131
AIX-MARSEILLE UNIVERSITE FACULTE DES SCIENCES DE LUMINY PH.D THESIS Presented for defence on 20 th January 2015 by KAPIL BAMBARDEKAR to obtain the degree of Doctor of Philosophy (Ph.D.) of the Aix-Marseille Université Specialization : Biophysics APPLYING OPTICAL TWEEZERS IN VIVO: A BIOPHYSICAL STUDY OF MECHANICAL FORCES IN DROSOPHILA MELANOGASTER AT THE ONSET OF GASTRULATION Thesis directed by PIERRE-FRANCOIS LENNE in the Insitute of Developmental Biology Marseille Luminy (IBDML) Referees: Emmanuel Courtade Sylvie Dufour Examiners: Hugues Giovanni Atef Asnacios Director: PIERRE FRANCOIS LENNE

applying optical tweezers in vivo: a biophysical study of mechanical

  • Upload
    vantram

  • View
    215

  • Download
    0

Embed Size (px)

Citation preview

Page 1: applying optical tweezers in vivo: a biophysical study of mechanical

AIX-MARSEILLE UNIVERSITE

FACULTE DES SCIENCES DE LUMINY

PH.D THESIS

Presented for defence on 20th January 2015 by

KAPIL BAMBARDEKAR

to obtain the degree of

Doctor of Philosophy (Ph.D.)

of the Aix-Marseille Université Specialization : Biophysics

APPLYING OPTICAL TWEEZERS IN VIVO:

A BIOPHYSICAL STUDY OF MECHANICAL

FORCES IN DROSOPHILA MELANOGASTER

AT THE ONSET OF GASTRULATION

Thesis directed by

PIERRE-FRANCOIS LENNE

in the Insitute of Developmental Biology Marseille Luminy (IBDML)

Referees: Emmanuel Courtade

Sylvie Dufour

Examiners: Hugues Giovanni

Atef Asnacios

Director: PIERRE FRANCOIS LENNE

Page 2: applying optical tweezers in vivo: a biophysical study of mechanical
Page 3: applying optical tweezers in vivo: a biophysical study of mechanical

SUMMARY

The goal of this thesis was to understand the nature and polarization of mechanical forces in

morphogenesis, for which the Drosophila embryonic epithelium was chosen as a model system. The

mechanical properties of cells and tissues have long been thought to play a crucial role in morphogenesis .

However there is a paucity of tools available to quantify and accurately measure mechanical forces in

vivo. Of all the approaches to understand mechanical forces, optical tweezers have shown the most

promise. Here, an optical tweezers setup was developed on a pre-existing single-plane illumination

(SPIM) setup. It was observed for the first time, that the cell-cell interface in embryonic epithelia could be

trapped and manipulated directly with optical tweezers. The interaction of the interface with the trap was

initially characterized at the end of cellularization where the tissue has minimal movements and acto-

myosin turnover. With a sinusoidal trap excursion, the interface amplitude was found to increase linearly

with applied laser power as well as trap amplitude and time period. To correlate the interface movement

to its tension, the trap stiffness was estimated in vivo with injected polystyrene beads. Furthermore, push

and pull experiments on the interface responding to a stationary trap, provided another way to address the

viscoelastic properties of the interface. The interface kinetics in stationary experiments could fit

adequately to a passive viscoelastic model. This model also explained well the linear response to trap

amplitude and time period, and formed the basis of estimating interface tension from its amplitude.

Moreover, the propagation of the sinusoidal movement to neighbouring interfaces decayed rapidly with

minimal phase lag in both experiments and the model. Having established a suitable regime of trapping

conditions, where interface deflection is small and linear, the mechanical anisotropy of the epithelium was

at the onset of gastrulation (early germband elongation). The interface tension increased by 2-3 fold

compared to end of cellularization, exhibiting both apico-basal and dorso-ventral polarization of tension,

concomitant with polarized accumulation of myosin. The role of myosin was established further through

ROCK-inhibition, which restored the tension. Perturbation of actin also decreased the interface tension,

additionally displaying creep-like behaviour. My work provides a crucial insight into the mechanical

behaviour of dynamic epithelia, as well as developing further the field of in vivo optical manipulation.

