29.04.2013 Views

Thesis (pdf) - Espci

Thesis (pdf) - Espci

Thesis (pdf) - Espci

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

Studies of dynamical phenomena<br />

in soft-matter and physical<br />

biology<br />

—————————————-<br />

Ph.D. thesis<br />

September 2008<br />

—————————————-<br />

Jan Brugués Ferré<br />

Ph.D. advisors:<br />

Jaume Casademunt Viader<br />

Pierre Sens<br />

Programa de doctorat de Física Avançada<br />

Bienni 2002-2004<br />

Departament d’Estructura i Constituents de la Matèria<br />

Universitat de Barcelona<br />

La Physique de la particule au solide<br />

Université Pierre et Marie Curie


Studies of dynamical phenomena<br />

in soft-matter and physical biology<br />

Ph.D. thesis, Jan Brugués Ferré<br />

Ph.D. advisors:<br />

Jaume Casademunt Viader<br />

Pierre Sens<br />

Programa de doctorat de Física Avançada<br />

Bienni 2002-2004<br />

Departament d’Estructura i Constituents de la Matèria<br />

Universitat de Barcelona<br />

La Physique de la particule au solide<br />

Université Pierre et Marie Curie<br />

September 2008


Acknowledgements<br />

This thesis has been something really hard, and has lasted probably too much.<br />

During all this time I met quite a lot of people which has somehow helped me<br />

finnish the thesis.<br />

I would first like to thank my two PhD advisors, Pierre Sens and Jaume<br />

Casademunt, who made this thesis possible.<br />

Jaume, moltes gràcies per la teva confiança i per haver-me donat l’oportunitat<br />

de fer la transició de física d’altes energies a biofísica. Gràcies per tot el que<br />

m’has ensenyat, per la teva paciència, per tots aquests bons moments, per<br />

preocupar-te per mi sempre i sobretot per la teva amistat.<br />

Pierre, it is difficult to express in words how grateful I am. To begin with,<br />

thanks for accepting me as a student, I am honored to be your first PhD student.<br />

I have learned a lot from you, both scientifically and personally. Thanks<br />

for your sense of humor and criticism. Thanks for pushing me when I needed,<br />

and being so patient with me during the writing of this thesis. I’ll never forget<br />

how much of your and your daughter time you have spent with me correcting,<br />

even if it was late in the night. Above all, thank you for being a friend, I will<br />

really miss all that time with you at ESPCI.<br />

M’agradaria també agraïr a en Quim, que va ser el meu director de tesis<br />

durant els primers anys en què vaig treballar sobre teoria de cordes. Gràcies<br />

per ser sempre sincer amb mi i pels teus bons consells, i per ser comprensiu<br />

malgrat la meva immaduresa.<br />

Thank you very much Jean-François, Guillaume, Benoit and François for<br />

the great scientific discussions and collaborations we had in Paris.<br />

Una de les persones responsables que hagi acabat fent biofísica és l’Otger,<br />

gràcies per presentar-me a en Pierre. Gràcies per la teva amistat, les converses<br />

sobre biofísica i noies, i els nostres kahales. Gràcies per estar al meu costat<br />

en els moments difícils de París. Durant tots aquests anys, moltes persones


ii Acknowledgements<br />

m’han ajudat i estat al meu costat, d’entre elles hi ha alguns dels meus millors<br />

amics i companys de feina. Gràcies Pau per ser com ets, les nostres crispis i<br />

aguantar-me en els meus moments ”pedra a la sabata”. Al Tonir, també per<br />

ser un bon amic i per tot el que em passat junts durant l’època de cordes.<br />

Gràcies també al Guillermo, al membranelo, al Carlos, al Tonim, al Minervo<br />

i tants d’altres que m’oblido però que han estat part de tots aquests anys.<br />

M’agradaria agraïr especialment a la Maria, que sempre ha estat al meu<br />

costat recolzant-me i ajudant-me en els pitjors moments, i per què la teva<br />

amistat és una de les coses més preuades i importants que tinc. Gràcies també<br />

a la Maria Sureda per ser una bona amiga i pels nostres cafés on ambdós ens<br />

hem desfugat a gust.<br />

Vull agraïr a la Julia tota la paciència infinita que ha tingut durant tots els<br />

darrers mesos de la tesi, en els quals no només m’ha ajudat i ha sacrificat tant<br />

del seu temps, sino que ha compartit amb mi part de la seva vida. Gràcies<br />

també a la Gandi.<br />

Moltes gràcies Ana Maria, per ser molt més que la meva professora de<br />

piano, pels teus bons consells, les nostres converses i per tot aquest temps<br />

en el que m’has ensenyat tantíssimes coses, entre les quals, a multiplicar el<br />

temps.<br />

Per acabar, vull agraïr molt profundament la meva família per suportarme<br />

i ajudar-me en els moments més complicats, i per, en general, ser-hi quan<br />

els he necessitat. Pels bons consells de la mimi (no n’hi ha prou en ser...) i la<br />

mami, i per estar sempre pendent que tot m’anés bé. Si he aconseguit el que<br />

he aconseguit és gràcies a vosaltres. Gràcies al nen i la nena i al meu pare per<br />

estar sempre accessibles. Gràcies al iaiu i al padrí, que per malgrat no ser-hi<br />

sempre m’han servit d’exemple a seguir. Aquesta tesi us la dedico a vosaltres.


Contents<br />

1 General introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1<br />

2 Probing elastic anisotropy from defect dynamics in Langmuir<br />

Monolayers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5<br />

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5<br />

2.2 Experimental setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />

2.2.1 The monolayer. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />

2.2.2 Brewster angle microscopy . . . . . . . . . . . . . . . . . . . . . . . 8<br />

2.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

2.4 Theoretical description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12<br />

2.4.1 Elastic free energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12<br />

2.4.2 Point defects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

2.4.3 Defect motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

2.4.4 Enhanced negative defect dissipation . . . . . . . . . . . . . . . 16<br />

2.4.5 Ratio of defect velocities . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

2.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

2.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20<br />

3 Self-organization and cooperativity of weakly coupled<br />

molecular motors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

3.1.1 Molecular motor modelization. . . . . . . . . . . . . . . . . . . . . 24<br />

3.1.2 The two state model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

3.2 Self-organization and cooperativity of weakly coupled<br />

molecular motors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

3.2.1 Simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

3.2.2 Mean-field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31


iv Contents<br />

3.2.3 Hard-core repulsion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

3.2.4 Long range interaction - transition to mean field . . . . . . 39<br />

3.2.5 Attractive interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41<br />

3.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

3.3.1 Force transmission . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

3.3.2 Motor efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45<br />

3.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

4 Membrane-cortex interactions: Micropipette experiments<br />

and blebbing cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51<br />

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52<br />

4.2 Modeling cell mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54<br />

4.2.1 Cellular membrane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54<br />

Energy cost of membrane deformation . . . . . . . . . . . . . . 54<br />

Membrane tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55<br />

4.2.2 Cytoskeletal cortex . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57<br />

Maxwell model for actin cortex viscoelasticity . . . . . . . 59<br />

Contractile stress in the cortex. . . . . . . . . . . . . . . . . . . . . 62<br />

Cytoskeletal cortex thickness . . . . . . . . . . . . . . . . . . . . . . 63<br />

Mechanical equilibrium of the cell . . . . . . . . . . . . . . . . . 65<br />

4.3 Membrane-cortex adhesion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

4.3.1 Kinetic model for membrane-cortex adhesion . . . . . . . . 67<br />

4.3.2 Micropipette experiments as membrane-cortex<br />

adhesion probes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

Influence of the link concentration . . . . . . . . . . . . . . . . . 73<br />

Effect of mysoin II activity . . . . . . . . . . . . . . . . . . . . . . . 74<br />

Adhesion energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76<br />

4.3.3 Energetic description of a bleb . . . . . . . . . . . . . . . . . . . . 80<br />

4.4 Dynamical response of a cell to a controlled mechanical<br />

perturbation using a micropipette set up . . . . . . . . . . . . . . . . . . . 86<br />

4.4.1 Micropipette experiments as membrane tension and<br />

adhesion probes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86<br />

4.4.2 Micropipette static suction pressure . . . . . . . . . . . . . . . . 91<br />

4.4.3 Effect of the loading rate . . . . . . . . . . . . . . . . . . . . . . . . . 92<br />

4.4.4 Can actin stop a moving membrane? . . . . . . . . . . . . . . . 93<br />

4.5 Experimental observation of the different regimes . . . . . . . . . . 98<br />

4.5.1 Non equivalence of initial aspiration and retraction . . . 99<br />

4.5.2 Micropipette suction regimes . . . . . . . . . . . . . . . . . . . . . . 100<br />

Saltatory and no cortex regime . . . . . . . . . . . . . . . . . . . . 101<br />

Cell oscillations in micropipette . . . . . . . . . . . . . . . . . . . 103


Contents v<br />

4.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105<br />

4.A Membrane flow dissipation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108<br />

4.A.1 Dissipation during membrane-cortex detachment . . . . . 110<br />

4.A.2 Dissipation during membrane flow inside a micropipette110<br />

5 General conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113<br />

A Resum en català . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117<br />

A.1 Auto-organització i cooperativitat de motors moleculars<br />

interactuant feblement. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117<br />

A.2 Mesura de l’anisotropia elàstica a través de la dinàmica de<br />

defectes en monocapes de Langmuir . . . . . . . . . . . . . . . . . . . . . 120<br />

A.3 Interaccions de membrana i còrtex: Experiments amb<br />

micropipeta i cèl . lules amb blebs . . . . . . . . . . . . . . . . . . . . . . . . . 123<br />

B Résumé en Français . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127<br />

B.1 Auto-organisation et coopérativité des moteurs moléculaires<br />

interagissant faiblement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127<br />

B.2 Mesure de anisotropie élastique a travers de la dynamique<br />

de défauts en monocapes de Langmuir . . . . . . . . . . . . . . . . . . . . 130<br />

B.3 Interactions de membrane et cortex : Expériences avec<br />

micropipette et cellules avec blebs . . . . . . . . . . . . . . . . . . . . . . . 133<br />

B.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137


1<br />

General introduction<br />

In recent years, research in biology has evolved towards a more quantitative<br />

analysis. The use of physical approaches have improved the understanding<br />

of many biological processes including membrane and cytoskeleton interactions,<br />

cell division, mechanotransduction and cell movement. In vitro approaches<br />

which try to identify the minimum amount of components needed<br />

to reproduce particular phenomena in living systems, are also an important<br />

tool for biological research.<br />

The study of biology is so vast that spans several orders of magnitude in<br />

both force and length. Typical length scales range from a few nanometers for<br />

a molecular motor, to a few meters corresponding to the size of a mammalian.<br />

Forces range from a few piconewtons 1 (stall force of a molecular motor), to<br />

several newtons (earth attraction on a human being is of about ∼ 500 N).<br />

We will focus on the cellular and sub-cellular scale. A cell has a typical<br />

diameter of 10 µm, and contains many organelles and proteins, including<br />

molecular motors, which are in the range of a few nanometers, capable of generating<br />

forces of piconewtons, and they are the smallest entities that we will<br />

consider. Our unit scale of energy is thus ∼ pN × nm, which is of the order<br />

of the thermal energy kBT ∼ 4 pN nm (the energy released by hydrolization<br />

of ATP is of about 8 kBT ). As a consequence we expect that thermal fluctuations<br />

play an important role at these microscopic scales. Moreover, living<br />

organisms continuously consume energy (ATP), which maintains them in an<br />

out of equilibrium steady state. Thermal equilibrium state corresponding to<br />

the death of the organism. Any physical approach at this scale must explicitly<br />

take into account these fluctuations and can not be described in terms of ener-<br />

1 1 pN = 10 −12 N


2 1 General introduction<br />

gies. In this context, out of equilibrium statistical physics is a good candidate<br />

in trying to model processes at this microscopic scale.<br />

At a more coarse grained level, the cell can be characterized by its shape<br />

and mechanical properties. Traditionally, the physics of the state of the matter<br />

(condensed matter), dealt with the macroscopic states of the matter that<br />

arise from the physical properties of the microscopic constituents (atoms or<br />

molecules) namely, solid and liquid. However, the state of cells and other<br />

materials such as polymers, colloids, foams or gels, do not correspond either<br />

to liquid or solid. They have a large degree of ordering and are definitely<br />

not simple liquids. Furthermore, they are fairly soft materials that can be deformed<br />

by thermal fluctuations. The ”softness” of these materials lead to a<br />

recent field in physics named soft matter. The typical energy scale for cell<br />

membrane deformation is of κ ∼ 10 − 100kBT , which is slightly larger than<br />

thermal fluctuations. Soft-matter is a good candidate to study coarse grained<br />

macroscopic properties of the cell where thermal fluctuations are averaged<br />

although play a role in determining shape and mechanical properties.<br />

At cellular scales, inertial forces are negligible when compared to viscous<br />

forces (the Reynolds number Re ≪ 1), and Aristotelian physics, where velocity<br />

and force are linearly related, govern generically cellular and sub-cellular<br />

phenomena (molecular motor motion, intracellular transport, cell deformation).<br />

In recent years, many qualitative advances in experimental techniques,<br />

ranging from microscopy, micromanipulation to gene transfection, have led<br />

to an enormous quantity of new data and observations at the cellular level<br />

which require fundamentally new concepts to help their understanding. In<br />

this context, physicists and biologists live an exciting period where physical<br />

concepts can help to describe and understand new general cellular properties,<br />

which at the same time can be directly accessed experimentally. Soft-matter<br />

and statistical physics are the obvious candidates to bridge between biology<br />

and physics, although new physical and biological tools will probably emerge<br />

as new phenomena is observed. My personal point of view is that the present<br />

state in biophysical research is qualitative similar to that in the early years of<br />

particle physics, where the first particle accelerators led to the discovering of<br />

new fundamental particles.<br />

The large amount of interesting and fundamental problems to be understood,<br />

the ease of experimental comparison, the beauty to relate simple concepts<br />

such as geometry, topology, or elasticity to biological phenomena, and<br />

the personal and scientific enrichment of the interdisciplinary work involved,


1 General introduction 3<br />

led me, irremediably, to devote this thesis and my future research to biophysics.<br />

This thesis do not respond to a pre-established scheme, although there is a<br />

logic underlying the pattern followed, which starts from a string theory PhD<br />

student during the first two years of my doctorate, and ends as a biophysicist.<br />

The transition between such a perpendicular fields has resulted in the following<br />

thesis, where aspects of purely soft matter concerning topological defects<br />

appear first, followed by a study of some cooperativity of molecular motors,<br />

and ends with a purely biophysical approach of membrane-cytoskeleton interactions<br />

which explicitly connects with biological experiments. Yet from<br />

the beginning I tried to work on problems that allowed me to interact with<br />

experiments and researchers from other fields.<br />

This thesis is composed of three chapters which are devoted to soft-matter<br />

and physical biology. The first chapter belongs to the field of topological defects.<br />

The study of topological defects is of general relevance in many areas of<br />

physics and biology, because their main universal properties can help understand<br />

systems that presumably exhibit topological defects, such as the mitotic<br />

spindle during cell division, using simple condensed matter experiments, like<br />

in liquid crystals. We consider the motion of defects and their interaction,<br />

which is not fully understood. In this regard, we study the dynamics of annihilation<br />

of two defects in Langmuir monolayers using a liquid crystal model.<br />

We are able to extract the dependence of the elastic anisotropy of the material<br />

on the surface pressure measuring only the ratio of velocities at which the<br />

defects approach each other.<br />

In the second chapter of the thesis, we consider some aspects of selforganization<br />

and cooperativity of molecular motors. Molecular motors are<br />

proteins that convert chemical energy into mechanical work at a molecular<br />

scale, and are responsible of many biological phenomena, ranging from muscle<br />

contraction to cellular transport and cell motility. Molecular motors usually<br />

cooperate to generate large forces, as in the case of muscle contraction.<br />

Although there are several theoretical works considering the coupling between<br />

molecular motors, there is not yet a full understanding of cooperativity<br />

and force distribution among clusters of molecular motors. We consider several<br />

representative weakly interactions between motors, namely, short range<br />

and long range repulsion, and weak attraction, and we study their collective<br />

behavior, force distribution and efficiency when a force is applied in the foremost<br />

motor. Our approach is mainly based on Langevin simulations using a<br />

two state model for a molecular motor.


4 1 General introduction<br />

Finally, in the third chapter of the thesis we consider the interaction between<br />

the cellular membrane and the cytoskeletal cortex. The plasma membrane<br />

and the cytoskeletal cortex are amongst the most important parts of the<br />

cell. They participate in providing cell shape, separate the extracellular media<br />

from the cytosol, act as mechanotransducers, and are responsible for cell motion.<br />

The adhesion between cortex and membrane is thus fundamental for the<br />

proper functionality of a cell. Failure or weakening of this adhesion induces<br />

a phenomenon named blebbing, which consist of highly dynamical spherical<br />

protrusions that nucleate as a consequence of membrane unbinding and<br />

retract with the appearance of a new cortex under the bare membrane. The<br />

properties of bleb formation and membrane-cortex adhesion can be experimentally<br />

addressed using mainly micropipette aspiration. We propose both<br />

a kinetic and energetic model for cortex and membrane adhesion which we<br />

use to understand the dependence of the adhesion strength on the active state<br />

of the cell and external perturbations such as osmotic shocks or micropipette<br />

aspirations. Once addressed the concept of adhesion energy, we consider the<br />

dynamical response of the membrane and cortex to a mechanical perturbation.<br />

In particular, we focus on the treadmilling properties of the cortex and<br />

its viscoelastic response using a Maxwell like description for the short time<br />

elastic and long time viscous behavior of the cortex. Using the micropipette<br />

aspiration set up as a model system, we theoretically identify all the possible<br />

aspiration regimes and we compare them with both existing experimental<br />

measurements, and experiments we have conducted on the Entamoeba histolytica<br />

system in collaboration with B. Mauguis and F. Amblart.


2<br />

Probing elastic anisotropy from defect dynamics in<br />

Langmuir Monolayers<br />

2.1 Introduction<br />

The study of topological defects is of general relevance in many areas of<br />

physics and biology, ranging from cosmology to superfluid helium or cell<br />

division. An important motivation arises from the universality of the defects,<br />

which occur in any system with order parameter. Their main properties are<br />

independent of the underlying physics, determined only by symmetries, the<br />

dimension of the order parameter and the defect charge. This universality has<br />

been lately exploited by condensed matter systems such as liquid crystals<br />

which allow to perform relatively easy experiments yielding knowledge in<br />

completely different physical areas.<br />

Of particular interest is the study of the motion of a defect, and the interaction<br />

and annihilation dynamics of defect pairs, which have been extensively<br />

studied by several groups (Ryskin and Kremenetsky, 1991; Svenˇsek<br />

and ˇZumer, 2002; Blanc et al., 2005). A remarkable feature of defect pairs annihilation,<br />

both observed experimentally and numerically, although not fully<br />

understood, is the enhanced mobility of the positively charged defect with respect<br />

to its negative counterpart. Elastic anisotropy (inequality of the elastic<br />

constants) and hydrodynamic effects arising from defect motion (backflow)<br />

have been shown to generically contribute to explain this asymmetry. In fact,<br />

the latter is dominant in the context of bulk (3-d) liquid crystals (Tóth et al.,<br />

2002; Svenˇsek and ˇZumer, 2002; Oswald and Ignés-Mullol, 2005; Blanc<br />

et al., 2005), thus hindering the possibility to develop a simple method to<br />

quantitatively relate material elasticity to defect dynamics, which would be<br />

an interesting alternative to traditional methods to determine the elastic constants<br />

(de Gennes and Prost, 1995; Oswald and Ignés-Mullol, 2005).


6 2 Probing elastic anisotropy from defect dynamics in Langmuir Monolayers<br />

We will study the opposite scenario by studying defect dynamics in Langmuir<br />

monolayers spread at the air/water interface (Hatta and Fischer, 2003),<br />

where the hydrodynamic effects can be neglected. Contrarily to bulk (3-d)<br />

liquid crystals, where rotational and translational viscosities are of the same<br />

order of magnitude, local molecular rotation without hydrodynamic effects is<br />

possible in Langmuir monolayers (2-d). Then, defect dynamics can be simply<br />

explained by the effect of material elasticity. As a result, measurements<br />

of differential defect mobilities enable us to probe the elastic anisotropy of<br />

the two-dimensional material.<br />

In this chapter, we will report on quantitative studies of defect dynamics in<br />

Langmuir monolayers with a polar nematic arrangement. The observed asymmetry<br />

of defect mobility will be explained only by topological properties of<br />

the defects, and a quantitative analysis in terms of a simple theoretical model<br />

that rules out backflow effects will be presented. The model is exploited to<br />

extract the dependence of the elastic anisotropy on surface pressure and to estimate<br />

the compressibility-dependent elastic constants of the 2-d system. This<br />

indirect procedure is demonstrated to be an advantageous alternative over direct<br />

measurements, unpractical in these quasi-two-dimensional systems.<br />

The results reported in this chapter correspond to the reference (Brugues<br />

et al., 2008a).<br />

2.2 Experimental setup<br />

Experiments are performed in a shallow thermostated Teflon cuvette where<br />

a monolayer is spread over an area delimited by two mobile barriers, which<br />

allows to control the surface pressure, Fig. 2.1.<br />

2.2.1 The monolayer<br />

The monolayer is composed of the photosensitive azobenzene amphiphile<br />

8Az3COOH. All classes of azobenzenes are composed of a common core<br />

of two phenyl rings linked by a N = N double bond 1 that can adopt two<br />

different geometrical configurations: the trans configuration, where the two<br />

phenyl rings are oriented in opposing directions, Fig. 2.2a (left), and the cis<br />

configuration, where the two phenyl rings are oriented in the same direction,<br />

Fig. 2.2a (right). The trans isomer exhibits an orientational order parameter<br />

due to its geometrical and elecrical properties.<br />

1 Therefore, the molecule cannot rotate in the double bond axis.


(a) (b)<br />

Π<br />

CCD CCD<br />

Brewster Angle<br />

Microscopy<br />

Π<br />

Chloroform<br />

solution<br />

Π<br />

2.2 Experimental setup 7<br />

Fig. 2.1. (a) Schematics of the experimental setup. Spread monolayers of the azobenzene amphiphile<br />

8Az3COOH are prepared by depositing drops of a chloroform solution on a surface<br />

of pure water contained in a Teflon cuvette. A lateral pressure Π is applied by adjusting the<br />

surface area. Orientational order inside the trans droplets is measured by a Brewster angle<br />

microscope (BAM) (Ignés-Mullol et al., 2004). (b) Images obtained using the experimental<br />

setup courtesy of Jordi Ignés-Mullol.<br />

Azobenzenes exhibit two main properties that can be exploited for our experimental<br />

purposes. On one hand, transitions between trans and cis isoforms<br />

can be induced by exposure to appropriate wavelengths of light (photoisomerization),<br />

Fig. 2.2a, and by thermal relaxation towards the trans configuration<br />

which has a favorable energetic configuration, Fig. 2.2b. Due to the dramatically<br />

different geometrical properties of the trans and cis conformations, mixtures<br />

of both isomers organize differently at the air/water interface. On the<br />

other hand, mesophases of the trans isomer, appear at relatively low area densities.<br />

As a consequence, monolayers composed of these azobenzenes exhibit<br />

long-range orientational order from the trans isoform, but positional disorder<br />

(Tabe and Yokoyama, 1994), which allow topological defects to easily<br />

move.<br />

(a) (b)<br />

N<br />

N<br />

hν<br />

Energy<br />

N<br />

hν kBT<br />

N<br />

′ , N<br />

N<br />

N N<br />

Configuration<br />

Fig. 2.2. Azobenzene photoisomerization. Transition between trans form (left) and cis form<br />

(right) can be induced using appropriate wavelengths of light (a). Alternatively, the cis configuration<br />

will thermally relax to the energetically stable trans configuration (b).


8 2 Probing elastic anisotropy from defect dynamics in Langmuir Monolayers<br />

Nonirradiated 8Az3COOH trans isomer monolayers exhibit complicated<br />

stripe like patterns (Crusats et al., 2004), Fig. 2.3a. If the azobenzene solution<br />

is exposed to room light, a mixture of cis and trans isomers is obtained which,<br />

after spreading a monolayer, is organized in circular droplets of a birefringent<br />

trans phase with a central topological defect embedded in an isotropic cis<br />

phase (Ignés-Mullol et al., 2005), Fig. 2.3b. We will use spread monolayers<br />

of trans and cis isomers to study annihilation of topological defects that arise<br />

from the fusion of a pair of droplets (see below).<br />

(a) (b)<br />

Fig. 2.3. (a) Nonirradiated textures of 8Az3COOH trans isomer monolayers, image modified<br />

from (Crusats et al., 2004). (b) Circular droplet of a trans isomer embedded in a cis phase.<br />

Both images are obtained using Brewster angle microscopy (BAM).<br />

Imaging of thin films at a gas/liquid interface can be achieved by several<br />

techniques (for instance fluorescence microscopy and AFM). Many of<br />

these techniques perturb the system, either by probe compounds (such as fluorescent<br />

dye) or by direct interaction (AFM). A non invasive technique that<br />

allows to study such systems without perturbing its original state is based on<br />

the Brewster angle, Fig. 2.4.<br />

2.2.2 Brewster angle microscopy<br />

For a beam of a p-polarized light (parallel to the plane of incidence), there<br />

is an angle of incidence θ (Brewster angle) at which no reflection occurs,<br />

Fig. 2.4 (left). When a thin film is placed between the two phases, the optical<br />

properties of the system change, and a small amount of light is reflected, Fig.<br />

2.4 (right). In general, the polarization of the reflected light differs from the<br />

incident light.<br />

A Brewster angle microscope (BAM), is built using a source of p-polarized<br />

light (a laser and a polarizer) and a receptor (a CCD camera and a polarizer),<br />

Fig. 2.1a. We use a calibration of the gray-scale distribution in the BAM im-


θ θ<br />

air<br />

thin film<br />

subphase<br />

Fig. 2.4. Scheme of the Brewster angle.<br />

2.3 Results 9<br />

ages and the orientational parameter as reported in (Ignés-Mullol et al., 2004),<br />

Fig. 2.5.<br />

Grey level<br />

φ(deg)<br />

Fig. 2.5. Gray level as a fucntion of the polar angle for a droplet of trans isomer in a matrix of<br />

cis isomer, image modified from (Crusats et al., 2004).<br />

2.3 Results<br />

For a wide range of experimental conditions, the organization of the amphiphilic<br />

molecules inside the circular domains is characterized by a uniform<br />

tilt (around 45 ◦ (Crusats et al., 2004)) with respect to the air-water interface<br />

normal, Fig. 2.6. Molecular ordering can thus be described by a twodimensional<br />

vector field n. Constant-angle anchoring at the droplet boundary<br />

(Crusats et al., 2004; Reigada et al., 2004) results in the inclusion of inner<br />

point defects of total charge +1 (Mermin, 1979; Tabe et al., 1999). Small<br />

enough droplets feature a single s =+1 defect near the center (in the core


10 2 Probing elastic anisotropy from defect dynamics in Langmuir Monolayers<br />

region of these singularities, the molecular tilt becomes normal to the interface<br />

(Tabe et al., 1999)), and a pure bend texture around the core (see Fig.<br />

2.8). Since the system is not chiral, equal amount of droplets with clockwise<br />

and counter-clockwise bending textures are present in the monolayer.<br />

(a) (b)<br />

θ<br />

x<br />

z<br />

φ n<br />

y<br />

Fig. 2.6. (a) Constant tilt of the amphiphilic molecules with respect to the air-water interface<br />

normal (z). The orientational order of the domains is described by the director field n, defined<br />

as the projection of the molecule on the monolayer plane (xy). (b) Director field on the<br />

monolayer plane (red) of a droplet with a positive +1 defect.<br />

Coalescence of droplets with the same chirality is hindered and seldom<br />

observed. On the other hand, droplets with opposed chirality fuse abruptly<br />

(Fig. 2.8), creating a pair of new point defects upon intersection. Fusion of<br />

droplets with the same chirality is energetically unfavorable since involves the<br />

creation of a line defect consisting of antiparallel arrangements of molecules<br />

in the contact line between droplets. The energy barrier to overcome scales<br />

then with the size of the droplet, the proportionality constant being the elastic<br />

splay constant of the material. For droplets with opposite chirality, the contact<br />

line between droplets have parallel director fields, Fig. 2.7.<br />

(a) (b)<br />

Fig. 2.7. Coalescence of two droplets of (a) opposite chirality and (b) same chirality.


2.3 Results 11<br />

Upon droplet fusion, a pair of defects appear at the surface of the newly<br />

formed droplet. Since the total charge must be preserved inside the closed domain,<br />

we consistently assign a −1/2 charge to each boundary defect. These<br />

semi-integer defects are necessarily confined to the droplet boundary since<br />

they are not allowed inside polar nematic domains. In the simplest route to<br />

collapse, typical in the fusion of droplets of different size, one of the originally<br />

central s =+1 defects is annihilated in a two-step mechanism, Fig.<br />

2.8A. First it migrates to merge at the boundary with one of the newly created<br />

singularities, resulting in a +1/2 boundary defect. The latter, and the<br />

remaining −1/2 boundary defect approach following an asymmetric dynamics<br />

along the contour of the fused droplet, with +1/2 defects always moving<br />

faster (see Fig. 2.8B for a representation of defect trajectories). Relaxation of<br />

the droplet shape following coalescence is much faster than defect dynamics.<br />

As a result, droplet circularity does not change measurably during the final<br />

stages of defect collapse, where a roughly constant linear velocity along the<br />

curved boundary is observed. The ratio of linear velocities in this regime measured<br />

for different surface pressures is used below to quantify the asymmetry<br />

in defect mobilities.<br />

-30 -20 -10 0<br />

time (s)<br />

! (C)<br />

!<br />

!<br />

" !<br />

!<br />

Fig. 2.8. (A) Experimental coalescence process of a clockwise and an anti-clockwise " bend<br />

domain at T = 32◦C and Π = 0.3mN m−1 leading to a droplet with a single s =+1 defect.<br />

BAM analyzer is set at 60◦ counterclockwise from the plane of incidence, which includes the<br />

vertical axis in the images. A sketch of the molecular field is shown below each experimental<br />

image. Merging of one of the inner s =+1 defects and one of the two −1/2 boundary defects<br />

created upon fusion (b) results in a ±1/2 pair of boundary defects that attract and annihilate<br />

(d) following the arclengths towards their meeting point shown in (B). The straight lines yield<br />

the average linear velocity in the pseudo-constant velocity regime prior to collapse. Elapsed<br />

times from panel (b) are 90 s (c) and 124.6 s (d). The ruler is 100µm long. Video image<br />

digitization and processing (Rasband, 2006) is used to set the quasi-circular droplet contour<br />

as fixed in space.<br />

L ,L<br />

+ -(µm)<br />

20<br />

0<br />

-20<br />

-40<br />

-60<br />

-80<br />

(B)


12 2 Probing elastic anisotropy from defect dynamics in Langmuir Monolayers<br />

2.4 Theoretical description<br />

The mutual elastic attraction between defects of opposite charge is the responsible<br />

of the annihilation of the boundary ±1/2 defects experimentally observed.<br />

The velocity at which the defects annihilate depends on the amount of<br />

energy dissipated as they approach each other. The defect movement involves<br />

a translational flow of the polar molecules along the direction of motion and<br />

reorientation of the director field through rotation of the polar molecules. The<br />

dissipation has then two main contributions that arise from backflow and the<br />

dissipation during molecule reorientation.<br />

Rotational viscosity in our system is of about 10 −10 kg s −1 (Feder et al.,<br />

1997). The effective translational viscosity results from the coupling between<br />

the monolayer and the subphase (Stone, 1995). The subphase has a viscosity<br />

of the order of 10 −3 kg m −1 s −1 and the dissipation occurs in a length scale<br />

of the order of the system, i.e., the droplet radius (∼ 50 µm), leading to an<br />

effective translational viscosity two orders of magnitude larger than the rotational<br />

viscosity. Recent numerical analysis (Svenˇsek and ˇZumer, 2002) has<br />

shown that the effect of elastic anisotropy and backflow cannot be decoupled<br />

in the study of asymmetric defect mobilities in bulk liquid crystals. Contrary<br />

to bulk liquid crystals, where rotational and translational viscosities are of<br />

the same order of magnitude, our system exhibits a larger effective translational<br />

viscosity with respect to the associated to molecular reorientations. As<br />

a consequence, the motion of the defect involves only reorientations of the<br />

director field, which allows to study the effects of the elastic properties of the<br />

material on the defect motion alone. Actually, it has recently been shown that<br />

complex reorientations of the molecular field can take place in this system<br />

without noticeable hydrodynamic effects (Burriel et al., 2006).<br />

2.4.1 Elastic free energy<br />

We have seen in the experimental section that for a wide range of experimental<br />

conditions, the organization of the amphiphilic molecules inside the<br />

circular confined domains is characterized by a uniform tilt with respect to the<br />

air-water interface normal (see Fig.2.6). Consequently, we describe our system<br />

with a two-dimensional director field n, and a generic elastic free energy<br />

that contains the lowest order in the director field distortions,<br />

<br />

F = d 2 x<br />

KS<br />

2 (∇ · n)2 + KB<br />

(∇ × n)2<br />

2<br />

<br />

, (2.1)<br />

where KS and KB are splay and bend constants (de Gennes and Prost, 1995;<br />

Oswald and Ignés-Mullol, 2005). Earlier studies on this system suggest that


2.4 Theoretical description 13<br />

the effect of anisotropic boundary conditions can be neglected with respect to<br />

inner elasticity in the regimes where pure bend textures are obtained (Ignés-<br />

Mullol et al., 2005). More generally, Eq. 2.1 should include contributions<br />

from a density order parameter so as to account for the finite compressibility<br />

of this system (Hinshaw et al., 1988; Tabe et al., 1999). Nevertheless, at a<br />

fixed lateral pressure, explicit density contributions can be renormalized into<br />

effective elastic constants and Eq. 2.1 is of general applicability, keeping in<br />

mind that the value of the elastic constants will depend on the applied pressure.<br />

a b<br />

v−<br />

Fig. 2.9. The motion of a negative defect along the curved boundary (a) can be mapped into<br />

that of a defect in rectilinear motion (b). This mapping is justified because the director field<br />

around ±1/2 boundary defects rotates when defects move along the curved boundary. In (a),<br />

we see a sketch of a −1/2 defect with a virtual −1/2 part outside the boundary (in dashed<br />

lines) moving at a velocity v− and its director field is rotated accordingly to its motion (the<br />

director normal to the boundary is preserved as the defect moves as indicated with the lines<br />

pointing to the center of the droplet). In addition, the effect of the curvature of the droplet on<br />

the velocities is negligible in our experimental conditions (droplet radii considered in the range<br />

30-60 µm ). In (b), a −1 defect is decomposed in a −1/2 defect and a virtual −1/2 defect in<br />

dashed lines. This corresponds to having a −1/2 defect in the boundary and consequently, the<br />

results of rectilinear ±1 defect motion can be extrapolated to that of semi-integer boundary<br />

defects.<br />

In the experimental section, we observed topological defects of charges<br />

±1 in the bulk and ±1/2 on the boundary, Fig. 2.8. Fractional defects in a<br />

nematic order parameter can not exist on the bulk by topological constraints<br />

and must be confined on the boundary. Fig. 2.8A shows that the director field<br />

around ±1/2 defects rotates when defects move along the curved boundary,<br />

that is, the angle between the ±1/2 director field and the tangent to the<br />

boundary is kept constant as the defect moves, Fig. 2.9a. Therefore, the director<br />

field around a defect moving along the boundary can be mapped into that<br />

of a defect in rectilinear motion. On the other hand, each boundary defect can<br />

be regarded as a ±1 with half its spatial extend being virtual. This allows to


14 2 Probing elastic anisotropy from defect dynamics in Langmuir Monolayers<br />

extrapolate the results of rectilinear ±1 defect motion to that of semi-integer<br />

boundary defects, Fig. 2.9b.<br />

2.4.2 Point defects<br />

Let us consider an isolated point defect in the director field of charge s and<br />

rotational symmetry. We use polar coordinates (r,φ) and express the director<br />

field using the angle ψ with respect to the radial direction, Fig. 2.10, nr =<br />

cosψ and nφ = sinψ. Minimization of the free energy (Eq. 2.1) gives,<br />

(1 + α cos2ψ) ∂ 2ψ = α sin2ψ<br />

∂φ2 ∂ψ<br />

∂φ<br />

2 <br />

− 1 , (2.2)<br />

where α ≡ (KS −KB)/(KS +KB) is a measure of the elastic anisotropy. α < 0<br />

(resp. > 0) favors splay (bend) textures, and α = 0 is the isotropic case.<br />

y<br />

r<br />

n ψ<br />

Ψ<br />

Fig. 2.10. Definition of the director field angular coordinates.<br />

(a) (b) (c) ψ<br />

Fig. 2.11. Solutions for s =+1 defects corresponding to the energetically favorable solution<br />

to equation 2.2 for (a), α < 0, which corresponds to a pure splay configuration, in this case<br />

ψ = 0. (b), α > 0, which corresponds to a pure bend configuration, ψ = π/2. And (c), α = 0,<br />

where the energetic cost of splay and bend configurations weight the same. The solution is<br />

a spiral, ψ = ψ0, where ψ0 ∈ (0,2π), that combines splay and bend contributions. In (c) the<br />

angle ψ between the director and the polar direction is sketched.<br />

φ<br />

x


2.4 Theoretical description 15<br />

For s =+1, the energetically favorable solution to Eq. 2.2 for α > 0 is<br />

ψ+ = ±π/2 (pure bend configuration), and ψ+ = 0,π for α < 0 (pure splay),<br />

Fig. 2.11. For s = −1, the director field around the defect is given in implicit<br />

form by (Landau et al., 1995)<br />

φ[ψ−,k−] ≡ k−<br />

ψ−<br />

0<br />

<br />

1 + α cos2x<br />

1 + αk2 1/2 dx, (2.3)<br />

− cos2x<br />

where the integration constant k− is determined by φ[−4π,k−]=2π, which<br />

is the topological condition for the director field of a −1 defect in polar coordinates,<br />

ψ[φ + 2π] =ψ[φ] − 4π, for φ = 0. Since a negative defect is a<br />

mixture of splay and bend configurations, for α = 0, the defect is deformed<br />

with respect the corresponding isotropic one, minimizing the splay or bend<br />

domain depending on the sign of the anisotropy, Fig. 2.12.<br />

(a) (b) 0<br />

(c)<br />

ψ<br />

-1<br />

-2<br />

-3<br />

-4<br />

-5<br />

-6<br />

0 0.5 1 1.5 2 2.5 3<br />

Fig. 2.12. Deformation of a −1 defect. (a) Negative defect for isotropic elastic constants. (b)<br />

The director angle ψ as a function of the polar angle φ, for isotropic elastic constants (dashed<br />

line) and for an anisotropy α = 0.9 (solid line). (c) Negative defect for an anisotropy α = 0.9.<br />

2.4.3 Defect motion<br />

During defect motion, the director field is distorted. We have shown that back<br />

flows are negligible in our system and consequently, in a quasistatic approximation,<br />

we can assume that Eq. 2.3 holds during motion and describes the<br />

instantaneous director field during rectilinear defect motion at velocity v.<br />

The dissipation rate per unit length associate to rotations of the director field<br />

is (Kleman and Laverntovich, 2003)<br />

<br />

Σ = γ<br />

dS<br />

φ<br />

2 ∂Ψ<br />

, (2.4)<br />

∂t


16 2 Probing elastic anisotropy from defect dynamics in Langmuir Monolayers<br />

where Ψ is the angle between the director field and the polar axis, Ψ = ψ +φ<br />

(see Fig. 2.10), and γ is the rotational viscosity of the medium. Because of<br />

the symmetry in the solutions of Eq. 2.3, the dissipation can be decomposed<br />

in the product of a radial and angular contribution, being this last part the<br />

responsible for differences in dissipation for s = ±1 defects,<br />

Σ = γv 2 <br />

R<br />

log<br />

rc<br />

2π<br />

0<br />

dφ sin 2 (φ − φ0)<br />

2 ∂Ψ<br />

, (2.5)<br />

∂φ<br />

where φ0 is the direction of movement, rc is the core radius of the defect and<br />

R is the system size, which in this case is the droplet radius. Regardless of the<br />

accuracy of the logarithmic dependence in the drag force expression (Ryskin<br />

and Kremenetsky, 1991), what is relevant for our analysis of the relative velocities<br />

(see below) is the fact that R and rc can be assumed to be roughly the<br />

same for the two defects of opposite charges 2 .<br />

Using the trigonometric equality sin 2 (φ − φ0) =1/2(1 − cos2(φ − φ0))<br />

and the defect ±1 symmetry, we have that (∂Ψ/∂φ) 2 [φ +π/2]=(∂Ψ/∂φ) 2 [φ]<br />

and consequently, the dissipation is independent of the direction of motion,<br />

and takes the simple form<br />

Σ = γv 2 <br />

R<br />

log<br />

rc<br />

<br />

1 2π<br />

dφ<br />

2 0<br />

2.4.4 Enhanced negative defect dissipation<br />

2 ∂Ψ<br />

. (2.6)<br />

∂φ<br />

The dissipation is proportional to the sum of the rotation squared of each polar<br />

molecule over the spatial extension of the defect. As a consequence, for<br />

any global fixed rotation of a defect, the least dissipative way to rotate each<br />

polar molecule would be, if possible, a uniform rotation. A uniform rotation<br />

of each polar molecule means a totally symmetric defect, which in the case<br />

of ±1 defects with elastic anisotropy, corresponds to the +1 defect. Since<br />

the negative defect is a mixture of bend and splay configurations, the elastic<br />

anisotropy introduces an inhomogeneous distortion on the director field and<br />

as a consequence the defect is squashed in the splay or bend domain depending<br />

on the sign of α (see Fig. 2.12). A global fixed rotation of the negative<br />

defect carries a local nonuniform rotation of the director field in the regions<br />

where the splay or bend has been increased or decreased. As a consequence,<br />

the dissipation will generically be larger for negative defects.<br />

2 More accurate estimations of those parameters result in a residual sublogarithmic dependence<br />

on the elastic anisotropy which can be neglected.


