can be loosely interpreted as the time it takes for a random walker to orient towards

an external cue.

To determine the mathematical meaning of directionality time and its applicability, log-

log EASD slope, β(t), is analytically derived for three directionally biased random walk

models:

(1) Drift Diffusion (DD)

(2) 2D Stepping Biased Random Walk (2D-SBRW)

(3) 1D Persistent Biased Random Walk in Continuous Time (1D-PBRW)

To optimize the flow of this chapter, detailed step by step derivations of β(t) for each

model have been relegated to Appendix A (Sec. 4.2) at the end of this chapter, while the

calculations and significant findings are summarized here.

DD (model 1) was a suitable starting point because these processes are readily under-

stood. For DD with a diffusion constant D, and drift speed u, the log-log EASD slope is

shown in Appendix A (Sec 4.2.2) to be

βDD(t) =1 + 2t

td

1 + ttd

(2.5)

where td = 2dDu2 defines directionality time. Note that β(t) begins at β(0) = 1 and asymp-

totes towards 2. This is the signature of a directionally biased random walk. Directionality

17

Page 44: The Mechanics of Cell Motility and a Unifying Theory for

time is the time at which β = 32 where the migration transitions from diffusive to direc-

tional. As D increases and/or u decreases, more observation time is required to determine

that the motion is directionally biased.

A 2D-SBRW (model 2) describes an object that steps from one discrete position, Rn,

to the next, Rn+1 (n = 0, 1, 2, ...), such that the displacements between successive steps are

biased towards a particular direction, ex (Fig. 2.1 c). Notationally, all random variables

are assigned capital letters. Using Ln and Θn to denote step lengths and polar angles

(orientation) respectively, the stepwise EASD can be shown to be

⟨R2n

⟩= n〈L2〉+ n(n− 1)〈L cos Θ〉2

(see Ref. [94] for the derivation). Note that the motion is diffusive (〈R2n〉 ∼ n) when n is

small and directional (〈R2n〉 ∼ n2) when n is large. This step by step random walk has no

persistent directionality at low n because the direction of motion changes with each step.

By defining a constant instantaneous speed v, the approximation n ≈ vt〈L〉 can be used to

derive EASD as a function of time t instead of step number n. In the time representation,

this is a model of a biased random walk (BRW) instead of an SBRW. Differentiating in

log-log coordinates gives the log-log EASD slope

βBRW(t; td) =1 + 2t

td

1 + ttd

(2.6)

which is functionally identical to βDD (Eq. 2.5), now with directionality time given by

td =〈L〉v

(〈L2〉 − 〈L cos Θ〉2

〈L cos Θ〉2

). (2.7)

The functional form of EASD slope is no different between models 1 and 2, only the math-

ematical constants that constitute directionality time have changed.

18

Page 45: The Mechanics of Cell Motility and a Unifying Theory for

This generalized directionality time equation can be understood by considering the

following example. Consider the case where the probability of measuring a step length L

between l and l + dl is given by a Poissonian probability density function (PDF), ρL(l) =

1lp

e− llp , and the orientation Θ is independent from step length. Then directionality time

simplifies to

td = tp

(2− c2

c2

)(2.8)

where c ≡ 〈cos Θ〉, and tp =lpv represents the reorientation time, the average time to take

one step in a new direction. The term c is tangent-bias correlation (similar to tangent-

tangent correlation). Values of c2 range from 0, corresponding to no orientational bias

(PDF ρΘ(θ) = 12π , where −π < θ ≤ π), to 1, corresponding to maximal orientational or

anti-orientational bias (PDF ρΘ(θ) = δ(θ) or δ(θ− π), where δ is the Dirac delta function).

Directionality time depends only on the reorientation time and the extent to which the

orientation is biased when a reorientation event occurs, increasing with the former and

decreasing with the latter. In particular, the term 2−c2c2

ranges from 1 at maximal bias, to

∞ at no bias. It may appear odd that td → tp (the equivalent of one step) when the system is

perfectly directional (c2 → 0). However, this is no more than a subtlety of stepping random

walks. There is no change in position defined at t < tp because of the way continuous time

was substituted in for discrete stepping number (t ≈ ntp). Therefore, the minimum time to

determine that movement is directionally biased will always be greater than or equal to tp.

No information can be gained from a random walker that has not yet taken any steps. Since

each step is accompanied by reorientation, this model cannot be used to derive a log-log

EASD slope equation that accounts for persistence.

In order to consider the relation between directionality time and persistence, a con-

tinuous time random walk model must be derived, noted as the PBRW (model 3). This

model is derived in 1D for simplicity using the biased telegrapher equation [93, 96]. To

put the biased telegrapher equation in context, the unbiased telegrapher equation is used

19

Page 46: The Mechanics of Cell Motility and a Unifying Theory for

to derive the dynamics and EASD of persistent random walks that describe the kinematics

of chemokinesis [41, 61], as well as the motion of grasshoppers and kangaroos [93]. In this

model, an object moves with constant speed v, either left (−x direction) or right (+x di-

rection) for some random run time (T (l) or T (r), respectively) before switching directions.

Bias is induced by drawing left and right run times from nonequivalent distributions and is

encapsulated by tangent-bias correlation c = 〈T (r)〉−〈T (l)〉〈T (r)〉+〈T (l)〉 . The log-log EASD slope of the

PBRW (Eq. 2.31) is more complicated than that for the BRW and DD because direction-

ality over short time scales caused by persistence induces a zero-time log-log EASD slope

βPBRW(0) = 2. As t increases and the orientational correlation of persistent motion is lost,

βPBRW(t) dips towards 1. Except when c = 0, βPBRW(t)→ 2 as t→∞, as is the signature

of directionally biased motion.

These βPBRW(t) curves are plotted Fig. 2.2 a for multiple values of c2 (solid curves).

In this plot, time is in units of λ−1+ , which is related to the average run time (persistence

time). At sufficiently large time scales

βPBRW(t > tBRW) ' βBRW(t; td) =1 + 2t

td

1 + ttd

(2.9)

where tBRW is the convergence time above which the difference between βPBRW and βBRW

is less than 5% (Fig. 2.2 b, curve). Directionality time can be measured by fitting βPBRW(t)

to the corresponding βBRW(t; td) at time scales larger than t = tBRW. The resulting mea-

surement of td from this fit is given by

td =1

λ+

2(1− 2c2)

c2. (2.10)

As with the BRW, td →∞ for a random walk that is unbiased. When the bias is sufficient

(12 ≤ c2 ≤ 1), the gap in time scales between short time scale persistent directionality and

long time scale biased directionality vanishes such that the random walk appears directional

at all time scales. By construction, we redefine negative values of directionality time to

20

Page 47: The Mechanics of Cell Motility and a Unifying Theory for

0 in this domain to be consistent with the interpretation that directionality time is the

observation time necessary to determine that motion is directional (Fig. 2.2 c).

0 20 40 60 80 1000.8

1.0

1.2

1.4

1.6

1.8

2.0

2.2

time, t (λ-1)+

log−

log

EASD

slo

pe, β

(t)

approximate (BRW)

exact (PBRW)

c2 = 0.00

c2 = 0.04

c2 = 0.16c2 = 0.36

c2 = 0.64

c2 =1.00

a b

c

0 0.5 10

20

40

60

80

dire

ctio

nalit

y tim

e, t d

(λ-1

)+

tangent-bias correlation, c2

0

5

10

<1%

<5%

<10%

>10%

15

conv

erge

nce

time,

t BRW

(λ-1

)+

Figure 2.2. Log-log EASD slopes of the 1D-PBRW and its correspon-dence to the directionality time model. (a) Log-log EASD slopes of the1D-PBRW model (βPBRW(t), solid curves) asymptotically approach the direc-tionality time model (βBRW(t; td), dashed curves) given a specific value of di-rectionality time td. These slopes are plotted against time for different valuesof the squared tangent-bias correlation, c2 = 〈cos Θ〉2. In these plots, timeis nondimensionalized by the persistence time-like parameter, λ−1

+ . Correlatedorientation over short time scales gives β(0) = 2 regardless of directional bias.Except when the motion is non-directional (i.e. c2 = 0), βPBRW(t) dips from2 to 1, before asymptoting back toward 2 as t → ∞. (b) The convergencetime above which the difference between between βPBRW(t) and βBRW(t; td) isless than 10, 5, and 1% are shown in gray scales plotted against c2. The 5%convergence time, denoted tBRW (black solid curve), is used for calculations inthis work. (c) The corresponding directionality time plotted against c2. Thetransition from random motion to directional motion

(βBRW(td) = 3

2

)does not

occur in the domain, 0.5 ≤ c2 ≤ 1. In this domain, directionality time is definedto be zero because the motion appears directional at all time scales.

The result βDD(t) = βBRW(t) ' βPBRW(t) at large time scales implies that one unique fit

function can be used to measure td from any set of idealized migration path measurements.

For brevity, we refer to this fit function as the βBRW-model. The βBRW-model is independent

21

Page 48: The Mechanics of Cell Motility and a Unifying Theory for

of the type of random walk process, returning values of td ranging from 0 when all time

scales appear directional, to ∞ when motion is completely unbiased.

2.3. Computational Modeling to Test Robustness. We used computer simula-

tions of 2D-PBRWs to test the robustness of the βBRW-model. A diagram of the 2D-PBRW

is shown in Fig. 2.3 a. Continuous time was simulated in time steps of δt. Run times,

Tn (times between reorientations), were drawn from Poisson distributions corresponding

to persistence times, tp, ranging from 0.4 to 3.6 s (Fig. 2.3 b). At the end of each run

time, a reorientation occurred. Polar angles, Θn, were drawn from von Mises distributions

[97] centered about θ = 0 in the ex direction. The von Mises distribution width, set by

parameter κ (Fig. 2.3 c), was selected from values between 0 (uniform distribution) and

10 (relatively narrow Gaussian-like curve). The correspondence between κ and 〈cos Θ〉 is

shown in Fig. 2.4. Positions along the 2D-PBRW were sampled periodically with sample in-

terval ∆t and Gaussian noise with width σm was added to each sampled position coordinate

so that simulated data would resemble experimental data similar to those of directionally

migrating neutrophils [1, 48].

To investigate the effects of measurement error on measures of directionality time, two

ensembles of simulated trajectories were computed, one with σm = 0 and other with σm =

1µm. These simulated trajectories were ergodic (data not shown) and EASD was equated

to MSD (β(t)↔ β(τ), t interchangeable with τ). The corresponding measurements of β(τ)

are shown Fig.2.3 d (black and red, respectively). Visually, the deviations between the two

data sets are greatest at short time scales (shaded region). Specifically, β(0) = 0 when

σm > 0, instead of β(0) = 2 when σm = 0. Effectively, measurement error hides the ability

to identify persistence.

Recall that βPBRW(t) converges to the βBRW-model above time tBRW (see Eq. 2.9 and

Fig.2.2 b). Measurement error sets an additional convergence time, tσm , above which the

βBRW-model is valid. This time is derived analytically in the Appendix B (Sec. 4.3) by

22

Page 49: The Mechanics of Cell Motility and a Unifying Theory for

00

0.5

1.0

1.5

-π π

κ=0,2,10

ρΘ

(ra

d-1

)

polar angle, θ (rad)

êy

êx

0 2 4 6 8 10

0.2

0

0.4

0.6

0.8tp = 0.4,2.0,

3.6s

run time, t (s)

ρT

(s-1

)

0 50 100 150 200 2500

0.5

1.0

1.5

2.0

2.5

σm = 1µm

τmin = 50.4s

σm = 0µm

βBRW-model

single parameter

exponential !tlo

g-l

og

MS

D s

lop

e, β

(τ)

time interval, τ (s)

Θn+1L

n Ln=vTn

Θn

δtσm

a

b d

c

Figure 2.3. Measuring directionality time from simulated 2D-PBRWmigration paths that resemble experimental data. (a) A schematic of the2D-PBRW model and its corresponding random variables. A simulated randomwalker travels with constant speed v in a straight line with polar angle Θn. Aftertraveling in one direction for a run time Tn, the random walker reorients andcontinues in a new direction, Θn+1. For each random walk, positions were sam-pled in increments of sampling interval ∆t. Measurement error σm was addedto each sampled position coordinate to generate migration paths resemblingexperimental data. (b) Run times, Tn, were randomly selected from Poissondistributions with average run times of tp, known as the persistence time. (c)Polar angles, Θn, were randomly selected from the von Mises distribution withbias factor κ. The value κ= 0 corresponds to unbiased motion. Increasing κcorresponds to more directional bias. (d) Log-log MSD slopes showing β(τ) fortwo values of centroid measurement error: σm=0, 1µm (black and red, respec-tively; ×’s for τ < τmin and •’s for τ > τmin; parameters tp = 3.6 s, κ = 1.5,v= 0.3µm/s, ∆t = 1 s; ensemble sizes of n = 400). The minimum time inter-val above which the βBRW-model fit decouples from the effects of measurementerror and persistence was estimated at τmin = 50.4 s. Fitting the βBRW-modelto data above this minimum time interval (green curve) gave virtually identicaldirectionality times for both values of measurement error. Data in the τ < τmin

domain (×’s in the shaded region) were not used for fitting. A single parameterexponential fit function, 2− e−t/td (cyan curve), also fit in the τ > τmin domain,shows that a heuristic fit model does not accurately measure directionality time.

23

Page 50: The Mechanics of Cell Motility and a Unifying Theory for

0 5 1 0

0

0.5

1.0

von Mises bias factor, κta

ng

en

t-b

ias

corr

ela

tio

n, <co

s(Θ

) >

Figure 2.4. Tangent-bias correlation, c = 〈cos Θ〉, plotted against thevon Mises bias factor, κ. This curve shows the correspondence betweenc used in the analytically derived biased random walk models, and κ used inthe simulated biased random walk model. This curve is calculated analytically

from the definition of the von Mises distribution. Specifically, 〈cos Θ〉 = I1(κ)I0(κ)

where In is the modified Bessel function of the first kind, order n. To sample csomewhat evenly, values κ = 0, 0.5, 1, 1.5, 2, 6, 10 were used in the simulationsof persistent biased random walks discussed in this manuscript.

adding a measurement error term to the EASD of the PBRW model and calculating the

time above which this modified PBRW model converges to within 5% of the βBRW-model:

tσm ≈ 4.5√

2dσm

vrms(∞). (2.11)

(Fig. 2.5) Here, d is the number of dimensions, and vrms(∞) = limτ→∞

vrms(τ) = |c|v is the

root mean squared (RMS) speed asymptote, a measurable quantity that is sampling interval

independent.

24

Page 51: The Mechanics of Cell Motility and a Unifying Theory for

0 0.5

ε=10

ε=5ε=2ε=0

1.00

50

100

conv

erge

nce

time,

t BRW

(λ-1

)

+tσm

exact

tangent-bias correlation, <cos(Θ)>2

Figure 2.5. Time scales at which βPBRW(t) converges to the βBRW-model when measurement error is nonzero (σm > 0). Times abovewhich the difference between βPBRW(t) and the βBRW-model is less than 5% are

shown as solid curves correspond to values of the constant ε =√

2dσmλ+

v, which

is proportional to measurement error. The convergence time is estimated by theequation λ+tσm ≈ 4.5C

c, as derived in Appendix A. When implementing a βBRW-

model fit to measure directionality time, β(τ) curves are fit at time intervalsτ > tσm to decouple directionality time from measurement error.

A minimum fitting time, τmin, is defined as the larger of tBRW and tσm . In this example,

τmin = tσm ≈ (4.5)(2) (1µm)(0.6)(0.3µm/s) = 50.4 s. Data in the τ > τmin domain (Fig. 2.3 d, •’s)

were fit to the βBRW-model (green curve) illustrating that the effects of measurement error

and persistence can be decoupled from measures of directionality time by leaving out data

below the time interval τmin. Although backwards extrapolation is required to measure

directionality time in this example, backwards extrapolation is not generally required (i.e.

when td > τmin).

The possibility of using a heuristic model to measure directionality time is investigated

by fitting the data in the τ > τmin domain to a single parameter exponential function, 2−

e−t/td (Fig. 2.3 d, cyan curve). This model does not fit the data as well as the βBRW-model,

returning overestimates of directionality time. Overall, measurements of directionality time

using a heuristic model such as this generally satisfy the overall objective of measuring the

time scale at which motion transitions from random to directional. However, lost when

25

Page 52: The Mechanics of Cell Motility and a Unifying Theory for

using a heuristic fit model is the consistency and interpretability of directionality time with

respect to an analytical framework that characterizes directional motion.

Simulated migration paths corresponding to a range of parameter combinations (tp, κ, σm)

were analyzed to answer the following questions related to the robustness of the βBRW-

model. For what type of motion does the βBRW-model fit well, and when does it fail? And

to what extent does directionality time decouple from measurement error by fitting at times

τ > τmin? Goodness of fit was tested using the reduced chi-squared metric, χ2ν (Fig. 2.6 a,

left column). Fits labeled “good” (unshaded regions) were those with χ2ν ≤ 1. All other fits

were labeled “problematic” (yellow shaded regions). The βBRW-model fits were good over

most of the parameter space. For non-directional motion (κ = 0), many of the βBRW-model

fits were problematic, although the model still returned characteristically large directional-

ity times. The βBRW-model fits were also problematic for directional motion with small tp

and td compared to the time scale of noise, tσm (i.e. tp and td � tσm). Measurements of

directionality time indicated that values of td increased slightly as σm increased (Fig. 2.6 a,

right column). This weak monotonic coupling between td and σm occurs because measure-

ment error decreases the value of β(τ) at all time scales, thus slightly increasing the value

of td obtained from fitting to the βBRW-model. This coupling was only significant at the

lower limits of tp (Fig. 2.6 b). Overall, these simulation results show that if motion is direc-

tional and tp and td are not masked by the time scale of noise, tσm , then the βBRW-model

fits robustly and returns measurements of directionality time that are negligibly skewed by

measurement error.

For comparison, the sampling interval dependent metrics (TAD, TAD persistence, tor-

tuosity, and tangent-tangent correlation) were calculated for the simulation trajectories

(Fig. 2.7). As expected, the numerical values associated with these metrics varied sig-

nificantly with sampling interval, coupling strongly to measurement errors at the smaller

sampling intervals.

26

Page 53: The Mechanics of Cell Motility and a Unifying Theory for

0 50 100

0.5

1.0

dire

ctio

nal m

igra

tion

e�

cien

cy,

<R(t) > / v

t

directionality time, td (tp)

σ m (µ

m)

012

+λ-1 (s)

3.60.4

<R(t)> vt0x

r

y ρR

σc = 2µm

σc = 1µm

σc = 0µm

von Mises bias factor, κ0 2 6 10 0 2 6 10

directionality time, td (s)

0 25 50

robustness

problematic �tgood �t

0.41.22.02.83.6

0.41.22.02.83.6

0.41.22.02.83.6pe

rsis

tenc

e tim

e, t p

(s)

a b

Figure 2.6. Additional data plots investigating the robustness of thedirectionality time model. (a) Goodness of fit and directionality time mea-surements across a range of parameter combinations (same parameters as inFig 2.3 d but ∆t = 10 s and tp, κ, σc all vary). Squares on phase diagramsindicate parameter combinations where ensembles of random walks were sim-ulated (n = 4000 migrations per ensemble). Goodness of fit was measured by

calculating reduced chi squared, χ2ν =

∑n

(β(τn)−βBRW(τn;td))2

σ2β(τn)

, where σβ is the

standard deviation (spread) on values of β(τn) used for fitting. A fit was cat-egorized as “good” if χ2

ν ≤ 1, otherwise it was categorized as “problematic.”The directionality time model worked robustly except near κ = 0 and when tpand td � τmin. The latter is the limit where measurements of directionalitytime can no longer be resolved. (b) Directional migration efficiency is inverselyproportional to directionality time (in units of tp). Directional migration effi-ciency is defined as the ensemble averaged distance travelled 〈R(t)〉 divided bythe distance a ballistic walker would have travelled if it had the same speed, vt.This plot shows that td is a proxy for the distance a walker will travel, and thattd best decouples from measurement error as persistence time tp increases.

2.4. Application to Real Data. Under idealized circumstances, the βBRW-model

may be used to fit ensembles of data. However, as we saw with the simulated data, adding

measurement error made the data less ideal. Small ensemble sizes, nonergodicity, and

changing experimental conditions all cause experimental data to vary. To improve the

27

Page 54: The Mechanics of Cell Motility and a Unifying Theory for

turning angle (rad)

−π π π−π

20

2

pro

ba

bil

ity

0

0.1

0.2

0.7

0 µm0.1µm

1s 4s 20s

sampling interval

∆tσc = ...

0 50 1000.2

0.4

0.620s

4s

1s

0.8

1.0

time interval, τ (s)

tan

ge

nt-

tan

ge

nt

corr

ela

tio

n

ba

c

1s 4s 20s0.8

0.9

1.0

pe

rsis

ten

ce

sampling interval

1s 4s 20s0.4

0.6

0.8

1.0

tort

uo

sity

sampling interval

Figure 2.7. Sampling interval dependent metrics applied to simulationdata (2D-PBRW, tp = 3.6 s, κ = 1.5, and v = 0.3µm/s). The position ofthe random walk was sampled every ∆t = 1, 4, or 20 s, with sampling errorsof σm = 0 (blue bars and thin blue curves) or 0.1µm (green bars and thickgreen curves). (a) Ensemble averaged turning angle distributions (TAD) basedon sampling intervals of 1 s (solid curves) and 4 s (short-dashed curves). TADpersistence is the fraction of all turning angles between ±π

2(inset, error bars

are standard error of the ensemble mean). When σm = 0, TAD persistence issmallest when ∆t ≈ tp. This is time scale of reorientation and consequently thesampling interval with which the motion appears most random. Small amountsof measurement error (σm = 0.1 µm) hide the persistence of motion that wouldotherwise be measured at the smallest sampling interval ∆t = 1 s. Thesedata show that TAD persistence is sampling interval dependent. (b) Ensembleaveraged tortuosity (error bars are standard error of the ensemble mean). Aswith TAD persistence, tortuosity is sampling interval dependent, increasing with∆t. (c) Tangent-tangent correlation curves. Tangents are calculated based onthe forward displacement between nearest sampled points. Therefore tangent-tangent correlation is sampling interval dependent. In particular, randomnessat short time scales are not resolved as ∆t increases. The persistence time,tp, can be back-measured from the tangent-tangent correlation curves if thereis sufficient temporal resolution (∆t ≤ tp). Each of these metrics are samplinginterval dependent and couple to measurement error at short sampling intervals.Hence, these metrics are not generally comparable from one experiment to thenext.

reliability and accuracy of the directionality time measurement, we now examine three

causes of deviation between experimental measurements of β(τ) and the βBRW-model. These

causes of deviation are: position variance, persistence, and nonergodicity. The causes of

28

Page 55: The Mechanics of Cell Motility and a Unifying Theory for

position variance can be further subdivided into three categories: implicit position variance,

measurement error, and parametric variance. Methods for handling these deviations are

addressed below.

Implicit position variance is already accounted for in the βBRW-model and does not

cause deviations. Deviations caused by persistence and measurement error induced position

variance can be decoupled from the βBRW-model by fitting above a minimum fit time τmin, as

discussed above (c.f. Fig. 2.3 d). Therefore, one only needs to be concerned with deviations

caused by parametric variance and nonergodicity.

Parametric variance, denoted σ2p, is the variance in distance traveled caused by variance

of random walk parameters across the ensemble (population heterogeneity). For example,

the variance in distance traveled at long time scales due to a spread in the instantaneous

speed parameter, δv, is σ2p = c2(δv)2t2. Parametric variance systematically increases mea-

surements of EASD and is the dominant cause of deviation between β(t) and the βBRW-

model at large time scales (see Eq. 2.37). Deviations caused by parametric variance cannot

be corrected by fitting above a minimum fit time, τmin.

Ergodicity is the conversion factor that maps EASD to TASD (and MSD), given in

Eq. 2.3. When a process is ergodic, EASD and MSD are equivalent, and so are calculations

of β based on either EASD or MSD. When a process is nonergodic, for example when the

instantaneous migration speed changes significantly over time across the entire ensemble,

measurements of β(τ) based on MSD will deviate from the βBRW-model which is derived

from the EASD. Deviations caused by ergodicity occur at all time scales and, like deviations

caused by parametric variance, cannot be corrected for by fitting above a minimum fit time,

τmin.

We show in Appendix B (Sec. 4.3, Eq. 2.44) that log-log MSD slope β(τ) can be de-

composed into three terms that correct for deviations caused by position variance and

29

Page 56: The Mechanics of Cell Motility and a Unifying Theory for

ergodicity:

β(τ) = βBRW(τ) + βσ(τ) + βξ(τ). (2.12)

The term βσ accounts for deviations caused by position variance (primarily parametric

variance), and βξ accounts for deviations caused by nonergodicity. The former is calcu-

lated by using experimentally observed variances in distance traveled to estimate how much

parametric variance and measurement error have an effect on MSD, and the corresponding

log-log MSD slope. The resulting position variance modification term is

βσ(τ) =d log

[〈r2(τ)〉

〈r2(τ)〉−σ2r(τ)

]d log[τ ]

(2.13)

where σ2r (τ) is the variance on the mean distance traveled at time interval τ (see Appendix B,

Sec. 4.3, for the derivation). The ergodicity term can also be calculated from experimental

data using the relation βξ(τ) = d log[ξ(τ)]d log[τ ] , but this term can be relatively noisy because

time averaging cannot be applied to reduce statistical noise. Often, migration data can be

made nearly ergodic by choosing to analyze truncated migration segments over which the

instantaneous speed remains relatively constant. In such cases, no ergodicity correction is

necessary. When the migration paths cannot be partitioned into ergodic segments, βξ(τ)

should be estimated by simulation rather than directly calculated from the experimental

data. A recipe for these simulations can be found in Appendix B (Sec. 4.3).

2.5. Neutrophil Chemotaxis. In this section, an implementation of the βBRW-model

to measure directionality time is demonstrated on migration paths of chemotactic human

polymorphonuclear neutrophils following the step-by-step procedure outlined in Fig. 2.8.

Neutrophil migration paths, r(t), and centroid measurement errors, σm, were obtained

from O’Brien et al. [1]. Two sets of data were analyzed, each containing the trajectories of

chemotactic neutrophils migrating on the 2D surfaces of polyacrylamide gels (Young’s modu-

lus: 10 kPa) towards a source of the chemoattractant formyl-methionyl-leucyl-phenylalanine.

30

Page 57: The Mechanics of Cell Motility and a Unifying Theory for

The difference between the two data sets was the coating on the gel surface, either human

fibrinogen (Fgn) or human type IV collagen (Col IV). Tangent-tangent correlation measure-

ments gave a persistence time, tp, upper bound of 5 s (Fig. 2.10 c). EASD was calculated but

was noisy. Therefore, MSD was also calculated to reduce the statistical noise (Fig. 2.9 a).

Data was checked for ergodicity by measuring the ensemble averaged instantaneous speed

(EAIS) over time. EAIS was approximately constant when the migration paths were trun-

cated at 400 s and no ergodicity correction was applied to this data. If EAIS were changing

significantly in time, the ergodicity correction term βξ(τ) could have been calculated using

the recipe in Appendix B (Sec. 4.3).

track migration:

determine positions, r(t)

& measurement error, σc

calculate EASD, <r2(t)>Is the data noisy?

calculate ergodicity, ξi2(τ)

& ergodicity correction βξ(τ)

via simulations

measure persistence time,

tp, from tangent-tangent

correlation

measure rms speed

asymptote, vrms(∞)

calculate minimum

βBRW-model !tting

time scale, τmin

Fit to βBRW-model to measure

directionality time, td

equate MSD with EASD,

<r2(τ)> = <r2(t=τ)>Is the data nonergodic?

calculate MSD, <r2(τ)>

yes

noyes

no

calculate log-log MSD slope

β(τ) and correct with terms

βσ(τ) and, if applicable, βξ(τ)

Figure 2.8. Step-by-step flow-chart for processing experimental mi-gration data to measure directionality time.

Next, vrms(∞) was calculated (Fig. 2.9 a, bottom-right inset): vrms(∞) ≈ 0.15 and

0.13µm/s for Fgn and Col IV, respectively. Taken together with tp, the minimum fit time

was calculated using Eq. 2.11 for both data sets: τmin ≈ 60 s and 69 s for Fgn and Col

IV, respectively. These minimum fit times were used later to fit the variance corrected

31

Page 58: The Mechanics of Cell Motility and a Unifying Theory for

a

b

101 20 30 50 200 300102

101

102

103

104

MS

D, <r

2> (µ

m2)

Fgn Col IV

50µm

0.1

0.2

0.3

100 2000 300

time interval, τ (s)

RM

S s

pe

ed

, vrm

s

(µm

/s)

time interval, τ (s)

slope = 1

slope = 2

0 50 100 150 200 250 300

0

0.5

1.0

1.5

2.0

2.5

log

-lo

g M

SD

slo

pe

, β(τ

)

time interval, τ (s)

0

20

30

10

Fgn Col IV

dir

ect

ion

ali

ty t

ime

, td (

s)

ligand

Fgn

Col IV

Figure 2.9. Measuring directionality time from experimental data. (a)The observed mean squared displacements (MSDs) of chemotactic human poly-morphonuclear neutrophils migrating on fibrinogen (Fgn, blue, circles, n=25)and collagen IV (Col IV, red, squares, n=19) coated polyacrylamide gels witha Young’s modulus of 10 kPa. The shaded regions indicate the standard errorsof the MSD. The corresponding trajectories (top-left) and RMS speed (bottom-right) are also shown. The RMS speed asymptotes, vrms(∞), for Fgn and ColIV are approximately 0.15 and 0.13µm/s, respectively. (b) Log-log MSD slopescorrected for position variance, β−βσ, are plotted against time interval τ and fitto the βBRW-model (solid curves) to obtain directionality times, td. The shadedregions indicate standard errors on log-log MSD slope at time intervals abovethe minimum fitting time. The resulting directionality times are depicted in thebar graph to the bottom-right, with error bars corresponding to 68% confidenceintervals.

32

Page 59: The Mechanics of Cell Motility and a Unifying Theory for

log-log MSD slope, β − βσ (see Fig. 2.9 b), to the βBRW-model and time intervals τ > τmin

(regions with shaded error bars). The resulting directionality times were 10.7 ± 1.0 s and

21.2 ± 5.1 s for Fgn and Col IV, respectively (Fig. 2.9 b, bottom-right inset). With 19

neutrophil trajectories in the Col IV data set, compared to 25 in the Fgn data set, the

Col IV fit was noisier. For comparison, the sampling interval dependent metrics (TAD,

TAD persistence, tortuosity, and tangent-tangent correlation) were calculated for this data

(Fig. 2.10). The numerical values of these measurements vary significantly with sampling

interval and their coupling to measurement errors is unknown.

3. Discussion and Future Direcitons

We have completed a three part description of directionality time, beginning with an

analytical derivation, checked by computational simulations, and applied to noisy real world

ensembles of directed migration paths. In comparison to the sampling interval dependent

metrics (see Figs. 2.7 and 2.10), the directionality time metric is nearly independent of

parameters constrained by the experimental apparatus and/or chosen arbitrarily by humans.

Whereas speed and persistence time are sufficient for characterizing non directed migration,

one additional metric, directionality time, along with speed and persistence time suffices to

characterize directed migration.

The theoretically derivation of the directionality time model and its corrections to ac-

count for measurement error (βσ) and nonergodic motion (βξ) are a step forward for the

analysis of directed motion. Of the two corrections, however, only measurement error was

fully demonstrated here. The experimental data used in this demonstration was nonergodic,

but made to be ergodic by constraining the analysis time window to 400 s. In Fig. 2.9 b,

the experimentally measured log-log MSD slope was corrected for measurement error using

Eq. 2.13. An explicit demonstration of the difference between the non corrected (β) and

measurement error corrected (β−βσ) is shown in Fig. 2.11. Without the measurement error

correction, the corresponding log-log MSD slopes asymptoted to values of approximately

33

Page 60: The Mechanics of Cell Motility and a Unifying Theory for

100 200 30000.2

0.4

0.6

0.8

1.0

time interval, τ (s)

tan

ge

nt-

tan

ge

nt

corr

ela

tio

n

60s

10s

turning angle (rad)

−π ππ−π

20

2

10s 60s0.6

1.0

0.8

turn

ing

an

gle

pe

rsis

ten

ce

∆t

pro

ba

bil

ity

60s

10s

0

0.05

0.10

0.15

0.20a b

c

ColFgn

10s 60s

sampling interval,

∆t

Fg

n

Co

l IV

tort

uo

sity

10s 60s0.4

0.6

0.8

1.0

∆t

Fg

n

Co

l IV

Figure 2.10. Sampling interval dependent metrics applied to chemo-tactic polymorphonuclear neutrophils on fibrinogen (Fgn, blue colors)and human collagen IV (Col IV, red colors) coated polyacrylamidegels with elastic modulus of 10 kPa. (a) Ensemble averaged turning an-gle distributions (TAD) are plotted based on measurements of turning anglesat time intervals of 10 s (solid curves) and 60 s (dashed curves). TAD persis-tence, calculated as the fraction of all turning angles between ±π

2, is shown in

the inset (error bars are standard error of the ensemble mean). (b) Ensembleaveraged tortuosity measured at the same sampling intervals (error bars arestandard error of the ensemble mean). (c) Tangent-tangent correlation curvesalso measured at sampling intervals ∆t = 10 and 60 s. Tangents are calculatedbased on the forward displacement between nearest sampled points. Thereforetangent-tangent correlation curves show more correlation when calculated atthe 60 s sampling interval, compared to the 10 s interval. For both Fgn andCol IV, tangent-tangent correlation curves drop towards their asymptote at apersistence time of tp < 10 s. Regardless of the sampling interval, chemotaxison Fgn is more correlated than chemotaxis on Col IV, a result that is consis-tent with our measurements of directionality time. Each of these metrics aresampling interval dependent and couple to measurement error. Hence, thesemetrics are not generally comparable from one experiment to the next.

1.7 and 1.85 (Col IV and Fgn, respectively), instead of 2 as predicted by the βBRW-model.

With the measurement error correction, these asymptotes came closer to a value of 2, at

1.8 and 1.9, respectively. Overall, the measurement error correction moved the experimen-

tally measured log-log MSD slope closer to the analytically derived βBRW-model, a partial

indication that the correction was valid.

34

Page 61: The Mechanics of Cell Motility and a Unifying Theory for

The erogidic correction, βξ, needs to be analyzed in a similar manner. First, the βξ cor-

rection should be confirmed using measurement error free (clean) nonergodic data generated

by computer simulation. Second, the βξ correction should be demonstrated explicitly with

nonergodic experimental data. These two tasks have begun in earnest, with the latter being

done on a mix or ergodic and nonergodic neutrophil migration paths collected by O’Brien

et al. to determine the precise role of extracellular matrix composition on mechanosensing

and neutrophil migration.

Part of this work is discussed in the next chapter, but there is still no demonstration

of the βξ correction because the findings of the next chapter were made at a time before

we took ergodicity into consideration. Instead, in the next chapter, we use a heuristically

derived directionality time model that fits the data well and provides a good relative com-

parison of directionality times across the datasets it is applied to. The heuristic measure of

directionality time is used to determine that neutrophils regulate chemotactic honing as a

function of substrate stiffness using the β2-integrin, CR3.

0 50 1001.0

1.5

2.0

time interval, τ (s)

β−βσ, Fgnβ−βσ, Col IV β, Fgn β, Col IV

log

-lo

g M

SD

slo

pe

, β(τ

)

Figure 2.11. Explicit demonstration of the measurement error correc-tion on experimentally measured log-log MSD slope. Log-log MSD slopedata correspond to Fig. 2.9. Compared to the non corrected measurements oflog-log MSD slope (β), the measurement error corrected slopes (β − βσ) havelong time interval asymptotes closer to 2. Therefore, the measurement error cor-rected data points are better explained by the βBRW model, which asymptotesto 2.

35

Page 62: The Mechanics of Cell Motility and a Unifying Theory for

4. Supplementary Material

4.1. Methods.

4.1.1. Migration Paths and Centroid Measurement Error. Differential interference con-

trast (DIC) image sequences capturing directionally migrating Polymorphonuclear Human

Neutrophils were obtained from Ref. [1] along with cell centroid positions, ri(tn), where i is

the migration path index and tn is the time measured in multiples of the sampling interval,

tn = n∆t (n = 1, 2, ...). Centroid measurement error was estimated as follows. A cell

was chosen at random and manually outlined five times. The five corresponding centroid

positions were determined using the regionprops algorithm in MATLAB (the MathWorks;

Natick, MA). The centroid measurement error of that cell, σm, was calculated as the RMS

displacement from the mean centroid position.

4.1.2. 2D PBRW Simulations and Data Fits. All simulation data were generated in

MATLAB. All data was fit using custom MATLAB software. This code is available for

download at http://www.github.com/aloosley (see Chapter 7, Sec. 1, pg. 170). Data fits

were calculated using the Levenburg-Marquardt least squared fitting algorithm [98] built

into MATLAB.

4.2. Appendix A: Analytical Modeling.

4.2.1. Appendix A Introduction and Notation. In this appendix, three random walk

models are discussed. Positions, R, are specified by a two dimensional (2D) Cartesian

coordinate system with unit vectors ex and ey (Fig. 2.1 c). The orientation of movement

is described by polar angle θ (π < θ ≤ π) defined with respect to the positive x-axis. The

random walker begins by moving with speed V1 in a particular direction Θ1 for a particular

time T1, before reorienting and moving with a new speed V2, in a new direction Θ2, for

another time period given by T2. This process continues and is altogether described by

speeds Vn, angles Θn, and step durations Tn (n = 1, 2, ...), which are random values of

speed v, polar angle θ, and time t, respectively. Each random variable is associated with a

36

Page 63: The Mechanics of Cell Motility and a Unifying Theory for

notation usageδt simulation time step

∆t sampling intervalt timeτ time interval

bold variables vectorsCAPITALIZED variables random variables

ρT(t)

probability density function (PDF),(the probability of measuring random variable

T between the values t and t+ dt)

Rrandom variable for position used in analytical models(corresponding to the PDF, ρ

R(r))

Θrandom variable for 2D orientation(the polar angle of the instantaneous velocity vector)

overline average over time t〈〉 average over ensemble or analytical PDF

c = 〈cos Θ〉 tangent-bias correlation (averaged over ρΘ(θ))

r experimentally measured positionv instantaneous velocity

vrms root mean squared speedφ turning angle (not to be confused with polar angle θ)

v(t+ τ) · v(t) tangent-tangent correlationtp persistence timetd directionality time

Table 2.1. Mathematical notation used to derive directionality timeand interpret cell motility.

probability density function (PDF), ρV (v), ρΘ(θ), and ρT (t). For each random walk model

described in this appendix, we define the relevant variables, parameters, and distributions

using the mathematical notation summarized in Table 2.1, and use these to determine the

ensemble averaged squared displacement (EASD), 〈R2(t)〉, and log-log EASD slope, β(t).

From beta(t), we derive the model specific meaning of directionality time, td, the measure

that is central to this work. A universal, random walk model invariant fit function is derived

to measure td from EASD data.

4.2.2. Model 1: Drift Diffusion (DD). A walker undergoing a Diffusion in 2D (without

drift) obeys the following rules:

37

Page 64: The Mechanics of Cell Motility and a Unifying Theory for

(1) Stepwise Position: The position after n steps is given by the random variable

Rn =n∑j=1

VjTj [cos(Θj)ex + sin(Θj)ey].

(2) Constant step duration: ρT (t) = δ(t− λ−1), i.e. Tj = λ−1 for all j

(3) No directional bias: ρΘ(θ) = 12π where −π < θ ≤ π

(4) Constant track speed: Vj = v

where δ(...) is the Dirac delta function, λ is the reorientation frequency, and the other terms

are defined in the Appendix A Introduction. The notion for average step duration, λ−1, is

used for consistency with the models below. The addition of drift to a diffusion process adds

an additional term to the stepwise position equation. Without loss of generality, one can

define a drift velocity along the x-axis, u = u ex. Subsequently, Rule 1 would be redefined

as

1’. Stepwise Position: The position after n steps is given by the random variable

Rn =n∑j=1

v

λ[cos(Θj)ex + sin(Θj)ey] +

nu

λex.

One example of DD would be the motion of an object trapped at an air-water interface

in the presence of a surface current (air or water). If the object were charged, drift could

also be induced by applying a uniform electric field (i.e. carrier drift) [99].

The ensemble averaged squared displacement (EASD) of a DD in d dimensions has been

shown to be ⟨R2(t)

⟩= 2dDt+ u2t2 (2.14)

where D ≡ v2

λ is the so-called diffusion constant [93]. At large times, motion is directional

(〈R2(t)〉 → u2t2 ∼ t2) with a root mean squared (RMS) speed, vrms(t) =

√〈R2(t)〉t , that

asymptotes to vrms(∞) = u. Differentiation in log-log coordinates

(d log[〈R2(t)〉]

d log[t] = t〈R2(t)〉

d〈R2(t)〉dt

)gives the log-log EASD slope

βDD(t; t0) =1 + 2 t

t0

1 + tt0

(2.15)

where time constant, t0 = 2dDu2 , incorporates both model parameters D and u. Defining

directionality time as the time when β = 32 provides a time scale where the migration

38

Page 65: The Mechanics of Cell Motility and a Unifying Theory for

transitions from diffusive to directionally drifting. This definition gives td = t0 or

td =2dD

u2. (2.16)

The greater the diffusion constant and/or lower the drift velocity, the more time is required

to observe a directional bias.

4.2.3. Model 2: 2-D Stepping Biased Random Walk (2D-SBRW). A walker undergoing

a 2-D Stepping Biased Random Walk (2D-SBRW) obeys the following rules:

(1) Stepwise position: The position after n steps is given by the random variable

Rn =

n∑j=1

Lj [cos(Θj)ex + sin(Θj)ey].

(2) Random step length: Lj is drawn from a PDF ρL(l)

(3) Directional Bias: Θj is drawn from a circular distribution (i.e. von Mises,

Wrapped Normal, etc.[97]) with PDF ρΘ(θ) that is symmetric about θ = 0 (−π <

θ ≤ π).

(4) Constant track speed: Vi = v

It can be shown that the EASD after n steps of a 2D-SBRW is given by

⟨R2n

⟩= n

⟨L2⟩

+ n(n− 1) 〈L cos Θ〉2 + 〈L sin Θ〉2 (2.17)

(see Ref. [94] for a derivation). If the orientation angles {Θj} are independent of the step

lengths {Lj} and the directional bias PDF is symmetric about θ = 0 such that 〈sin Θ〉 = 0

(c.f. rule 3), then the EASD simplifies to

⟨R2n

⟩= n

⟨L2⟩

+ n(n− 1)〈L〉2c2 (2.18)

where c = 〈cos Θ〉 is called tangent-bias correlation because it correlates the random walk

orientation (tangent vectors) to the direction of bias (θ = 0). Squared tangent-bias correla-

tion, c2, ranges from 0, corresponding to no orientational bias(ρΘ(θ) = 1

2π where − π < θ ≤ π),

to 1, corresponding to maximal orientational bias (ρΘ(θ) = δ(θ) or δ(θ − π)).