Page 4: applying optical tweezers in vivo: a biophysical study of mechanical

La mécanique des cellules et des tissus joue un rôle crucial durant la morphogenèse. Cependant, les outils disponibles pour mesurer les forces in vivo sont très rares. Parmi les approches potentielles, nous avons retenu les pinces optiques qui ont été largement utilisées sur des systèmes in vitro, molécules individuelles ou cellules isolées. Nous avons développé un dispositif combinant pinces optiques et imagerie par feuillet de lumière. Nous montrons que les interfaces cellulaires de l’épithélium précoce de l’embryon de Drosophile peuvent être piégées et manipulées directement avec des pinces optiques. Dans un premier temps, la manipulation optique est réalisée à la fin de la cellularisation, processus par lequel des membranes cellulaires séparent les noyaux pour donner naissance à un épithélium ; à ce stade, les mouvements cellulaires sont minimes et les cellules ont des formes hexagonales similaires. En imposant un mouvement sinusoïdal au piège perpendiculairement à une interface, nous étudions la déflection de l’interface en fonction de la puissance laser, de l’amplitude du mouvement du piège et de la fréquence d’oscillation. La réponse est linéaire pour les petites déformations (amplitude de déflection de l’interface inférieure à 500 nm). Afin d’extraire des valeurs de tension, nous estimons la raideur du piège en comparant la déflection imposée directement par le laser à celle produite par des billes individuelles piégées forçant le déplacement de l’interface. Les tensions mesurées sont de l’ordre de quelques dizaines à quelques centaines de pN. En outre, des expériences de déflection–relaxation par déplacement instantané puis arrêt du piégeage, ont été réalisées, fournissant une alternative à l’analyse fréquentielle pour étudier les propriétés viscoélastiques de l’interface. Un modèle de type solide linéaire standard rend compte des observations et permet d’extraire les paramètres viscoélastiques de l’interface. Nous mettons également en évidence que la déflection imposée à une interface se propage aux interfaces voisines en s’affaiblissant exponentiellement sur une distance d’une à deux cellules. Cette technique étant établie, nous l’utilisons pour mesurer les tensions durant l’extension de la bandelette germinale. Les tensions sont anisotropes, les jonctions parallèles à la direction dorsoventrale ayant une tension trois fois plus élevée que celles perpendiculaires. Nous mesurons également des tensions aux interfaces cellulaires plus grandes dans le plan des jonctions adhérentes que dans les plans plus basaux de l’épithélium. Ces tensions sont significativement réduites par inhibition de l’activité du moteur moléculaire Myosine-II. Ce travail fournit pour la première fois des mesures absolues des tensions intercellulaire et un outil pour l’étude quantitative de la mécanique épithéliale in vivo.

Page 5: applying optical tweezers in vivo: a biophysical study of mechanical

CONTENTS

CHAPTER 1 : INTRODUCTION ............................................................. 1

1.1 OVERVIEW.………...………………………………………….......... 3

1.2 ORGANIZATION OF THE TISSUE AND UNDERLYING MECHANICAL ELEMENTS

………………………………………………………….……………….. 5

1.3 CURRENT MECHANICAL UNDERSTANDING OF SINGLE CELLS AND TISSUES IN VITRO ………..……………………………….…………………………..9

1.4 MORPHOGENESIS: TISSUE DYNAMICS DURING EMBRYONIC DEVELOPMENT

…………………………………………..……….................................... 13

1.5 MECHANICAL CHANGES UNDERLYING MORPHOGENETIC MOVEMENTS .... 17

1.6 IN VIVO TOOLS FOR DIRECT UNDERSTANDING OF FORCES ……….......... 21

1.7 GERMBAND ELONGATION IN DROSOPHILA MELANOGASTER AS A MODEL

SYSTEM FOR UNDERSTANDING THE MECHANICAL FORCES IN MORPHOGENESIS

……........................................................................................................... 25