2.4 Theoretical description 17<br />

To prove this statement more rigorously, we consider the angular part<br />

of the dissipation, 2π<br />

0 dφ (∂Ψ/∂φ)2 . Recalling the definition of the defect<br />

charge,<br />

2π<br />

0<br />

dφ ∂Ψ<br />

∂φ<br />

= Ψ(2π) −Ψ(0) ≡ 2πs, (2.7)<br />

we show that the overall absolute value of the rotation of the director field<br />

Ψ in both positive and negative defects is the same. In the case of a positive<br />

defect, the derivative of the director field with respect to the angular direction<br />

is equal to 1 independently of the anisotropy, whereas in the negative case,<br />

it is a function of the angular direction and the anisotropy (Eq. 2.3). As a<br />

consequence,<br />

2π<br />

0<br />

dφ<br />

2 ∂Ψ+<br />

=<br />

∂φ<br />

1<br />

2π<br />

dφ<br />

2π 0<br />

∂Ψ±<br />

2 2π<br />

≤ dφ<br />

∂φ 0<br />

2 ∂Ψ−<br />

, (2.8)<br />

∂φ<br />

where in the last step we have used a general inequality of functional analysis<br />

(the continuous version of the discrete triangular inequality). The immediate<br />

implication of this result is that, due to elastic anisotropy, isolated positive<br />

defects are always faster than negative ones (independently of the anisotropy<br />

sign).<br />

2.4.5 Ratio of defect velocities<br />

The power supplied per unit length needed to maintain the defect moving at<br />

a fixed velocity v is P = fv, where f is the drag force, and must be equal to<br />

the dissipation rate Σ. This leads to the expression<br />

f− 1<br />

2 γv−<br />

⎡<br />

<br />

R 2π<br />

log dφ ⎣1+ 1<br />

<br />

1+αk2 ⎤2<br />

− cos2ψ− ⎦ ,<br />

rc<br />

0<br />

k−<br />

rc<br />

1+α cos2ψ−<br />

<br />

R<br />

f+ πγv+ log , (2.9)<br />

for the negative and positive defect respectively. Although the attractive force<br />

between defects is of elastic origin and can be in general a complicated function<br />

of the position, the anisotropy α, and the viscosity of the material can<br />

vary as a function of the lateral pressure, to compare the ratio between velocities<br />

we only need to know that both the defects feel the same force f<br />

in opposite directions (Newton’s third law), and the same material viscosity.<br />

Using the expression for the drag force in Eq. 2.9 for the +1 and −1 defects,


18 2 Probing elastic anisotropy from defect dynamics in Langmuir Monolayers<br />

v+<br />

=<br />

v−<br />

1<br />

⎡<br />

2π<br />

dφ ⎣1 +<br />

2π 0<br />

1<br />

<br />

1 + αk<br />

k−<br />

2 ⎤2<br />

− cos2ψ− ⎦ .<br />

1 + α cos2ψ−<br />

(2.10)<br />

As seen in Fig. 2.13, there is a wide range of anisotropy α for which the<br />

ratio of velocities is roughly constant, and the asymmetry is only significant<br />

for |α| tending to 1. The ratio v+/v− is invariant if KB and KS are interchanged<br />

(Svenˇsek and ˇZumer, 2002), i.e., if α us replaced by −α, since the<br />

change α →−α results only in a rotation of π/2 of the solution of Eq. 2.2.<br />

Thus the overall rotation of n due to defect motion is the same, leaving invariant<br />

the integrand in Eq. 2.10.<br />

2.5 Discussion<br />

v−/v+<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

|α|<br />

Fig. 2.13. Plot of the solution of Eq. 2.10.<br />

We can now use Eq. 2.10 to determine the ratio KS/KB by measuring defect<br />

velocities under different monolayer conditions.<br />

To this end we have conducted experiments varying the surface pressure,<br />

which is kept below 4 mN m −1 . Above this value, the range of motion of<br />

the slow defect cannot be resolved in our system. Similarly, temperature is<br />

maintained at 32 ◦ C, a compromise between the hindered mobility at lower<br />

temperatures and the fast dynamics above this value. Droplet radii were considered<br />

in the range of 30 - 60 µm, with no significant effect in the results.<br />

Experimental ratios of defect velocities (defined as the linear velocity<br />

along the curved boundary) as a function of lateral pressure are shown in Fig.<br />

2.14 in terms of the mean and standard deviation of several experiments. The


2.5 Discussion 19<br />

ratio v−/v+ can be transformed, using Eq. 2.10, to yield α vs. Π and, therefore,<br />

the ratio of the elastic constants as a function of Π (Fig. 2.15). Our measurements<br />

can be combined with the results of Feder et al. (Feder et al., 1997)<br />

whose analysis of orientation fluctuations in this system state that the value<br />

for the geometric mean of KS and KB is of about K ≡ √ KSKB ∼ (40±25)kBT ,<br />

and insensitive to pressure variation. This way, we can estimate KS and KB as<br />

a function of Π, (Fig. 2.15), using<br />

− K2<br />

α = K2 S<br />

. (2.11)<br />

+ K2<br />

K 2 S<br />

In the zero-Π limit (lowest anisotropy) KS KB 10 −19 J. As Π increases,<br />

there is a rapid increase (resp. decrease) of KS (resp. KB) until Π 1 mN/m,<br />

where further variation of these parameters cannot be resolved. These values<br />

are consistent with estimations found in the literature (Tabe et al., 1999) and<br />

values extrapolated for thin suspended liquid crystal films (Rosenblatt et al.,<br />

1979).<br />

v−/v+<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 1 2 3 4<br />

Π (mN m −1 )<br />

Fig. 2.14. Experimental measurements of the relative velocities of boundary defects moving<br />

towards their annihilation as a function of the lateral surface pressure.<br />

We can qualitatively relate these results to the known tendency of azobenzene<br />

derivatives to form supramolecular aggregates (H-aggregates), whose<br />

presence can be revealed here by spectroscopy methods (Pedrosa et al., 2002).<br />

In short, aggregates form in such a way that the azobenzene planes of neighboring<br />

molecules are parallel, stacking perpendicularly to the molecular tilting<br />

direction. Because of this, a splay arrangement of the molecular field<br />

inside the axi-symmetric droplets would be energetically more demanding<br />

than its bend counterpart, since the former would impose a certain curvature


20 2 Probing elastic anisotropy from defect dynamics in Langmuir Monolayers<br />

α<br />

!<br />

1.00<br />

0.95<br />

0.90<br />

0.85<br />

0.80<br />

0<br />

K S, K B (J)<br />

1<br />

10 -17<br />

10 -18<br />

10 -19<br />

10 -20<br />

10 -21<br />

0<br />

2<br />

"(mM m -1 )<br />

Π (mN m −1 )<br />

1 2 3<br />

"(mM m -1 Π (mN m ) −1 )<br />

Fig. 2.15. Anisotropy parameter, α, as a function of Π, as obtained from the data in Fig. 2.14.<br />

The inset shows the value of the elastic constants KS (), and KB (◦) estimated as described<br />

in the text.<br />

on the H-aggregates. The known increase of the extension of aggregates with<br />

Π (Ignés-Mullol et al., 2005) can therefore justify the observation that splay<br />

distortions are increasingly less favorable. In fact, molecules cease to behave<br />

as individual monomers at Π 2 mN/m, which may justify the levelling off<br />

of the elastic constants at moderate pressures. Nevertheless, this qualitative<br />

argument cannot explain why the behavior is so dramatically different even<br />

with modest extension of the aggregation. Molecular dynamics simulations<br />

could be employed to address these issues (Allen, 1996).<br />

2.6 Conclusions<br />

We have studied Langmuir monolayers where defect dynamics can be analyzed<br />

in the absence of backflow. As a consequence, differential defect mobilities<br />

can be traced back to elastic anisotropy. We have shown that this simplified<br />

scenario can be exploited with the aid of a simple model to use dynamical<br />

measurements to gain quantitative knowledge of the dependence of<br />

rather elusive material parameters, here the elastic constants, on a basic thermodynamic<br />

control parameter of the monolayer such as the surface pressure.<br />

3<br />

4<br />

4


Acknowledgments<br />

2.6 Conclusions 21<br />

The work in this chapter has been done in close collaboration with J. Ignés-<br />

Mullol and F. Sagués, in particular, the experimental measurements performed<br />

on the 8Az3COOH monolayers.


3<br />

Self-organization and cooperativity of weakly coupled<br />

molecular motors<br />

3.1 Introduction<br />

The collective behavior of molecular motors plays an important role in many<br />

biological phenomena, including muscle contraction, cellular transport, cell<br />

motility and cell division.<br />

In recent years, the appearance of new in vitro experiments allowed the<br />

observation of single motor proteins. These experiments included gliding essays,<br />

optical tweezers and micromechanical force measurements. Their results<br />

revealed new insights into the basic principles of single molecular motors,<br />

and inspired new models for cooperative motors. In that context, molecular<br />

motors attached to actin (myosin), microtubule (kynesin or dynein) filaments<br />

or rigid cargoes, were modeled as rigidly coupled motors (i.e., sharing<br />

equally any external force applied) (Leibler and Huse, 1993; Juelicher<br />

and Prost, 1995; Julicher et al., 1997; Badoual et al., 2002; Klumpp and<br />

Lipowsky, 2005; Klumpp et al., 2006). These models resulted in non linear<br />

force-velocity relationships and dynamic transitions with spontaneous bidirectional<br />

movement in motors lacking intrinsic directionality (as observed<br />

experimentally in the case of NK11, a mutant of NCD (Endow and Higuchi,<br />

2000)). However, the assumption of rigid coupling may result inappropriate<br />

in certain conditions such as in gliding essays for instance, when the distance<br />

between motors is large enough. Relaxing the backbone rigidity and letting<br />

the motors interact elastically, leads to a cross-over behavior of the motors<br />

from non linear to linear force-velocity curves (Vilfan et al., 1998).<br />

More recent experiments have emphasized that for intracellular traffic,<br />

the cooperative action of groups of processive motors is required when they<br />

are attached to fluid-like cargos, such as vesicles or membrane tubes (Koster<br />

et al., 2003; Leduc et al., 2004). In this context, motors are not rigidly cou-


24 3 Self-organization and cooperativity of weakly coupled molecular motors<br />

pled, and although they may interact, they are free to move independently<br />

from each other. This new scenario of motor cooperativity inspired new theoretical<br />

models based on lattice formulation, which consider both attractive<br />

and repulsive interactions (Campas et al., 2006). These models reproduce<br />

and explain qualitatively the ability of groups of motors to pull on membrane<br />

tubes, but are unable to predict the force distribution among the motors since<br />

they are based on discrete hoping rates and not on mechanistic models.<br />

Inspired on the case of motors pulling a liquid cargo, we will consider<br />

the situation of weakly coupled molecular motors that cooperate to overcome<br />

an external force that is applied to only one or a few motors in the front.<br />

Our approach will be based on the simplest possible two state model for<br />

a motor (Julicher et al., 1997). At the price of missing a detailed description<br />

of specific cases, our mechanistic approach (in contrast to the lattice approach)<br />

allows a quantitative analysis of cluster efficiency and force transmission<br />

among the motors. We aim to gain new insights into generic phenomena<br />

concerning force transmission between motors and the connection between<br />

motor interactions, velocity-force curves and the collective performance of<br />

motor clusters.<br />

3.1.1 Molecular motor modelization<br />

Following (Julicher et al., 1997), we will briefly review the theoretical models<br />

of translationary molecular motors. Molecular motors are basically proteins<br />

that convert, at a molecular scale, chemical energy into mechanical work.<br />

Translationary molecular motors move unidirectionally along polar filaments<br />

made of periodic structures (actin filaments and microtubules). Many models<br />

for molecular motors are inspired on ”thermal ratchets” by Feynman (Feynman<br />

et al., 1966) which explain unidirectionality out of periodic distribution<br />

of temperatures with appropriate asymmetry (Büttiker, 1987; Landauer,<br />

1988). This approach, although interesting, it is not appropriate for molecular<br />

motors which are isothermal machines.<br />

The general mechanism behind the non vanishing displacement of a brownian<br />

particle in an asymmetric periodic structure without any thermal gradient<br />

or macroscopic force, consists in driving the system out of the thermodynamic<br />

equilibrium, violating detailed balance, in which case energy is<br />

constantly added to the system (ATP consumption in the case of molecular<br />

motors), and the time symmetry is broken. There are three different mechanisms<br />

that maintain the system out of thermodynamic equilibrium in a periodic<br />

isothermal ratchet (Julicher et al., 1997),


3.1 Introduction 25<br />

Fluctuating forces. The particle is submitted to a periodic potential and fluctuating<br />

forces that, although having a zero averaged value, have non trivial<br />

temporal correlations that do not satisfy a fluctuation-dissipation theorem,<br />

and encode the energy source.<br />

Fluctuating potentials. The particle is submitted to a periodic potential which<br />

value depends on time. In this case, the fluctuating forces are Gaussian white<br />

noise that obey a fluctuation-dissipation theorem. In this example the energy<br />

source is implicitly manifested in the time dependence of the periodic potential.<br />

Fluctuating state transitions. The particle has different states which transition<br />

rates do not satisfy detailed balance (the ratio of the two state rates it is<br />

not given by the energy difference between the two states). The dynamics of<br />

transitions between states is added independently and may reflect chemical<br />

reactions (ATP consumption as example for molecular motors). The fluctuating<br />

forces satisfy again a fluctuation-dissipation theorem.<br />

We will focus on a simple description of the third mechanism with only<br />

two states which basically represent an attached state and a weakly bound<br />

state, aiming to extract generic features of the motion of a motor which will<br />

allow a straight forward extension to the case of many motors.<br />

3.1.2 The two state model<br />

We consider the simplest fluctuating state transitions model, which is a<br />

molecular motor with only two states, corresponding to the bound state<br />

(power stroke) and a weakly bound state (detached state). The equation of<br />

motion for the position of the motor, x, is a Langevin equation,<br />

dx<br />

λi<br />

dt = −∂xUi(x)+ fi(t), (3.1)<br />

where λi is a friction coefficient that in general may depend on the state of<br />

the motor, and fi(t) is a fluctuating force that satisfy a fluctuation-dissipation<br />

theorem,<br />

〈 fi(t)〉 = 0, 〈 fi(t) f j(t ′ )〉 = 2λikBT δ(t −t ′ )δij. (3.2)<br />

The potential Ui has the filament symmetries and reflects the interaction between<br />

the motor and the filament. As a consequence the potential must be<br />

both periodic and asymmetric. The conformational change of the protein<br />

which leads to the motor movement is represented by the sliding over the


26 3 Self-organization and cooperativity of weakly coupled molecular motors<br />

asymmetric potential, Fig. 3.1. In this model we assume only a conformational<br />

change which leads to the power stroke of the motor. Any other conformational<br />

change is considered instantaneous when compared to chemical<br />

reaction times of the order of milliseconds.<br />

(a)<br />

(b)<br />

ATP<br />

ATP ADP P<br />

P ADP<br />

ADP<br />

Unbinding Diffussion Binding Power stroke<br />

(c) α1 β2<br />

α2<br />

Fig. 3.1. Equivalence between a molecular motor cycle (a) and the two state model (b-c)<br />

explained in the text.<br />

In Fig. 3.1 we schematically show the equivalence between a motor cycle<br />

and the two state model. Initially, the motor is attached to the filament in<br />

what is called ”rigor state” until a molecule of ATP binds to the motor, which<br />

detaches from the filament and hydrolyzes the ATP molecule,<br />

β1<br />

ATP ⇋ ADP + P, (3.3)<br />

this reaction represents the transition from the bound state (1) to the unbound<br />

state (2). Associated to this reaction there is a chemical potential difference<br />

∆µ ≡ µAT P − µADP − µP which measures the free-energy change per ATP<br />

consumed. At the unbound state, when a phosphate molecule P is released<br />

the motor attaches to the filament and performs the power stroke after the<br />

molecule of ADP is released. The transition from the unbound state (2) to the<br />

bound state (1) is thermal (passive),<br />

M − ADP − P ⇋ M + ADP + P. (3.4)


Transition rates<br />

3.1 Introduction 27<br />

There are two transition rates associated to each of the two different reactions<br />

(Eq 3.3 and 3.4), Fig. 3.1c. α1,2 correspond to the hydrolysis of ATP,<br />

and their ratio depends on the difference of energy between the two states,<br />

∆U ≡ U1 −U2, and the difference of chemical energy ∆µ. β1,2 correspond to<br />

the thermally induced release of phosphate and ADP, and their ratio satisfies<br />

detailed balance,<br />

<br />

α1(x) −∆U(x)+∆µ<br />

= exp<br />

(3.5)<br />

α2(x) kBT<br />

<br />

β1(x) −∆U(x)<br />

= exp<br />

, (3.6)<br />

β2(x) kBT<br />

The global transition rates between the two states w1,2, include both α and β<br />

rates: wi ≡ αi + βi, which in terms of α2 ≡ α, and β2 ≡ β,<br />

<br />

−∆U<br />

w1(x) = (α exp∆µ + β)exp<br />

(3.7)<br />

kBT<br />

w2(x)=α + β, (3.8)<br />

which do not satisfy detailed balance unless ∆µ = 0. When a cell has excess<br />

of ATP (∆µ > 0), the motors are driven out of thermodynamic equilibrium<br />

and perform a directional movement along its associated polar filaments. In<br />

physiological conditions, ∆µ ∼ 10kBT , the system is far from equilibrium,<br />

and α exp∆µ ≫ β. If in addition ∆U ≫ kBT , which is equivalent to neglect<br />

thermally excited transitions with respect to conformational changes,<br />

w1(x)=α exp∆µ (3.9)<br />

w2(x)=α + β. (3.10)<br />

For simplicity we can assume U2 to be a flat potential, and de-exitations (corresponding<br />

to the rate w2) to be spatially delocalized. Moreover, it is reasonable<br />

to assume that the transition from the bound to the unbound state, which<br />

depends on ATP binding, is localized at the minimum of U1, corresponding<br />

to the fact that ATP is not likely to bind to a motor during the ”power stroke”.<br />

In Fig. 3.2, we schematically represent the potentials and the transition rates<br />

between the two states.


28 3 Self-organization and cooperativity of weakly coupled molecular motors<br />

Ui<br />

wi<br />

δ<br />

ℓ<br />

Fig. 3.2. Schematics of the molecular motor as a ratchet. The bound state corresponds to an<br />

asymmetric periodic potential U1, with a periodicity ℓ. The bound state corresponds to a flat<br />

potential U2. The transition rate from U1 to U2 state, w1 is localized at the minimum of the<br />

bound state potential (with a width δ). De-excitations are delocalized, w2 =constant.<br />

3.2 Self-organization and cooperativity of weakly coupled<br />

molecular motors<br />

We are interested in the cooperativity of N infinitely processive motors moving<br />

along a one-dimensional filament. An external force F opposing the motion<br />

is applied only on the foremost motor. The force transmitted to the rest of<br />

the motors is given by motor-motor interactions. We consider the interaction<br />

between motors to be a short-ranged (non strongly-binding) potential including<br />

a hard-core repulsive part. By non bindging potential we mean that the<br />

mean distance between motors diverges at the steady state, in contrast to the<br />

case of strongly coupled motors, Fig. 3.3.<br />

We use a two state model as described in section 3.1.2, which is characterized<br />

by a bound and unbound potential and their respective transition rates.<br />

We represent the ratchet potential as Ui(xi,ki), where the index i ∈ (1,N) is a<br />

label for the i-th motor, xi is its position, and ki ∈ (1,2) is a discrete stochastic<br />

variable representing the state of the motor. Finally, the interaction between<br />

motors is represented by the potential W(xi − xk). The equation of motion for<br />

the motors is described by a set of Langevin equations equivalent to Eq. 3.1,<br />

λ ˙xi = −U ′<br />

i (xi,ki) −∑ W<br />

k=i<br />

′ (xi − xk) − Fδ1i + ζi(t), (3.11)<br />

where λ is a friction coefficient, which is taken to be the same for both bound<br />

and unbound states. ζi(t) is the thermal noise that satisfy 〈ζi(t)ζ j(t ′ )〉 =<br />

w1<br />

U2<br />

U1<br />

x<br />

w2<br />

x


3.2 Self-organization and cooperativity of weakly coupled molecular motors 29<br />

W (r)<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

-0.5<br />

0.1 0.2 0.3 0.4 0.5<br />

Fig. 3.3. Several examples of interacting potentials between motors. In black, hard-core potential;<br />

in blue, a weak-attractive (see text for a definition of weak-attractive) potential; in green,<br />

a long-range potential, and in dotted red, an example of strongly binding potential.<br />

2kBT λδijδ(t −t ′ ) (Gaussian process), and its strength is defined as the diffusion<br />

coefficient, D = kBT /λ.<br />

We restrict our present analysis to motors that move on the same filament,<br />

and as a consequence, they share the same ratchet potential, Ui(xi,ki) ≡<br />

U(xi,ki), that represent the interaction between filament and motor. Each motor<br />

switches states independently, that is, transition rates of motor i depend<br />

only on xi (Julicher et al., 1997), in contrast to previous studies where motors<br />

switch at the same time (Derényi and Ajdari, 1996).<br />

For simplicity we define U(x,1) ≡U1(x) as a completely asymmetric sawtooth<br />

potential with period ℓ and height U, and a sliding velocity v = U/(λℓ),<br />

which represents the ”power stroke” of the motor. U(x,2) ≡ U2 is modeled as<br />

a flat potential, Fig. 3.4.<br />

Hereinafter we define U1 and U2 as the bound and unbound states respectively.<br />

The lifetime of U2 is τ and de-exitations are delocalized. Transitions<br />

from U1 to U2 are localized at the minimum of U1(x) (see section 3.1.2) and<br />

are much faster than any other process during a motor cycle (τ and the ”power<br />

stroke”), and are considered instantaneous, which corresponds to a large excess<br />

of ATP (∆µ ≫ 0).<br />

3.2.1 Simulations<br />

We will numerically solve the set of Langevin equations using a standard<br />

Euler method for the integration of differential equations. The discretized<br />

version of the set of equations, Eq. 3.11, reads,<br />

r


30 3 Self-organization and cooperativity of weakly coupled molecular motors<br />

Ui<br />

U<br />

ℓ<br />

v<br />

√ Dτ<br />

Fig. 3.4. Two state potential of a molecular motor. The state U1 is a totally asymmetric potential<br />

with period ℓ with localized transitions at each minima. The unbound state U2 is modeled<br />

as a flat potential with delocalized transitions after a deterministic time τ.<br />

xi( j + 1)=xi( j)+<br />

<br />

U2<br />

U1<br />

x<br />

δki1v −∑ W<br />

k=i<br />

′ (xi( j) − xk( j)) − Fδi1<br />

<br />

∆t (3.12)<br />

+(−2Dlog(χ)∆t) 1/2 cos(2πχ), (3.13)<br />

where the last term is a discretization of the thermal noise ξi(t), using a uniform<br />

distributed stochastic variable, χ ∈ (0,1) (Sancho et al., 1982). For the<br />

transitions between states U1 and U2, we define the following rules: From U1<br />

to U2,<br />

If(ℓ + δ >| mod(xi,ℓ) |>ℓ − δ), ki = 2, (3.14)<br />

which corresponds to a transition localized in a small interval δ from the<br />

minimum of the potential (see Fig.3.2). For the transition from U2 to U1, the<br />

motor remains at U2 for a deterministic interval of time τ, and then changes<br />

to U1.<br />

Simulation parameters<br />

The fixed simulation parameters are<br />

• The time step ∆t is fixed to 5 × 10 −5 s.<br />

• The periodicity of the potential U1 is fixed to 8 nm which is roughly the<br />

length of a heterodimer of α-β tubulin (a microtubule monomer).<br />

• The interval around the potential minimum for motor transition is fixed to<br />

δ = 0.002ℓ.<br />

• The de-excitation time of the state U2, τ = 1/w2 is fixed to 50 ms.


3.2 Self-organization and cooperativity of weakly coupled molecular motors 31<br />

• The value of the maximum of the U1 potential is chosen U/kBT = 20,<br />

which is roughly the energy released for ATP hydrolysis.<br />

• The motor friction coefficient is fixed to λ = 4×10 −4 kg s −1 , which leads<br />

to a sliding velocity over the potential U1 of about 0.2µm s −1 , resulting in<br />

a motor velocity consistent with KIF1A velocities (Okada and Hirokawa,<br />

1999; Nishinari et al., 2005; Endow and Higuchi, 2000).<br />

In the following sections, we will consider the cooperativity and selforganization<br />

of molecular motors for different interaction potentials, namely,<br />

hard-core repulsion, weak-attraction and long-range repulsion. We will compare<br />

their velocity-force curves to a mean field calculation (that we will define<br />

in the following). We will see that depending on the interaction between motors,<br />

clusters of motors of sizes different from N self-organize and move at<br />

different velocities. Finally, will consider the distribution of forces transmitted<br />

among the motors and the cluster efficiency.<br />

3.2.2 Mean-field<br />

The degree of cooperativity between motors depends on the correlations between<br />

positional and internal degrees of freedom of different motors. Consider<br />

the case of two motors that are connected by a rigid rod, Fig. 3.5. Their<br />

collective motion depends critically on the state of each motor. There are four<br />

instantaneous velocities that can be defined depending on the state of each<br />

(see Fig. 3.5). This case corresponds to the maximum cooperativity between<br />

motors and the internal degrees of freedom of the motors (state 1 or 2) are<br />

correlated with the positional degree of freedom. Their collective behavior<br />

and extension to N motors have been extensively studied in (Juelicher and<br />

Prost, 1995; Julicher et al., 1997; Badoual et al., 2002) and reveal nonlinear<br />

deviations of the force-velocity relationship.<br />

Consider now the opposite case: two motors that interact with a soft (nonbinding)<br />

potential and an opposing force 1 F acting on the first motor. If the<br />

resulting interacting force varies slowly with respect to the period of the periodic<br />

potential ℓ: ℓW ′′ (r)/W ′ (r) ≪ 1, the interaction force remains insensible<br />

to a motor cycle and its spatial fluctuations, so the motors feel the same<br />

interacting force 〈−W ′ (r)〉 −W ′ (〈r〉) =F/2 irrespective to the particular<br />

motor state, and move at an average velocity v that corresponds to the velocity<br />

of a single motor with an applied force F/2, Fig. 3.6. The intuitive picture<br />

corresponds to represent the motion of a motor as a vehicle (which is over<br />

1 We need to apply a force otherwise the mean distance between motors diverges and they<br />

do not interact with each other.


32 3 Self-organization and cooperativity of weakly coupled molecular motors<br />

(a) (b) (c)<br />

Fig. 3.5. Schematics of two motors connected by a rigid rod. When both motors are in the<br />

diffusive or bound state (a) and (b), they behave as single a motor (with half the total force).<br />

(b) In the crossed configurations, the advancing probability of the motor in the diffusive state<br />

do not rely on noise since the bound motor is pushing (pulling), and the motor advances<br />

deterministically.<br />

damped), with its engine that has an asymmetric combustion cycle (the cycle<br />

of the motor) but moves at a constant velocity. The two motor case would correspond<br />

to two cars that are pushing together against a force, sharing exactly<br />

the same force, and hence moving at the velocity they would move if they<br />

were alone pushing against half of the applied force (Fif. 3.6b). This softly<br />

enough interacting potential results in the intuitive result that a force applied<br />

to a cluster of N motors is equally shared and their velocity VN is related to<br />

the velocity of a single motor, VN(F)=V1(F/N).<br />

(a) (b)<br />

F<br />

F/2 F/2<br />

〈 ˙x〉 = v v<br />

Fig. 3.6. Schematics of the interaction of two motors satisfying mean field. (a) the spring<br />

connecting the two motors represents a soft connection that is insensible over a period ℓ. In<br />

(b), the car analogy, where an over damped car is moving at a constant velocity although its<br />

internal motor has asymmetric combustion cycles.<br />

More rigorously, we define the mean field approximation for a cluster of<br />

N motors as satisfying the relation:<br />

〈−W ′ (r)〉 −W ′ (〈r〉), (3.15)<br />

so that correlations between positional and internal degrees of freedom of<br />

different motors are neglected. The steady state solution of Eq. 3.11 for<br />

the mean field approximation implies that the force between two motors is<br />

F


3.2 Self-organization and cooperativity of weakly coupled molecular motors 33<br />

constant, and, consequently, each motor behaves as an isolated motor subject<br />

to an equal share F/N of the external force. The collective dynamics<br />

is then reduced to the single-motor problem with VN(F) =V1(F/N) and<br />

FN(V )=NF1(V ). The physical picture corresponds to N ballistic motors defined<br />

solely by V1(F).<br />

It is illustrative to obtain the exact velocity-force curve for N motors in<br />

mean field approximation, or equivalently, for a single motor at a force F/N.<br />

We use these values later as a reference to compare with interacting potentials<br />

that not satisfy the conditions for mean field approximation.<br />

The mean velocity over a time t corresponding to m cycles is given by<br />

〈 ˙x〉 = ∆x<br />

t ≡ ∑m i=0 δxi<br />

∑ m i=0 δti<br />

= 〈δx〉<br />

, (3.16)<br />

〈δt〉<br />

where ∆x is the total distance traveled by the motor during the time t. This<br />

distance can be decomposed in steps of distances δxi, which correspond to<br />

the advanced distance per cycle. Similarly, the time t is the sum of intervals<br />

δti corresponding to the time lapse of the i-th cycle. Finally, the mean velocity<br />

is the ratio of the mean displacement and the mean time lapse per cycle. We<br />

compute the mean velocity in two steps: first, the mean displacement and<br />

second the corresponding mean time per cycle,<br />

Mean displacement per cycle<br />

We consider the initial position of the motor as an arbitrary minimum of the<br />

asymmetric potential U1, Fig. 3.7. The motor instantaneously changes its state<br />

to U2 and starts diffusing for a time τ. During this diffusive state, the motor is<br />

drifted by the external force at a velocity F/λ. After a time τ the motor deexitate<br />

to the state U2 at a position x with a gaussian probability distribution<br />

centered at x0 ≡−Fτ/λ,<br />

P(x)= 1<br />

<br />

(x + Fτ/λ)2<br />

√ exp −<br />

2πa 2a2 <br />

, (3.17)<br />

where a2 ≡ Dτ. At the state U2, the motor diffuses and slides over the ramp of<br />

the potential towards the minimum. The probability of going forward towards<br />

the minimum of the potential (and not going backwards) is given by (Parmeggiani<br />

et al., 1999)<br />

<br />

<br />

(F −U1/ℓ)z<br />

P→ 1 − exp − . (3.18)<br />

kBT


34 3 Self-organization and cooperativity of weakly coupled molecular motors<br />

When the external force approaches the force due to the power stroke, U1/ℓ,<br />

the advancing probability is considerably reduced. For the sake of simplicity,<br />

we will neglect here the noise in the U1 state, and the motor always move<br />

forward towards the minimum, advancing a multiple of the potential period,<br />

nℓ, each cycle.<br />

Expressing the transition position from the state U2 to U1, x ≡ (n−1)ℓ+z<br />

(see Fig. 3.7), the mean displacement per cycle is 〈δx〉 = ℓ∑ ∞ n=−∞ np(n), with<br />

p(n) the probability of falling at any position between n(ℓ − 1) and nℓ,<br />

〈δx〉 = ℓ<br />

∞<br />

∑<br />

n=−∞<br />

ℓ<br />

n<br />

0<br />

<br />

<br />

1 (ℓ(n − 1)+z + Fτ/λ)2<br />

dz√<br />

exp − . (3.19)<br />

2πDτ 2Dτ<br />

For an external force F = λℓ/(2τ), the motor at state U1 after the transition<br />

time τ is positioned at half the period of the potential from the initial position.<br />

In this case, the probability of advancing forward or backwards (+nℓ or<br />

−nℓ) is exactly the same and the mean displacement is δx = 0. This force<br />

corresponds to the stall force of the motor if λℓ/(2τ) < U/ℓ, otherwise, the<br />

stall force is U/ℓ. Defining the two dimensionless parameters α = vτ/ℓ, the<br />

ratio of the lifetime of the state U2 and the sliding time on the state U1, and<br />

β ≡ ℓ/(4Dτ) 1/2 , the ratio of the potential period and the diffusive displacement<br />

in U2, we can express the stall force as,<br />

F 0<br />

s = λv min<br />

<br />

1, 1<br />

2α<br />

where the super-index 0 stands for mean field.<br />

Mean time lapse per cycle<br />

<br />

, (3.20)<br />

The time lapse corresponding to advancing a distance nℓ is the sum of the U2<br />

life-time τ and the sliding time on the U1 state, δt = τ +ts, with ts =(ℓ−z)/v,<br />

see Fig. 3.7. The mean time lapse is then 〈δt〉 = ∑ ∞ n=−∞ δt(n)p(n), with p(n)<br />

the probability of falling at any position between n(ℓ − 1) and nℓ,<br />

〈δt〉 =<br />

∞<br />

∑<br />

n=−∞<br />

ℓ<br />

0<br />

<br />

dz τ +<br />

ℓ − z<br />

v<br />

<br />

<br />

<br />

1 (ℓ(n − 1)+z + Fτ/λ)2<br />

√ exp −<br />

2πDτ 2Dτ<br />

(3.21)<br />

Finally the mean velocity V 0 ≡〈˙x〉 is the ratio of Eq. 3.19 and Eq. 3.21,<br />

which in terms of the dimensionless parameters α, β and the dimensionless<br />

force f ≡ F/(λv), can be expressed as


3.2 Self-organization and cooperativity of weakly coupled molecular motors 35<br />

F τ/λ<br />

(i)<br />

t =0<br />

(ii)<br />

z<br />

(iii)<br />

0 nℓ − 1 x nℓ<br />

Fig. 3.7. Schematics of the meanfield calculation for the force-velocity curve of N motors,<br />

which corresponds to a single motor with a rescaled force. (i) transition from the U1 to U2<br />

state and drift due to the external force. (ii) the motor de-excitate to the state U1 at a distance x<br />

from the origin with an associated probability (see text). (iii) the motor slides over the potential<br />

U1 until a new minimum at nℓ is reached. As explained in the text, backwards transitions are<br />

neglected in the limit of vanishing noise.<br />

where<br />

φ( f ) ≡ 1<br />

2<br />

<br />

erfc(αβ f )+<br />

V 0 φ( f )<br />

( f ) = (1 − f ) , (3.22)<br />

α + φ( f )<br />

∞<br />

∑<br />

n=1<br />

(erfc(β(n + α f )) − erfc(β(n − α f )))<br />

<br />

,<br />

(3.23)<br />

and erfc(x) ≡ x<br />

0 dxexp(−x2 ) is the complementary error function. The veloc-<br />

ity of a free motor ( f = 0) is given by the simple expression V 0 (0)= v<br />

1+2α .<br />

In this calculation we have neglected the noise contribution during the sliding<br />

in the U1 state. This contribution might be important, specially near the<br />

stall force if Fs ∼ U/ℓ, in which case the probability of a backward step tends<br />

to unity (see Eq. 3.18). In general, our approximation should be valid for<br />

U/ℓ ≫ λℓ/(2τ) or a small noise intensity such that the length diffused in a<br />

typical sliding time ∼ ℓ/v is much smaller than the period of the potential,<br />

D ≪ vℓ. The parameters we are considering correspond to this last condition.<br />

In Fig. 3.8, we plot velocity-force curves for different noise intensities<br />

and de-excitation τ, showing that the disagreement with our approximation is<br />

fairly small for small noise. Hereinafter, we will use the values F 0<br />

s and V 0 (0)<br />

calculated to compare with the non mean field cases.