39

Page 66: The Mechanics of Cell Motility and a Unifying Theory for

Note that the motion is diffusive (〈R2n〉 ∼ n) when n is small and directional (〈R2

n〉 ∼ n2)

when n is large. This step by step random walk has no persistent directionality at low n

because the direction of motion changes with every step, n. In order to calculate log-log

EASD slope, β(t), an average track speed v is defined and the relation n ≈ vt〈L〉 is used to

transform 〈R2n〉 → 〈R2(t)〉. In the time representation, this is a model of a biased random

walk (BRW). Converting the step lengths into step durations (L = vT ) gives

〈R2(t)〉 ≈ v2

[(〈T 2〉〈T 〉2

− c2

)〈T 〉t+ c2t2

]. (2.19)

Like with DD, the t2 term dominates at large times and RMS speed asymptotes to

vrms(∞) = |c|v. (2.20)

Differentiating in log-log coordinates

(d log[〈R2(t)〉]

d log[t] = t〈R2(t)〉

d〈R2(t)〉dt

)gives the log-log EASD

slope

βBRW(t; t0) =1 + 2t

t0

1 + tt0

(2.21)

where t0 = 〈T 〉

(〈T2〉〈T 〉2−c2

c2

). Note that βBRW has the same functional form as βDD (Eq. 2.15).

Only the mathematical constants have changed. Analogous to the DD treatment, defining

directionality time at the transition point, β = 32 , gives td = t0, or

td = 〈T 〉

〈T 2〉〈T 〉2 − c

2

c2

. (2.22)

This generalized directionality time equation is more easily understood by considering an

example. Consider the case where the probability of measuring a step length T between t

and t+ dt is given by a Poissonian PDF, ρT (t) = 1tp

e− ttp . Then, 〈T 〉 = tp, 〈T 2〉 = 2t2p, and

directionality time simplifies to

td = tp

(2− c2

c2

). (2.23)

40

Page 67: The Mechanics of Cell Motility and a Unifying Theory for

Importantly, directionality time is independent of speed v. It depends only on the average

step duration tp and the extent to which the orientation is biased when a reorientation

occurs (given by c2). Directionality time decreases when either the reorientation rate (1/tp)

or orientational bias (c2) increase. In particular, the term 2−c2c2

scales from 1 at maximal

bias, to ∞ at no bias. It may appear odd that td → tp (the average time of one step) when

the system is perfectly directional (c → 0). However, this is no more than a subtlety of

stepping random walks. Based on the way EASD was converted from the stepping number

domain to the continuous time domain, positions are only allowed to change in increments

of 〈T 〉 = tp. Therefore, the minimum time to determine movement is directionally biased

will always be greater than or equal to tp. No information can be gained from a random

walker that has not yet taken any steps. Since each step is accompanied by reorientation,

this model cannot be used to derive a log-log EASD slope equation that accounts for short

time scale directionality known as persistence. In order to consider the relation between

directionality time and persistence, we consider a continuous time random walk model in

the following section.

4.2.4. Model 3: 1-D Persistent Biased Random Walk (1D-PBRW) in Continuous Time.

A walker undergoing a 1-D Persistent Biased Random Walk (1D-PBRW) obeys the following

rules:

(1) Stepwise position: The position after n steps is given by the random variable

Xn =

n∑j=(odd)

VjT(r)j −

n∑i=(even)

VjT(l)j

, where l and r indicate movement to the

left (-x direction) and right (+ x direction), respectively.1

(2) Variable leftward and rightward step durations: The leftward and rightward

durations are drawn from Poissonian PDFs: ρT

(k)(t) = λke−λkt, where k = l, r.

1Although the stepwise position is useful for introducing the model, below we derive the positiondistribution in continuous time.

41

Page 68: The Mechanics of Cell Motility and a Unifying Theory for

(3) Directional Bias: Set by the difference between the average rightward and left-

ward step durations in rule 2: c ≡ 〈cos Θ〉 = 〈T (r)〉−〈T (l)〉〈T (r)〉+〈T (l)〉 = λl−λr

λl+λr

(4) Constant track speed: Vj = v

The 1D-PBRW walker alternates between two possible directional states, traveling left (l)

or right (r). To allow for persistence, a PDF for the position of the PBRW object, ρX (x, t),

is derived in continuous time from a partial differential equation (as opposed to from the

stepwise position). It can be shown (see Refs. [93, 96]) that PDFs of a 1D-PBRW are

solutions of the biased telegrapher equation

∂2ρX∂t2

+ (λl + λr)∂ρX∂t

+ v (λl − λr)∂ρX∂x

= v2∂2ρX∂x2

(2.24)

where parameters λ−1l and λ−1

r are the average walk times in the leftward and rightward

direction (between alternations), and v is the instantaneous speed (c.f. rules 2 and 4).

Ensemble averaged displacement (EAD), 〈X(t)〉 =∫∞−∞ xρX (x, t)dx, can be calculated in

two steps. Step one, multiply both sides of Eq. 2.24 by x. Step two, integrate by parts over

the entire x-axis. Assuming the position distribution ρX (x, t) and its first spatial derivative

∂ρX

∂x tend to zero as x→ ±∞, integration by parts gives the ordinary differential equation

d2 〈X(t)〉dt2

+ λ+d 〈X(t)〉

dt= vλ− (2.25)

where parameters λ± ≡ λl±λr are introduced to simplify the equations. The frequency λ−

defines the degree of directional bias: leftward when λ− < 0 and rightward when λ− > 0.

The frequency λ+ is inversely related to the average step duration, tp = 12

(λ−1l + λ−1

r

),

known as persistence time [93].

Solving Eq. 2.25 with initial conditions 〈X(t)〉|t=0= 0 and d〈X(t)〉dt |t=0= 0 gives the EAD

〈X(t)〉 = vλ−λ+

[t− 1

λ+

(1− e−λ+t

)]. (2.26)

Note that when there is no bias (i.e. λ− = 0), EAD is identically equal to 0.

42

Page 69: The Mechanics of Cell Motility and a Unifying Theory for

The EASD, 〈X2(t)〉 =∫∞−∞ x

2ρX (x, t)dx, is calculated analogously to the EAD. Mul-

tiplying both sides of the biased telegrapher equation (Eq. 2.24) by x2 and integrating

gives an ordinary differential equation containing 〈X(t)〉 and 〈X2(t)〉 terms. Substitution

of 〈X(t)〉 using Eq. 2.26 gives the ordinary differential equation

d2⟨X2(t)

⟩dt2

+ λ+d⟨X2(t)

⟩dt

= 2v2

{1 +

λ2−λ+

[t− 1

λ+

(1− e−λ+t

)]}. (2.27)

Solving for 〈X2(t)〉 using the initial conditions 〈X2(t)〉|t=0= 0 and d〈X2(t)〉dt |t=0= 0 gives the

EASD

⟨X2(t′)

⟩=

v2

λ2+

[2(3c2 − 1

) (1− e−t′

)+ 2

(1− c2

(2 + e−t

′))

t′ + c2t′2]

(2.28)

where t′ ≡ λ+t is nondimensionalized time and c = λ−λ+

is the tangent-bias correlation as

described in the previous section (c.f. rule 3). When there is bias (c2 > 0),⟨X2(t)

⟩→ c2v2t2

in the large time scale limit (t′ → ∞), corresponding to directional motion with an RMS

speed asymptote

vrms(∞) = |c|v (2.29)

just like in the BRW model (model 2). At maximal bias, c2 = 1, the walker travels in a

straight line with speed v. If there is no bias, c2 = 0, the t′2 term vanishes and the t′ term

indicative of diffusive motion dominates at large time scales (t′ →∞). In this special case,

EASD simplifies to the persistent random walk (PRW) equation commonly used to model

the movement of chemokinetic cells,

⟨X2(t)

⟩= 2v2tp

[t− tp

(1− e−t/tp

)](2.30)

[59, 61, 93], where tp ≡ λ−1+ = 1

2〈T(l)〉 = 1

2〈T(r)〉 is the so-called persistence time, half the

average time the walker moves in one direction before switching to the other.

43

Page 70: The Mechanics of Cell Motility and a Unifying Theory for

The log-log EASD slope

(d log[〈X2(t)〉]

d log[t] = t〈X2(t)〉

d〈X2(t)〉dt

)is

βPBRW(t′; c) =(2− 4c2)t′ + 2c2t′2 + (4c2 − 2)t′e−t

′+ 2c2t′2e−t

(6c2 − 2) + (2− 4c2)t′ + c2t′2 + (2− 6c2)e−t′ − 2c2t′e−t′. (2.31)

These curves are plotted for multiple values of tangent-bias correlation c in Fig. 2.2 a

(solid curves). Of particular interest is the form of a directionality time that would be

calculated from the log-log EASD slope in Eq. 2.31. Persistence over short time scales

causes a βPBRW(0) = 2 and instead of a monotonic increase of βPBRW(t) from 1 to 2 as

with βDD(t) and βBRW(t) (Eqs. 2.15 and 2.21, respectively), βPBRW dips towards a value of

1 before eventually asymptoting back towards 2 as t→∞. At sufficiently large time scales,

the constant terms and terms with decaying exponentials in Eq. 2.31 become negligible and

βBRW(t′; t′0 > t′BRW) ≈1 + 2 t

t′0

1 + t′

t′0

(2.32)

where t′0 = 2(1−2c2)c2

(Fig. 2.2 a, dashed curves). This is known as the βBRW-model. The

convergence time scale above which the βPBRW(t) is approximated by the βBRW-model in

Eq. 2.32, denoted t′BRW, is defined as the time above which∣∣∣βPBRW(t′)−βBRW(t′; t′0)

βBRW(t′; t′0)

∣∣∣ < 5%

(Fig. 2.2 b, solid black curve). For the purpose of fitting the βBRW-model to PBRW data,

choosing to fit data at times above the convergence time scale is important. Fitting at times

above 10λ−1+ will produce good fits because 10λ−1

+ > tBRW for all but the most unbiased

motion, in which case directionality time has little meaning regardless.

Notably, at time scales t′ > t′BRW, βPBRW has the same functional form as the βDD

and βBRW (Eqs. 2.15 and 2.21, respectively). Analogous to both DD and the 2D-SBRW,

defining directionality time as the time when βBRW = 32 gives t′d = t′0 = 2(1−2c2)

c2. In

the interpretation that directionality time is a proxy for the observation time necessary to

distinguish directionality from randomness, we redefine directionality time to vanish in the

44

Page 71: The Mechanics of Cell Motility and a Unifying Theory for

12 < c2 < 1 domain, signifying directional motion at all time scales:

td =

1λ+

2(1−2c2)c2

, 0 < c2 < 12

0 12 < c2 < 1

(2.33)

(Fig. 2.2 c). Importantly, directionality time depends only on the reorientation time scale,

λ−1+ , and reorientation bias, c2. Directionality time does not depend on speed v.

4.2.5. Model 3A: 1-D Persistent Biased Random Walk with measurement error (PBR-

WwE). The 1D-PBRW (model 3) can be modified to account for measurement errors,

σm. Assuming positions along the PBRW are sampled and each position coordinate has

a Gaussian distribution with standard deviation σm, the change to EASD is: 〈X2(t)〉 →

〈X2(t)〉+ 2σ2m, where 〈X2(t)〉 is given in Eq. 2.28. The log-log EASD slope becomes

βPBRWwE(t′; c) =(2− 4c2)t′ + 2c2t′2 + (4c2 − 2)t′e−t

′+ 2c2t′2e−t

(6c2 − 2) + (2− 4c2)t′ + c2t′2 + (2− 6c2)e−t′ − 2c2t′e−t′ +2σ2

mλ2+

v2︸ ︷︷ ︸≡ε2

(2.34)

where the term added with respect to Eq. 2.31 is marked by an underbrace and defined as

ε2 for brevity. As with the 1-D PBRW without measurement error, the 1-D PBRWwE also

converges to βBRW(t). The convergence times above which the difference between βPBRWwE

and βBRW is less than 5% are shown in Fig. 2.5 (solid curves) for different values of the

constant ε. These convergence times increase with increasing centroid measurement error

and/or decreasing instantaneous speed. When ε2 is sufficiently large, the convergence time is

estimated by comparing ε2 to c2t′2 (see the denominator of Eq. 2.34), giving tσm ≈ 4.5√

2σmvrms(∞) =

4.5 ελ+c

(Fig. 2.5, dashed curves), where vrms(∞) = |c|v as derived in Eq. 2.29. The factor of

4.5 was chosen so the convergence time scale corresponds to the 5% difference threshold.

A larger factor can be chosen to lower the percent difference threshold. This convergence

time scale can be used to set a minimum fit time (c.f. τmin in Fig. 2.3 d). Fitting the

45

Page 72: The Mechanics of Cell Motility and a Unifying Theory for

βBRW-model at times above the minimum fit time significantly decouples directionality time

measurements from measurement error. Note that tσm is sampling interval independent.

As the dimensionality of the random walk increases, so does the effect of measurement

error. Specifically, 〈X2(t)〉 → 〈X2(t)〉 + 2dσ2m where d is the number of dimensions. Also

taking into the account the convergence time tBRW due to persistence over short time scales,

the generalized convergence time scale, known as the minimum fit time tmin, is the greater

of:

tmin = max

tσm ≈ 4.5

√2dσm

vrms(∞)

tBRW ≈ 10 1λ+≈ 10tp

(2.35)

both of which are readily measured and sampling interval independent. Fitting experimental

data to the βBRW-model at times above tmin decouples measurements of directionality time,

td, from the effects of persistence and measurement error.

4.3. Appendix B: Deviations Caused by Variances and Ergodicity. In Appen-

dix A, we showed that biased random walks could be described by the βBRW-model

βBRW(t; td) =1 + 2t

td

1 + ttd

(2.36)

which can be used to measure directionality time, td. This analytical fit model implicitly

accounts for the variance of position, R, caused by variance in the step time distribution,

ρT (t), and the orientation distribution, ρΘ(θ) (c.f. Appendix A, rules 2 and 3). However, it

does not account for other sources of variance associated with measurements of real data,

or variance of parameters such as speed, v, which can take on a different value for each

random walker in an ensemble. We will refer to these three categories of variances as:

implicit variance, measurement error, and parametric variance, respectively. The implicit

variance is given by, σ2R(t) = 〈R2(t)〉 − 〈R(t)〉2. The variance attributed to measurement

error is 2dσ2m, where d is the number of dimensions. Finally, parametric variance is denoted

σ2p(t). Adding the variances in quadrature and using the 1D-PBRW model to calculate

46

Page 73: The Mechanics of Cell Motility and a Unifying Theory for

implicit variance, 〈X2(t)〉 − 〈X(t)〉2, from Eqs. 2.26 and 2.28, the composite variance of

position X is

σ2X,real(t) =

v2

λ2+

(5c2 − 2) +2v2

λ+(1− c2)t+ O

(e−λ+t

)+ O

(te−λ+t

)︸ ︷︷ ︸

implicit variance,

σ2X

+

2σ2m︸︷︷︸

measurement variance

+ c2(δv)2t2︸ ︷︷ ︸parametric variance,

σ2p

(2.37)

where O(f(t)) denotes a term of order f(t). Note here that d = 1. For demonstrative

purposes, only one type of parametric variance is written in Eq. 2.37, the speed variance (at

large time scales). There can be other parametric variance terms but in general, σ2p(t) ∼

t2. Therefore, at large time scales, parametric variance is the dominant cause of position

variance.

The EASDs derived analytically in Appendix A assumed zero measurement error and

parameter variance. When measurement error and parameter variance are nonzero as is

the case with experimental measurements, EASD is increased with respect to our analytical

models

〈R2(t)〉 = 〈R2(t)〉BRW + σ2(t) (2.38)

where σ2(t) ≡ 2dσ2m+σ2

p(t) incorporates all measurement and parameter uncertainties, and

it has been assumed that each of these uncertainties is random and does not change 〈R(t)〉.

Ensemble averaged data is often noisy. As more migration trajectories are tracked,

the standard errors of ensemble averaged statistics decrease. However, ensemble averaged

statistics such as EASD are often still too noisy to fit with many experimental designs.

Statistical noise can be reduced by first time averaging the data before calculating the

47

Page 74: The Mechanics of Cell Motility and a Unifying Theory for

ensemble average. The time averaged squared displacement (TASD), for example, is given

by

R2i (τ) =

1

(tdur,i − τ)

∫ (tdur,i−τ)

0|Ri(t+ τ)−Ri(t)|2 dt (2.39)

where i is an integer path index (i.e. path 1, 2, 3,...) and tdur,i is the total duration of the

ith path. The difference between TASD and EASD is described by

R2i (τ) = ξi(τ)〈R2(t = τ)〉 (2.40)

where the factor ξi is known as ergodicity. Mean squared displacement (MSD) is calculated

by ensemble averaging TASD (averaging over all trajectories, i)

〈R2(τ)〉 = 〈ξi(τ)〉〈R2(t = τ)〉 (2.41)

For β-curves measured from MSD instead of EASD is

β(τ) =d log

[〈R2(τ)〉

]d log[τ ]

. (2.42)

Rewriting 〈R2(τ)〉 in terms ξ, 〈R2〉BRW, and σ2 using Eqs. 2.41 and 2.38 gives

β(τ) =d log

[〈ξ(τ)〉〈R2(τ)〉BRW

(1 + σ2(τ)

〈R2(τ)〉BRW

)]d log[τ ]

. (2.43)

The log-log MSD slope can thus be expanded into its final form

β(τ) = βBRW(τ) + βσ(τ) + βξ(τ) (2.44)

where βBRW(τ) is given by Eq. 2.36, βσ(τ) is given by

βσ(τ) =d log

[1 + σ2(τ)

〈R2(τ)〉BRW

]d log[τ ]

(2.45)

48

Page 75: The Mechanics of Cell Motility and a Unifying Theory for

and βξ(τ) is given by

βξ(τ) =d log [〈ξ(τ)〉]d log[τ ]

. (2.46)

The variance correction βσ(τ) becomes negligible as time increases unless σ2(τ) is of

order τ2 or greater such that σ2(τ)→∞ faster than 〈R2(τ)〉BRW →∞. At large time scales,

the position variance is dominated by parametric variance and not intrinsic variance (c.f.

Eq. 2.37). Therefore, the standard deviations of the mean distance traveled can be used to

calculate σ2(τ) in the variance correction term, βσ(t). At large time scales, the variance

correction term can be completely determined from the experimental data by using Eq. 2.38

to substitute 〈R2(τ)〉BRW for 〈R2(τ)〉 − σ2(τ) into Eq. 2.45, giving

β(τ)−d log

[〈R2(τ)〉

〈R2(τ)〉−σ2(τ)

]d log[τ ]

− d log [〈ξ(τ)〉]d log[τ ]

= βBRW(τ). (2.47)

Here, all terms that can be measured experimentally have been moved to the left of

the equality, while the fit model has been moved to the right. Eq. 2.47 is only valid

at sufficiently large time intervals when the effects of persistence on β(τ) are negligible

(i.e. βPBRW(τ) ≈ βBRW(τ), c.f. Eqs. 2.31 and 2.21), and when the effect of implicit

variance is small relative to the combined measurement and parametric variances (i.e.[〈r2(τ)〉 − 〈r(τ)〉2

]�[σ2m + σ2

p(τ)], c.f. Eq. 2.37). The minimum fitting time interval

τmin defined in Eq. 2.35 can be used to satisfy these conditions.

As a final note, calculating the ergodicity correction term βξ(τ) from experimental data

is not always helpful because its uncertainty is often relatively large and ergodicity cannot

be time averaged to reduce statistical noise. Therefore, the correction term βξ(τ) should

only be calculated if the system is believed to be nonergodic (i.e. when the parameters that

describe an experimentally measured migration change significantly over the duration of

that migration). If the data is nonergodic and βξ(τ) is unacceptably noisy when calculated

from experimental data, sometimes truncating the migration paths to equalize (and/or

shorten) their durations will make the data nearly ergodic. This was the procedure used

49

Page 76: The Mechanics of Cell Motility and a Unifying Theory for

in order to ignore the ergodicity correction term (βξ(τ) = 0) in the chemotactic neutrophil

example from the main text (Fig. 2.9).

If the data is still nonergodic after truncation, βξ(τ) can be simulated as follows. First,

make an initial measurement of the persistence time and directionality time from the trun-

cated migration paths. Based on these initial measurements, parameters tp and κ = κ(c2)

can be chosen for a corresponding 2D-PBRW simulation that will recapitulate the experi-

mental data. Second, calculate the ensemble averaged instantaneous speed (EAIS) of the

truncated migration paths over their entire duration 0 ≤ t ≤ tdur. When the data is non-

ergodic, there will typically be an overall acceleration or deceleration associated with the

EAIS. Third, run 2D-PBRW simulations with instantaneous speeds, V (t), randomly se-

lected from the experimentally measured EAIS distribution each time a reorientation event

occurs. Simulate enough trajectories so that a smooth ensemble averaged ergodicity curve as

a function of time interval, ξ(τ), can be calculated. Finally, calculate ergodicity correction

term βξ(τ) using Eq. 2.46.

50

Page 77: The Mechanics of Cell Motility and a Unifying Theory for

CHAPTER 3

Directionality Time Analysis Identifies Rigidity Sensing

Pathway Through αMβ2 Integrins

Portions of this chapter have been published, XM O’Brien, AJ Loosley, KE Oakley,

JX Tang, and JS Reichner. J Leukocyte Biol 96, 993-1004 (2014), Copyright 2014 by the

Federation of American Societies for Experimental Biology. Adaptations of text and figures

from this publication are presented here with permission from the Federation of American

Societies for Experimental Biology, and XM O’Brien. All supplementary material can be

found online.

51

Page 78: The Mechanics of Cell Motility and a Unifying Theory for

Forward. In this chapter, we return to the usage of turning angle distributions (TADs),

tortuosity, and directionality time (td) to study the role of integrins in regulating directed

neutrophil motility in 2D as a function of substrate stiffness. The fit model used to measure

directionality time in this chapter is different compared to the last chapter. Instead of the

βBRW-model derived from random walk theory (Eq. 2.9), a heuristically derived exponential

version (denoted βexp) is used instead. At the time that we analyzed the data that went

into this chapter, directionality time was only a partially developed idea for a sampling

interval independent metric slated to be a proverbial change of direction for a field that

far too often used sampling interval dependent metrics to characterize cell motility. While

the βexp-model lacks a bottom-up derivation and precise mathematical meaning of its fit

parameters, the βexp-model does generate intuitive, heuristically interpretable data that

can be compared across the data sets it is applied to. In this chapter, the directionality

time metric revealed that stiffness dependent 2D honing depended on the engagement of

the β2, CR3. To give these results more context, an extended overview on integrins has

been written into this chapter.

1. Introduction

1.1. Neutrophils. Neutrophils serve as the body’s first line of defense against invading

pathogens, transmigrating though vascular endothelium and chemotaxing through vicinal

extracellular matrix (ECM) to reach the target area of infection or injury [8, 100]. Improper

neutrophil activity has grave clinical consequences: insufficient activity results in recurrent

life-threatening infections and impaired wound-healing, whereas excessive activity leads to

an exaggerated immune response, resulting in tissue damage.

1.2. Neutrophil Mechanosensing. Mechanosensing of cells within tissues refers to

the ability of the cells to perceive differences in the mechanical properties of its environment.

52

Page 79: The Mechanics of Cell Motility and a Unifying Theory for

The mechanical environment has been shown to influence gene expression, proliferation, cy-

toskeletal organization, and survival significantly in a number of cell types [45, 101–104].

Neutrophil emigration into tissue is a multistep process, culminating with entry into notably

different tissue microenvironments. In so doing, cells interpret a number of biochemical cues

that hone their directed movement and regulate their host-defence functions within injured

or infected tissues. The physical properties of the tissue microenvironment contribute sig-

nificantly to the regulation of neutrophil migration speed and generation of traction [48, 49].

Relatively few host-defense functions take place in circulating or nonadherent neutrophils

nor does the physiological context in which they occur resemble rigid surfaces, such as glass

or plastic (Young’s modulus of ∼ 106 to 107 kPa [105]). With the use of polyacrylamide

gels of physiologically relevant stiffnesses coated with the ECM protein fibronectin (Fn),

Oakes et al. [48] found that human neutrophils migrating up a concentration gradient of

the chemoattractant Formyl-Methionyl-Leucyl-Phenylalanine (fMLP) were faster but less

directed on softer gels (Young’s modulus of 10 kPa) than they were on stiffer gels (Young’s

modulus of 100 kPa).

1.3. Integrins and Mechanosensing Mechanisms. Integrins are transmembrane

adhesion-receptors that play a significant role in mechanosensing [29]. Integrins dynami-

cally form and break bonds with the ECM as well as several intermediate proteins that link

integrins to the actin cytoskeleton [46]. Depending on factors such as integrin conformation

and distribution on the cell surface, traction forces can be applied to the ECM through

integrin mediated coupling to the actin cytoskeleton. Typically, force transduction occurs

through the formation of integrins clusters of at least 3 to 5 integrins, the minimal number

required to couple to the actin cytoskeleton [106]. Traction forces are normally contractile

due to the inward direction of actin retrograde flow. The bonds and proteins that link the

ECM to integrins, and integrins to the actin cytoskeleton, form a constitutive system that

effectively behaves like a catch bond between the ECM and the actin cytoskeleton. As the

53

Page 80: The Mechanics of Cell Motility and a Unifying Theory for

force on the adhesion complex increases, the effective catch bond strengthens. Furthermore,

other integrins diffusing through the plasma membrane aggregate near the adhesion com-

plex overtime, bind ECM, and further strengthen the overall link between ECM and the

cytoskeleton.

Measuring rigidity fundamentally equates to determining the stress required to strain

a substrate by a specific amount, or vice verse, determining the strain required to stress a

substrate by a specific amount. In some cases, it is the magnitude of applied strain that

appear to be conserved across changing substrate stiffness [48, 53, 107, 108], while in other

cases it is the magnitude of the applied stress that appears to be conserved [51, 109]. One

example of a mechanosensing mechanism is stress induced protein unfolding that exposes

new effector binding or phosphorylation sites [110, 111]. The downstream response to

increased effector binding and/or phosphorylation provides an indirect measure of the strain

response with respect to the applied stress, which therefore makes it an indirect measure of

substrate rigidity. The forces involved between single integrins and the ECM range from 1

to 30 pN, but as we will see in chapter 5, the size of contractile forces from mature adhesion

complexes are as large as 100 nN in neutrophils.

1.4. Integrin Ligand Engagement. Integrins are distinguishable by their α and β

subunits (Fig. 3.1) [26, 112, 113]. Fn is a canonical β1-integrin ligand, and α5β1 is a key

mediator of neutrophil migration on this substrate. β1 integrins have been shown to play

a role in the cellular mechanosensing apparatus in cells other than neutrophils [114, 115].

Fn, however, is also recognized by several other receptors, one of which is the β2 integrin,

αMβ2 (CR3). To examine the contribution of CR3 in transducing mechanotactic cues,

we analyzed the migration of human neutrophils up a concentration gradient of fMLP on

polyacrylamide gels, ranging in stiffness from 10 to 100 kPa, coated with Fn or the CR3-

restricted ligand, fibrinogen (Fgn), which is known to be associated with the ECM under

inflammatory conditions, forming the provisional fibrin matrix at virtually all sites of tissue

54

Page 81: The Mechanics of Cell Motility and a Unifying Theory for

αM αX

α1α2

α5β2

β1

Figure 3.1. Schematic of integrin types. Integrins are transmembrane re-ceptors that can be distinguished by their α and β subunits. The integrinsrelevant to this study (highlighted yellow) are those that bind the ligands colla-gen, fibronectin, and fibrinogen. These include α1β1 (VLA-1) and α2β1 (VLA-2) which bind collagen, α5β1 (VLA-5) which binds fibronectin, αXβ2 (CR4)which binds fibrinogin, and lastly, αMβ2 (CR3) which binds both fibronectinand fibrinogen. This figure was adapted from Ref. [113] with information fromRef. [112].

damage [116]. We also extended our study of neutrophil migration to include gels coated

with type IV Col (Col IV), an ECM ligand recognized by β1 but not β2 integrins.

In this chapter, the role of specific groups of integrins in the mechanoregulation of

neutrophil chemotaxis is determined from observations of chemotaxis on the ECM pro-

teins, Fn, Fgn, and Col IV, which bind to specific integrin subgroups. Knowing which

integrins engaged during chemotaxis, we decompose the migration trajectories into param-

eters that characterize honing, persistence, and speed to determine that the ligand-specific

mechanosensitivities for each parameter appear to be regulated independently. The roles of

specific integrins are further elucidated by performing integrin blocking experiments. Direc-

tionality time measurements with and without β2 integrin blocking show that β2 integrins

modulate honing on a 2D surface in response to substrate stiffness.

55

Page 82: The Mechanics of Cell Motility and a Unifying Theory for

2. Materials and Methods

2.1. Reagents. Anti-Fgn (85D4) and anti-Fn (FN-15) antibodies were purchased from

Sigma-Aldrich (St. Louis, MO, USA); anti-type IV Col pAb and anti-integrin β1-blocking

antibody (P5D2) were purchased from Millipore (Billerica, MA, USA); and anti-integrin β2-

blocking antibody (TS1/18) was purchased from Pierce Biotechnology (Rockford, IL, USA).

L-15 and HBSS were purchased from Invitrogen (Carlsbad, CA, USA), and polymyxin B and

sulfo-SANPAH were purchased from Pierce Biotechnology. Human Fn isolated from plasma

(>95% purity) and human type IV Col from human placenta (>95% purity) were purchased

from Sigma-Aldrich. Human Fgn isolated from plasma (>95% purity) was purchased from

Molecular Innovations (Detroit, MI, USA), and IRDye800-conjugated goat anti-mouse IgG

was obtained from Rockland Immunohistochemicals (Gilbertsville, PA, USA).

2.2. Neutrophil Isolation. Under the approval and guidelines of the Rhode Island

Hospital Institutional Review Board, neutrophils were isolated from healthy human vol-

unteers by collection into EDTA-containing Vacutainer tubes (BD Biosciences, San Jose,

CA, USA). Histopaque 1077 (Ficoll Histopaque) was used for the initial cell separation,

followed by gravity sedimentation through 3% dextran (average 400-500 kDa MW). Con-

taminating erythrocytes were removed by hypotonic lysis, yielding a neutrophil purity of

> 95%. Neutrophils were suspended in HBSS (without Ca2+ or Mg2+) on ice until use in

the experiments. All reagents contained < 0.1 pg/ml endotoxin, as determined by Limulus

amoebocyte lysate screening (BioWhittaker, Walkersville, MD, USA).

2.3. Substrate Preparation. Migration experiments were carried out in heatable

glass-bottom DeltaT dishes (Bioptechs, Butler, PA, USA) on 10 kPa, 50 kPa, or 100 kPa

gels, prepared following a method described by Pelham and Wang [101]. Briey, a solution

of acrylamide and bisacrylamide was polymerized using tetramethylethylenediamine and

ammonium persulfate. Gels were made in DeltaT dishes using AbGene frames (AbGene,

Epsom, UK) as molds. The gels were allowed to polymerize at room temperature. Once

56

Page 83: The Mechanics of Cell Motility and a Unifying Theory for

polymerized, gels were soaked overnight in PBS, allowing unpolymerized acrylamide to

diffuse out. The final size of the gels was ∼1 cm× 1 cm× 300µm.

Gel stiffness was regulated by varying the percentage of bisacrylamide in relation to the

percentage of acrylamide in the initial mixture [44], and elasticity was confirmed with an

AR2000 oscillating plate rheometer (TA Instruments, New Castle, DE, USA) [117]. The

gels were coated with Fgn, Col IV, or Fn using the chemical cross-linker sulfo-SANPAH,

which was allowed to bond covalently to the acrylamide gel for ∼1 h at room temperature.

Gels were washed three times with 50 mM HEPES (pH 7) and then incubated in 200µL of

20µg/mL Fgn, Col IV, or Fn, all diluted in 50 mM HEPES (pH 7), while exposed to UV

using a Foto/Prep I transilluminator (Fotodyne, Hartland, WI, USA), equipped with 15 W,

312 nm bulbs with an energy output of ∼ 4000µW/cm2 for 12 min at room temperature,

washing three times with PBS before use.

The elasticity of the gels has been shown to be unaffected by the protein-coating pro-

cedure, and the density of protein on the surface of the gel is unaffected by the elasticity

of the gels [43, 48]. The uniformity of the protein coating on the acrylamide substrate was

confirmed by immunofluorescence. Gels were cast on activated glass slides and coated with

Fgn, Col IV, or Fn, as described above or 90:10 and 10:90 mixtures by volume of 20µg/ml

Fgn and 20µg/ml Fn. Gels were blocked with 200 µL of 2% BSA at room temperature for

30 min before being incubated with a 100µg/ml solution of mouse anti-human Fgn, Col IV,

or Fn antibody as appropriate in 2% BSA for 1 h at room temperature. The gels were then

washed three times with PBS. Next, gels incubated with 1:10,000 IRDye800-conjugated

anti-mouse IgG in 2% BSA for 1 h at room temperature. After incubation, the gels were

washed three times with PBS and scanned using the Odyssey Infrared Imaging System

(LI-COR, Lincoln, NE, USA). Each gel was divided into four, 5 mm2 regions of interest.

Fluorescence intensity from regions of interest was obtained for six gels from two to four

independent experiments/condition.

57

Page 84: The Mechanics of Cell Motility and a Unifying Theory for

2.4. Neutrophil Chemotaxis. For chemotactic migration assays, 1-2×106 neutrophils

were resuspended at 37C in 2 mL L-15-glu. Neutrophils were allowed to settle for approxi-

mately 2 min. A femtotip (Eppendorf North America, New York, NY, USA) was filled with

5µL of a 1-mM concentration of bacterial fMLP and placed with the tip at the migration

surface within the field of view. When indicated, neutrophils were pretreated on ice for

30 min with 10 µg/ml β2-blocking antibody (TS1/18), β1-blocking antibody (P5D2), or

isotype control in L-15-glu, which was maintained for the duration of the experiment.

2.5. Microscopy. A Nikon TE-2000U inverted microscope (Nikon, Melville, NY,

USA) coupled to a CoolSNAP HQ CCD camera (Roper Scientific, Martinsried, Germany)

or an iXonEM + 897E back-illuminated Electron Multiplying CCD camera (Andor, Belfast,

UK), outfitted with a Bioptechs (Butler, PA, USA) stage heater and a 20x Nikon Plan

apochromat objective, was used for all experiments. DIC images were captured over 30 min

on a 10 s interval using Elements (Nikon). All data were analyzed using Excel (Microsoft,

Redmond, WA, USA), ImageJ (U.S. National Institutes of Health, Bethesda, MD, USA),

and MATLAB (MathWorks, Natick, MA, USA) computational software.

2.6. Cell Tracking Tools and Analytics. The neutrophils chosen for analysis ex-

hibited movement for at least 300 s. Under the conditions analyzed in this study, 25 to 30%

of cells migrated to the fMLP point source in a given 30-min observation period. Cells were

tracked frame-by-frame using custom MATLAB software [48]. Individual cell boundaries

were determined through thresholding of the DIC images with respect to a median image,

calculated by taking the median intensity of all images in a sequence. Each cell centroid

was determined through a center of mass calculation, based on the cell border. Finally, cell

migration trajectories, ~r(t), were assembled from the time-course of all centroid positions.

Mean squared displacement (MSD),⟨r2(τ)

⟩, was calculated as a function of time

interval, τ , as described in Chapter 2 (Eq. 2.2). From MSD, the log-log MSD slope,

58

Page 85: The Mechanics of Cell Motility and a Unifying Theory for

β(τ) =d[〈r2(τ)〉]

dlog[τ ] , was determined. The directionality of the cells in this study were charac-

terized fitting β(τ) to a three parameter exponential fit function,

βexp = β∞ − (β∞ − β0) exp

(−τtd

)(3.1)

where fit parameter td characterizes the time scale at which migration transitions from

random to directed, fit parameter β0 heuristically describes the randomness of motion at

τ = 0, and β∞ heuristically describes the directedness of motion at long time intervals

(τ →∞).

To further analyze the directionality transition, the time averaged quantities, β− and β+

were calculated as the time averages of βexp at time intervals τ < td and τ > td respectively.

Root mean squared speed (RMSS) was calculated by taking the square root of MSD

at time interval, τ , and dividing by τ . TADs were calculated as the distribution of angles

between sequential displacement vectors ~r(t + ∆t) − ~r(t) binned in 10◦ increments. The

sampling interval of ∆t = 60 s was used for all TAD calculations. Positive angles indicate

counter-clockwise turns, whereas negative angles indicate clockwise turns. Tortuosity was

calculated as the fraction: end-to-end displacement of the migration path divided by to-

tal migration path length. Like TAD, tortuosity depends on sampling interval ∆t and is

routinely reported for ∆t = 10 s, as we do here.

2.7. Statistics. Data were pooled from a minimum of three independent experiments

representing three to six different donors, with an n equal to 15-40 cells for each condition.

ANOVA with Newman-Keuls post hoc analysis, as appropriate, was performed using MAT-

LAB (MathWorks) or Excel (Microsoft) running the statistiXL data package (Needlands,

Western Australia). The null hypothesis was rejected if P<0.05.

2.8. Online Supplemental Material. Supplemental Fig. 1 shows that protein coat-

ing is unaffected by gel stiffness, and mixtures of Fn and Fgn are cross-linked to the gel

59

Page 86: The Mechanics of Cell Motility and a Unifying Theory for

in proportion to their percentage in the applied protein mixture. In Supplemental Fig.

2, neutrophils migrating on Col IV-coated substrates toward fMLP show no difference in

MSD, td, β−, or β+ upon blocking of β2 integrins. Supplemental Fig. 3 shows that blocking

β1 integrins do not change MSD or transition significantly from random to directed motion

in cells migrating on Fgn-coated gels toward fMLP. Video 1 shows neutrophils migrating

on Fgn-coated gels of 10 kPa and 100 kPa stiffness toward a fMLP point source; Video 2,

neutrophils migrating on Col-coated gels of 10 kPa and 100 kPa stiffness toward a fMLP

point source; Video 3, neutrophils migrating on Fn-coated gels of 10 kPa and 100 kPa stiff-

ness toward a fMLP point source; and Video 4, neutrophils migrating on Fn-coated gels

of 10 kPa+β2-integrin block and 100 kPa+β2-integrin block stiffness toward a fMLP point

source.

3. Results and Discussion

In this work, we demonstrate that the process of human neutrophil chemotaxis toward a

fMLP point source is ligand-directed and is underpinned additionally by distinct constituent

processes directing cellular speed, direction, and persistence of motion that can be mod-

ulated independently by the tissue microenvironment and integrin engagement. A novel

feature of this work is the development of biophysical tools, some introduced for the first

time here, which allow fine dissection of the mechanisms that lead to the mechanosensitive

differences in neutrophil chemotaxis.

Previous work from our lab [48] found that the mechanical properties of a Fn-coated

substrate strongly affected neutrophil morphology, with cells spreading over a larger area

and more quickly on substrates of greater stiffness. In addition, they also observed that

neutrophil function was affected on stiffer substrates, resulting in slower migration, but

with an increase in directedness, such that a greater net distance was traveled over time.

Their finding that the elastic properties of the substrate dictate the ability and efficiency of

the neutrophils to adhere and migrate, taken in the context that the mechanical properties

60

Page 87: The Mechanics of Cell Motility and a Unifying Theory for

of cells and tissues can be altered dramatically in states of disease or inflammation [118],

suggests that studies of cell adhesion and migration on flexible substrates represent a more

physiologically appropriate and biologically relevant approach than similar studies on plastic

or glass. Physiologically, these findings also suggest that neutrophil function may be subject

to the relative tonicity of the tissue in which the response is taking place.

We were interested in investigating if these findings were restricted to Fn or if they

could be extended to matrices more reflective of an inflammatory site and recognized by

other integrin families. To that end, polyacrylamide gels of 10 kPa, 50 kPa, and 100

kPa stiffness were coated with Fgn, Col IV, or Fn using the photoactivatable cross-linker,

sulfo-SANPAH. In these experiments, ligand density was held constant across conditions

to allow the isolation of substrate stiffness as an independent variable. Immunological

detection was used to confirm that the protein coating was uniform and proportional and

that the density of protein was independent of substrate stiffness (Supplemental Fig. 1

found online, see [1]). Neutrophils do not migrate through a matrix-coated polyacrylamide

gel, and therefore, they provide us an experimental system that allows for the interrogation

of cellular mechanosensing, based purely on the mechanical properties of the substrate at

constant ligand density.

3.1. Neutrophil morphology is dependent on substrate stiffness and inde-

pendent of ligand coating. Consistent with cell morphology on Fn, neutrophils adhered

to coated substrates show distinct changes in morphology that are dependent on substrate

stiffness and independent of ligand coating. A dramatic increase in spreading was observed

on 100 kPa gels compared with that on 10 kPa gels, regardless of ligand Fgn, Col IV, or

Fn at 120 s after initial cell adhesion, a time when cells typically reach their stable spread

area [48] (Fig. 3.2 a). Strikingly, the spread area of cells on 100 kPa substrates was more

than double that of cells on 10 kPa substrates (Fig. 3.2 a).

61

Page 88: The Mechanics of Cell Motility and a Unifying Theory for

10s

20s

30s

40s50s

60s

70s

b c

time interval, τ (s)

Fn Fgn1

0 k

Pa

10

0 k

Pa

Cola

g

f

1000

1000 0

100

10

10

1100

Slope = 1

Slope =

2

MS

D (

µm

)2

Linear Motion:

300

Random Motion:

time interval, τ (s)

*

turn

ing

an

gle

dis

trib

uti

on

angle (deg) angle (deg) angle (deg)

1000

1000 0

100

10

d

MS

D (

µm

)2

1200

1000

800

600

400

200

MSD at 150s

*

MS

D (

µm

2)

**

**

**

**

**

*

10

1

100 300

RM

SS

m/s

)

0

0.0

0.1

0.2

0.3

0.4

e

Col

10 100 10 100

Fn

10 100

Fgn

50 50 50

Slope = 1

Slope =

2

10s

20 30s

40 0s

60

70s

Figure 3.2. Mobility and TAD of neutrophil chemotaxis toward fMLPon surfaces of varying stiffness depend on ligand coating. (a) Humanneutrophils adhered to Fgn-, Col IV-, orFn-coated gels show distinct changes inmorphology that are dependent on substrate stiffness. Inset on each micrographis the average spread area of migrating neutrophils at 120 s after adhesion (20×bright-field magnification; bar=50µm). (b) For a migrating cell, trajectory plotswere generated at 10 s increments between cell centroids. (caption continued onpg. 64)

62

Page 89: The Mechanics of Cell Motility and a Unifying Theory for

3.2. The MSD mechanosensitivity of the neutrophil chemotaxis toward

fMLP is ligand-dependent. Unlike the ligand-independent effect in cell morphology,

the migration dynamics on these single substrates were found to have three distinct pat-

terns of behavior in response to mechanotactic cues. Neutrophils were allowed to migrate

on Fgn-, Col IV-, or Fn-coated substrates toward a fMLP point source, and migrating

neutrophils were tracked over a 30-min period using time-lapse DIC images acquired every

10 s (Supplemental Videos 1-3 found online). With the use of these images, we were able

to track individual cells, calculate their centroid, and generate cell-migration trajectories

(Fig. 3.2 b). We used these trajectories to quantify migration dynamics.