1.8 SUMMARY………………………………………............................. 27

CHAPTER 2 : MATERIALS AND METHODS………........................ 29

2.1 MATERIALS………………………………………………………... 30

2.2 SAMPLE PREPARATION…………………………………………. 30

2.3 EXPERIMENTAL SETUP………………………………………….. 31

2.4 OPTICAL TWEEZERS EXPERIMENTS…………………………... 32

2.5 DATA ANALYSIS………………………………………………….. 34

2.6 EXPERIMENTAL LIMITATIONS…………………………………. 34

2.7 QUANTIFICATION OF E-CADHERIN, MYOSIN II AND LIFEACT.36

CHAPTER 3 : ESTABLISHING OPTICAL TWEEZERS AS AN IN VIVO TOOL TO UNDERSTAND CELL-CELL INTERFACE

MECHANICS ....................................................................................... 37

3.1 MOTIVATION……………………………………………………... 39

3.2 RESULTS………………………………………………………….. 41

3.2.1 Characterizing the deflection of cell-cell interfaces imposed by

optical tweezers………………………………………………….…... 41

3.2.2 Dynamics of cell-cell interface interaction with optical tweezers

Page 6: applying optical tweezers in vivo: a biophysical study of mechanical

as a measure of position offset and phase lag………………………. 45

3.2.3 Relaxation of interface deformation ..…………………….…… 47

3.2.4 Propagation of interface deformation ..…………………….…… 47

3.3 DISCUSSIONS……………………………………………………… 50

3.3.1 Origin and nature of in vivo optical forces…………………… 50

3.3.2 Towards a mechanistic understanding of the experiments….. 53

3.3.3 Developing a mechanical model…………………………….. 55

3.4 SUMMARY….……………………………………………………… 58

CHAPTER 4 : NATURE AND ORIGIN OF MECHANICAL FORCES

IN DROSOPHILA GERMBAND ELONGATION ................................ 59

4.1 MOTIVATION……………………………………………………… 61

4.2 RESULTS…………………………………………………………… 63

4.2.1 Probing the anisotropy of mechanical forces in early gastrulation.. 63

4.2.2 Effect of actin perturbation on interface tension…………..….. 68

4.2.3 Temporal and spatial dynamics: Position offset and phase lag

measurements in different stages and perturbations ………………… 72

4.3 DISCUSSIONS……………………………………………………… 74

4.3.1 Interface tension is regulated by myosin-II activity……….… 74

4.3.2 Actin cortex is responsible for interface tension………….. …. 75

4.3.3 Creep and sinuosoidal loading………………..……………….. 76

3.4 SUMMARY….……………………………………………………… 77

5 CONCLUSIONS AND PERSPECTIVES ............................. ……… 79

ACKNOWLEDGEMENTS .....................................................................................I

REFERENCES .......................................................................................................... II

MANUSCRIPT (IN PRESS PNAS)

Page 7: applying optical tweezers in vivo: a biophysical study of mechanical

LIST OF FIGURES

Figure 1.1 Organization of the tissue…………………………………………… 4

Figure 1.2 Mechanical elements in cell and tissue organization………………. 6

Figure 1.3 Terminology and methods to understand cell and tissue

mechanics………………………………………………………………………... 8

Figure 1.4 Mechanical models for cells…………...……………………………10

Figure 1.5 Cell movements and rearrangements in development…………….. 12

Figure 1.6 Actomyosin dynamics in morphogenesis………………………...….14

Figure 1.7 Measuring mechanical forces in morphogenesis…………….……... 16

Figure 1.8 Mechanical models of morphogenesis…………………………..... 18

Figure 1.9 Methods for in vivo force measurements…….…………………... 20 Figure 1.10 Overview of optical tweezers and particle tracking rheology….. 22

Figure 1.11 Germband elongation: molecular and cellular events…..……... 24

Figure 1.12 Mechanical forces during germband elongation.................……... 26