36 3 Self-organization and cooperativity of weakly coupled molecular motors<br />

V1/V1(0)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

F/Fs<br />

Fig. 3.8. Velocity-force curves for the mean field calculation (solid line) (Eq. 3.22), the corresponding<br />

simulation with noise turned off in the U1 state (solid symbols), and simulation<br />

with noise in both states (empty symbols). The parameters used are: τ = 0.05 and D = 0.25<br />

(back and circles); τ = 0.05 and D = 0.025 (red and squares); τ = 0.175 and D = 0.25 (blue<br />

and diamonds); τ = 0.175 and D = 0.025 (green and triangles). Both velocities and forces are<br />

normalized to the analytical zero force velocity and the stall force (see text).<br />

3.2.3 Hard-core repulsion<br />

Significant deviations from mean field are expected generally for shortranged<br />

repulsive interactions (Campas et al., 2006) and for small or moderate<br />

noise strength in intermediate force range. In our model, we expect correlations<br />

between motors to enhance the motor performance in general as in the<br />

rigidly bound motors (see Fig. 3.5). However, not all the motor configurations<br />

in short-range repulsive interactions contribute to increase the velocity with<br />

respect to mean field, Fig. 3.9.<br />

We will study the case of a hard-core potential between motors, modeled<br />

by a truncated Lennard-Jones potential (see Fig. 3.3),<br />

W(r)= 4ε<br />

<br />

σ 6 σ 12<br />

−<br />

r6 r12 <br />

r ≤ 2 1/6 σ<br />

(3.24)<br />

ε r > 2 1/6 σ.<br />

We analyze the dependence of the deviation from mean field with respect to<br />

(i) β, the dimensionless parameter which is the ratio of the potential period<br />

and the diffusive displacement in the state U2. Since the period of the potential<br />

is fixed (ℓ ∼ 8 nm), varying β stands for varying noise intensity, Fig. 3.10. (ii)<br />

the hard-core parameter σ, which represents effectively the size of the motor


3.2 Self-organization and cooperativity of weakly coupled molecular motors 37<br />

Fig. 3.11. We generally obtain enhanced motor performance with respect to<br />

mean field, VN(F) > V1(F/N), for most of the parameter values. However,<br />

there is a small region of parameters for which the deviation with respect to<br />

mean field vanishes or even become negative.<br />

Deviation from mean field<br />

The deviation from mean field can be explained by considering the 2 motor<br />

case with each of the possible motor-motor state combinations, Fig. 3.9.<br />

There are two configurations of the motors that correspond to a mean field<br />

contribution: the two motors in the bound state U1, Fig. 3.9(b) and the two<br />

motors in the state unbound U2. These two configurations satisfy that the<br />

mean force is equally shared by the two motors and the velocity corresponds<br />

to the velocity a single motor would have in half the force applied. The<br />

crossed configurations are the responsible for mean field deviation and illustrate<br />

the effect of the correlations between motors. In Fig. 3.9a, the foremost<br />

motor is in the unbound state and the second motor is sliding in the bound<br />

state. A single motor in U2 would drift at a velocity −F/λ. In the case of two<br />

motors, the second transmits the power stroke to the foremost, which finally<br />

moves at a velocity v−F/(2λ) (> 0 if F is smaller than the stall force) which<br />

deterministically enhances the advancing probability of the motor which do<br />

not rely on noise, leading to an effect that persists for small noise, O(D 0 ),<br />

while V1(F) falls as O( √ Dexp(−1/ √ D)). As a consequence, this motor configuration<br />

always contribute towards gain in motor performance. For the last<br />

configuration, corresponding to the first motor in the U1 state and the second<br />

in the U2 state, the first motor is moving alone since the second motor is diffusing<br />

and in mean, not moving. As a consequence, the first motor moves at a<br />

velocity v − F/λ, which is smaller than the mean field velocity v − F/(2λ),<br />

and the deviation from mean field is negative.<br />

The relative weights between configurations U1-U2 and U2-U1 determine<br />

the amount and sign of deviation from mean field. A large positive deviation<br />

from mean field will occur when the configuration U1-U2 dominates,<br />

this is, when the time during which the second motor pushes the first one<br />

is maximized. This time depends strongly on the particle size, or, σ. For<br />

mod(σ,ℓ) ∼ 1, this effect disappears completely since the pushing distance<br />

tends to 0: if both motors are in contact, they coincide in the U1 minimum<br />

and consequently they change to the U2 state together. On the contrary, configuration<br />

U2-U1 is in general present, so we expect a negative deviation from<br />

the mean field for σ close to the period length ℓ. For intermediate values<br />

of σ, we expect a gain with respect to mean field, because the second mo-


38 3 Self-organization and cooperativity of weakly coupled molecular motors<br />

tor will generically push against the first motor. In particular, due to symmetry,<br />

the maximum performance will correspond to mod(σ,ℓ) ∼ 1/2, Fig.<br />

3.11. The amount of deviation from mean-mean field, depends on the correlations<br />

between motors. If entropic repulsion between motors dominates,<br />

D<br />

(V1(0)−V1(F))ℓ<br />

≫ 1, correlations between motors are fully suppressed, and we<br />

recover the mean field approximation results. For very weak noise, the increased<br />

motor performance becomes dramatic, Fig. 3.10.<br />

(a) (b) (c) (d)<br />

v − F/(2λ) −F/(2λ)<br />

v − F/(2λ)<br />

v − F/λ<br />

Fig. 3.9. Schematics of cooperativity in the case of two motors with hard-core repulsion. (a)<br />

the second motor pushes the foremost motor which is at the diffusive state. This configuration<br />

represents a net gain with respect to mean field, since the advancing mechanism for the foremost<br />

motor do not rely on thermal fluctuations as in the mean field case. (b) the two motors<br />

slide over the U1 potential, sharing the external load as in mean field. (c) the foremost motor<br />

takes over all the load when sliding over the U1 potential whilst the second motor diffuses.<br />

This configuration corresponds to a loss with respect to mean field since a larger force is applied<br />

on the foremost motor. (d) the two motors diffuse on the U2 state sharing the external<br />

load as in mean field approximation.<br />

N motors<br />

For large N, we find that VN( f ) converges to a limiting curve VN → V∞( f ),<br />

that corresponds to a force per motor f that tends to a constant. One would<br />

be tempted to state that the force per motor tends the corresponding mean<br />

field approximation, that is, the force per motor is equally distributed and<br />

f = F/N. As we will see in section 3.3.1, the force distribution is not homogeneous<br />

throughout the motors, although its rescaled value saturates to a<br />

constant. In this sense, we recover the extensive scaling as in mean field, but<br />

with a V∞( f ) = V1( f ), Fig. 3.12. This is an important difference with respect<br />

to the prediction of the lattice model in (Campas et al., 2006), where VN(F)<br />

(the non rescaled velocity-force curve) was found to saturate with N to a limiting<br />

curve VN(F) → ˜V (F), i.e. they obtained a non-extensive scaling of force,<br />

except near stall. Consequently even with an unlimited supply of motors, the<br />

cluster could not advance against an arbitrarily large force, contrarily to our


3.2 Self-organization and cooperativity of weakly coupled molecular motors 39<br />

VN/V 0<br />

1 (0)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

F/(NF 0 s )<br />

Fig. 3.10. Velocity-force relationship for N = 1, 2 motors, α = 0.15 and hard-core repulsion<br />

(repulsive part of a Lennard-Jones potential truncated at r = 2 1/6 σ, with ε = 1 and σ = 0.2ℓ).<br />

Circles correspond to β = 10; empty circles, N = 1 and N = 2 with mean field; solid circles, N<br />

= 2. Squares correspond to β = 44.7; empty squares, N = 1 and N = 2 with mean field; solid<br />

squares, N = 2. Axes are normalized by D = 0 values of the 1-motor stall force and the 1-motor<br />

free velocity (see text).<br />

V2(2F0)/V1(F0)<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0 0.5 1 1.5 2 2.5 3<br />

Fig. 3.11. Dependence on the hard-core size σ (in units of ℓ) of the gain with respect to mean<br />

fied for an external force F0 = 1 3 F0<br />

s .<br />

result which states that any arbitrary force can be overcome if enough motors<br />

are supplied.<br />

3.2.4 Long range interaction - transition to mean field<br />

Early in section 3.2.2, we described the criteria for mean field approximation.<br />

In the last section, we showed that increasing noise leads to mean field<br />

σ


40 3 Self-organization and cooperativity of weakly coupled molecular motors<br />

VN/V 0<br />

1 (0)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.5<br />

0 5 10 15<br />

N<br />

20 25 30<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

F/(NF 0 s )<br />

Fig. 3.12. Convergence with the number of motors N, of velocity vs force per motor, for<br />

α = 0.15 and β = 14.1. Diamond, N = 1; inverted triangle, N = 2; triangle, N = 3; empty<br />

square, N = 5; empty circle, N = 10; solid square, N = 15; solid circle, N = 30. In the inset,<br />

the convergence for a fixed force F0 = 1/3F 0<br />

s<br />

(see Fig. 3.10). Another route to mean field is to impose a soft interacting potential<br />

such that the variation of the force over a ratchet period is negligible,<br />

ℓW ′′ (r)/W ′ (r), which ensures the decoupling of internal and spatial degrees<br />

of freedom between motors. This condition is easily achieved for a (longranged)<br />

interaction of the form W(r) ∼ exp(−r/λ), with λ ≫ ℓ. Adding to<br />

this potential a hard-core part to prevent motor-motor crossing, we end up<br />

with a long range interaction (see Fig. 3.3),<br />

<br />

σ 6 σ 12<br />

4ε −<br />

W(r)= r6 r12 <br />

+ Aλ exp − r<br />

<br />

r ≤ 2<br />

λ<br />

1/6 σ<br />

<br />

Aλ exp − r<br />

<br />

r > 2<br />

λ<br />

1/6 (3.25)<br />

σ,<br />

where the parameter A, sets the intensity force of the long-range potential.<br />

We expect a crossover force between hardcore repulsion and long-range interaction,<br />

if the force provided by the long-range part is not enough to overcome<br />

the force per motor: −W ′ (2 1/6 σ) < F/N. In Fig. 3.13, we show several<br />

plots of the 2-motor velocity-force curves for different long-range intensities<br />

intensity. The transition region from mean field to cooperative behavior<br />

corresponds qualitatively to a cross-over force F ∗ 2A/λ exp(−2 1/6 σ/λ).<br />

Above the cross-over force, the velocity-force curve approaches the hardcore<br />

curve (region in between the top and bottom curve in Fig. 3.13). The gain<br />

in velocity from respect to mean field in this region grows linear with the force<br />

until it reaches the hard-core region, inset of Fig. 3.13, with a slope that is<br />

VN (NF0)/V1(F0)<br />

2.5<br />

2<br />

1.5<br />

1


3.2 Self-organization and cooperativity of weakly coupled molecular motors 41<br />

independent of the long-range intensity. This linear growth can be attributed<br />

to the cooperativity of the motors which slowly grow (since λ ≫ σ) until<br />

it saturates to the maximum value given by the close packing of the motors<br />

(hard-core part). The fact that the cooperativity increases linearly with the<br />

force in the long-range interaction may have influence in the distribution of<br />

forces in larger clusters (N > 2), since an unevenly distribution of forces in a<br />

motor cluster could lead to the same motor velocity due to the cooperativity<br />

increase in the more loaded regions (see section 3.3.1).<br />

V2/V 0<br />

1 (0)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

∆V2/V 0<br />

1 (0)<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

0 0.1 0.2 0.3 0.4<br />

∆F/(2F 0 s )<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

F/(2F 0 s )<br />

Fig. 3.13. Velocity force curves for long range interaction for A = 0.5,1,1.5,2 and 10. In the<br />

inset, velocity gain with respect to mean field for the transition region between the mean field<br />

curve (inferior curve) and hard-core curve (superior curve) as a function of the increase of<br />

force from the cross-over, for each value A. In all cases, λ = 10ℓ.<br />

3.2.5 Attractive interaction<br />

We have seen in section 3.2.3 that the enhanced cooperativity from mean field<br />

arises from the interplay between the gain from the crossed configurations<br />

U1-U2, where the trailing motor pushes the diffusing leading motor (see Fig.<br />

3.9a), and the loss from the crossed configurations U2-U1, where the leading<br />

motor moves alone and therefore slower than mean field (see Fig. 3.9c). In<br />

the case of rigidly coupled motors, the crossed configuration U2-U1 gives<br />

also a net gain because the leading motor pulls the diffusive trailing motor<br />

increasing the advancing probability dramatically (see Fig. 3.5b).<br />

If a weak (non-binding) attractive interaction is added to the hard-core repulsion<br />

(a Lennard-Jones potential in our case), a new scenario emerges in


42 3 Self-organization and cooperativity of weakly coupled molecular motors<br />

which the collective behavior differs qualitatively from both short-ranged interactions<br />

and rigidly coupled motors. Consider first the case of 2 motors. For<br />

vanishing external force, the attractive interaction is not binding (the mean<br />

distance between motors diverge with time), and the velocity is equal to the<br />

free motor. As we increase the force in the leading motor, the effective attractive<br />

potential between motors increases, Fig. 3.14a and the push-and-pull<br />

mechanism is increased. As a consequence, we find that when increasing the<br />

force, the mean velocity of the pair is increased with respect to the free motor<br />

velocity, Fig. 3.14b.Upon further increase of the load, the velocity V2 must<br />

eventually decrease below V1(0), Fig. 3.14b.<br />

(a) (b)<br />

W (r)+Fr<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Binding<br />

Non binding<br />

0 0.5 1 1.5 2<br />

r<br />

V2(F )<br />

2.6<br />

2.5<br />

2.4<br />

2.3<br />

2.2<br />

2.1<br />

2<br />

0 0.2 0.4 0.6 0.8 1<br />

Fig. 3.14. (a) Effective attractive potential resulting of the sum of a non binding attractive<br />

interaction between motors and a non vanishing external force (red). Non binding attractive<br />

potential in the absence of external force (black). (b) Non monotonous velocity-force curve<br />

for 2 attractive motors.<br />

This behavior has strong implications for the self-organization in the case<br />

where many motors are available.<br />

Cluster dynamics<br />

In the case of N motors, we expect a non-trivial self-organization of motors<br />

as a consequence of the non-monotonous velocity-force curve for 2 motors<br />

(see Fig. 3.14b), that exhibits velocities that are larger than the free motor<br />

velocity. In fact, for forces such that V2(F) > V1(0), a cluster formed by the<br />

two first motors escape from the rest moving at faster velocity. When F is<br />

increased beyond the value for which V2(F)=V1(0), the motor following the<br />

cluster is captured and the new cluster of 3 motors will start increasing its<br />

velocity with the force again as in the 2 motor case. As a consequence, if<br />

N motors are available, the curve VN(F) exhibits N − 2 ”rebounds” at V1(0)<br />

F


3.2 Self-organization and cooperativity of weakly coupled molecular motors 43<br />

before decreasing below that value (when no more motors are available), each<br />

one consisting in the capture of an additional motor by the cluster, Fig. 3.15.<br />

VN/V1(0)<br />

1.15<br />

1.1<br />

1.05<br />

1<br />

0.95<br />

0 0.5 1 1.5 2 2.5<br />

Fig. 3.15. Velocity-force curve for the case of attractive forces with α = 0.15, β = 14.1, ε = 1<br />

and σ = 0.2ℓ.<br />

The remarkable behavior depicted in Fig. 3.15 (for which the depth of the<br />

attractive well is only 10% of U) shows a genuinely self-organized collective<br />

behavior that has no counterpart in single-motor dynamics or in rigidly<br />

coupled motors.<br />

Transient bimodality and hysteresis<br />

In the neighborhood of force values Fi, where VN(Fi) =V1(0), we expect<br />

transient stable clusters of i + 1 and i + 2 motors, where i is the i − th rebound<br />

corresponding to the force Fi. If we consider the case of 3 motors, we<br />

expect to observe clusters of both 3 and 2 motors near the force for which<br />

V3(F)=V1(0). For smaller forces, although the stationary cluster is formed<br />

by 2 motors, a finite long lived cluster of 3 motors will appear transiently if<br />

the motors start close enough. If we plot an histogram of velocities as a function<br />

of time, we indeed observe initially a peak at a velocity V3 corresponding<br />

to the cluster of 3 motors, that eventually disappears while another peak of<br />

velocity V2 corresponding to the cluster of 2 motors, remains stationary for<br />

large times, Fig. 3.16b. This transient bimodality implies the appearance of<br />

metastability and hysteresis as we move along the force coordinate since the<br />

capture of a motor (when F surpasses Fi from below) or the release of one<br />

motor (when F decreases below Fi from above) takes a finite time (see Fig.<br />

3.16a)<br />

F


44 3 Self-organization and cooperativity of weakly coupled molecular motors<br />

(a) (b)<br />

V3<br />

2.5<br />

2.4<br />

2.3<br />

2.2<br />

2.1<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4<br />

F<br />

Fig. 3.16. (a) Velocity-force curve for 3 motors. Arrows indicate hysteric transitions from the<br />

metastable extensions for N = 2 and N = 3. (b) Transient bimodality for the decay of the<br />

3-motor metastable cluster to the 2-motor (downward arrow in (a)). In all cases, α = 0.15,<br />

β = 14.1 and σ = 0.2ℓ.<br />

3.3 Discussion<br />

3.3.1 Force transmission<br />

An important aspect for motor kinetics is how the external force is transmitted<br />

through the cluster of motors, since force affects dramatically the motor<br />

detachment rate. If each motor in the cluster shares the same amount of force,<br />

the cluster will be more stable and able to sustain larger forces than in the<br />

case of unevenly shared force, where the motors sustaining more load will<br />

tend to detach faster. Depending on the motor reservoir, this effect could lead<br />

to a catastrophe event in which all the motors in the cluster would detach, decreasing<br />

dramatically the stall force of the cluster. An extension of the present<br />

work dealing with cluster kinetics including detachment dependent on force<br />

is work in progress (Brugues and Casademunt, 2008). In this section, we are<br />

interested in the force distribution in the cluster depending on the interaction<br />

potential. In Fig. 3.17a, we show the force per motor due only to interaction 2<br />

as a function of the position, N = 1 being the first motor, without adding the<br />

power stroke part. Surprisingly, the force distribution is not in general uniform.<br />

For hard-core interaction, there exist a plateau of maximal sustained<br />

force near the front of the cluster that exceeds the force corresponding to<br />

mean field approximation, and a zone of about 3 motors that sustain smaller<br />

2 We do not add the force due to the ratchet potential since the total force per motor must be<br />

the same in order for the cluster to move at the same velocity.<br />

t<br />

V2<br />

V3<br />

V


3.3 Discussion 45<br />

forces than mean field. Using the long-range interaction, we have seen that<br />

the cooperativity gain increases with the interaction force between motors<br />

until it reaches a saturation value for motors which are very close to each<br />

other. This effect is enough to explain the non uniform distribution of forces<br />

between motors. In Fig.3.17b, we show the force distribution for the case of 7<br />

motors. We can see that for a large long-range intensity (Eq. 3.25), we recover<br />

mean field, this is, the force is uniformly distributed among the motors. As<br />

soon as we decrease the intensity, the external force will reach the cross-over<br />

force and some motors will begin to cooperate (in particular, the motors in the<br />

front), being able to sustain larger forces than the non cooperating motors, for<br />

the same cluster velocity (see Fig.3.13). The increase of cooperativity is linear<br />

for a range of forces, until it saturates, in which case the force distribution<br />

corresponds to the hard-core case (see Fig. 3.17b). For the attractive case, we<br />

observe that the force is nearly uniformly distributed.<br />

We can reduce the effect of the cooperativity to 2 motors: For the case of<br />

hard-core interaction, the deviation from mean field corresponds to the rear<br />

motor ”pushing” against the first motor. As a consequence, provided that any<br />

motor in the cluster has a motor behind it, it would perform better. The last<br />

motor of the cluster though, it is not pushed by any other motor, and as a<br />

consequence, it will perform worse, and will also push less efficiently. Since<br />

the whole cluster move at the same velocity, the rear motors must sustain<br />

lower forces to compensate its lower performance. Thus, the non uniform<br />

distribution arises naturally from the motor cooperativity.<br />

For the case of attractive interaction, the deviation from mean field corresponds<br />

to both motors pushing and pulling (respectively) to each other, and<br />

as a consequence, the cooperativity is a symmetric effect (not only a rear motor<br />

pushing the foremost one), and as a consequence, all motors in the cluster<br />

perform nearly equal. The last one being slightly less efficient because it lacks<br />

the pushing effect.<br />

In general, one could experimentally imagine to be able to test the kind<br />

of motor interaction by testing not only the force-velocity curves, but also<br />

the stability of the cluster which depends dramatically on the motor force<br />

distribution (Brugues and Casademunt, 2008).<br />

3.3.2 Motor efficiency<br />

The efficiency of a motor is defined as the ratio of mechanical work performed<br />

to chemical energy consumed,


46 3 Self-organization and cooperativity of weakly coupled molecular motors<br />

0.4 (a) (b)<br />

(fi − fmf )/fmf<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

-0.1<br />

0 2 4 6 8 10 12 14 16 18 20<br />

(fi − fmf )/fmf<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

-0.4<br />

1 2 3 4 5 6 7<br />

Cluster position Cluster position<br />

Fig. 3.17. Relative averaged force distribution with respect to mean field among the motors<br />

forming a cluster with an applied external force F = N ( fmf = 1). (a) Distribution for hardcore<br />

interaction for 4, 7, 8, 19 and 20 motors (circles), and for attractive interaction for 10<br />

motors (squares). (b) Distribution of forces for 7 motors for attractive interaction (squares),<br />

hardcore interaction (empty circles) and long-range interaction with intensity A = 0.5 (solid<br />

blue circles), A = 3 (solid red circles) and A = 10 (solid black circles).<br />

η(F)= FV(F)<br />

, (3.26)<br />

r(F)∆µ<br />

where r(F) is the chemical reaction rate, that in our case is the number of<br />

excitations per unit time (Julicher et al., 1997; Parmeggiani et al., 1999),<br />

which corresponds to transitions from the U1 to U2 state. In a mean field<br />

approximation, N motors consume N times as molecules of ATP per unit<br />

time as a single motor, rN(F) =Nr1(F/N), and they sustain N times the<br />

force of a motor for a given velocity, FN(V )=NF1(V ). The efficiency of a<br />

motor cluster within mean field approximation reads ηN(F) =η1(F/N): a<br />

motor cluster has proportionally more motive power, but it is as an efficient<br />

machine as a single motor.<br />

In the case of short-ranged repulsion and weak-attraction, the efficiency<br />

of the motors is generically increased, more strongly for the attractive case,<br />

Fig 3.18a. The enhanced efficiency, is a result of both the increase of power<br />

with respect to mean field VN(F)F ≥ V1(F/N)F, and the decrease of the rate<br />

of ATP consumption rN(F) ≤ Nr1(F/N), Fig. 3.18b.<br />

The counter-intuitive decrease in ATP consumption when the motors<br />

move faster 3 can be qualitatively explained by the same cooperativity mechanism<br />

that increases the velocity with respect to mean field (see section 3.2.3).<br />

Consider the case of two motors. In mean field, the two motors are independent,<br />

move at the same velocity and we only need to consider the consump-<br />

3 One would expect that if a motor moves faster, completes more cycles in less time, and<br />

therefore, should in principle consume more ATP.


10 (a) (b)<br />

ηN /η max<br />

1<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 2 4 6 8 10 12 14<br />

F<br />

r(F )/r(0)<br />

2<br />

1.5<br />

1<br />

0.5<br />

3.3 Discussion 47<br />

0<br />

0 2 4 6 8 10 12 14<br />

Fig. 3.18. (a) Efficiency normalized to the maximum value for N = 1. Empty circles, N = 1 and<br />

N = 3,5 for mean field; squares, N = 3, triangles N = 5 (solid symbol, attractive case, empty<br />

symbol, repulsive case). (b) Rate of ATP consumption normalized to the rate of consumption<br />

at vanishing force, corresponding to 5 motors. Empty circles for mean field, empty triangles<br />

for hard-core repulsion, and solid triangles for attraction. The parameters used are α = 0.15,<br />

β = 14.1 and ε = 1.<br />

tion of one of them. If we focus on the process of advancing one period of<br />

the potential, Fig.3.19a, after the transition from the U1 to U2 state, the motor<br />

diffuses with a negative drift due to the external force and the probability to<br />

de-exitate to the previous period is higher than advancing to the next potential<br />

period. For small forces and small noise, when the motor de-excites to<br />

the previous period, the distance to the minimum U1, z ∼ Fτ/λ is smaller<br />

than the period length z ≪ ℓ and hence, the time for re-excitation to state<br />

U2 is much shorter than the sliding time corresponding to advance a period.<br />

As a consequence, there are many fast transitions between state U1 and U2<br />

before the motor advances one period at which many ATP molecules are consumed<br />

(one for each transition). At increasing forces this effect is increased<br />

because the probability of fast transitions is increased, until a threshold force<br />

at which the consumption exhibits a maximum and increasing the force has<br />

the effect of reducing the sliding time in the previous fast transitions until the<br />

force is equal to the ratchet force and the motor do not slide anymore (in our<br />

parameters, the stall force is given by the siding force).<br />

In the case of motors cooperating with each other (attractive and hardcore),<br />

the scenario is completely different. The probability of advancing a<br />

period is increased by the cooperation between motors, Fig. 3.19b, and the<br />

motor performs nearly a single slow cycle per ATP consumed. This effect is<br />

enhanced with the force since the motors are kept in contact and it is even<br />

more important in the attractive case (as discussed in earlier sections).<br />

F


48 3 Self-organization and cooperativity of weakly coupled molecular motors<br />

(a)<br />

(b)<br />

Many fast cycles<br />

+<br />

one slow cycle ∼ one slow cycle<br />

Fig. 3.19. Qualitative explanation of the differences in ATP consumption between mean field<br />

and cooperative case. In (a), a single motor representing mean field performs many fast cycles<br />

between states U1 and U2 corresponding to de-excitations to the previous period due to the<br />

force driven drift, before advances a period performing a slow cycle. As a consequence, most<br />

of the ATP is consumed corresponds to fast cycles, increasing the rate of consumption. In<br />

(b), cooperativity between motors enhances the probability to advance one period, and as a<br />

consequence, most of the ATP consumed correspond to slow cycles, decreasing the rate of<br />

consumption.<br />

For large clusters, we have seen that the distribution of forces (the average<br />

force of a single motor in the cluster) is not uniform. The foremost motors<br />

sustain larger forces than the rear ones, but move at the same velocity. As a<br />

result, motors in the front have an enhanced cooperativity which is translated<br />

in a higher motor efficiency (Brugues and Casademunt, 2008).<br />

3.4 Conclusions<br />

We have considered the collective behavior of molecular motors for hard-core<br />

repulsion, weak attractive interaction and long-range repulsive interaction.<br />

Their collective performance is generally optimized with respect to mean<br />

field. For the case of attractive motors, new cluster dynamics phenomena<br />

arise which involves hysteresis and transient bimodality. If a large enough<br />

reservoir is available, the cluster moves at a velocity greater than the free<br />

motor, and the cluster size is dynamically readjusted depending on the force<br />

applied. Finally, we have considered the cluster efficiency and force distribution<br />

which is a crucial assumption in some recent works concerning collective<br />

behavior (Campàs et al., 2008).<br />

It remains an open question whether biological systems might take advantage<br />

of these effects, possibly combining them with the regulation of processivity.<br />

Nevertheless, our results may be relevant in the broad context of<br />

nonequilibrium physics, and could be used for artificial design of motors. It<br />

is worth remarking that some inputs of our model, such as the interactions<br />

and the friction coefficient (defining the noise strength), involve not only the


3.4 Conclusions 49<br />

motor itself but also the cargoes. Consequently, it should be feasible to experimentally<br />

tune parameters in broad ranges, and in particular find regimes<br />

where the present predictions are observable.<br />

A limitation of the present analysis for biological or biomimetic applications<br />

is that finite processivity and force-dependent kinetics have been<br />

neglected. This could be easily incorporated in the model (Brugues et al.,<br />

2008a).


4<br />

Membrane-cortex interactions: Micropipette<br />

experiments and blebbing cells


52 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

4.1 Introduction<br />

The plasma membrane is essential for the cell. It encloses the cell, defines<br />

its boundaries, mediates the exchange of materials between the cytosol and<br />

the extracellular media, and contains proteins that act as sensors of external<br />

signals which allow the cell to react and adapt its behavior in response to environmental<br />

changes. Underneath the membrane lies the cytoskeletal cortex, a<br />

highly dynamic network of actin filaments that reorganizes continuously. The<br />

cortex works in conjunction with the membrane, and is primary responsible<br />

for cell shape, cytokinesis, movement and mechanotransduction. Interactions<br />

between cortex and membrane are fundamental for all these functions. The<br />

plasma membrane and cortex are held together by both molecular interaction<br />

between both lipids and membrane proteins and cytoskeletal proteins. The<br />

loss of adhesion induces the formation of blebs, which are highly dynamical<br />

spherical deformations of the cell surface lacking cytoskeleton. If the cell is<br />

to survive, those patches of unbound membrane must eventually heal with<br />

the appearance of a new cortex that retracts the protrusion. Although in most<br />

cases blebs are an indication of cell death, there are some examples where<br />

cells take advantage of blebs to move.<br />

In this chapter, we first investigate the mechanism of cortex and membrane<br />

adhesion. Our approach is based on two completely different descriptions.<br />

On one hand, we properly consider the kinetics of the membrane and cortex<br />

linkers, at the price of losing spatial correlations between links. On the other<br />

hand, we use a coarse grained description for the adhesion in terms of an<br />

effective macroscopic density of adhesion energy. While the first description<br />

allows a better understanding of the dynamical properties of the adhesion,<br />

and in particular, the dependence of the adhesive strength in terms of the<br />

active state of the cell, it fails to predict any scale of bleb nucleation. The<br />

second description do not treat properly the link dynamics, but it is able to<br />

qualitatively describe bleb nucleation and bleb statistics.<br />

The second part of the chapter is devoted to the study of the dynamical<br />

response of the membrane and cortex to mechanical perturbations, that<br />

eventually lead to an unbinding of the membrane from the cortex. In particular,<br />

we consider how the viscoelasticity (elastic behavior at short time and<br />

viscous behavior at longer time) and the treadmilling (polymerization and<br />

depolymerization) of the cortex affects the membrane-cortex instability. We<br />

propose a simple Maxwell like model for the cortex which allows us to determine<br />

all possible dynamical regimes under a controlled perturbation using<br />

a micropipette aspiration set up. We compare our theoretical predictions with


4.1 Introduction 53<br />

previous works found in the literature and our experiments specially conducted<br />

to observe the different dynamical regimes.


54 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

4.2 Modeling cell mechanics<br />

4.2.1 Cellular membrane<br />

The cellular membrane is a self-assembled bilayer containing lipids and many<br />

other proteins (Alberts et al., 2002). Membrane lipids have a polar head and a<br />

hydrophobic tail which is sandwiched between the hydrophilic head groups.<br />

Within a vesicle, the lateral diffusion of a lipid is of about 1µm 2 s −1 (Alberts<br />

et al., 2002), which means that an average lipid molecule would diffuse the<br />

typical cell length (∼ µm) within a few seconds. In biological membranes,<br />

this diffusion is reduced by a factor of 3-10, due to complex structures in the<br />

membrane such as lipid rafts (Dietrich et al., 2001) and other obstacles like<br />

membrane proteins and cysoskeleton.<br />

In this section we briefly describe the mechanical properties of the membrane:<br />

energy cost of membrane deformation and membrane tension.<br />

Energy cost of membrane deformation<br />

Any deformation of the membrane can be decomposed in a longitudinal deformation<br />

(which includes stretching and shearing) and an out-of-plane deformation.<br />

Each mode of deformation has its energetic cost, and contributes<br />

to the total energy of a membrane deformation.<br />

Shearing corresponds to a deformation on the plane that conserves area,<br />

which microscopically speaking corresponds to a rearrangement of the lipids<br />

in the leaflet. Since the membrane behaves as a two dimensional fluid, the<br />

energetic cost of shearing the membrane is negligible in comparison to the<br />

other two modes of deformation.<br />

Stretching corresponds to an in plane deformation that changes the area<br />

of an element of the surface, which corresponds to varying the lipid density<br />

by direct expansion or contraction of area per molecule. Defining the relative<br />

change in area α ≡ ∆A/A (areal strain), its associated energy per unit surface<br />

is<br />

es = 1<br />

2 Kaα 2 , (4.1)<br />

where Ka is the elastic moduli for direct area expansion with values for typical<br />

lipid vesicles of about 10 −1 Nm −1 (Evans and Rawicz, 1990).<br />

Bending, corresponds to the energy cost of out of plane deformations.<br />

Microscopically, the cost to bend a membrane is related with the different<br />

amounts of direct stretching of molecules in each leaflet, the density of


4.2 Modeling cell mechanics 55<br />

molecules being changed locally depending on the curvature of the deformation.<br />

The bending energy density of a two dimensional membrane can<br />

be expressed in terms of two invariants involving the two principal curvatures<br />

(c1 ≡ 1/R1 and c2 ≡ 1/R2) of a two dimensional surface (Helfrich and<br />

Servuss, 1984),<br />

eb = 1<br />

2 κ(c1 + c2 − c0) 2 + ¯κc1c2. (4.2)<br />

The spontaneous curvature c0 is in general nonzero whenever the two sides of<br />

the membrane are not identical. Microscopically, this corresponds to a different<br />

lipid composition or density, or it is associated to a different composition<br />

of the aqueous media facing the two leaflets of the membrane. The sum of the<br />

principal curvatures, or mean curvature, is associated with an elastic modulus<br />

κ, and the product of principal curvatures or Gaussian curvature with<br />

¯κ. The Gauss-Bonnet theorem shows that the Gaussian curvature is a topological<br />

invariant, and consequently, we will neglect its contribution since we<br />

will restrict ourselves to surfaces with spherical topology and constant bending<br />

modulus ¯κ. Typical measured values for the bending modulus are κ ∼10<br />

kBT (Evans and Rawicz, 1990), and consequently, thermal fluctuations induce<br />

large undulations.<br />

Membrane tension<br />

As opposed to the bending rigidity and the elastic moduli for direct area expansion,<br />

the membrane tension, which by definition is the derivative of the<br />

membrane energy with respect to the area, it is not a well-defined microscopic<br />

quantity. The total microscopic area, the area per lipid times the total number<br />

of lipids, is not experimentally measurable. The reason for that are the thermal<br />

induced fluctuations in the membrane which are expected to be important<br />

since the the bending modulus is of the order the thermal energy. Some<br />

fluctuation modes will be experimentally accessible whereas others will not<br />

be optically resolved (modes of wavelength smaller ∼ µm). The apparent<br />

membrane area then will not correspond to the microscopic total area, Fig.<br />

4.1. Imposing tension in the membrane increases the apparent area by removing<br />

fluctuation modes, each of them carrying the same amount of energy<br />

(Equipartition theorem), and therefore, there is an energetic cost associated<br />

to the increase of apparent area even though the total amount of microscopic<br />

area is conserved. We can define then an effective tension that will be a superposition<br />

of an area increase due to reduction of membrane undulations plus a


56 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

Fig. 4.1. Schematic representation of a fluctuating membrane at different length scales. In<br />

solid line, the real membrane area. In dashed line, the apparent membrane area.<br />

direct expansion in area per molecule. The effective membrane tension is related<br />

to the areal strain by (Helfrich and Servuss, 1984; Fournier et al., 2001;<br />

Evans and Rawicz, 1990),<br />

α − α0 = kBT<br />

8πκ log<br />

<br />

κπ2 /A + γ<br />

κπ 2 /A0 + γ0<br />

<br />

+ γ − γ0<br />

Ka<br />

, (4.3)<br />

where A and A0 are the actual and a reference apparent area respectively, γ0 is<br />

the initial tension corresponding to A0, and α − α0 is the difference between<br />

the actual and initial relative excess area between the apparent and real area.<br />

The tension initially grows exponentially with the relative increase of area<br />

(entropic regime) and its followed by a linear increase that is related with the<br />

direct increase of area per molecule (elastic regime). The crossover tension<br />

between the two regimes is of the order γ ∼ 510 −4 N/m, or equivalently,<br />

α ∼ 2-5% (Evans and Rawicz, 1990), which is not very far from the tension<br />

needed to disrupt the membrane.<br />

Cells have their own mechanisms to regulate membrane tension to new environmental<br />

conditions in order to prevent, for instance, burst under a sudden<br />

decrease of external pressure. There are mainly two regulatory mechanisms:<br />

membrane and bulk exchange through the membrane, Fig. 4.2. The former<br />

involving both vesicular transport between the interior of the cell (Golgi apparatus)<br />

and the cell membrane (endocytosis and exocytosis), and sequestered<br />

membrane in form of membrane invaginations, like in the case of caveolae<br />

(Alberts et al., 2002). The second one utilizes several types of channels<br />

embedded in the membrane that by either active (involving energy consumption)<br />

and/or passive (tension or carrier mediated gates) transport, induce a<br />

bulk flux between the interior and exterior of the cell.