The MSD of tracked cells is a measure of the net distance that an average cell will

travel during a particular interval of time. The dependence of the MSD on the time interval

portrays the type of underlying motion: for a simple random walk, the MSD is linear with

time, whereas for directed motion along a straight line, the MSD increases quadratically

with time. Generally, MSD is proportional to τα, where exponent α characterizes the

directionality of motion. When MSD is plotted against time interval τ in log-log coordinates,

the exponent α can be estimated by the slope of the curve, β (Fig. 3.2 c, see Chapter 2

Eq. 2.4). A value of β = 1 indicates motion that appears random. A value of β = 2 indicates

purely directed motion at that given time scale (Fig. 3.2 c). At long-time intervals, where

the displacement of the migrating cell approximates a straight line from the starting position

to the fMLP point source, the slope of the MSD curve approaches two, whereas at shorter

time intervals, which reveal the contours of the natural path wiggle of the cell migration,

the slope tends toward one. Overall, β increases from one to two as the time interval τ

increases.

(Fig. 3.2 caption continued) (c) Idealized MSD plots in log-log coordinates. The slope characterizesthe type of motion. A constant value of one indicates diffusive motion, whereas a constant valueof two indicates ballistic motion. (d) MSD plotted, based on the average migration trajectories ofhuman primary neutrophils migrating on Fgn-, Col IV-, and Fn-coated gels of 10 kPa, 50 kPa, or 100kPa stiffness toward a fMLP point source. Migration was tracked over a 30 min period by time-lapseDIC images acquired every 10 s (Supplemental Videos 1-3 found online). Error bars represent thestandard error of the mean (SEM). (e) The average MSD at a time interval of 150 s and corresponding

63

Page 90: The Mechanics of Cell Motility and a Unifying Theory for

The MSDs for cells migrating on Fgn-, Col IV-, or Fn-coated substrates are plotted

in Fig. 3.2 d. The corresponding MSDs and RMS speed at a time interval of 150 s are

shown in Fig. 3.2 e. This time interval corresponds to constant speed ballistic motion

under all conditions studied and is appropriate for comparing the overall relative motion of

cells migrating toward the chemotactic source. Cells migrating on Fn- and Col IV coated

substrates have mechanosensitive changes in MSD, with cells showing greater displacements

on softer substrates. Cells migrating on Fgn-coated surfaces show the largest displacements,

which are independent of substrate stiffness.

3.3. The TAD mechanosensitivity of neutrophil chemotaxis toward fMLP is

ligand-dependent. Prior studies from this lab demonstrated mechanosensitive differences

in TAD of neutrophils migrating on 10 kPa and 100 kPa Fn-coated substrates [48]. The TAD

during cell migration is determined by measuring the angle of displacement with respect

to the previous step at a given time interval (shown in Fig. 3.2 f for a 10-s time interval).

The TAD for cells migrating as a function of ligand and substrate stiffness at τ = 60 s is

shown in Fig. 3.2 g. When analyzing TAD, a comparison of histogram height at 0◦, which

represents the percentage of turning angles that fall between ±5◦, enables a measure of

migration persistence relative to treatment conditions (Fig. 3.2 g, inset bar graphs; denoted

TAD at 0◦). Consistent with our previous data [48], cells migrating on Fn-coated gels

demonstrate mechanosensitive TAD, with cells migrating on 100 kPa substrates showing

a significantly greater persistence than cells migrating on 10 kPa substrates. In contrast,

cells migrating on Fgn-coated surfaces show strongly directed motion with no significant

root mean squared speed (denoted RMSS) for each condition. Error bars on MSD represent SEM,whereas RMS speed data are plotted in quartiles by Box and Whisker Plot. P<0.55 and P<0.05versus Fgn-coated surfaces of all stiffnesses. (F) Turning angles during cell migration are definedby the angles between subsequent displacement vectors that make up the migration trajectory. TheTAD implicitly depends on the sampling interval used to determine the displacement vectors. (g)TADs at a 60-s sampling interval of cells migrating on Fgn-coated substrates (left), Col IV-coatedsubstrates (middle), and Fn-coated substrates (right) are shown. (Insets) The percentage of turningangles between ±5o for each coating, denoted TAD at 0o. Error bars represent SEM. P<0.05, 10kPa Fn-versus 100 kPa Fn-coated surfaces. nd, No difference.

64

Page 91: The Mechanics of Cell Motility and a Unifying Theory for

dependence on substrate stiffness by this measure. In contrast, cells migrating on Col IV-

coated substrates demonstrated a TAD independent of substrate stiffness and migration

trajectories significantly less directed than all but cells migrating on the softest Fn-coated

substrate.

3.4. The tortuosity of neutrophil chemotaxis toward fMLP is ligand-dependent

and independent of substrate stiffness. Another measure of migration persistence is

tortuosity, calculated as a ratio of the net distance a cell migrates toward the chemotactic

point source to the total migration path length (Fig. 3.3 a). With the use of a 10 s sampling

interval to calculate the total migration path, tortuosity values for neutrophils migrating

on Fgn-, Col IV-, or Fn-coated gels showed no mechanosensitive variation (Fig. 3.3 b). The

value of tortuosity depended on ligand coating and varied significantly between ligands,

with cells on Fgncoated substrates migrating with the highest tortuosity, followed by cells

migrating on Fn-coated substrates, and with cells on Col IV coated substrates migrating

with the lowest tortuosity. This differs from our TAD results. We attribute these incon-

sistencies in persistence to the sampling interval dependence of TAD and tortuosity and

an incomplete representation of the data based on using a single sampling interval for the

characterization of persistence. TAD and tortuosity depend on the sampling interval, as

the turning angles and total migration path length, respectively, depend on the sampling

interval. The subjectivity of choosing a single sampling interval introduces inconsistency,

making comparisons across current persistence measurements potentially inaccurate. This

variability led us to develop a novel, sampling interval-independent metric, called direction-

ality time to quantify directionality without the subjective user input of a sampling interval.

Directionality time is measured by fitting the MSD in log-log coordinates to the exponential

function given in Eq. 3.1. The fit parameters td (directionality time), β0 and β∞, charac-

terize the migration paths based on data at all possible sampling intervals, instead of just

one. Directionality time, which corresponds to the decay time of the exponential function,

65

Page 92: The Mechanics of Cell Motility and a Unifying Theory for

10s

20s 30s

40s 50s

60s

70s

0

0.2

0.4

0.6

0.8

Col

10 100 10 100

Fn

10 100

Fgn

a

btortuosity

nd1.0

50 50 50

*

tortuosity = Displacement / Path

*

nd

nd

*

Figure 3.3. Tortuosity of neutrophil chemotaxis toward fMLP isligand-dependent but independent of substrate stiffness. (a) Tortu-osity is calculated as the ratio of total displacement (dotted gray line) to totalpath (solid black line), as shown in the schematic, and represents the overalldirectionality of the entire migration pathway. Tortuosity depends on time in-terval associated with the displacement vectors that make up the path. (b) Theaverage tortuosity for neutrophils migrating on Fgn-, Col IV-, and Fn-coatedgels toward a fMLP point source is shown as a Box and Whisker Plot, show-ing mean and quartile data. Tortuosity did not vary significantly by substratestiffness among gels of a given protein coating. tortuosity did vary significantlyby protein coating with cells on Fgn-coated gels migrating most directedly, fol-lowed by Fn-coated gels, and Col IV-coated gels migrating the least directedly.P<0.05 versus all other protein coatings.

describes the minimum observation time necessary to determine that chemotactic motion

up the fMLP gradient is directional. All things equal, a larger value for td indicates more in-

herent “wiggle” to a cell’s migration path. We further characterize the paths by calculating

parameters β− and β+, which are time averages of βexp at time intervals τ < td and τ > td,

respectively. These parameters give a measure for the amount of randomness (values near

one) of the migration path at short time scales compared with the amount of directedness

(values near two) in the migration path at large time scales (see Materials and Methods).

66

Page 93: The Mechanics of Cell Motility and a Unifying Theory for

3.5. Neutrophils migrating on Fgn-coated substrates toward fMLP show

mechanosensitive changes in td and β−. Cells migrating on Fgn-coated substrates

toward a fMLP point source show no mechanosensitive difference in MSD (Fig. 3.4 a and

Supplemental Video 1 found online) or similarly, RMS speed (Fig. 3.2 e). Values of td , β−,

and β+ were derived from the β(τ) over time-interval curves (Fig. 3.4 b). For cells migrating

on Fgn-coated gels, we have identified a mechanosensitive shift in td, with cells migrating

on 100 kPa substrates showing significantly more directional behavior than cells migrating

on 10 kPa substrates (Fig. 3.4 c). Values of β−, and β+ are represented in Fig. 3.4 d as

the directionality transition. As we increase the time interval of observation beyond td, the

essential directedness of migration becomes apparent for all conditions that we have studied,

with the common β+ of approximately 1.8. When looking at time intervals smaller than td,

we see a measure of how closely the migrating cell stays fixed on its path. This value shows

significant differences between conditions studied and may represent the interplay between

a “homing” and a “surveying” phenotype. Cells on Fgn-coated substrates have significantly

more directionality on 100 kPa substrates at time intervals shorter than td than do cells

migrating on 10 kPa substrates (Fig. 3.4 d).

3.6. Neutrophils migrating on Col-coated substrates toward fMLP show a

td independent of substrate stiffness. Cells migrating on Col IV-coated substrates

toward a fMLP point source show a mechanosensitive difference in MSD at 150 s, with

cells migrating on softer substrates covering a greater distance per time interval (Fig. 3.5 a

and Supplemental Video 2 found online). Measurements of td, β−, and β+, derived from

the β curves (Fig. 3.5 b), plotted in Fig. 3.5, c and d, show that migration directionality

is independent of substrate stiffness. Cells migrating on Col IV-coated substrates show td

values significantly larger than cells migrating on Fgn-coated substrates, indicating that

cells show more locally diffusive motion.

67

Page 94: The Mechanics of Cell Motility and a Unifying Theory for

time interval, τ (s)

1000

1000 0

100

10

101

100

MSD

(µm

)2

300

MSD at 150s

**

time interval, τ (s)

time

aver

aged

β(τ

)

log-

log

MSD

slo

pe, β

(τ)

directionality time, td (s)

a b

c d

τ < t d τ > t d

β− β+

directionality transition

Figure 3.4. Cells migrating on Fgn-coated gels toward fMLP have sim-ilar MSD and RMS speed but show a mechanosensitive change in tdand β−. Human primary neutrophils migrating on Fgn-coated gels of 10 kPa,50 kPa, or 100 kPa stiffness toward a fMLP point source were tracked over a 30-min period (Supplemental Video 1 found online). (a) The MSD of neutrophilmigration paths is plotted as a function of time between steps. (Inset) Theaverage MSD at τ = 150 s for each condition. MSD and RMS speed for cellsmigrating on Fgn-coated gels are independent of substrate stiffness. (b) Thelog-log MSD slope of each condition is plotted over time interval, τ . (c) Meanvalues of confidence intervals on measurements of directionality time, td. Cellsmigrating on Fgn-coated gels show a stiffness-dependent change in td, with cellsmigrating on 100 kPa gels transitioning to directed motion at a significantlyshorter time interval than those on 10 kPa gels. (d) Values of β− and β+, whichquantify the degree of randomness in the migration path over short time in-tervals τ < td and the degree of directedness in the migration path over longtime intervals τ > td, respectively. These data are plotted as the directionalitytransition. Cells migrating on 100 kPa gels have significantly more directed β−than cells on 10 kPa. β+ of cells on Fgn-coated gels is independent of substratestiffness. Error bars represent SEM. P<0.05, 10 kPa versus 100 kPa.

68

Page 95: The Mechanics of Cell Motility and a Unifying Theory for

1000

1000 0

100

10

101

100

MS

D (

µm

)2

300

MSD at 150s

*

directionality time, td (s)

time interval, τ (s) time interval, τ (s)

log

-lo

g M

SD

slo

pe

, β(τ

)

a b

c d

tim

e a

ve

rag

ed

β(τ

)

τ < t d τ > t d

β+

directionality transition

Figure 3.5. Neutrophils migrating on Col IV-coated substrates towardfMLP show td, β−, and β+ independent of substrate stiffness. Humanprimary neutrophils migrating on Col IV-coated gels of 10 kPa, 50 kPa, or 100kPa stiffness toward a fMLP point source were tracked over a 30-min period(Supplemental Video 2 found online). (a) The MSD of neutrophil migrationpaths is plotted as a function of time between steps. (Inset) The average MSDat a τ = 150 s for each condition. Cells migrating on 10 kPa Col IV-coated gelsshow a significant increase in RMS speed over cells on 50 kPa or 100 kPa gels.P<0.05, 10 kPa versus 50 kPa and 100 kPa. (b) The log-log MSD slope of eachcondition is plotted over time interval, τ . (c) Mean values of confidence intervalson measurements of directionality time, td. The td of cells migrating on Col IV-coated gels was independent of substrate stiffness. (d) Values of β− and β+,which quantify the degree of randomness in the migration path over short timeintervals τ < td and the degree of directedness in the migration path over longtime intervals τ > td, respectively. These data are plotted as the directionalitytransition. Cells migrating on Col IV-coated gels have no significant differencesin β− or β+. Error bars represent SEM.

3.7. Neutrophils migrating on Fn-coated substrates toward fMLP show

mechanosensitive differences in td and β−. Cells migrating on Fn-coated substrates

toward a fMLP point source show a mechanosensitive difference in MSD at 150 s, with

69

Page 96: The Mechanics of Cell Motility and a Unifying Theory for

cells migrating on stiff substrates covering the least distance per time interval (Fig. 3.6 a

and Supplemental Video 3 found online). Again, βexp was fit for each stiffness (β plotted

in Fig. 3.6 b) to calculate td (Fig. 3.6 c), and both β− and β+ (Fig. 3.6 d). Cells mi-

grating on 100 kPa Fn-coated substrates show significantly smaller td compared with those

migrating on 10 kPa substrates (Fig. 3.6 c). Cells migrating on 100 kPa substrates also

showed significantly larger β−. Taken together, cells on stiffer Fn-coated gels migrate more

directionally.

3.8. The mechanosensitive component of directionality for neutrophils mi-

grating on Fn-coated substrates toward fMLP is dependent on β2 integrins. For

cells migrating on Fn-coated surfaces that engage β1 and β2 integrins, we were able to gain

some insight into a β2-integrin-dependent modulation of chemotactic parameters under our

experimental conditions. Neutrophils were pretreated on ice for 30 min with 10 µg/ml

β2-blocking antibody (TS1/18, which binds the I-domain and blocks the function of all

β2-integrins) or isotype control (data not shown), which was maintained for the duration of

the experiment. Cells migrating on Fn-coated substrates toward a fMLP point source after

β2-integrin blocking show significant differences in MSD (Fig. 3.7 a), with cells on both

stiffnesses showing slower relative motion. Directionality measures, td (Fig. 3.7 b and Sup-

plemental Videos 3 and 4 found online) and β− (Fig. 3.7 c), demonstrate significantly less

persistent migration after β2-integrin blocking that is statistically indistinguishable from

that of untreated cells migrating on Col IV-coated substrates. This shift in directedness

also holds for our other measures of persistence, tortuosity (Fig. 3.7 d) and TAD (data

not shown). Cells migrating on Col-coated substrates toward a fMLP point source after

β2-integrin blocking show no difference in relative motion or persistence measures (Sup-

plemental Fig. 2 found online). Additionally, blocking β2 integrins of cells on Fn-coated

gels did not significantly change their spread areas, 120 s after cell adhesion (data not

shown). Cells pretreated with the β1-blocking antibody (P5D2) and allowed to migrate on

70

Page 97: The Mechanics of Cell Motility and a Unifying Theory for

1000

1000 0

100

10

101

100

MSD

(µm

)2

300

MSD at 150s

*

**

τ < t d τ > t d

directionality time, td (s)

time interval, τ (s) time interval, τ (s)

log-

log

MSD

slo

pe, β

(τ)

a b

c d

time

aver

aged

β(τ

)

β+

directionality transition

Figure 3.6. Neutrophils migrating on Fn-coated substrates towardfMLP show mechanosensitive differences in MSD, td, and β−. Hu-man primary neutrophils migrating on Fn-coated gels of 10 kPa, 50 kPa, or 100kPa stiffness toward a fMLP point source were tracked over a 30-min period(Supplemental Video 3 found online). (a) The MSD of neutrophil migrationpaths is plotted as a function of time between steps. (Inset) The average MSDat τ = 150 s for each condition. Cells migrating on 100 kPa Fn-coated gels showa significant decrease in MSD when compared with cells on 50 kPa or 10 kPagels. P<0.05 100 kPa versus 50 kPa or 10 kPa. (b) The log-log MSD slope ofeach condition is plotted over time interval, τ . (c) Mean values of confidenceintervals on measurements of directionality time, td. The td of cells migrating onFn-coated gels show a stiffness-dependent change in td, with cells migrating on100 kPa gels transitioning to directed motion at significantly shorter time scalescompared to cells on 10 kPa gels. (d) Values of β− and β+, which quantify thedegree of randomness in the migration path over short time intervals τ < td andthe degree of directedness in the migration path over long time intervals τ > td,respectively. These data are plotted as the directionality transition. Cells mi-grating on 100 kPa gels were significantly more directed than cells on 10 kPa.The β+ of cells on Fn-coated gels was independent of substrate stiffness. Errorbars represent SEM. P<0.05, 10 kPa versus 100 kPa.

71

Page 98: The Mechanics of Cell Motility and a Unifying Theory for

Fgn-coated surfaces showed no significant changes in MSD (Supplemental Fig. 3A found

online), td (Supplemental Fig. 3B), or β (Supplemental Fig. 3C). Cells pretreated with the

β1-blocking antibody (P5D2) neither adhere nor migrate on Col IV- or Fn-coated surfaces

(data not shown). These data suggest that for human neutrophils on compliant surfaces,

β1 integrins are driving adhesion and migration on Fn, whereas β2-integrin engagement is

directing a mechanosensitive enhancement of migration directionality. Neutrophils do not

regularly encounter ECM ligands in isolation. By necessity, cells are incorporating input

of varying ligand ratios and substrate compliance. The interplay of β1-integrin mechano-

tactic shifts in RMS speed and β2-integrin-dependent modulation of directionality suggests

a mechanism by which the cellular mechanosensing apparatus can be tuned to fine, incre-

mental changes in the microenvironment through which the cell is migrating.

These different patterns of substrate-dependent mechanotactic sensitivity are particu-

larly interesting in that they point to the fine interpretation of extracellular signals required

to regulate a neutrophil during chemotaxis: integrating mechanical properties of the sub-

strata, nature, and concentration of chemotactic signals, as well as the composition of the

ECM and ligand density. For example, regulation of the migratory response may be more

subject to mechanical cues under conditions of suboptimal chemokine concentrations rather

than under conditions of saturating chemokine concentrations. Alternatively, an examina-

tion of different chemotactic signals, such as IL-8 or C5a, might yield more pronounced

differences in migration between the stiffnesses of Fgn-coated substrates examined here.

Perhaps Fgn, as a nondiffusible cue of an inflamed matrix, and the chemotactic stimulus of

the fMLP gradient may coordinate to increase efficiency of neutrophil targeting to a degree

that overwhelms mechanotactic input. Mechanotactic shifts in RMS speed may prove to be

dependent on β1-integrin engagement.

Our work reported here, using the predominantly β1-integrin ligands Fn and Col IV

and the β2-restricted ligand Fgn, identifies an extraordinary level of complexity in the

72

Page 99: The Mechanics of Cell Motility and a Unifying Theory for

1000

1000 0

100

10

10

1

100

MS

D (

µm

)2

300

MSD at 150s

*

*

*

*

0

0. 2

0. 4

0. 6

0. 8

Col

10 kPa

Fn

1. 0*

FnFn-β2

Fn-β2Col

100 kPa

*

directionality time, td (s)

directionality transition

time interval, τ (s)

tort

uo

sity

a

b

c

dti

me

av

era

ge

d β

(τ)

sub

stra

te s

ti!

ne

ss (

kPa

)τ < t d τ > t d

β+

Figure 3.7. The mechanosensitive component of directionality for neu-trophils migrating on Fn-coated substrates toward fMLP is dependenton β2 integrins. Human primary neutrophils migrating on Fn-coated gels of10 kPa or 100 kPa stiffness toward a fMLP point source were treated with β2-blocking antibody and tracked over a 30-min period (Supplemental Videos 3 and4 available online). (a) The MSD of neutrophil migration paths is plotted as afunction of time between steps. (Inset) The average MSD at a 150-s time intervalfor each condition. Blocking β2 integrins significantly decreased MSD and RMSspeed on Fn-coated gels of 10 kPa and 100 kPa. (b) Log-log MSD slope, β(τ)was fit to the βtextupexp-model yielding mean values of td that represent thetime interval at which migration transitions from random to directed. Blockingβ2 integrins increased td on 100 kPa Fn-coated gels. β2-Integrin-blocked cellsmigrating on Fn-coated gels have a td equivalent to untreated cells migratingon Col IV-coated gels. (caption continued on pg. 74)

73

Page 100: The Mechanics of Cell Motility and a Unifying Theory for

characteristic measured metric Fgn Col Fn Fn+β2-blockmorphology spread area X X X X

relative motion MSD or RMSS • X X X

sampling interval dependent TAD • • X •measures of path “wiggle” tortuosity • • • •

sampling interval independent td X • X •measures of path “wiggle” β− X • X •

β+ • • • •

Table 3.1. Mechanosensitive parameters of human neutrophil chemo-taxis toward fMLP. Chemotactic parameters found to be mechanosensitivefor different coating and treatment conditions. A check mark represents param-eter that varies significantly by stiffness under the condition indicated. A bulletrepresents a parameter that does not vary significantly with stiffness.

mechanical regulation of neutrophil migration dependent on the nature of the matrix ligand

(Table 3.1). The RMS speed and directionality of neutrophil migration are both affected

by substrate stiffness on Fn. Persistence alone is affected by surface stiffness on the β2-

restricted ligand Fgn, whereas only the RMS speed alone is altered by surface stiffness

on the β1-restricted ligand Col IV. Therefore, under the conditions studied, differences

between the parameters of migration dynamics demonstrate a selectivity among integrin

ligands that allows for differential regulation of MSD and directionality in response to

mechanotactic cues. Additionally, the metrics that we derived to examine the time scale of

the transition from diffusive to directed migration reveal a previously unrecognized intricacy

in the directionality of neutrophil migration that can be mechanodirected by β2-integrin

engagement, underscoring the need for sensitive quantitative measures of chemotaxis.

(Fig. 3.7 caption continued) (c) Values of β− and β+, which quantify the degree of randomness in themigration path over short time intervals τ < td and the degree of directedness in the migration pathover long time intervals τ > td, respectively. These data are plotted as the directionality transition.Blocking β2 integrins significantly changes β− and alters the directionality transition profile to oneindistinguishable from untreated cells on Col IV-coated gels. (d) Cells migrating on Fn-coated gelsafter β2-integrin blocking have significantly lower tortuosities that are statistically equivalent tothose of cells migrating on Col-coated gels. The average tortuosity for neutrophils migrating on Fn-and Col IV-coated gels toward a fMLP point source is shown as a Box and Whisker Plot, showingmean and quartile data. Error bars represent SEM. P<0.05 untreated versus β2-integrin block onFn.

74

Page 101: The Mechanics of Cell Motility and a Unifying Theory for

Other investigations into mechanosensing focus on the lower range of stiffness, 0.3-3

kPa [49], which fits well with reported measures of endothelial cell stiffness. The 10- and

100-kPa substrates used in these studies represent the low and high end of physiological

stiffness reported in diverse tissue types [44, 119, 120]. In particular, we are interested in

discovering how mechanotactic changes in tissue, such as in formation of some tumors, com-

bine with ligand composition and chemotactic influences to progress or support a pathologic

or diseased tissue environment. This is likely a key step in mediating chronic inflammation

and fibrosis, where the inflammatory response itself changes the tissue environment. In

the context of a resolving site of inflammation, where matrix deposition is dense, the local

environment is also stiffened. It is plausible that increased stiffness may affect matrix re-

modeling, such as through increased matrix metalloproteinase release, but this hypothesis

may be confounded by the inability to dissociate the effect of matrix stiffness from ligand

density. Moreover, this may vary among tissues, where the response to injury may result

in deposition of different matrix components. Under conditions where an injury response

includes progressive tissue stiffening, such as Acute Respiratory Distress Syndrome, tumor

growth, or fibrosis, the changes in rigidity would be temporally regulated. The mechanosen-

sitive response of the extravasated neutrophil would be expected to vary according to the

time in the host response that the cell entered the site of injury. The synchronicity of

β1- and β2-integrin engagement may additionally afford a transition between surveying and

homing phenotypes in chemotaxing cells. β1-integrin engagement supports decreased cell

motility on stiffer substrates responding to the tonicity of a tissue as a marker of increased

injury or inflammation, while maintaining the local diffusivity of a surveying phenotype.

With β1- and β2-integrin engagement, we see both modulation of RMS speed and direction-

ality, with β2-integrin engagement supporting a homing phenotype of enhanced cell motility

and increasing directionality. Our data suggest that Fgn may be serving as a biologically

75

Page 102: The Mechanics of Cell Motility and a Unifying Theory for

significant CR3 ligand in this context by acting as a nondiffusible cue of the inflamed ma-

trix that aids neutrophil targeting by increasing β2-integrin engagement and by extension,

migration directionality up a fMLP gradient. These data combine to show that neutrophil

mechanosensitivity is not only ligand-dependent but also that mechanosensitive shifts in

migration dynamics are not binary. Even at this simple level, the parameters of neutrophil

response are finely tunable with environmental cues. For example, it may be that cells

entering a site of progressive stiffness, resulting from a wound, are mechanosensitive for Col

IV, whereas cells encountering Fn and/or Fgn deposition are likely to be associated with

an acute injury before scarring.

Current techniques to parameterize the trajectory of a chemotaxing cell most com-

monly pair migration speed with some measure of persistence/directionality [48, 49, 90].

The root MSD, divided by time, gives RMS speed over a particular time interval, whereas

persistence/directionality is typically characterized by measuring the TAD and/or tortuos-

ity. TAD and tortuosity implicitly depend on an arbitrarily chosen time interval, causing

such measures to skew potentially from report to report. Furthermore, a particular time

interval used to calculate TAD or tortuosity for one experiment may not be applicable to

other experiments if image-acquisition frequency is constrained. The distinction of true

randomness in the migration path from randomness caused by uncertainty in determining

the centroid positions is challenging. As our data represent here, migration dynamics can

vary significantly by ligand coating, and it follows that the optimal time interval by which to

analyze and compare migration conditions may also vary. To address this concern, we intro-

duce directionality time, a sampling-interval invariant measure of directionality motivated

by MSD fitting that incorporates trajectory characteristics for all time intervals. Direc-

tionality time characterizes the time scale at which migration transitions from random to

directed. Conceptually, it measures the time it takes for a cell that has veered off of its

directed course to reorient itself to its chemotactic path. A larger value for td suggests that

76

Page 103: The Mechanics of Cell Motility and a Unifying Theory for

there is more inherent wiggle to a cell’s migration path evocative of a surveying phenotype.

Conversely, small td suggests a more directed homing phenotype. All else equal, migration

trajectories are more persistent when td decreases. As a result of its incorporation of data

across all time intervals, td enables a global, less subjective characterization of migration

directionality that can uncover subtle shifts in migration dynamics not reflected accurately

using TAD or tortuosity.

In conclusion, to migrate effectively, neutrophils must integrate many divergent signals,

including chemotactic, mechanotactic, and substrate context, to initiate an appropriate cel-

lular response to injury or inflammation. This migration necessarily incorporates elements

of speed, direction, and persistence of motion. Neutrophils are mechanosensitive but may

acquire mechanosensitivity by different mechanisms while migrating on different matrices.

The understanding of the underlying mechanisms that regulate directed neutrophil migra-

tion dynamics, the teasing apart of the constituent contributions, and the knowledge of

how they interact can provide insight into comprehending how subtle shifts in migration

dynamics and overall phenotype caused by environmental cues lead to significant shifts

in cellular behavior and clinical outcome. Sensitive analytical methods introduced in this

work that are capable of capturing these fine changes in cellular behavior may be key to the

identification of novel nodes for clinical intervention and immune modulation, which can be

targeted, not just to cell type but also to the relevant tissue microenvironment.

77

Page 104: The Mechanics of Cell Motility and a Unifying Theory for

Part 2

Integrating Cell Kinematics with Cell

Mechanics

Page 105: The Mechanics of Cell Motility and a Unifying Theory for

CHAPTER 4

Cell Mechanics and Motility: Springs and Stick-Slip

Adhesion

Portions of this chapter have been published, AJ Loosley and JX Tang. Phys Rev E

86, 031908 (2012), Copyright 2012 by the American Physical Society. All adaptations of

text and figures from this publication have been done with permission from the American

Physical Society.

79

Page 106: The Mechanics of Cell Motility and a Unifying Theory for

Forward. Part 1 covered cell migration and mechanosensing. Cell mechanics were

discussed as possible mechanisms for the migration characteristics that were observed. In

part 2, the scope of part 1 is appended to include a direct investigation of the role of

cell mechanics in cell motility. This investigation begins here with a set of multi-element

toy models that give insight into the mechanical interactions and subsequent cell shape

dynamics and contractile forces that arise as part of the cell motility process.

1. Introduction

Cells are the building blocks of life and their migration is crucial to the biological

functions that sustain life. For example, tissue and nervous system formation depends on

the coordinated migration of pre-differentiated stem cells [121–123], whereas host immune

response depends on leukocyte migration to sites of infection and injury [8, 124]. Thus,

understanding the mechanisms of cell migration is important to the field of biology as well

as to advancing the frontier of medicine.

Many cells migrate by crawling along a particular substratum. The mechanisms that

generate cell crawling dynamics can be generally described in two steps. Step one, actin

polymerization occurs at the leading edge of a cell (lamellipodium) causing the cell to

protrude forward [6, 7]. New adhesion sites form at the leading edge during this process.

Step two, contractile forces generated within the cytoskeleton act to pull the rear of the cell

body forward in concert with graded adhesion between the cell and substratum [6]. Cell

shape may also play a role in cell crawling. The subject of how cell shape is determined

based on intra- and extracellular factors has been studied extensively both experimentally

and mathematically [21, 44, 48, 125–127]. There are also studies of the reverse problem in

the context of how cell shape affects focal adhesion site formation, traction forces, and cell

polarization [54, 128, 129], but the specific effects of cell shape on locomotion are as of yet

poorly understood.

80

Page 107: The Mechanics of Cell Motility and a Unifying Theory for

A variety of cell shape dynamics can occur depending on the type of crawling cells as well

as intra- and extracellular factors [21, 130, 131]. For example, leukocytes and fibroblasts

exhibit fairly nondeterministic ruffling- and bubbling-like shape dynamics [48, 132, 133].

Other cells, particularly fish epithelial keratocytes, exhibit shape dynamics that appear

periodic and coherent [20, 31, 125, 134]. Such dynamical periodicity and regularity over

many cell lengths of migration make the latter cell type, fish keratocytes, a prototypical

system for studying cell shape dynamics and motility [21, 135, 136].

Periodic shape dynamics observed in crawling fish keratocytes are caused by alternating

stick-slip motions localized at opposite sides of the cell’s broad trailing edge [31, 137]. In

fast moving keratocytes, ones that move roughly 0.1µm/s or faster, these sticking and

slipping cycles are often observed to be coherent but opposite in phase [135]. Hence, one

side of the trailing edge sticks while the other slips in what is known as bipedal locomotion.

Barnhart et al. recently introduced a two dimensional mechanical spring model with stick-

slip adhesion to capture the dynamics of bipedal locomotion in fish keratocytes [135]. This

model consists of four point-like elements located at regions of prevalent shape dynamics of

the cell. One element represents the cell leading edge, a region where forces responsible for

cell migration are generated by complex cytoskelatal processes such as actin polymerization

and retrograde flow [138–141]. Two elements represent opposite sides of the cell trailing

edge, regions that exhibit periodic sticking and slipping motions. These three elements are

connected by a particular spring configuration that incorporates a fourth element in the

central region of the cell. The springs represent either cytoskeletal elasticity or coupling

between the cytoskeleton and the nucleus and act to restore overall cell shape in response

to mechanical perturbation.

We build on the model by Barnhart et al. by analyzing different possible spring con-

figurations that recapitulate the shape dynamics of crawling fish keratocytes and use the

results of this analysis to determine how these dynamics are dependent on cell elasticity,

81

Page 108: The Mechanics of Cell Motility and a Unifying Theory for

size, and aspect ratio. The central element is now interpreted to be the cell nucleus and

we compare its motion to experimentally observed nucleus lateral displacements. Based on

assumptions such as symmetry about the axis of motion and confinement of the nucleus to

the central region of the cell, we determine that there are only four viable spring config-

urations, including the one studied in the previous work. We analyze the dynamics of all

four configurations and choose one deemed most mechanically representative of the real cell

that also generates realistic dynamics. Using this configuration, we identify three principal

parameters representing lamellipodial elasticity, cell length, and cell width that are signif-

icant determinants of the amplitude and period of cell shape oscillations. Varying these

principal parameters over a realistic range, we show that this simple spring model generates

shape dynamics corresponding to coherent bipedal, coherent non-bipedal, and decoherent

crawling cells.

2. Methods

2.1. Model overview. Similar to the previous work by Barnhart et al. [135], we model

the fish keratocyte in 2-D using four elastically coupled point-like elements representing dif-

ferent dynamic regions of the cell. To introduce the assumptions and physics underlying this

elastic coupling model, we begin with a demonstrative 1-D version illustrated in Fig. 4.1 a.

In this version, the front end, represented by x1, extends forward with velocity vf (dashed

line indicates cell protrusion). It is assumed that this forward propulsion is maintained by

the formation of new adhesions to the substrate. The trailing edge of the cell, at position

x2, is coupled to the front by a spring of equilibrium length L0 and stiffness K that is

representative of the cell length and elasticity of the actin cytoskeleton, respectively. The

assumption of a linearly elastic cytoskeleton is justified under physiologically normal strains

[51, 142].

82

Page 109: The Mechanics of Cell Motility and a Unifying Theory for

The trailing edge element experiences two types of drag forces, adhesion (sticking) and

viscous shear (slipping). Adhesion occurs due to stochastic binding and unbinding of ad-

hesion proteins between the cell and its substrate (c.f. Chapter 3, Sec. 1.3: integrin borne

adhesion complexes) [134, 143, 144]. The associated free energy landscape that influences

the adhesion proteins is modelled by quadratic potential wells with minima corresponding

to binding sites on the substrate [143, 144]. Equivalently, transient attachments of adhesion

proteins between the cell and its substrate can be thought of as springs (see Fig. 4.1 a

overlay). If the average spring constant for each adhesive bond is κ, then the force against

the direction of motion due to a particular adhesion bond that forms at time tbindi is ap-

proximately

Fi ≈ κx2(t− tbindi ), (4.1)

where x2 is the trailing edge speed, t is time, and index i refers to the ith adhesion bond.

Equation 4.1 is valid only between the binding time, tbindi , and some particular unbinding

time, tunbindi , when the spring detaches. Times tbindi and tunbindi are stochastic variables

with distributions that depend on the trailing edge velocity [143]. Upon summation over

all binding events, the time averaged adhesive drag force is found to scale linearly with x2

and κ at low trailing edge speed. At sufficiently high trailing edge speed, the adhesive force

vanishes because adhesion proteins do not spend enough time within the capture region of

conjugate binding sites to form bonds. At high trailing edge speed, the drag force is also

thought to scale linearly with x2 due to the hydrodynamics of low Reynolds number viscous

shear. However, the constant of proportionality is much smaller than that associated with

adhesion.

One can define two drag coefficients: α for slipping, and β for sticking, which incorpo-

rates κ. The overall stick-slip drag force as a function of trailing edge velocity is modelled

83

Page 110: The Mechanics of Cell Motility and a Unifying Theory for

by

Fd[x2] =

−βx2 , x2 < v1

v1−x2v1−v2

(βv1 − αv2)− βv1 , v1 < x2 < v2

−αx2 , v2 < x2

(4.2)

Here, sticking occurs when x2 < v1 (stick domain) due to adhesion bonds, slipping occurs

when x2 > v2 (slip domain), and some combination of sticking and slipping occurs when

v1 < x2 < v2 (transition domain). In the transition domain, the drag force is modelled by

linear interpolation (Fig. 4.1 b) though the shape of the curve in this transition region has

little effect on the resulting dynamics.

An additional consideration taking into account the time it takes the cell to switch

from sticking to slipping, and vice versa, is captured by a small inertia-like parameter, g,

the physical meaning of which is fully discussed in Ref. [135]. Including this g-factor, the

equations of motion for the one dimensional model are

x1 = vf (4.3)

gx2 − Fd [x2]−K(L− L0) = 0 (4.4)

where L = x1 − x2, and Fd[x2] is the stick-slip drag force given in Eq. 4.2.

Solutions to Eqs. 4.3 and 4.4 are limit cycles in the phase space of scaled cell-length,

L−L0K , and trailing edge velocity, x2. Fig. 4.1 c shows two such shape-cycle trajectories

plotted in this phase space. When the inertia-like term is removed (g = 0), spring force

must be balanced by drag force. In the stick domain (x2 < v1 < vf ), the trailing edge

velocity is less than the velocity of the extending leading edge. Consequently, the spring

representing cell length extends, increasing the forward force applied to the trailing edge

element. As this force increases, so too does the trailing edge velocity. When x2 increases to

be infinitesimally greater than v1, the sticking drag force is insufficient to balance against

the forward force of the spring. The trailing edge therefore accelerates instantaneously

84

Page 111: The Mechanics of Cell Motility and a Unifying Theory for

a

b

c

0 v1 vf v2

0

sca

led

ce

ll le

ng

th

velocity of trailing edge x2

βv1

αv2g = 0 nN s2/µm

g = 0.3 nN s2/µm

L-L0

K

0

0

α, slipping

sticking

transition

α<<β

β

v2v1 vf

Fd(v)

βv1

αv2

x2velocity of trailing edge

dra

g fo

rce

K,L0

κ

vf

L

x2 x1

Figure 4.1. 1-D crawling cell model and stick-slip adhesion definitions.(a) The leading edge, x1, moves forward with constant velocity, vf , representinga region where the lamellipodium extends forward. The trailing edge, x2, iselastically coupled to the leading edge by a spring of elasticity K and rest-lengthL0 representing cytoskeletal elasticity and extension, respectively. A stick-slipdrag force underneath the trailing edge is modelled by many small springs withaverage spring constant κ. (b) Drag force-velocity curve. At low trailing edgevelocity, x2 < v1 (stick domain), drag force is generated by adhesion complexesforming between the cell membrane and substrate. To good approximation, suchadhesion generated drag force scales linearly with velocity characterized by dragcoefficient, β. At high trailing edge velocity, x2 > v2 (slip domain), adhesioncomplexes no longer form. The drag force in this domain is purely viscousin nature and characterized by the relatively small linear drag coefficient, α(α � β). At intermediate velocities, v1 < x2 < v2 (transition domain), dragforce is generated by a mixture of the sticking and slipping mechanisms. Theoverall drag-velocity curve is continuous in all domains. (c) Cell length-velocityphase space trajectories with and without the inertia term, g. Data pointsare separated by a constant time step equal to one fiftieth of the limit cycleperiod (T/50). Therefore, rapid changes in velocity and cell length are notedby relatively large distances between consecutive data points.

until force balance is re-established by the slipping drag force (cyan trajectory). When the

inertia-like term is applied, e.g. g = 0.3 s2 nN/µm, force balance is not instantaneously

85

Page 112: The Mechanics of Cell Motility and a Unifying Theory for

required and abrupt acceleration does not occur. Hence, the limit cycle trajectory in phase

space appears rounded (green trajectory). In both cases, the shape of the drag force-velocity

curve in the transition domain (Eq. 4.2, v1 < x2 < v2) has negligible effect on the resulting

dynamics because the dynamical variable x2(t) remains within this domain over a duration

that is negligible compared to the limit cycle period.

This model is extended into 2-D as shown in Fig. 4.2. The trailing edge, where bipedal

locomotion occurs, is represented by two elements located at ~xl and ~xr. The drag force

in Eq. 4.2 is vectorially applied to both elements in the opposite direction of motion. The

nucleus is represented by an element located at ~xn. Drag on this element is intermediate

between sticking and slipping drags associated with the trailing edge. The front element

that drives the system forward is now represented by location vector ~xf , instead of ~x1 as

in the 1-D model. Later in this work, we replace the front element with a rod-like element

that better represents the wide extent of the protruding edge of the lamellipodium. Spring

and drag forces are combined into a set of 2-D equations of motion. The 2-D equations of

motion and a discussion about initial conditions is provided in the Supplementary Material

(Sec. refsec:sprMdlEqOfMotion). A reference diagram for this model is shown in Fig. 4.2 b

listing the spring constants, spring lengths, and drag coefficients. Cell lengths ∆yl and

∆yr are two of the dynamical variables used to characterize bipedal locomotion. They

are defined as the distance from the front element to the left and right trailing elements,

respectively, projected onto the axis of forward motion (y-axis). Fig 4.2 b also defines a

cell width, ∆x, as the distance between the trailing edge elements projected on the axis

perpendicular to forward motion (x-axis).

2.2. Simulation methods and criteria for characterizing dynamics. Solutions

to the 2-D equations of motion (Eqs. 1-4, 7, 8 in the Supplementary Material) were found

by numerical integration using the Runge-Kutta algorithm built into MATLAB R2010b

(The Mathworks, Natick, MA). A dynamical solution was considered periodic if the left

86

Page 113: The Mechanics of Cell Motility and a Unifying Theory for

vf

∆yl(t)∆yr(t)

∆x(t)

∆zlam

KN, N

KL, L

K L, L

K D, D

KD , D

vf

, Lxn

xf

xl

xr

KW, W

z

y

xx

y

z

α, β, g

α, β, g

γ

a b

Figure 4.2. Schematic of four element elastic coupling model in 2-D.(a) Side profile of the cell, ∆zlam is the lamellipodium thickness. (b) Top downreference diagram of the 2-D elastic coupling model. Elements, depicted byovals, are located at ~xf , ~xl, ~xr, and ~xn. The front element moves with constantvelocity ~vf . Spring lengths and elasticity are indicated next to each spring.Element specific drag coefficients are shown in rectangular boxes. The cell ismodelled symmetrically about the axis of forward motion, ~vf . Cell lengths ∆yland ∆yr are defined as the distance between ~xf and either ~xl or ~xr respectively,projected onto the axis of forward motion.

and right side cell length dynamics stabilized into periodic motion within 800 s. For a typi-

cal limit cycle period of 40 , this equates to 20 periods. Fourier transformation was used to

measure frequency. For solutions deemed periodic, phase differences between ∆yl and ∆yr

were calculated. Frequency was determined by locating the first harmonic of the Fourier

transform while phase was determined by the complex argument of the Fourier transform at

this harmonic. Fourier transforms were calculated using the MATLAB fast Fourier trans-

form algorithm. We also measured amplitudes of cell length modulation and nucleus lateral

displacement, which is defined as the distance of the central element, ~xn, from the axis

of forward motion. Simulation dynamics were considered bipedal if cell length oscillations

were periodic and the phase difference between ∆yl and ∆yr was between 0.45 and 0.55

periods. Dynamics were otherwise labelled as either periodic or irregular. Bipedal dynam-

ics are said to be realistic if the following three conditions are satisfied, which are based

on experimental observations of fish keratocyte dynamics discussed in the Supplementary

Material (Figs. 4.6 and 4.7):

87

Page 114: The Mechanics of Cell Motility and a Unifying Theory for

(1) amplitude of cell length modulation from 1 to 3µm;

(2) amplitude of nucleus lateral displacement from 0.3 to 1.2µm;

(3) period of limit cycle from 30 to 70 s.

Throughout this paper, simulation results are benchmarked against experimental analysis

of keratocytes discussed in the Supplementary Material.

2.3. Choice of model parameters. Parameter values for α, β, v1, and v2 were chosen

based on estimates made from the previous work [135]. A summary of the parameters used

in this model, including numerical ranges based on measurements of cell size, aspect ratio,

and other dynamical quantities, is shown in Table 4.1. The elastic modulus of a keratocyte,

E, has been measured to be between 10 and 150 nN/µm2 [142, 145, 146], and is thought

to increase from anterior to posterior. The model was analyzed over this range of E by

varying the stiffness of springs that correspond to different regions of the cell. These spring

stiffnesses were calculated using the relation,

k =ES

d, (4.5)

where S is the cross section area and d is the spring length. For example, to calculate

KD, we set S ≈ L∆zlam and d = D (see Fig. 4.2). Using ∆zlam ∼ 0.1µm [146, 147]

and L/D ∼ 0.5 − 1 based on keratocyte cell shapes measurements (see Supplementary

Material) [21, 125, 135] yields a spring constant range KD ∼ 0.5− 15nN/µm. The viability

of this model was tested using spring constants varied from 0 to 10 nN/µm. Spring lengths

were chosen in conjunction with spring constants so that simulated cell width and length

corresponded to the shapes of fish keratocyte cells observed in previous publications (see

Supplementary Material), though cell shape range need not have been restricted in this

manner.

88

Page 115: The Mechanics of Cell Motility and a Unifying Theory for

parameter meaning range units refsα slipping drag coefficient

0.15 - 0.5 nN s/µm(viscous shear)

β sticking drag coefficient20 - 100 nN s/µm

(adhesion under trailing edge)

γ nuclear drag coefficient1 - 20 nN s/µm

(adhesion under cell nucleus)

g inertia term0 - 0.8 nN s2/µm [135](sets switching time-scale

between sticking and slipping)

v1 critical sticking velocity0.08 µm/s

(upper limit of the sticking domain)

v2 critical slipping velocity1 µm/s

(lower limit of the slipping domain)

vf leading edge velocity0.2 µm/s

(stick-slip dynamics require vf > v1)

KN

spring constants 0 - 10 nN/µmKD [142, 145]

KL [135, 146]

KW

N 1 - 20

µmD spring lengths 5 - 35 [21, 125, 135]

L (determines cell shape) 10 - 30 [148, 149]

W 18 - 60

R handle bar rod length 0-30 µm

Table 4.1. List of parameters and physiologically viable values for the2D stick-slip model. Parameter ranges correspond to experimentally observedcell velocity, elasticity, etc. as determined by estimation or measurements re-ported in previous work. Each parameter range is justified by the referencesgiven here, except for rod length R, which we scale with the width of the cell’sperceived leading edge. Some spring constants are not applicable depending onwhich configuration is used. Here, spring lengths were selected to permit propercell shape.

2.4. Simulation benchmarking. Phase contrast movies of eleven motile fish kerato-

cytes, five undergoing bipedal locomotion, were analyzed to measure cell sizes, aspect ratios,

and other dynamical quantities used to benchmark simulation dynamics (Fig. 4.6). These

movies were obtained from the supplementary materials of Refs. [21, 125, 135, 150]. Movies

were converted to image sequences using Virtual Dub (Avery Lee) or MPEG Streamclip

(Squared 5) depending on file format. Custom MATLAB software was used to determine

89

Page 116: The Mechanics of Cell Motility and a Unifying Theory for

image by image cell symmetry axes and trajectories of the leading edge, trailing edge, and

nucleus centroid. We measured nucleus lateral displacement to be the distance from the

nucleus centroid to a line of best fit (Fig. 4.7). Experimental cell length oscillations were

measured as the distance between the center of the leading edge and either of the trailing

edge elements, projected onto the cell symmetry axis. There are minor discrepancies be-

tween these measurements and simulated cell length oscillations because the experimental

symmetry axis does not always correspond to the axis of forward motion. An example of

cell length oscillation measurements is shown in Movie S1.

3. Results

3.1. Viable spring configurations. There are several ways to elastically couple the

elements that make up the two dimensional model (Fig. 4.2 b), in particular by adding or

removing springs to form different spring configurations. By assuming symmetry about the

axis of forward motion and by requiring the cell to maintain a reasonable shape with width

and length comparable to observations, the number of possible configurations is constrained

to the four illustrated in Fig. 4.3 a. Config. 1 is the simplest possible configuration that can

generate bipedal locomotion whereas Configs. 2 through 4 generate bipedal locomotion with

one added element that represents the cell nucleus. The dynamics of all four configurations

are discussed at length in the Supplementary Material (Fig. 4.8).

Briefly, Config. 1 can generate dynamics that are similar to bipedal locomotion, though

the single direct coupling between trailing edge elements through the KW spring leads to

aberrant motions at the trailing edge. Specifically, slipping of one trailing edge element

extends the KW spring causing momentary aberrant slipping of the opposite trailing edge

element. There is also no possibility for Config. 1 to describe the observed lateral displace-

ment of a keratocyte nucleus. Adding a central element allows for indirect elastic coupling

between the trailing edge elements that supplements the direct KW connection. Config. 2

is like Config. 1 except a central element is added and all four elements are directly coupled

90

Page 117: The Mechanics of Cell Motility and a Unifying Theory for

to each other. This configuration can generate bipedal locomotion and realistic nucleus

lateral displacement if one interprets the central element to be the nucleus. However, such

dynamics are not robust under parameter variation compared to configurations with fewer

springs. Config. 2 works best near the KD → 0 or KW → 0 limits, i.e. Config. 3 or Config. 4,

respectively.

Config. 3 is the spring arrangement considered by Barnhart et al. They found that

stable bipedal locomotion occurs over a range of KW and g-values. During bipedal locomo-

tion, the central element, ~xn, oscillates in the lateral direction entrained to the bipedal limit

cycle. Although Config. 3 produces realistic bipedal dynamics, we have no physical inter-

pretation of a spring directly coupling the trailing edge elements. In contrast, we introduce

an alternative configuration, Config. 4, and use a spring orientation argument to suggest

that it better captures the mechanical properties of the actin cytoskeleton. Config. 4 is dif-

ferent from Config. 3 by the removal of the KW spring (KW = 0), and the addition of two

springs that couple each trailing edge element to the leading edge element (KD > 0). In this

model, the KD springs tend to orient with angles similar to the known orientation angles

of actin filaments that make up the lamellipodial actin network in keratocytes [151, 152].

Specifically, actin filaments in keratocytes under physiological conditions show long range

orientation order with angles between ±25o and ±45o with respect to the direction of lead-

ing edge protrusion. The KD springs capture the anisotropy of network elasticity [153] in

the direction parallel to filament orientation. Cell width is now maintained by both the KD

and KL springs, instead of spring KW as in Config. 3. Springs, KL, coupling the trailing

edge to the nucleus can be interpreted in the context of the contractile actin-myosin bundle

at the rear of the cell [154], though a KW spring can also be interpreted in the same way.

Configuration 4 generates dynamics similar to those of Config. 3, in some cases with

a slightly larger nucleus lateral displacement closer to experimentally observed values. An

example time lapse showing the dynamics of Config. 4 is shown in Fig. 4.3 b. Cell length

91

Page 118: The Mechanics of Cell Motility and a Unifying Theory for

po

siti

on

m)

position (μm) position (μm) position (μm)

po

siti

on

m)

0

5

10

15

20

25 26 s 33 s 40 s

0

5

10

15

20

25

-10 -5 0 5 10

47 s

-10 -5 0 5 10

54 s

-10 -5 0 5 10

61 s

dc

b

a

-1

0

1

time (seconds)100806040200n

ucl

eu

s la

tera

l

dis

pla

cem

en

t (μ

m)

time (seconds)0 20 40 60 80 100

1012141618

cell

len

gth

, ∆y

m)

xn

xxl r

xf

xf

xl xr l r

xn

x xl r

xf

∆y

l r

xn

xrlx

xf

Con!guration 1 Con!guration 2

Con!guration 3 Con!guration 4

Figure 4.3. Viable 2-D spring configurations and the dynamics of Con-fig. 4. (a) Diagrams of the four viable 2-D spring configurations. Each viableconfiguration maintains reasonable cell shape and is symmetric about the axisof forward motion. Config. 4 is the preferred configuration proposed in thiswork. (b) Time lapse of the simulated dynamics of Config. 4. showing bipedallocomotion. The time lapse is shown in 7 s increments and corresponds to timeplots of cell length and nucleus lateral displacement shown in (c) and (d). Thissimulation corresponds to a cell with a time averaged length and width of 13and 19µm, respectively. The amplitude of length oscillations is 3.0µm with aperiod of 40.5 s. The amplitude of nucleus lateral displacement is 0.6µm. A con-tinuous motion time lapse of this simulation is found in Movie S2. Parameterscorresponding to this simulation are listed in Fig. 4.8.

92

Page 119: The Mechanics of Cell Motility and a Unifying Theory for

and nucleus lateral displacement time plots corresponding to the time lapse are shown in

Fig. 4.3 c and d, respectively. In this example, the nucleus undulates laterally in a series

of exponential decays that have a period of 40.5 s, and an amplitude of 0.6µm, consistent

with observations. Nucleus lateral displacement can be made more sinusoidal if the drag

force (parameter γ) is reduced compared to the spring forces acting on the nucleus, though

it is unclear if this would be more realistic.

To assess the viability of Config. 4, we investigated how its bipedal dynamics changed

in response to varying mechanical parameters, KD, KL, and KN , drag parameters, α, β,

γ, and g, and cell shape parameters, D, L, and N . Fig. 4.4, a and b, are phase diagrams

of the dynamical responses plotted in the g-KD and g-KL spaces, respectively. Hatched

areas indicate regions of realistic bipedal dynamics for two choices of drag coefficient γ.

The dynamical response is characterized by amplitudes of cell length and nucleus lateral

displacement oscillations (blue curves), and the overall limit cycle period (green curve)

shown in Fig. 4.4, c and d. Bipedal locomotion occurred for spring stiffnesses, KD &

0.4 nN/µm and KL & 0.5 nN/µm. The model fails at lower spring stiffnesses because there

is not enough rigidity between elements to maintain normal cell shape. In the case of lower

elasticity, KL < 0.5 nN/µm, element ~xl can unrealistically swing over from the left side to

the right side of the cell (and vice versa for element ~xr). In the normal cell shape regime,

response characteristics changed very little under variation of KL, in contrast to variation of

KD. Therefore, we identify lamellipodial spring stiffness, KD, as a principal parameter that

tunes the length- and time-scales of limit cycle behaviour, more so than other mechanical

parameters in this model. Fig. 4.9 in the Supplementary Material shows that response

characteristics are also sensitive to variation of sticking coefficient, β, though less sensitive

to variation of drag coefficient, γ, and inertia-like parameter, g.

Not shown in Fig. 4.4 or in the Supplementary Material is the effect of spring elasticity

KN on the dynamics of Config. 4. Spring KN in conjunction with drag on the nucleus

93

Page 120: The Mechanics of Cell Motility and a Unifying Theory for

can be used to fix the average displacement between the front element and the nucleus

element. Spring elasticity KN is required to be greater than 1 nN/µm in order to maintain

cell shape. As KN increases, the nucleus element is drawn toward the front of the cell,

thus also drawing the trailing edge elements inwards. Bipedal locomotion still occurs with

slightly altered period and amplitudes.

0 1 2 3 40

1

2

3

4

0

15

30

45

60

75

KL (nN/µm)

pe

riod

(s)

NLD

cell length

period

0 1 2 3 40

1

2

3

4

0

15

30

45

60

75

am

plit

ud

e (µ

m)

KD (nN/µm)

cell lengthperiodNLD

g (

nN

s2/µ

m)

KD (nN/µm)0 1 2 3 4

0

0.1

0.2

0.3

0.4

KL (nN/µm)0

0

0.1

0.2

0.3

0.4

1 2 3 4

amplitude of celllength oscillations, both ∆yl(t) and ∆yr(t)

amplitude of nucleus lateral displacement(NLD), ∆xn(t)

period of oscilations

realistic bipedal, γ = 5 nN s/µmbipedal

non bipedal realistic bipedal, γ = 2 nN s/µm

a

c

b

d

vf

∆xn(t)

∆yl(t) ∆yr(t)KD KDKN

KLKL

Figure 4.4. Characterized dynamical responses of Config. 4 with re-spect to parameters KD, KL and g. The green shaded regions of the g−KD

(a) and g−KL (b) parameter spaces indicate bipedal dynamics and hatch pat-terns indicate realistic bipedal dynamics for two choices of γ. (c, d) Amplitudesof cell length and nucleus lateral displacement (NLD) oscillations (blue curves),as well as the limit cycle period (green curve), plotted against mechanical pa-rameters KD and KL. Lamellipodium spring elasticity, KD, significantly alterslimit cycle amplitudes and periods, whereas trailing edge spring elasticity, KL,does not. The model fails when either spring constant is too low (. 0.5 nN/µm)where then point-like elements delocalize leading to a loss of normal shape. Ineach chart, parameters that are not varied are listed in Fig. 4.8 under Config. 4.

94

Page 121: The Mechanics of Cell Motility and a Unifying Theory for

3.2. Crawling dynamics depend on cell aspect ratio. Using Config. 4, we system-

atically varied lamellipodial elasticity, cell size, and cell aspect ratio to analyze their effects

on crawling dynamics. Cell width was varied by changing spring length L (width ≈ 2L), and

lamellipodial elasticity was varied via the KD parameter, which is the principle mechanical

parameter that tunes the dynamics of this model. Two different cell lengths were studied

based on the experimental cell length distribution in Fig. 4.6 a: short, 〈∆y(t)〉 = 11µm

(N = 8µm), and long, 〈∆y(t)〉 = 16µm (N = 12µm). Mathematically, diagonal spring

length D was made functionally dependent on L, N , and KD in order to hold the cell length

constant under variation of the dependent parameters (Fig. 4.2).

The results of this analysis are shown as cases 1 and 2 in Fig. 4.5. Case 3 is a modification

of Config. 4 where a rod-like element is used at the leading edge instead of point-like element,

~xf . In the sense that the KD springs represent a center of mean elasticity on the two sides of

the lamellipodial actin network, it is likely more realistic that these springs should couple to

two points at the leading edge that are symmetrically displaced from the axis of symmetry,

instead of to a point at ~xf . A rod-like element allows us to modify the endpoints of springs

KD in just this way. The rod also better aligns the KD springs with the known long range

angular orientation of the cytoskeleton discussed above. The modified equations of motion

for this configuration, which we call the handlebar model, are found in the Supplementary

Material.

Diagrams of the three cases discussed above are shown in Fig. 4.5 a. For each case, we

investigated the effects of varying cell width and lamellipodium elasticity. Dynamical output

is characterized by amplitude of cell length oscillations and dynamical behaviour as shown

in Fig. 4.5 b and c, respectively. The amplitude maps can be broken down into three key

regions. Regions of red indicate relatively large amplitudes (3-10µm). These amplitudes are

greater than those of most coherent keratocytes and occur when lamellipodium elasticity

is low, causing the trailing edge to stick longer. Regions of dark blue indicate relatively

95

Page 122: The Mechanics of Cell Motility and a Unifying Theory for

small amplitudes (0-1µm). These amplitude describe most smooth gliding keratocytes

that have small stick-slip events observed at the trailing edge (for example, in Movie S1).

Regions of both light blue and yellow indicate amplitudes of realistic bipedal locomotion (1-

3µm). In cases 1 and 2, there appear to be “anomalous” amplitude variations for wide cells

when KD & 6 nN/µm. The phase diagrams indicate that these anomalous regions of the

amplitude maps correspond to irregular behaviour (Fig. 4.5 c). Such regions could represent

the phase space for fast moving decoherent cells. Smooth gliding cells are described by all

dynamical behaviours with small amplitudes. Exceptions shown here are cells that fail

to maintain proper cell shape, which are indicated in white at the top-right corner of the

phase diagram for case 1, although there are other examples beyond these parameter ranges.

Realistic bipedal locomotion occurs in regions of overlap between those labelled bipedal on

the phase diagrams and those where amplitudes of cell length oscillations fall between 1

and 3µm. Overall, one can use the phase diagrams in conjunction with the amplitude maps

to characterize the dynamical responses of the model. These diagrams describe how cell

crawling dynamics are dependent on cytoskeletal elasticity, extension, and cell aspect ratio.

4. Discussion

Cell motility models typically consist of a set of dynamical equations that describe the

biochemistry (i.e., diffusion and flow of biomolecules that regulate myosin motors, actin

polymerization, etc.) and/or the biomechanics (i.e., adhesion between cell and substrate,

cortical tension, etc.) of a system to varying degrees of complexity [126, 127, 135, 140, 155–

158]. That a simple mechanical model involving only four elements coupled by passive

springs is able to significantly recapitulate the motion of these highly complicated systems

is surprising. There are alternative models describing the shape dynamics of keratocytes,

such as one proposed by Ziebert et al. [158], in which cell length oscillations result from

filament orientation and overall cell shape. In another model proposed by Barnhart et al.

[127], the implications of substrate adhesion strength on keratocyte motility were studied

96

Page 123: The Mechanics of Cell Motility and a Unifying Theory for

Amplitude of Cell Length Oscillations

L (µm)10 15 20 25 30

10

8

6

4

2

periodic

bipedal

10 15 20 25 30L (µm)

10

8

6

4

2

irregular

bipedal

periodic

10 15 20 25 30L (µm)

2

4

6

8

10

KD (µ

m)

irregular

bipedal

periodic

2

4

6

8

10

3025201510L (µm)L (µm)

2

4

6

8

10

3025201510L (µm)

2

4

6

8

10

3025201510

KD (µ

m)

0

2

4

6

8

10

(µm)

Phase Diagrams

<∆y>=16µm

Case 3

<∆y>=16µm

Case 2

<∆y>=11µm

Case 1

∆yr(t)∆yl(t)

b

a

c

Figure 4.5. Dynamical responses of Config. 4 and the handlebar modelwith respect to lamellipodial elasticity, cell size, and cell aspect ratio.(a) Diagrams indicating average cell length, 〈∆y〉, and mechanical model: eitherConfig. 4 (first and second columns) or the handlebar model (third column).Cell width is varied by changing spring length L. Cell length is varied bychanging middle spring length, N , and then choosing lamellipodial spring lengthD = D(L,KD, N) such that 〈∆y(t)〉 remains constant. Parameter KD sets thelamellipodial elasticity. In case 3, the length of the handle bar is set equal toL. (b) Color maps indicating the amplitude of cell length oscillations undervariation of KD and L. Saturated red (larger amplitudes occurring when KD

is relatively small) indicates amplitudes greater than those typically observed.Dark blue (smaller amplitudes occurring when KD is relatively large) indicatesamplitudes that are small and difficult to measure experimentally. Light blueand yellow (light gray shades) indicate amplitudes corresponding to realisticbipedal dynamics. (caption continued on pg. 98)

by considering the interplay between actin polymerization, myosin II transport, myosin II

generated actin retrograde flow, and linear adhesion forces between the cell and its substrate.

Their model generates the characteristic cell shapes and migration speeds recorded in fish

97

Page 124: The Mechanics of Cell Motility and a Unifying Theory for

keratocytes, but it does not account for bipedal locomotion. Our model describes general

cell crawling dynamics and its dependency on mechanical properties but does not account

for the specific effects of adhesion strength because the physics of adhesion at the leading

edge and other more complicated factors such as actin polymerization and actin retrograde

motion are all contained in the self-propulsion parameter, ~vf , which prescribes the cell’s

locomotion speed. We ignore the finer details of these factors in exchange for a simple

model to study cell crawling dynamics with a focus distinct from that of Ref. [127].

In our analysis of cell crawling movies from previous publications, it is apparent that

the leading edge velocity of bipedally crawling cells fluctuates, although these fluctuations

are not as large as stick-slip induced velocity fluctuations at the trailing edge. As such,

assuming a constant velocity, ~vf , at the leading edge is an oversimplification. Instead, one

could model the forward propulsion by introducing a propulsion force in place of velocity

~vf . Such a force would have to be anchored by adhesion formation under the ventral surface

of the cell that would lead to a net forward spring force large enough to propel the trailing

edge elements into the slipping domain without allowing the cell to stall. The force at

the leading edge could be calculated from empirical force-velocity relations such as the one

recently measured by Heinemann et al. [147] using slow crawling keratocytes. In this case,

the load force would be assumed to scale with the total spring force acting against leading

edge and the force-velocity relation would then be used to calculate the leading edge velocity.

At this time, however, it is unknown how the force-velocity curves measured by Heinemann

et al. are different from those of fast crawling keratocytes, which have distinctly different

leading edge characteristics. Measurements and models of force-velocity curves for other

systems such as listeria [140, 159] are strikingly different from those in Ref. [147]. In listeria,

(Fig. 4.5 caption continued) Amplitude maps can be interpreted in conjunction with the corre-sponding phase diagrams beneath in (c). The dynamical response of the system is categorized intothree behaviours: bipedal, periodic, and irregular. Bipedal regions includes both realistic and otherbipedal locomotion. Both periodic and bipedal regions correspond to coherent gliding-like kerato-cytes, whereas irregular dynamics with anomalously large amplitudes may correspond to decoherentcells. The white region (case 1) indicates solutions where the cell fails to maintain a reasonableshape.

98

Page 125: The Mechanics of Cell Motility and a Unifying Theory for

the protrusion velocity is nearly independent of load under high loading conditions, whereas

the opposite is true for slow crawling keratocytes. Given the variability of force-velocity

relationships in the literature, we stay with the constant velocity approximation as a first

step to mechanically modelling cell dynamics, instead of invoking a protrusion velocity that

requires knowledge of a force velocity curve.

Beyond investigating the stick-slip dynamics of the keratocyte trailing edge, our analy-

sis also probes the role of the cell nucleus. It is well known that epithelial keratocytes are

complex systems with mechanical properties that depend on the actin cytoskeleton [160–

162], but here we find that coupling to a central element is required to generate realistic

nucleus lateral displacement. This implies a mechanical landscape where the trailing edge

and lamellipodium are both elastically coupled to the cell nucleus. The coupling scheme

made up of all possible spring connections among the four elements (Config. 2) can generate

bipedal locomotion; however, the system is not robust under parameter variation. Because

this configuration is made up of more than the minimal number of springs required to main-

tain proper cell shape, the trailing edge elements are more sensitive to sudden motions that

propagate through multiple spring pathways during stick-slip transitions. Hence, adding

additional springs over-constrains the system.

Both the third and fourth configurations generate realistic bipedal dynamics including

nucleus lateral displacement oscillations. In the fourth configuration, the KD springs cap-

ture the elasticity of the lamellipodial actin cytoskeleton in the direction parallel to overall

filament orientation. There is a general robustness in terms of parameter ranges over which

realistic bipedal locomotion occurs, though some parameters (KD and β) significantly mod-

ify the corresponding time-scales and amplitudes. Cell shape dynamics are highly dependent

on spring constant KD, whereas they are far less dependent on spring constant KL. This

implies that locomotion is sensitive to elasticity of the actin cytoskeleton, but not elasticity

of couplings between the trailing edge and the nucleus. Blebbistatin, a myosin II inhibitor,

99

Page 126: The Mechanics of Cell Motility and a Unifying Theory for

is known to inhibit actin network flow at the rear of keratocytes [154] but does not greatly

change keratocyte stick-slip dynamics [135]. This result is consistent with our model, which

predicts that cell crawling dynamics are relatively insensitive to variation of elasticity at

the rear of the cell. By the same logic, the model does not contradict findings suggesting

that the rear of a keratocyte is stiffer than its lamellipodium [146].

When interpreting Fig. 4.5, one should consider how KD scales with cell size. Holding

the elastic modulus, E, constant, one can estimate KD as a function of cell width parameter,

L, using Eq. 4.5. This function is obtained by solving KD ≈ EL∆zlamD(L,KD) , where D(L,KD) is

the spring length necessary to maintain constant cell length when L and KD are varied.

Solutions to this equation show that many limit cycles with abnormally large amplitudes in

Fig. 4.5, cases 1 and 2, are not possible given realistic values of E and ∆zlam. This mechanics

argument gives insight into why keratocytes are not observed with these excessively large

stick-slip cycles.

In addition to recapitulating the dynamics of fast moving keratocytes, the model is also

applicable to slow moving ones. The transition between slow and fast is set by parame-

ter v1, where in this work we have analyzed cells modelled by vf > v1. Figure 4.6 shows

distributions representing the wide variety of cell speeds, sizes, and stick slip amplitudes

recorded in fast moving keratocyte. Observed crawling dynamics can are categorized into

three groups: Coherent bipedal, coherent non-bipedal, and decoherent locomotion. The

model presented in this work reproduces these observations. We have shown how lamel-

lipodial elasticity, cell size, and cell aspect ratio can determine crawling behaviour even

before consideration for more complicated biological mechanisms. These findings suggest

the existence of mechanically preferred cell shapes for cells that need to move quickly and

efficiently. The mechanical model presented in this work should be applicable to other

fan shaped cells such as gliding human fibrosarcoma cells [132] and the ameboid sperm of

ascaris [163]. More complicated cell shapes and shape dynamics are possible by adding

100

Page 127: The Mechanics of Cell Motility and a Unifying Theory for

more stick-slip elements to the model. Therefore, this model may also be applicable to cells

such as leukocytes and fibroblasts that undergo more complicated, highly variable, shape

dynamics.

5. Supplementary Material

5.1. Analysis of fish keratocyte movies. Movies of crawling fish keratocytes from

previous publications [21, 125, 135] were analyzed as discussed in the methods section to

determine realistic ranges of dynamical variables that could be used to test and benchmark

our model. Cell aspect ratio, speed, and oscillation amplitude distributions derived from

the analysis are plotted in Fig. 4.6. Additionally, each keratocyte was categorized into

one of two main groups: decoherent and coherent. Decoherent cells typically crawl slowly

(vf < 0.1µm/s) with a blebbing or irregular leading edge. Coherent cells often crawl

faster than decoherent cells and maintain relatively smooth leading edge. Coherent cells

were further categorized as either bipedal or non bipedal. Bipedal cells exhibited stick-

slip events that alternated on both sides of the trailing edge, whereas these events do

not alternate in nonbipedal cells. The majority of cells were 20 to 50µm wide and 10

to 20µm long, crawled with cell length oscillation amplitudes from 0 to 4µm, and with

speeds from 0.1 to 0.4µm/s. Movie S1 shows an example of the data analysis we did on

a coherent non-bipedal keratocyte movie published in Ref. [21]. The cell in this movie

changes its direction of motion to the right (clockwise), causing the cell length to be greater

on the left than on the right (Movie S1 B). This cell corresponds to the data point at

[width,∆yamp] = [40.4± 2.6µm, 0.8± 0.8 µm] in Fig. 4.6 b. We used the length-width

distribution to determine a realistic range of cell aspect ratios that were tested with our

spring model to answer the question: How is cell crawling dependent on cell aspect ratio?

5.2. Shape dynamics of fast crawling keratocytes. A time lapse of a bipedally

crawling fish keratocyte migrating with a leading edge velocity of 0.17µm/s is shown in

101

Page 128: The Mechanics of Cell Motility and a Unifying Theory for

1 2 3 40.1

0.2

0.3

0.4

20 40 600.1

0.2

0.3

0.4

5030

1 2 3 40

1

2

3

4

20 40 600

1

2

3

4

30 50

10

15

20

25

30

20 40 605030

coherent non-bipedaldecoherent

values used for model

coherent bipedal

(a)

(b)

(d) (e)

(c)

aver

age

cell

leng

th,

<∆y>

(µm

)le

adin

g ed

ge

spee

d, v

f (µm

/s)

ampl

itude

of c

ell l

engt

h os

cilla

tions

, ∆y

amp (

µm)

cell width (µm) [cell width] / <∆y>

vf

width

∆yl(t) ∆yr(t)

Figure 4.6. Cell aspect ratio, speed, and amplitude of cell length oscil-lation distributions measured from previously published movies of fishkeratocytes. Cells are categorized into two groups, coherent and decoherent(dark green squares) as described in this work. Coherent cells are subcatego-rized as either bipedal (light green circles) or non-bipedal (blue triangles). (a)Cell length versus width distribution. Data points and error bars represent av-erages and standard deviations of measurements made at five randomly chosenframes. Most cells were 20 to 50 µm wide and 10 to 20µm long. Our modellingwork covers two cell length, 11 and 16µm, in addition to this entire range of cellwidths. (b and c) Cell length oscillation amplitude is plotted against cell widthand the width / 〈∆y〉 ratio. Horizontal error bars represent standard deviationssimilar to those in (a). (caption continued on pg. 103)

Fig. 4.7 a. Periodic retractions on opposite sides of the trailing edge are manifestations of

cycling between sticking and slipping. Here, these periodic retractions are opposite in phase

102

Page 129: The Mechanics of Cell Motility and a Unifying Theory for

indicative of a cell undergoing bipedal locomotion. Bipedal sticking and slipping on both

sides of the trailing edge occur every 50 to 65 s. Cell outlines drawn in 25 s increments depict

this behaviour, with black arrows indicating regions of distinguishable sticking occurring

over successive time frames.

The alternating stick and slip events at the trailing edge lead to nucleus undulation

perpendicular to the direction of migration and phase-locked to stick-slip cycle. Red dots

in Fig. 4.7 b indicate locations of the nucleus centroid corresponding to the time points

shown. The trajectory of the nucleus is fit to a third order polynomial (Fig. 4.7 c). Here,

the amplitude of nucleus lateral displacement ranges from 0.3 to 1.2µm with an average

of 0.7µm (Fig. 4.7 d). The period of nucleus lateral displacement ranges from 50 to 65 s

(Fig. 4.7 e), same as the stick-slip period. Of the bipedally crawling keratocytes we analyzed,

nucleus lateral displacement amplitudes varied greatly and it was often unclear whether

lateral undulations were closer to a periodic exponential relaxation, or a sinusoidal-like

curve as plotted in Fig.4.7 d. The average amplitude of nucleus lateral displacement was

0.98µm with a standard deviation of 0.60µm. This data is compared against the results of

our modelling work.

5.3. The two dimensional model. In two dimensions, we analyze the shape dynam-

ics for a variety of spring configurations that describe the general model depicted in Fig. 4.2.

In the notation of this paper, ~xi is the position of the ith element where index i denotes

one of the following abbreviations: f for the leading edge element, n for the nuclear region

element, l for the left trailing edge element, and r for the right.

(Fig. 4.6 caption continued) Vertical error bars are estimated based on the standard deviation in celllength change caused by individual stick-stip events, and based on the average cell length uncertainty,measured frame by frame. The amplitudes of coherent bipedal cells are largest, whereas the othercells have small amplitudes with large error bars. (d and e) Leading edge speed versus cell widthand width / 〈∆y〉 ratio.

103

Page 130: The Mechanics of Cell Motility and a Unifying Theory for

10μm

1007550250

1007550250

0 10 20 30 4005

10152025

x (µm)

y (µ

m)

(c) (d) (e)

(b)

(a)

0 100 200-3

0

3

time (seconds)nucl

ear l

ater

aldi

spla

cem

ent (µm

)

0 0.02 0.0400.20.40.60.8

frequency (Hz)

Four

ier a

mpl

itude

Figure 4.7. Bipedal locomotion and lateral oscillations of the nucleusobserved in fish keratocytes. (a) Time-lapse of a fish keratocyte moving fromtop-left to bottom-right, shown in 25 s increments. Cell outlines in fading greyrepresent the same cell 25 and 50 s prior to current position. Periodic sticking(arrows) and slipping alternate across opposite sides of the trailing edge. (b)Time lapse of the same cell in A, now showing the trajectory of the nucleus aswell as cell outlines for every time increment. (c) Nucleus position of the cellanalyzed in A (red) fit to a third order polynomial (grey). (d) Lateral motionof the nucleus calculated as the distance between nucleus position and fit in thedirection normal to the fit. Lateral motion is plotted with respect to time. (e)Fourier transform of (d) shows a frequency corresponding to a period between50 to 65 s.

5.3.1. Equations of Motion. The general equations of motion for the two dimensional

model are

~xf = vf y (4.6)

γ~xn −∑i=f,l,r

Kin(|~xn − ~xi| − Lin,0)~xn − ~xi|~xn − ~xi|

= 0 (4.7)

g~xl − Fd[|~xl|]~xl

|~xl|−∑

i=f,n,r

Kil(|~xl − ~xi| − Lil,0)~xl − ~xi|~xl − ~xi|

= 0 (4.8)

g~xr − Fd[|~xr|]~xr

|~xr|−∑i=f,n,l

Kir(|~xr − ~xi| − Lir,0)~xr − ~xi|~xr − ~xi|

= 0 (4.9)

104

Page 131: The Mechanics of Cell Motility and a Unifying Theory for

where γ is the drag coefficient of the central element, and matrix elements Kij and Lij,0

are the spring elasticity and equilibrium length, respectively, for the spring connecting the

ith and jth elements. Assuming symmetry about the axis of forward motion, the spring

elasticity and equilibrium length matrix elements are given by

Kij =

f n l r

0 KN KD KD

KN 0 KL KL

KD KL 0 KW

KD KL KW 0

ij

(4.10)

and

Lij,0 =

f n l r

0 N D D

N 0 L L

D L 0 W

D L W 0

ij

(4.11)

Different spring configurations are analyzed by varying elements of the spring matrix and

a range of cell shapes can be analyzed by varying both Kij and Lij . A spring is removed

by setting its elasticity coefficient to zero. A movie of one spring configuration in which

KW = 0 is shown in Movie S2. This particular configuration is discussed in detail in the

main article and is referred to as Config. 4.

5.3.2. Handlebar model. In the standard 2-D model above, the diagonal springs (KD)

couple both trailing edge elements to a point-like leading edge element along the symmetry

axis of the cell. In the sense that the KD springs represent a center of mean elasticity of

the actin cytoskeleton in the lamellipodium, it is likely more realistic that the KD springs

105

Page 132: The Mechanics of Cell Motility and a Unifying Theory for

should couple to two points at the leading edge symmetrically displaced from the axis of

symmetry. We therefore introduce a handlebar model in which a rod perpendicular to the

axis of symmetry represents the cell leading edge instead of point-like element ~xf . The KD

springs couple to the ends of the rod and the KN spring couples to center, (see Fig. 4.5 a,

right diagram). Equations of motion 4.6 and 4.7 remain unchanged while Eqs. 4.8 and 4.9

become

g~xl − Fd[|~xl|]~xl

|~xl|−

∑i=f+,n,r

Kil(|~xl − ~xi| − Lil,0)~xl − ~xi|~xl − ~xi|

= 0 (4.12)

g~xr − Fd[|~xr|]~xr

|~xr|−

∑i=f−,n,l

Kir(|~xr − ~xi| − Lir,0)~xr − ~xi|~xr − ~xi|

= 0, (4.13)

where ~xf± = ~xf ± R2 x. This model is discussed in association with Fig. 4.5 of Sec. 3.2.

5.3.3. Initial Conditions. In general, the solutions of a dynamical system depend on

initial conditions. In the context of this work, choosing initial conditions is akin to deter-

mining the initial cell shape and regional velocities. Depending on the choice of parameters,

many solutions converge to bipedal limit cycles. To speed up this convergence, we chose

initial conditions that were asymmetric about the axis of symmetry. For instance, when

solving the equations of motion for Config. 4, the following initial conditions were applied:

[xf , yf ] |t=0 = [0, 0]

[xn, yn] |t=0 = [0,−N ]

[xl, yl] |t=0 = [−L,−(N + 1)]

[xr, yr] |t=0 = [+L,−(N − 1)]

[xl, yl] |t=0 = [xr, yr] |t=0

(4.14)

In the special case where fully symmetric initial conditions are applied (i.e., xl|t=0 =

−xr|t=0 and xl|t=0 = −xr|t=0), cell length modulation that would otherwise be bipedal is

in-phase across the trailing edge elements. Such an in-phase limit cycle is unstable in that

any small driving force applied to either trailing edge element causes the system to shift

106

Page 133: The Mechanics of Cell Motility and a Unifying Theory for

from the in-phase limit cycle to the bipedal limit cycle. These unstable states are ruled out

and not discussed further.

5.4. Comparing 2-D Elastic Configurations. There are several ways to elastically

couple the elements that make up the two dimensional model. Four viable configurations

that capture the prevalent shape dynamics of fish keratocytes are illustrated in the top

row of Fig. 4.8. In the previous work [135], it was shown that elastic coupling Config. 3

generates realistic bipedal dynamics over a range of spring elasticities, KW , and inertia-like

parameter, g. We investigate the dynamics of three alternative elastic coupling configura-

tions to determine if they reproduce realistic bipedal locomotion of crawling cells and lateral

displacement of the cell nucleus (Fig. 4.7 d).

5.4.1. Configuration 1. The simplest possible configuration that can produce bipedal

locomotion consists of two elastically coupled trailing edge elements both coupled to the

leading edge (Config. 1). The trailing edge elements, ~xl and ~xr, undergo bipedal dynamics,

while the front element, ~xf , pulls the cell forward at a constant velocity. A representative

dynamical response of this configuration is shown in the first column of Fig. 4.8. The

dynamics of cell length, ∆y, follow a bipedal limit cycle with a modulation amplitude of

1.9µm and a period of 35.3 s.

There are, however, three notable deviations between the simulation dynamics of Con-

fig. 1 and the actual bipedal dynamics of keratocytes. First, an aberrant secondary slip

occurs during the sticking phase on one side of the trailing edge coincident with slipping

of the opposite side. As the slipping side element, say ~xl, is propelled forward, increased

spring force through the KW spring causes momentary slipping of ~xr, which would other-

wise remain in the sticking phase. This secondary slip is a result of direct coupling between

the two trailing edge elements. More precise measurements of the dynamics of keratocyte

shape are necessary to determine whether these aberrations occur in real cell crawling.

107

Page 134: The Mechanics of Cell Motility and a Unifying Theory for

Second, stable bipedal limit cycles occur over a range of physiologically relevant param-

eters only when the ratio of spring equilibrium lengths, W/D, is near 2 or greater. For

W/D < 1.6, the range of parameters corresponding to stable bipedal limit cycles is signif-

icantly reduced if existing at all. In Fig. 4.8, the corresponding ratio is W/D = 1.6. The

average cell length in this simulation is approximately 10µm, less than that of a typical

keratocyte, and cannot be increased without widening the model cell as required to main-

tain the ratio of W/D ≥ 1.6. Alternatively, to generate bipedal locomotion for a cell of this

length without increasing cell width, one can increase the drag force coefficient of sticking,

β, or decrease spring constant, KD. Although these variations allow for a more realistic

overall cell aspect ratio, the amplitude of cell length modulation becomes unrealistically

large. After translating W and D into cell width and length, this configuration generally

does not reproduce the dynamics of a bipedally crawling cell unless the cell width to length

ratio is at least 1.5. We measure width to length ratios of bipedally crawling cells to range

from 1 to 3 (Fig. 4.6). Thus, while it is possible to simulate bipedal locomotion of wider fish

keratocytes, there are many aspect ratios that are not appropriately modelled by Config.

1.

Finally, this spring configuration does not recapitulate the smooth periodic nucleus

lateral displacement observed. Unlike the other configurations, Config. 1 has no centrally

coupled element representative of the cell nucleus. At best, one could interpret the nucleus

centroid to be the location of a fictitious spot marked at the middle of the posterior spring.

Mathematically, this is ~xn = (~xl+~xr)/2. The lateral motion of this spot is plotted as lateral

displacement for Config. 1 in Fig. 4.8. In this interpretation, the nucleus jumps laterally

from one side to the other every time one of the trailing edge elements slips. Without

unrealistically overdamping the system, smooth sinusoidal-like or exponential relaxation-

like nucleus lateral displacement cannot be generated.

108

Page 135: The Mechanics of Cell Motility and a Unifying Theory for

xn

xxl r

xfxf

xl xr l r

xn

x xl r

xf

la

tera

l d

isp

lace

me

nt

(μm

) ce

ll le

ng

th, ∆

y (

μm

)

∆y

Con!guration 1 Con!guration 2 Con!guration 3 Con!guration 4

l r

xn

xrlx

xf

0.01 0.02 0.03 0.040

0.25

0.5

0.75

1

frequency (Hz)

0.01 0.02 0.03 0.040

0.25

0.5

0.75

1

frequency (Hz)

0.01 0.02 0.03 0.040

0.25

0.5

0.75

1

frequency (Hz)

0 20 40 60 80 1008

10

12

14

16

18

time (seconds)

0 20 40 60 80 1008

10

12

14

16

18

time (seconds)

Fo

uri

er

am

plit

ud

e

0 20 40 60 80 1008

10

12

14

16

18

time (seconds)

0.01 0.02 0.03 0.040

0.25

0.5

0.75

1

frequency (Hz)

0 20 40 60 80 1008

10

12

14

16

18

time (seconds)

-1.5-1

-0.50

0.51

1.5

time (seconds)

100806040200-1.5

-1-0.5

00.5

11.5

time (seconds)

100806040200-1.5

-1-0.5

00.5

11.5

time (seconds)

100806040200-1.5

-1-0.5

00.5

11.5

time (seconds)

100806040200

left ∆yl

∆yrright

g α β γ KN KD KL KW N D L W

Config. (nN s2

µm) (nN s

µm) (nN s

µm) (nN s

µm) ( nN

µm) ( nN

µm) ( nN

µm) ( nN

µm) (µm) (µm) (µm) (µm)

1 0.1 0.15 12 n/a 0 1 0 1 n/a 15 n/a 242 0.1 0.5 50 5 1 0.9 1 1 8 13 12 203 0.1 0.5 25 5 1 0 1 10 3 n/a 10 204 0.3 0.5 50 5 1 0.7 1 0 8 13 10 n/a

Figure 4.8. Representative dynamical responses of four select config-urations of the 2-D cell crawling model. Config. 1 is the simplest arrange-ment of springs necessary for bipedal locomotion. This configuration lacks thefourth element necessary to represent smooth oscillations of the nucleus in thelateral direction. Configuration 2 is the most general spring arrangement whena central element is added to the model. Configurations 3 and 4 are specialcases of Config. 2, where particular springs are removed by setting their corre-sponding spring constants to zero. Oscillations in cell length, ∆y (see secondrow), are shown for both the left side (green) and right side (blue) trailing edgeelements. Fourier transforms (third row) correspond to cell length oscillations.Lateral displacement (fourth row) corresponds to the average position of ~xl and~xr in the case of Config. 1, and the central element in the cases of Configs. 2-4.