(a) (b)<br />

(c)<br />

4.2 Modeling cell mechanics 57<br />

Fig. 4.2. Different mechanisms of membrane tension regulation. In (a), several examples of the<br />

presence of both active and passive channels in the membrane, which regulate the membrane<br />

tension. (b) Electron micrograph of a fibroblast showing the presence of caveolae in the cellular<br />

membrane. As tension is applied on the membrane, the caveolae may unfold maintaining<br />

the membrane tension constant. (c) Electron micrographs of the formation of clathrin-coated<br />

vesicles from the plasma membrane during endocytosis. Membrane recycling regulates membrane<br />

tension by changing the amount of lipids in the bilayer. (Images from (Alberts et al.,<br />

2002)).<br />

4.2.2 Cytoskeletal cortex<br />

Actin cortex is involved in many mechanical processes of the cell, ranging<br />

from large scale phenomena such as cell locomotion and cell shape remodeling,<br />

to small scale phenomena including endocytosis. Actin forms a dynamical<br />

cross-linked gel that is attached to the membrane, where predominantly<br />

polymerizes 4.3. Actin is cross-linked by several proteins, the most common<br />

being filamin and α-actinin, which concentration affects dramatically the cortex<br />

elastic properties. The average polarization of the cortex points away from<br />

the membrane.<br />

There are other actin associated proteins of the ERM family such ezrin,<br />

which take part on the anchoring of the cortex to the membrane. Finally, actin<br />

filaments can associate to myosin motors forming the so called actomyosin<br />

complexes or microfilaments. Myosin II self-assemble into bipolar filaments<br />

that act on actin generating stress in the cortex by sliding each filament with<br />

respect to each other, Fig. 4.3c. Actomyosin complexes can form very highly<br />

organized structures capable of generating large amount of stress, as in the<br />

case of myofibrils in muscle contraction. In the cortex, actomyosin complexes<br />

organize randomly as part of the highly crosslinked network of actin filaments<br />

with a mesh size of about 10-100 nm.<br />

Actin polymerization is regulated by several cortex proteins such as<br />

formin, profilin or the multiprotein complex Arp2/3 which promote actin<br />

polymerization, and by certain lipids composing the plasma membrane such


58 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

as PIP2 (Raucher et al., 2000; Sheetz et al., 2006). Although the precise mechanism<br />

remains unclear, it is widely accepted that actin in the cortex polymerizes<br />

mainly under the membrane and depolymerizes along the cortex thickness.<br />

As a consequence, the cortex treadmills from the membrane at a velocity<br />

which is roughly the polymerization velocity, Fig. 4.3.<br />

(a) (b)<br />

(c)<br />

h ∼ 500 nm ξ ∼ 100 nm<br />

+<br />

−<br />

−<br />

− − Contractile stress<br />

−<br />

Fig. 4.3. Schematics of the cytoskeletal cortex. (a) The cortex is a thin layer of cross-linked<br />

actin filaments (red) formed under the membrane. (b) The actin network forming the cortex<br />

is characterized by a constant treadmill of material from the membrane. (c) Molecular motors<br />

(Myosin) slide actin filaments with respect to each other creating contractile stress in the<br />

cortex.<br />

Viscoelastic properties of the cysotkeleton<br />

Several rheological techniques may be used to characterize the mechanical<br />

properties of the cortex and the cytoskeleton of the cell in general. These<br />

methods include laser tweezers, AFM and magnetic bead rheology, among<br />

others (Pullarkat et al., 2007). The complex structure and heterogeneity of<br />

the cytoskeleton makes quantitative rheological measurements particularly<br />

difficult. Moreover, elastic properties may be expected to depend on the active<br />

state of the cell. However, the viscoelastic behavior under an external<br />

deformation consisting of an elastic response at short time and a viscous flow<br />

at long time is a general property of the cytoskeleton. Unfortunately, linear<br />

response characterized by a single relaxation time scale (see below) is not sufficient<br />

to account for the cytoskeleton viscoelasticity at all time scales, and<br />

a crossover from a linear viscoelastic behavior to a power law stress stiffening<br />

regime is experimentally observed (Fernández et al., 2006; Trepat et al.,<br />

2007), which corresponds to a continuum of relaxation time scales.<br />

+<br />

+<br />

+<br />

vp<br />

vd<br />


Theoretical modeling of the cytoskeleton<br />

4.2 Modeling cell mechanics 59<br />

The modeling on cell mechanical properties can be divided into two main<br />

approaches, which consider passive, viscoelastic properties of the cell and<br />

active mechanisms of cell mechanics, respectively. The split of the mechanics<br />

of the cell into an active and passive regimes is justified by the different<br />

response times of the cell in rheological experiments, seconds to minutes for<br />

a passive response and several minutes for an active response (Thoumine and<br />

Ott, 1997).<br />

The purely passive elastic response has been modeled considering basically<br />

two different frameworks. Tensegrity, which stands for structures consisting<br />

of elements under tension (actin cables) and elements under compression<br />

(microtubules) (Ingber, 1993), and soft glass rheology, where the competition<br />

between entropy in crosslinked actin network and bending of individual<br />

filaments can account for the crossover from linear elasticity to non-linear<br />

stiffening (Wilhelm and Frey, 2003; Head et al., 2005; Head et al., 2003a;<br />

Head et al., 2003b; Head et al., 2003c)<br />

The model of active mechanisms consists on generalized hydrodynamic<br />

theories of polar maxwell viscoelastic gels formed by polar filaments (actin<br />

or microtubule filaments) which are maintained out of equilibrium by constant<br />

consumption of energy (ATP) (Kruse et al., 2004; Juelicher et al., 2007;<br />

Joanny et al., 2007).<br />

Here, we are interested in the actin cortex, which is a highly dynamic<br />

thin layer, Fig. 4.3. We aim to use the simplest model that accounts for viscoelasticity<br />

and treadmilling, and which includes an active stress driven by<br />

actomyosin complexes. We consider that polymerization occurs mainly in the<br />

cortex, and we assume that the interaction between motors and actin leads to<br />

a contractile effective stress which depends only on myosin concentration.<br />

Maxwell model for actin cortex viscoelasticity<br />

We will work assuming a linear viscoelastic regime for the cytoskeleton,<br />

which is at least observed experimentally for certain conditions (Thoumine<br />

and Ott, 1997). The simplest viscoelastic model that incorporates short time<br />

scale elastic behavior and large time scale fluid behavior is the so called<br />

Maxwell model. For short time scales, the cortex behaves as a solid, and<br />

the internal stress σij obeys σij = 2Euij (Landau et al., 1995), where uij ≡<br />

1/2(∂iu j + ∂ jui) is the strain tensor, ui being the deformation field,and E the<br />

Young modulus of the cortex. In the opposite limit, the cortex behaves as<br />

a fluid, and the stress is given by σij = 2ηuij ˙ (Landau and Lifshitz, 1987),


60 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

where η is the viscosity of the cortex. Defining τ as the crossover time scale<br />

between the two regimes, the viscosity and Young modulus are related by<br />

η ∼ τE, which follows from equating both elastic and viscous stress at the<br />

crossover time τ. The linear interpolatory equation that describe both limiting<br />

cases is given by,<br />

<br />

1 d<br />

+ σij = 2E<br />

τ dt<br />

duij<br />

. (4.4)<br />

dt<br />

In order to illustrate that this equation represent both solid and fluid behavior,<br />

consider a periodic deformation where σij and uij are proportional to e −iωt .<br />

Eq. 4.4 reads,<br />

σij = 2Euij<br />

. (4.5)<br />

1 + i/ωτ<br />

For large frequencies ω ≫ 1/τ, the stress is σ 2Euij and the deformation<br />

corresponds to an elastic solid. For small frequencies ω ≪ 1/τ, the stress is<br />

σ −2iωτEuij = 2τEuij, ˙ which is the usual expression for a viscous fluid<br />

with viscosity Eτ. The microscopic origin of the relaxation time τ is the kinetics<br />

of the cross-linking proteins, which bind and unbind to actin filaments,<br />

creating and breaking crosslinks continuously, relaxing any tension stored<br />

between filaments. Typical orders of magnitude for the cortex parameters are<br />

E = 103 Pa, τ = 10 s, and η = 104 Pa s (Wottawah et al., 2005).<br />

Treadmilling in the cortex induces an inwards convective flow of material<br />

from the membrane with a velocity related to the polymerization velocity, Fig.<br />

4.3. In order for Eq. 4.4 to account for treadmilling, we need to transform the<br />

time derivative d/dt in a convective time derivative, d/dt ≡ ∂/∂t + vi∂i.<br />

The presence of myosin induce a contractile stress in the cortex. At<br />

equilibrium, the myosin stress σm and the cortex elastic stress balance, and<br />

∇σ + ∇σm = 0, where σm > 0 since myosin stress is compressive, see Fig.<br />

4.4.<br />

Considering a gel growing in a cylinder, the force balance in the angular<br />

direction is automatically satisfied, and the radial component reads<br />

∂<br />

∂r r (σrr + σm) − (σθθ + σm)=0, (4.6)<br />

where we are assuming the myosin contraction to be isotropic. The radial<br />

stress applied on the membrane is σrr(R)+σm(R). Myosin needs a characteristic<br />

time ∼ 1/kon to attach to actin, and since the gel grows from the<br />

membrane, this corresponds to a characteristic length from the membrane


(a) (b)<br />

θ<br />

r<br />

h<br />

σmdS3<br />

4.2 Modeling cell mechanics 61<br />

σmdS4<br />

σmdS2<br />

σmdS1<br />

Fig. 4.4. Sketch of the coordinates in the cortex (a), and stress applied by the surrounding<br />

myosin on a piece of cytoskeletal cortex (b).<br />

∼ vp/kon, where myosin has not been able to attach to the membrane. As a<br />

consequence, the density of myosin at the membrane is zero, so the stress applied<br />

on the membrane is only an elastic stress, which depends on the myosin<br />

concentration throughout the gel,<br />

σrr(R)= 1<br />

R<br />

dr(σθθ + σm) ≡<br />

R R−h<br />

γc + γm<br />

, (4.7)<br />

R<br />

where γc and γm are the lateral and myosin tension in the cortex. Similarly,<br />

for spherical geometry the normal stress at the membrane is<br />

σrr(R)= 2<br />

R2 R<br />

rdr (σθθ + σm). (4.8)<br />

R−h<br />

A gel growing away from a curved surface needs to either concentrate or<br />

dilate in order to accommodate the curvature, which induce stress in the gel<br />

that depend on its thickness (Noireaux et al., 2000; Sekimoto et al., 2004).<br />

The contribution of this effect to the normal stress of order O (h/R) 2 .<br />

The cortex thickness is of about 500 nm, which is much smaller than<br />

the radius of the cell, R ∼ 20µ. Using a thin layer approximation, we can<br />

treat the cortex as a two dimensional surface, averaging all quantities along<br />

the radial direction. We can also neglect the geometric contribution to the<br />

stress in comparison to the active stress generated by myosin, which is of<br />

the order O (h/R) (see next section). Therefore, we write first an effective<br />

maxwell equation for a flat surface, and then for small deformations of the<br />

cortex, we relate the normal and lateral stress by force balance. The tangential<br />

component of the stress tensor is


62 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

<br />

1 ∂<br />

+<br />

τ ∂t + vα∂<br />

<br />

β + vp∂z σαβ = 2E ˙u αβ, (4.9)<br />

where the greek indices stand for longitudinal directions along the surface,<br />

and z is the normal direction in which the material convects at a veloc-<br />

ity vp. Averaging this<br />

<br />

last equation with respect to the slab thickness h<br />

, in a thin film approximation, we obtain<br />

Ā ≡ 1/h h<br />

0 A(z)dz<br />

<br />

1 d<br />

+ ¯σ<br />

τ∗ αβ = 2E ˙ū<br />

dt<br />

αβ, (4.10)<br />

where we have assumed that actin polymerizes at the membrane surface with<br />

no lateral stress, σ αβ(z = h) =0, and we have used that in the thin shell<br />

approximation σ αβ(z = 0) ∼ ¯σ αβ. The total time derivative now stands for<br />

the convective derivative over the longitudinal directions, d/dt ≡ ∂t + vα∂ β ,<br />

and τ ∗ ≡ τ + h/vp is an effective relaxation time that includes the effect of<br />

the treadmilling along the normal direction.<br />

As already mentioned in this section, we consider the actomyosin complexes<br />

to create an effective stress which depends only on the density of motors.<br />

Its particular value is a function of the relative orientation of the actin<br />

filaments and myosin concentration in the cortex (Juelicher et al., 2007). The<br />

stress profile of the actomyosin complexes depends both on the treadmilling,<br />

which limits the amount of myosin present in the cortex, and on its bulk density.<br />

In the following section we explicitly calculate the active stress in the<br />

cortex using linear kinetics for myosin.<br />

Contractile stress in the cortex.<br />

We can use linear kinetics to write the local mass conservation for the density<br />

ρ of actomyosin, which is related with the total amount of active stress in the<br />

cortex,<br />

dρ<br />

dt = kon(ρs − ρ) − ko f f ρ, (4.11)<br />

where kon and ko f f are rates of attachment and detachment, and ρs is the maximum<br />

density of actomyosin in the gel, which is related to the density of actin<br />

filaments in the gel, and the total number of bulk myosins. The time derivative,<br />

d/dt ≡ ∂/∂t + v·∇ is a convective derivative that takes into account the<br />

treadmilling of material towards the interior of the cell at a velocity vp. The<br />

steady state profile of myosin is given by


ρ(z)= ˜ρs<br />

<br />

1 − exp<br />

4.2 Modeling cell mechanics 63<br />

<br />

− z<br />

λ<br />

<br />

, (4.12)<br />

where z is the coordinate normal to the membrane, ˜ρs ≡ ρskon/(kon + ko f f )<br />

is the mean density of myosin in the absence of actin turnover, and λ ≡<br />

vp/(kon + ko f f ) is a typical length scale below which convection dominates<br />

myosin binding.. When actin turnover is slow, the attachment length becomes<br />

small and the whole cortex is decorated with myosin. On the contrary, if actin<br />

turnover is very fast compared to attachment kinetics, the attachment length<br />

can be of the order of the cortex thickness The parameter that controls the<br />

amount of density in the cortex is then h/λ, h being the cortex thickness.<br />

Assuming a linear relationship between the actin stress and the myosin<br />

density, and defining ε as the amount of stress per actomyosin, the total lateral<br />

tension due to myosins is,<br />

h<br />

γm ≡ ε<br />

0<br />

ρ(z)=σmsλ<br />

<br />

h<br />

λ +<br />

<br />

exp − h<br />

<br />

− 1 , (4.13)<br />

λ<br />

where σms ≡ ε ˜ρs is the stress that the mean density of myosin would perform<br />

on the gel in the absence of actin turnover. If the gel thickness is very small<br />

compared to the typical length needed for a myosin to attach to the gel (λ ∼<br />

kon/vp), myosin density is strongly limited by the treadmilling process and<br />

the lateral tension is<br />

γm ∼ σms(h 2 /λ), (4.14)<br />

In the opposite limit, myosin kinetics is fast and the maximum stress is<br />

reached almost on the entire gel thickness,<br />

γm ∼ σmsh. (4.15)<br />

Actin turnover and gel thickness affects dramatically the amount of stress<br />

stored in the cortex. The cortex thickness is in turn regulated by several mechanisms,<br />

including the stress stored in the cortex. In next section, we briefly<br />

review the mechanisms that may fix the cytoskeleton thickness.<br />

Cytoskeletal cortex thickness<br />

Actin filaments grow preferentially from the membrane due to the presence<br />

of actin polymerization promoters in the membrane (PIP2 as an example<br />

(Raucher et al., 2000; Sheetz, 2001)). After a certain length, actin filaments<br />

are cross-linked by proteins present in the cortex (mostly filamin and


64 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

α-actinin (Alberts et al., 2002)) becoming a meshwork with gel like properties<br />

that treadmills at a velocity given by the polymerization velocity vp. The<br />

equilibrium between polymerization and depolymerization fixes the steady<br />

state thickness of the cortex. Considering that the main contribution from<br />

actin depolymerization takes place at the gel surface,<br />

˙h = vp − vd(h), (4.16)<br />

where vp and vd are the polymerization and depolymerization velocities respectively.The<br />

gel will reach its steady state when the depolymerization velocity<br />

at the gel surface equals the polymerization velocity, vp = vd(h). Three<br />

mechanisms may control the thickness of the gel: biochemical regulation,<br />

monomer diffusion, and stress induced depolymerization.<br />

The amount of actin monomer concentration determines the polymerization<br />

and depolymerization velocity of actin. Moreover, several proteins including<br />

Cofilin regulate the turnover rate of actin by adjusting actin monomer<br />

concentration (Carlier et al., 1997), providing the cell with a mechanism to<br />

regulate its cortex thickness.<br />

Since the cortex polymerizes mainly from the surface, actin monomers<br />

must diffuse along the cortex thickness, and hence, the polymerization velocity<br />

can be limited by this process (Noireaux et al., 2000).<br />

All symmetry-breaking models (Sekimoto et al., 2004; van Oudenaarden<br />

and Theriot, 1999; Mogilner and Oster, 2003) consider a coupling between<br />

tensile stress and gel dissociation. The microscopic origin of the dissociation<br />

can be the unbinding of cross-links in the cortex, which disrupts the gel-like<br />

structure, or direct depolymerization of actin filaments. Both depolymerization<br />

and disruption of the gel structure can be effectively described with a<br />

depolymerization velocity which increases with the tensile stress. Kramers<br />

rate theory (Kramers, 1940; Plastino et al., 2004) suggests an exponential<br />

dependence of the depolymerization velocity with the stress (Noireaux et al.,<br />

2000; Plastino et al., 2004; Gerbal et al., 1999; Juelicher et al., 2007),<br />

vd = v 0 d exp<br />

<br />

σ<br />

, (4.17)<br />

where v 0 d is the depolymerization velocity at vanishing stress, and σ0 is<br />

a reference stress related to actin and cross-linker kinetics. We will assume<br />

that the polymerization velocity at the membrane is constant and that the gel<br />

is growing without lateral stress. Notice that the a finite thickness solution<br />

is ensured by the increase of active contractile tension due to myosin as a<br />

function of the cortex thickness (see Eq. 4.13) (Juelicher et al., 2007).<br />

σ0


Mechanical equilibrium of the cell<br />

4.2 Modeling cell mechanics 65<br />

In section 4.2.2, we have considered the amount of stress generated by myosin<br />

in the cortex (see Eq. 4.13). Myosin stress is transmitted to the membrane via<br />

deformation of the actin cortex and via proteins that link the actin cortex<br />

to the membrane. The normal stress transmitted to the membrane is given<br />

by 2(γc + γm)/R, where γc is the lateral tension in the cortex and γm is the<br />

myosin tension (see Eq. 4.7), and where we are neglecting corrections of<br />

order O (h/R) 2 .<br />

At steady state, the shape of the cell is determined by the balance of the<br />

outward forces from the osmotic pressure and the inward forces from the<br />

membrane and cortex, Fig. 4.5. This static force balance (membrane not moving),<br />

leads to the Laplace law including the tension in the cortex,<br />

∆P = 2 γ + γc + γm<br />

. (4.18)<br />

R<br />

At static equilibrium, the strain rate vanishes ( ˙R = 0), and as a consequence,<br />

the lateral stress in the cortex is σθθ = 0 1 (see Eq. 4.10). Then, the only<br />

tension contribution at equilibrium from the cortex corresponds to myosin<br />

tension, γm. A cell may thus be pressurized by increasing myosin activity or<br />

density.<br />

Laplace law stated above (Eq. 4.18), is only valid if the cortex stress is<br />

transmitted to the membrane. The cortex is adhered to the membrane by actin<br />

associated proteins which are responsible for force transmission. Failure of<br />

these links leads to separation of the membrane from the cytoskeleton. The<br />

next chapter is devoted to the stability of the membrane and cortex adhesion<br />

and its consequences to the cell behavior.<br />

1 Here we are neglecting the stress created by the curvature which is of the order O (h/R) 2 .


66 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

(a) (b)<br />

Pc<br />

P0<br />

Fig. 4.5. Schematics of the forces involved in the cell at steady state. In (a), the internal, Pc,<br />

and external, P0, pressures exert a normal force on the membrane and cortex. The normal projection<br />

of the myosin tension in the cortex (b) is transmitted to the membrane through proteins<br />

that link the cortex and the membrane, and balances the resulting force at the membrane.


4.3 Membrane-cortex adhesion<br />

4.3 Membrane-cortex adhesion 67<br />

In this section we describe the membrane-cortex adhesion using two different<br />

approaches:<br />

(i) The first approach uses a kinetic model for its linkers, where we will<br />

consider the cortex as a flat and rigid surface, and we will leave for forthcoming<br />

sections the effect of the cytoskeleton visco-elasticity. We will find the<br />

adhesion stability conditions as a function of kinetic parameters of the linkers<br />

and the total force applied on the membrane, which may result from pressure,<br />

cytoskeleton tension or any external perturbation (such as an osmotic<br />

shock or the action of a micropipette aspiration). As a result, we will obtain<br />

that cooperativity between bonds determines the stability of membranecortex<br />

adhesion: below a critical force links can to the load and the adhesion<br />

of membrane and cytoskeleton is stable. Above a critical force, there is a<br />

global link failure and an abrupt transition from a stable adhesive membrane<br />

to an unstable membrane detached from the cytoskeleton. Finally, we will<br />

show how adhesion depends critically on the pre-stressed state of the cortex<br />

and the density of linkers, and we will compare our model with quantitative<br />

experiments found in the literature (Merkel et al., 2000).<br />

(ii) The second approach is an energetic description of the detachment<br />

process, which considers the cost of creating a protrusion detached from the<br />

cortex. Balancing the energy cost of membrane-cytoskeleton detachment and<br />

the membrane tension with the energy gain from the pressure, we find an<br />

abrupt transition for membrane detachment.<br />

4.3.1 Kinetic model for membrane-cortex adhesion<br />

The sketch of the system is depicted in Fig. 4.6. The density of links ρb between<br />

membrane and cortex is represented by springs that attach and detach<br />

continuously at rates kon and ko f f respectively. A force per unit area f is<br />

shared by the bonds, that are stretched by a small amount x, which is a molecular<br />

length scale. When the link and membrane forces are not balanced, the<br />

membrane flows and the stretching increases at a rate given by viscous dissipation.<br />

In appendix 4.A, we show that the dominant term in the dissipation<br />

is due to the flow of cytosol through the cortex meshwork, which in terms<br />

of dissipation per unit area is ˙E ∼ ˙x 2 ηch/ξ 2 , where ηc is the cytosol viscosity<br />

(∼ 3.10 −3 − 2.10 −1 Pa s (Charras et al., 2008)), h is the thickness of the<br />

cortex (∼ 500 nm), and ξ is the cortex mesh size (∼ 10 − 100 nm).<br />

Defining the effective viscosity per unit length, η ≡ ηch/ξ 2 , a free membrane<br />

under a pressure ∆P would move at a velocity ˙x ∼ ∆P/η, and would


68 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

(a) (b)<br />

x<br />

ξ<br />

kxρb<br />

f<br />

η ˙x<br />

δ<br />

koff<br />

Fig. 4.6. (a) Schematics of the links between membrane and cortex. (b) Linkers have an association<br />

and dissociation rate that depend on the energy profile of the link (black curve). When<br />

a force is applied on the bond, its energy landscape and associated rates are modified (red<br />

curve).<br />

need a typical time t ξ ∼ ξη/∆P to inflate to a distance ξ , which is the mean<br />

distance between links (see Fig. 4.6a). For typical pressure values in micropipette<br />

experiments, ∆P ∼ 100 − 1000 Pa (Merkel et al., 2000; Rentsch<br />

and Keller, 2000), t ξ is of about 10 −6 − 10 −3 s. If we compare this time with<br />

a dissociation attempt frequency starting at about ∼ 10 9 − 10 10 s −1 (Evans,<br />

2001), we realize that the inflation time needed for spatial correlation between<br />

neighboring links (that is, the time needed to inflate to a distance ξ ) is several<br />

orders of magnitude larger than the typical link kinetics time scale 2 . As a<br />

consequence, we can assume that links are not spatially correlated although<br />

they cooperate to overcome the force applied on the membrane. Basically,<br />

there are a lot of attachment and detachment events while the membrane is<br />

moving. This image breaks down if the detached patch of membrane is sufficiently<br />

large so that the links in the center have no chance to rebind as the<br />

patch of membrane inflates.<br />

We will use the mean density of bonds when considering the force balance<br />

at the membrane, which is justified by the fast kinetics of attachment and<br />

detachment,<br />

kon<br />

η dx<br />

dt = f − kxρb, (4.19)<br />

where k is the stiffness of the links, which we take of about 10 −2 pN nm. The<br />

linear kinetics equation for the number of attached links reads,<br />

2 The threshold pressure for spatial correlation is calculated considering the pressure needed<br />

to inflate the membrane a length ∼ ξ in a time ∼ 1/kon, and is of about 10 7 Pa. We will see<br />

that the links typically become unstable for much smaller pressures.


4.3 Membrane-cortex adhesion 69<br />

dρb<br />

dt = kon(ρ0 − ρb) − ko f f ρb, (4.20)<br />

where ρ0 is the saturation density for bound links (ρ0 ∼ 1/ξ 2 ). The kinetic<br />

rates kon and ko f f are determined by the energy landscape of the link, Fig.<br />

4.6b. The energy barrier to overcome in order to dissociate a link changes as<br />

an external force is applied (or equivalently, an external potential −xf), Fig.<br />

4.6b, and the off rate increases exponentially with the applied force (Kramers,<br />

1940; Kampen, 2007),<br />

ko f f = k 0 o f f exp[kxδ/kBT ], (4.21)<br />

where δ ∼nm is a molecular scale of the link which is related to the maximum<br />

of the energetic barrier (see Fig. 4.6b), and kx is the stretching force on a<br />

link. The rate of rebinding, kon depends in general on the substrate and link<br />

stiffness (Evans, 2001), but for the sake of simplicity, we assume it here to<br />

be constant 3 .<br />

Before considering the dynamical system given by equations Eq. (4.19)<br />

and (4.20), it is convenient to rescale the variables with the relevant scales to<br />

identify the dimensionless parameters that govern the stability of the system.<br />

Time is rescaled by the attachment time scale tb ≡ 1/kon, and τ ≡ t/tb. The<br />

stretching length scale is related with the typical force per link and the stiffness<br />

ls ≡ f /(kρ0), and z ≡ x/ls. Finally, we rescale the link density by the<br />

saturation density ¯ρb ≡ ρb/ρ0, and the set of equations Eq. (4.19) and (4.20)<br />

become,<br />

ζ dz<br />

dτ = 1 − z ¯ρb, (4.22)<br />

d ¯ρb<br />

dτ = −(1 + χ expαz) ¯ρb + 1, (4.23)<br />

where ζ ≡ η/(ktb) is the ratio between hydrodynamic relaxation and link<br />

kinetics time scales (ζ ≫ 1), χ ≡ k 0 o f f /kon compares the on and off rates, and<br />

α =<br />

f /ρ0<br />

, (4.24)<br />

kBT /δ<br />

is the ratio of the typical force per link and a typical force from thermal noise.<br />

We investigate the linear stability of the stationary points of these two nonlinear<br />

equations, ˙z = ˙¯ρb = 0, given by<br />

3 We have considered several dependences of kon on both geometry and force and the qualitative<br />

behavior of our analysis remain unchanged.


70 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

¯ρbs = 1<br />

, (4.25)<br />

zs<br />

zs − 1 = χ expαzs. (4.26)<br />

A stationary stretching that is finite and linearly stable, corresponds to a stable<br />

adhesion between membrane and cortex. The complete unbinding of the<br />

membrane from the cortex corresponds to a vanishing density of links (or<br />

equivalently, a diverging stretching (Eq. 4.25)). ). Equations Eq. (4.25) and<br />

(4.26) do not have solutions in general for every value of the parameters χ<br />

and α. The parameter space that defines the region of stationary solutions is<br />

given by the condition α < α ∗ (Fig. 4.7a), with<br />

χ =<br />

1<br />

α∗ exp(1 + α∗ . (4.27)<br />

)<br />

The space of stationary solutions ( Eq. (4.27)), that corresponds to a mem-<br />

(a) 3<br />

(b) 30<br />

χ<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

SOLUTION<br />

SPACE<br />

NO SOLUTION<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

α<br />

z<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

NO<br />

SOLUTION<br />

0.1 0.15 0.2 0.25 0.3 0.35 0.4<br />

Fig. 4.7. (a) Parameter space for the existence of stationary solutions. Notice that the only relevant<br />

parameters are the ratio of on and off rates χ = k 0 o f f /kon and the rescaled force per bond<br />

α = f /(kBT /δ). (b) Stationary normalized stretching z as a function of the normalized force<br />

α for χ = 0.9. Unstable solution (dotted red), stable solution (solid green) and the condition<br />

for solution stability, z < (1 + α)/α (solid black).<br />

brane bound to the cortex, may contain regions of stable or unstable solutions<br />

depending on the value of the different parameters. The use of linear stability<br />

analysis allows to understand the behavior of the stationary solutions under<br />

small perturbations δρb ≡ ρb − ρbs and δz ≡ z − zs. The evolution of a linear<br />

perturbation around the stationary solution of Eq. (4.22) and (4.23) is given<br />

by δ ˙z<br />

δ ˙ρb<br />

<br />

= 1<br />

<br />

− 1/ζ<br />

zs α(1 − zs)<br />

− z 2 s /ζ<br />

− z 2 <br />

δz δz<br />

≡ Λ , (4.28)<br />

s δρb δρb<br />

α


4.3 Membrane-cortex adhesion 71<br />

The stability is obtained from the eigenvalues of Λ, or equivalently, its determinant<br />

and trace . Since the trace Tr(Λ), is always negative, the stationary<br />

solution is stable if the determinant is positive, det(Λ) > 0, or equivalently, if<br />

1 + α(1 − zs) > 0. Using Eq. 4.26 we find that the condition that defines the<br />

stable parameter space is exactly the same condition for solution existence<br />

defined in Eq. 4.27 . Moreover, since there are always a pair of stationry<br />

solutions, (z 1 s ,z 2 s ), the condition for positive determinant implies that within<br />

this pair of stationary points, the smaller corresponds to the stable solution<br />

whereas the larger to the unstable one, see Fig. 4.7b. In other words, the solution<br />

that corresponds to the links being closer to the cortex and equivalently,<br />

larger link density, is the stable solution, and ceases to exist for α > α ∗ (see<br />

Eq. 4.27).<br />

It is striking that the linear stability of the solution does not involve the<br />

friction η, or in terms of rescaled quantities, ζ . Intuitively one should expect<br />

that links are stable if within the time of a detachment and rebinding event,<br />

the membrane barely move, otherwise, the stretching of the link becomes too<br />

large to be sustained. In other words, one would expect that the stability of the<br />

bonds is strongly dependent on the ratio between hydrodynamic relaxation<br />

(how fast the membrane moves) and bond kinetics, which is the definition<br />

of ζ . The reason for ζ not playing any role in the stability of the stationary<br />

solutions, comes from the fact that at linear level, the stability is dictated by<br />

the sign of the slope in the perturbations close to the stationary point (whether<br />

the tendency is to go towards or away the stationary point), but not by its value<br />

(in this case, given by ζ ). When the perturbation is strong enough, we need<br />

to include non-linear terms in order to know the evolution towards the stable<br />

or unstable branch. The non-linear analysis would depend on the values of<br />

ζ since it determines whether the bond kinetics is fast enough to drive the<br />

system back to equilibrium or, whether, if the membrane has moved too far<br />

for the bonds to drive it back. The nonlinear effects results in a regime near<br />

the unbinding transition that is nonlinearly unstable.<br />

Our discussion reveals that membrane-cortex adhesion is stable for a<br />

limited range of forces that depend only on the link kinetic parameters<br />

(χ = k 0 o f f /kon), Fig. 4.7, and becomes unstable for a critical force per link,<br />

that as a function of the critical rescaled force α ∗ , is<br />

f ∗ b = kBT<br />

δ α∗ , (4.29)<br />

where α ∗ is the critical rescaled force given by Eq. 4.27. The origin of the<br />

force f on the membrane is in general the pressure applied on the membrane,<br />

which origin include external perturbations (the action of micropipette aspi-


72 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

ration or osmotic shocks) and active terms (myosin activity in the cytoskeleton).<br />

From section 4.2.2, we know that the force per unit area that acts on the<br />

bonds at equilibrium, ∆P − 2γ/R is proportional to the tension generated by<br />

the presence of myosin in the cortex (Eq. 4.18), feq = 2γm/R, and the rescaled<br />

force is<br />

<br />

α = ∆P − 2 γm<br />

R<br />

<br />

δ<br />

ρ0kBT<br />

, (4.30)<br />

where R is the radius of curvature. High internal pressure, a low membrane<br />

tension or a high myosin activity can all contribute to membrane unbinding.<br />

4.3.2 Micropipette experiments as membrane-cortex adhesion probes<br />

Micropipette experiments allow to quantitative probe membrane-cortex adhesion.<br />

The size of the detachment area is fixed by the radius of the pipette, and<br />

the pressure in the micropipette is controlled. We use the model discussed in<br />

the previous section, which treats properly the link kinetics, to interpret some<br />

experimental measurements (Merkel et al., 2000).<br />

During a micropipette experiment, a drop of pressure is applied on a region<br />

of the cell (∼ R 2 p) and the force per unit area with respect to the equilibrium<br />

state,<br />

feq = 2 γm<br />

R<br />

γ<br />

= ∆P − 2 , (4.31)<br />

R<br />

is increased by an amount Pe − Pp ≡ ∆P ′ , where Pe is the pressure of the external<br />

cellular media and Pp is the pressure inside the micropipette. Assuming<br />

the perturbation to be instantaneous 4 , and the cytoskeleton to be a rigid (non<br />

deformable) surface, the total force per unit area acting on the links after the<br />

micropipette suction is,<br />

f = ∆P ′ + 2 γm<br />

, (4.32)<br />

R<br />

where the second term in the right hand side is the pre-stressed state of the<br />

cell, or equivalently, the equilibrium force per unit area.<br />

Experimental results are often expressed in terms of the percentage of<br />

cells for which the plasma membrane is detached from the cytoskeleton at<br />

a given pressure (Merkel et al., 2000). We expect the membrane to abruptly<br />

detach from the cortex above a critical pressure. These results vary significantly<br />

with both Talin (a protein that links actin and membrane) and myosin<br />

concentration, see Fig. 4.8.<br />

4 In section 4.4 we will treat carefully the response of the cortex.


Influence of the link concentration<br />

4.3 Membrane-cortex adhesion 73<br />

The critical rescaled force per link α ∗ depends only on χ, the ratio between<br />

on and off rates, which is an intrinsic property of the links (Eq. 4.27). For a<br />

given cell type, we can assume that χ is fixed, and consequently α ∗ . In the<br />

previous section we have seen that there is detachment for a critical force<br />

per link. Recalling the relation between the critical rescaled force α ∗ and the<br />

critical force per unit area f ∗ (Eq. 4.24),<br />

f ∗ ∗ kBT<br />

= ρ0α , (4.33)<br />

δ<br />

we realize that the attachment can be weakened either by decrease of the<br />

available density of links (in the mentioned experiments (Merkel et al., 2000),<br />

Talin), or by decreasing the cortex density. In both cases, the critical force for<br />

unbinding should decrease linearly.<br />

decreasing the available density of links, which can be obtained by decreasing<br />

the total amount of actin associated proteins (in the mentioned experiments,<br />

Talin (Merkel et al., 2000)), or decreasing the linking sites available<br />

in the actin cortex (by weakening the cortex using latrunculin A or cytochalasin<br />

D; or decreasing filamin concentration which increases the mesh size),<br />

should decrease linearly the force needed to induce the instability.<br />

In wild type amoeba (Merkel et al., 2000), Fig. 4.8a, there is a clear shift<br />

of about ∼ 150−200 Pa between the pressure needed to unbind membrane at<br />

the front and at the rear part of the cell, which corresponds to the more active<br />

and the more stable part of the cell respectively. This shift can be attributed to<br />

a different concentration of Talin between these two regions.This is supported<br />

by experiments performed in a mutant lacking Talin, for which the threshold<br />

pressure is the same in the rear and the front, and also exactly equal to the<br />

front of the wild type cell, Fig. 4.8b.<br />

In the case of a mutant lacking Myosin II, the values for the micropipette<br />

critical pressure may differ from the wild type cell because the total force<br />

per unit area is a combination of both pressure and myosin driven tension<br />

(see paragraph below), but the shift in pressure between a myosin null cell<br />

with Talin and without Talin should be of the same order as in the wild type<br />

cell, according to our predicted linear relation between force and link density<br />

(Eq. 4.33). In Fig. 4.8c-d, those results are reported obtaining a difference<br />

between the front and the rear part of the cell of about 150 − 500 Pa, which<br />

is of the same order as in the wild type cell. On the other hand, the absolute<br />

value in critical micropipette pressures in the wild type and the null-myosin<br />

cell differs by a factor of five. This increase in the stability of membranecortex<br />

adhesion in the absence of Myosin can also be accounted with our


74 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

(a)<br />

aspirated cells (%) aspirated cells (%)<br />

suction pressure (Pa)<br />

−Talin<br />

(c)<br />

(b) (d)<br />

suction pressure (Pa)<br />

aspirated cells (%)<br />

aspirated cells (%)<br />

−MII<br />

suction pressure (Pa)<br />

−MII<br />

−Talin<br />

suction pressure (Pa)<br />

Fig. 4.8. Percentage of cells for which the membrane is unbound from the cytoskeleton in<br />

the micropipette for wild-type (a), Talin null (b), Myosin null (c), and Myosin and Talin null<br />

cells (d). Figure modified from (Merkel et al., 2000). The circles correspond to the cell front,<br />

and triangles to the cell rear. Knockout of Talin leads to the same threshold pressure for the<br />

cell front, but changes the cell rear to the same value as the cell front, indicating that Talin<br />

concentrates on the cell rear. In the absence of myosin the pressure for unbinding is much<br />

larger.<br />

model considering the explicit relation between the force per unit area and<br />

the micropipette pressure plus myosin driven cortical tension, as we explicitly<br />

discuss in the following paragraph.<br />

Effect of mysoin II activity<br />

The effect of myosin activity in micropipette experiments leads to intriguing<br />

results. As reported in (Merkel et al., 2000), the absolute value in critical micropipette<br />

pressures in a wild type and a myosin null cell differs by a factor of<br />

five. Consequently, the stability of membrane-cortex adhesion is enhanced by<br />

reducing myosin activity, an effect that in (Merkel et al., 2000) is attributed<br />

to the myosin induced softening of the cortex.<br />

This effect can be explained very naturally by our model: given an available<br />

density of links, ρ0, the critical force per unit area, f ∗ , is fixed by Eq.<br />

4.33. Since the force is a combination of myosin driven cortical tension and


4.3 Membrane-cortex adhesion 75<br />

micropipette pressure, (Eq. 4.32), we expect that a decrease of myosin activity<br />

(which reduces the cortical tension) should lead to an increase of the<br />

critical micropipette pressure,<br />

∆P ′ ∗ kBT<br />

= ρ0α<br />

δ<br />

− 2γm . (4.34)<br />

R<br />

Because of Myosin activity, in a wild-type cell, the cortical tension in the<br />

cortex is different from zero, and the links are pre-stressed. As a consequence,<br />

a weaker pressure suction is needed to unbind the membrane from the cortex,<br />

which occurs for moderate pressures of about 200 − 300 Pa. Contrarily, in a<br />

myosin-null cell, the cortical tension in the cortex vanishes and the links that<br />

are not initially pre-stressed, leading to a tighter adhesion between membrane<br />

and cortex. A larger micropipette pressure difference is needed in this case in<br />

order to unbind the membrane from the cortex, see Fig. 4.9.<br />

Comparing the critical pressures in both wild type and myosin-null cells<br />

(Fig. 4.8 a-c or b-d), we can estimate the amount of myosin driven tension in<br />

a cell,<br />

γm = R<br />

2 (∆P−MII − ∆P), (4.35)<br />

where R stands for a radius between the cell and micropipette radius. The typical<br />

amoeba radius in the reported experiments is of about 10 µm (Merkel<br />

et al., 2000), resulting in a myosin tension of about γm ∼ 5×10 −3 N/m, which<br />

is at least two orders of magnitude of the typical membrane tension of a vesicle,<br />

contributing to the 80% of the force needed to unbind the membrane from<br />

the cortex in non pre-stressed links, or in terms of aspiration pressure, ∼ 800<br />

Pa from the ∼ 1000 Pa needed to detach the membrane. Myosin driven tension<br />

is thus the main force opposing the outward pressure from the cell and<br />

giving shape to the cells, the membrane acting only to transmit the internal<br />

pressure to the cortex.<br />

A slightly increase of myosin activity, in particular a 25% increase in<br />

myosin driven contraction, which in terms of aspiration pressure represents<br />

∼ 200 Pa, leads to instability (∼ 1000 Pa, see Fig. 4.8). As a consequence,<br />

amoebas are close to the instability threshold in normal conditions.<br />

The advantage of being close to the instability threshold using myosin<br />

contraction, that accounts for the 80% of the critical force, could be related<br />

with the ability of the cell to fine tune its response under external stimuli<br />

(food or undesired environmental conditions) initiating cell movement by a<br />

localized influx of calcium through cell membrane. Asymmetric distribution<br />

of Talin is another possible way to drive direct motility of cells.


76 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

Using the myosin-null experiments, Fig. 4.8c, in which case the micropipette<br />

pressure is directly related with the available density of links (Eq.<br />

4.34) we can estimate the relative concentration of Talin ρt with respect to<br />

the total linker concentration ρ0,<br />

ρt<br />

ρ0<br />

= ∆P−MII − ∆P −MII,T<br />

∆P −MII , (4.36)<br />

which is of about 10 − 30% of the saturation density ρ0. Assuming the saturation<br />

density to be limited by the mesh size, ρ0 ∼ 1/ξ 2 ∼ 100 links/µm 2 ,<br />

Talin density should be of about 10 links/µm 2 . This small density of Talin<br />

links seems to be enough to drive direct motion in amoebas by its asymmetric<br />

distribution. It remains to be unveiled how those proteins are distributed<br />

unevenly within the membrane.<br />

We have not estimated yet the values for α ∗ or χ. Using a value of 100<br />

links/µm 2 for the saturation density of links and a link length of the order of<br />

nanometers, we find that<br />

α ∗ = ∆P−MII δ<br />

, (4.37)<br />

kBT ρ0<br />

is of about 4. In other words, the critical force per link (Eq. 4.29) is four times<br />

the thermal force of the link, and is of the order of 16 pN. From the estimates<br />

of α ∗ , we can use Eq. 4.27 to evaluate the ratio of on and off rates, χ ∼ 10 −3 .<br />

We have seen that understanding membrane-cortex adhesion require a<br />

proper account of myosin activity and link density.How this microscopic kinetic<br />

description links with a continuum coarse grained description by means<br />

of an adhesion energy is addressed in the following paragraph. Many experiments<br />

claim to measure adhesion energy as if it was a constitutive parameter<br />

of the cell. We have shown that this is not the case, and adhesion depends on<br />

the state of the cell, in particular on myosin activity. We will give an explicit<br />

dependence in terms of myosin tension and link density, in the framework of<br />

a rigid cortex 5 .<br />

Adhesion energy<br />

Small interlink distances (∼ 100 nm) compared to a typical micropipette radius<br />

of few µm, suggests a continuum energetic description for the membranecortex<br />

adhesion (Sheetz, 2001; Merkel et al., 2000; Brochard-Wyart et al.,<br />

5 Considering the cortex as a rigid surface leads to an over estimation of the adhesion energy<br />

but qualitatively predicts the same behavior.