5.4.2. Configuration 2. To address the discrepancies between actual keratocyte dynam-

ics and the dynamics of Config. 1, a central element (location ~xn) is introduced to the 2-D

109

Page 136: The Mechanics of Cell Motility and a Unifying Theory for

crawling cell model. The central element allows for alternative coupling pathways between

the two trailing edge elements, rather than the direct coupling through spring KD that

leads to aberrant secondary slips in the cell length dynamics. Config. 2 is the case where all

elements are elastically interconnected. Configurations 3 and 4, discussed below, are special

cases of Config. 2 each corresponding to the removal of particular springs. We assume sym-

metry about the axis of forward motion and require the cell to maintain a reasonable shape

with width and length comparable to our benchmarks. This shape requirement is achieved

by the appropriate assignment of equilibrium spring lengths and elasticities. Configura-

tions 2-4, illustrated in Fig. 4.8, are the only three spring configurations that satisfy these

constraints. Unsuitable Configs. include those that do not have the necessary springs to

maintain cell width {KD = KW = 0, KL = KW = 0, or KN = KW = 0}. For each Config.

in this set, the nuclear and trailing edge elements move from their initial locations to the

axis of forward motion indicating an unrealistic collapse of cell shape into one dimension.

Config. 2 generates bipedal locomotion over many regions of parameter space. For ex-

ample, choosing parameters similar to those of Configs. 1, 3 or 4 results in a comparable

dynamical response despite the additional elastic coupling. In Fig. 4.8, Config. 2 parameters

are chosen similar to those listed for Config. 4 (discussed below) and the result is realistic

bipedal dynamics with cell length modulation and nucleus lateral displacement amplitudes

of 2.9 and 0.53µm, respectively, and a period of 40.0 s. Unlike cell length limit cycle shown

for Config. 1, here there is no secondary slip. Due to the maximal number of springs in

this configuration, however, there are many cases of aberrated bipedal limit cycles associ-

ated with Config. 2 despite this representative result in Fig. 4.8 not being one such case.

Compared to configurations with fewer springs, the dynamical behaviours produced with

Config. 2 are more sensitive to parameter variation, especially spring length variations. In

110

Page 137: The Mechanics of Cell Motility and a Unifying Theory for

the KD → 0 and KW → 0 limits, we found that Config. 2 robustly generates bipedal lo-

comotion without secondary aberrations. With this insight, we analyze the KD = 0 and

KW = 0 cases in more detail, which correspond to Configs. 3 and 4, respectively.

5.4.3. Configuration 3. Config. 3 is the spring arrangement considered by Barnhart et

al. [135]. They found that stable bipedal locomotion occurs for choices of spring elasticity

KW > 3 nN/µm (0.01 < g < 10 nN s2/µm), consistent with benchmark values. Not dis-

cussed in the previous work, one finds that the central element, ~xn, smoothly oscillates in

the lateral direction entrained to bipedal limit cycle. The amplitude of these oscillations

is 0.55µm, comparable to experimental observation (Fig. 4.7). This amplitude can be ad-

justed by varying the nuclear drag coefficient, γ, but due to the interplay of spring and

nuclear drag forces, cannot be increased much greater than 0.6µm.

An interesting transition in dynamics occurs when the nuclear drag coefficient, γ, is

increased to make the nucleus insensitive to the spring forces through the KL springs. One

gets a crude estimate of this critical value, γc, by using the dynamics shown in Config. 3 to

determine the drag force necessary to counterbalance these spring forces. Here, two springs

of elasticity KL ∼ 1 nN/µm apply forces to the central element, which in turn laterally

traverses ∼ 1µm every half period ∼ 20 s, of the bipedal limit cycle. Combining these

numbers gives γc ≈ 10 nN s/µm. When the nucleus is immobilized with respect to forces

from the KL springs, the system should behave like Config. 1 in which case element ~xn

assumes the role of the leading edge element, ~xf . As predicted, we find that the dynamical

responses of Config. 3 approach those of Config. 1 when γ > 7 nN s/µm. In essence, Config. 3

can be considered an extension of Config. 1 in which an extra spring at the front is added

to extend the cell length. Whereas cell shape dynamics produced by Config. 1 were not

robustly bipedal when W/D < 1.6, the front spring length of Config. 3 can be specifically

tuned to account for proper cell length.

111

Page 138: The Mechanics of Cell Motility and a Unifying Theory for

5.4.4. Configuration 4. Although Config. 3 produces realistic bipedal dynamics, we pro-

pose Config. 4 because the springs are easier to interpret with respect to the framework of

the actin cytoskeleton. As discussed in the main article, we determined that this spring con-

figuration generates realistic crawling dynamics that are sensitive to changes in KD, but not

KL (Fig. 4.4 b). Sensitivity with respect to drag force parameters β, γ, and g is investigated

here. Drag force on either trailing edge element is pointed in the direction opposing motion,

which at any given time, primarily opposes spring forces from the KD springs. Therefore,

tuning the sticking coefficient β is similar to tuning the spring elasticity KD, in terms of

changes to the dynamical behaviour of Config. 4. These dynamical behaviours are plotted

in Fig. 4.9 a. As stick drag force is scaled up by increasing β, the trailing edge elements

stick longer giving rise to monotonically increasing amplitudes of cell length modulation,

nucleus lateral displacement (NLD), and period. Fig. 4.9 b shows how increasing nucleus

drag coefficient, γ, decreases nucleus lateral displacement. Finally, Fig. 4.9 c shows that

the dynamics are relatively insensitive to variation of parameter g, which sets the timescale

between sticking and slipping.

Movie Legends (Movies Online)

Movie S1. Live tracking of the shape dynamics of a coherent non-bipedalkeratocyte. This movie is representative of our analysis of previously published moviesof fish keratocytes. (a) Overlaid on raw footage of the crawling cell are the perceivedcoordinates of the left and right side trailing edge elements, the nucleus element, and theleading edge element. The line approximates the axis of symmetry. The scale bar is 20µm.Cell aspect ratio, width, cell length oscillations, and cell speed are measured based on thepositions and movements of these elements. (b) Plot of the left and right side cell lengthdynamics. Left cell length is greater than right cell length because the cell is turning to theright with respect to its direction of motion. Despite noisy data, small stick slip events aredetected in this plot including one caused by collision with a fragment at 285 s.

Movie S2. Two dimensional simulation movie representative of a fish kerato-cyte undergoing bipedal locomotion. (a) Animation showing the four element model(Config. 4). Dashed lines indicate springs. (b) Cell length limit cycles for the trailing left(green) and right (blue) elements corresponding to (a). Cell length, ∆y, is defined as thevertical distance between the front element and either one of the trailing edge elements.This simulation corresponds to the timelapse and dynamics shown in Fig. 4.8, Config. 4,and Fig. 4.3. Parameter values are listed in Fig. 4.8.

112

Page 139: The Mechanics of Cell Motility and a Unifying Theory for

0 0.2 0.4 0.6 0.80

1

2

3

4

0

15

30

45

60

75a

mp

litu

de

m)

g (nN s2/µm)

NLD

cell length

period

20 40 60 80 1000123456

0

15

30

45

60

75

am

plit

ud

e (µ

m)

β (nN s/µm)

NLD

cell length

period

0 4 8 12 16 200

1

2

3

4

0

15

30

45

60

75

γ (nN s/µm)

pe

riod

(s)

NLD

cell length

period

(a)

(c)

(b)

realistic bipedal,

γ = 5 nN s/µm

amplitude of celllength oscillations, both ∆yl(t) and ∆yr(t)

amplitude of nucleus lateral displacement(NLD), ∆xn(t)

period of oscilations

vf

∆xn(t)

∆yl(t) ∆yr(t)KD

KLKL

KD

α,β,gα,β,g

γ

Figure 4.9. Drag force sensitivity analysis of Config. 4. Amplitudes ofcell length and nucleus lateral displacement (NLD) oscillations (blue curves),as well as the limit cycle period (green curve), are plotted with respect to dragparameters β (a), γ (b), and g (c). Cell length oscillation amplitude and periodmonotonically increase with the increase of all three drag parameters, but aremost sensitive to the stick drag coefficient, β. Dynamics are insensitive to freeparameter g, which sets the time of switching between sticking and slipping. Pa-rameters not varied in each chart are kept fixed at the values listed for Config. 4in Fig. 4.8.

113

Page 140: The Mechanics of Cell Motility and a Unifying Theory for

CHAPTER 5

Immune System Modifier β-Glucan Regulates Motility

Through the Actin Cytoskeleton

The contents of this chapter are intended to be part of a future manuscript.

114

Page 141: The Mechanics of Cell Motility and a Unifying Theory for

Abstract and Forward. Neutrophil motility is an integral component of the innate

immune system. As they migrate, neutrophils exert cytoskeletal driven contractile forces

on their environment through integrin borne adhesions, but the relationship between these

mechanical outputs and immune system potentiating biological response modifiers has yet

to be determined. We use live cell imaging and traction force microscopy to study the

relationship between neutrophil motility, force exertion, and the biological response modifier

β-glucan, a substance that has shown promise as a clinic grade therapeutic. Subsequently,

we find that β-glucan modifies the mechanical output of neutrophils on soft, 5 and 25 kPa

2D substrates, decreasing both the contractile force and energy applied as the concentration

of β-glucan is increased over a clinically relevant range. Despite these changes in mechanical

output, contractile energy is conserved across substrate stiffness for all concentrations of

β-glucan tested. We also report an overall inverse correlation between migration speed

and mechanical output. While there has been some speculation that a specific force range

exists for facilitating optimal cell migration, we conclude that optimal cell migration is not

associated with a specific magnitude of contractile force in the 1-100 nN range. Instead,

increased motility is associated with diminished contraction.

1. Introduction

1.1. Neutrophils. Neutrophils are the first responders to injury and infection and

are integral components of the innate immune system [8]. To reach sites of injury and

infection, neutrophils must transmigrate through the endothelial barrier of the blood vessel

and migrate within the extracellular matrix (ECM). During migration, neutrophils interact

mechanically with their surrounding environment through integrin borne adhesions [164].

Forces exerted by neutrophils on their substrate are a fundamental component of the sub-

strate stiffness dependent regulation of migration [48, 50, 165], which ultimately affects

wound healing [84] and other host defense functions [166–168].

115

Page 142: The Mechanics of Cell Motility and a Unifying Theory for

1.2. Traction Forces. The net force required to propel a neutrophil in typical cell

media one cell length per minute without friction is on the order of 10−6 to 10−5 nN.1 Yet

the forces that many motile neutrophils exert on their substrate in vitro in 2D and 3D

range from 1 to 100 nN in magnitude [48, 50, 165], up to 8 orders of magnitude larger than

the frictionless propulsion force. There are two main reasons for such large forces on the

substrate. First, cells apply force to their environment as a means of probing its rigidity

[29], which in turn regulates many cell characteristics from morphology to migration speed

[48, 49]. This detection of the extracellular rigidity is termed mechanosensing. Second,

as adhesion complexes form, a significant amount of friction must be overcome to sustain

motility (c.f. Fig. 4.1). Several models of cell motility would indicate that an intermediate

range of contractile force exertion is optimal for facilitating cell migration. Relatively large

forces are thought to indicate motility-impairing levels of adhesion, whereas relatively small

forces are thought to indicate an inability to form mature adhesion complexes [41, 42, 169].

1.3. Force Motility Relationships. Bangasser et al. have recently compiled the

migration speeds and contractile forces of several motile cell types crawling on soft substrates

[42]. Their analysis suggests that many cell types exhibit a biphasic speed-stiffness and

force-stiffness dependence. They adapt a motor-clutch model [170] to fit the latter relation.

Missing is a detailed analysis for the force/energy-speed dependence. With neutrophils

migrating on fibronectin coated polyacrylamide gels with stiffnesses between 3 and 13 kPa,

Stroka et al. have shown that migration speed peaks on stiffnesses near 7 kPa, depending

on fibronectin density [41]. The peak root mean squared (RMS) speed asymptote was

roughly vrms(∞) = 0.05-0.1µm/s (see Materials and Methods for definition). In similar

experiments, Oakes et al. and O’Brien et al. have shown that the neutrophil RMS speed

asymptote is vrms(∞) = 0.15-0.2µm/s on gels of 10 kPa stiffness, and drops monotonically

1The net force required to propel a neutrophil without friction is estimated here using Stokes’ law,F = 6πµRv. Plugging in order of magnitude estimates corresponding to neutrophil motility in atypical cell media, µ 10−3 kg/m/s, R 10−5m, v 3 ·10−7 m/s, gives F ≈ 10−6 nN, more than six ordersof magnitude smaller than the forces reliably measured with traction force microscopy.

116

Page 143: The Mechanics of Cell Motility and a Unifying Theory for

as stiffness increases from 5 to 100 kPa [1, 48]. Furthermore, the average traction stress

of neutrophils was shown to increase with substrate stiffness over the 5 to 50 kPa range.

Taken together with other neutrophil speed and traction measurements from [49] and [50],

it is possible to reconstruct a rudimentary force-speed distribution. But, since only the

ensemble averages are available, much of the force-speed distribution is missing, and there

is no way to determine the energy-speed distribution. These distributions would be highly

useful for modeling cell motility. In this manuscript, we measure the force-speed and energy-

speed distribution explicitly for neutrophils migrating on compliant substrates with Young’s

moduli of 5 and 25 kPa.

1.4. β-Glucan. To further examine the connection between cell mechanics and speed

in a context relevant to the immune system, we investigate the effect of the pathogen

associated molecular pattern (PAMP), β-glucan (BG) [76]. BG is a complex polysaccharide

ubiquitous in the cell walls of bacteria and fungi such as Saccharomyces cerevisiae [171, 172].

Several forms of BG are also the main element of fiber in grains such as oats and barley [173].

BG from yeast can be purified into a clinical grade, water soluble form called PGG-glucan

[69], which has been shown to be a biological response modifier in vivo [174], improving

wound healing in mice [70], showing promise as an anti-cancer agent [71–73], and enhancing

chemotactic honing of neutrophils on glass [75, 76]. Several clinical trials have shown that

0.5-2.25 mg/kg intravenous dosages of PGG-glucan to humans with colorectal cancer or post

surgical infections significantly improved recovery, highlighting its potential as a therapeutic

drug for humans [74].

While BG has potential to be a positive pharmacological intervention that improves the

function of the innate immune system, the role of cell mechanics as part of this positive

intervention and immunological response has yet to be studied. BG is known to bind and

allosterically regulate the affinity of Complement Receptor 3 (CR3, also known as αMβ2,

Mac-1, and CD11b/CD18) [77, 79], an integrin expressed on the surface of neutrophils that

117

Page 144: The Mechanics of Cell Motility and a Unifying Theory for

plays a significant role in mechanosensitive neutrophil behavior [1, 48]. Little is known about

the link between biological response modifiers and the cell mechanics of immune cells, but it

is rational to propose that BG (and possibly other biological response modifiers) regulates

their mechanical output. This hypothesis is tested below using traction force measurements

of motile neutrophils treated with soluble BG.

2. Materials and Methods

2.1. Reagents. 40% acrylamide solution, 2% bis-acrylamide solution, N,N,N’,N’ -

tetramethylethylenediamine (TEMED) and ammonium persulfate (APS) were purchased

from Bio-Rad Laboratories (Mississauga, ONT, USA). Dulbecco’s PBS, Lebovitz’s L-15

medium, Hanks Balance Salt Solution without Ca2+/Mg2+ (HBSS−/−), and 0.5µm Fluo-

spheres carboxylate-modified microspheres were purchased from Invitrogen (Carlsbad, CA,

USA). Polymyxin B and sulfo-SANPAH were purchased from Pierce Biotechnology. Human

fibronectin isolated from plasma (>95% purity), histopaque 1077, dextran for neutrophil

prep (average molecular weight, 400-500 kDa), dextran for osmolarity control experiments

(Mw 5000), formyl-methionyl-leucyl-phenylalanine (fMLP), and sterile dimethyl sulfoxide

(DMSO) were purchased from Sigma-Aldrich. Pharmaceutical-grade purified, endotoxin

free, soluble yeast β-glucan (Imprime PGG) in citrate buffer was generously provided by

Biothera (Eagan, MN, USA). β-glucan is referred to as BG throughout this work.

2.2. Substrate Preparation. Polyacrylamide gel (PAG) substrates with Young’s

moduli of 5 and 25 kPa were prepared as described previously [48]. Acrylamide and bis-

acrylamide solutions (Bio-Rad) were mixed with 0.5µm fluorescent spheres (Invitrogen)

and polymerized using TEMED and APS (Bio-Rad). Gels were cast on silanized Bioptechs

Delta-T glass dishes inside air-tight molds made from gene frames and plastic coverslips.

Immediately after catalyzing PAG polymerization and sealing the air-tight mold (before

polymerization), gel mixtures were centrifuged at 1000 g at 4◦C for 25 minutes to sediment

118

Page 145: The Mechanics of Cell Motility and a Unifying Theory for

fluorescent spheres along a subapical layer at the top surface of the PAG (Fig. 5.1 a). After

centrifugation, the gels were allowed to polymerize at room temperature. The final size

of the gels was ∼ 1 cm× 1 cm× 300µm. At ∼ 300µm thick, the underlying glass has no

bearing on the stiffness detected by cells plated on top [175], nor do applied traction forces

in the cellular force range of 1 to 100 nN penetrate through the gel down to the glass [37].

Gel stiffness measurements and substrate functionalization with fibronectin (Fig. 5.1 a) was

done exactly as described in Ref. [1] .

2.3. Neutrophil Preparation. Under the approval and guidelines of the Rhode Is-

land Hospital Institutional Review Board, neutrophils were isolated from healthy human

volunteers by collection into EDTA-containing Vacutainer tubes (BD Biosciences, San Jose,

CA, USA). Histopaque 1077 (Ficoll Histopaque) was used for the initial cell separation,

followed by gravity sedimentation through 3% dextran (average molecular weight, 400-

500 kDa). Contaminating erythrocytes were removed by a single hypotonic lysis, yielding

a neutrophil purity of >95%. Neutrophils were suspended in HBSS−/− on ice until use in

the experiments. All reagents contained <0.1 pg/ml endotoxin, as determined by Limulus

amoebocyte lysate screening (BioWhittaker, Walkersville, MD, USA). When indicated, neu-

trophils were pretreated on ice for 30 min with 7µL (∼ 1 mg/mL) of monoclonal anti-CR3

antibody, clone 44abc, isolated from hybridoma cells (American Type Culture Collection,

Manassas, VA, USA) per 106 cells (total volume = 500µL).

For traction-migration assays, approximately 106 neutrophils were resuspended at 37◦C

in 2 mL L-15 with 2 mg/mL glucose added. When indicated, this media was modified

to contain 10µg/mL β-glucan (BG10), 50µg/mL β-glucan (BG50), or 50µg/mL dextran

(Dex50) in citrate buffer. These doses roughly correspond to the BG weight to blood volume

ratios of humans intravenously dosed with BG during recent phase III clinical trials [74].

Neutrophils were allowed to settle for at least 5 min before introducing fMLP (10 nM final

concentration) to activate and induce migration.

119

Page 146: The Mechanics of Cell Motility and a Unifying Theory for

2.4. Microscopy. A Nikon TE-2000U inverted microscope (Nikon, Melville, NY,

USA) coupled to an iXonEM + 897E back-illuminated Electron Multiplying CCD cam-

era (Andor, Belfast, UK), outfitted with a Bioptechs (Butler, PA, USA) stage heater, a

Nikon 40× CFI S Plan Fluor ELWD objective, and a FITC dichroic optical cube (Nikon)

was used for all live cell imaging experiments. DIC and fluorescent bead field images were

captured over 80-100 min in 10 s intervals using the Elements program (Nikon). To minimize

photo damage to the cell, fluorescent images were captured with long exposures (> 1 s) at

low illumination intensity rather than short exposures at high illumination intensity. All

data were analyzed using ImageJ (U.S. National Institutes of Health, Bethesda, MD, USA),

and MATLAB (MathWorks, Natick, MA, USA) computational software.

2.5. Cell Tracking and Migration Analysis. Perceived cell boundaries were traced

from DIC images using custom MATLAB software and a stylus and touch screen (Fig. 5.1 b,

white boundaries). Sampling intervals after manual tracing were either 20, 40, 80, or 160 s

(usually the latter). We found this manual tracing method more accurate compared to

using automatic tracking software. From these cell boundary measurements, the regionprops

algorithm in MATLAB was used to calculate cell centroid ~r(t), spread area, and aspect ratio.

The turning angles between successive displacements (e.g. the angle between ~r(t+∆t)−~r(t)

and ~r(t)−~r(t−∆t), where ∆t is the sampling time interval) were also tracked. Persistence

fractions were calculated as the fraction of all turning angles between ±30◦ calculated with

a sampling interval of 160 s.

Mean squared displacement was defined in the usual way [1, 48, 95],⟨r2c (τ)

⟩=⟨

|~rc(t+ τ)− ~rc(t)|2⟩c, where the overline indicates a time average over t, τ is the time

interval over which displacements are measured, c = 1, 2, 3, ... is the index of the migration

trajectory of a specific cell, and 〈...〉c indicates an ensemble average of all cell trajectories,

c. To minimize nonergodic contributions to MSD and increase the comparability of this

statistic, all trajectories were truncated at t = 1200 s (20 min) before calculation.

120

Page 147: The Mechanics of Cell Motility and a Unifying Theory for

2.6. Traction Force Microscopy. Images of bead fields were compared to a reference

image before cells adhered to the substrate. Corresponding displacement field ~u(~x, t), and

traction stresses ~T (~x, t), were calculated in 2D space (~x) and time (t) using digital area

correlation and Fourier-transform traction cytometry [33] as described previously [48].

0.5µmpolyacrylamide gel

10−25µmintegrin�bronectinactin cytoskeleton�uorescent sphere

a

b

c

5kPa

BG0 BG50

25kPa

BG0 BG50

10µm

100nN

10µm

stress islands

20nN

traction stress (10-3nN/µm2) traction stress (10-3nN/µm2)0 375 750 1125 15000 125 250 375 500

Figure 5.1. Traction force microscopy diagram and representative DICand traction field frames of cells on soft and stiff (5 kPa and 25 kPa)substrates with β-glucan concentrations of 0 (BG0) and 50µg/mL(BG50). (a) A neutrophil on the surface of a 2D polyacrylamide gel substratecoated with fibronectin (triangles). Contractile forces on the substrate are facil-itated by integrins (rectangles) that couple the substrate to the retrograde flowof the actin cytoskeleton (inward cyan arrows). (b) Representative images ofneutrophils and (c) corresponding traction stress maps show examples of stressislands (outlined in black) and their corresponding forces (black arrows). Colorindicates the absolute value of traction stress, whereas small white arrows showtraction stress direction. Traction stresses were larger on the 25 kPa substrates(note the change in color bar scale).

2.7. Force and Energy Measurements. To minimize the measurement of noise,

traction stresses below 0.1 and 0.3 nN/µm2 noise thresholds were ignored on the soft and

121

Page 148: The Mechanics of Cell Motility and a Unifying Theory for

stiff gels (5 and 25 kPa), respectively. Stress islands were defined either as distinct closed

regions of traction stress above the noise threshold, or partitioned subregions from one closed

region above the noise threshold. Examples of both cases can be seen in Fig. 5.1 c. The

5 kPa BG50 shows three distinct stress islands whereas the other panes show one traction

region partitioned into either two or three stress islands. Where appropriate, partitions

were made along lines of local traction stress minima.

At times, stress islands encompassed areas outside the perceived cell boundary. Nonva-

nishing stress measurements outside the perceived boundary occur for three reasons: First,

the perceived cell boundary differs from the actual cell boundary as visualized in Fig. 5.7

with a plasma membrane stain. Second, the displacements are Gaussian filtered, which

slightly blurs the stress islands. While blurring slightly extends the area of each stress

island, the integrated force is not affected. Third, we determine the traction stress field by

solving the inverse Bousinesq equation, which assumes point-like traction sources [33]. It

has been shown that the assumption of finite sized traction sources on the order of the area

of a small integrin cluster leads to stress regions that are more contained within the per-

ceived cell boundary [35], but this analytical method was not applied here. Taken together,

we take a justified as-is approach to analyze the stress field.

Once stress islands were identified, the force from the nth island was calculated as

~Fn(t) =∫n~T (~x, t)d~x, where ~T is the 2D traction stress field, ~x is the position in 2D, and

t is time. The average stress island magnitude, denoted as contractile force throughout

this work, was calculated as⟨|~Fn(t)|

⟩n, where 〈...〉n is an average over each stress island.

Contractility time was defined as the time that contractile forces were above thresholds

of 8 nN on a 5 kPa substrate and 25 nN on a 25 kPa substrate. The energy required to

stress the substrate (denoted contractile energy throughout this work) was calculated as

E =∑n

1

2

∫n

~T (~x) · ~u(~x)d~x ≈∑n

∆z

2Y

∫n

∣∣∣~T (~x)∣∣∣ d~x, where ~u is the 2D displacement field.

122

Page 149: The Mechanics of Cell Motility and a Unifying Theory for

2.8. Dipole Moment Analysis. Similar to the procedure outlined in [176], dipole

moments were calculated from the stress dipole tensor, Mij =∫S xiTjdS, where S is a

80µm× 80µm bounding box centered at the cell centroid that captures the relevant stresses,

and xi is the ith element of the position vector (i = 1, 2), and Tj is the jth element of the

traction stress vector (j = 1, 2). The diagonal terms M11 and M22 correspond to linear

extensile or contractile stress in the x1 and x2 directions, respectively. The off-diagonal

terms correspond to off-axis stress. In the absence of net torque within the bounding

box, the dipole tensor is symmetric and diagonalizable with orthogonal eiganvectors. The

principle dipole moments, denoted ~µM and ~µm for the major and minor dipole moments,

respectively, were calculated by diagonalizing the dipole tensor, Mij , and multiplying each

unit eigenvector by their corresponding eigenvalue. The major dipole moment was the

vector with the larger magnitude.

2.9. Statistics. Significance has been determined by assessing the overlap of standard

error measurements. All shaded regions on time plots represent ± one standard error of

the mean drawn about the mean. All error bars represent standard error of the mean.

Pearson product-moment correlation coefficients (PCCs) between two paired variables, for

example, A1 and A2, were calculated as ρ = 〈A1A2〉−〈A1〉〈A2〉σA1

σA2, where σ2

Ai= 〈A2

i 〉−〈Ai〉2 is the

variance (i = 1, 2). PCCs characterize the linear correlation between two paired variables,

but not the nonlinear correlation. To measure the nonlinear correlation, alternating condi-

tional expectation (ACE) [177] calculations were made using an ACE algorithm described

in Ref. [178].

3. Results

3.1. Spread Area and Motility Biphasically Dependent on BG Concentra-

tion. To tease out the force-speed and energy-speed relationships, we tracked the spread ar-

eas, migration trajectories, and traction stresses of fMLP induced chemokinetic neutrophils

123

Page 150: The Mechanics of Cell Motility and a Unifying Theory for

(example shown in Movie S1). Ensemble averages of the MSD, spread area, contractile

energy, and contractile force are plotted in Fig. 5.2. Zero time, t = 0, refers to the moment

fMLP was added, approximately five minutes after neutrophils were plated. Notationally,

BG0, BG10, and BG50 refer to soluble BG concentrations of 0, 10, and 50µg/mL, respec-

tively. BG is shown to alter spread area and MSD in a biphasic manner. Specifically, MSD

and spread area were smallest with BG10 cells compared to the other BG concentrations,

irrespective of substrate stiffness (Fig. 5.2, a and b). Interestingly, while BG10 cells showed

decreased cell speed with increasing substrate stiffness, this trend reversed with the BG50

cells. Also interesting, substrate stiffness did not affect cell migration speed under the BG0

condition. In Ref. [48], speed was affected by substrate stiffness but the difference they

found may be attributed to not accounting for nonergodicity in their MSD calculations.

Finally, there was a small relative increase in spread area as substrate stiffness increased,

irrespective of BG concentration.

3.2. Mechanotactic Conservation of Energy and Contractility Time. Sum-

ming up the total energy required to strain the substrate at a given point in time, the strain

energy (also denoted contractile energy) applied by neutrophils significantly decreased with

increasing concentrations of BG (Fig. 5.2 c). Contractile force, measured by integrating

traction stress over discrete stress islands and averaging the magnitudes of those forces (see

Materials and Methods), also decreased with increasing concentrations of BG. Osmolar-

ity control experiments using 50µg/mL dextran showed that decreases in the mechanical

output of neutrophils was not attributable to the added osmolarity of adding BG (Suppla-

mentary Material, Fig. 5.8). Futhermore, measurements of adhesion rates, percent active,

and percent of active cells contractile, showed no change between the BG0 and BG50 con-

centrations (Supplamentary Material, Fig. 5.9). In context, the mean contractile force of

BG0 neutrophils on 5 kPa substrates varied from 20 to 30 nN, values remarkably similar

to the reported mean RMS force of 28 ± 10 nN associated with neutrophils migrating on

124

Page 151: The Mechanics of Cell Motility and a Unifying Theory for

25 kPa5 kPa

a

c

b

d

e f

50

100

150

200

250

300

350

spre

ad

are

a (µ

m2)

time, t (min) time, t (min) frequency

time interval, τ (s) time interval, τ (s) frequency

0 10 20 300

20

40

60

80

100

120

con

tra

ctile

fo

rce

(n

N)

0 10 20 30

MS

D (µ

m2)

102 103

102

103

102 103

0

2

4

6

8

10

12

con

tra

ctile

en

erg

y (

nNµ

m)

0 5 10

0

50

100

150

0 5 10 150

10

20

30

0 10 20

200

400

600

500

300

100

0 20 40

0

2000

4000

6000BG0

BG10

BG50

5 kPa 25 kPa0

10

20

30

40

50

con

tra

ctili

ty t

ime

(m

in)

0

10

20

30

40

5 kPa 25 kPa

pe

rsis

ten

ce f

ract

ion

(%

)

Figure5.2. Neu-trophil motil-ity and me-chanical out-put depend onBG concen-tration andsubstrate stiff-ness. (captionon pg. 126)

125

Page 152: The Mechanics of Cell Motility and a Unifying Theory for

9 kPa PAG substrates coted with a mixture of E-selectin and ICAM-1 [165]. Contractile

forces on the 25 kPa substrates were less than five times larger than their 5 kPa counter-

parts (Fig. 5.2 d), indicating that neither the magnitudes of the substrate displacements,

nor the corresponding magnitudes of the contractile forces, were conserved across substrate

stiffness. This result is consistent with the force-stiffness observations of several other cell

types [42, 51] including Madin-Darby canine kidney cells [47], mouse embryonic fibroblasts,

and NIH3T3 fibroblasts [109]. Other than decreasing the absolute magnitudes of contrac-

tile forces, BG did nothing to alter the relative force-stiffness dependence as there was a

three-fold relative increase in contractile force magnitude going from 5 to 25 kPa substrates,

irrespective of BG concentration.

Most interestingly, the contractile energy was conserved across substrate stiffness, and

this conservation held for all concentrations of BG. The conservation of contractile energy

across substrate stiffness was first demonstrated with NIH 3T3 fibroblasts very recently by

Oakes et al. [179]. This conservation has never been demonstrated with neutrophils or with

a biological response modifier.

(Fig. 5.2 caption continued) For all time series plots, the shaded regions correspond to the meanplus and minus the standard error of the mean. The notation BG0, BG10, and BG50 refers to 0,10 and 50µg/mL concentrations of soluble BG. (a) Mean Squared Displacement (MSD) plots inlog-log coordinates. To minimize nonergodic contributions to MSD and ensure the contributions ofeach time average were unbiased, all migration trajectories were truncated at 20 min for the purposeof calculating MSD. All log-log MSD slopes range from 1 to values less than 1.5 indicating slightlysuper diffusive chemokinetic motion at asymptotic time intervals. Cell speed changes biphasicallyas a function of BG concentration on both soft and stiff substrates, with stiffness slightly tuningthe biphasic response. (b) Spread area plotted against time. As with MSD, the spread area alsobiphasically responds to BG concentration. (c) Strain energy, the potential energy loaded into thesubstrate by the cell at a particular point in time. Strain energy is conserved across substratestiffnesses, regardless of BG concentration. Increasing BG leads to decreased strain energy and(d) contractile force. In each of (a-d), histograms to the right correspond to each distribution ofdata at 560 s on the 25 kPa substrate. (e) Contractility time is the time contractile force remainsabove 8 nN on the 5 kPa substrates, and above 42 nN on the 25 kPa substrates. No significantchange in contractility time was observed between the BG0 and BG10 cells, but contractility timesignificantly decreased with 50µg/mL BG concentrations compared to either BG0 or BG10 cells,on both substrate stiffnesses. (f) Persistence fraction is the fraction of turning angles between±30◦. Like MSD and spread area, persistence time trends biphasically with BG concentration.n = 19, 21, 20, 18, 29, 23 for the 5 kPa BG0, BG10 and BG50, and the 25 kPa BG0, BG10, andBG50 conditions, respectively.

126

Page 153: The Mechanics of Cell Motility and a Unifying Theory for

Contractility time is defined to be the amount of time a cell contracts its substrate (see

Materials and Methods). No significant change in contractility time was observed between

the BG0 and BG10 cells, but contractility time significantly decreased going from either

BG0 or BG10, to BG50 cells (Fig. 5.2 e). Contractility time, like contractile energy, was

conserved across substrate stiffness, irrespective of BG concentration.

The ideal measure of the persistence of chemokinetic trajectories in 2D is persistence

time [56, 60, 62], the average time a cell crawls before changing direction. However, the

persistence times of our trajectories were not resolvable with confidence because they were

much smaller than the sampling time. To get a measure of the relative persistence, we

instead calculate turning angle distributions. Turning angles in our context were defined

as the angles between successive trajectory tangent vectors measured in 160 s increments

(see Materials and Methods), and the persistence fraction was defined as the fraction of all

turning angles within ±30◦ (Fig. 5.2 f). These persistence fractions are not a measure of how

long a cell is likely to travel before turning, but rather the appearance of directionality at

160 s time scales. The relative values of persistence fraction showed a biphasic relationship to

BG concentration similar to that of the MSD and spread area. As with contractile energies

and contractility times, persistence fractions were conserved across substrate stiffnesses,

irrespective of BG concentration.

The correlations between spread area, motility characteristics, and the mechanical out-

put are graphed in Fig. 5.3. The six nodes correspond to variables that were simultaneously

measured for each cell (ensemble averages plotted in Fig. 5.2). Weighted lines connecting

each node indicate Pearson product-moment correlation coefficients (PCCs), which measure

the linear correlation between the pairs of variables. Negative PCCs (red lines) indicate

negative correlations and positive PCCs (black lines) indicate positive correlations. MSD,

spread area, and persistence fraction were all positively correlated, irrespective of stiffness.

Essentially, persistent cells traveled further than non persistent cells in a given amount of

127

Page 154: The Mechanics of Cell Motility and a Unifying Theory for

time, and the greater the spread area, the more persistent the cell was. PCCs indicate

correlation, not causality. However, we suspect that cells with greater spread areas are

more persistent because of increased integrin engagement and the accompanying requisite

to break more integrin bonds to change direction. Negative correlations exist between BG

concentration and energy, energy and MSD, MSD and contractility time, and contractility

time and BG. In the following section, we study these correlations in more detail to deter-

mine the links between energy and motility. One final note about PCCs, PCCs reflect only

the linear portion of correlation between pairs of variables. To complement the PCCs, alter-

nating conditional expectation (ACE) values are tabulated in the Supplementary Material

(Table 5.1). ACE values reflect the probability that any function, linear or nonlinear, can

be used to describe the relationship between pairs of variables. For example, whereas the

PCC between spread area and contractile energy on the 25 kPa substrate may indicate no

linear relationship between the two variables, the corresponding ACE value indicates that

a nonlinear relationship may exist.

3.3. Optimal Motility Correlated Inversely With Mechanical Output. The

PCC values graphed above indicate the possibility of a correlation between motility and

energy/force exertion. At present, little is known about the energy/force-speed dependency

and its elucidation would be useful for cell motility modeling. We find that the contractile

energy of faster moving neutrophils (MSD at 560 s> 500µm2) is significantly lower than

the contractile energy of slower moving cells (MSD at 560 s≤ 500µm2, Fig. 5.4, a and b).

The exception to this finding is with BG50 cells on 5 kPa substrates, where there were very

few significantly motile cells in the ensemble. Insets (panel b) show the entire distribution of

time average contractile energy against MSD at 560 s. The inverse speed energy correlation

is also borne out by the PCC values shown as ρ in the inset. As with the conservation of

energy across substrate stiffnesses, the inverse correlation between energy and speed was

also conserved in the absence of BG. The corresponding force-speed distributions are plotted

128

Page 155: The Mechanics of Cell Motility and a Unifying Theory for

beta glucan

contractile energy

MSD at 560 s

contractility time

persistence fraction

spread area

0.200.20

0.05

0.05

-0.2

7-0

.27

0.0

8

0.3

8

-0.08-0.08

-0.290.06

0.090.09

-0.07-0.07

-0.26

0.16

0.290.29

0.030.03

beta glucan

contractile energy

contractility time

persistence fraction

spread area

-0.07-0.07

0.12

0.12

-0.1

9-0

.19 -0

.03

-0.15-0.15

-0.22-0.0

5

0.610.61

-0.06-0.060.070.07

-0.12

0.490.6

2-0.18-0.18

0.39

MSD at 560 s

25 kPa5 kPa ba

Figure 5.3. Two graphs showing the Pearson product-moment corre-lation coefficients (PCCs) between variables describing the motilityand mechanical output of neutrophils. Each graph corresponds to onestiffness, (a) 5 kPa on the left, and (b) 25 kPa on the right. Each node cor-responds to a variable measured in Fig. 5.2. Values of MSD, spread area, andcontractile energy correspond to measurements at 560 s. PCCs have values from-1 to 1 estimating the linear portion of the correlation between the two variables.Line thickness is proportional to the magnitude of the PCC, with red denot-ing a negative correlation (PCC< 0), and black denoting a positive correlation(PCC> 0).

in Fig. 5.4 c. Contractile forces were determined by taking the average magnitude of vector

stress integrated over each stress island, whereas contractile energies were calculated by

integrating scalar traction stresses. Insomuch that one quantity is a vector sum and the

other is a scale sum, there was no reason to expect the force-speed distributions to mirror

the energy-speed distributions. However, the force-speed distributions were, overall, very

similar to the energy-speed distributions, showing the same inverse correlations with similar

PCC values. Overall, optimal motility correlated inversely with mechanical output, even

when the mechanical output was diminished by the biological response modifier, soluble

BG.

3.4. Dipole Case Study: Forces Perpendicular to the Direction of Motion. To

examine the spatiotemporal distribution of contractile forces that occured during migration,

129

Page 156: The Mechanics of Cell Motility and a Unifying Theory for

0

2

4

6

8

10

12

cont

ract

ile e

nerg

y (n

Nµm

)

25 kPa5 kPa

fastBG0

slowBG0

0 10 20 300 10 20 300

2

4

6

8

10

12

cont

ract

ile e

nerg

y (n

Nµm

)

slowBG50fastBG50

time (min) time (min)

a

b

c

cont

ract

ile e

nerg

y (n

Nµm

)MSD at 560s

(103µm2)

cont

ract

ile e

nerg

y (n

Nµm

)

cont

ract

ile fo

rce

( nN

)

MSD at 560s (103µm2)

0 1 2 3 4 50

4

8

12 ρ = -0.24ρ = -0.24ρ = -0.52ρ = -0.52

0 1 2 3 4 50

4

8

12 ρ = -0.24ρ = -0.24ρ = -0.08ρ = -0.08

1 2 3 50 4MSD at 560s (103µm2)

ρ = -0.20ρ = -0.20ρ = -0.49ρ = -0.49

0

20

40

60

80

100

120

140

1 2 3 50 4MSD at 560s (103µm2)

ρ = -0.26ρ = -0.26ρ = -0.10ρ = -0.10

Figure 5.4. Slower moving cells exert more energy contracting theirsubstrate than faster moving cells. Slower cells were gated from faster cellsusing a threshold MSD of 500µm2 (measured at 560 s). The ensemble averagedcontractile energies are plotted (a) in the absence of BG (BG0), and (b) inthe presence of 50µg/mL BG (BG50). Shaded (fast moving cells) and striped(slow moving cells) regions indicate the mean plus and minus one standard errorof the mean. Except with BG50 cells on 5 kPa substrates, slower moving cellsexerted significantly more energy than faster moving cells. With the BG50 cellson 5 kPa substrates, no significant difference in contractile energy was observed.(caption continued on pg. 132)

130

Page 157: The Mechanics of Cell Motility and a Unifying Theory for

dipole moment analysis was applied to the traction stress fields of two motile neutrophils

on 5 kPa subtrates, one with the BG0 concentration (Movie S1) and one with the BG50

concentration (Movie S2). Major and minor dipole moments (µM and µm, respectively)

were determined from principal moment analysis of the dipole tensor (see Materials and

Methods). Two variables were used to characterize the spatial distribution of forces: First

the relative angle between the major dipole moment and the direction of motion (angle θ,

Fig. 5.5 a). Second, the difference in the absolute magnitudes of the major and minor dipole

moments, |~µM | − |~µm|. Small differences indicate a relatively isotropic traction stress field

(such as an annulus, see Chapter 6, Fig. 6.7 e for an example), whereas large differences

indicate an anisotropic stress field that is principally oriented along the axis of the major

dipole moment vector, ~µM (as in Fig. 5.5 a). Forces on the BG0 cell were situated toward

the uropod, just as reported in [48, 50, 165], with the principal stress orientation often

nearly perpendicular to the direction of motion (Fig. 5.5, b and c, see Movie S1). As

reported previously, we observed reorganization of the stress island centers prior to changes

in direction [165]. However, once the force centers had stabilized into a perpendicular

mode, changes to the dipole axis trailed changes to the direction of motion. Therefore,

contractile force reorganization appears to be the driving force behind sharp changes in

direction, whereas smooth, shallow changes in direction are not the result of contractile

force reorientation. This observation is characterized in Fig. 5.5 b. Here the cell crawls with

an overall counterclockwise migration trajectory (Movie S1). The major dipole orientation

was most often between -50 and -70◦ with respect to the direction of motion, indicating that

changes to the dipole moment orientation lagged behind changes to the direction of motion.

With the exception of anomalous dipole calculations between 0 and 10 min (Fig. 5.5 d),

and during the time of significant traction stress reorientation close to the 15 min mark,

the traction stress distribution was primarily anisotropic in nature, as seen by the nearly

131

Page 158: The Mechanics of Cell Motility and a Unifying Theory for

three-fold difference between the major and minor dipole magnitudes over the 19 to 35 min

range.

We also analyzed a BG50 cell on a 5 kPa substrate. As reflected by the ensemble av-

eraged plots (Fig. 5.2), the BG50 cell migrated less persistently, had smaller spread area,

contractile forces, and contractile energies, compared to the BG0 cell (BG50 cell in Movie S2

compared with BG0 cell in Movie S1). With these differences though, the major dipole mo-

ment was still primarily either perpendicular or parallel to the direction of motion (Fig. 5.5, e

and f). Dipole moments became less pronounced ahead of turns made. As the persistence

of this BG50 neutrophil was much smaller, it was difficult to determine whether stabilized

dipole moments preceded or proceeded changes in direction. Overall, the spatial distri-

bution of the stresses was less organized, and the three-fold difference between major and

minor dipole moment magnitude was not observed (Fig. 5.5 g).

3.5. Estimating Active Contractile Force and Cytoskeletal Stiffness. Marcq

et al. recently introduced a simple 1D spring model to interpret the contractile forces ex-

erted by cells on soft substrates [51]. In their model, the cell is represented by a spring

with stiffness kC , and the substrate is represented by two surrounding springs of stiffness

ksub (Fig. 5.6 a). Inward “active” forces, ~FA, represent the total myosin mediated con-

tractile force transduced from the actin cytoskeleton to the substrate. Note, the force

measured by detecting substrate displacements is only a fraction of the total active force

because of the cytoskeletal spring. Using this model and converting the spring stiffnesses to

Young’s moduli, the magnitude of the measured contractile force follows the Hill function

F = FA

(Ysub

YC+Ysub

), where YC is the estimated Young’s modulus of the cytoskeleton, and

Ysub is the Young’s modulus of the substrate. Fitting the model using contractile force

(Fig. 5.4 caption continued) Insets show scatter plots of time averaged contractile energy versusMSD at 560 s for BG0 (yellow triangles) and BG50 (blue circles) cells on 5 kPa (left) and 25 kPa(right) substrates. The striped region in the scatter plots corresponds to slower moving cells belowthe MSD threshold. (c) Scatter plots of time averaged contractile force versus MSD at 560 s. Colorsand shapes correspond to the insets in (b).

132

Page 159: The Mechanics of Cell Motility and a Unifying Theory for

BG0

BG50

g

f

d

c

b

e

a

5% 11% 17% 23% 29%0% 90−90

0relative dipoleangle, θ(deg)

-375 > µM-250 > µM > -175-125 > µM > -250

µM > -125

major dipole moment (nN µm)

0 10 20 30−500

−250

0

250

time (min)

dipo

le m

omen

t (n

N µ

m)

µm

µM

−90−60−30

0306090

rela

tive

dipo

le

angl

e, θ

(deg

)

θ

minor dipole, µM

major dipole,µ

m

−90−60−30

0306090

rela

tive

dipo

le

angl

e, θ

(deg

)

90−90

0relative dipoleangle, θ(deg)

4% 8% 13% 17% 21%0%

dipo

le m

omen

t (n

N µ

m)

0 10 20 30−500

−250

0

250

time (min)

µM

µm

Figure 5.5. Dipole moment analysis of a motile BG0 cell migratingon a 5 kPa substrate. (a) Schematic of the major (red) and minor(blue)dipole moment axes. Stress island force vectors are shown for comparison. Theangle between the major dipole moment and the direction of motion, known asthe relative dipole angle, is defined to be θ. This PMN corresponds to movieS1 at 21 min. The scale bar illustrates 5µm. (b) Histogram of the relativedipole angles. On average, the dipole moment aligns nearly perpendicular tothe direction of motion. (c) Relative dipole angle plotted over time. (d) Majorand minor dipole moments plotted over time. Negative moments correspond tocontractile motion. Large differences between the major and minor momentsindicate a definitive axis of contractility, whereas small differences indicate anannular like traction field.

measurements on both 5 and 25 kPa substrates, the estimated cytoskeletal Young’s modu-

lus was determined to be within a range from 10 to 20 kPa, without significant dependence

on BG concentration (Fig. 5.6 b). Comparatively, the cytoskeletal moduli was one order

133

Page 160: The Mechanics of Cell Motility and a Unifying Theory for

0

10

20

30

40

50

0

50

100

150

BG10 BG50BG0

kC

FA FA

ksubksub

cyto

ske

leta

l mo

du

lus,

EC (

kPa

)

act

ive

fo

rce

, FA (

nN

)

BG10 BG50BG0

a

b c

Figure 5.6. Estimating total myosin mediated contractile force andcell stiffness. (a) A schematic of a 1D spring model proposed by Marcq et al.to estimate the intracellular contractile force, FA, and stiffness, kC , given theobserved contractile forces on a range of substrate stiffnesses, ksub. Solving thespring model at varying concentrations of BG show that, (b) the cytoskeletalstiffness does not change significantly with BG, and (c) the active contractileforce decreases with increasing BG. Error bars represent the standard errors ofthe mean.

of magnitude smaller than that predicted from Madin-Darby Canine Kidney fibroblasts

(∼ 100 kPa) [51], comparable to the cytoskeletal stiffness of fish keratocytes (10-150 kPa)

[2, 135, 180] but intermediate of Young’s moduli estimates for cortical actin (0.01-0.1 kPa)

[145] and stress fibres (∼ 1000 kPa) [142]. Increasing BG concentration decreased the ac-

tive contractile force (Fig. 5.6 c), suggesting either a loss of coupling strength between the

cytoskeleton and the substrate, or a decrease in myosin generated contraction. The mea-

surement of contractile forces on a wider range of substrate stiffnesses would improve the

confidence placed on these numbers, YC and FA.

4. Discussion

We have shown that contractile energy is conserved across substrate stiffness, and this

conservation remains with the addition of the biological response modifier, BG. The energy

and force per integrin can be estimated by assuming that roughly 300 integrins per µm2

134

Page 161: The Mechanics of Cell Motility and a Unifying Theory for

populate a stress island [29], and that each stress island ranges in area from 30-50µm2

(∼ 100µm2 total average stress island area). The result is an energy per integrin on the

order of 10 pNnm, and a forces per integrin of approximately 2.5 and 5 pN for neutrophils on

5 and 25 kPa substrates, respectively. These force per integrin estimates are slightly smaller

than values of 13-50 pN obtained by measuring the rupture force of VLA5 to fibronectin

[181, 182]. Assuming a similar rupture force for CR3 (both VLA5 and CR3 bind fibronectin),

then there is a significantly higher capacity for integrin borne force generation than what

we have observed. Indeed, a very small fraction of the neutrophil ensemble was able to

maintain forces as much as 4 to 5 times the mean. Another explanation for the small force

per integrin estimate is that only a sub-population of integrins bind stably and transduce

force [79]. Thus, the force per stably bound integrin would be greater, as would the energy

per integrin.

For neutrophils in suspension with human serum, the amount of active CR3 and total

CR3 expression of BG50 treated neutrophils was approximately 1.26±0.18 times greater

than BG0 neutrophils [79]. If this result can also be shown for neutrophils treated with

both fMLP and BG, then one can rule out a reduction in active integrin as a mechanism

for BG induced decreased mechanical output. Instead, reduction of mechanical output

with increasing BG concentration may be the effect of intracellular factors leading to a

decrease in total myosin contraction, or a disruption of the proteins that bind integrins to

the cytoskeleton (e.g. talin, vinculin, etc.). Regardless of the mechanism, such data would

shed light on the amount of force and energy transduction per integrin, and may set a new

lower bound on the force per integrin required for mechanosensors within the neutrophil to

detect substrate stiffness.2

In addition to experiments probing for differential integrin expression as BG concentra-

tion is ramped up, a causal link between BG and changes in the mechanical output would

2These experiments can be done by flow cytometry, and should be completed shortly.

135

Page 162: The Mechanics of Cell Motility and a Unifying Theory for

be better established by imaging the actin cytoskeleton. The organization of the actin cy-

toskeleton may be suggestive of a mechanism to explain the differences in BG mediated

mechanical output (Fig. 5.2 c), or differences in the total active myosin generated contrac-

tile force determined from the Marcq model (Fig. 5.6 c). It has recently become possible to

introduce primary human neutrophils with a small fluorophore fused peptide called LifeAct

that binds F-actin without inhibiting actin dynamics [183], thus allowing for real time imag-

ing of the actin cytoskeleton. The most informative experiment would be to visualize the

actin cytoskeleton in real time using LifeAct, while doing TFM, at varying concentrations

of BG. However, there is yet to be compelling data showing that LifeAct does not alter

neutrophil motility and function. Therefore, TFM followed by cell fixation and staining

of the actin cytoskeleton with fluorophore fused phalloidin (a compound that binds at the

interface between F-actin subunits and prevents depolymerization [184]) is likely the most

feasible course of action at this time.3

We have also shown numerous correlations between variables characterizing morphology,

motility, and mechanical output. Correlation does not imply causality, it merely implies

the possibility of causality. Some of the correlations and possible causalities are discussed

here. As discussed above, spread area was positively correlated to persistence. We suspect

that cells with greater spread areas were more persistent because of increased integrin

engagement. Cells with greater spread likely require more energetically costly bond rupture

events in order to change direction, as speculated by Oakes et al. who also observed a

positive correlation between spread area and persistence. O’Brien et al., have shown that

persistence of neutrophils crawling on 2D substrates decreases, irrespective of substrate

stiffness and cell spread area, when substrate engagement is restricted to β1 integrins [1].

Conversely, fibronectin used in this study engages both the β1 integrin VLA5 and the β2

integrin CR3. Thus, we can hypothesize that persistence in 2D motility is likely causally

correlated to CR3 ligation that is increased by greater spread area. This hypothesis would

3These simple experiments will be completed shortly.

136

Page 163: The Mechanics of Cell Motility and a Unifying Theory for

be best supported using integrin labeling and adhesion detection from a TIRF microscope.

However, integrin labeling with primary cells such as neutrophils is extremely difficult and

a PAG substrate does not have an optical index of refraction suitable for TIRF microscopy.

This experiment could instead be done with neutrophil-like DMSO differentiated HL-60

transfected with GFP-CR3, plated onto a silicon gel optically corrected for TIRF microscopy

[40].

Other than its dependence on spread area, persistence fraction appears to be indepen-

dent of all other variables, indicating that neither mechanical output nor the addition of

BG affects persistence on soft substrates. However, BG has been shown to improve 2D

neutrophil honing towards weak chemoattractant sources on glass [75]. TFM experiments

have shown evidence of a principal force orientation from perpendicular the direction of

motion to parallel the direction of motion that occurs in chemotactic cells (force orientation

data shown in Chapter 6). This, in contrast to the dipole moment analysis only showing

perpendicular principal force orientation, suggests that chemotactic honing is facilitated by

a mode of mechanical output that does not occur during chemokinesis. Taken together with

the significant effect of BG on mechanical output, it is reasonable to suggest that honing

and mechanical output are linked, and that changes in mechanical output caused by BG are

either the cause of, or the result of, differential honing, thus also explaining why BG appears

to modify neutrophil honing, but not persistence. These observations can be confirmed us-

ing dipole moment analysis to determine the principal orientation of chemotactic forces and

should be combined with directionality time analysis, which gives an unbiased measure of

the extent of honing during directed cell migration [185] (c.f. Chapter 2). Higher order

multipole moments such as quadrupoles have also been shown to be indicative of direction-

ality in dictyostelium discoideum motility [176], and may be informative in the analysis of

neutrophil motility.

137

Page 164: The Mechanics of Cell Motility and a Unifying Theory for

In summary, the chemokinetic contractile forces of neutrophils were localized towards

the posterior of the cell, as reported previously [48, 165]. Dipole moment analysis rigorously

indicated that the principal orientation of these contractile forces were perpendicular to the

direction of motion. This contractile force orientation is similar to that of keratocytes [5, 31],

but differs from that of other eukaryotic cells such as fibroblasts which generate forces at

the leading edge that typically align parallel the direction of motion [32]. We did not find

that motility was optimally dependent on an intermediate range of contractile force, as

suggested by Bangasser et al. [42]. Instead, we found the cell speed scaled inversely with

both contractile force and energy, and that the strength of this correlation was the same on

both 5 and 25 kPa substrates. As also shown previously, neither contractile displacements

nor contractile forces were conserved against substrate stiffness [48]. We show, for the

first time, evidence that contractile energy is a substrate stiffness conserved quantity for

neutrophils. The addition of the biological response modifier BG, to model the role of cell

mechanics in pharmacological intervention, decreased contractile energy, yet energy was

still conserved across substrate stiffness. In cases where BG increased cell motility (25 kPa

substrate), there was also a significant inverse correlation between contractile energy and

cell speed. Taken together with the conserved inverse correlation to cell speed, it appears

that consideration for total contractile energy should be an important component of a cell

motility model.

5. Supplementary Material

5.1. Supplementary Movie Legends.

Movie S1. BG0 neutrophil undergoing chemokinesis on a 5 kPa substrate. Time-

0 corresponds to the addition of fMLP (10 nM). (top-left) Cell migration trajectory and

spread area were tracked from differential contrast images shown here. The white outline is

the perceived cell boundary, and scatter points are centroids calculated from these outlines.

138

Page 165: The Mechanics of Cell Motility and a Unifying Theory for

Scatter point color indicates time. The white arrow originating from the cell centroid shows

the instantaneous velocity of the cell time averaged over an interval enclosing 1 min before

and after the time shown. The instantaneous velocity legend (white arrow, top-left) cor-

responds to 0.125µm/s. (middle-left) Spread area calculated based on the perceived cell

boundaries. (bottom-left) The average of stress island force magnitudes are plotted here

above the total contractile energy. (right) Stress maps of the neutrophil. Color describes

traction stress (color bar shown in Fig. 5.1 c). The white outline is again the perceived cell

boundary. Black arrows originating from the center of stress islands correspond to the stress

island force. The stress island force legend (black arrow, top-left) corresponds to 10 nN.

139

Page 166: The Mechanics of Cell Motility and a Unifying Theory for

Movie S2. BG50 neutrophil undergoing chemokinesis on a 5 kPa substrate.

Time-0 corresponds to the addition of fMLP (10 nM). (top-left) Cell migration trajectory

and spread area were tracked from differential contrast images shown here. The white

outline is the perceived cell boundary, and scatter points are centroids calculated from

these outlines. Scatter point color indicates time. The white arrow originating from the

cell centroid shows the instantaneous velocity of the cell time averaged over an interval

enclosing 1 min on before and after the time shown. The instantaneous velocity legend

(white arrow, top-left) corresponds to 0.125µm/s. (middle-left) Spread area calculated

based on the perceived cell boundaries. (bottom-left) The average of stress island force

magnitudes are plotted here above the total contractile energy. (right) Stress maps of

the neutrophil. Color describes traction stress (color bar shown in Fig. 5.1 c). The white

outline is again the perceived cell boundary. Black arrows originating from the center of

stress islands correspond to the stress island force. The stress island force legend (black

arrow, top-left) corresponds to 10 nN.

5.2. Supplementary Tables and Figures.

5 kPa r 25 kPa A CT E PF MSDA × 0.71 0.87 0.80 0.81

CT 0.75 × 0.68 0.73 0.69E 0.81 0.72 × 0.72 0.76

PF 0.93 0.67 0.78 × 0.83MSD 0.94 0.71 0.78 0.99 ×

Table 5.1. Table of Alternating Conditional Expectations (ACE) val-ues. ACE values range from 0 to 1. The larger the value, the greater theprobability there exists a function (either linear or nonlinear) that describes thedistribution between the two variables. ACE values are tabulated in matrixform to correlate spread area (A), contractility time (CT), contractile energy(E), persistence fraction (PF), and mean squared displacement (MSD). Thebottom-left off-diagonal terms correspond to ACE values on 5 kPa substrates.The top-right off-diagonal terms correspond to ACE values on 25 kPa substrates.This table is meant to be used in relation to the Pearson product-moment cor-relation coefficients graphed in Fig. 5.3.

140

Page 167: The Mechanics of Cell Motility and a Unifying Theory for

DIC DiI (PM)DIC (trace)

a b c

Figure 5.7. The perceived cell boundary may differ from the true cellboundary. The perceived cell boundary of a neutrophil captured by (a) dif-ferential interference contrast microscopy and (b) traced by eye. The trace isbased not only on the current, but also on prior and subsequent images to makea best guess of the perceived cell boundary. (c) The same cell, showing DiIplasma membrane (PM) staining shows that the perceived cell boundary candiffer from the true cell boundary.

a

b

time (min) time (min)

0 10 20 30

BG0

Dex50

BG50 + CR3 block

BG0

Dex50

BG50 + CR3 block

0 10 20 300

20

40

60

80

100

120

con

tra

ctile

fo

rce

(n

N)

0

2

4

6

8

10

12

con

tra

ctile

en

erg

y (

nNµ

m)

25 kPa5 kPa

Figure 5.8. The mechanical output during control experiment for BGassociated change in osmolarity. (a) Contractile energy and (b) contrac-tile force measurements of neutrophils migrating in purified 50µg/mL dextran(Dex50) in citrate buffer show that the osmolarity and sugar concentration ofthe BG50 condition does not account for the changes in mechanical output.The mechanical output was nearly completely blocked when BG50 cells werepretreated with CR3 blocking antibody. n = 20, 15 for Dex50 5 and 25 kPasubstrates, respectively.

141

Page 168: The Mechanics of Cell Motility and a Unifying Theory for

0

50

100

5kPa 25kPa

BG0

5kPa 25kPa

BG50

5kPa 25kPa

BG50+CR3block

percent

P(adhere) P(active|adhere) P(contractile|active)

Figure 5.9. Neutrophil adhesion rate, activity rate, and contractilityrate. The percentage P (adhere) is the percent of neutrophils that contact andadhere to the substrate divided by all cells the make contact with the substrate.The conditional percentage P (active|adhere) is the percent of adhered cells thatappear to be active, meaning they are extending pseudopodia and/or motile(determined by eye). The conditional percentage P (contractile|active) is thepercent of all active cells that apply visible displacements to the substrate. BG0and BG50 refer to 0 and 50µg/mL β-glucan, respectively, and BG50+CR3blockrefers to BG50 cells pretreated with CR3 integrin blocking antibody. These ratesact as internal controls indicating similar neutrophil preparation and proteincoating conditions across experiments. CR3 block treated cells show a significantdecrease in adhesion-, activity-, and contractility rate. Error bars representstandard error of the mean.

142

Page 169: The Mechanics of Cell Motility and a Unifying Theory for

CHAPTER 6

Future Directions and Closing Remarks

143

Page 170: The Mechanics of Cell Motility and a Unifying Theory for

Forward. The contents of this chapter discuss the preliminary data of two ongoing

projects that show strong potential as avenues of future study. The first is an extension

of the findings in Chapter 5 regarding the orientation of traction forces during chemotaxis.

Dipole moment analysis revealed that traction forces during chemokinesis were principally

oriented perpendicular to the direction of motion. In another data set described in this

chapter, we measured the traction forces of directionally migrating neutrophils crawling up

a chemoattractant gradient toward a point source. In agreement with previous findings, the

contractile forces during chemotaxis were localized toward the cell posterior [48, 50, 165].

What has not been well characterized before is the spatiotemporal orientation of these con-

tractile forces. We found that forces during chemotaxis begin small and perpendicular to the

direction of motion. Over time, they build in magnitude and transition (at approximately a

magnitude of 20 nN) from perpendicular to parallel. A model that describes these contrac-

tile force observations is proposed in Sec. 1, and we find that this same model may also be

able to mechanistically account for measurements of directionality time. A computational

algorithm with sample solutions that can be replicated by a beginning to intermediate level

scientific programmer is also contained below.

The second ongoing project address the basis of a reported chemotactic defect of neu-

trophils in patients with sepsis that may be associated with pathological sequela and mortal-

ity. Again, the perpendicular to parallel force transition is observed, but cells isolated from

patients with sepsis also show large isotropic stress distributions. These findings, discussed

briefly in Sec. 2, go hand in hand with our BG finding indicating that biological response

modifiers may act through the modulation of cell mechanics and motility.

This chapter finishes with a short closing remark section that ties together findings

presented in this dissertation and places them in the context of the current state of the

field.

144

Page 171: The Mechanics of Cell Motility and a Unifying Theory for

Receptor

Chemoattractant

Filaments

NucleusPseudopod

Figure 6.1. Schematic showing the chemical signaling of chemotaxisand the downstream cellular components that drive cell migration.This figure has been borrowed from Ref. [186].

1. Future Project 1: Signaling During Early Stage Chemotaxis

1.1. Introduction. Chemotaxis is the process in which a cell aligns its migration path

to the concentration gradient of a specific external chemoattractant. Unlike most prokary-

otic cells which detect concentration gradients over time, eukaryotic cells including neu-

trophils detect chemoattractant gradients spatially over the extent of the cell body. This

detection occurs when chemoattractant molecules (and/or other external chemical cues)

bind to specific receptors on the surface of the cell membrane. Figure 6.1 shows an exam-

ple of chemoattractant (green circles) binding transmembrane receptors (blue horns). In

Chapter 5, the chemoattractant was fMLP and the receptors were formyl peptide receptor 1

(FPR1). Chemical signaling mechanisms downstream of bound receptors lead to processes

such as actin filament polymerization and myosin contraction that drive pseudopod forma-

tion, ultimately leading to cell migration. In the context of a chemoattractant gradient, the

distribution of bound receptors about the cell membrane is asymmetric and this asymmetry

determines the direction of chemotactic transmigration.

The chemotaxis process can be divided into three steps. The first step is quantifying

signal detection. What is the spatial distribution of bound receptors given a distribution of

chemoattractant molecules? Statistical mechanics provides the tools to analytically model

receptor occupation [186]. The second and third steps are signal processing and the imple-

mentation of biomechanical processes that drive cell migration. A good signal processing

145

Page 172: The Mechanics of Cell Motility and a Unifying Theory for

mechanism must be able to separate the average chemoattractant concentration, S, from

the concentration gradient, ∇S, measured over the length of the cell, l. Robust directed

cell migration like that of neutrophils requires that the cell respond over a broad range of

average chemoattractant concentrations, S. Essentially, the cell needs to adapt to different

average chemoattractant concentrations. Consider a test model of a cell with motility ma-

chinery coupled to the extracellular chemoattractant concentration by a factor of ξ ∝ l∇SS

.

In a linear gradient, the motility machinery would change its state as the cell migrates and

S increases (ξ decreases). Using this test model, the conditions under which chemotaxis

occurs would be severely restricted. Therefore, this test model is not suitable for describing

chemotaxis under physiological conditions. A better model of chemotaxis signal processing

leaves a large range of S decoupled from the motility machinery. If ξ is completely decoupled

from S, it is called perfect adaptation.

Several mathematical models exist for predicting the distribution of signaling proteins

associated with cell migration. Most of these models achieve something close to perfect

adaptation using a design principle called local excitation and global inhibitor (LEGI).

A LEGI model can be described in terms of four chemical components, the signal, two

factors turned on by the signal that activate and inhibit a downstream response, and a

downstream response that activates the machinery of cell motility. The activator is locally

confined and fast acting whereas the inhibitor is diffusible and slow acting. Each component

effective represents signaling proteins and other compounds associated with chemotaxis in

cells. Because the response is what drives the motility machinery, it is the response that

must perfectly adapt to the chemoattractant signal. To achieve perfect adaptation, the

signal up-regulates the fast activator, which causes the response to increase. The signal

also up-regulates the slow inhibitor, which eventually causes the response to relax back to

its original equilibrium value. The magnitude and timescale of this transitory response spike

depends on the reaction rates between the catalytic reactions that regulate the response.

146

Page 173: The Mechanics of Cell Motility and a Unifying Theory for

However, at steady state, the response does not depend on average signal concentration

across the cell, S. Thus perfect adaptation is achieved. We hypothesize that the duration

of the transitory response spike in reaction to a change in chemotactic cue can be used to

mechanistically model results found using our directional motility metric.

As amoeboid cells migrate, they apply contractile forces to their substrate. In our recent

experiments using traction force microscopy to measure the forces applied by neutrophils

during chemotaxis, we observed an initial period of small contractile forces that were prin-

cipally oriented perpendicular to the direction of migration. After a period of one to two

minutes during which time the magnitude of contractile forces increased, the principal force

orientation transitioned from a perpendicular mode to a parallel mode (Fig. 6.2). Although

the average scalar traction and RMS contractile force of chemotactic neutrophils have been

characterized in the past [48, 50, 165], the orientational dynamics of these force modes has

yet to be studied. Due to the timing and transient nature of the perpendicular force mode

as we observe it, and the premise that cells are designed to be adaptive to chemotactic sig-

nals [187], it is plausible that our observations can also be explained by a process featuring

perfect adaptation. With this in mind, we solve two LEGI models below in order to study

the nature of the transient response factor. In neutrophils, the elements of several LEGI

models have been suggested to correspond to particular signaling proteins such as phospho-

inositide 3-kinase (PI3K), and small G-protein [188]. A solution to this model provides both

an experimentally testable mechanism for the perpendicular to parallel force transition, and

should give mechanistic insights into why contractile force exertions are significantly altered

from early to later stages of cell adhesion [189].

1.2. Computational Model.

1.2.1. Reaction Diffusion Equations. There are several LEGI models in existence but

here we use one chosen one by Levchenko and Iglesias [190]. The model is described as

follows: A local activator of concentration A found along the cell membrane catalyzes the

147

Page 174: The Mechanics of Cell Motility and a Unifying Theory for

1 min 2 min 3 min 4 min

a

b

c

d

0 100 200 300 400 500traction (Pa)

0 1 2 3 402460

20

40

60

0

1

time (min)

en

erg

y

(pJ)

forc

e

(nN

)

dire

ctio

na

lity

ind

ex [γ]

Figure 6.2. Classification of neutrophil traction stress fields during 2Dchemotaxis. (a) Traction field time-lapse of a chemotactic neutrophil in oneminute increments. White lines indicate the cell determined from correspond-ing DIC images (not shown here). Black dashed lines outline traction islands.Black vectors represent the forces calculated by integrating over the local trac-tion island region. Two modes are used to describe the orientation of islandforces: perpendicular (⊥) and parallel (‖). Perpendicular and parallel modessignify traction forces oriented nearly perpendicular or parallel to direction ofmotion, respectively. (b) Average traction island force and directionality indexplotted against time. After a period of low force output from 0 to 1 minute,island forces increase in magnitude and the traction force mode transitions fromperpendicular to parallel by 2 minutes. (c) Work done on the substrate plot-ted against time (neglecting energy exerted to stress the substrate in the 3rddimension). (d) Schematic showing the general force mode progression (newunpublished result). Scale bars show 10µm.

conversion of an inactive response protein, R, to an active response protein, R∗ (Fig. 6.3).

The presence of active response protein leads to cell locomotion in a manner not described

here. The active response protein can also return to its inactive state as catalyzed by a

globally diffusing inhibitor, I. Below is a brief derivation of reaction kinematics similar to

the derivation in Ref. [190]. First consider the enzymatic reactions,

R+Akc1ku1

U1ka1⇀ R∗ (6.1)

148

Page 175: The Mechanics of Cell Motility and a Unifying Theory for

R∗ + Ikc2ku2

U2ka2⇀ R. (6.2)

Corresponding reaction rates are derived using the law of mass action as follows,

dR

dt= −kc1RA+ ku1U1 + ka2U2 (6.3)

dR∗

dt= −kc2R∗I + ku2U2 + ka1U1 (6.4)

dU1

dt= −kc2R∗I + ku2U2 + ka1U1 (6.5)

dU2

dt= −kc2R∗I + ku2U2 + ka1U1 (6.6)

where U1 and U2 are substrate-enzyme reaction intermediates and the k’s are rate constants

[191]. A more general treatment for the reaction rate equations 6.3-6.6 imposes Hill functions

to account for the total free enzyme concentration, A and I. Here, we have simplified the

reaction equations by assuming that the fraction of bound enzyme is very small compared

to the amount of total enzyme. Assuming a quasi-steady state solution for the reaction

intermediates, Eqn. 6.4 simplifies to,

dR∗

dt= −k−RIR∗ + kRAR. (6.7)

Applying similar logic (see Ref. [190]), the reaction rates for activator A and inhibitor I

are,

dA

dt= −k−AA+ kAS (6.8)

dI

dt= −k−II + kIS. (6.9)

Finally, Eqns. 6.7-6.9 are simplified by non-dimensionalization and scaling using the fol-

lowing parameters: τ ≡ k−At, a ≡ (kR/k−A)A, i ≡ (kRkAk−I/kIk2A)I, r ≡ R∗/Rtot,

α ≡ k−I/k−A, and β ≡ (kIk−Rk−A/k−IkRkA). The resulting equations define the reaction

rates of the response protein fraction r and scaled enzyme concentration a and i localized

149

Page 176: The Mechanics of Cell Motility and a Unifying Theory for

to the cell membrane at location m:

da(m)

dτ= −(a− s) (6.10)

di(m)

dτ= −α(i− s) (6.11)

dr(m)

dτ= −βir + a(1− r). (6.12)

Allowing for global diffusion of inhibitor i, Eqn. 6.11 is a boundary condition for the global

diffusion equation,

∂i(~x, t)

∂τ=

D

k−A∇i(~x, τ) (6.13)

where D is the diffusion coefficient of the diffusible inhibitor in cytosol, and ~x is the 2-D

spatial coordinate within the cell.

I

S S(t)

R*/Rtot(t)

t

t

R

R*

A

a b

Figure 6.3. The LEGI model. (a) Chemical reaction circuit diagram. Thesignal, S, activates a response activator, A, and a response inhibitor, I, whichcatalyze the forward and reverse reactions between an inactive and active re-sponse protein, R and R∗, respectively. In this model, each element is confinedto the cell membrane except the inhibitor, I, which can diffuse through thecellular cytosol. Active response factor, R∗, up-regulates cell motility. (b) Per-fect Adaptation. In this LEGI model, when the reaction rates for the reactionsS ⇀ A and S ⇀ I have similar form in their dependence on signal concen-tration S, the equilibrium fraction of active response does not change when aspatially homogeneous signal is altered. In response to the signal change, thereis a transient increases in active response protein. The magnitude and durationof this response factor spike depend on the ratio between inhibitor productionand activator production. Colors: cyan indicates low signal concentration, redindicates high signal concentration.

150

Page 177: The Mechanics of Cell Motility and a Unifying Theory for

1.2.2. Finite Difference Equations - Cartesian Coordinates. The model can be solved

on a disk of radius P in Cartesian coordinates (x, y) using the forward time centered space

(FTCS) method. The coordinate system is discretized using xp = p∆x and yq = q∆x, both

with dimensions of µm and p, q = −P∆x ,

−P∆x + 1, ..., 0, ..., P∆x − 1, P∆x such that the domain is

a square with side length 2P . Time is discretized as τn = n∆τ , with dimensions of k−1−A.

In this coordinate system, the diffusion equation is approximated by the finite difference

equation,

i(xp, yq, τn+1) = i(xp, yq, τn) (1− 4ζ) +

+ζ [i(xp+1, yq, τn) + i(xp−1, yq, τn) + i(xp, yq+1, τn) + i(xp, yq−1, τn)](6.14)

where ζ ≡ D∆τk−A(∆x)2 .1 The initial condition for the inhibitor is zero everywhere on the

domain. The disk boundary defined by the circle p2 + q2 = P 2/(∆x)2. In discretized space,

a list of boundary coordinates is generated by populating the set

{p, q} ∈ round(√

p2 + q2)

=P

∆x(6.15)

This is the boundary where differential Eqns. 6.10-6.12 are applied (using 2nd order Runge

Kutta algorithm). Equation. 6.14 is solved everywhere on the domain before i is updated

at the boundary using a Runge Kutta step based on Eqn. 6.11.

1.2.3. Finite Difference Equations - Polar Coordinates. The model can be solved on a

disk of radius P using the FTCS method. The disk is discretized in polar coordinates (ρ, θ)

such that ρx = x∆ρ and θy = y∆θ, with dimensions of µm and degrees ( ◦) respectively.

Time is discretized again as τn = n∆τ , with dimensions of k−1−A. In these polar coordinates,

1As a side note, using ζ = 14 gives the Jacobi iterative method for solving the Poisson equation (i.e.

D = 1, ∆τ = 0.01, and ∆x = 0.2).

151

Page 178: The Mechanics of Cell Motility and a Unifying Theory for

the diffusion equation for inhibitor i is

∂i(ρ, θ, τ)

∂τ=

D

k−A

[∂2i

∂ρ2+

1

ρ

∂i

∂ρ+

1

ρ2

∂2i

∂θ2

]. (6.16)

Using the definition ζ ≡ D∆τk−A(∆ρ)2 , Eqn. 6.16 corresponds to the following FTCS finite

difference equation,

i(ρx, θy, τn+1) = i(ρx, θy, τn)

[1− 2ζ − 2ζ

(x∆θ)2

]+

+ [i(ρx−1, θy, τn) + i(ρx+1, θy, τn)] (ζ +ζ

2x) +

+[i(ρx, θ(y−1), τn) + i(ρx, θ(y+1), τn)

] ζ

(x∆θ)2. (6.17)

valid for x = 1, 2, 3, ..., P−∆ρ∆ρ , y = 0, 1, 2, ..., 360◦−∆θ

∆θ , and n = 1, 2, 3, ...,τf∆τ . Notice that the

center of the disk must be treated carefully due to the 1ρ∂i∂ρ and 1

ρ2∂2i∂θ2 terms in Eqn. 6.16.

While ∂2i∂θ2 is identically zero, 1

ρ∂i∂ρ must be bound in the limit that ρ → 0. So long as any

mathematical source of i is bound throughout the disk of radius P , the following boundary

condition can be derived at the pole ρ = 0 using arguments found in Ref. [192],

∂i

∂ρ(0+, θ, τ) = − ∂i

∂ρ(0−, θ, τ) (6.18)

which corresponds to the following finite difference equation,

I(0, θ, τn+1) =1

2

[i(∆ρ, 0, τn+1) + i(∆ρ,

180◦

∆θ, τn+1)

](6.19)

In [186], a Neumann boundary condition for inhibitor i is cited in lieu of the boundary

condition in Eqn. 6.11. That condition is

D

P

∂i(m)

∂n= kis(m)− k−ii(m) (6.20)

152

Page 179: The Mechanics of Cell Motility and a Unifying Theory for

where n represents the direction normal to the outer boundary (cell membrane). The

corresponding finite difference equation is

i(P, θy, τn+1) = i(P −∆ρ, θy, τn+1) +τ1ki

1 + τ1k−is(P, θy, τn+1) (6.21)

using the definition τ1 ≡ P∆ρ/D. The diffusion equation is solved above using polar

coordinates instead of Cartesian coordinates in order to simplify the Neumann boundary

condition. Assuming that diffusion of the inhibitor occurs along the membrane, one can

plug the Neumann boundary condition, Eqn. 6.21, into the diffusion reaction equation to

get an initial value problem with tangential diffusion at the boundary. Using this equation,

a source term for i can be defined, in addition to a continuity equation that accounts for

inhibitor flux at the boundary. This avenue of mathematical analysis of the model is not

pursued in this work since my derivation of the model could not recapitulate the Neumann

boundary condition in Eqn. 6.20.

1.3. Results.

1.3.1. Cartesian Coordinate Model. Solutions to linear algebraic equation 6.14 were it-

erated and 2nd order Runge Kutta was used to solve Eqns. 6.10-6.12 along the circular

boundary. The solutions for a, i, and r, solved using Runge-Kutta at FTCS are plotted

against time in Fig. 6.4. Here, a periodic signal, s(θ) = cos2( 2πθ720◦ ), is initially turned on at

τ = 0.5. Later, the signal is made homogeneous and increased in two steps at τ = 25 and

τ = 50. During the period signal, the response reaches a positive steady state along the

entire membrane except where these is no signal. As the signal increases further, perfect

adaptation is observed. The duration of the transient response before steady state varies

inversely with k−I/k−A (not shown here). Parameters used to generate this solution are

found in the figure caption.

1.3.2. Polar Coordinate Model. Solutions to linear algebraic equation 6.17 were iter-

ated. Along the cell membrane, ρx = P , the initial value problems for activator a(θ, τ)

153

Page 180: The Mechanics of Cell Motility and a Unifying Theory for

time (1/k-A) time (1/k-A)

con

cen

tra

tio

n (

AU

)

inhibitor

an

gle

) /

de

ga

ng

le (θ

) /

de

ga

ng

le (θ

) /

de

g

activator

response

c

d

e

b

a

P

θ

90270

180

0

Pyq

xp

-P-P

-P

Figure 6.4. LEGI model results using Cartesian coordinate system.Modeling Dynamics for inputs of P = 10µm, kA = 10.0 s−1, k−A = 10.0 s−1,kI = 1.0 s−1, k−I = 1.0 s−1, kR = 1.0 s−1, k−R = 5.0 s−1, D = 10.0µm2 s−1,s(θ, τ < 0.5) = 0, s(θ, 0.5 ≤ τ < 25.) = cos2

(2πθ720◦

), s(θ, 25. ≤ τ < 50.) =

2, s(θ, 50. ≤ τ) = 2.5, and all initial conditions are zeroed. The integrationparameters are ∆x = 0.5µm and ∆τ = 0.01. (a) Schematic of the 3-D (2-space,1-time) model cell in discretized Cartesian coordinate. The coordinate systemshown here is left handed to be consistent with Python matrices (this should bechanged in any future work). (b) Time course of the concentrations of signal,activator, inhibitor, and response protein at ρ = P (membrane), and θ = 0◦.After an initial steady state response is established after the onset of signal(τ = 20−25), perfect adaptation is maintained. (c), (d), and (e) The activator,inhibitor, and response along the membrane are plotted, a(θ, τ), i(ρ = P, θ, τ)and r(θ, τ) respectively. Color represents concentration in arbitrary units.

and response r(θ, τ) (Eqns. 6.10 and 6.11) were solved using 2nd order Runge-Kutta and

Euler methods, respectively. Figure 6.5 shows a solution to the model in which a signal,

s(θ) = cos2( 2πθ720◦ ), is turned on at τ = 0.5 (dashed blue line). Following the onset of the

signal, the activator exponentially relaxes toward its steady state. The response variable

transiently increases before relaxing back to zero. While this is indicative of perfect adap-

tation, this adaptation is achieved by the means described in section 1.2.1. Instead, the

154

Page 181: The Mechanics of Cell Motility and a Unifying Theory for

inhibitor dynamically blows up to infinity after the diffusive inhibitor reaches the origin,

and adaptation is achieved by inhibitor saturation (Fig. 6.5, b and d, and Movie 2). The

inhibitor saturation is the result of computationally solving the diffusion equation in polar

coordinates and not realistic. Interestingly though, inhibitor saturation is a perfectly fair

means of decoupling the response from the average signal concentration. Unfortunately, an

inhibitor saturated cell also loses the ability to detect the signal gradient.

In an attempt to “fix” the inhibitor instability at the origin (see Movie 2), the diffusion

operator was redefined at the origin two different ways. First, a no flux condition was

applied, ∂i∂ρ |ρ=0 = 0. Second, the treatment in Eqn. 6.19 was applied. Both treatments lead

to the same instability.2

1.4. Discussion. In this work, the LEGI model showed that adaptations to changes in

chemoattractant signal could be achieved by two mechanisms. First (Section 1.3.1), a finite

valued concentration of inhibitor competing against an activator led to perfect adaptation.

Second (Section 1.3.2), inhibitor saturation gave perfect adaptation because the response

was completely cut off from the signal. The latter mechanism was generated by accident

because of a computational pole instability that should be fixed by anyone continuing this

project.

Hidden in the way the Cartesian coordinate model was presented above, the diffusion

equation was not coupled to the boundary. While the inhibitor from the membrane was

allowed to diffuse into the bulk, bulk inhibitor did not interact with membrane inhibitor.

A more realistic model would couple the globally diffusing inhibitor to the membrane using

a no flux boundary condition,

∂i

∂n= −α(i− s). (6.22)

2An applied mathematician suggested the boundary condition ∂i∂t = 2D

[∂2i∂ρ2 + 1

ρ2∂2i∂θ2

]at the origin.

This may fix the r = 0 singularity.

155

Page 182: The Mechanics of Cell Motility and a Unifying Theory for

b

con

cen

tra

tio

n (

AU

)

time (1/k-A)

inhibitor

time (1/k-A)

an

gle

, θ (

de

g)

an

gle

, θ (

de

g)

an

gle

, θ (

de

g)

activator

response

c

d

e

a

ρx

P

θy

90270

180

0

Figure 6.5. LEGI model results using polar coordinate system. Mod-eling Dynamics for inputs of P = 10µm, kA = 1.0 s−1, k−A = 10.0 s−1,kI = 1.0 s−1, k−I = 1.0 s−1, kR = 1.0 s−1, k−R = 1.0 s−1, l = 1.0µm,D = 1.0µm2 s−1, s(θ, τ < 0.5) = 0, s(θ, τ ≥ 0.5) = cos2

(2πθ720

), and all initial

conditions to zero. The integration parameters are ∆ρ = 0.5µm, ∆θ = 1.0 ◦,and ∆τ = 0.01. (a) Schematic of the 3-D (2-space, 1-time) model cell in dis-cretized polar coordinates. (b) Time course of the concentrations of signal,activator, inhibitor, and response protein at ρ = P (membrane), and θ = 25 ◦.Though the system of partial differential equations is well-posed, the system isunstable in i the way it is currently coded, after the signal is turned on, i blowsup to infinity. Nonetheless, the response has the expected form after the signalis turned on, initially increasing before relaxing back to zero. This is adaptationby inhibitor saturation. (c), (d), and (e) The activator, inhibitor, and responsealong the membrane are plotted, a(θ, τ), i(ρ = P, θ, τ) and r(θ, τ) respectively.Color represents concentration in arbitrary units. All color bars saturate at aconcentration of 1 AU. The dashed white lines correspond to the plots in (a).

Here the right hand side is the source, and n represents the direction normal to the boundary.

When the source at the boundary is zero, the inhibitor flux, u = −Dn · ∇i = 0, as desired.

Therefore, the Neumann boundary condition presented previously [186] is correct. Taken

together, the Cartesian coordinates model solved above (Fig. 6.4) is valid but the diffusion of

156

Page 183: The Mechanics of Cell Motility and a Unifying Theory for

the inhibitor did not play a role in the response. In the polar coordinates model (Fig. 6.5),

inhibitor diffusion was properly coupled to the membrane but this model had a pesky

computational instability at the pole that needs to be fixed.

In the future, one should also consider the mechanics of the cell membrane. It has been

proposed [186] that membrane speed, vn, couple to the LEGI model as follows:

dvndt

= f(R)− γκ− λvn. (6.23)

Here, f(R) is some function of the response protein, R, described by Eqn. 6.11, γ is the

membrane bending rigidity, κ is the membrane curvature, and λ is a frictional coefficient

[193]. More generally, any number of terms can be added to Eqn. 6.23 to capture the physics

of the system (See Ref. [126] for examples).

Implementation of the LEGI model will be highly useful because it can lead to an

experimental design for testing the signaling mechanism underlying the recently observed

orientation transitions in chemotactic neutrophils. Furthermore, the adaptation time of the

response in this model may make it possible to mechanistically explain measurements of

directionality time [1, 185] (c.f. Chapter 2). One of the data sets we have collected involves

neutrophil chemotaxis on soft substrates with the inhibition of motility activator PI3K. The

LEGI model is ideally suited for modeling the findings made from this data set.

1.5. Pseudocode. The following pseudocode is intended to help a beginning to inter-

mediate computational physicist solve the LEGI model. Sample code written in Python is

also available online (see Chapter 7, Sec. 1 for a link to my github account).