∆Pp(Pa)<br />

1800<br />

1600<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

15 Pa µm 2<br />

4.3 Membrane-cortex adhesion 77<br />

0<br />

0 20 40 60 80<br />

ρt =0<br />

100 120<br />

ρ0(1/µm 2 )<br />

1/ξ 2<br />

∆P −MII<br />

p<br />

∆P −MII,T<br />

p<br />

∆Pp<br />

∆P −T<br />

p<br />

Fig. 4.9. Theoretical predictions for the critical aspiration pressure in a micropipette experiment<br />

as a function of the available density of links ρ0 according to Eq. 4.34. In solid black<br />

and red, a cell containing and lacking myosin II respectively. The slope is the same in both<br />

cases and equal to 15 Pa µm 2 (see Eq. 4.34). The red dashed line shows the case for Talin null<br />

mutants (as reported in (Merkel et al., 2000), whereas black dashed line corresponds to a cell<br />

with an available link density of order one linker per mesh size.<br />

2006). The value of this adhesion energy is considered as a parameter accessible<br />

through micropipette or tether extrusion experiments. However, the<br />

value of the adhesion energy should in general be a function of the density of<br />

links, which in turn depend on the state of the cell. As a consequence, the relationship<br />

between pressure applied (in the case of a micropipette aspiration)<br />

and the adhesion energy is far from being straight forward.<br />

In our kinetic model, the work needed to detach the membrane is the external<br />

force applied on the membrane times the displacement before unbinding,<br />

which is given by the difference of the critical stretching x ∗ and the initial<br />

equilibrium stretching xeq, see Fig. 4.10,<br />

w = f ∗ (x ∗ − xeq), (4.38)<br />

where we assume that x ∗ −xeq is very small. In other words, we need to bring<br />

the membrane to a distance from the cortex that corresponds to the critical<br />

stretching (see Fig. 4.7b). As in the case of the critical aspiration pressure<br />

in a micropipette experiment, the adhesion energy will depend on the level<br />

of stress in the cytoskeleton, which determine the mean number of attached<br />

links to be broken, and consequently, it will depend on the available density


78 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

of links ρ0 and on the myosin driven cortical tension γm. In what follows<br />

we will derive the expression of the adhesion energy and we will obtain a<br />

non linear relation with the link density and the myosin tension, contrarily<br />

to the intuitive linear dependence on the available density of links. As an<br />

example, we will consider the adhesion energies that would correspond to the<br />

experiments done by (Merkel et al., 2000) in wild type and mutant amoebas<br />

lacking myosin II or Talin.<br />

f<br />

δx<br />

xeq<br />

Fig. 4.10. Schematics of the work needed to unbind the membrane. Initially, the cell is at<br />

equilibrium and the stretching is given by xeq. In order to unbind the membrane, we need to<br />

apply a force that drives the links towards the critical stretching x ∗ .<br />

In terms of rescaled quantities (z = xkρ0/ f , ¯ρ0 = ρ0ξ 2 and α = f δ/(kBT ρ0)),<br />

we can express the adhesion per unit area (Eq. 4.38) as<br />

<br />

w = w0 ¯ρ0 1 − zeqαeq<br />

z∗α ∗<br />

<br />

, (4.39)<br />

where w0 ≡ (kBT /δ) 2 α ∗ (1 + α ∗ )/(kξ 2 ) is the maximum adhesion energy<br />

that corresponds to non pre-stressed links in the cortex (no myosin activity),<br />

and, zeq and αeq are the equilibrium rescaled stretching and force respectively.<br />

At equilibrium, the rescaled force depends on the myosin driven cortical tension<br />

which sets the equilibrium of the cell and the force per link by Eq. (4.32),<br />

feq = 2γm/(Rρ0),<br />

x ∗<br />

αeq = 2γm δ<br />

, (4.40)<br />

ρ0R kBT<br />

which using the values obtained in the previous paragraph for myosin tension<br />

γm ∼ 5.10 −3 Nm −1 , leads to a rescaled force of about 3, where we have<br />

assumed the available density of links to be ∼ 1/ξ 2 . In the case of the mutant<br />

lacking Talin, the link density is reduced by a 10 − 30%, and αeq ∼ 4.


4.3 Membrane-cortex adhesion 79<br />

The equilibrium rescaled stretching is related with the equilibrium rescaled<br />

force by Eq. 4.26. Finally, the critical rescaled force and stretching are given<br />

respectively by Eq. (4.26) and (4.27).<br />

2 w<br />

ρ0ξ<br />

w0<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

-Talin<br />

0<br />

0 1 2 3 4<br />

αeq<br />

-MII<br />

-T, -MII<br />

Wild type<br />

Fig. 4.11. Characterization of the effective binding energy. Reducing the link density (-T) has<br />

two effects: decrease the global factor in the adhesion energy that is proportional to the link<br />

density, but also increase the pre-stressed state of the cell. Myosin perturbations (-MII) set the<br />

value of the binding energy in the curve.<br />

The expression for the adhesion energy per unit area, Eq. (4.39), contains<br />

the intuitive linear dependence with the saturation number of links, w ∼ w0 ¯ρ0,<br />

which only depends on intrinsic properties of the links, namely the on and off<br />

rates, and the stretching constant of the links; and a correction that takes into<br />

account the pre-stressed state of the cell. As the link density is reduced, not<br />

only the adhesion energy decreases linearly, but also the level of pre-stress on<br />

each link increases, thus reducing significantly the adhesion energy, Fig. 4.11.<br />

The linear dependence in the link density of the adhesion energy w0 ¯ρ0 should<br />

only be observed in the absence of myosin II. As a consequence, it does not<br />

make sense to infer from a particular experiment an absolute adhesion energy<br />

for a certain cell, since the particular measured value will depend in general<br />

on the state of the cell which is in most cases unknown. In the experiments<br />

done in wild type and mutant amoebas by (Merkel et al., 2000), we expect<br />

four different values for the adhesion energy for the wild type and the three<br />

different mutants. Using for the link stiffness k ∼ 10 −2 pN nm −1 (Evans,<br />

-T


80 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

2001), the maximum (bare) adhesion energy w0 is about 3 mJ m −2 , and this<br />

is what would be measured in a mutant lacking myosin II, i.e., an equilibrium<br />

state which is not pre-stressed. For a mutant lacking Talin and myosin II<br />

(αeq = 0), the adhesion energy is reduced by a 10 − 30%. For a wild type cell<br />

and a mutant lacking Talin (αeq ∼ 3 and 4 respectively) the adhesion energies<br />

are reduced by a 60% and 80% respectively, see Fig. 4.11 and Table. 4.1. The<br />

dramatic increase in the adhesion energy for a cell lacking myosin activity<br />

but with the same density of links available, which can be of the order of<br />

the 250%, illustrates the importance of the equilibrium pre-stressed state of a<br />

cell in the experimental measurements of adhesion energies and detachment<br />

pressures.<br />

Cell type w (mJ m −2 )<br />

Wild type 1.2<br />

Talin-null 0.6<br />

Myosin II-null 3<br />

Myosin II and Talin-null 2.4<br />

Table 4.1. Table for the predicted adhesion energies w for wild type and mutants lacking<br />

myosin or Talin, corresponding to the experiments by (Merkel et al., 2000).<br />

So far we have considered a description for the membrane-cortex adhesion<br />

that describes properly the dynamics of the linkers. For the sake of simplicity<br />

we have considered the cortex as a flat rigid surface, which yields to an<br />

overestimation of the adhesion energy and the threshold for membrane unbinding.<br />

The effect of the cortex viscoelasticity will be explicitly considered<br />

in forthcoming sections. Additionally, membrane unbinding usually involves<br />

a characteristic nucleation length scale, as in the case of blebbing cells (Charras<br />

et al., 2006; Charras et al., 2008), which our present mean field description<br />

considering no spatial information is unable to describe. Adding lateral correlation<br />

in the model is straightforward conceptually (Garrivier et al., 2002) but<br />

involves extra nonlinearities which do not allow analytical treatment. Instead,<br />

in the next section we propose a coarse grained description based on energy<br />

considerations (Sheetz et al., 2006; Charras et al., 2008).<br />

4.3.3 Energetic description of a bleb<br />

The previous description of membrane-cortex adhesion assumes no lateral<br />

correlation between the links, which corresponds to small deformations of<br />

the membrane. When a large enough patch of membrane detaches (Sb ≫ ξ 2 ),


4.3 Membrane-cortex adhesion 81<br />

see Fig. 4.12, there is lateral correlation between links, and deformation satisfies<br />

a force balance that involves pressure, membrane bending and membrane<br />

tension must be considered. In this case, a macroscopic energetic approach<br />

that takes in account the bleb shape is justified for nucleation sizes larger<br />

than the mesh size ξ . The notions of link density and on and off rates are replaced<br />

by an effective adhesion energy density w which is the energetic cost<br />

per unit area of detaching membrane from the cytoskeleton 6 .<br />

(a) (b) (c)<br />

Vb<br />

u<br />

Sb<br />

Fig. 4.12. Nucleation of a bleb. (a) The detachment of a large enough area Sb can lead to the<br />

formation of a bleb (b) or an eventually resealing of the membrane.<br />

Assuming for now that the pressure and the membrane tension are constant,<br />

and taking into account that the curvature contribution to the energy is<br />

negligible compared to the tension contribution (κ/(γRb) ∼ 10 −8 ), the cost<br />

of energy of forming a bleb is simply<br />

∆E = −∆PVb + γ∆S + wSb, (4.41)<br />

where ∆S is the increase of membrane area when the bleb is formed. In the<br />

linear regime (u ≪ Rb), (see Fig. 4.12b) , the volume and the excess area<br />

are Vb ∼ S 2 b /R and ∆S ∼ S2 b /R2 respetivelly. Minimization of the energy, Eq.<br />

4.41, with respect to the radius of the bleb leads to Laplace law, Rb = 2γ/∆P,<br />

and the energy in terms of the bleb detached area becomes<br />

∆E = wSb − ∆P2<br />

4γ S2 b. (4.42)<br />

For small detached area, the energy increases linearly with the area, whereas<br />

for large bleb detached area, the quadratic term dominates and the energy<br />

decreases monotonously, Fig. 4.13. Consequently, there is an energy barrier<br />

∆E ∗ corresponding to the nucleation of a bleb of detached area S∗ b given by<br />

6 The adhesion energy w has to be thought of depending of the pressure and cell activity, as<br />

we have seen in the previous section. However here


82 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

∆E<br />

∼ wSb<br />

S ∗ b = 2 wγ<br />

,<br />

∆P2 (4.43)<br />

∆E ∗ = w2γ .<br />

∆P2 (4.44)<br />

∆E ∗ = w2 γ<br />

∆P 2<br />

S ∗ b =2 wγ<br />

∆P 2<br />

2 ∆P<br />

∼−<br />

4γ S2 b<br />

Fig. 4.13. Energy profile of a bleb in terms of detached area Sb. For small bleb detached area<br />

(Sb ≪ γw/∆P 2 ), the energy grows linearly with the adhesion energy. As the bleb surface becomes<br />

larger, the gain of energy driven by the pressure dominates over the adhesion energy and<br />

the energy decreases monotonously. The maximum of energy, ∆E ∗ ∼ w 2 γ/∆P 2 , corresponds<br />

to an energy barrier to nucleate a bleb and the detached area, S ∗ b = 2wγ/∆P2 , corresponds to<br />

the critical area above which a bleb will nucleate.<br />

Above the critical detached surface S∗ b , forming a bleb is energetically<br />

favorable, whereas for smaller surfaces a bleb will not form. Associated<br />

to the energetic barrier ∆E ∗ , there is a characteristic time for bleb nucleation<br />

τn. The waiting time for a fluctuation to overcome a barrier grows exponentially<br />

with the energy according to Kramers theory (Kramers, 1940),<br />

τn ∼ exp(∆E ∗ /kBT ), with a pre-factor that depends on the unbinding properties<br />

of the links ,<br />

τn = τ 0 <br />

w2γ n exp<br />

∆P2 1<br />

ET<br />

Sb<br />

<br />

, (4.45)<br />

where ET is an energy scale that can be related with local pressure, thermal,<br />

or active fluctuations. The nucleation is an indication of the likelihood of<br />

bleb formation. When the argument of the exponential in Eq. 4.45 is smaller<br />

than one, a bleb is likely to occur, whereas for larger values the waiting time


4.3 Membrane-cortex adhesion 83<br />

diverges and a bleb event occurs very unlikely. The nucleation of a bleb is<br />

in principle a very complicated process, and we do not aim to give an exact<br />

derivation, but an order of magnitude for the critical pressure,<br />

∆P ∗ <br />

w2γ ∼<br />

ET<br />

1/2<br />

. (4.46)<br />

According to Eq. 4.42, once the bleb is nucleated, it grows indefinitely.<br />

This is because we have considered a bleb as a small perturbation of the<br />

cell decoupling both cell and bleb energies and considering the former as a<br />

constant. Stabilization of the bleb can result from an increase of membrane<br />

tension or a decrease of internal pressure as the bleb grows.<br />

Mean number of blebs in a cell<br />

Using the concept of bleb nucleation time, we can write a qualitatively coarse<br />

grained description for the mean number of blebs in a cell. Blebs nucleate at<br />

an average rate given by the inverse of the nucleation time τn. At the same<br />

time, blebs have a life time, which includes a fast inflation and slower retraction<br />

τb, Fig. 4.14. The equation governing the mean number of blebs is given<br />

by<br />

˙<br />

Nb = Nt − Nb<br />

τn<br />

− Nb<br />

, (4.47)<br />

τb<br />

where Nt represents the maximum number of blebs in the cell, corresponding<br />

to the number of blebs that would occupy all the cell area.<br />

Fig. 4.14. Although blebs are highly dynamical, one can define an average number of blebs as<br />

a balance between bleb nucleation and bleb retraction.


84 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

At steady state, the mean number of blebs is constant and is given by<br />

Nb = τb<br />

Nt. (4.48)<br />

τn + τb<br />

In general, the life time of a bleb τb, depends on the volume of the mature<br />

bleb. We consider the life time of the bleb to be dominated by the retraction<br />

process which is controlled by the growth and contraction of a new actomyosin<br />

cortex. The contraction time is limited by myosin activity, and we<br />

consider it to be constant and of the order of minutes(Charras et al., 2008).<br />

As we have shown previously, the nucleation time τn depends exponentially<br />

on the energy cost to create a bleb, τn = τ0 n exp w2γ/(∆P 2ET ) . The presence<br />

of blebs should affect the mechanical state of the cell, both pressure<br />

and its tension, ∆P(Nb), γ(Nb). In the former case, a bleb reduces the drop<br />

of pressure when the cytosol and small vesicles flow through the cell cortex<br />

towards the bleb. Big organelles and vesicles remain in the interior of the cell<br />

an the osmotic pressure in the bleb decreases. In the case of the membrane<br />

tension, the increase of area of the cell, which is linear with the number of<br />

blebs, reduces membrane fluctuations and eventually stretching of the membrane,<br />

which causes an increase of the membrane tension (see section 4.2.1).<br />

Linearly we can express both effects as<br />

′<br />

∆P = ∆P0 1 − β Nb , (4.49)<br />

γ = γ0(1 + β ′′ Nb), (4.50)<br />

where ∆P0 and γ0 are the pressure and tension the cell would have in the<br />

absence of blebs .Both the drop of pressure and increase of membrane tension<br />

tend to reduce the probability of forming more blebs,<br />

τn ∼ τ 0 <br />

w2γ0 n exp<br />

∆P2 0<br />

1 + βNb<br />

ET<br />

<br />

, (4.51)<br />

where β = 2β ′ + β ′′ . Plugging this equation to Eq. 4.48, we obtain a selfconsistent<br />

expression for the number of blebs, or equivalently, the pressure or<br />

tension in the cell,<br />

Nb = Nt<br />

<br />

1 + τ0 <br />

n w2γ0 exp<br />

τb ∆P2 0<br />

1 + βNb<br />

ET<br />

−1<br />

. (4.52)<br />

This equation relates the number of blebs with the value of the pressure<br />

and tension that the cell would have in the absence of blebs, which can be<br />

seen as a control parameter that can be varied by external (such as osmotic<br />

shock) and internal (such as myosin activity) stress.


4.3 Membrane-cortex adhesion 85<br />

Below we arbitrary focus on the pressure variation and keep the membrane<br />

tension constant. Considering the variation of the membrane tension could<br />

be easily applied and leads to the same qualitatively results. We can solve<br />

Eq. 4.52 as a function of the normalized pressure in the absence of blebs,<br />

∆P0/ w2γ0/ET ) 1/2 , which we use as a control parameter, Fig. 4.15. For<br />

small values of the control parameter, blebs are not formed. The transition<br />

from a non-blebbing cell to a blebbing cell is given by Nb ∼ 1. After blebs<br />

start to form, the pressure remains nearly constant because blebs buffer the<br />

cell internal pressure, even if the pressure the cell would have in the absence<br />

of blebs is increased (see Fig. 4.15),<br />

∆Pmax ∼<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

w 2 γ0<br />

ET<br />

1<br />

log(τ0 1/2<br />

. (4.53)<br />

n /τb)<br />

(a) (b) (c)<br />

∆P/ w 2 γ0/kBT 1/2<br />

0.0<br />

0.0<br />

0.0 0.5 1.0 1.5 2.0 0.0 0.5 1.0 1.5 2.0<br />

∆P0/ w 2 γ0/kBT 1/2<br />

Fig. 4.15. Cell pressure and number of blebs as a function of the normalized pressure in the<br />

absence of blebs (b). Initially, no blebs are present in the cell and the cell pressure is sensitive<br />

to external perturbations, ∆P ∼ ∆P0. Above a pressure threshold (dotted line), blebs start to<br />

form and the pressure saturates to a constant value as long as new blebs can be formed. In<br />

(a), an Entamoeba histolytica with a single bleb corresponding to the region near the pressure<br />

threshold. In (c), M2 cells from (Charras et al., 2006), corresponding to the region of many<br />

blebs.<br />

Blebs can be seen in non healthy cells, possibly because of weak cytoskeleton<br />

adhesion, but can also serve physiological purposes such as motility<br />

(Charras and Paluch, 2008; Coudrier et al., 2005). In this case, blebbing<br />

produces a diffusion like motion (Coudrier et al., 2005).<br />

In this chapter, we have considered both kinetic and energetic properties of<br />

the membrane-cortex adhesion. Any visco-elastic property of the cortex has<br />

been neglected for simplicity. In the following chapter, we use a micropipette<br />

set up both experimentally and theoretically to study the reaction of a cell to<br />

a mechanical perturbation.<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Nb/Nt


86 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

4.4 Dynamical response of a cell to a controlled mechanical<br />

perturbation using a micropipette set up<br />

In section 4.3 we analyzed the stability of the membrane-cortex adhesion using<br />

both a simple kinetic model for the protein linkers and a coarse grained<br />

energetic description. We have deliberately neglected the viscoelastic properties<br />

of the cortex in order to simplify the analysis and focus only on the<br />

instability. In this section we study the reactivity of the cytoskeletal cortex<br />

to dynamical perturbations, which depends both on cortex viscoelasticity and<br />

actin polymerization.<br />

Generically, the cytoskeletal cortex behaves as a viscoelastic material,<br />

with an elastic response at short time , and a fluid response at longer time.<br />

Under micropipette aspiration or any other pressure perturbation such as an<br />

osmotic shock, pressure builds up a transient elastic stress on the cortex that<br />

eventually relaxes after a time τ ∗ (see Eq. 4.5). During the elastic deformation,<br />

proteins that link the membrane to the cortex are put under tension. The<br />

elastic stress built in the cortex may eventually reach the critical force per link<br />

at which the cortex-membrane adhesion becomes unstable (see section 4.3).<br />

The section is organized as follows: First, we consider the case of an instantaneously<br />

applied pressure in the micropipette. This case illustrates the<br />

effect of the elastic deformability of the cortex on the force experienced by the<br />

linkers. Second, we consider a pressure that is applied at a constant rate. The<br />

comparison between pressure loading rate and cortex relaxation rate fixes a<br />

maximum transient force experienced by the linkers, which strongly depends<br />

on the loading rate. As a consequence, there is not an absolute value of micropipette<br />

pressure at which the membrane-cortex adhesion become unstable<br />

but a continuos range of pressures depending on the time scale to build up<br />

pressure in the micropipette.<br />

4.4.1 Micropipette experiments as membrane tension and adhesion<br />

probes<br />

There are several techniques that may serve as a direct measure of membrane<br />

tension, membrane tether extraction, which consists of pulling a membrane<br />

tether out of the cell membrane using optical tweezers (Hochmuth et al.,<br />

1996), micropipette aspiration (Evans and Rawicz, 1990), and analysis of<br />

fluctuation spectrum (Pécréaux et al., 2004), to cite some examples.<br />

Micropipette aspiration is able to both perturb and measure the response<br />

to a perturbation, while using a membrane tether allows only to measure<br />

the membrane tension. For practical reasons, micropipette aspiration (Fig.


4.4 Dynamical response of a cell to a controlled mechanical perturbation using a micropipette set up 87<br />

4.16) is more convenient to measure the membrane tension area dependence<br />

and to be able to observe the cross-over between the entropic and elastic<br />

regime (Evans and Rawicz, 1990) (see Eq. 4.3), because the area drown in<br />

the micropipette is a sizable fraction of the total area of the cell (typical micropipette<br />

radii are of about 5-10 µm). We first review some of the results of<br />

micropipette aspiration experiments in vesicles and in cells.<br />

Vesicles<br />

Experiments on vesicles have been extensibly done for many years (Evans<br />

and Rawicz, 1990; Merkel et al., 2000; Rentsch and Keller, 2000; Rentsch<br />

and Keller, 2000; Sit et al., 1997). Initially the vesicle is at equilibrium and<br />

force balance (Laplace law) is satisfied,<br />

Pc − P0 = 2 γ<br />

, (4.54)<br />

R<br />

where Pc and P0 are the internal and external pressures respectively (see Fig.<br />

4.16). When a drop of pressure is applied locally inside the micropipette,<br />

the vesicle deforms, its radius of curvature increases, and eventually a new<br />

equilibrium is reached, Fig. 4.16,<br />

<br />

1<br />

∆P = 2γ<br />

Rp<br />

− 1<br />

<br />

, (4.55)<br />

R(L)<br />

where Rp and R are the micropipette and vesicle radii, L is the length of the<br />

tongue inside the micropipette, and ∆P ≡ P0 − Pp is the difference between<br />

the external pressure and the pressure inside the micropipette (see Fig. 4.16).<br />

Notice that equation (4.55) only has a stable solution if the membrane tension<br />

γ increases with the tongue length L (or area), since as the radius of the vesicle<br />

R decreases by volume conservation as the vesicle is being sucked, so does<br />

the force opposing force suction, which is the constant micropipette curvature<br />

(1/Rp) minus the vesicle curvature (1/R).<br />

The membrane tension does increase with the vesicle area (Eq. 4.3), and<br />

an equilibrium shape for a given tongue length can always be obtained until<br />

the radius of the cell reaches the radius of the pipette for which the vesicle is<br />

completely swallowed. Using equation (4.55), which determines the tension<br />

needed for the new equilibrium, and the relationship between the increase of<br />

tongue length and the areal strain α (Evans and Rawicz, 1990),<br />

α ∼ 1<br />

2<br />

Rp 2 <br />

3<br />

Rp ∆L<br />

−<br />

, (4.56)<br />

R R<br />

Rp


88 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

R0<br />

Pc<br />

R(L)<br />

P0<br />

Rp<br />

L<br />

Pp<br />

Fig. 4.16. Schematics of a micropipette experiment on a vesicle.<br />

we can measure experimentally the relationship between the tension and the<br />

areal strain (Eq. 4.3). Micropipette experiments have been used to show the<br />

cross-over between a regime where the tension is dominated by thermal fluctuations<br />

(entropic) and a regime where the tension is dominated by direct<br />

stretching of lipids (elastic) (Evans and Rawicz, 1990) (see Eq. 4.3).<br />

In vesicles, the new equilibrium is reached by the balance of the micropipette<br />

pressure and the area dependent tension (Eq. 4.55). In a cell, the<br />

cytoskeletal cortex also contributes to the force balance of the cell (see section<br />

4.2.2 and Eq. 4.18), and any elastic deformation contributes to balancing<br />

the pressure. The stationary force balance equivalent to the case of the vesicle,<br />

Eq. 4.55, is<br />

<br />

1<br />

∆P = 2(γ(L)+γm) −<br />

Rp<br />

1<br />

<br />

. (4.57)<br />

R(L)<br />

We are interested in the way the cell reaches this new equilibrium. There are<br />

two possible scenarios taking into account the cortex-membrane stability:<br />

(i) The membrane-cortex complex deforms continuously until the new equilibrium<br />

is reached, Fig. 4.17a→b.<br />

(ii) The membrane-cortex deforms until the critical force on the linkers is<br />

reached and the membrane unbinds from the cortex. At this stage, the<br />

bare membrane reaches a different equilibrium with a larger tongue length<br />

resulting either from the increase of membrane tension, Fig. 4.17a→c<br />

or from the formation of a new cortex which stops the membrane, Fig.<br />

4.17a→e. In the former case, after a characteristic polymerization time, a<br />

new cortex is finally formed under the bare membrane. In the presence of<br />

myosin, contractile tension is built in the cortex which presumably brings<br />

the membrane-cortex to the same final equilibrium than in the previous<br />

process, Fig. 4.17c→d and 4.17f→b.


4.4 Dynamical response of a cell to a controlled mechanical perturbation using a micropipette set up 89<br />

(c)<br />

(d)<br />

(a)<br />

(b)<br />

Fig. 4.17. In a micropipette experiment, the cell reaches the new equilibrium through three<br />

different scenarios. The initial state of the cell corresponds to a the membrane attached to the<br />

micropipette under moderate micropipette suction. When the suction increases, the cell deforms<br />

and a membrane-cortex tongue advances towards the interior of the micropipette. If the<br />

transient elastic tension in the cortex due to the deformation is smaller than the critical tension<br />

for membrane-cortex instability, the cell reaches the new equilibrium deforming continuously<br />

from (a) to (b). If the tension in the cortex exceeds a critical value, the membrane detaches<br />

from the cortex. A new transient equilibrium is reached either by increase of membrane tension<br />

(c) or polymerization of a new cortex (e). In the former case, a new cortex eventually<br />

repolymerizes underneath the bare membrane, and myosin driven contraction drives the cell<br />

to the final equilibrium, (c) to (d), which corresponds to the final configuration (b). In the second<br />

case, as the new cortex is polymerized, myosin contraction drives the cell to the final state<br />

(f) which coincides with the final state (b).<br />

These scenarios differ on the transient elastic tension built on the cortex,<br />

which results from the elongation of the tongue inside the micropipette, and<br />

the way the membrane is transiently stopped. In order to quantitatively describe<br />

these regimes we need to both consider the maximum elastic stress<br />

stored in the cortex during the aspiration and compare the force per link<br />

with the force threshold (Eq. 4.29) obtained in membrane-cortex adhesion<br />

analysis, section 4.3, and consider the conditions for which the membrane is<br />

stopped either by a new cortex or by the increase of membrane.<br />

In our analysis, we will consider that myosin driven tension in the cortex<br />

is constant 7 , and thus can be obtained by the initial equilibrium before<br />

aspiration, Fig 4.17a,<br />

L<br />

7 Keeping the myosin tension constant assumes that the thickness of the gel is roughly constant<br />

during the aspiration process, and second, myosin adapts fast to the new cytoskeletal<br />

Rp<br />

R0<br />

R<br />

(e)<br />

(f)


90 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

<br />

1<br />

∆P(0)=2(γm + γ(0)) −<br />

Rp<br />

1<br />

<br />

,<br />

R0<br />

(4.58)<br />

where ∆P(0) is the pressure of a reference state, where the membrane is<br />

attached to the micropipette.<br />

When the micropipette pressure ∆P is decreased, a tongue starts to advance<br />

at a velocity that is dictated by viscous dissipation. The dominant term<br />

in the dissipation (see Appendix 4.A) should results from membrane flow<br />

through links to the cortex, and leads to an effective viscosity that increases<br />

linearly with the length of the tongue, η(L) ∼ µbRp(Rp + L)/ξ 2 , where µb<br />

is a two dimensional viscosity for the bilayer of about 3 × 10−6 Pa m s<br />

(Hochmuth et al., 1982), and ξ is the mean distance between links which is<br />

roughly the mesh size of the cortex. The dynamical force balance including<br />

the dissipative force is given by<br />

µb<br />

Rpξ 2 (L + Rp)˙L<br />

<br />

1<br />

= ∆P − 2(γ(L)+γc + γm)<br />

Rp<br />

− 1<br />

<br />

. (4.59)<br />

R(L)<br />

The radius of the cell is related with the increase of tongue length assuming<br />

volume conservation, R(L) ∼ R0(1 − R2 p∆L/(4R3 0 )). For typical experimental<br />

conditions, R0 ∼ 20µm, Rp ∼ 5µm, and the increase of tongue<br />

length is a few Rp. As a consequence, the cell curvature can be considered as<br />

constant R ≈ R0. The membrane tension depends generically on the tongue<br />

length through the areal strain (see section 4.4.1), α ∼ 1/2∆L/Rp((Rp/R) 2 −<br />

(Rp/R) 3 ). For our experimental values the areal strain is of about 0.02∆L/Rp (Brugues<br />

et al., 2008b). We know from experiments (Evans and Rawicz, 1990) that<br />

the crossover from the entropic to the elastic regime in vesicles occurs for<br />

γ ∼ 5 × 10−4 N/m, which corresponds to α ∼ 2-5%, so for an increase of the<br />

membrane tongue of about two micropipette radii, we would start to stretch<br />

directly the lipids of the membrane in a vesicle and the tension would vary<br />

linearly with the areal strain (see Eq. 4.3 in section 4.2.1). In a cell, we do not<br />

know the amount of available membrane, and the dependence of the tension<br />

in the areal strain does not need to coincide with the case of a vesicle. However,<br />

we will assume linear response, with a stretching modulus ¯Ka which is<br />

expected to be smaller than the actual lipid stretching modulus,<br />

γ ∼ γ0 + ¯Ka<br />

Rp 2 <br />

3<br />

Rp ∆L<br />

−<br />

≡ γ0 + ζ∆L, (4.60)<br />

2 R R Rp<br />

area and its density is not affected by the elastic stress. These restrictions could be easily<br />

relaxed, but the main results remain qualitatively the same.


4.4 Dynamical response of a cell to a controlled mechanical perturbation using a micropipette set up 91<br />

where ζ is the increase of tension per unit length of about 450 N m−2 . For<br />

pressures of about 1000 Pa, the membrane tension balances the pressure for<br />

lengths between two and three times the radius of the micropipette (Merkel<br />

et al., 2000; Rentsch and Keller, 2000; Brugues et al., 2008b), and ζ should be<br />

of the order of 450 N m−2 , which roughly corresponds to direct stretching of<br />

the lipids. We consider the dissipation to be independent of the micropipette<br />

length, and write Eq. 4.59 as<br />

<br />

µb ˙<br />

1<br />

∆L = ∆Pp − 2(ζ∆L + γc)<br />

ξ 2<br />

Rp<br />

− 1<br />

<br />

, (4.61)<br />

R<br />

where now ∆Pp ≡ Pp − Pp(0) is the difference of pressure between the reference<br />

pressure and the final applied pressure, and we have used Eq. (4.58)<br />

for the initial force balance. Assuming for simplicity that the cortex do not<br />

propagate inside the cell, since that would involve shear stress, the strain rate<br />

in the micropipette is simply ˙L/L. The time evolution of the cortical tension<br />

is given by the maxwell like equation<br />

<br />

1 d<br />

+<br />

τ∗ dt<br />

<br />

γc = Eh ˙L<br />

. (4.62)<br />

L<br />

The equations 4.60 and 4.62 will be solved numerically below. The most important<br />

feature is that the stress reaches a maximum connected to the viscous<br />

relaxation at large time. We will give analytical approximations for this maximum<br />

in the elastic and viscous limits.<br />

4.4.2 Micropipette static suction pressure<br />

For an instantaneous applied pressure, the maximum stress occurs in the elastic<br />

regime and is relaxed after a time τ∗ . The solution of Eq. 4.62 in the elastic<br />

regime gives<br />

<br />

L<br />

γc ∼ Ehlog , (4.63)<br />

L0<br />

where L0 ≡ L(0) ∼ Rp. We have seen that for tongue lengths of about two<br />

or three times the radius of the micropipette, the membrane tension is large<br />

enough to balance micropipette pressures of about 1000 Pa, which is the typical<br />

order of magnitude of pressure needed to induce membrane-cortex detachment.<br />

In this approximate treatment, we linearize the expression of the<br />

cortical tension, γc ∼ Eh∆L/Rp, and the equation of motion for the increase<br />

of tongue Eq. 4.61, reads


92 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

<br />

˜η ∆L ˙ = ∆Pp − 2 ζ + Eh<br />

<br />

∆L<br />

,<br />

Rp ˜Rp<br />

(4.64)<br />

where 1/ ˜Rp ≡ (1/Rp − 1/R), and ˜η ≡ µb/ξ 2 . The solution for the increase<br />

of tongue length is<br />

∆L(t)=<br />

<br />

∆PRp ˜Rp<br />

1 − exp<br />

2(ζ Rp + Eh)<br />

− 2(Rpζ + Eh)<br />

˜ηRp ˜Rp<br />

t<br />

<br />

. (4.65)<br />

The maximum stress stored in the cortex is then given by γc(τ ∗ )=Ehlog(1+<br />

∆L(τ ∗ )/Rp). Although the pressure in the micropipette is set instantaneously,<br />

the flow of membrane and cortex is not instantaneous, but fixed by dissipation.<br />

As a consequence, the rate of increase of cortex tension, and its maximum<br />

value, are limited by this dissipation.<br />

4.4.3 Effect of the loading rate<br />

Experimentally it is never possible to set an instantaneous pressure. Instead,<br />

a characteristic time scale to build the final value is required. This time scale<br />

may affect dramatically the response of the cortex to the pressure, and consequently,<br />

the force on the links and the probability of membrane unbinding.<br />

Applying the same final pressure with two different rates may result in a<br />

membrane-cortex that is stable if the loading rate is small and unstable if the<br />

loading rate is larger.<br />

We model the pressure ramp as an exponential relaxation to the final value<br />

which interpolates a linear ramp plus a saturation towards the final pressure,<br />

<br />

∆Pp(t)=∆P∞ 1 − e −βt<br />

, (4.66)<br />

where β is the characteristic pressure rate. Depending on the value of β, we<br />

expect the cortex to behave as a solid or as a fluid. For βτ ∗ ≫ 1, the final<br />

pressure is reached instantaneously compared to the relaxation of the cortex,<br />

and we can use the results obtained above, Eq. 4.65. For βτ ∗ ≪ 1, the cortex<br />

reacts as a viscous fluid,<br />

γc ∼ τ ∗ Eh˙L/L, (4.67)<br />

and the equation of motion for the tongue length in the micropipette reads,<br />

<br />

˜η ∆L ˙ = ∆P∞ 1 − e −βt<br />

<br />

− 2 ζ∆L + τ ∗ Eh ˙L<br />

<br />

1<br />

. (4.68)<br />

L ˜Rp


4.4 Dynamical response of a cell to a controlled mechanical perturbation using a micropipette set up 93<br />

Linearizing the expression of the viscous tension γc, and defining an effective<br />

viscosity ηv ≡ ˜η + 2τ ∗ Eh/(Rp ˜Rp) ∼ 10 8 Pa s m −1 as the sum of the<br />

membrane and cortex viscosities, we can solve the increase of tongue length<br />

in the micropipette, Eq. 4.68<br />

∆L(t)= ˜Rp∆P∞<br />

2ζ<br />

1<br />

1 − β/β ζ<br />

<br />

1 − e −βt + β<br />

<br />

e<br />

βζ −β <br />

ζ t<br />

− 1<br />

<br />

, (4.69)<br />

where β ζ ≡ 2ζ /( ˜Rpηv) ∼ 10 s −1 . Since τ ∗ ∼ 10 s, and we are in the regime<br />

βτ ∗ ≪ 1, we have that β ≪ β ζ and the tongue length is roughly<br />

∆L(t) ∼ ˜Rp∆P∞<br />

2ζ<br />

<br />

βt + β<br />

<br />

e<br />

βζ −β <br />

ζ t<br />

− 1<br />

<br />

. (4.70)<br />

The maximum stress is obtained imposing ˙γc = 0 in Eq. 4.67. The time at<br />

which the stress reaches its maximum is given by,<br />

with c ≡ 1 +<br />

tmax = ˜Rpηv<br />

2ζτ∗ −c<br />

−c − ProductLog(−e ) , (4.71)<br />

4Rpζ 2<br />

β ˜R 2 pηv∆P∞ . The maximum stress is given by γc(tmax), Fig. 4.18.<br />

In order to obtain the detachment pressure on the micropipette as a function<br />

of the loading rate, we equate the maximal critical force per link ξ 2 (γm +<br />

γc(max))/Rp, to the value obtained in section 4.3, 16 pN. In Fig. 4.19, we<br />

plot the critical pressure as a function of the loading rate.<br />

If the critical force per link is reached, the membrane detaches from the<br />

cortex. After a transient membrane flow, the tongue eventually stops. Two<br />

possible mechanisms may contribute to stop the membrane tongue, increase<br />

of tension (see Fig. 4.17c) and repolymerization of new cortex (see Fig. 4.17e.<br />

In the following section we discuss both contributions.<br />

4.4.4 Can actin stop a moving membrane?<br />

There are mainly two mechanisms that may contribute to stopping the membrane<br />

tongue. First, as the tip grows the membrane tension increases with<br />

the total apparent membrane area in the pipette (see Eq. 4.60 and section<br />

4.2.1), and a force balance is reached that only involves membrane tension.<br />

Second, actin polymerization underneath the membrane is promoted by nucleators<br />

(Charras et al., 2006), and eventually, a new cytoskeletal cortex may<br />

be formed under the moving membrane, which stress contributes to balancing


94 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

γmax/γ ∞ max<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

γc (KPa µm)<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

0 10 20 30 40 50 60<br />

t (sec)<br />

0<br />

0 20 40 60 80 100<br />

βτ ∗<br />

Fig. 4.18. Maximum tension stored in the cortex normalized to the tension under infinite loading<br />

rate, as a function of the loading rate times the relaxation time, for a final pressure of 1 KPa<br />

and for τ∗ = 15 sec (blue), 50 sec (red) and 100 sec (green). The maximum tension saturates<br />

as a result of membrane dissipation, ∆L ˙ ∼ ∆Pf / ˜η. In the inset, the evolution of the tension in<br />

the cortex as a function of time for different loading rates, β = 1, 0.25, 0.1, 0.05 and 0.025<br />

sec−1 .<br />

P (KPa)<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

UNBINDING<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

β (s −1 )<br />

Fig. 4.19. Critical micropipette pressure for membrane unbinding as a function of the loading<br />

rate β for a typical degradation time of 30 seconds. The critical force per bond used is 16 pN,<br />

as obtained in section 4.3.


4.4 Dynamical response of a cell to a controlled mechanical perturbation using a micropipette set up 95<br />

the pressure. In what follows, we will consider the conditions for cortex repolymerization<br />

under a moving membrane, and we will compare the relative<br />

effect of re-polymerizing a new cortex and the increase of membrane tension<br />

in stopping a moving tongue.<br />

In section 4.2.2, we discussed the mechanisms that control the cortex<br />

thickness. A simple dynamical evolution for the thickness is given by Eq.<br />

4.16, ˙h = vp − vd(h). Its stationary solution (˙h = 0), results from the balance<br />

between polymerization and depolymerization velocities, which in general<br />

depend on stress and monomer concentration.<br />

Actin filaments nucleate at the membrane and grow towards the interior<br />

of the cell (Alberts et al., 2002; Charras et al., 2008). When they reach a<br />

characteristic size, which is roughly the mesh size of the gel ξ , filaments<br />

overlap and start crosslinking, leading to an elastic gel structure. In order to<br />

have a gel that is able to sustain and exert stress, the thickness must reach<br />

a critical value which is at least the crosslinking length ξ . This condition<br />

defines a threshold for the depolymerization velocity above which a stable<br />

layer of new cytoskeleton can not be formed: the polymerization velocity is<br />

equal to the depolymerization velocity at a thickness equal to the mesh size,<br />

vp ∼ vd(ξ ).<br />

In general, the balance between polymerization and depolymerization velocities<br />

depends on the stress stored in the cortex and the actin monomer<br />

concentration and diffusion. Since a cortex is always formed once the membrane<br />

has stopped (Merkel et al., 2000; Charras et al., 2008), and experiments<br />

using photobleaching recovery suggest a large diffusion constant for g-actin<br />

of D ∼ 10 2 µm 2 /s(Lanni and Ware, 1984) 8 , we will not consider any effect of<br />

the monomer concentration and difusion here.<br />

In our case, the gel grows on a membrane that is moving. Assuming the<br />

gel grows under no stress under the membrane, it becomes stressed as it ages<br />

away from the membrane, Fig. 4.20. The total amount of stress stored in the<br />

cortex depends on how fast the gel is recycled, or equivalently, the time during<br />

which a layer of gel is stretched before it is depolymerized, t ∼ h/vp. The<br />

depolymerization velocity vd increases exponentially with the stored stress<br />

(see Eq. 4.17), so a gel is expected to be thinner in a moving membrane.<br />

The effective stress in the gel is in general given by a viscoelastic maxwell<br />

like equation, Eq. 4.10, with a relaxation time scale τ ∗ which includes both<br />

the effect of cytoskeleton viscoelasticity and treadmilling (polymerization<br />

and depolymerization). Here we consider very thin gels, so the relaxation<br />

8 Which corresponds to mean displacement of 10µm after one second, much larger than the<br />

typical advancing velocity ˙L < µm/s


96 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

(a) (b)<br />

L<br />

˙L<br />

vm =0 vm = ˙ vp<br />

L<br />

l<br />

ξ<br />

vd<br />

l + δ<br />

Fig. 4.20. Schematics of a gel growing on a moving membrane. (a) The membrane tongue<br />

in the micropipette moves as a consequence of the suction pressure. (b) Actin filaments grow<br />

under the membrane and cross-link at a distance of about the mesh size of the gel. The resulting<br />

cortex is stretched due to the membrane displacement.<br />

time τ ∗ is dominated by treadmilling, in which case τ ∗ ∼ h/vp, and we<br />

expect it to be smaller than the loading time scale due to the membrane<br />

flow tη ∼ ηRp/∆P. As a consequence, the gel behaves as a viscous fluid<br />

(σ ∼ Eτ ∗ ∇ ˙u), with a viscoelastic stress at the outer layer of the gel,<br />

σ ∼ E h<br />

vp<br />

˙L<br />

L + σm(h). (4.72)<br />

Intuitively, the amount of deformation u stored at a distance h from the<br />

membrane is given by the increase of tongue length during a time equal to<br />

the relaxation time τ ∗ , u ∼ τ ∗ ˙L. Since we are considering very thin gels we<br />

can assume that h is smaller than λ, where λ is the characteristic binding<br />

length of myosin (see section 4.2.2), myosin tension is strongly limited by<br />

the treadmilling process and σm(h) ∼ σmsh/(2λ), where σms is the stress corresponding<br />

to the maximum density of myosin attached to the cortex (see<br />

section 4.2.2).<br />

The condition that allows the cortex to grow thicker than the typical mesh<br />

size ξ becomes,<br />

vp ≥ vd(ξ )=v 0 d exp<br />

<br />

E ξ<br />

σ0 L<br />

˙L<br />

vp<br />

+ σmsξ<br />

<br />

, (4.73)<br />

σ02λ<br />

where σ0 is a reference stress related to actin and cross-linker kinetics, and v 0 d<br />

is the stress free depolymerization velocity. From this last relation we obtain<br />

the condition the tongue velocity ˙L, and its length L, must satisfy in order for a<br />

new crosslinked gel to be able to polymerize under the membrane. Neglecting<br />

any variation of the pressure or the tension, while the tongue advances in the<br />

micropipette, the membrane velocity is given by η ˙L ∼ ∆P − γ/Rp 9 . Since<br />

9 There is no elastic term arising from the cytoskeleton because it is yet to be formed.


4.4 Dynamical response of a cell to a controlled mechanical perturbation using a micropipette set up 97<br />

the tongue velocity is bounded, the stress (∼ ˙L/L) will decrease in time as the<br />

tongue length increases. As a consequence, there is a tongue length scale L σ c<br />

above which the visco-elastic stress is small enough to allow the appearance<br />

of a new cytoskeletal cortex,<br />

∆L σ c = Rp∆P − γ<br />

ηRp<br />

ξ<br />

vp σ0 log vp/v0 d<br />

E<br />

. (4.74)<br />

− σmsξ /2λ<br />

In absence of membrane motion, myosin stress is in general an important<br />

contribution to regulate cortex thickness (see section 4.2.2). Typical gel thicknesses<br />

are of about 500 nm. Since we aim to find the conditions for gel generation<br />

(ξ ≪ 500 nm), the elastic stress resulting from membrane motion must<br />

be larger than myosin stress and we can typically neglect the myosin contribution<br />

contribution in Eq. 4.74.<br />

Once the condition for polymerizing a gel is fulfilled, the new cortex slows<br />

down the membrane, which reduces the tongue velocity and in turn, favors<br />

gel growth. As the thickness increase, myosin is recruited, which increases<br />

the stress in the cortex (see Eq. 4.72). After the initial slowing down of the<br />

membrane tongue due to visco-elastic stress, the cortex is able to finally stop<br />

the membrane if the tension created by myosin10 balances the difference of<br />

pressure, Rp∆P = γ + γm, which gives a lower bound for the gel thickness h,<br />

necessary to stop the membrane: γm(h) ≥ Rp∆P − γ,<br />

h ≥ λ<br />

<br />

1 + Rp∆Pp − γ<br />

+ ProductLog<br />

σms<br />

<br />

−exp<br />

<br />

−1 − Rp∆P − γ<br />

σms<br />

<br />

≡ hs,<br />

(4.75)<br />

where we have used Eq. 4.13 for the myosin tension. For small suction<br />

pressures, the gel thickness needed is very small h ≪ λ, and the critical<br />

thickness grows as the squared root of the micropipette pressure, hs ∼<br />

( λ<br />

σms Rp∆P − γ) 1/2 . For larger suction pressures, a thicker cortex is required,<br />

and the myosin kinetics is fast enough to reach the maximum stress almost in<br />

the entire gel thickness (h ≫ λ), and the critical thickness is linear with the<br />

pressure, hs ∼ (Rp∆P − γ)/σms.<br />

The critical length scale for cortex repolymerization must be compared to<br />

the length at which the membrane is stopped by the increase of membrane<br />

tension. If the length for a gel to grow is smaller, the gel is able to stop the<br />

membrane by itself, otherwise, the increase of membrane tension stops the<br />

membrane before a gel has grown.<br />

10 Note that if the membrane stops, the elastic stress induced on the cortex is only myosin<br />

stress after the relaxation time.


98 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

The evolution of the tongue in absence of cortex stress is η ˙L ∼ ∆P−(γ0 +<br />

ζ∆L)/Rp, where ζ is the rate at which the tension increases with the increase<br />

of tongue length L (see Eq. 4.60). This leads to the characteristic length L γ c,<br />

at which the membrane is stopped by the increase of membrane tension,<br />

The ratio of the two length scales is<br />

Lσ c<br />

L γ c<br />

∆L γ c = Rp∆P − γ0<br />

. (4.76)<br />

ζ<br />

∼ Eξζ<br />

log<br />

ηRpσ0vp<br />

v 0 d<br />

vp<br />

<br />

. (4.77)<br />

Taking the reference stress σ0 as of the same order of magnitude than the<br />

Young modulus, the ratio of lengths is of about 1 − 10, so in general, the<br />

gel is able to grow before the increase of membrane tension stops the membrane.<br />

However, both lengths can be of the same order of magnitude, and the<br />

different phenomena may not be distinguishable.<br />

4.5 Experimental observation of the different regimes<br />

We have theoretically considered the unbinding dynamics of the membranecortex<br />

complex focussing on micropipette experiments, although the results<br />

are of general applicability for any dynamical perturbation in the cell. As<br />

discussed in section 4.3, there is a critical force per bond at which the<br />

membrane-cortex attachments become unstable (Eq. 4.29 in section 4.3).<br />

This force was translated to a static critical suction pressure by the relation<br />

f ∗ =(∆P + 2γm/R)ξ 2 . Comparison with experiments (Merkel et al., 2000)<br />

allowed to determine the energy adhesion and a quantitative understanding<br />

of the discontinuous transition between stable and unstable attachments, as a<br />

function of linker concentration and myosin activity.<br />

Earlier in this chapter, we have seen that those concepts are not static<br />

and that the stability criteria depends on the dynamics of the perturbation<br />

(see Fig. 4.19). As an indication of this phenomenon, (Merkel et al., 2000)<br />

reports an apparent counter intuitive observation: after the initial instability<br />

in which the membrane unbinds from the cytoskeletal cortex, the growth of<br />

a new cortex is able to retract the tongue back to a position that was initially<br />

unstable, although the pressure in the micropipette is kept constant, Fig. 4.21.<br />

In the original work, (Merkel et al., 2000) propose as an explanation that the<br />

density of linkers is increased in order to sustain larger pressures.


4.5 Experimental observation of the different regimes 99<br />

4.5.1 Non equivalence of initial aspiration and retraction<br />

Our theoretical analysis propose an alternative explanation that does not involve<br />

any active response of the cell, but relies on the dynamical origin of the<br />

perturbation and viscoelastic behavior of the cortex, Fig. 4.21.<br />

Initially, the micropipette aspirates the cell at a certain rate (supposedly<br />

instantaneous), and the cortex is elastically deformed. As already explained,<br />

the membrane detaches as a consequence of the force applied on the links,<br />

which is the sum of the elastic deformation due to the suction process and the<br />

myosin tension. Consequently, the final pressure and the loading rate must<br />

lie on the unstable region described in Fig. 4.19, which we represent as the<br />

process (i) circle in Fig. 4.21b. When the membrane reaches its new equilibrium<br />

length (by increase of membrane tension or by cortex repolymerization),<br />

a new cortex grow under the membrane, which after recruiting myosin,<br />

it retracts the tongue towards the cell. In this case, the loading rate at which<br />

the force is applied is given by the actin sliding velocity driven by myosin.<br />

Molecular motor velocities are in the range of several nanometers per second<br />

(see chapter 3), which is at least one order of magnitude smaller than the initial<br />

velocity of the tongue dictated by viscous dissipation. As a consequence,<br />

during the retraction process, the contribution from the elastic deformation in<br />

the cortex is smaller than in the aspiration case (see Fig. 4.18). In terms of the<br />

phase diagram (Fig. 4.19), the retraction process corresponds to the same final<br />

pressure but to a smaller loading rate. Eventually, the rate of retraction (or<br />

null rate) could correspond to a stable attachment (process (ii) of Fig. 4.21b),<br />

in which case, once the new equilibrium is reached (the retracted tongue), the<br />

membrane would not detach again.<br />

In contrast, cellular blebbing is mainly caused by myosin contraction,<br />

which implies that the loading rate is the same than the retraction rate. In<br />

other words, the myosin tension alone must overcome the critical value for<br />

membrane unbinding, which corresponds to the criteria derived in section 4.3.<br />

The periodicity arises from the fact that during both detachment and retraction,<br />

the force per link remains the same, contrarily to the micropipette case,<br />

where the initial suction process induces an extra transient force in the links,<br />

which is not present during retraction.<br />

From the previous analysis, we expect other possible regimes, which summarizing<br />

include: saltatory increase of the tongue during the suction process<br />

(Rentsch and Keller, 2000), corresponding to several repolymerization<br />

of cortical layers which become unstable; oscillatory behavior near the equilibrium<br />

tongue length, corresponding to a cortical layer reaching the unbinding<br />

transition while the tongue is retracting; and a regime where no cortex


100 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

(a)<br />

L<br />

P<br />

γm<br />

Cytoskeleton<br />

Saturation from contraction<br />

membrane tension<br />

Unbinding<br />

t<br />

t<br />

P (KPa)<br />

(b)<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

(ii)<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

β (s −1 )<br />

Fig. 4.21. Non equivalence between the suction process and the cortex retraction. (a) Evolution<br />

of the membrane tongue, micropipette pressure and myosin tension as a function of<br />

time. Initially, the membrane unbinds from the cortex and the tongue grows until it reaches a<br />

saturation value due to increase of membrane tension or cortex repolymerization. During the<br />

retraction process, the membrane may not detach from the cortex although the pressure in the<br />

micropipette is the same than during the extraction. (b) Same phenomenon explained using the<br />

phase diagram for critical pressure (see Fig. 4.19). Initially (i), the pressure is applied at a high<br />

rate, and this triggers membrane unbinding (region in red). After saturation, the final pressure<br />

is kept constant but the loading rate is dictated by myosin activity, and is much smaller, so the<br />

new point in the phase diagram corresponds to a stable attachment (ii).<br />

can be formed, which corresponds to the case where the membrane velocity<br />

is above the polymerization threshold and the cell is finally engulfed inside<br />

the micropipette.<br />

To this end, we have collaborated with Benoit Mauguis and François Amblart<br />

(Brugues et al., 2008b), in order to conduct several experiments on the<br />

Entamoeba histolytica system (Coudrier et al., 2005), to experimentally observe<br />

several of the possible predicted regimes.<br />

4.5.2 Micropipette suction regimes<br />

We can distinguish three phases during a micropipette experiment, which<br />

are the initial growth of the tongue, saturation, and retraction processes (see<br />

Fig 4.21).<br />

The first phase may occur while the membrane is still bound to the cytoskeleton,<br />

or after its unbinding.As already discussed the second phase may<br />

be due to an increase of membrane tension or to the contractile activity of the<br />

cytoskeleton. Since we have already discussed the role of membrane tension<br />

(i)


4.5 Experimental observation of the different regimes 101<br />

earlier in this chapter, we will neglect the effect of the increase in membrane<br />

tension in the following discussion.<br />

Saltatory and no cortex regime<br />

We have seen in section 4.4.4, that a new cortex can repolymerize under a<br />

moving membrane above a threshold tongue length, which is determined by<br />

the suction pressure, or equivalently, the velocity at which the tongue moves<br />

(Eq. 4.74). If the condition is fulfilled, a new cortex is polymerized and a<br />

force opposing the motion slows down the membrane. For a critical thickness<br />

hs (Eq. 4.75), the cortex is able to stop the membrane, in which case we reach<br />

the phase two of retraction. However, we have seen that myosin contraction<br />

can lead to membrane unbinding 11 , if the critical force per link is surpassed,<br />

or equivalently, if a threshold cortex thickness is reached,<br />

hb Rp<br />

σmsξ 2 f ∗ , (4.78)<br />

where we have used the limit where the myosin kinetics is faster than the cortex<br />

recycling (h > λ, see Eq. 4.13). If the unbinding thickness hb is smaller<br />

than the stopping thickness, the new cortex is able to slow down the membrane<br />

transiently until the detachment increases the tongue velocity again,<br />

leading to a saltatory like movement (Fig. 4.25). This regime is experimentally<br />

observed in Fig. 4.22a→e. The theoretical prediction is depicted in Fig.<br />

4.22e, and is based in the force balance between the pipette pressure and the<br />

myosin tension, Eq. 4.61, and the stress dependent dynamical evolution of<br />

the gel thickness, Eq. 4.16.<br />

Interestingly, the micropipette suction force increases as the tongue length<br />

increases, because this corresponds to a decreasing cell radius, and a correspondingly<br />

increasing Laplace pressure that drives the cell towards the micropipette.<br />

If the increase of membrane tension is not sufficient to stop the<br />

membrane, the amoeba will be irremediably swallowed by the micropipette,<br />

with an increasing membrane velocity as more amoeba is sucked. The maximum<br />

velocity is roughly ∆ ˜P/ ˜η with ∆ ˜P is the Laplace pressure 12 , in which<br />

case, the threshold length for cortex polymerization is (Eq. 4.74)<br />

∆L σ c ∼ ∆ ˜Pξ<br />

˜ηvp<br />

E<br />

σ0 log(vp/v 0 d ).<br />

(4.79)<br />

11 Decreasing the micropipette radius increases the force on the bonds for the same cortex<br />

stress.<br />

<br />

12 ∆ ˜P 1Rp<br />

1<br />

≡ ∆P − γ − .<br />

Rc(L)


102 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

(a)<br />

(c)<br />

(f)<br />

10µm<br />

10µm<br />

(b)<br />

(d)<br />

10µm<br />

10µm<br />

(e)<br />

∆L(µm)<br />

30<br />

25<br />

20<br />

15<br />

10<br />

(a)<br />

(c)<br />

(b)<br />

(d)<br />

5<br />

0 5 10<br />

t (sec)<br />

15 20<br />

5µm 5µm 5µm<br />

Fig. 4.22. Saltatory regime (a-e), and transition to no cortex regime (f). In (a-d) sequences<br />

of cortex repolymerization (a) and (c), which stops the advancing tongue, followed by detachment<br />

processes (b) and (d) with an abrupt acceleration of the tongue motion. We define<br />

this regime as saltatory since although the membrane nearly stops, it does not retract. In (e) a<br />

qualitative theoretical prediction of the saltatory process in which the critical cortex thickness<br />

needed to stop the tongue hs corresponds to the critical thickness for bond unbinding hb, using<br />

the parameters, vp = 50 nm s −1 , hc = 200 nm and ∆P = 400 Pa . (f) In the same experiment,<br />

we observe a transition from the saltatory regime to the no cortex regime in which several attempts<br />

of repolymerizing a gel (white arrows) are frustrated and the tongue advances until the<br />

whole amoeba is engulfed into the micropipette. In these series of snapshots, the micropipette<br />

has been moved<br />

If the threshold length is similar to the length of the amoeba inside the micropipette<br />

(∼ 40µm, ∆P ∼ 1000 Pa) for the maximal velocity, we do not<br />

expect to observe cortex repolymerization. In particular, a transition from a<br />

saltatory regime (cortex polymerization) to a no cortex regime could be observable<br />

since the threshold length increases as the amoeba is swallowed.<br />

In Fig. 4.22f, we report, during the same suction experiment as in the Fig.<br />

4.22a→d, a transition to a non cortex regime, in which several partial cortex<br />

repolymerization are observed (white arrows), before the amoeba was finally<br />

swallowed.<br />

Another aspiration regime is observed when the cortex is able to repolymerize,<br />

with a thickness below the unbinding transition, but unable to stop<br />

the membrane. In this case, one should observe an initial slope of the membrane<br />

trajectory, corresponding to the motion of the free membrane, followed<br />

by a change in the slope due to the opposing force from the new cortex layer<br />

(Fig. 4.25). Experimentally, this case is difficult to observe unless filamentous


4.5 Experimental observation of the different regimes 103<br />

actin is fluorecently labeled, since it is not possible to distinguish by confocal<br />

microscopy alone if the membrane decreases its velocity by increase of<br />

membrane tension or due to the formation of a new cortex.<br />

Cell oscillations in micropipette<br />

Cortex unbinding may also occur after retraction has been initiated (hs < hb <<br />

h), in which case, we expect to observe an oscillatory regime, similar to the<br />

saltatory regime, which approaches the saturation length by a series of retractions<br />

and expansions (Fig. 4.25). This regime is experimentally observed, in<br />

Fig. 4.23 we show two snapshots of the oscillatory behavior with the corresponding<br />

tracking. The theoretical prediction is depicted and plotted with the<br />

experimental measurements, showing a qualitatively good agreement. Notice<br />

that this regime corresponds with the aspiration phase, since the mean position<br />

of the tongue increases and eventually saturates (Fig. 4.23).<br />

38<br />

(a) (b)<br />

10µm<br />

10µm<br />

L(µm)<br />

36<br />

34<br />

32<br />

30<br />

28<br />

26<br />

0 20 40 60 80 100 120 140 160<br />

t (sec)<br />

Fig. 4.23. Tracking of the amoeba oscillations inside a micropipette of 10 µm using a suction<br />

pressure of about 400 Pa. (a) two snap shots corresponding to a maximum and a minimum of<br />

a period (top and bottom respectively). (b) plot of the tracked oscillations (tracking of images<br />

using ImageJ (Rasband, 2006)) and theoretical prediction (solid line), with the mean length<br />

evolution (dotted line). The parameters used are ∆P ∼ 400 Pa, vp ∼ 50nm s −1 , hc ∼ 600 nm.<br />

The period of the measured oscillations is well defined, Fig. 4.24, and is<br />

given by the time needed for the gel to reach the critical thickness at which<br />

the threshold force is reached and the gel detaches. It is roughly given by<br />

τ ∼ hb/vp, or<br />

τ ∼ Rp<br />

vp<br />

f ∗ b . (4.80)<br />

σmsξ 2


104 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

Remarkably, for a given cell, the critical unbinding force, the myosin density,<br />

cross-linkers density and polymerization velocity are fixed, so the period of<br />

oscillation depends only on the radius of the micropipette, but not on the micropipette<br />

pressure. Using the value for σms and the critical force obtained in<br />

section 4.3 (σms = γm/h ∼ 10 4 Pa, f ∗ ∼ 16 pN), and typical values for polymerization<br />

velocity vp ∼ 50nm s −1 and crosslinker distance ξ ∼ 100 nm, we<br />

obtain an oscillation period of about 15 seconds for the micropipette radius<br />

used in the experiments (Rp ∼ 5µm), which is in good agreement with the<br />

period measured of about 13 ± 9 seconds (Fig. 4.24).<br />

Number<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 5 10 15 20 25 30 35<br />

Period (s)<br />

Fig. 4.24. Histogram of the oscillation period obtained for 4 different amoeba with a pipette<br />

radius Rp =5µm. The average period τ = 13 sec would correspond to a gel thickness of about<br />

250µm, which is within the range of typical cortex thickness.<br />

More exhaustive experiments varying the radius of the micropipette or<br />

using drugs affecting polymerization velocity or crosslinker density would<br />

allow to quantitatively test our prediction for the oscillation period, specially<br />

the fact that it does not depend on the micropipette pressure. An indication<br />

of this phenomena is given by the sharp period measured in different cells in<br />

which the micopipette pressure was not the same, but varied in order to reproduce<br />

the oscillations (Fig. 4.24). Unfortunately, changing the radius of the<br />

micropipette Rp is difficult, because small Rp requires large suction pressure<br />

in order to see the effects, and large Rp allows the cell to crawl by itself inside<br />

the micropipette, even without applying any pressure.<br />

Finally, during the retraction phase, two regimes can be observed. On one<br />

hand, the stable retraction of the tongue occurs when the final thickness of the<br />

gel is below the critical unbinding thickness. In this case, the tongue reaches<br />

an equilibrium length given by the force balance Rp∆P − γ(L)=γm. On the


4.6 Conclusions 105<br />

other hand, the oscillations reported in the aspiration phase will continue once<br />

the saturation phase has been reached since the final thickness will still be<br />

larger than the critical thickness, but with a vanishing mean displacement.<br />

hb<br />

h<br />

Retraction<br />

Oscillatory<br />

hb >hs<br />

Saltatory<br />

h<br />

Steady<br />

flow<br />

hs >hb<br />

Fig. 4.25. Phase diagram of the different cortex regimes depending on the thickness at which<br />

the links remain stable hb and the thickness needed to stop the membrane hs. The diagonal<br />

divides the space in a region where hb > hs, upper region, and a region where hs > hb, lower<br />

region. In this diagram, we choose an arbitrary value of h. Depending on the value of hs and<br />

hb we observe four different regimes. The retraction regime, corresponding to hb > h > hs.<br />

The oscillatory regime, corresponding to h > hb > hs. The saltatory regime, which differs<br />

from the oscillatory regime in the fact that the former do not stop the membrane at any time,<br />

hs > h > hb. The steady flow regime, where h < (hs,hb), and the tongue is stopped by the<br />

increase of membrane tension.<br />

In Fig. 4.25, we show a phase diagram for all possible regimes observed<br />

in a micropipette experiment.<br />

4.6 Conclusions<br />

We have exhaustively considered the phenomenon of membrane-cortex adhesion<br />

and its stability under pressure perturbations. We have followed two<br />

main steps:<br />

First, we have used two different approaches based on both a model for<br />

the membrane-cortex linkers, which takes properly into account the kinetics<br />

of attachment and detachment, and an energetic description of the detachment<br />

process, to describe the adhesion between the membrane and cortex.<br />

We have shown that within both approaches, there is a global link failure and<br />

an abrupt transition from a stable adhesive membrane to an unstable membrane<br />

detached from the cytoskeleton for a critical force on the links.<br />

hs


106 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

The strength of the adhesion depends strongly on both myosin concentration,<br />

which sets a force on the links that increase the likelihood of membrane<br />

detachment, and link concentration, which determines the amount of force<br />

per link. Our results show that the adhesion energy is not an absolute quantity<br />

experimentally accessible, but depends on the active state of the cell, which<br />

can account for differences in the measured adhesion energy of up to 250%.<br />

We give a comprehensive explanation of previous works found in the literature<br />

(Merkel et al., 2000), where myosin and link perturbations lead to a priori<br />

surprising results concerning the values of the critical unbinding pressure and<br />

adhesion energy.<br />

We have described the mean number of blebs in a cell using a coarse<br />

grained kinetic model. We are able to distinguish two qualitative different<br />

regimes: a diffusive state where the cell uses blebbing to move, characterized<br />

by a mean number of blebs close to one, and a state where the cell blebs<br />

extensively, in which case blebs serve as a pressure buffer keeping the internal<br />

pressure constant.<br />

Second, we have considered the dynamical response of a cell to mechanical<br />

perturbations, and in particular, how the viscoelasticity and the treadmilling<br />

(polymerization and depolymerization) of the cortex affects the transmission<br />

of force to the links and the membrane-cortex instability. Based on<br />

experimental evidence, we have used the simplest Maxwell like model that<br />

describes the short time elastic behavior and long time viscous behavior of<br />

the cytoskeletal cortex, and takes explicitly into account both the treadmilling<br />

properties of the cortex and the active stresses generated by myosin concentration<br />

in the cortex. We found that the unbinding instability is highly sensitive<br />

to the rate at which a pressure perturbation is applied in comparison with<br />

the relaxation time scale of the cytoskeletal cortex.<br />

Focussing on micropipette experiments, which allow to both measure mechanical<br />

properties of the cell and control the suction pressure, we have identified<br />

several possible regimes during a micropipette extraction: A stable deformation<br />

where cortex and membrane deform while remaining adhered. A<br />

deformation which involves an initial detachment followed by a repolymerization<br />

of new cortex with stable adhesion. In this last regime, the cortex may<br />

eventually become unstable as the elastic stress in the cortex increases due<br />

to myosin activity, leading to either a saltatory or oscillatory regime. If the<br />

adhesion of the new cortex remains stable, the tongue inside the micropipette<br />

retracts towards the cell. In order to understand both the effect of the repolymerization<br />

of a new cortex and the increase of membrane tension in the process<br />

of stopping the detached membrane, we have modeled the growth of the


4.6 Conclusions 107<br />

cortex thickness under a moving membrane. We found that there is a critical<br />

membrane velocity above which a stable gel can not be formed.<br />

We have tested our theoretical predictions on experiments found in the<br />

literature (Merkel et al., 2000), and on experiments we have conducted on the<br />

Entamoeba histolytica. We have experimentally identified several predicted<br />

regimes and we have found good agreement with the theoretical prediction.<br />

Our results represent a full characterization of a mechanical perturbation<br />

on the membrane cortex interface and provides a better understanding of the<br />

adhesion between membrane and cortex, ruling out the concept of an absolute<br />

value for the adhesion energy during experimental measurements.<br />

Aknowledgements<br />

The work in this chapter has been done in close collaboration with B. Mauguis<br />

and F. Amblart, in particular, the experimental measurements conducted<br />

in the Entamoeba histolytica system.


108 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

4.A Membrane flow dissipation<br />

In this appendix, we will consider the energy dissipation of a membrane that<br />

flows away from the cortex. The membrane velocity depends on this dissipation<br />

and it is involved in several sections of the present chapter, and it is<br />

particularly important in determining the loading rate on the cortex which<br />

determines the elastic stresses stored in the cortex, and consequently, its response<br />

to external perturbations such as osmotic shocks and micropipette experiments.<br />

The appendix is organized in two parts that correspond to the two<br />

relevant geometries that we encounter in our analysis: the dissipation during<br />

the initial membrane-cortex detachment, which can be extrapolated to the<br />

dissipation during bleb inflation, and the dissipation during membrane flow<br />

inside a micropipette (see Fig 4.27).<br />

Membrane motion during bleb inflation or inside a micropipette involve<br />

flow of both cytosol and membrane. The increase of volume in both blebs and<br />

tongues in micropipette aspiration, needs a continuous income of membrane<br />

and cytosol from the rest of the cell. In general, the energy dissipation of a<br />

liquid flow depends on its velocity gradients and viscosity,<br />

<br />

˙E = µ dV (∇ · v) 2 , (4.81)<br />

where µ is a viscosity, v the velocity of the flow, and the integral runs over<br />

the liquid volume. In the case of interest, we need to add the contributions<br />

from both the membrane and cytosol.<br />

Membrane dissipation<br />

The membrane behaves as a two dimensional incompressible fluid. As a<br />

consequence, the velocity profile through the neck defined by the area of<br />

membrane-cortex detachment, is given by lipid conservation within the membrane<br />

area,<br />

2πr˙r = ˙A, (4.82)<br />

where r is the distance from the the center of the detached area (see Fig.<br />

4.26a), and A is the area of the bleb or micropipette tongue. Since the membrane<br />

coming from the rest of the cell is attached to the cortex, it must flow<br />

through protein linkers, where the velocity must vanish, Fig. 4.26b. As a result,<br />

there is a velocity gradient between the protein linkers of about ˙r/ξ ,<br />

being ξ the mean distance between links. The total dissipation due to membrane<br />

motion is thus given by the sum of a term that comes from the global


4.A Membrane flow dissipation 109<br />

gradient of the flow (∼ ˙A/r 2 ) and a term from the local gradient of the flow<br />

at the protein linkers (∼ ˙A/(ξ r)),<br />

<br />

1 Acell<br />

˙Em ∼ µb log +<br />

ξ 2 S<br />

1<br />

<br />

˙A<br />

S<br />

2 , (4.83)<br />

where S is the area of the detached patch, µb ∼ 3 × 10 −6 Pa s m (Hochmuth<br />

et al., 1982) is the two dimensional viscosity of the membrane (µb ≡ eηb,<br />

where e and ηb are respectively the thickness of the bilayer and the viscosity<br />

of the membrane). The first term in the membrane dissipation Eq. (4.83)<br />

corresponds to the local dissipation at the binders and the second term corresponds<br />

to the global dissipation. The former dominates for patches that are<br />

larger than the mean distance between linkers (S > ξ 2 ), and it is always the<br />

case for blebs (S ∼ µm 2 ) and micropipette experiments (Rp ∼ µm).<br />

(a) (b)<br />

˙r<br />

Fig. 4.26. Sketch of the flow of lipid membrane from the cell body to the patch of detach<br />

area (dotted line). In (a), the velocity profile depicted in lines. (b) The flow must vanish at the<br />

protein linkers, and as a result, there is a gradient of velocity between protein linkers.<br />

Cytosol dissipation<br />

Assuming the cytosol behaves as an incompressible fluid, the velocity profile<br />

is given by mass conservation, ˙r ∼ ˙V /r 2 , where V is the volume of the bleb<br />

or tongue in the micropipette. If the cortex remains stable under the patch of<br />

detached membrane, which is the case for the initial detachment, the cytosol<br />

must flow through the network of actin filaments that form the cortex, which<br />

induces velocity gradients of about ∼ ˙r/ξ (see Fig. 4.27), as in the case of the<br />

membrane flowing through linkers, assuming that the mean distance between<br />

linkers and the cortex mesh size are roughly the same. The total dissipation<br />

contains a global gradient of the flow (∼ ˙V /r 3 ) and a global contribution<br />

(∼ ˙V /(r 2 ξ )), restricted to the thickness of the cortex, h,<br />

r<br />

ξ


110 4 Membrane-cortex interactions: Micropipette experiments and blebbing cells<br />

<br />

h 1<br />

˙Ec ∼ ηc +<br />

Sξ 2 S3/2 <br />

˙V 2 , (4.84)<br />

where ηc ∼ 3 × 10 −3 -2 × 10 −1 Pa s is the cytosol viscosity (Charras et al.,<br />

2008). As in the case of the membrane, the dissipation resulting from the<br />

cytosol flow through the cortex dominates.<br />

4.A.1 Dissipation during membrane-cortex detachment<br />

The dissipation from the membrane deformation that follows the detachment<br />

from the cortex, see Fig. 4.27, has a contribution from the membrane flow<br />

through linkers and a contribution from cytosol flow through the cortex (Eq.<br />

4.83 and 4.84),<br />

˙E ∼ µb<br />

ξ 2 ˙A 2 + ηch<br />

Sξ 2 ˙V 2 . (4.85)<br />

For small deformation of the membrane, ˙V ∼ S ˙x and ˙A ∼ x ˙x, where x is the<br />

normal coordinate to the cortex. For small enough distances (x ∼ 10 nm),<br />

the relevant contribution to the dissipation is the flow of cytosol through the<br />

cortex. For larger distances between the cortex and the membrane, the dissipation<br />

is mainly due to the membrane drag through the linkers. For initial<br />

detachment, which is relevant in the kinetic stability analysis in section 4.3,<br />

the dissipation is then<br />

and the effective viscosity η = Sηch/ξ 2 .<br />

˙E ∼ Sηch<br />

ξ 2 ˙x 2 , (4.86)<br />

4.A.2 Dissipation during membrane flow inside a micropipette<br />

During micropipette aspiration experiments the cortex and membrane may<br />

form a tongue inside the pipette that advances at a certain velocity, Fig. 4.27<br />

(see section 4.4). In this geometry, the main contributions to the dissipation<br />

are the membrane flow through the binders, and the cytosol flow towards the<br />

micropipette (in this case there is no flow through the cortex as it is moving<br />

with the membrane). The former main contribution comes from the cancelation<br />

of the membrane velocity at each binder of the cytoskeletal network, and<br />

it is related with the rate of area increase in the micropipette, Eq. 4.83. For the<br />

micropipette geometry, ˙A = Rp ˙L, where L is the tongue of membrane inside


(a) (b)<br />

4.A Membrane flow dissipation 111<br />

Fig. 4.27. Schematics of the membrane (green) and cytosol (black) flows in the initial detachment<br />

(a) and micropipette experiment (b) geometry. During the initial detachment (a),<br />

membrane must be dragged from the cell (green arrows) and cytosol must flow (black arrows)<br />

through the cortex (in red). In the micropipette experiment (b), membrane is dragged from the<br />

cell and cytosol flows towards the micropipette. In this case, the cortex is deformed from the<br />

micropipette-cell contact, and follows the motion of the membrane.<br />

the micropipette, and the membrane flow in the cell ˙r at a distance r from<br />

the micropipette contact is ˙r ∼ Rp ˙L/r. At the interior of the micropipette, the<br />

cortex is deformed as the membrane moves and the gradient of velocity at<br />

the cortex is ˙L/L. The incompressibility of the membrane leads to a constant<br />

velocity inside the micropipette ˙L. The relative velocity between the membrane<br />

and cortex which is the responsible of the viscous dissipation inside<br />

the micropipette is then ˙L − z˙L/L, where z is the coordinate along the micropipette.<br />

The resulting viscous dissipation is then the sum of the membrane<br />

drag from the cell and inside the micropipette and the flow of cytosol towards<br />

the micropipette (second term in Eq. (4.84)),<br />

˙E ∼ Rpµb<br />

ξ 2<br />

<br />

Rp log<br />

Rcell<br />

Rp<br />

<br />

+ L<br />

<br />

˙L<br />

3<br />

2 + Rpηc ˙L 2 . (4.87)<br />

The cytosol term is typically four orders of magnitude smaller than the contribution<br />

from the membrane flow, and the effective viscosity η(L) is then<br />

η(L) ∼ µbRp<br />

ξ 2 (Rp + L). (4.88)


5<br />

General conclusions<br />

In this work we have studied three problems from soft matter to physical<br />

biology.<br />

In the first part, we have studied defect dynamics in Langmuir monolayers.<br />

An intrinsic asymmetry between velocities of positive and negative<br />

defects is observed during the annihilation process. We have shown that this<br />

asymmetry results from the different topology of the negative and positive<br />

defects, and the asymmetry of the elastic constants of the material. Using a<br />

simple quantitative model based on liquid crystals, we have been able to relate<br />

dynamical measurements (in this case, the defect velocity), to thermodynamic<br />

properties of the material (in this case, the elastic constants), as a function of<br />

the surface pressure. This procedure opens new ways to measure material<br />

properties which are otherwise very difficult to obtain. A possible continuation<br />

of the present work is to consider similar analysis of the wave generation<br />

phenomena induced by light. As explained in the chapter, these materials are<br />

photosensitive, and irradiation using the appropriate wave length induces a<br />

transition between a polar to a non polar molecule. This effect is interesting<br />

because it generates dynamical wave patterns (Crusats et al., 2005; Burriel<br />

et al., 2006) that depend on the topological charge and their separation. A<br />

similar approach as the used in this chapter could lead to new strategies to<br />

measure the elastic properties of the material using irradiation of light. More<br />

interestingly, the conformational change induced by light, resembles the conformational<br />

change used by molecular motors to produce work (Browne and<br />

Feringa, 2006). In the spirit of biomimetic experiments, the study of artificial<br />

molecular motors build from azobenzene molecules would be of special<br />

interest in understanding cooperative phenomena or self organization of real<br />

molecular motors in a system which is can be both much more controlled and<br />

tuned.