• preamble, import

– math, numpy, scipy, matplotlib

– progressbar (so I can see how efficiently the program is running)

• initialize BVP object

– declare variables

157

Page 184: The Mechanics of Cell Motility and a Unifying Theory for

– set default parameters

– set default initial conditions

– set boundary location equation (I set the boundary to the circumference of a

disk in Cartesian coordinates)

• Euler forward step and initial value problem iterations (i.e. Eqns. 6.14, 6.17,etc)

until time t reaches end-time tf

– diffusion equation solved using C++ code embedded into Python via the

scipy.weave package (originally coded in Python without C++ but the pro-

gram took way too long to complete)

– update order for each iteration

(1) update a by 2nd order Runge Kutta (RK)

(2) update i by FTCS everywhere

(3) update i on the boundary (membrane) by 2nd order RK (Linear update

in polar coordinate code)

(4) update r by 2nd order RK (Euler method update used in polar coor-

dinate mode)

(5) update τ = τ + ∆τ

• output a progress bar during iteration algorithms (C++ about 45-50 times faster

than straight Python)

• if necessary, convert i from polar coordinates to Cartesian coordinates and make

a colorplot movie of i(2-D space) vs. time

• save files

• other plots

158

Page 185: The Mechanics of Cell Motility and a Unifying Theory for

2. Future Project 2: Neutrophils From Patients With Sepsis Show Novel

Stress Distributions During Chemotaxis

2.1. Introduction. Gram negative bacterial sepsis is caused in part by an overabun-

dance of lipopolysaccharide (LPS), an endotoxin found in the outer cell membrane of Gram

negative bacteria [194]. LPS activates several immune system pathways including the pro-

duction of proinflammatory cytokines that elicits neutrophil diapedesis (Fig. 6.6, lower left).

In this manner, high levels of LPS from a Gram negative bacterial infection can lead to a

systemic inflammatory response, known as sepsis. In this case, tissue damage and systemic

organ failure resulting from excessive neutrophil activity is often more harmful than the

initial infectious insult [195].

Several treatments focusing on the blockage of LPS induced extracellular cues that

mediate neutrophil activity have been clinically unsuccessful [195, 196], and the mortality

rate of patients with sepsis remains an issue [197, 198]. New treatments for sepsis are

necessary and the effect of sepsis on the neutrophil itself may elucidate new treatment

possibilities. It has been reported that neutrophil chemotaxis is impaired in patients with

sepsis, and that this impairment correlates to expected rate of survival. In the following

section, we analyze the migration paths of chemotactic human neutrophils from both healthy

donors and patients with sepsis. To better understand the differences in migration and

intracellular signaling between the two populations, the cellular traction forces were also

measured using traction force microscopy (All relevant Materials and Methods can be found

in Chapter 5, all substrates are 5 kPa in stiffness).

2.2. Results. In the previous chapter, we showed that the traction forces of chemoki-

netic neutrophils were localized at the uropod and oriented primarily perpendicular the

direction of motion during motility. This was no surprise because traction forces at the uro-

pod of the neutrophil perpendicular to the direction of motion had been reported previously

[50, 165]. With chemotactic neutrophils, we found that overtime from the initial onset of

159

Page 186: The Mechanics of Cell Motility and a Unifying Theory for

lipopolysaccharide (LPS)neutrophil chemotaxis

to sources of infection

Figure 6.6. Neutrophil activation, endothelium transmigration, andchemotaxis through the extracellular matrix. Lipopolysaccharide in-duces proinflammatory cytokine production that leads to neutrophil activationand transmigration through the endothelium. Once in the extracellular matrix,neutrophils migrate by chemotaxis to sites of infection and injury. This figurewas adapted from Ref. [199].

a chemotactic signal, the magnitude of contractile force increased to a point where a force

orientation transitioned from a perpendicular mode to a parallel mode. An example of this

phenomenon is shown in the TFM time lapse depicted in Fig. 6.7 a (perpendicular force

mode denoted by ⊥, parallel force mode denoted by ‖). Here, the contractile force magni-

tude increased until it peaks near 25 nN, at which time the force orientation transitioned

from perpendicular to parallel (Fig. 6.7 b, blue data points). Directionality index, defined

as the average cosine between stress island forces and direction of motion, is also plotted

to quantify the force directionality.3 Directionality index transitioned from an intermediate

value between 0 and 1, to 1 at approximately 2 min after the initial onset of chemoattractant

signal (Fig. 6.7 b, green data points). The contractile energy is shown in Fig. 6.7 c, and a

schematic of the force orientation transition is shown in Fig. 6.7 d.

3In the future, this should be reanalyzed using dipole moment analysis instead of directionality indexto obtain an unbiased measure of the principal force orientation.

160

Page 187: The Mechanics of Cell Motility and a Unifying Theory for

1 min 4 min 7 min 10 min

1 min 2 min 3 min 4 min

Healthy

Septic

{ {

a

b

c

e

f

g

i

h

d

0 100 200 300 400 500

traction (Pa)

cell outline

stress island and

force vectorforce

directionalityindex

en

erg

y

(nN

. µm

)fo

rce

(n

N)

dire

ctio

na

lity

ind

ex [γ]

time (min)0 2 4 6 8 10 12

0

5

0

20

40

60

0

1

en

erg

y

(nN

. µm

)fo

rce

(n

N)

dire

ctio

na

lity

ind

ex [γ]

time (min)0 2 4 6 8 10 12

0

5

0

20

40

60

0

1

isotropic

Figure 6.7. Differential Traction Force Mapping of representative neu-trophils from healthy (H) and septic (S) donors. Healthy donor: (a)Traction field time-lapse of a migrating healthy cell in 1 min increments. Whitelines indicate the cell outline determined from corresponding DIC images (notshown). Black dashed lines outline stress islands. Black vectors represent theforces corresponding to local traction islands. Orientation of island forces areperpendicular (⊥) or parallel (‖) relative to direction of motion. (caption con-tinued on pg. 162)

161

Page 188: The Mechanics of Cell Motility and a Unifying Theory for

In chemotactic neutrophils from patients with sepsis, contractile forces again begin ori-

ented perpendicular to the direction of motion and slowly increased in magnitude. After an

initial period in the perpendicular mode, the force orientation transitioned to a mode de-

scribed as a cycling between the parallel mode and an additional isotropic force mode (6.7 e,

isotropic mode denoted �). Because neutrophils effectively produce zero net force (and

torque) on their substrate during migration [176] (c.f. Chapter 5, Sec.refsec:mechTFM),

measurements of the stress island force magnitude during periods of the isotropic mode are

not applicable (Fig. 6.7 f, grey shaded periods). Beyond the differential force mode, the

contractile energies of neutrophils from septic donors were significantly greater than the

contractile energies of neutrophils from healthy donors (Fig. 6.7, g compared to c). Neu-

trophils from septic donors also spread more, migrated slower, and honed less effectively

than neutrophils from healthy donors (Fig. 6.8).

2.3. Discussion. Combining the chemotaxis results shown here with the chemokinesis

results from Chapter 5, one can describe the cell mechanics of neutrophil motility as follows

(Fig. 6.9): First, cells apply contractile forces even when they are not motile. When motile

under chemokinetic conditions, the contractile forces are remarkably smaller compared to

cells that contract but are not motile. Motile neutrophils show forces localized at the uro-

pod as reported previously [48, 50, 165], and these forces are perpendicular to the direction

(Fig. 6.7 caption continued) (b) Average traction island force and directionality index plotted againsttime. After a period of low force output from 0 to 1 min, island forces increase in magnitude whilethe directionality of the traction force mode transitions from perpendicular to parallel after 2 min.(c) Energy exerted by a migrating neutrophil from a healthy donor. (d) Schematic showing theforce mode transitions from perpendicular to parallel direction. Septic Donor: (e) Traction fieldtime-lapse of a migrating neutrophil from a septic donor shown in 4 min increments. Markings areas in (a), with the addition of another force mode, isotropic (�). The isotropic mode correspondsto a single donut shaped traction island with negligible net force. (f) Forces oscillate and tend tozero during the isotropic mode. (g) The energy exerted by a migrating neutrophil from a septicdonor is greater than healthy cell. (h) Schematic showing the general force mode progression ofa neutrophil from a septic donor. Perpendicular force mode abates and the cell cycles betweenparallel and isotropic force exertion modes. (i) Legend including traction field color bar. Scale barsare 10µm. Similar findings were found from 11 neutrophils collected from 3 patients with sepsis,and 9 cells from 3 healthy donors.

162

Page 189: The Mechanics of Cell Motility and a Unifying Theory for

are

a [

A]

(µm

2)

Spread Area

Migration Speed

Mean Square Displacement

rati

o [

x/y

]Aspect Ratio

Directionality Time

spe

ed

[Ι∆dΙ/∆

t]

(µm

/min

)ti

me

[t p

] (m

in)

xy

A

∆d

t

t+∆t

time (min)

MS

D

exp

on

en

t

0 5 100123

td

0

2

4

6

1

1.5

2

100

200

300

5

10

H S

H S

a

d

e

b

c

20 40 80 160

101

102

103

SH

squ

are

dis

pla

cem

en

t

[<d(t,∆t)2>

t] (µ

m2)

time increment [∆t] (s)

t

t+∆t

t+2∆t

Figure 6.8. Neutrophils fromhealthy (H) and septic (S)donors differ in morphol-ogy and migration speed,and honing. (a) S-neutrophilsspread more than H-neutrophils.(b) H-neutrophils trend towardsa greater range of aspect ra-tios than S-neutrophils. (c) and(d) S-neutrophils move slowerthan H-neutrophils. (e) Thedirectionality time was greaterfor S-neutrophils compared to H-neutrophils, indicating less hon-ing efficiency by S-neutrophils.n=11 neutrophils from 3 donorswith sepsis; 9 cells from 3 healthydonors.

of motion under chemokinetic conditions. Also reported previously, the spatial force profile

undergoes significant reorientation prior to sharp changes in direction [165]. Not reported

previously, dipole moment analysis quantitatively demonstrates that the principal force ori-

entation is perpendicular to the direction of motion during chemokinesis. Further, changes

163

Page 190: The Mechanics of Cell Motility and a Unifying Theory for

in force orientation lag behind changes in motility direction during smooth turns, in con-

trast to changes in force that occur prior to sharp turns. We present arguments suggesting

that persistence is likely dependent on spread area and not mechanical output.

With neutrophils undergoing chemotaxis, the early spatio-orientational distribution of

the contractile forces resembles that of chemokinetic cells. Initially, forces are small and

perpendicular to the direction of motion. However, the similarities between chemokinetic

mechanical output and chemotactic mechanical output only last on the order of minutes.

Over time from the onset of the chemotactic gradient, contractile forces are observed to

increase and a transition from perpendicular to parallel force orientation occurs. Perhaps

the mechanical output of a chemokinetic cell would resemble the mechanical output of a

chemotactic cell if the chemokinetic cell was highly persistent and moved in one direction

long enough for the adhesion centers to mature. With neutrophils from septic donors, a

parallel force mode was also observed following an initial perpendicular force mode. But

neutrophils from septic donors also displayed an isotropic force mode characterized by rel-

atively large contractile energies. A significant difference in honing efficiency was observed

between neutrophils from septic and healthy donors. Based on our observation that parallel

force modes are associated with chemotaxis, but not chemokinesis, and we hypothesize that

honing, not persistence, is highly dependent on the strength, distribution, and orientation

of contractile forces.

3. Closing Remarks

Cell motility is required for many important processes in nature including inflammation,

wound healing, sexual reproduction, embryonic development, cancer metastasis, and more.

Determining the mechanisms that underlie cell motility is important for advancing the

frontier of medicine, but this is not a simple task. The cell is an extremely complex system

with cell motility being a spontaneous emergent behavior that is the culmination of many

stochastic molecular processes occurring within this system. Further complicating matters,

164

Page 191: The Mechanics of Cell Motility and a Unifying Theory for

force modes

changing direction

smooth turns sharp turnsnon motile

chemokinetic

(with and without BG)

chemotactic

(healthy)

chemotactic

(septic) { {cell outline

migration trajectory

stress island and force vector

force modes

Figure 6.9. Schematics summarizing all the force modes of motile andnon motile neutrophils. Symbols ∴, ⊥, ‖, and � indicate anomalous forceorientations (anomalous mode), orientations perpendicular to the direction ofmotion (perpendicular mode), orientations parallel to the direction of motion(parallel mode), and isotropic orientation (isotropic mode), respectively. Inchemokinesis, when a neutrophil makes a smooth turn, force reorientation lagsbehind changes in direction. In a sharp turn, force orientation is lost prior tothe sharp turn before being re-established.

motility also couples to extracellular cues. To gain a full understanding of the processes

underlying cell motility would require the tracking of millions of molecular reactions between

an incredibly large number of signaling proteins. The experimental apparatus to fully

elucidate a system as complex as the cell simply does not exist at this time.

On a whole, this dissertation is a combination of experimental analysis of cell motility

and the implementation of tools from classical physics and random walk theory to create

models that predict and/or characterize the motile behaviors of cells. The models we

165

Page 192: The Mechanics of Cell Motility and a Unifying Theory for

presented were simple in nature, reducing the complex cell to a small set of elements that

capture the essential biomechanical and biochemical factors that affect cell motility.

Cellular motion has been routinely characterized by parameters such as speed, tor-

tuosity, and persistence time. However, most of these parameters are not reproducible

because their numerical values depend on technical parameters like sampling interval and

measurement error. In Part 1, Chapter 2 of this dissertation, we addressed the need for a

reproducible metric that did not depend on technical parameters by analytically deriving a

novel metric called directionality time. Directionality time was interpreted as the minimum

observation time required to identify motion as directionally biased, or similarly, the time

required for a cell to adapt to and hone towards an external cue. Measured based on fit-

ting mean squared displacement in log-log coordinates, we showed that the corresponding

fit function was approximately model invariant and applicable to a variety of directionally

biased motions, including processes that were nonergodic. Simulations were used to further

show the robustness of directionality time measurements and their decoupling from mea-

surement errors. Finally, we demonstrated, step-by-step, how to measure the directionality

time of noisy, nonergodic experimental data. Directionality time should have broad appli-

cability across many fields of inquiry as a robust metric for characterizing the motions of

single particles, cells, animals, and other motile systems.

In Chapter 3 of this dissertation, the focus then switched to using the directionality

time metric to characterize human neutrophil chemotaxis. Neutrophil chemotaxis is well

understood to occur in response to biochemical stimuli such as chemoattractants binding

cognate receptors and leading to directional motility. We showed the physical properties of

the underlying substrate also contribute significantly to the regulation of human neutrophil

chemotaxis by characterizing directed migration toward fMLP on fibrinogen (Fgn), type IV

collagen (Col-IV), and fibronectin (Fn)-coated gels of varying stiffness. Using directionality

time together with MSD to quantify migration speed, we found that mechanoregulation

166

Page 193: The Mechanics of Cell Motility and a Unifying Theory for

of migration varied significantly by ligand-coating. We showed that neutrophils on the β1

integrin ligand Col-IV demonstrated decreasing migration speed but no change in direc-

tionality time as substrate stiffness increased. In contrast, neutrophils on the β2 integrin

ligand Fgn demonstrated no change in migration speed but decreased directionality time as

substrate stiffness increased. For neutrophils on Fn, recognized by both β1 and β2 integrins,

both speed and directionality time decreased with increasing substrate stiffness leading us

to hypothesize that β1 and β2 integrin engagement were necessary for the mechanosensitive

regulation of migration speed and honing efficiency, respectively. Further substantiating

our hypothesis, blocking β2 integrins of cells migrating on Fn altered migration dynamics,

with directionality time becoming statistically indistinguishable from those of untreated

cells migrating on the β1 integrin ligand Col-IV. These data demonstrate that individ-

ual components of the neutrophil chemotactic response are tunable with different ligand,

mechanotactic, and chemotactic cues and that this tuning is integrin dependent.

In Part 2 of this dissertation, we explored the role of cell mechanics in determining

the dynamics of motile cells. Crawling motile cells in particular exhibit a variety of cell

shape dynamics ranging from complex ruffling and bubbling to oscillatory protrusion and

retraction. For example, periodic shape changes during cell migration have been recorded

in fast moving fish epithelial keratocytes where sticking and slipping at opposite sides of the

cell’s broad trailing edge generate bipedal locomotion. In Chapter 4, we modeled crawling

cells using a 2D construction of elements representing linkages between the cytoskeleton and

the underlying substrate. The mechanical properties of the cell were modeled as follows:

First, the elements were connected with springs representing the actin cytoskeleton. Second,

a front element in the lamellipodial region was propelled in the direction of motion. Third,

stick slip adhesion, generating frictional forces opposing the direction of motion, was applied

to elements at the posterior of the cell. By benchmarking the cell shape dynamics of

four spring configurations, each representing a possible elastic configuration of the actin

167

Page 194: The Mechanics of Cell Motility and a Unifying Theory for

cytoskeleton, against the cell shape dynamics of crawling keratocytes, our analysis showed

that elastic coupling to the cell nucleus was necessary to generate the observed motion.

Based on this finding, we select a configuration to study the effects of cell elasticity, size, and

aspect ratio on crawling dynamics. This configuration predicted that shape dynamics are

highly dependent on the lamellipodial elasticity, but less sensitive to elasticity at the trailing

edge. The model could be used to predicted a wide range of dynamics, from coherent bipedal

shape dynamics as seen in crawling keratocytes, to decoherent chaotic shape dynamics as

observed in amoeboid cells. In general, this model could also predict the traction force

applied by cells on their substrate. This work highlights how the dynamical behavior of

crawling cells can be derived from a mechanical properties through which biochemical factors

may operate to regulate cell motility.

Finally, in Chapter 5, we combined our tools for characterizing cell motility from Part 1,

to study the traction forces of both chemokinetic and chemotactic neutrophils. In addition

to seeking the effect of neutrophil traction forces as a function a substrate stiffness, we

sought to measure traction forces in the context of a biological response modifier to assess

the role of cell mechanics in pharmacological intervention, something that had not been

studied previously. β-glucan was used as a model biological response modifier, a substance

that has shown promise as a clinic grade therapeutic. We found that β-glucan modified the

mechanical output of human neutrophils on soft, 5 and 25 kPa 2D substrates, decreasing

both the contractile force and energy applied as the concentration of β-glucan was increased

over a clinically relevant range. Despite these changes in mechanical output, contractile

energy was conserved across substrate stiffness for all concentrations of β-glucan tested. We

also reported an overall inverse correlation between migration speed and mechanical output.

While there has been some speculation that a specific force range existed for facilitating

optimal cell migration, we concluded that optimal cell migration was not associated with

168

Page 195: The Mechanics of Cell Motility and a Unifying Theory for

a specific magnitude of contractile force in the 1-100 nN range. Instead, increased motility

was associated with diminished contraction.

The work presented in this dissertation underscores the synergistic connection between

physics and biology. Here, data from an otherwise biology oriented experimental project

studying cell motility was analyzed using a novel metric developed from random walk theory.

Meanwhile, the principals of fluid and continuum mechanics were merged to measure cellular

traction forces and determine the interdependencies between contractile forces and motility.

As we discuss in the Future Directions section, the link between physics and biology is

further strengthened in that the cellular traction forces of neutrophils from sepsis patients

may indicate new avenues of treatment for sepsis. Indeed, there are many interesting and

clinically relevant opportunities for physicists to apply their skill set to research projects in

the field of biology.

169

Page 196: The Mechanics of Cell Motility and a Unifying Theory for

CHAPTER 7

Extras

1. Cell Trajectory Simulator Class (MATLAB)

Part of Ph.D. training for a modeler and data analyst is learning code, but one quickly

discovers that being a good coder does not necessarily translate into being a good researcher.

Graduate students are not often motivated to write humanly understandable code, often

instead opting to use a “slap it together until it works” approach. As such, reading through

just about any graduate student’s computer code is a surefire way to induce excessive

cringing. To demonstrate what I hope is one exception to this rule, this appendix contains

publishable quality computer code of a MATLAB object class that when instantiated, gen-

erates simulated cell trajectories based on a 2-D persistent biased random walk. Sampling

error and sampling rate are free parameters that make the simulated data output mimic

experimental data. This code was instrumental for work on the directionality time model

and is generically useful for studying cell motility from the perspective of biased random

walk theory.

This modelling tool along with the LEGI model (c.f. Chapter 6), cell tracking, traction

force microscopy, and other modeling tools (all written in either MATLAB or Python) can

be found at, http://www.github.com/aloosley.

170

Page 197: The Mechanics of Cell Motility and a Unifying Theory for

classdef cellTrajSim < handle

% cellTrajSim: Simulates a biased random walk representative of something

% that moves randomly by nature and is biased by an external cue to move in

% a particular direction. An example is a cell undergoing chemotaxis.

%

% The biased random walker choose to change direction based on a Poissonian

% process with magnitude lambda [1/s]. When a reorientation decision is made,

% the orientation angle is drawn from a von Mises distribution with

% directionality factor, kappa. The random walk is sampled every tSample

% [s] with an error factor representing centroid measurement error [um].

% The direction of external bias is set by angMean, and can change using

% the angMeanChgRate [1/s] variable. The simulation time interval is tInt [s],

% and simulations continue until tEnd [s].

%

% Usage

% -----

% **default constructor**

% cellTrajSim(lambda, kappa, speedInit, speedChgRate, centMeasErr,...

% angMean, angMeanChgRate, tInt, tSample, tEnd)

%

% Arguments (input)

% ----------------

% lambda - the average frequency of deciding to change direction [1/s]

% kappa - directional factor (inverse width of von Mises Distribution)

% speed - instanteous speed during random walk [um/s]

% centMeasErr - centroid measurement error [um]

% angMean - mean bias angle

% angMeanChgRate - mean bias angle change rate [1/s]

% tInt - simulation time interval [s]

% tSample - sampling time interval [s]

% tEnd - end simulation time [s]

%

% Public Methods List

% -------------------

% simulateTraj(); % Determines simulation trajectory,

% % coordinates saved to xyArray, sampled

% % coordinates with centroid measurement error

% % are saved to xySampArray

% calcSqDisp(); % Calculates squared displacement from xySampArray

% fitTASDexp(fitTime); % fits TASD to exponential model, stopping fit

% % a fitTime [s]

% fitTASDbeta(fitTime); % fits TASD to beta model, stopping fit a fitTime [s]

% fitTASDbetaCME(fitTime); % fits TASD modified to acount for centroid

% % measurement error (CME),to beta model,

% % stopping fit a fitTime [s]

%

%

% Examples

% --------

% sim = cellTrajSim % Instantiates random walk simulation object with

% % default parameters

%

% sim = cellTrajSim(lambda, kappa, speedInit, speedChgRate, centMeasErr,...

171

Page 198: The Mechanics of Cell Motility and a Unifying Theory for

% angMean, angMeanChgRate, tInt, tSample, tEnd)

% sim.simulateTraj()

% % Instantiates random walk simulation object and simulates

% % trajectory with argument parameters

%

% Variables Returned

% ------------------

% All output variables are publicly accessable. For example, to access

% the raw simulated trajectory (2D x-y coordinates) use,

% simTraj sim.xyArray

%

% List of public object variables:

% tArray - N by 1 array of time coordinates, where N is the simulation

% length, given by tEnd/tInt

% xyArray - N by 2 array of x-y coordinates

% speedArray - N by 1 array of speed over time

% tSampArray - M by 1 array of sampled time coordinates, where M is

% tEnd/tSample

% xySampArray - M by 2 array of sampled coordinates. Noise has been

% added to each sampled coordinate using a radial Gaussian

% with a width given by argument centMeasErr, which

% corresponds to real world centroid measurement error

% angleArray - N by 1 array of the orientation of the cell at each timestep

% turnsMAde - N by 1 array tracking whether or not a decision to turn has

% been made at each timestep

%

%

% Author: Alex J Loosley

% e-mail address: [email protected]

% Release: 2

% Release date: 6/12/14

% Please acknowledge the usage of this code

properties(GetAccess = ’public’, SetAccess = ’private’)

% **Default Simulation Parameters**

lambda = 0.5; % [1/s] Average number of pseudopod events per second

speedInit = 0.2033; % [um/s] Speed at time = 0

speedChgRate = 0.00001 % [um/s] Linear acceleration over time

angMean = 0*(pi/180); % [rad] Overall direction [rad] (model invariant to choice)

angMeanChgRate = 0*(pi/180)/300; % [rad/s] Rate

kappa = 2; % Unitless factor, kappa, in Von Mises distribution

centMeasErr = 2; %[um]

% Simulation Discretization and Timing

tInt = 1; % time interval [s]

tSample = 10; % experimental sampling interval [s]

tEnd = 12*60; % [s]

% Tracking

turnsMade

tArray

tSampArray

172

Page 199: The Mechanics of Cell Motility and a Unifying Theory for

xyArray

xySampArray % trajectories sampled from simulation (i.e. every tSample = 10 seconds)

angleArray

speedArray % new as of 02/09/2014

% Tracking Event timing (Poisson Dist double checked)

evDur % Event Duration Timer

evDurOut % Formatted for output (zeros truncated off end)

% Dynamic Variables (time, spatial coordinates, walker orientation

% angle, speed)

t

x

y

angle

speed

% Indices

tIdx

tSampIdx

evIdx

% SD, TASD, MSD measures

% (SD = squared displacement, TA = time averaged)

sqDispTime

sqDisp

TAsqDisp

TAsqDisp_StE

sqDispExp

TAsqDispExp

% SD and TASD fitting and GOF

expFit

expFitGOF

expFitOutput

expFitRChi2

betaFit

betaFitGOF

betaFitOutput

betaFitRChi2

betaFit_dirTime

expFit_dirTime

expFit_asymp

expFit_asympOff

TASDexpFit

TASDexpFitGOF

TASDexpFitOutput

TASDexpFitRChi2

TASDbetaFit

TASDbetaFitGOF

TASDbetaFitOutput

173

Page 200: The Mechanics of Cell Motility and a Unifying Theory for

TASDbetaFitRChi2

TASDbetaFit_dirTime

TASDexpFit_dirTime

TASDexpFit_asymp

TASDexpFit_asympOff

TASDbetaFitCME

TASDbetaFitCMEGOF

TASDbetaFitCMEOutput

TASDbetaFitCMERChi2

TASDbetaFitCME_dirTime

% Other

tempAngles % see comment below - work around for Von Mises distribution error

end

methods

% Class Constructor

function self = cellTrajSim(lambda, kappa, speedInit, speedChgRate,...

centMeasErr, angMean, angMeanChgRate, tInt, tSample, tEnd)

if nargin > 0

self.lambda = lambda;

self.kappa = kappa;

self.speedInit = speedInit;

self.speedChgRate = speedChgRate;

self.centMeasErr = centMeasErr;

self.angMean = angMean;

self.angMeanChgRate = angMeanChgRate;

self.tInt = tInt;

self.tSample = tSample;

self.tEnd = tEnd;

end

self.turnsMade = zeros(round(self.tEnd/ self.tInt)+1,1);

self.tArray = zeros(round(self.tEnd/ self.tInt)+1,1);

self.xyArray = zeros(round(self.tEnd/ self.tInt)+1,2);

self.angleArray = zeros(round(self.tEnd/ self.tInt)+1,1);

self.xySampArray = zeros(round(self.tEnd/ self.tSample),2);

self.speedArray = zeros(round(self.tEnd/ self.tSample),1);

end

end

methods(Access = private)

% Sets ICs to the origin of space and time (initial angle)

function initialize(self)

self.t = 0;

self.x = 0;

self.y = 0;

self.angle = self.angMean;

self.speed = self.speedInit;

self.tArray(1) = self.t;

174

Page 201: The Mechanics of Cell Motility and a Unifying Theory for

self.xyArray(1,1) = self.x;

self.xyArray(1,2) = self.y;

self.angleArray(1) = self.angle;

self.speedArray(1) = self.speed;

self.tIdx = 1;

self.tSampIdx = 1;

self.evIdx = 1;

end

function addGaussNoiseXY(self)

self.xySampArray(self.tSampIdx,1) = self.x + normrnd(0,self.centMeasErr);

self.xySampArray(self.tSampIdx,2) = self.y + normrnd(0,self.centMeasErr);

self.tSampIdx = self.tSampIdx + 1;

end

% trajectory Step based on a Poissonian process

function trajStep(self)

r = rand(1,1);

if r <= self.lambda*self.tInt % CHANGE DIRECTION

self.turnsMade(self.tIdx) = 1; % Notes that a turn was made

self.evDur(self.evIdx) = self.evDur(self.evIdx) + self.tInt;

% The above line: Minimum event time is dt

self.evIdx = self.evIdx + 1; % Initializes next event

self.angMean = atan2(-self.y, 200-self.x);

self.tempAngles = vmrand(self.angMean + self.angMeanChgRate*self.t,...

self.kappa,2,1);

self.angle = self.tempAngles(1);

% Above line: I requested 2 random numbers at a time because

% vmrand errors from time to time when only one is requested

% (cause of error currently unknown)

self.speed = random(’exp’, self.speedChgRate*self.t + self.speedInit);

% check angle limit - not necessary using Von Mises distribution,

% "vmrand", uncomment for a pseudowrapped normal distribution.

%{

chk = 1;

while chk == 1

if abs(self.angle)> pi

tempAngles = normrnd(self.angMean + self.angMeanChgRate*self.t, self.kappa);

self.angle = tempAngles(1); % I requested 2 random numbers at a time because vmrand errors from time to time when only one is requested (cause of error unknown)

else

chk = 0;

end

end

%}

% self.angle = mod(self.angle+pi,2*pi)-pi; % sets angle between -pi and pi

self.x = self.x + self.speed*cos(self.angle)*self.tInt;

self.y = self.y + self.speed*sin(self.angle)*self.tInt;

self.xyArray(self.tIdx+1,1) = self.x;

self.xyArray(self.tIdx+1,2) = self.y;

self.angleArray(self.tIdx+1) = self.angle;

self.speedArray(self.tIdx+1) = self.speed;

175

Page 202: The Mechanics of Cell Motility and a Unifying Theory for

else % Otherwise DON’T CHANGE DIRECTION (or speed)

self.evDur(self.evIdx) = self.evDur(self.evIdx) + self.tInt;

% The above line updates event interval duration timer

self.x = self.x + self.speed*cos(self.angle)*self.tInt;

self.y = self.y + self.speed*sin(self.angle)*self.tInt;

self.xyArray(self.tIdx+1,1) = self.x;

self.xyArray(self.tIdx+1,2) = self.y;

self.angleArray(self.tIdx+1) = self.angle;

self.speedArray(self.tIdx+1) = self.speed;

end

% Update time counters

self.t = self.t + self.tInt;

self.tArray(self.tIdx + 1) = self.t;

self.tIdx = self.tIdx + 1;

end

end

methods

function simulateTraj(self)

self.initialize()

self.evDur = zeros(round(self.tEnd/self.tInt),1);

while self.t < self.tEnd

if mod(round(self.t/self.tInt),round(self.tSample/self.tInt)) == 0

self.addGaussNoiseXY() % If the above is true, this samples

% the trajectory point and adds Gaussian noise

end

self.trajStep()

end

self.evDurOut = self.evDur(1:find(self.evDur==0,1)-1);

self.tSampArray = 0:self.tSample:self.tEnd;

end

% Square displacement (SD, TASD) and square displacement

% exponent calculations

function calcSqDisp(self)

nSamp = size(self.xySampArray,1);

self.sqDispTime = (self.tSample : self.tSample : (nSamp-1)*self.tSample)’;

% Square Displacement, reiterate simulation to get

% ensemble average

self.sqDisp = NaN(nSamp-1,1);

for k_int = 1:nSamp - 1

xd = self.xySampArray(1+k_int,1)-self.xySampArray(1,1);

yd = self.xySampArray(1+k_int,2)-self.xySampArray(1,2);

self.sqDisp(k_int) = xd^2 + yd^2;

end

176

Page 203: The Mechanics of Cell Motility and a Unifying Theory for

% Time Averaged Square Displacement

self.TAsqDisp = NaN(nSamp-1,1);

self.TAsqDisp_StE = NaN(nSamp-1,1);

for k_int = 1:nSamp-1

rSq = NaN(nSamp-k_int,1);

for k =1:nSamp-k_int

xd = self.xySampArray(k+k_int,1)-self.xySampArray(k,1);

yd = self.xySampArray(k+k_int,2)-self.xySampArray(k,2);

rSq(k)=xd^2+yd^2;

end

self.TAsqDisp(k_int)=nanmean(rSq);

self.TAsqDisp_StE(k_int)=nanste(rSq);

end

% meanSqDisp Exponent calculated by taking the forward

% difference derivative of MSD vs time in loglog space

self.sqDispExp = NaN(size(self.sqDisp,1)-1,1);

self.TAsqDispExp = NaN(size(self.TAsqDisp,1)-1,1);

for kk = 1:size(self.sqDisp,1)-1

self.sqDispExp(kk) = log(self.sqDisp(kk+1)/self.sqDisp(kk)) / log((kk+1)/kk);

self.TAsqDispExp(kk) = log(self.TAsqDisp(kk+1)/self.TAsqDisp(kk)) / log((kk+1)/kk);

end

end

% Fits the simulated MSD using data points up to time tEndFit (seconds)

function fitTASDexp(self,tEndFit)

model = fittype(’a-b*exp(-time/dirTime)’,’independent’,’time’,...

’coefficients’,{’a’,’b’,’dirTime’});

Value = ’NonlinearLeastSquares’;

opts = fitoptions(’method’, Value);

opts.StartPoint=[2,1,20];

opts.Upper = [2.5,2.5,tEndFit];

opts.Lower = [1.2,0.1,1 ];

[self.TASDexpFit, self.TASDexpFitGOF, self.TASDexpFitOutput] =...

fit(self.sqDispTime(1:round(tEndFit/self.tSample)), ....

self.TAsqDispExp(1:round(tEndFit/self.tSample)), model, opts);

self.TASDexpFit_dirTime = self.TASDexpFit.dirTime;

self.TASDexpFit_asymp = self.TASDexpFit.a;

self.TASDexpFit_asympOff = self.TASDexpFit.b;

self.TASDexpFitRChi2 = self.TASDexpFitGOF.rmse^2;

end

function fitTASDbeta(self,tEndFit)

model = fittype(’(1+2*(time/dirTime))/(1+(time/dirTime))’,...

’independent’,’time’,’coefficients’,{’dirTime’});

Value = ’NonlinearLeastSquares’;

opts = fitoptions(’method’, Value);

opts.StartPoint= [20];

opts.Upper = [tEndFit];

opts.Lower = [1 ];

[self.TASDbetaFit, self.TASDbetaFitGOF, self.TASDbetaFitOutput] =...

fit(self.sqDispTime(1:round(tEndFit/self.tSample)), ...

177

Page 204: The Mechanics of Cell Motility and a Unifying Theory for

self.TAsqDispExp(1:round(tEndFit/self.tSample)), model, opts);

self.TASDbetaFit_dirTime = self.TASDbetaFit.dirTime;

self.TASDbetaFitRChi2 = self.TASDbetaFitGOF.rmse^2;

end

function fitTASDbetaCME(self,tEndFit)

model = fittype(’(1+2*(time/dirTime))/(1+(time/dirTime))’,...

’independent’,’time’,’coefficients’,{’dirTime’});

Value = ’NonlinearLeastSquares’;

opts = fitoptions(’method’, Value);

opts.StartPoint= [20];

opts.Upper = [tEndFit];

opts.Lower = [1 ];

[self.TASDbetaFitCME, self.TASDbetaFitCMEGOF, self.TASDbetaFitCMEOutput] =...

fit(self.sqDispTime(1:round(tEndFit/self.tSample)),...

self.TAsqDispExp(1:round(tEndFit/self.tSample)).*...

(self.TAsqDisp(1:round(tEndFit/self.tSample))./...

(self.TAsqDisp(1:round(tEndFit/self.tSample))-self.centMeasErr^2)),...

model, opts);

self.TASDbetaFitCME_dirTime = self.TASDbetaFitCME.dirTime;

self.TASDbetaFitCMERChi2 = self.TASDbetaFitCMEGOF.rmse^2;

end

function fitExpIn(self,tIn,yIn,tEndFit)

model = fittype(’a-b*exp(-time/dirTime)’,’independent’,’time’,...

’coefficients’,{’a’,’b’,’dirTime’});

Value = ’NonlinearLeastSquares’;

opts = fitoptions(’method’, Value);

opts.StartPoint=[2,1,20];

opts.Upper = [2.5,2.5,tEndFit];

opts.Lower = [1.2,0.1,1 ];

[self.expFit, self.expFitGOF, self.expFitOutput] =...

fit(tIn(1:round(tEndFit/self.tSample)), yIn(1:round(tEndFit/self.tSample)),...

model, opts);

self.expFit_dirTime = self.expFit.dirTime;

self.expFit_asymp = self.expFit.a;

self.expFit_asympOff = self.expFit.b;

self.expFitRChi2 = self.expFitGOF.rmse^2;

end

function fitBetaIn(self,tIn,yIn,tEndFit)

model = fittype(’(1+2*(time/dirTime))/(1+(time/dirTime))’,...

’independent’,’time’,’coefficients’,{’dirTime’});

Value = ’NonlinearLeastSquares’;

opts = fitoptions(’method’, Value);

opts.StartPoint=[20];

opts.Upper = [tEndFit];

opts.Lower = [1 ];

[self.betaFit, self.betaFitGOF, self.betaFitOutput] =...

fit(tIn(1:round(tEndFit/self.tSample)), yIn(1:round(tEndFit/self.tSample)),...

model, opts);

178

Page 205: The Mechanics of Cell Motility and a Unifying Theory for

self.betaFit_dirTime = self.betaFit.dirTime;

self.betaFitRChi2 = self.betaFitGOF.rmse^2;

end

end

end

2. How to Save $150 On A New Mercury Lamp

Earlier this year, our fluorescent mercury arc lamp began flickering at the far too

young age of 1144 hours. After a long back and forth with the manufacturer about

honoring the 1200 hour warranty, a stalemate ensued because the warranty was only

valid if the bulb failed to ignite completely. Regardless of the warranty conditions, a

flickering arc lamp greatly impaired our ability to do experiments so it was time to

bring out the big guns (Fig. 7.1), an intensity kymograph that led to a well deserved

$165 discount on a new bulb. Small victories.

time length along yellow line in

KymographBright

frames 58 & 59

frames 37 & 38

Dark

NOTESa) There �uctuations a persistentb) These �uctuations began happening like this sometime around Dec 10th, there was no smooth transition from proper lamp output to this behaviourc) A di�erent lamp (same model) with 800hrs instead of 1144 hrs works perfectly in the same unit, everything else equald) This is one movie, of manye) In all cases, lamp has been left on >15 min after ignitef ) Fluctuations are not at the frequency of AC �uctuations, they last 10-30sg) Intensity drops are 10-20% (see arrows above right of Kymograph)h) Fluctuations clearly visible through multiple cameras and by eye through microscope eyepiece

Intensity Fluctuation

10µm

2min

Figure 7.1. The $165 kymograph. Still frames (left) and corresponding pixelintensity kymograph (right) showing flickering of our too young to die fluorescentbulb. Arrows along the y-axis of the kymograph correspond to decreased bulboutput.

179

Page 206: The Mechanics of Cell Motility and a Unifying Theory for

Bibliography

[1] X. M. O’Brien, A. J. Loosley, K. E. Oakley, J. X. Tang, and J. S. Reichner,

“Technical advance: Introducing a novel metric, directionality time, to quantify

human neutrophil chemotaxis as a function of matrix composition and stiffness,”

J Leukocyte Biol, vol. 95, pp. 993–1004, 2014.

[2] A. J. Loosley and J. X. Tang, “Stick-slip motion and elastic coupling in crawling

cells,” Phys Rev E, vol. 86, p. 031908, 2012.

[3] M. Amercrombie, “The croonian lecture, 1978: The crawling movement of meta-

zoan cells,” Proc Roy Soc London Ser B, vol. 207, pp. 129–147, 1980.

[4] G. Li and J. X. Tang, “Accumulation of microswimmwers near a surface me-

diated by collision and rotational brownian motion,” Phys Rev Lett, vol. 103,

p. 078101, 2009.

[5] T. Oliver, M. Dembo, and K. Jacobson, “Traction forces in locomoting cells,”

Cell Motil Cytoskel, vol. 31, pp. 225–240, 1995.

[6] T. J. Mitchison and L. P. Cramer, “Actin-based cell motility and cell locomo-

tion,” Cell, vol. 84, pp. 371–379, 1996.

[7] T. D. Pollard and G. G. Borisy, “Cellular motility driven by assembly and

disassembly of actin filaments,” Cell, vol. 112, pp. 453–465, 2003.

[8] C. Nathan, “Neutrophils and immunity: challenges and opportunities,” Nat Rev

Immunol, vol. 6, pp. 173–182, 2006.

[9] P. Martin, “Wound healing – aiming for perfect skin regeneration,” Science,

vol. 276, pp. 75–81, 1997.

[10] B. Eckes, R. Nischt, and T. Krieg, “Cell-matrix interactions in dermal repair

and scarring,” Fibro Tissue Repair, vol. 3, p. 4, 2010.

180

Page 207: The Mechanics of Cell Motility and a Unifying Theory for

[11] P. Martin and J. Lewis, “Actin cables and epidermal movement in embryonic

wound healing,” Nature, vol. 360, pp. 179–183, 1992.

[12] X. Trepat, M. R. Wasserman, T. E. Angelini, E. Millet, D. A. Weitz, J. P.

Butler, and J. J. Fredberg, “Physical forces during collective cell migration,”

Nat Phys, vol. 5, pp. 426–430, 2009.

[13] X. Serra-Picamal, V. Conte, R. Vincent, E. Anon, D. T. Tambe, E. Bazellieres,

J. P. Butler, J. J. Fredberg, and X. Trepat, “Mechanical waves during tissue

expansion,” Nat Phys, vol. 8, pp. 628–634, 2012.

[14] G. A. Dunn and G. E. Jones, “Cell motility under the microscope: Vorsprung

durch technik,” Nat Rev Mol Cell Bio, vol. 5, pp. 667–672, 2004.

[15] O. Shimomura, F. H. Johnson, and Y. Saiga, “Extraction, purification and prop-

erties of aequorin, a bioluminescent protein from the luminous hydromedusan,

aequorea,” J Cell Compar Physl, vol. 59, pp. 223–239, 1962.

[16] V. Lecaudey and D. Gilmour, “Organizing moving groups during morphogene-

sis,” Curr Opin Cell Biol, vol. 18, pp. 102–107, 2006.

[17] B. Akiyoshi, K. K. Sarangapani, A. F. Powers, C. R. Nelson, S. L. Reichow,

H. Arellano-Santoyo, T. Gonen, J. A. Ranish, C. L. Asbury, and S. Big-

gins, “Tension directly stabilizes reconstituted kinetochore-microtubule attach-

ments,” Nature, vol. 468, pp. 576–579, 2010.

[18] F. Julicher, K. Kruse, J. Prost, and J.-F. Joanny, “Active behavior of the cy-

toskeleton,” Phys Rep, vol. 449, pp. 3–28, 2007.

[19] F. Gittes, B. Mickey, J. Nettleton, and J. Howard, “Flexural rigidity of micro-

tubules and actin filaments measured from thermal fluctuations in shape,” J

Cell Biol, vol. 120, pp. 923–934, 1993.

[20] J. Lee, A. Ishihara, J. A. Theriot, and K. Jacobson, “Principles of locomotion

for simple-shaped cells,” Nature, vol. 362, pp. 167–171, 1993.

[21] K. Keren, Z. Pincus, G. M. Allen, E. L. Barnhart, G. Marriott, A. Mogilner,

and J. A. Theriot, “Mechanism of shape determination in motile cell,” Nature,

vol. 453, pp. 475–480, 2008.

181

Page 208: The Mechanics of Cell Motility and a Unifying Theory for

[22] J. L. Ross, M. Y. Ali, and D. M. Warshaw, “Cargo transport: Molecular motors

navigate a complex cytoskeleton,” Curr Opin Cell Biol, vol. 20, pp. 41–47, 2008.

[23] J. R. Bamburg, H. E. Harris, and A. G. Weeds, “Partial purification and char-

acterization of an actin depolymerizing factor from brain,” FEBS Lett, vol. 121,

pp. 178–182, 1980.