114 5 General conclusions<br />

In the second part, we have considered the collective behavior of molecular<br />

motors for several weak interactions, namely, hard-core repulsion, weak<br />

attractive interaction and long-range repulsive interaction. We have shown<br />

that their collective performance is generally optimized with respect to the<br />

corresponding meanfield scenario, which underestimates the cooperativity of<br />

molecular motors in terms of velocity and efficiency. The deviation from a<br />

mean field picture can be understood within a two state model. We obtain a<br />

nearly uniform distribution of forces along the motor cluster. Force distribution<br />

along the cluster is important since dynamics of motor attachment and<br />

detachment depend critically on the applied force. In this chapter, we only<br />

wanted to elucidate some generic cooperativity phenomena, so we neglected<br />

the finite processivity and force dependent kinetics of molecular motors. A<br />

more accurate model accounting for real molecular motors should be considered,<br />

able to deal with dimeric motors. A future direction of the present work<br />

will carefully examine the force distribution on the cluster of motors and its<br />

implications on self-organization. A non uniform distribution of forces (as<br />

in the case of long-range interaction) will presumably lead to a dynamically<br />

unstable cluster that will break towards a smaller stable cluster with a larger<br />

(hard-core), but uniformly distributed forces. It would be interesting also to<br />

determine how general are these cooperative effects and whether is possible<br />

to recover the same effects using simple cluster rules concerning the non trivial<br />

additivity of velocities (a n = 2 cluster goes faster than a n = 1 cluster<br />

with half the applied force), which could lead to shock waves and other interesting<br />

phenomena. The main flaw of this chapter is the completely lack<br />

of direct experimental comparison. In this regard, experiments using a GFP<br />

labelled monomeric molecular motor KIF1A, attached to different types of<br />

cargoes, for instance magnetic beads, allowing to fine tune the motor-motor<br />

interaction, will be of special interest to test our predictions.<br />

In the third part, we have considered the adhesion between plasma membrane<br />

and cytoskeletal cortex and its stability under pressure perturbations.<br />

We have used two different approaches based on both a kinetic model for<br />

membrane-cortex linkers, and an energetic description of the detachment process.<br />

We have shown that within both approaches, there is a global link failure<br />

and an abrupt transition from a stable adhesive membrane to an unstable detached<br />

membrane for a critical force on the links. We have shown that the stability<br />

of the adhesion depends strongly on the state of the cell, and in particular,<br />

on both myosin concentration and link concentration. Our results suggest<br />

that the adhesion energy is not an absolute quantity which is experimentally<br />

accessible, but depends on the state of the cell. We have described the mean


5 General conclusions 115<br />

number of blebs in a cell using a coarse grained kinetic model. We are able to<br />

distinguish two qualitatively different regimes: a state where cells use blebbing<br />

to move, corresponding to a state where there are rarely more than one<br />

bleb at a given time, and a state where the cell blebs extensively, in which case<br />

blebs serve as a mechanical buffer keeping the internal pressure or membrane<br />

tension constant. Finally, we have considered the dynamical response of a cell<br />

to mechanical perturbations, and in particular, how the viscoelasticity and the<br />

treadmilling process of the cortex affects the membrane-cortex stability. We<br />

have shown that the instability of the membrane-cortex depends on the rate<br />

at which a pressure perturbation is applied in comparison with the relaxation<br />

time scale of the cytoskeletal cortex. Focussing on the micropipette aspiration<br />

set up, we have identified several possible regimes during a micropipette<br />

extraction: A stable deformation where cortex and membrane deform in conjunction.<br />

A deformation which involves an initial detachment followed by<br />

a repolymerization of new cortex with stable adhesion. In this last regime,<br />

the cortex may eventually become unstable as the elastic stress in the cortex<br />

increases due to myosin activity, leading to both a saltatory and oscillatory<br />

regimes. If the adhesion of the new cortex remains stable, the cell may retract<br />

all the way to the initial state although the perturbation remains applied. We<br />

have modeled the growth of the cortex thickness under a moving membrane<br />

and we have found that there is a critical membrane velocity above which a<br />

stable gel can not be formed. We have tested our theoretical predictions on<br />

experiments found in the literature and on experiments we have conducted<br />

on the Entamoeba histolytica in close collaboration with B. Mauguis and F.<br />

Amblart.<br />

This chapter is specially relevant because it covers many different physical<br />

possibilities regarding adhesion and micropipette aspiration, which can<br />

thoroughly be compared to experiments. However, more experiments need to<br />

be done, specially regarding the dependence of adhesion on myosin and link<br />

density. Labeling of actin in the micropipette experiments will provide evidence<br />

of the several reported regimes, specially for the threshold for cortex<br />

repolymerization. The difficulty of using several radii of micropipettes, do not<br />

allow to clearly demonstrate the dependence of the oscillations in the radius<br />

of curvature alone. It will be also interesting to experimentally address the<br />

threshold pressure for membrane detachment as a function of the loading rate<br />

using the same micropipette set up used with the Entamoeba histolytica. We<br />

are currently working towards this direction in collaboration with B. Mauguis<br />

and F. Amblart.


116 5 General conclusions<br />

As general remarks, this thesis served more as a path or transition towards<br />

a purely biophysical research, than a closed and definite project. The knowledge<br />

and tools learned will hopefully help me to address new biological problems.<br />

I personally believe the future directions in physical biology will tend<br />

to point towards the relation between forces and functionality, ranging from<br />

the coupling between external stimuli and gene network regulation, to the relation<br />

between development and cell morphology and stress. One example<br />

of future direction with involve concepts considered in this thesis and also a<br />

large amount of available data and observations, as well as important medical<br />

applications, is the study of the stability of the mitotic spindle both theoretically<br />

and experimentally. There, concepts from soft-matter (liquid crystals)<br />

and statistical physics (molecular motors) could help to elucidate how such a<br />

dynamical structure auto-organizes and is able to sustain and generate forces<br />

in the range of nano newtons. In addition, the variety of available techniques<br />

allows to compare any possible model prediction with in vivo and reconstituted<br />

systems.


A<br />

Resum en català<br />

El present treball versa sobre aspectes dinàmics de fenòmens relacionats<br />

amb la biofísica i la matèria condensada. L’estructura de la tesi està dividida<br />

en tres parts principals: en primer lloc estudiem la dinàmica de defectes<br />

topològics en monocapes de Langmuir, amb l’intenció de relacionar les propietats<br />

dinàmiques (velocitats) amb els paràmetres materials del sistema (constants<br />

elàstiques). La segona part, està dedicada a la cooperativitat de motors<br />

mol . leculars. Estudiem el mecanisme físic responsable de l’increment<br />

en l’eficiència de motors cooperatius en funció del tipus d’interacció entre<br />

motors. Finalment, estudiem la interacció del cortex i la membrana cel . lular,<br />

donant ènfasi a la relació entre la dinàmica dels lligams entre la membrana i<br />

còrtex, i la viscoelasticitat d’aquest darrer.<br />

A.1 Auto-organització i cooperativitat de motors moleculars<br />

interactuant feblement.<br />

Els motors moleculars són proteïnes capaces de convertir l’energia provinent<br />

de la hidròlisi de l’ATP en treball mecànic, que és utilitzat per generar forces<br />

que permeten una bona part dels moviments cel . luars. El seu comportament<br />

col . lectiu juga un paper important en molts processos biològics, incloent contracció<br />

muscular, transport, motil . litat i divisió cel . lular.<br />

En el darrers anys, el desenvolupament de nous experiments ”in vitro”, ha<br />

permés d’observar motors mol . leculars aïllats. Aquest experiments incloen<br />

assaigs lliscants, pinces òptiques i mesures micromecàniques de força. Els<br />

resultats d’aquests experiments han revelat noves idees sobre els principis<br />

bàsics del motors mol . leculars i han inspirat nous models pel seu comportament<br />

col . lectiu. En aquest context, existeixen força models considerant els<br />

motors enganxats a filaments rígids, i per tant, rígidament acoblats. Aquest


118 A Resum en català<br />

models prediuen un comportament no lineal de la relació entre força aplicada<br />

i velocitat dels motors, que ha estat mesurada experimentalment.<br />

L’assumpció de motors rígidament lligats, no és sempre vàlida. En experiments<br />

recents centrats en tràfic intercel . lular, s’ha demostrat que l’acció<br />

conjunta de grups de motors mol . leculars és necessària quan aquests estan enganxats<br />

a càrregues de tipus fluid (com ara vesícules o tubs de membrana).<br />

En aquest context, però, els motors no estan rígidament lligats (malgrat poder<br />

interactuar entre ells), i poden moure’s de manera indepèndent respecte els<br />

altres motors. Aquest nou escenari de cooperativitat ha inspirat nous models<br />

de cooperació entre motors basats en models discrets, capaços d’explicar<br />

qualitativament l’habilitat de grups de motors d’estirar tubs de membrana<br />

conjuntament. Entre les limitacions d’aquests models, rau la impossibilitat<br />

d’obtenir la distribució de forces de cada motor.<br />

Motivats pel cas de motors estirant una càrrega líquida, en el present<br />

treball estudiarem la situació corresponent a motors interactuant feblement,<br />

sotmesos a una força aplicada només al primer motor, que seria l’equivalent<br />

al front del tub de membrana. La nostre aproximació al problema es basarà<br />

en l’ús del model més simplificat possible de dos estats per un motor. El<br />

nostre model mecanístic ens permetrà obtenir informació quantitativa sobre<br />

l’eficiència i transmissió de força de grups de motors. El nostre objectiu en<br />

aquest treball és aclarir fonomens genèrics de transmissió de força entre motors,<br />

la seva connexió amb el tipus d’interacció entre motors, corves de forçavelocitat<br />

i rendiment col . lectiu de grups de motors.<br />

El model de dos estats per un motor mol . lecular, es basa en el fet que hi ha<br />

dues principals configuracions durant el moviment del motor, corresponents<br />

a un estat lligat, responsable de fer avançar el motor, i un estat deslligat,<br />

on el motor difon. Durant l’estat lligat, el motor està sotmés a un potencial<br />

periòdic que representa la periodicitat del filament on el motor es mou. El<br />

potencial corresponent a l’estat deslligat és pla. La transició de l’estat lligat<br />

a l’estat deslligat es produeix al voltant del mínim de potencial, i correspon<br />

a l’absorció d’una molècula d’ATP. La transició de l’estat deslligat al lligat<br />

està deslocalitzada i té una vida mitja fixada, representant l’alliberació d’un<br />

fosfat un cop l’ATP ha estat hidrolitzat.<br />

En aquest model, el moviment dels motors mol . leculars és descrit amb una<br />

equació de Langevin, que integrarem númericament per tres tipus d’interacció:<br />

repulsiva de curt abast, feblement attractiva i repulsiva de llarg abast.<br />

El grau de cooperativitat entre motors, depèn de la correl . lació entre graus<br />

de llibertat posicionals i ineterns. En el cas de dos motors fortament lligats,<br />

la cooperació es fa evident, ja que quan els estats dels motors són creuats (un


A.1 Auto-organització i cooperativitat de motors moleculars interactuant feblement. 119<br />

en estat lligat i l’altre en estat deslligat), el motor que esta deslligat, enlloc<br />

de difondre, és empés o estirat per l’altre motor en l’estat lligat. Per tant,<br />

el motor en l’estat deslligat avança un periode de manera determinista, i no<br />

degut a fluctuacions com en el cas d’un motor aïllat. En la situació oposada,<br />

on dos motors interactuen amb un potencial molt suau, petites variacions en<br />

la posició dels motors de l’ordre de la periodicitat del filament, es descorrellacionen<br />

amb els estats del motor i perdem la cooperativitat. Aquest cas, el<br />

definirem com a camp mitjà, ja que els motors en l’estat estacionari senten<br />

el mateix potencial d’interacció i la seva velocitat és equivalent a la velocitat<br />

d’un motor amb una força externa N vegades més petita. En el límit de soroll<br />

petit, es pot calcular la corva força-velocitat de manera exacta pel cas de<br />

camp mitjà, que utilitzarem com a referència per comparar amb els altres<br />

casos d’interacció.<br />

En primer lloc, considerem una interacció repulsiva de volum exclós, que<br />

modelem amb un potencial de Lennard-Jones truncat. Genèricament obtenim<br />

un guany respecte el cas de camp mitjà, que depèn del tamany dels motors<br />

moleculars respecte a la periodicitat del filament. A mida que augmentem el<br />

nombre de motors, el guany relatiu va disminuint fins que satura a un valor<br />

constant.<br />

En segon lloc, considerem un potencial d’interacció repulsiu de llarg<br />

abast, que modelem amb el potencial de volum exclós, més un potencial que<br />

decau exponencialment amb un paràmetre que mesura la suavitat. Quan la<br />

força externa supera un cert valor, el potencial suau no és suficient per compensar<br />

la força i els motors comencen a veure el potencial de volum exclós.<br />

Per tant, per un rang de forces, observem que els motors es comporten com<br />

a camp mitjà, però a partir d’un cert valor, els motors comencen a cooperar i<br />

observem una transició cap al comportament descrit pel potencial de volum<br />

exclós.<br />

Finalment, considerem una interacció feblement attractiva. En aquest cas,<br />

observem un fenomen en principi poc intuitiu.<br />

Conclusions<br />

Hem considerat el comportament col . lectiu de motors moleculars per interaccions<br />

de volum exclós, attracció feble i interacció repulsiva de llarg abast.<br />

El redniment col . lectiu és genéricament superior a la del cas de camp mitjà.<br />

Pel cas de motors atractius, observem un nou fenomen de dinàmica de grup<br />

de motors que involucra histèresi i bimodalitat transitòria. Si la reserva de<br />

motors es prou gran, el tamany del grup de motors es reajusta dinàmicament<br />

depènent de la força aplicada. Finalment, hem considerat l’eficiència i la distribució<br />

de força per cada motor del grup. Una limitació del present treball pel


120 A Resum en català<br />

que fa a l’aplicació biològica o biomimètica, és que hem negligit la processivitat<br />

i la depèndència en la força de les transisions entre estats dels motors.<br />

Aquests aspectes els poden incorporar fàcilment en el model i representen la<br />

futura direcció del present treball.<br />

A.2 Mesura de l’anisotropia elàstica a través de la dinàmica de<br />

defectes en monocapes de Langmuir<br />

L’estudi de defectes topològics és genèricament rellevant en força àrees tant<br />

de física com en biologia, comprenent desde cosmologia a heli superfluid<br />

o divisió cel . lular. Una part important de la motivació pel seu estudi ve del<br />

seu grau d’universalitat, com passa amb qualsevol sistema amb paràmetre<br />

d’ordre. Les seves propietats principals són indepèndents de la underlying<br />

física, i estan només determinades per simetries, la dimensió del paràmetre<br />

d’ordre, el defecte i el sistema. Aquest grau d’universalitat ha estat utilitzat<br />

per sistemes pertanyents a la matèra condensada, com ara els cristalls<br />

líquids, que permeten, a través d’experiments relativament senzills, obtenir<br />

informació sobre àrees de la física comletament diferents.<br />

L’estudi del moviment de defectes, la seva interacció i anihilació és particularment<br />

interessant i ha estat estudiat extensivament per nombrosos grups<br />

d’investigació. Un aspecte sorprenent del fenomen d’anihilació, que malgrat<br />

no estar completament entès, és observat tan experimentalment com<br />

mitjançant càlcul numèric, és la sistemàtica diferència de velocitats entre<br />

defectes negatius i positius, sent el darrer sempre més ràpid. S’ha demostrat<br />

que efectes hidrodinàmics i l’anisotropia de les constants elàstiques<br />

(diferència entre les constants elàstiques) contribueixen genèricament a explicar<br />

aquesta assimetria en les velocitats. De fet, en el context de cristalls liquids<br />

3-dimensionals, l’efecte hidrodinàmic domina per sobre de l’anisotropia<br />

elàstica, impedint qualsevol intent de relacionar quantitativament l’elasticitat<br />

del material amb la dinàmica dels defectes, cosa que seria una alternativa<br />

força interessant als mètodes tradicionals per determinar constants elàstiques.<br />

El nostre estudi considerarà la situació oposada, on els efectes hidrodinàmics<br />

són negligibles. Per aquest fí, hem considerat la dinàmica de defectes en<br />

monocapes de Langmuir exteses a l’interface aire/aigua. Contràriament als<br />

cristalls líquids 3-dimensionals, on la viscositat rotacional i tranlacional són<br />

del mateix ordre de magnitud, en monocapes de Langmuir (2-dimensionals),<br />

les molècules poden rotar sense generar efectes hidrodinàmics. En aquest<br />

cas, la dinàmica de defectes, i més concretament, l’assimetria de velocitats<br />

es pot explicar només a partir de les propietats elàstiques del material. Com


A.2 Mesura de l’anisotropia elàstica a través de la dinàmica de defectes en monocapes de Langmuir 121<br />

a conseqüencia, mesures de la diferència en mobilitat dels defectes negatius<br />

i positius, ens permetrà d’obtenir informació sobre l’anisotropia elàstica del<br />

material.<br />

En el present treball està dividit en dos parts. En primer lloc, explicarem<br />

el dispositiu experimental utilitzat per mesurar la dinàmica de defectes. A<br />

continuació, demostrarem que l’observada assimetria en la mobilitat dels defectes<br />

pot ser explicada mitjançant únicament propietats topològiques dels<br />

defectes. Presentarem també un model quantitatiu que permetrà extreure la<br />

depèndència de l’anisotropia elàstica de la pressió superficial, i estimar les<br />

constants elàstiques del sistema.<br />

Dispositiu experimental<br />

La part experimental del treball ha estat feta mitjançant un recipient de tefló<br />

on la monocapa s’exten sobre una àrea limitada per dos barreres mòvils, que<br />

hem utilitzat per controlar la pressió superficial.<br />

La monocapa és composa del fotosensible azobenzé amfifílic 8Az3COOH.<br />

Totes les classes d’azobenzé estan constituides d’un nucli format per dos<br />

anells de fenil lligats a un doble enllaç de nitrogen, que poden adoptar dos<br />

configuracions geomètriques diferents: la configuració trans i la configuració<br />

cis. Degut a les propetats geomètriques i elèctriques, l’isomer trans mostra un<br />

paràmetre d’ordre orientacional.<br />

Els azobenzens presenten dos propietats que usarem pels nostres propòsits<br />

experimentals. Per una banda, es poden induir transisions entre les configuracions<br />

cis-trans irradiant les molècules amb llum d’una longitud d’ona<br />

apropiada. Les molécules cis relaxen espontàniament a la configuració trans<br />

per fluctuacions tèrmiques. Degut a les propietats geomètriques completament<br />

diferents, les barreges de isòmers cis i trans s’organitzen de manera<br />

diferent a l’interfície d’aire/aigua. Per altra banda, les mesofases de l’isomer<br />

trans, tenen una densitat superfícial relativament petita. Consqüentment, les<br />

monocapes compostes d’aquest isomer presenten un ordre orientacional de<br />

llarg abast, però un desordre posicional, cosa que permet que els defectes es<br />

moguin amb facilitat.<br />

Barrejes de tran-cis isomers on la concentració de cis és dominant, resulta<br />

en la formació de gotes d’isomer trans, que presenten un defecte de càrrega<br />

+1 al centre. Utilitzarem aquestes gotes per estudiar l’anihilació de defectes<br />

resultant de la fusió d’un parell de gotes.<br />

El métode utilitzat per la visualització del paràmetre d’ordre, està basat<br />

en el concepte d’angle de Brewster, que és l’angle d’incidència en el qual un<br />

raig de llum polaritzat es reflecta completament. Si dipositem una capa de


122 A Resum en català<br />

material molt fina al pla d’incidència, una petita part de la llum es reflectirà,<br />

presentant en general una polarització diferent a la de la llum incident. Calibrant<br />

apropiadament la senyal rebuda, es pot mesurar unívocament la direcció<br />

del paràmetre d’ordre orientacional.<br />

Els resultats experimentals revel . len les molècules amfifíliques dins de<br />

les gotes, es caracteritzen piter formar un angle constant respecte la normal<br />

de l’interfície aire/aigua. Per tant, podem descriure l’organització molecular<br />

utilitzant un vector 2-dimensional. Al fusionar-se dues gotes, degut a la<br />

conservació de càrrega topològica, apareixen dos defectes de càrrega fraccionaria<br />

−1/2 a la frontera. Típicament, un dels dos defectes fraccionaris es<br />

fuciona amb un dels dos defectes centrals, creant un defecte +1/2 a la frontera.<br />

A continuació, els dos defectes de la frontera s’atrauen degut a la càrrega<br />

oposada i s’acaben fusionant. El defecte positiu sempre es mou més depressa<br />

que el negatiu.<br />

Descripció teòrica<br />

L’atracció elàstica de defectes de càrrega oposada és la responsable de l’anihilació<br />

dels defectes ±1/2 a la frontera. La velocitat a la que els defectes es mouen<br />

depèn de la quantitat d’energia dissipada. En monocapes de Langmuir, la<br />

viscositat rotacional deguda a la reorientació de les molécules, és dominant<br />

respecte a la viscositat translacional de les molècules. Per tant, l’explicació<br />

de l’assimetria de velocitats entre els defectes positius i negatius rau només<br />

en l’anisotropia de les constants elàstiques.<br />

Per tal de descriure els defectes, utilitzem una energia lliure que inclou<br />

els ordres més baixos en la distorsió de la direcció de les molècules.<br />

L’energia conté dos termes caracteritzats per dos constants elàstiques: un<br />

terme d’obertura, anomenat ”splay”, que penalitza les configuracions de<br />

molècules, que partint d’un mateix punt, formen un angle entre elles, i un<br />

terme de curvatura entre molècules, anomenat ”bend”, que penalitza configuracions<br />

curvades de molècules paral . leles.<br />

La forma de les solucions de defectes positius no depèn de l’anisotropia de<br />

les constants, mentres que pel cas de defectes negatius, aquests es deformen<br />

depèndent de l’anisotropia de les constants.<br />

La dissipació que els defectes creen al moure’s, es pot demostrar que és<br />

l’energia que costa moure un defecte al voltant del seu eix. Per el cas en<br />

que l’anisotropia és nul . la, cada molècula, tant del defecte negatiu com positiu,<br />

rota la mateixa quantitat i per tant, els dos defectes dissipen la mateixa<br />

quantitat d’energia. Per contra, quan l’anisotropia és diferent de zero, la deformació<br />

del defecte negatiu implica que hi ha certes molècules es mouen


A.3 Interaccions de membrana i còrtex: Experiments amb micropipeta i cèl . lules amb blebs 123<br />

menys i d’altres, que s’han de moure més respecte al cas isòtrop. Degut al<br />

fet que la dissipació és una funció quadràtica en la rotació del defecte, els<br />

defectes negatius sempre dissipen més, i per tant, es mouen més lentament.<br />

Utilitzant l’expressió de la dissipació dels defectes demostrem que el quocient<br />

de velocitats del defecte negatiu i positiu depèn només de l’anisotropia<br />

de les constants elàstiques. Per tant, mesurant experimentalment les respectives<br />

velocitats dels defectes quan s’anihilen, podem obtenir el valor de<br />

l’anisotropia. Modificant la pressió superficial mitjançant el nostre dispositiu<br />

experimental, obtenim la depèndència de l’anisotropia en funció de la pressió.<br />

Utilitzant el fet que la mitjana geomètrica de les constants elàstiques s’ha demostrat<br />

en la literatura que és indepèndent de la pressió, podem relacionar<br />

directament el valor absolut de les constants amb l’anisotropia que obtenim.<br />

Els nostres resultats mostren que la contribució ”splay” incrementa amb la<br />

pressió, mentres que la contribució ”bend” disminueix. Aquests resultats són<br />

consistents amb la literatura i amb el fonomen anomenat ”aggragats H”, que<br />

consisteix en la preferència de les molècules d’azobenzé de formar agregats<br />

on els plans de les molècules són paral . lels. Per tant, augmentant la pressió,<br />

s’afavoreix aquest fenòmen que penalitza la contribució ”splay”, en contra de<br />

la contribució ”bend”.<br />

Conclusions<br />

Hem estudiat monocapes de Langmuir, on la dinàmica de defectes pot ser explicada<br />

únicament per l’anisotropia de les constants elàstiques. Hem mostrat<br />

que el nostre sistema, juntament amb un model senzill, permet d’utilitzar<br />

mesures dinàmiques per obtenir informació de la depèndència de paràmetres<br />

del material (en aquest cas, les constants elàstiques), generalment difícils de<br />

mesurar directament, en funció de paràmetres termodinàmics bàsics de la<br />

monocapa com la pressió superficial.<br />

Aquest treball ha estat fet en col . laboració amb els doctors J. Ignés-Mullol<br />

i F. Sagués.<br />

A.3 Interaccions de membrana i còrtex: Experiments amb<br />

micropipeta i cèl . lules amb blebs<br />

La membrana cel . lular és essencial per la cèl . lula. Serveix d’envolcall, defineix<br />

les fronteres, fa de mediadora de l’intercanvi de materials entre el cistosol<br />

i el medi extracel . lular i conté proteïnes que actuen com a sensors dels<br />

senyals externs que permeten a la cèl . lula reaccionar i adaptar el seu comportament<br />

en resposta a canvis en l’ambient. Sota la membrana hi ha el còrtex,


124 A Resum en català<br />

format per una xarxa d’actina que es reorganitza constantment. El còrtex es<br />

responsable, entre d’altres coses, de la forma de la cèl . lula, el moviment i el<br />

transport cel . lular intern. La membrana i el còrtex estan adherits mitjançant<br />

interaccions febles de lípids de membrana i proteïnes del còrtex i per interaccions<br />

fortes de proteïnes. La pèrdua d’adhesió té com a conseqüència la<br />

formació de blebs, que són unes protuberàncies de forma esfèrica que es formen<br />

quan un troç de membrana es desenganxa del còrtex. Normalment, la<br />

presencia de blebs és indicació de mort cel . lular, encara que hi ha exemples<br />

on les cèl . lules s’aprofiten d’aquest fenòmen per moure’s. En aquesta part de<br />

la tesi, hem estudiat el mecanisme d’adhesió de la membrana i el còrtex, i la<br />

seva resposta dinàmica a perturbacions mecàniques que eventualment tenen<br />

com a resultat la separació de la membrana del còrtex.<br />

Descripció teòrica<br />

Per modelitzar l’adhesió entre la membrana i el còrtex, hem considerat els<br />

lligams, que normalment són proteïnes, com a molles, i per tant, exercint una<br />

força que es proporcional a la seva extensió. La velocitat a la que la membrana<br />

escapa del còrtex ve donada per el balanç de força entre la pressió aplicada a<br />

la membrana i la força recuperadora dels lligams. Els lligams s’enganxen i es<br />

desenganxen constantment de la membrana a una certa freqüència que depèn<br />

de la força que se’ls hi aplica. Considerant l’equació pel balanç de força en<br />

funció de la densitat de lligams i l’equació de conservació del nombre total de<br />

lligams, que determina la fracció de lligams enganxats, obtenim que existeix<br />

una força crítica a partir de la qual els lligams són incapaços de compensar la<br />

pressió i acabant fallant alliberant la membrana del còrtex.<br />

Si considerem ara una descripció energètica del mateix problema, on comparem<br />

el guany d’energia de desenganxar la membrana degut a la pressió i la<br />

pérdua d’energia degut a la pérdua de lligams i a la tensió de la membrana,<br />

obtenim, com en el cas anterior, una inestabilitat on la membrana es desenganxa<br />

per una pressió crítica. La transició de l’estat enganxat a l’estat desenganxat<br />

ve donada per una barrera energètica que disminueix amb la pressió.<br />

Podem definir un temps característic de pas que depèn exponencialment de la<br />

barrera energètica. El temps de pas ens dóna una estimació de la freqüència a<br />

la que un bleb es crea. Al mateix temps, els blebs es retrauen i desapareixen<br />

després d’un temps característic que depèn del temps que es triga a formar<br />

un nou còrtex sota la membrana desenganxada. El balanç entre els blebs que<br />

es formen i els que desapareixen dóna el número mitjà de blebs en una cèllula,<br />

que depèn de la seva pressió interna. Podem identificar dos règims: per<br />

una pressió moderada, la cèl . lula es manté estable i no apareix cap bleb. Quan


A.3 Interaccions de membrana i còrtex: Experiments amb micropipeta i cèl . lules amb blebs 125<br />

la pressió augmenta suficientment, comencen a aparèixer blebs i la pressió<br />

es manté constant. Aquests dos règims poden correspondre a cèl . lules que<br />

utilitzen pocs (un o dos) blebs per moure’s en el cas de pressió moderada,<br />

ocèl . lules que estan morint, on s’observen blebs per tota la superfície de la<br />

cèl . lula.<br />

Experiments basats en perturbacions mecàniques demostren que les cèllules<br />

es comporten elàsticament per temps petits i una de forma fluida per<br />

temps llargs. Aquest comportament, anomenat viscoelàstic, es pot modelitzar<br />

fàcilment amb un model que interpola la descripció d’un sòlid i un liquid,<br />

degut a Maxwell. Hem utilitzat aquesta descripció per modelar el còrtex de la<br />

cèl . lula. Aplicant una pressió externa que creix a una certa velocitat, que utilitzem<br />

com a paràmetre, podem demostrar que la comparació entre el temps<br />

típic de relaxació del còrtex, que determina la transició entre el comportament<br />

elàstic i fluid de la cèl . lula, i l’escala de temps a la que la pressió augmenta<br />

determina la força màxima transmesa als lligams. Intuitivament, la força aplicada<br />

a un lligam serà la quantitat de deformació del còrtex en el temps durant<br />

el qual el còrtex s’ha comportat elàsticament. Aquesta força ha de ser comparada<br />

amb la força crítica a partir de la qual els lligams es trenquen i la<br />

membrana es desenganxa del còrtex.<br />

Degut a que la força transmesa als lligams depèn tant de la pressió final<br />

que apliquem a la membrana com del ritme al qual incrementem la pressió,<br />

obtenim un diagrama d’estabilitat que ens diu que encara que apliquem pressions<br />

molt fortes, els lligams són capaços de no trencar-se si el ritme al que<br />

apliquem la pressió es prou lent. Igualment, per una pressió moderada, podem<br />

aconseguir trencar els lligams si la velocitat a la qual augmentem la pressió<br />

es prou elevada. Aquest resultat és força rellevant en el cas d’experiments<br />

realitzats amb micropipetes, ja que aplicar una pressió mai és un procés instantani.<br />

Finalment, usant els conceptes anteriors, hem identificat tots els possibles<br />

règims d’aspiració mitjançant una micropipeta: Si el ritme d’aspiració i la<br />

pressió són prou petits, la membrana i el còrtex són deformats fins que arriben<br />

a un nou equilibri sense que la membrana i el còrtex es desenganxin.<br />

Si el ritme d’aspiració es prou ràpid, la membrana es densenganxa del còrtex<br />

i flueix fins que un nou còrtex es format. En auqest cas, hi ha dos possibilitats.<br />

Si quan el nou còrtex es contrau crea una força suficientment gran, els<br />

lligams poden tornar-se a trencar generant un moviment oscil . latori on el trencament<br />

i la formació de còrtex es succeeixen consecutivament. Si la força de<br />

contracció del nou còrtex es suficientment petita, la membrana i còrtex es<br />

retrauen contínuament fins al nou equilibri.


126 A Resum en català<br />

Juntament amb un grup experimental de l’Institut Curie, hem comprobat<br />

experimentalment els règims predits teòricament i em reproduit quantitativament<br />

les oscil . lacions observades en el sistema de l’ameba Entamoeba histolytica.<br />

Conclusions<br />

Hem estudiat exhaustivament el fenomen d’adhesió entre la membrana i el<br />

còrtex i la seva estabilitat, sotmesos a una perturbació en la pressió. Hem<br />

usat una descripció microscòpica dels lligams i una descripció energètica<br />

de l’adhesió. En ambdós casos hem obtingut una inestabilitat global on els<br />

lligams es trenquen per una força crítica. Hem descrit el número mitjà de<br />

blebs en una cèl . lula en funció de la pressió, obtenint dos règims qualitativament<br />

diferents. Per una banda, per pressions moderades, la cèl . lula roman<br />

sense blebs, mentre que per pressions superiors, es creen blebs que mantenen<br />

la pressió interna de la cèl . lula constant.<br />

Hem considerat la resposta dinàmica de la cèl . lula a perturbacions mecàniques.<br />

En particular, em emfatitzat com el caràcter viscoelàstic de la cèl . lula afecta<br />

la transmissió de força als lligams i l’estabilitat d’aquests. Hem obtingut que<br />

la pressió crítica en un experiment depen fortament de la velocitat a la qual<br />

s’aplica la pressió. Com a conseqüència, hem identificat tots els possibles<br />

règims dinàmics durant l’aspiració amb una micropipeta, que inclouen: Una<br />

deformació on la membrana i el còrtex romanen enganxats. Una deformació<br />

on la membrana es desenganxa del còrtex i posteriorment el nou còrtex retrau<br />

fins la nova posició d’equilibri, i finalment, una deformació on la membrana<br />

es desenganxa durant la retracció de forma periòdica.<br />

Utilitzant mesures obtingudes de la literatura i experiments realitzats especialment,<br />

hem pogut comprobar les nostres prediccions teòriques.<br />

Aquest treball ha estat fet en col . laboració amb l’estudiant de tesi B. Mauguis<br />

i el doctor F. Amblart.


B<br />

Résumé en Français<br />

Ce travail traite d’aspects dynamiques de certains phénomènes reliés à la<br />

biophysique et la matière condensée. La thèse est divisée en trois parties<br />

principales. D’abord, nous étudions la dynamique des défauts topologiques<br />

des monocouches de Langmuir, avec l’intention de relier les proprietés dynamiques<br />

(vitesse) avec les paramètres matériels du système (constants élastiques).<br />

La deuxième partie est dédiée à la coopérativité des moteurs moléculaires.<br />

Nous étudions le mécanisme physique responsable du partage de force entre<br />

moteurs coopératifs en fonction du type d’interaction entre moteurs. Finalement,<br />

nous étudions l’interaction entre le cortex et la membrane cellulaire, en<br />

mettant l’accent sur le rapport entre la dynamique des liens entre la membrane<br />

et le cortex, et la viscoélasticité de ce dernier.<br />

B.1 Auto-organisation et coopérativité des moteurs moléculaires<br />

interagissant faiblement<br />

Les moteurs moléculaires sont des protéines capables de transformer l’énergie<br />

provenant de l’hydrolyse de l’ATP en travail mécanique. Les forces ainsi<br />

générées permettent une grand partie des mouvements cellulaires. Les processus<br />

collectifs jouent un rôle important dans beaucoup de processus biologiques,<br />

dont la contraction musculaire, le transport, la motilité et la division<br />

cellulaire.<br />

Ces dernières années, le développement des nouvelles expériences ”in<br />

vitro” à permit d’observer l’activité de moteurs moléculaires isolés. Ces<br />

expériences incluent ”gliding assays”, pinces optiques et mesures micromécaniques<br />

de force. Les résultats de ces expériences ont révélé de nouvelles idées<br />

sur les principes basiques des moteurs moléculaires et ont inspiré de nouveaux<br />

modèles pour comprendre leur comportement collectif. Dans ce con-


128 B Résumé en Français<br />

texte, de nombreux modèles suppose que les moteurs sont rigidement liés<br />

les un aux autres. Ces modèles prédisent un comportement non linéaire<br />

de la relation entre la force appliquée et la vitesse des moteurs, qui a été<br />

mesuré expérimentalement. L’hypothèse de moteurs rigidement liées n’est<br />

pas toujours valable. De récentes expériences concentres sur le traffic intracellulaire<br />

ont démontrées que la mise en mouvement de charges fluides<br />

(comme des vésicules ou des tubes de membrane) nécessitait également<br />

l’action de plusieurs moteurs moléculaires. Dans ce contexte, les moteurs<br />

ne sont pas rigidement lies entre eux et peuvent en principe se déplacer de<br />

manière indépendante. Ce nouveau scénario de coopérativité a inspiré de<br />

nouveaux modèles d’interaction entre moteurs basés en modèles discrets, capables<br />

d’expliquer qualitativement l’habileté de groupes de moteurs d’étirer<br />

tubs de membrane conjointement.<br />

Motivés par le cas de moteurs tirant un chargement liquide, dans cet travail<br />

nous allons étudier la situation qui correspond aux moteurs interagissant<br />

faiblement, soumis a une force appliqué seulement au premier moteur<br />

(par exemple, celui qui se trouve a l’extrémité du tube). Nous supposons un<br />

modèle à deux états pour chaque moteur. Notre modèle mécanique nous permet<br />

obtenir des informations quantitatives sur l’efficacité et la transmission de<br />

force dans un groupe de moteurs. Notre objectif dans ce travail est comprendre<br />

des phénomènes génériques de transmission de force entre moteurs, sa<br />

connexion avec le genre d’interaction entre moteurs, courbes de force-vitesse<br />

et efficacité collectif d’un groupes de moteurs.<br />

Le modèle de deux états pour un moteur moléculaire est basé dans le fait<br />

qu’il y a deux configurations principales pendant le mouvement du moteur,<br />

correspondant à un état lié, responsable de faire avancer le moteur, et un état<br />

délié, dans lequel le moteur diffuse. Dans l’état lié, le moteur est soumis à<br />

un potentiel périodique qui représente la périodicité du filament sur lequel<br />

le moteur marche. Le potentiel qui correspond à l’état délié est plain. La<br />

transition de l’état lié à l’état délié correspond a l’absorption d’une molécule<br />

d’ATP. L’état délié a une durée de vie moyenne fixé, représentant la libération<br />

d’un phosphate, une fois l’ATP hydrolysé.<br />

Dans ce modèle, le mouvement des moteurs moléculaires est décrit avec<br />

une équation de Langevin, que nous intégrons numériquement pour trois genres<br />

d’interaction : répulsive à court portée, faiblement attractive et répulsive à<br />

long portée. Le degré de cooperativité entre moteurs, dépend de la corrélation<br />

entre degrés de liberté positionel et interne. Dans le cas de deux moteurs<br />

fortement lies, la coopération est triviale, puisque un moteur dans un état<br />

délié ne diffuse pas, mais se déplace à la vitesse d’un moteur lié voisin. Par


B.1 Auto-organisation et coopérativité des moteurs moléculaires interagissant faiblement 129<br />

conséquence, le moteur en l’état délié avance de manière déterministe et non<br />

a cause des fluctuations comme dans le cas d’un moteur isolé.<br />

Dans la situation opposé, ou deux moteurs interagissent avec un potentiel<br />

très doux, de petites variations en la position des moteurs de l’ordre de la<br />

périodicité du filament ne sont pas corrélés avec l’état du moteur et perdent<br />

la cooperativité. Cet cas, nous allons le définir comme a champ moyenne,<br />

étant donné que les moteurs en l’état stationnaire soufrent le même potentiel<br />

d’interaction et sa vitesse est équivalent à la vitesse d’un moteur avec la force<br />

externe N fois plus petite. Dans le limite de bruit petit, on peut calculer la<br />

courbe force-vitesse de manière exacte pour le cas de champ moyen, que<br />

on utilisera comme référence pour faire la comparaison avec les autres cas<br />

d’interaction.<br />

D’abord, nous considérons une interaction répulsive de volume exclus<br />

que nous modelons avec un intervalle du potentiel de Lennard-Jones. Nous<br />

obtenons un bénéfice par rapport au cas de champ moyen, qui dépend de la<br />

taille des moteurs moléculaires par rapport a la périodicité du filament. Fur à<br />

mesure que on augmente le numéro de moteurs, le bénéfice relatif diminues<br />

jusque s’arrête a un valeur constant.<br />

En deuxième lieu, nous considérons un potentiel d’interaction répulsive<br />

de long portée que nous modelons avec le potentiel de volume exclus, avec<br />

un potentiel que décline de façon exponentiel avec un paramètre que mesure<br />

la douceur. Quand la force externe surpasse un certaine valeur, le potentiel<br />

doux n’est pas suffisante pour compenser la force et les moteurs commencent<br />

a voir le potentiel de volume exclus. Par conséquence pour un rang de forces,<br />

nous observons que les moteurs se conduisent comme a champ moyen, mais a<br />

partir d’un certaine valeur, les moteurs commencent à coopérer et nous observons<br />

une transition vers au comportement décrit par le potentiel de volume<br />

exclus. Finalement, nous considérons une interaction faiblement attrayante.<br />

Dans ce cas, nous observons un phénomène, en principe, peu intuitif.<br />

Conclusions<br />

Nous avons considéré le comportement collective de moteurs moléculaires<br />

par interactions de volume exclus, attraction faible et interaction répulsive<br />

de long portée. Le rendement collectif est supérieur au du champ moyen.<br />

En le cas de moteurs attrayantes, nous observons un nouveau phénomène de<br />

dynamique de groupe de moteurs que implique hysteresis et bimodalité transitoire.<br />

Si la réserve de moteurs est assez grand, la taille du groupe de moteurs<br />

se réajuste dynamiquement en dépendant de la force appliqué. Finalement,<br />

nous avons considéré l’efficacité et la distribution de force pour chaque


130 B Résumé en Français<br />

moteur du groupe. Une limitation du présent travail pour ce qui concerne a<br />

l’application biologique ou biomemitique, c’est que nous avons négligeait la<br />

processivité et la dépendance en la force des transitions entre les états des<br />

moteurs. Ces aspects on les peut incorporer facilement dans le modèle et<br />

représentent la future direction de ce travail.<br />

B.2 Mesure de anisotropie élastique a travers de la dynamique de<br />

défauts en monocapes de Langmuir<br />

L’étude des défauts topologiques est relevant en beaucoup de zones aussi<br />

de physique comme de biologie,comprennent des la cosmologie a heli superfluide<br />

ou division cellulaire. Une partie important de la motivation pour<br />

son étude vient de son degré d’universalité, comme passe t’il avec quelque<br />

système avec paramètre d’ordre. Ses propretés principaux sont seulement<br />

déterminés per symétries, la dimension du paramètre d’ordre, le défaut et<br />

le système. Cet degré d’universalité aété utilise par systèmes appartenants à<br />

la matière condensée, comme les cristaux liquides, qui permettent, a travers<br />

d’expériences relativement simples, obtenir information sur matières de la<br />

physique différents.<br />

L’étude du mouvement de défauts, sa interaction et annihilation est particulièrement<br />

intéressant et il a était étudié par des nombreux groupes de<br />

recherche. Un aspect surprenant du phénomène d’annihilation, qui , malgré<br />

ne pas être entendu, est observé tant expérimentalement comme grâce a un<br />

calcul numérique, est la systématique différence de vitesses entre défauts<br />

négatifs et positifs, soient le dernière toujours le plus rapide. On a démontré<br />

que les effets hydrodynamiques et l’anisotropie des constants élastiques<br />

(différence entre les constants élastiques) contribuaient à expliquer cette<br />

asymétrie en les vitesses. En fait, dans le contexte de cristaux liquides 3dimensionels,<br />

l’effet hydrodynamique domine au dessus de l’anisotropie<br />

élastique, empêchant tous les essais pour rapporter quantitativement l’élasticité<br />

du matériel avec la dynamique des défauts, que serait une alternative très<br />

intéressant aux méthodes connus pour déterminer constants élastiques. Notre<br />

étude prends en considération la situation opposée, ou les effets hydrodynamiques<br />

sont négligeables. Pour cet but, nous avons considéré la dynamique<br />

des défauts en monocapes de Langmuir étendues à l’interface aire/eau. Contrairement<br />

aux cristaux liquides 3-dimensionnels, ou la viscosité rotacional et<br />

translacional sont du même ordre de magnitude, en monocapes de Langmuir(2dimensionnel),<br />

les molécules peuvent roter sans gérer effets hydrodynamiques.<br />

En ce cas, la dynamique de défauts, et en concret, l’asymétrie de vitesses,


B.2 Mesure de anisotropie élastique a travers de la dynamique de défauts en monocapes de Langmuir 131<br />

peut être expliqué seulement a partir des proprettes élastiques de la matière.<br />

Comme a conséquence, mesures de la différence en mobilité des défauts<br />

négatives et positives, nous permettra d’obtenir information sur l’anisotropie<br />

élastique du matériel. Dans cet travail il est divise en deux parties. D’abord,<br />

nous allons expliquer le dispositif expérimental utilise pour mesurer la dynamique<br />

de défauts. En suite, nous démontrerons que l’asymétrie observée<br />

dans la mobilité des défauts peut être explique seulement grâce aux proprettes<br />

topologiques des défauts. Nous allons aussi présenter un modèle quantitatif<br />

qui permettra extraire la dépendance de la anisotropie élastique de la pression<br />

superficiel, et estimer les constantes élastiques du système.<br />

Dispositif expérimental<br />

La partie expérimental du travail à été fait a travers un récipient de Téflon ou<br />

la monocape s’étends sur une zone limitée par deux barrières mobiles, que<br />

nous avons utilise pour surveiller la pression superficielle. La monocape est<br />

composé du photosensible azobenzene amphiphilique 8Az3COOH. Tour les<br />

classes d’azobenzene sont constitues d’un nucleus formé pour deux anneaux<br />

de fenil lies a un double liaison de nitrogène, que peuvent adopter deux configurations<br />

géométriques et électriques, l’isomère trans montre un paramètre<br />

d’ordre d’orientation.<br />

Les azobenzeness présentent deux propretés que nous utiliserons pour<br />

nôtres finalités expérimentales. D’un coté, on peut induire transitions entre<br />

les configurations cis-trans irradiant les molécules avec une lumière d’une<br />

longueur d’onde convenable. Les molécules cis relaxent spontanément à la<br />

configuration trans par fluctuations thermiques. Dû aux propretés géométriques<br />

complètement différent, les mélanges d’isomères cis et trans, s’organisent<br />

de manière différent à l’interphase d’aire/eau. De l’autre coté, les mesofaces<br />

de l’isomère trans,ont une densité superficielle relativement petite. Par<br />

conséquence, les monocapes composes avec cet isomère, présentent un défaut<br />

de chargement +1 au centre. Nous allons utiliser ces gouttes pour étudier<br />

l’annihilation des défauts résultants de la fusion de deux gouttes. Le métope<br />

utilisé pour la visualisation du paramètre d’ordre, est basé en le concept<br />

d’angle de Brewster, que c’est l’angle d’incidence dans lequel un rayon de<br />

lumière polarisé réflexe complètement. Si on dépose une cape de matérielle<br />

très fine au plain d’incidence, une petite partie de la lumière sera reflectie,<br />

présentant en général une polarisation différent a la de la lumière incident.<br />

En calibrant convenablement le signal reçu, on peut mesurer sans erreur, la<br />

direction du paramètre d’ordre d’orientation.<br />

Les résultats expérimentales révèlent les molécules amphiphiliques audedans<br />

des gouttes, se caractérisent pour être partie d’un angle constant


132 B Résumé en Français<br />

par rapport la normal de l’interphase aire/eau. Par consequence, nous pouvons<br />

decrire l’organisation moleculaire utilitsant un vecteur 2-dimensionnel.<br />

Quand deux gouttes se fusionnent, grâce a la conservation du chargement<br />

topologique, apparaîtrez deux défauts de chargement fractionnaire -1/2 a la<br />

frontière. Typiquement, un des deux défauts fractionnaires se fusionne avec<br />

un des deux défauts centrales, en créant un défaut +1/2 a la frontière. Tout<br />

suite, les deux défauts de la frontière s’attirent a cause de le chargement opposé<br />

et s’y fusionnent. Le défaut positif se remue toujours plus vite que le<br />

négatif.<br />

Description théorique<br />

L’attirance élastique de défauts de chargement opposé est la responsable de<br />

l’annihilation des défauts ±1/2 à la frontière. La vitesse a laquelle les défauts<br />

se meuvent dépend de la quantité de l’énergie dissipé. En monocapes de<br />

Langmuir, la viscosité rotationelle a cause de la réorientation des molécules,<br />

est dominant par rapport à la viscosité translationelle des molécules. Par<br />

conséquence, l’interprétation de l’asymétrie de vitesses entre les défauts positifs<br />

et négatifs est seulement en l’anisotropie des constants élastiques.<br />

A fin de décrire les défauts, nous utilisons une énergie libre que inclure les<br />

ordres plus baisses en la distorsion de la direction de les molécules. L’énergie<br />

contient deux termes caractérisés par deux constants élastiques : un terme<br />

d’oberture, nommé ”splay”, qui pénalise les configurations de molécules, que<br />

partant d’un même point, forment un angle entre elles , et un terme de courbature<br />

entre molécules, nommé ”bend”, qui pénalise configurations courbées<br />

de molécules parallèles.<br />

La forme des solutions de défauts positifs ne dépend pas de l’anisotropie<br />

des constants, tandis que en le cas des défauts négatifs, ces-ci se déforment en<br />

dépendant de l’anisotropie des constants. La dissipation que les défauts créent<br />

en se déplaçant, on peut démontrer que c’est l’énergie que coûte déplacer un<br />

défaut au tour de son axe. Dans le cas en que l’anisotropie est nulle, chaque<br />

molécule, soit d’un effet négatif comme positif, rota la même quantité et, par<br />

conséquence, les deux défauts dissipent la même quantité d’énergie. Par contre,<br />

quand l’anisotropie est différent de zéro, la déformation du défaut négatif<br />

implique qu’il y a certaines molécules qui se déplacent moins et d’autres<br />

qu’on de se déplacer plus au cas isotrope. A cause du fait que la dissipation<br />

est une fonction quadratique en la rotation du défaut, les défauts négatifs se<br />

dissipent toujours plus, et par conséquence, se déplacent plus lentement.<br />

Utilisant l’expression de la dissipation des défauts, nous démontrons que<br />

le quotient de vitesses du défaut négatif et positif dépend seulement de


B.3 Interactions de membrane et cortex : Expériences avec micropipette et cellules avec blebs 133<br />

l’anisotropie des constants élastiques. Par conséquence, mesurant expérimentalement<br />

les respectives vitesses des défauts quand elles s’annihilent, nous pouvons<br />

obtenir le valeur de l’anisotropie. En modifiant la pression superficiel a travers<br />

notre dispositif expérimental, nous obtenons la dépendance de l’anisotropie<br />

en fonction de la pression. En utilisant le fait que la moyenne géométrique<br />

des constantes élastiques on a démontré en la littérature que est indépendant<br />

de la pression, nous pouvons mètre en relation directement le valeur absolu<br />

des constants avec l’anisotropie que nous obtenons. Nos résultats nous montrent<br />

que la contribution ”splay” accroisse avec la pression, tandis que la contribution<br />

”bend” diminue. Ces résultats sont consistent avec la littérature et<br />

avec le phénomène nommé ”agrégées H”, que consistent en la préférence<br />

des molécules d’azobenzene de former agrées ou les plans des molécules<br />

sont parallèles. Par conséquence, augmentant la pression, se favorise cet<br />

phénomène que pénalise la contribution ”splay” par rapport a la contribution<br />

”bend”.<br />

Conclusions<br />

Nous avons étudié des monocapes de Langmuir, ou la dynamique des défauts<br />

peut être expliqué uniquement par l’anisotropie des constants élastiques.<br />

Nous avons montré que notre système, joint avec un modèle simple, permet<br />

d’utiliser mesures dynamiques pour obtenir information de la dépendance<br />

de paramètres du matériel)en ce cas, les constants élastiques), généralement<br />

difficiles de mesurer directement, en fonction des paramètres thermodynamiques<br />

basiques de la monocape comme la pression superficiel. Cet travail<br />

a été fait en collaboration avec les docteurs J.Ignés-Mullol et F. Sagués.<br />

B.3 Interactions de membrane et cortex : Expériences avec<br />

micropipette et cellules avec blebs<br />

La membrane cellulaire est essentiel pour la cellule. Elle sert d’enveloppe,<br />

définie les frontières, elle permet l’échange de matériaux entre le cytosol et<br />

le milieu extracellulaire et elle contient des protéines qui agissent comme<br />

senseurs des signaux externes qui permettront à la cellule de réagir et d’adapter<br />

son comportement en réponse à un changements d’environnement. Sous la<br />

membrane, il y a le cortex, formé par un réseau d’actine que se réorganise<br />

constamment. Le cortex influence la forme de la cellule, son déplacement<br />

et le transport intra-cellulaire. La membrane et le cortex sont liés par des<br />

interactions faibles entre les lipides et les protéines du cortex, et par des interactions<br />

fortes entre protéines. La perte d’adhérence a comme conséquence


134 B Résumé en Français<br />

la formation de blebs, qui sont des protubérances sphériques formées quand<br />

un morceau de membrane se détache du cortex. La présence de blebs est souvent<br />

l’indication de la mort cellulaire, bien que certaines cellules utilisent ce<br />

phénomène pour se déplacer. Dans cette partie de la thèse, nous avons étudié<br />

les aspects statiques et dynamiques du détachement membranaire en réponse<br />

à des perturbations mécaniques.<br />

Description théorique<br />

Pour modéliser l’adhérence entre la membrane et le cortex, nous avons relié<br />

la cinétique d’attachement et de détachement d’une liaison entre membrane<br />

et cortex à la contrainte mécanique exercée par les forces de pression et la<br />

tension du cortex. Dans ce modèle, la membrane plasmique joue un rôle stabilisateur<br />

en compensant une partie des forces de pression. Les différents sites<br />

de liaison agissent de concert pour retenir la membrane, et lorsque l’un des<br />

liens cède, les autres sont soumis à une force plus élevée et ont plus de chance<br />

de céder. Nous avons montré que ce phénomène conduit à une rupture globale<br />

de tous les liens, et au détachement de la membrane, au-delà d’une pression<br />

critique.<br />

Nous avons également étudié les aspects énergétiques de la formation<br />

de bleb. En comparant l’énergie nécessaire pour décrocher la membrane<br />

à l’énergie gagnée lorsqu’un bleb gonfle sous l’effet de la pression, nous<br />

obtenons, comme en le cas antérieur, une instabilité où la membrane se<br />

décroche pour une pression critique. La formation d’un bleb est un processus<br />

de nucléation, caractérisé par une barrière d’énergie qui diminue quand la<br />

pression augmente. On peut définir un temps caractéristique de nucléation qui<br />

dépend de façon exponentielle de la barrière énergétique. Une fois formée, les<br />

blebs se rétractent et disparaissent après un temps caractéristique nécessaire<br />

à la formation d’un nouveau cortex sous la membrane décrochée. L’équilibre<br />

entre formation et rétraction des blebs donne le nombre moyen de bleb à la<br />

surface de la cellule. Ce nombre augmente avec la pression interne, et diminue<br />

avec la tension membranaire et l’intensité de l’adhésion. Lorsque des blebs<br />

sont présent, la pression interne et la tension membranaire sont contrôlés par<br />

l’énergie d’adhésion entre membrane et cortex.<br />

Nous avons ensuite appliqué notre étude à des expériences de micropipette,<br />

lors desquelles une cellule est localement aspirée dans une micropipette. La<br />

cellule peut répondre de plusieurs manières à cette perturbation, y compris<br />

en formant un bleb à l’intérieur de la pipette. Nous avons montré que la<br />

réponse de la cellule dépendait de façon critique de la réponse dynamique<br />

du cytosquelette. Il est expérimentalement établi qu’une cellule soumise à


B.4 Conclusions 135<br />

une perturbation mécanique se comporte de manière élastique à temps court,<br />

et visqueuse à temps long. Nous avons utilisé un modèle très simple de fluide<br />

de Maxwell pour modéliser le comportement viscoélastique du cortex de la<br />

cellule. Soumise à une dépression locale, la cellule se déforme tout d’abord<br />

de manière élastique, puis commence à couler comme un fluide visqueux. La<br />

transition entre ces deux comportements correspond à la contrainte élastique<br />

maximum dans le cortex et sur les liens avec la membrane. Plus la perturbation<br />

est rapide, plus la contrainte est élevée. Cette contrainte doit ensuite<br />

être comparée avec la force critique à partir de laquelle les liaisons se cassent<br />

pour savoir si la membrane va se détacher du cortex. La formation d’un bleb<br />

dépend donc autant de la vitesse avec laquelle une perturbation est appliquée<br />

que de l’intensité de cette perturbation. Une fois un bleb formé, il peut être<br />

stabilisé par une augmentation globale de la tension membranaire. Un nouveau<br />

cortex peut alors se former, et le bleb se rétracte. A ce point, nous avons<br />

déterminé les conditions dans lesquelles cette rétraction peut être presque totale,<br />

et les conditions pour qu’un nouvel épisode de détachement se produise<br />

lors de la rétraction, donnant ainsi lieu à des oscillations périodiques de la cellule<br />

dans la pipette. Nous avons également déterminé dans quelles conditions<br />

un cortex polymérisant sous une membrane en mouvement est en mesure de<br />

freiner l’avancée de la membrane.<br />

Nous avons finalement obtenu un “diagramme de phase” complet du mouvement<br />

d’une cellule sous aspiration locale, qui inclue aspiration complète,<br />

aspiration suivie d’une rétraction complète ou partielle, oscillations, et mouvement<br />

saltatoire.<br />

Conjointement avec un groupe expérimental de l’Institut Curie, nous<br />

avons vérifié expérimentalement les régimes prédits théoriquement et nous<br />

montré que notre modèle reproduit quantitativement les oscillations observés<br />

dans le système de l’amibe Entamoeba histolyca.<br />

B.4 Conclusions<br />

Nous avons exhaustivement étudié le phénomène d’adhésion entre la membrane<br />

et le cortex et la stabilité de l’adhésion lorsque la cellule est soumise à<br />

une perturbation de pression. Nous avons employé une description cinétique<br />

et une description énergétique de l’adhésion. Dans les deux cas, nous avons<br />

obtenu une instabilité globale ou les liaisons se cassent au-delà d’une force<br />

critique. Nous avons décrit le nombre moyen de blebs sur une cellule en fonction<br />

de la pression.


136 B Résumé en Français<br />

Nous avons considéré la réponse dynamique de la cellule à des perturbations<br />

mécaniques. En particulier, nous avons montré que le caractère<br />

viscoélastique de la cellule affecte fortement la stabilité de la membrane.<br />

Nous avons également identifié les différents régimes dynamiques possibles<br />

lors de l’aspiration d’une cellule par une micropipette, que incluent : une<br />

déformation où la membrane et le cortex restent attachés, une déformation<br />

où la membrane se décroche du cortex et est ultérieurement rétractée par<br />

un nouveau cortex, ainsi qu’une déformation où la membrane effectue des<br />

mouvements périodiques. En utilisant des résultats de la littérature et des<br />

expériences réalisées durant cette thèse, nous avons pu vérifier nos prédictions<br />

théoriques. Ce travail était fait en collaboration avec l’étudiant de thèse B.<br />

Mauguis et le docteur F. Amblart.


References<br />

Alberts, B., Johnson, A., Lewis, J., Raff, M., Roberts, K., and Walter, P.<br />

(2002). Molecular Biology of the Cell, 4th Edition. Garland Science.<br />

Allen, M. (1996). Molecular dynamics calculation of elastic constants in<br />

gay–berne nematic liquid crystals. The Journal of chemical physics.<br />

Badoual, M., Juelicher, F., and Prost, J. (2002). Bidirectional cooperative<br />

motion of molecular motors. Proc Natl Acad Sci USA, 99(10):6696–<br />

6701.<br />

Blanc, C., Svenˇsek, D., ˇZumer, S., and Nobili, M. (2005). Dynamics of nematic<br />

liquid crystal disclinations: The role of the backflow. Phys Rev<br />

Lett, 95(9):97802.<br />

Brochard-Wyart, F., Borghi, N., Cuvelier, D., and Nassoy, P. (2006). Hydrodynamic<br />

narrowing of tubes extruded from cells. Proceedings of the<br />

National Academy of Sciences, 103(20):7660.<br />

Browne, W. R. and Feringa, B. L. (2006). Making molecular machines work.<br />

Nature nanotechnology, 1(1):25–35.<br />

Brugues, J. and Casademunt, J. (2008). In preparation.<br />

Brugues, J., Ignés-Mullol, J., Casademunt, J., and Sagues, F. (2008a). Probing<br />

elastic anisotropy from defect dynamics in langmuir monolayers. Phys<br />

Rev Lett, 1:037801.<br />

Brugues, J., Mauguis, B., Amblart, F., and Sens, P. (2008b). In preparation.<br />

Burriel, P., Ignés-Mullol, J., Reigada, R., and Sagués, F. (2006). Collective<br />

molecular precession induced by rotating illumination in photosensitive<br />

langmuir monolayers. Langmuir : the ACS journal of surfaces and colloids,<br />

22(1):187–93.<br />

Büttiker, M. (1987). Z. Phyis. B, 68:161.


138 REFERENCES<br />

Campas, O., Kafri, Y., Zeldovich, K. B., Casademunt, J., and Joanny, J.-F.<br />

(2006). Collective dynamics of interacting molecular motors. Phys Rev<br />

Lett, 97(3):038101.<br />

Campàs, O., Leduc, C., Bassereau, P., Casademunt, J., Joanny, J.-F., and<br />

Prost, J. (2008). Coordination of kinesin motors pulling on fluid membranes.<br />

Biophysical Journal, 94(12):5009–17.<br />

Carlier, M. F., Laurent, V., Santolini, J., Melki, R., Didry, D., Xia, G. X.,<br />

Hong, Y., Chua, N. H., and Pantaloni, D. (1997). Actin depolymerizing<br />

factor (adf/cofilin) enhances the rate of filament turnover: implication in<br />

actin-based motility. J Cell Biol, 136(6):1307–22.<br />

Charras, G. and Paluch, E. (2008). Blebs lead the way: how to migrate without<br />

lamellipodia. Nat Rev Mol Cell Biol, page 7.<br />

Charras, G. T., Coughlin, M., Mitchison, T. J., and Mahadevan, L. (2008).<br />

Life and times of a cellular bleb. Biophys J, 94(5):1836–53.<br />

Charras, G. T., Hu, C.-K., Coughlin, M., and Mitchison, T. J. (2006).<br />

Reassembly of contractile actin cortex in cell blebs. J Cell Biol,<br />

175(3):477–490.<br />

Coudrier, E., Amblard, F., Zimmer, C., Roux, P., Olivo-Marin, J.-C., Rigothier,<br />

M.-C., and Guillén, N. (2005). Myosin ii and the gal-galnac lectin<br />

play a crucial role in tissue invasion by entamoeba histolytica. Cell Microbiol,<br />

7(1):19–27.<br />

Crusats, J., Albalat, R., Claret, J., Ignés-Mullol, J., Reigada, R., and Sagués,<br />

F. (2005). Photoinduced molecular reorientation dynamics in confined<br />

domains of langmuir monolayers. The Journal of chemical physics,<br />

122(24):244722.<br />

Crusats, J., Albalat, R., Claret, J., Ignés-Mullol, J., and Sagues, F. (2004).<br />

Influence of temperature and composition on the mesoscopic textures of<br />

azobenzene langmuir monolayers. Langmuir, 20(20):8668–8674.<br />

de Gennes, P. G. and Prost, J. (1995). The Physics of Liquid Crystals 2nd ed.<br />

Oxford University Press.<br />

Derényi, I. and Ajdari, A. (1996). Collective transport of particles in a ”flashing”<br />

periodic potential. Physical review E, Statistical physics, plasmas,<br />

fluids, and related interdisciplinary topics, 54(1):R5–R8.<br />

Dietrich, C., Bagatolli, L., Volovyk, Z., Thompson, N., Levi, M., Jacobson,<br />

K., and Gratton, E. (2001). Lipid rafts reconstituted in model membranes.<br />

Biophysical Journal, 80(3):1417.<br />

Endow, S. and Higuchi, H. (2000). A mutant of the motor protein kinesin that<br />

moves in both directions on microtubules. Nature.


REFERENCES 139<br />

Evans, E. (2001). Probing the relation between force - lifetime - and chemistry<br />

in single molecular bonds. Annual review of biophysics and<br />

biomolecular structure, 30:105–128.<br />

Evans, E. and Rawicz, W. (1990). Entropy-driven tension and bending elasticity<br />

in condensed-fluid membranes. Phys Rev Lett.<br />

Feder, A., Tabe, Y., and Mazur, E. (1997). Orientational fluctuations in a<br />

two-dimensional smectic-c liquid crystal with variable density. Phys<br />

Rev Lett, 79(9):1682–1685.<br />

Fernández, P., Pullarkat, P. A., and Ott, A. (2006). A master relation defines<br />

the nonlinear viscoelasticity of single fibroblasts. Biophysical Journal,<br />

90(10):3796–805.<br />

Feynman, R. P., Leighton, R. B., and Sands, M. (1966). The Feynman Lectures<br />

on Physics. Addison-Wesley, Reading, MA), Vol. I, Chap. 46.<br />

Fournier, J., Ajdari, A., and Peliti, L. (2001). Effective-area elasticity and<br />

tension of micromanipulated membranes. Phys Rev Lett.<br />

Garrivier, D., Decave, E., Brechet, Y., Bruckert, F., and Fourcade, B. (2002).<br />

Peeling model for cell detachment. Eur Phys J E, 8(1):79–97.<br />

Gerbal, F., Noireaux, V., Sykes, C., Jülicher, F., Chaikin, P., Ott, A., Prost, J.,<br />

Golsteyn, R. M., Friederich, E., Louvard, D., Laurent, V., and Carlier,<br />

M. F. (1999). On the ‘listeria’propulsion mechanism. Pramana.<br />

Hatta, E. and Fischer, T. (2003). Splitting of an s= 1 point disclination into<br />

half-integer disclinations upon laser heating of a . . . . J. Phys. Chem. B.<br />

Head, D. A., Levine, A. J., and MacKintosh, F. C. (2003a). Deformation<br />

of cross-linked semiflexible polymer networks. Phys Rev Lett,<br />

91(10):108102.<br />

Head, D. A., Levine, A. J., and MacKintosh, F. C. (2003b). Distinct regimes<br />

of elastic response and deformation modes of cross-linked cytoskeletal<br />

and semiflexible polymer networks. Physical review E, Statistical, nonlinear,<br />

and soft matter physics, 68(6 Pt 1):061907.<br />

Head, D. A., Levine, A. J., and MacKintosh, F. C. (2005). Mechanical response<br />

of semiflexible networks to localized perturbations. Physical review<br />

E, Statistical, nonlinear, and soft matter physics, 72(6 Pt 1):061914.<br />

Head, D. A., MacKintosh, F. C., and Levine, A. J. (2003c). Nonuniversality<br />

of elastic exponents in random bond-bending networks. Physical review<br />

E, Statistical, nonlinear, and soft matter physics, 68(2 Pt 2):025101.<br />

Helfrich, W. and Servuss, R. (1984). Undulations, steric interaction and cohesion<br />

of fluid membranes. Nuovo Cimento D, 3(1):137–151.<br />

Hinshaw, G., Petschek, R., and Pelcovits, R. (1988). Modulated phases in<br />

thin ferroelectric liquid-crystal films. Phys Rev Lett, 60(18):1864–1867.


140 REFERENCES<br />

Hochmuth, F. M., Shao, J. Y., Dai, J., and Sheetz, M. P. (1996). Deformation<br />

and flow of membrane into tethers extracted from neuronal growth<br />

cones. Biophys J, 70(1):358–69.<br />

Hochmuth, R., Wiles, H., Evans, E., and Mccown, J. (1982). Extensional flow<br />

of erythrocyte membrane from cell body to elastic tether. ii. experiment.<br />

Biophysical Journal, 39(1):83.<br />

Ignés-Mullol, J., Claret, J., Albalat, R., Crusats, J., Reigada, R., Romero, M.<br />

T. M., and Sagués, F. (2005). Texture changes inside smectic-c droplets<br />

in azobenzene langmuir monolayers. Langmuir : the ACS journal of<br />

surfaces and colloids, 21(7):2948–55.<br />

Ignés-Mullol, J., Claret, J., and Sagués, F. (2004). Star defects, boojums,<br />

and cardioid droplet shapes in condensed dimyristoylphosphatidylethanolamine<br />

monolayers. J. Phys. Chem. B, 108(2):612–619.<br />

Ingber, D. E. (1993). Cellular tensegrity: defining new rules of biological<br />

design that govern the cytoskeleton. J Cell Sci, 104 ( Pt 3):613–27.<br />

Joanny, J.-F., Juelicher, F., Kruse, K., and Prost, J. (2007). Hydrodynamic<br />

theory for multi-component active polar gels. New J Phys, 9:–.<br />

Juelicher, F., Kruse, K., Prost, J., and Joanny, J.-F. (2007). Active behavior of<br />

the cytoskeleton. Phys Rep, 449:3–28.<br />

Juelicher, F. and Prost, J. (1995). Cooperative molecular motors. Phys Rev<br />

Lett, 75(13):2618–2621.<br />

Julicher, F., Ajdari, A., and Prost, J. (1997). Modeling molecular motors. Rev.<br />

Mod. Phys., 69(4):1269–1281.<br />

Kampen, N. V. (2007). Stochastic processes in physics and chemistry.<br />

books.google.com.<br />

Kleman, M. and Laverntovich, O. D. (2003). Soft Matter Physics: An Introduction.<br />

Springer.<br />

Klumpp, S. and Lipowsky, R. (2005). Cooperative cargo transport by several<br />

molecular motors. Proceedings of the National Academy of Sciences.<br />

Klumpp, S., Muller, M., and Lipowsky, R. (2006). Cooperative transport by<br />

small teams of molecular motors. Arxiv preprint q-bio.SC.<br />

Koster, G., Vanduijn, M., Hofs, B., and Dogterom, M. (2003). From the cover:<br />

Membrane tube formation from giant vesicles by dynamic association<br />

of motor proteins. Proceedings of the National Academy of Sciences,<br />

100(26):15583.<br />

Kramers, H. (1940). Brownian motion in a field of force and the diffusion<br />

model of chemical reactions. Physica.


REFERENCES 141<br />

Kruse, K., Joanny, J.-F., Juelicher, F., Prost, J., and Sekimoto, K. (2004).<br />

Asters, vortices, and rotating spirals in active gels of polar filaments.<br />

Phys Rev Lett, 92:078101.<br />

Landau, L. D. and Lifshitz, E. M. (1987). Fluid mechanics. page 552.<br />

Landau, L. D., Lifshitz, E. M., Kosevich, A. M., and Pitaevskii, L. P. (1995).<br />

Theory of Elasticity 3rd Edition. Butterworth-Heinemann.<br />

Landauer, R. (1988). J. Stat. Phys., 53:233.<br />

Lanni, F. and Ware, B. R. (1984). Detection and characterization of actin<br />

monomers, oligomers, and filaments in solution by measurement of fluorescence<br />

photobleaching recovery. Biophysical Journal, 46(1):97.<br />

Leduc, C., Campas, O., Zeldovich, K., Roux, A., Jolimaitre, P., Bourel-<br />

Bonnet, L., Goud, B., Joanny, J.-F., Bassereau, P., and Prost, J. (2004).<br />

From the cover: Cooperative extraction of membrane nanotubes by<br />

molecular motors. Proceedings of the National Academy of Sciences,<br />

101(49):17096.<br />

Leibler, S. and Huse, D. (1993). Porters versus rowers - a unified stochasticmodel<br />

of motor proteins. J Cell Biol, 121(6):1357–1368.<br />

Merkel, R., Simson, R., Simson, D., Hohenadl, M., Boulbitch, A., Wallraff,<br />

E., and Sackmann, E. (2000). A micromechanic study of cell polarity<br />

and plasma membrane cell body coupling in dictyostelium. Biophys J,<br />

79:707–719.<br />

Mermin, N. D. (1979). Rev. Mod. Phys., 51:591–648.<br />

Mogilner, A. and Oster, G. (2003). Force generation by actin polymerization<br />

ii: the elastic ratchet and tethered filaments. Biophysical Journal,<br />

84(3):1591–605.<br />

Nishinari, Okada, Y., Schadschneider, A., and Chowdhury, D. (2005). Intracellular<br />

transport of single-headed molecular motors kif1a. Phys Rev<br />

Lett.<br />

Noireaux, V., Golsteyn, R. M., Friederich, E., Prost, J., Antony, C., Louvard,<br />

D., and Sykes, C. (2000). Growing an actin gel on spherical surfaces.<br />

Biophysical Journal, 78(3):1643–54.<br />

Okada, Y. and Hirokawa, N. (1999). A processive single-headed motor: Kinesin<br />

superfamily protein kif1a. Science.<br />

Oswald, P. and Ignés-Mullol, J. (2005). Backflow-induced asymmetric collapse<br />

of disclination lines in liquid crystals. Phys Rev Lett, 95(2):27801.<br />

Parmeggiani, A., Juelicher, F., Ajdari, A., and Prost, J. (1999). Energy transduction<br />

of isothermal ratchets: Generic aspects and specific examples<br />

close to and far from equilibrium. Phys Rev E, 60(2):2127–2140.


142 REFERENCES<br />

Pécréaux, J., Döbereiner, H.-G., Prost, J., Joanny, J.-F., and Bassereau, P.<br />

(2004). Refined contour analysis of giant unilamellar vesicles. The European<br />

physical journal E, Soft matter, 13(3):277–90.<br />

Pedrosa, J., Romero, M., and Camacho, L. (2002). Organization of an amphiphilic<br />

azobenzene derivative in monolayers at the air-water interface.<br />

J. Phys. Chem. B.<br />

Plastino, J., Lelidis, I., Prost, J., and Sykes, C. (2004). The effect of diffusion,<br />

depolymerization and nucleation promoting factors on actin gel growth.<br />

European Biophysics Journal.<br />

Pullarkat, P., Fernandez, P., and Ott, A. (2007). Rheological properties of the<br />

eukaryotic cell cytoskeleton. Physics Reports, 449(1-3):29–53.<br />

Rasband, W. S. (1997-2006). Imagej. U.S. N.I.H, Bethesda, Maryland.,<br />

http://rsb.info.nih.gov/ij/.<br />

Raucher, D., Stauffer, T., Chen, W., Shen, K., Guo, S., York, J., Sheetz, M.,<br />

and Meyer, T. (2000). Phosphatidylinositol 4,5-bisphoshate functions as<br />

a second messenger that regulates cytoskeleton-plasma membrane adhesion.<br />

Cell, 100(2):221–228.<br />

Reigada, R., Mikhailov, A. S., and Sagués, F. (2004). Nonequilibrium orientational<br />

patterns in two-component langmuir monolayers. Physical review<br />

E, Statistical, nonlinear, and soft matter physics, 69(4 Pt 1):041103.<br />

Rentsch, P. and Keller, H. (2000). Suction pressure can induce uncoupling of<br />

the plasma membrane from cortical actin. Eur. J. Cell Biol.<br />

Rosenblatt, C., Pindak, R., Clark, N., and Meyer, R. (1979). Freely suspended<br />

ferroelectric liquid-crystal films: Absolute measurements of polarization,<br />

elastic constants, and viscosities. Phys Rev Lett, 42(18):1220–1223.<br />

Ryskin, G. and Kremenetsky, M. (1991). Drag force on a line defect moving<br />

through an otherwise undisturbed field: Disclination line in a nematic<br />

liquid crystal. Phys Rev Lett, 67(12):1574–1577.<br />

Sancho, J., Miguel, M., Katz, S., and Gunton, J. (1982). Analytical and numerical<br />

studies of multiplicative noise. Phys Rev A, 26(3):1589–1609.<br />

Sekimoto, K., Prost, J., Juelicher, F., Boukellal, H., and Bernheim-<br />

Grosswasser, A. (2004). Role of tensile stress in actin gels and a<br />

symmetry-breaking instability. Eur Phys J E, 13(3):247–259.<br />

Sheetz, M. P. (2001). Cell control by membrane-cytoskeleton adhesion. Nat<br />

Rev Mol Cell Biol, 2(5):392–6.<br />

Sheetz, M. P., Sable, J. E., and Döbereiner, H.-G. (2006). Continuous<br />

membrane-cytoskeleton adhesion requires continuous accommodation<br />

to lipid and cytoskeleton dynamics. Annual review of biophysics and<br />

biomolecular structure, 35:417–34.


REFERENCES 143<br />

Sit, P., Spector, A., Lue, A., Popel, A., and Brownell, W. (1997). Micropipette<br />

aspiration on the outer hair cell lateral wall. Biophys J, 72(6):2812–<br />

2819.<br />

Stone, H. (1995). Fluid motion of monomolecular films in a channel flow<br />

geometry. Physics of Fluids.<br />

Svenˇsek, D. and ˇZumer, S. (2002). Hydrodynamics of pair-annihilating<br />

disclination lines in nematic liquid crystals. Phys Rev E, 66(2):21712.<br />

Tabe, Y., Shen, N., Mazur, E., and Yokoyama, H. (1999). Simultaneous observation<br />

of molecular tilt and azimuthal angle distributions in spontaneously<br />

modulated liquid-crystalline langmuir monolayers. Phys Rev<br />

Lett, 82(4):759–762.<br />

Tabe, Y. and Yokoyama, H. (1994). Novel liquid crystalline structure in langmuir<br />

monolayers of amphiphilic azobenzene derivative. Journal of the<br />

Physical Society of Japan.<br />

Thoumine, O. and Ott, A. (1997). Time scale dependent viscoelastic and<br />

contractile regimes in fibroblasts probed by microplate manipulation. J<br />

Cell Sci, 110 ( Pt 17):2109–16.<br />

Tóth, G., Denniston, C., and Yeomans, J. (2002). Hydrodynamics of topological<br />

defects in nematic liquid crystals. Phys Rev Lett, 88(10):105504.<br />

Trepat, X., Deng, L., An, S. S., Navajas, D., Tschumperlin, D. J., Gerthoffer,<br />

W. T., Butler, J. P., and Fredberg, J. J. (2007). Universal physical<br />

responses to stretch in the living cell. Nature, 447(7144):592–+.<br />

van Oudenaarden, A. and Theriot, J. A. (1999). Cooperative symmetrybreaking<br />

by actin polymerization in a model for cell motility. Nat Cell<br />

Biol, 1(8):493–9.<br />

Vilfan, A., Frey, E., and Schwabl, F. (1998). Elastically coupled molecular<br />

motors. Eur Phys J B, 3(4):535–546.<br />

Wilhelm, J. and Frey, E. (2003). Elasticity of stiff polymer networks. Phys<br />

Rev Lett, 91(10):108103.<br />

Wottawah, F., Schinkinger, S., Lincoln, B., Ananthakrishnan, R., Romeyke,<br />

M., Guck, J., and Käs, J. (2005). Optical rheology of biological cells.<br />

Phys Rev Lett, 94(9):098103.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!