[24] M. V. Troys, L. Huyck, S. Leyman, S. Dhaese, J. Vandekerkhove, and C. Ampe,

“Ins and outs of adf/cofilin activity and regulation,” Eur J Cell Biol, vol. 87,

pp. 649–667, 2008.

[25] A. Huttenlocher and A. R. Horwitz, “Integrins in cell migration,” Cold Spr

Harbor Perspec Biol, vol. 3, p. a005074, 2011.

[26] E. Puklin-Faucher and M. P. Sheetz, “The mechanical integrin cycle,” J Cell

Sci, vol. 122, pp. 179–186, 2009.

[27] M. L. Gardel, B. Sabass, L. Ji, G. Danuser, U. S. Schwarz, and C. M. Waterman,

“Traction stress in focal adhesions correlates biphasically with actin retrograde

flow speed,” J Cell Biol, vol. 183, pp. 999–1005, 2008.

[28] Y. Li, P. Bhimalapuram, and A. R. Dinner, “Model for how retrograde actin flow

regulates adhesion traction stresses,” J Phys-Condens Mat, vol. 22, p. 194113,

2010.

[29] S. W. Moore, P. Roca-Cusachs, and M. P. Sheetz, “Stretchy proteins on stretchy

substrates: The important elements of integrin-mediated rigidity sensing,” Dev

Cell, vol. 19, pp. 194–206, 2010.

[30] A. K. Harris, P. Wild, and D. Stopak, “Silicone rubber substrata: a new wrinkle

in the study of cell locomotion,” Science, vol. 208, pp. 177–179, 1980.

[31] J. Lee, M. Leonard, T. Oliver, A. Ishihara, and K. Jacobson, “Traction forces

generated by locomoting keratocytes,” J Cell Biol, vol. 127, pp. 1957–1963,

1994.

[32] M. Dembo and Y. L. Wang, “Stresses at the cell-to-substrate interface during

locomotion of fibroblasts,” Biophys J, vol. 76, pp. 2307–2316, 1999.

182

Page 209: The Mechanics of Cell Motility and a Unifying Theory for

[33] J. P. Butler, I. M. Tolic-Nørrelykke, B. Fabry, and J. J. Fredberg, “Traction

fields, moments, and strain energy that cells exert on their surroundings,” Am

J Physiol Cell Physiol, vol. 282, pp. C595–C605, 2002.

[34] B. Sabass, M. L. Gardel, C. M. Waterman, and U. S. Schwarz, “High resolution

traction force microscopy based on experimental and computational advances,”

Biophys J, vol. 94, pp. 207–, 2008.

[35] J. Huang, X. Peng, L. Qin, T. Zhu, C. Xiong, Y. Zhang, and J. Fang, “Determi-

nation of cellular tractions on elastic substrate based on an integral boussinesq

solution,” J Biomech Eng, vol. 131, p. 061009, 2009.

[36] J. Toyjanova, E. Bar-Kochba, C. Lopez-Fagundo, J. Reichner, D. Hoffman-

Kim, and C. Franck, “High resolution, large deformation 3d traction force mi-

croscopy,” PLoS One, vol. 9, p. e90976, 2014.

[37] S. A. Maskarineca, C. Franck, D. A. Tirrell, and G. Ravichandran, “Quantifying

cellular traction forces in three dimensions,” Proc Natl Acad Sci USA, vol. 106,

pp. 22108–22113, 2009.

[38] J. C. del Alamo, R. Meili, B. Alvarez-Gonzalez, B. Alonso-Latorre, E. Bastou-

nis, R. Firtel, and J. C. Lasheras, “Three-dimensional quantification of cellular

traction forces and mechanosensing of thin substrata by fourier traction force

microscopy,” PLoS One, vol. 8, p. e69850, 2013.

[39] C. Lopez-Fagundo, E. Bar-Kochba, L. Livi, D. Hoffman-Kim, and C. Franck,

“Traction forces of schwann cells on compliant substrates,” J R Soc Interface,

vol. accepted, 2014.

[40] E. Gutierrez, E. Tkachenko, A. Besser, P. Sundd, K. Ley, G. Danuser, M. H.

Ginsberg, and A. Groisman, “High refractive index silicone gels for simultaneous

total internal reflection fluorescence and traction force microscopy of adherent

cells,” PLoS One, vol. 6, p. e23807, 2011.

[41] K. M. Stroka and H. Aranda-Espinoza, “Neutrophils display biphasic relation-

ship between migration and substrate stiffness,” Cell Motil Cytoskeleton, vol. 66,

pp. 328–341, 2009.

183

Page 210: The Mechanics of Cell Motility and a Unifying Theory for

[42] B. L. Bangasser, S. S. Rosenfeld, and D. J. Odde, “Determinants of maximal

force transmission in a motor-clutch model of cell traction in a compliant mi-

croenvironment,” Biophys J, vol. 105, pp. 581–592, 2013.

[43] C.-M. Lo, H. Wang, M. Dembo, and Y. li Wang, “Cell movement is guided by

the rigidity of the substrate,” Biophys J, vol. 79, pp. 144–152, 2000.

[44] T. Yeung, P. C. Georges, L. A. Flanagan, B. Marg, M. Ortiz, M. Funaki, N. Za-

hir, W. Ming, V. Weaver, and P. A. Janmey, “Effects of substrate stiffness

on cell morphology, cytoskeletal structure, and adhesion,” Cell Motil Cytoskel,

vol. 60, pp. 24–34, 2005.

[45] A. J. Engler, S. Sen, H. L. Sweeney, and D. E. Discher, “Matrix elasticity directs

stem cell lineage specification,” Cell, vol. 126, pp. 677–689, 2006.

[46] V. Vogel and M. P. Sheetz, “Local force and geometry sensing regulate cell

functions,” Nat Rev Mol Cell Bio, vol. 7, pp. 265–275, 2006.

[47] M. Ghibaudo, A. Saez, L. Trichet, A. Xayaphoummine, J. Browaeys, P. Sil-

berzan, A. Buguin, and B. Ladoux, “Traction forces and rigidity sensing regu-

late cell functions,” Soft Matter, vol. 4, pp. 1836–1843, 2008.

[48] P. W. Oakes, D. C. Patel, N. A. Morin, D. P. Zitterbart, B. Fabry, J. S. Reichner,

and J. X. Tang, “Neutrophil morphology and migration are affected by substrate

elasticity,” Blood, vol. 114, pp. 1387–1395, 2009.

[49] R. A. Jannat, G. P. Robbins, B. G. Ricart, M. Dembo, and D. A. Hammer,

“Neutrophil adhesion and chemotaxis depend on substrate mechanics,” J Phys-

Condens Mat, vol. 22, p. 194117, 2010.

[50] R. A. Jannat, M. Dembo, and D. A. Hammer, “Traction forces of neutrophils

migrating on compliant substrates,” Biophys J, vol. 101, pp. 575–584, 2011.

[51] P. Marcq, N. Yoshinaga, and J. Prost, “Rigidity sensing explained by active

matter theory,” Biophys J, vol. 101, pp. L33–L35, 2011.

[52] S.-Y. Tee, J. Fu, C. S. Chen, and P. A. Janmey, “Cell shape and substrate

rigidity both regulate cell stiffness,” Biophys J, vol. 100, no. 5, pp. L25–L27,

2011.

184

Page 211: The Mechanics of Cell Motility and a Unifying Theory for

[53] S. Ghassemi, G. Meacci, S. Liu, A. A. Gondarenko, A. Mathur, P. Roca-

Cusachs, M. P. Sheetz, and J. Hone, “Cells test substrate rigidity by local

contractions on submicrometer pillars,” Proc Natl Acad Sci USA, vol. doi,

p. 1119886109, 2012.

[54] A. R. Houk, A. Jilkine, C. O. Mejean, R. Boltyanskiy, E. R. Dufresne, S. B.

Angenent, S. J. Altschuler, L. F. Wu, and O. D. Weiner, “Membrane tension

maintains cell polarity by confining signals to the leading edge during neutrophil

migration,” Cell, vol. 148, no. 1, pp. 175–188, 2012.

[55] P. W. Oakes, Y. Backham, J. Stricker, and M. L. Gardel, “Tension is required

but not sufficient for focal adhesion maturation without a stress fiber template,”

J Cell Biol, vol. doi, p. 10.1083/jcb.201107042, 2012.

[56] G. A. Dunn, “Characterising a kinesis response: time averaged measures of cell

speed and directional persistence,” Agents Actions Suppl, vol. 12, pp. 14–33,

1983.

[57] R. T. Tranquillo, D. A. Lauffenburger, and S. H. Zigmond, “A stochastic model

for leukocyte random motility and chemotaxis based on receptor binding fluc-

tuations,” J Cell Biol, vol. 106, pp. 303–309, 1988.

[58] C. L. Stokes, D. A. Lauffenburger, and S. K. Williams, “Migration of individual

microvessel endothelial cells: Stochastic model and parameter measurement,” J

Cell Sci, vol. 99, pp. 419–430, 1991.

[59] M. R. Parkhurst and W. M. Saltzman, “Quantification of human neutrophil

motility in three-dimensional collagen gels. effect of collagen concentration,”

Biophys J, vol. 61, pp. 306–315, 1992.

[60] E. A. Codling, M. J. Plank, and S. Benhamou, “Random walk models in biol-

ogy,” J R Soc Interface, vol. 5, pp. 813–834, 2008.

[61] C. A. Reinhart-King, M. Dembo, and D. A. Hammer, “Cell-cell mechanical

communication through compliant substrates,” Biophys J, vol. 95, pp. 6044–

6051, 2008.

185

Page 212: The Mechanics of Cell Motility and a Unifying Theory for

[62] P.-H. Wu, A. Giri, S. X. Sun, and D. Wirtz, “Three-dimensional cell migration

does not follow a random walk,” Proc Natl Acad Sci USA, vol. 111, pp. 3949–

3954, 2014.

[63] T. R. Mempel, S. E. Henrickson, and U. H. von Andrian, “T-cell priming by

dendriticcells in lymph nodes occurs in three distinct phases,” Nature, vol. 427,

pp. 154–159, 2004.

[64] G. Shakhar, R. L. Lindquist, D. Skokos, D. Dudziak, J. H. Huang, M. C. Nussen-

zweig, and M. L. Dustin, “Stable t cell-dendritic cell interactions precede the

development of both tolerance and immunity in vivo,” Nat Immunol, vol. 6,

pp. 707–714, 2005.

[65] R. G. A. Gautreau, “Quantitative and unbiased analysis of directional persis-

tence in cell migration,” Nat Protoc, vol. 9, p. 19311943, 2014.

[66] E. Batschelet, Circular statistics in biology. London, UK: Academic Press Inc,

1981.

[67] T. Nattermann and L.-H. Tang, “Kinetic surface roughening. i. the kardar-

parisi-zhang equation in the weak-coupling regime,” Phys Rev A, vol. 45,

pp. 7156–7161, 1992.

[68] N. Gal, D. Lechtman-goldstein, and D. Weihs, “Particle tracking in living cells:

a review of the mean square displacement method and beyond,” Rheol Acta,

pp. 1–19, 2013.

[69] M. Gawronski, J. T. Park, A. S. Magee, and H. Conrad, “Microfibrillar structure

of pgg-glucan in aqueous solution as triple-helix aggregates by small angle x-ray

scattering.,” Biopolymers, vol. 50, pp. 569–578, 1999.

[70] B. W. LeBlanc, J. E. Albina, and J. S. Reichner, “The effect of pgg-β-glucan on

neutrophil chemotaxis in vivo.,” J Leukocyte Biol, vol. 79, pp. 667–675, 2006.

[71] B. Li, D. J. Allendorf, R. Hansen, J. Marroquin, C. Ding, D. E. Cramer, and

J. Yan, “Yeast β-glucan amplifies phagocyte killing of ic3b-opsonized tumor

cells via complement receptor 3-syk-phosphatidylinositol 3-kinase pathway,” J

Immunol, vol. 177, pp. 1661–1669, 2006.

186

Page 213: The Mechanics of Cell Motility and a Unifying Theory for

[72] F. Hong, J. Yan, J. T. Baran, D. J. Allendorf, R. D. Hansen, G. R. Ostroff,

P. X. Xing, N. K. Cheung, and G. D. Ross, “Mechanism by which orally admin-

istered β-1,3-glucans enhance the tumoricidal activity of antitumor monoclonal

antibodies in murine tumor models,” J Immunol, vol. 173, pp. 797–806, 2004.

[73] C. Qi, Y. Cai, L. Gunn, C. Ding, B. Li, G. Kloecker, K. Qian, J. Vasilakos,

S. Saijo, Y. Iwakura, J. R. Yannelli, and J. Yan, “Differential pathways regulat-

ing innate and adaptive antitumor immune responses by particulate and soluble

yeast-derived β-glucans,” Blood, vol. 117, pp. 6825–6836, 2011.

[74] Biothera; Biothera. Study of Imprime PGGr in Combination With Cetuximab

in Subjects With Recurrent or Progressive KRAS Wild Type Colorectal Cancer

(PRIMUS). In: ClinicalTrials.gov [Internet]. Bethesda (MD): National Library

of Medicine (US). 2000-2014. Available from: http://clinicaltrials.gov/

show/NCT01309126 NLM Identifier: NCT01309126. Several of these clinical trial

records can be found by searching the ClinicalTrials.gov database for “imprime

PGG”.

[75] M. B. Harler, E. Wakshull, E. J. Filardo, J. E. Albina, and J. S. Reichner,

“Promotion of neutrophil chemotaxis through differential regulation of β1 and

β2 integrins,” J Immunol, vol. 162, pp. 6792–6799, 1999.

[76] L. M. Lavigne, J. E. Albina, and J. S. Reichner, “Beta-glucan is a fungal de-

terminant for adhesion-dependent human neutrophil functions,” J Immunol,

vol. 177, pp. 8667–8675, 2006.

[77] Y. Xia and G. D. Ross, “Generation of recombinant fragments of cd11b ex-

pressing the functional β-glucan-binding lectin site of cr3,” J Immunol, vol. 162,

pp. 7285–7295, 1999.

[78] H. S. Goodridge, A. J. Wolf, and D. M. Underhill, “Beta-glucan recognition by

the innate immune system,” Immunol Rev, vol. 230, pp. 38–50, 2009.

[79] X. M. OBrien, K. E. Heflin, L. M. Lavigne, K. Yu, M. Kim, A. R. Salomon,

and J. S. Reichner, “Lectin site ligation of cr3 induces conformational changes

and signaling,” J Biol Chem, vol. 287, pp. 3337–3348, 2012.

187

Page 214: The Mechanics of Cell Motility and a Unifying Theory for

[80] K. Guevorkian and J. James M. Valles, “Swimming paramecium in magnetically

simulated enhanced, reduced, and inverted gravity environments,” Proc Natl

Acad Sci USA, vol. 103, pp. 13051–13056, 2006.

[81] B. L. Taylor, I. B. Zhulin, and M. S. Johnson, “Aerotaxis and other energy-

sensing behavior in bacteria,” Ann Rev Microbiol, vol. 53, pp. 103–128, 1999.

[82] C.-M. Lo, H.-B. Wang, M. Dembo, and Y. li Wang, “Cell movement is guided

by the rigidity of the substrate,” Biophys J, vol. 79, pp. 144–152, 2000.

[83] J. B. McCarthy, S. L. Palm, and L. T. Furcht, “Migration by haptotaxis of a

schwann cell tumor line to the basement membrane glycoprotein laminin,” J

Cell Biol, vol. 97, no. 3, pp. 772–777, 1983.

[84] R. F. Diegelmann and M. C. Evans, “Wound healing: An overview of acute,

fibrotic and delayed healing,” Front Biosci, vol. 9, pp. 283–289, 2004.

[85] C.-L. Lin, R. M. Suri, R. A. Rahdon, J. M. Austyn, and J. A. Roake, “Den-

dritic cell chemotaxis and transendothelial migration are induced by distinct

chemokines and are regulated on maturation,” European Journal of Immunol-

ogy, vol. 28, no. 12, pp. 4114–4122, 1998.

[86] D. R. Levitan and C. Petersen, “Sperm limitation in the sea,” Trends Ecol Evol,

vol. 10, pp. 228–231, 1995.

[87] G. Valentin, P. Haas, and D. Gilmour, “The chemokine sdf1a coordinates tissue

migration through the spatially restricted activation of cxcr7 and cxcr4b,” Curr

Biol, vol. 17, no. 12, pp. 1026–1031, 2007.

[88] E. C. Woodhouse, R. F. Chuaqui, and L. A. Liotta, “General mechanisms of

metastasis,” Cancer, vol. 80, pp. 1529–1537, 1997.

[89] C. R. Cooper and K. J. Pienta, “Cell adhesion and chemotaxis in prostate cancer

metastasis to bone: a minireview,” Prostate Cancer P D, vol. 3, pp. 6–12, 2000.

[90] F. Lin, C. M.-C. Nguyen, S.-J. Wang, W. Saadi, S. P. Gross, and N. L. Jeon,

“Effective neutrophil chemotaxis is strongly influenced by mean il-8 concentra-

tion,” Biochem Bioph Res Co, vol. 319, pp. 576–581, 2004.

188

Page 215: The Mechanics of Cell Motility and a Unifying Theory for

[91] L. F. Richardson, “The problem of contiguity: An appendix to statistic of

deadly quarrels,” Gen Syst, vol. 61, pp. 139–187, 1961.

[92] B. Mandelbrot, “How long is the coast of britain? statistical self-similarity and

fractional dimension,” Science, vol. 156, no. 3775, pp. 636–638, 1967.

[93] H. G. Othmer, S. R. Dunbar, and W. Alt, “Models of dispersal in biological

systems,” J Math Biol, vol. 26, pp. 263–298, 1988.

[94] L. M. Marsh and R. E. Jones, “The form and consequences of random walk

movement models,” J Theor Biol, vol. 133, pp. 113–131, 1988.

[95] W. Luo, C. han Yu, Z. Z. Lieu, J. Allard, A. Mogilner, M. P. Sheetz, and A. D.

Bershadsky, “Analysis of the local organization and dynamics of cellular actin

networks,” J Cell Biol, vol. 202, pp. 1057–1073, 2013.

[96] S. Goldstein, “On diffusion by discontinuous movements, and on the telegraph

equation.,” J Mech Appl Math, vol. 6, pp. 129–156, 1951.

[97] M. Evans, N. Hastings, and B. Peacock, Statistical Distributions. Wiley, 3 ed.,

2000.

[98] D. W. Marquardt, “An algorithm for least-squares estimation of nonlinear pa-

rameters,” J Soc Ind Appl Math, vol. 11, pp. 431–441, 1963.

[99] S. M. Sze and M.-K. Lee, Semiconductor Devices: Physics and Technology.

Wiley, 3 ed., 2012.

[100] T. N. Mayadas and X. Cullere, “Neutrophil β2 integrins: moderators of life or

death decisions,” Trends Immunol, vol. 26, pp. 388–395, 2005.

[101] R. J. Pelham and Y.-L. Wang, “Cell locomotion and focal adhesions are reg-

ulated by substrate flexibility,” Proc Natl Acad Sci USA, vol. 94, pp. 13661–

13665, 1997.

[102] L. A. Flanagan, Y.-E. Ju, B. Marg, M. Osterfield, and P. A. Janmey, “Neurite

branching on deformable substrates,” Neuroreport, vol. 13, pp. 2411–2415, 2002.

[103] A. J. Engler, M. A. Griffin, S. Sen, C. G. Bonnemann, H. L. Sweeney, and

D. E. Discher, “Myotubes differentiate optimally on substrates with tissue-like

stiffness : pathological implications for soft or stiff microenvironments,” J Cell

189

Page 216: The Mechanics of Cell Motility and a Unifying Theory for

Biol, vol. 166, pp. 877–887, 2004.

[104] A. J. Engler, M.Sheehan, H. L. Sweeney, and Dennis.E.Discher, “Substrate

compliance versus ligand density in cell on gel responses.,” Biophys J, vol. 86,

pp. 617–628, 2004.

[105] B. Qi, N. Tessier-Doyen, and J. Absi, “Youngs modulus evolution with tem-

perature of glass/andalusite model materials: Experimental and numerical ap-

proach,” Comp Mater Sci, vol. 55, pp. 44–53, 2012.

[106] F. Coussen, D. Choquet, M. P. Sheetz, and H. P. Erickson, “Trimers of the

fibronectin cell adhesion domain localize to actin filament bundles and undergo

rearward translocation,” J Cell Sci, vol. 115, pp. 2581–2590, 2002.

[107] A. Saez, A. Buguin, P. Silberzan, and B. Ladoux, “Is the mechanical activity

of epithelial cells controlled by deformations or forces?,” Biophys J, vol. 89,

pp. L52–L54, 2005.

[108] F. M. Hameed, A. J. Loosley, D. van Noort, and M. P. Sheetz, “Cells test rigidity

of nanopillars by local contractions to uniform displacements.” submitted to

Biophys J 2013, 2013.

[109] A. K. Yip, K. Iwasaki, C. Ursekar, H. Machiyama, M. Saxena, H. Chen,

I. Harada, K.-H. Chiam, and Y. Sawada, “Cellular response to substrate rigidity

is governed by either stress or strain,” Biophys J, vol. 104, pp. 19–29, 2013.

[110] Y. Sawada, M. Tamada, B. J. Dubin-Thaler, O. Cherniavskaya, R. Sakai,

S. Tanaka, and M. P. Sheetz, “Force sensing by mechanical extension of the

src family kinase substrate p130cas,” Cell, vol. 127, pp. 1015–1026, 2006.

[111] A. del Rio, R. Perez-Jimenez, R. Liu, P. Roca-Cusachs, J. M. Fernandez, and

M. P. Sheetz, “Stretching single talin rod molecules activates vinculin binding,”

Science, vol. 323, pp. 638–641, 2009.

[112] M. A. Williams and J. S. Solomkin, “Integrin-mediated signaling in human

neutrophil functioning,” J Leukocyte Biol, vol. 65, pp. 725–736, 1999.

[113] R. O. Hynes, “Integrins: Bidirectional, allosteric signaling machines,” Cell,

vol. 110, pp. 673–687, 2002.

190

Page 217: The Mechanics of Cell Motility and a Unifying Theory for

[114] S. Gehler, M. Baldassarre, Y. Lad, J. L. Leight, M. A. Wozniak, K. M. Riching,

K. W. Eliceiri, V. M. W. abd David A. Calderwood, and P. J. Keely, “Filamin

aβ1 integrin complex tunes epithelial cell response to matrix tension,” Mol Biol

Cell, vol. 20, pp. 3224–3238, 2009.

[115] J. B. Litzenberger, J.-B. Kim, P. Tummala, and C. R. Jacobs, “β1 integrins

mediate mechanosensitive signaling pathways in osteocytes,” Calcif Tissue Int,

vol. 86, pp. 325–332, 2010.

[116] D. C. Altieri, R. Bader, P. M. Mannucci, and T. S. Edgington, “Oligospecificity

of the cellular adhesion receptor mac-1 encompasses an inducible recognition

specificity for fibrinogen,” J Cell Biol, vol. 107, pp. 1893–1900, 1988.

[117] O. Mueller, H. E. Gaub, M. Baermann, and E. Sackmann, “Viscoelastic moduli

of sterically and chemically cross-linked actin networks in the dilute to semidi-

lute regime: measurements by oscillating disk rheometer,” Macromolecules,

vol. 24, pp. 3111–3120, 1991.

[118] D. E. Discher, P. A. Janmey, and Y. li Wang, “Tissue cells feel and respond to

the stiffness of their substrate,” Science, vol. 310, pp. 1139–1143, 2005.

[119] T. Matsumoto, H. Abe, T. Ohashi, Y. Kato, and M. Sato, “Local elastic mod-

ulus of atherosclerotic lesions of rabbit thoracic aortas measured by pipette

aspiration method,” Physiol Meas, vol. 23, pp. 635–648, 2002.

[120] A. L. McKnight, J. L. Kugel, P. J. Rossman, A. Manduca, L. C. Hartmann,

and R. L. Ehman, “Mr elastography of breast cancer: preliminary results,” Am

J Roentgenol, vol. 178, pp. 1411–1417, 2002.

[121] E. M. Leise, “Modular construction of nervous systems: A basic principle of

design for invertebrates and vertebrates,” Brain Res Rev, vol. 15, pp. 1–23,

1990.

[122] N. M. L. Douarin and M.-A. M. Teillet, “Experimental analysis of the migra-

tion and differentiation of neuroblasts of the autonomic nervous system and of

neurectodermal mesenchymal derivatives, using a biological cell marking tech-

nique,” Dev Biol, vol. 41, no. 1, pp. 162 – 184, 1974.

191

Page 218: The Mechanics of Cell Motility and a Unifying Theory for

[123] B. Alberts, A. Johnson, J. Lewis, M. Raff, K. Roberts, and P. Walter, Molecular

biology of the cell, pp. 969–982. New York: Garland Science, fourth ed., 2002.

[124] V. Witko-Sarsat, P. Rieu, B. Descamps-Latscha, P. Lesavre, and L. Halbwachs-

Mecarelli, “Neutrophils: Molecules, functions and pathophysiological aspects,”

Lab Invest, vol. 80, pp. 617–653, 2000.

[125] C. I. Lacayo, Z. Pincus, M. M. VanDuijn, C. A. Wilson, D. A. Fletcher, F. B.

Gertler, A. Mogilner, and J. A. Theriot, “Emergence of large-scale cell mor-

phology and movement from local actin filament growth dynamics,” PLoS Biol,

vol. 5, no. 9, p. e233, 2007.

[126] M. Herant and M. Dembo, “Form and function in cell motility: from fibroblasts

to keratocytes,” Biophys J, vol. 98, no. 8, pp. 1408–1417, 2010.

[127] E. L. Barnhart, K.-C. Lee, K. Keren, A. Mogilner, and J. A. Theriot, “An

adhesion-dependent switch between mechanisms that determine motile cell

shape,” PLoS Biol, vol. 9, no. 5, p. e1001059, 2011.

[128] C. A. Lemmon and L. H. Romer, “A predictive model of cell traction forces

based on cell geometry,” Biophys J, vol. 99, no. 9, pp. L78–L80, 2010.

[129] A. D. Rape, W. Guo, and Y. Wang, “The regulation of traction force in relation

to cell shape and focal adhesions,” Biomaterials, vol. 32, no. 8, pp. 2043–2051,

2011.

[130] R. W. Carthew, “Adhesion proteins and the control of cell shape,” Curr Opin

Genet Dev, vol. 15, no. 4, pp. 358 – 363, 2005.

[131] M. T. Cabeen and C. Jacobs-Wagner, “Bacterial cell shape,” Nat Rev Microbiol,

vol. 3, pp. 601–610, 2005.

[132] S. Paku, J. Tovari, Z. Lorincz, F. Timar, B. Dome, L. Kopper, A. Raz, and

J. Timar, “Adhesion dynamics and cytoskeletal structure of gliding human fi-

brosarcoma cells: a hypothetical model of cell migration,” Exp Cell Res, vol. 290,

no. 2, pp. 246–253, 2003.

[133] C. Rotsch, K. Jacobson, and M. Radmacher, “Dimensional and mechanical

dynamics of active and stable edges in motile fibroblasts investigated by using

192

Page 219: The Mechanics of Cell Motility and a Unifying Theory for

atomic force microscopy,” P Natl Acad Sci USA, vol. 96, no. 3, pp. 921–926,

1999.

[134] C. W. Wolgemuth, “Lamellipodial contractions during crawling and spreading,”

Biophys J, vol. 89, no. 3, pp. 1643–1649, 2005.

[135] E. L. Barnhart, G. M. Allen, F. Julicher, and J. A. Theriot, “Bipedal locomotion

in crawling cells,” Biophys J, vol. 98, pp. 933–942, 2010.

[136] T. M. Svitkina, A. B. Verkhovsky, K. M. McQuade, and G. G. Borisy, “Analysis

of the actin-myosin ii system in fish epidermal keratocytes: mechanism of cell

body translocation,” J Cell Biol, vol. 139, no. 2, pp. 397–415, 1997.

[137] M. Dembo, T. Oliver, A. Ishihara, and K. Jacobson, “Imaging the traction

stresses exerted by locomoting cells with the elastic substratum method,” Bio-

phys J, vol. 70, pp. 2008–2022, 1996.

[138] M. Welch, A. Mallavarapu, J. Rosenblatt, and T. Mitchison, “Actin dynamics

in vivo,” Curr. Opin. Cell. Bio., vol. 9, pp. 54–61, 1997.

[139] J. A. Theriot and T. J. Mitchison, “Actin microfilament dynamics in locomoting

cells,” Nature, vol. 352, pp. 126–131, 1991.

[140] A. Mogilner and G. Oster, “Force generation by actin polymerization ii: the

elastic ratchet and tethered filaments,” Biophys J, vol. 84, pp. 1591–1605, 2003.

[141] C. Jurado, J. R. Haserick, and J. Lee, “Slipping or gripping? fluorescent speckle

microscopy in fish keratocytes reveals two different mechanisms for generating

a retrograde flow of actin,” Mol Biol Cell, vol. 16, pp. 507–518, 2005.

[142] S. Deguchi, T. Ohashi, and M. Sato, “Tensile properties of single stress fibers iso-

lated from cultured vascular smooth muscle cells,” J Biomech, vol. 39, pp. 2603–

2610, 2006.

[143] A. E. Filippov, J. Klafter, and M. Urbakh, “Friction through dynamical forma-

tion and rupture of molecular bonds,” Phys. Rev. Lett., vol. 92, p. 135503, Mar

2004.

[144] S. Walcott and S. X. Sun, “A mechanical model of actin stress fiber formation

and substrate elasticity sensing in adherent cells,” Proc Natl Acad Sci USA,

193

Page 220: The Mechanics of Cell Motility and a Unifying Theory for

vol. 107, no. 17, pp. 7757–7762, 2010.

[145] F. Wottawah, S. Schinkinger, B. Lincoln, R. Ananthakrishnan, M. Romeyke,

J. Guck, and J. Kas, “Optical rheology of biological cells,” Phys Rev Lett,

vol. 94, p. 098103, 2005.

[146] V. M. Laurent, S. Kasas, A. Yersin, T. E. Schaffer, S. Catsicas, G. Dietler,

A. B. Verkhovsky, and J.-J. Meister, “Gradient of rigidity in the lamellipodia of

migrating cells revealed by atomic force microscopy,” Biophys J, vol. 89, no. 1,

pp. 667–675, 2005.

[147] F. Heinemann, H. Doschke, and M. Radmacher, “Keratocyte lamellipodial pro-

trusion is characterized by a concave force-velocity relation,” Biophys J, vol. 100,

pp. 1420–1427, 2011.

[148] Z. Pincus and J. A. Theriot, “Comparison of quantitative methods for cell-shape

analysis,” J Microsc, vol. 227, pp. 140–156, 2007.

[149] P. Lenz, K. Keren, and J. A. Theriot, “Biophysical aspects of actin-based cell

motility in fish epithelial keratocytes,” in Cell Motility, Biological and Medical

Physics, Biomedical Engineering, pp. 31–58, Springer New York, 2008.

[150] J. A. Theriot, “Theriot lab movies.” http://cmgm.stanford.edu/theriot/

movies.htm, May 2011.

[151] A. B. Verkhovsky, O. Y. Chaga, S. Schaub, T. M. Svitkina, J.-J. Meister, and

G. G. Borisy, “Orientational order of the lamellipodial actin network as demon-

strated in living motile cells,” Mol Biol Cell, vol. 14, pp. 4667–4675, 2003.

[152] T. E. Schaus, E. W. Taylor, and G. G. Borisy, “Self-organization of actin fil-

ament orientation in the dendritic-nucleation/array-treadmilling model,” Proc

Natl Acad Sci USA, vol. 104, pp. 7086–7091, 2007.

[153] F. Fleischer, R. Ananthakrishan, S. Eckel, H. Schmidt, J. Kas, T. M. Svitk-

ina, V. Schmidt, and M. Beil, “Actin network architecture and elasticity in

lamellipodia of melanoma cells,” New J Phys, vol. 9, p. 420, 2007.

[154] S. Schaub, S. Bohnet, V. M. Laurent, J.-J. Meister, and A. B. Verkhovsky,

“Comparative maps of motion and assembly of filamentous actin and myosin ii

194

Page 221: The Mechanics of Cell Motility and a Unifying Theory for

in migrating cells,” Mol Biol Cell, vol. 18, no. 10, pp. 3723–3732, 2007.

[155] Y. Lin, “Mechanics model for actin-based motility,” Phys Rev E, vol. 79,

p. 021916, 2009.

[156] K. Larripa and A. Mogilner, “Transport of a 1d viscoelastic actinmyosin strip

of gel as a model of a crawling cell,” Physica A, vol. 372, pp. 113–123, 2006.

[157] A. Mogilner and D. W. Verzi, “A simple 1-d physical model for the crawling

nematode sperm cell,” J Stat Phys, vol. 110, pp. 1169–1189, 2003.

[158] F. Ziebert, S. Swaminathan, and I. S. Aranson, “Model for self-polarization and

motility of keratocyte fragments,” J Roy Soc Interface, vol. (online in adv of

print), pp. 1–9, 2011.

[159] J. L. McGrath, N. J. Eungdamrong, C. I. Fisher, F. Peng, L. Mahadevan, T. J.

Mitchison, and S. C. Kuo, “The force-velocity relationship for the actin-based

motility of listeria monocytogenes,” Curr Biol, vol. 13, pp. 329–332, 2003.

[160] A. F. Straight, A. Cheung, J. Limouze, I. Chen, N. J. Westwood, J. R. Sellers,

and T. J. Mitchison, “Dissecting temporal and spatial control of cytokinesis

with a myosin ii inhibitor,” Science, vol. 299, pp. 1743–1747, 2003.

[161] P. T. Yam, C. A. Wilson, L. Ji, B. Hebert, E. L. Barnhart, N. A. Dye, P. W.

Wiseman, G. Danuser, and J. A. Theriot, “Actin-myosin network reorganiza-

tion breaks symmetry at the cell rear to spontaneously initiate polarized cell

motility,” J Cell Biol, vol. 178, no. 7, pp. 1207–1221, 2007.

[162] M. L. Gardel, J. H. Shin, F. C. MacKintosh, L. Mahadevan, P. Matsudaira,

and D. A. Weitz, “Elastic behavior of cross-linked and bunded actin networks,”

Science, vol. 304, pp. 1301–1305, 2004.

[163] J. E. Italiano, T. M. Roberts, M. Stewart, and C. A. Fontana, “Reconstitution

in vitro of the motile apparatus from the amoeboid sperm of ascaris shows

that filament assembly and bundling move membranes,” Cell, vol. 84, no. 1,

pp. 105–114, 1996.

[164] A. Arnaout, “Structure and function of leukocyte adhesion molecules,” Blood,

vol. 75, pp. 1037–1050, 1990.

195

Page 222: The Mechanics of Cell Motility and a Unifying Theory for

[165] L. A. Smith, H. Aranda-Espinoza, J. B. Haun, M. Dembo, and D. A. Hammer,

“Neutrophil traction stresses are concentrated in the uropod during migration,”

Biophys J, vol. 92, pp. L58–L60, 2007.

[166] J. A. Swanson, M. T. Johnson, K. Beningo, P. Post, M. Mooseker, and N. Araki,

“A contractile activity that closes phagosomes in macrophages,” J Cell Sci,

vol. 112, pp. 307–316, 1999.

[167] M. Herant, V. Heinrich, and M. Dembo, “Mechanics of neutrophil phagocytosis:

behavior of the cortical tension,” J Cell Sci, vol. 118, pp. 1789–1797, 2005.

[168] M. Herant, V. Heinrich, and M. Dembo, “Mechanics of neutrophil phagocytosis:

experiments and quantitative models,” J Cell Sci, vol. 119, pp. 1903–1913, 2006.

[169] R. Meili, B. Alonso-Latorre, J. C. del Alamo, R. A. Firtel, and J. C. Lasheras,

“Myosin ii is essential for the spatiotemporal organization of traction forces

during cell motility,” Mol Biol Cell, vol. 21, pp. 405–417, 2010.

[170] C. E. Chan and D. J. Odde, “Traction dynamics of filopodia on compliant

substrates,” Science, vol. 322, pp. 1687–1691, 2008.

[171] A. Bacic, G. Fincher, and B. Stone, eds., Chemistry, Biochemistry, and Biology

of 1-3 Beta Glucans and Related Polysaccharides. Academic Press, 2009.

[172] G. C. Chan, W. K. Chan, and D. M. Sze, “The effects of beta-glucan on human

immune and cancer cells,” J Hematol Oncol, vol. 2, p. 25, 2009.

[173] S. Y. Kim, H. J. Song, Y. Y. Lee, K.-H. Cho, and Y. K. Roh, “Biomedical issues

of dietary fiber beta-glucan,” J Korean Med Sci, vol. 21, pp. 781–789, 2006.

[174] G. D. Brown and S. Gordon, “Fungal β-glucans and mammalian immunity,”

Immunity, vol. 19, pp. 311–315, 2003.

[175] A. Buxboim, K. Rajagopal, A. E. X. Brown, and D. E. Discher, “How deeply

cells feel: methods for thin gels,” J Phys-Condens Mat, vol. 22, p. 194116, 2011.

[176] H. Tanimoto and M. Sano, “A simple force-motion relation for migrating cells

revealed by multipole analysis of traction stress,” Biophys J, vol. 106, pp. 16–25,

2014.

196

Page 223: The Mechanics of Cell Motility and a Unifying Theory for

[177] L. Breiman and J. H. Friedma, “Estimating optimal transformations for multiple

regression and correlation,” J Am Stat Assoc, vol. 80, pp. 580–598, 1985.

[178] H. Voss and J. Kurths, “Reconstruction of nonlinear time delay models from

data by the use of optimal transformations,” Phys Rev A, vol. 234, pp. 336–344,

1997.

[179] P. W. Oakes, S. Banerjee, C. M. Marchetti, and M. L. Gardel, “Geometry

regulates traction stresses in adherent cells,” Biophys J, vol. 107, pp. 825–833,

2014.

[180] V. M. Laurent, S. Kasas, A. Yersin, T. E. Schaffer, S. Catsicas, G. Dietler,

A. B. Verkhovsky, and J.-J. Meister, “Gradient of rigidity in the lamellipodia

of migrating cells revealed by atomic force microscopy,” Biophys J, vol. 89,

pp. 667–675, 2005.

[181] O. Thoumine, P. Kocian, A. Kottelat, and J.-J. Meister, “Short-term bind-

ing of fibroblasts to fibronectin: optical tweezers experiments and probabilistic

analysis,” Eur Biophys J, vol. 29, pp. 398–408, 2000.

[182] F. Li, S. D. Redick, H. P. Erickson, and V. T. Moy, “Force measurements of the

α5β1 integrin-fibronectin interaction,” Biophys J, vol. 84, pp. 1252–1262, 2003.

[183] J. Riedl, A. H. Crevenna, K. Kessenbrock, J. H. Yu, D. Neukirchen, M. Bista,

F. Bradke, D. Jenne, T. A. Holak, Z. Werb, M. Sixt, and R. Wedlich-Soldner,

“Lifeact: a versatile marker to visualize f-actin,” Nat Methods, vol. 5, p. 605,

2008.

[184] J. A. Cooper, “Effects of cytochalasin and phalloidin on actin,” J Cell Biol,

vol. 105, pp. 1473–1478, 1987.

[185] A. J. Loosley, X. M. O’Brien, J. S. Reichner, and J. X. Tang, “Describing

directionally biased paths by a characteristic directionality time,” submitted,

vol. xx, pp. xxx–xxx, 2014.

[186] H. Levine and W.-J. Rappel, “The physics of eukaryotic chemotaxis,” Phys

Today, vol. 66, pp. 24–30, 2013.

197

Page 224: The Mechanics of Cell Motility and a Unifying Theory for

[187] K. Takeda, D. Shao, M. Adler, P. G. Charest, W. F. Loomis, H. Levine, A. Gro-

isman, W.-J. Rappel, and R. A. Firtel, “Incoherent feedforward control governs

adaptation of activated ras in a eukaryotic chemotaxis pathway,” Sci Signal,

vol. 5, p. ra2, 2012.

[188] D. M. Lehmann, A. M. P. B. Seneviratne, and A. V. Smrcka, “Small mole-

cule disruption of g protein beta gamma subunit signaling inhibits neutrophil

chemotaxis and inflammation,” Mol Pharmacol, vol. 73, pp. 410–418, 2008.

[189] M. A. Fardin, O. M. Rossier, P. Rangamani, P. D. Avigan, N. C. Gauthier,

W. Vonnegut, A. Mathur, J. Hone, R. Iyengar, and M. P. Sheetz, “Cell spreading

as a hydrodynamic process,” Soft Matter, vol. 6, pp. 4788–4799, 2010.

[190] A. Levchenko and P. Iglesias, “Models of eukaryotic gradient sensing: applica-

tion to chemotaxis of amoebae and neutrophils,” Biophys J, vol. 82, pp. 50–63,

2002.

[191] L. Michaelis and M. L. Menten, “Die kinetik der invertinwirkung,” Biochem Z,

vol. 49, pp. 333–369, 1913.

[192] A. Bruno-Alfonso, L. Cabezas-Gomez, and H. A. Navarro, “Alternate treat-

ments of jacobian singularities in polar coordinates within finite-difference

schemes,” World J Model Sim, vol. 8, pp. 163–171, 2012.

[193] I. Hecht, M. L. Skoge, P. G. Charest, E. Ben-Jacob, R. A. Firtel, W. F. Loomis,

H. Levine, and W.-J. Rappel, “Activated membrane patches guide chemotactic

cell motility,” PLoS Comp Biol, vol. 7, p. e1002044, 2011.

[194] E. Lien, T. K. Means, H. Heine, A. Yoshimura, S. Kusumoto, K. Fukase, M. J.

Fenton, M. Oikawa, N. Qureshi, B. Monks, R. W. Finberg, R. R. Ingalls, and

D. T. Golenbock, “Toll-like receptor 4 imparts ligand-specific recognition of

bacterial lipopolysaccharide,” J Clin Invest, vol. 105, pp. 497–504, 2000.

[195] R. S. Hotchkiss and I. E. Karl, “The pathophysiology and treatment of sepsis,”

N Engl J Med, vol. 348, pp. 138–150, 2003.

[196] C. G. Leon, R. Tory, J. Jia, O. Sivak, and K. M. Wasan, “Discovery and devel-

opment of toll-like receptor 4 (tlr4) antagonists: A new paradigm for treating

198

Page 225: The Mechanics of Cell Motility and a Unifying Theory for

sepsis and other diseases,” Pharm Res, vol. 25, pp. 1751–1761, 2008.

[197] H. E. Wang, N. I. Shapiro, D. C. Angus, and D. M. Yealy, “National estimates of

severe sepsis in united states emergency departments,” Crit Care Med, vol. 35,

pp. 1928–1936, 2007.

[198] E. K. Stevenson, A. R. Rubenstein, G. T. Radin, R. S. Wiener, and A. J.

Walkey, “Two decades of mortality trends among patients with severe sepsis:

A comparative meta-analysis,” Crit Care Med, vol. 42, pp. 625–631, 2014.

[199] V. Kumar, A. K. Abbas, N. Fausto, and R. Mitchell, Robbins Basic Pathology.

Elsevier/Saunders, 2007.

⊙199