Figure 2.1 Optical setup combining light sheet microscopy and optical

tweezers…………………………………………………………………………. 31

Figure 2.2 Calibration of trap power and displacement……………………... 32

Figure 2.3 Noise in interface position detection…………………………….. 35

Figure 3.1 Characterizing the deflection of cell-cell interfaces imposed

by optical tweezers……………………………………………………………. 40

Figure 3.2 Interface deflection as a function of trap movement

amplitude and time period of trap movement oscillation…………………... 42

Figure 3.3 Offset between laser and interface positions……………………. 44

Figure 3.4 Phase lag between laser and interface movement…………………46

Figure 3.5 Mechanical model of the interface and tissue response …………... 48

Figure 3.6 Interface deflection induced by cytoplasmic trap………………… 49

Figure 3.7 Quantitative Phase Microscopy Image obtained on early

Drosophila embryo ………………………………………………………….…50

Figure 3.8 Comparison of interface deformation with and without beads… 51

Page 8: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 3.9 Viscosity measurements obtained from bead trajectories………… 52

Figure 3.10 Relaxation of interfaces after trap release. Initial speed of

relaxation is related to the friction coefficient………………………………….54

Figure 3.11 Elongation of interfaces adjacent to the optically deformed

interface…………………………………………………………………………55

Figure 4.1 Tension at cell contacts before and during germband

elongation in normal embryos…………………………………………………. 62

Figure 4.2 Polarity of interface tension at different stages for interface

orientation along the dorso-ventral plane……………………………………… 64

Figure 4.3 Apicobasal polarity of interface tension at different stages……... 65

Figure 4.4 Apicobasal polarity in the lateral plane at the end of

cellularization……………………………………………………………………66

Figure 4.5 Effect of myosin II perturbation on interface tension…………… 67

Figure 4.6 Influence of actin intensity at different stages on interface

tension………………………………………………………………………….. 69

Figure 4.7 Effect of actin perturbation on interface tension………………… 70

Figure 4.8 Phase lag and offset in different embryos ………………………… 71

Figure 4.9 Time dependence and creep after cytochalsin treatment…………. 73

Page 9: applying optical tweezers in vivo: a biophysical study of mechanical

LIST OF ABBREVIATIONS

ECM – extracellular matrix

FA – focal adhesions

AFM – atomic force microscopy

FRET – forster resonance energy transfer

AOTF – acousto-optical tunable filter

NA – numerical aperture

SPIM – selective plane illumination microscopy

GUI – graphical user interface

ROCK – rho-associated protein kinase

PSF – point spread function

– interface sinusoidal amplitude

– trap sinusoidal amplitude - trap stiffness

ini_off - initial offset between trap and interface

fin_off – final offset between trap and interface

int_mov - interface movement when trap is switched on

SD – standard deviation of mean

– trapping force

T – interface tension

l0 – initial junction length

– trap position

x – interface position - interface friction coefficient

SLS – standard linear solid

A/P – anterio-posterior

Page 10: applying optical tweezers in vivo: a biophysical study of mechanical

D/V – dorso-ventral

lat A – latrunculin A

cyto D – cytochalasin D

WT – not injected / control

Page 11: applying optical tweezers in vivo: a biophysical study of mechanical

CHAPTER 1

INTRODUCTION

1

Page 12: applying optical tweezers in vivo: a biophysical study of mechanical

2

Page 13: applying optical tweezers in vivo: a biophysical study of mechanical

1.1 Motivation

Over the last few decades, significant progress has been made in understanding single cell mechanics.

Because cells can be isolated and cultured in vivo, biological mechanisms such as cell migration,

adhesion as well cell-division, tissue growth can be studied from a reductionist perspective. Such an

approach has also facilitated the development of techniques to manipulate and probe mechanical forces in

these events. Particularly, this has led to a detailed understanding in the role of the acto-myosin

cytoskeleton in imparting dynamic mechanical properties to the cell which is crucial in developing and

maintaining an equilibrium of forces with the extracellular environment. The rich diversity of mechanical

models that has developed has also helped in extending this knowledge at the tissue level. In recent years

a lot of focus has shifted in understanding these properties in the natural tissue environment. Here I give

an overview of current understanding in the field and explain the background and context of my work.

3

Page 14: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 1.1 Organization of the tissue

a) A schematic of the typical organization of the cytoskeletal filaments inside an eukaryotic cell.

b) Cell–cell and cell–ECM junctions in epithelia. (a Pullarkat et al. 2007, b Rodriguez et al. 2013)

4

Page 15: applying optical tweezers in vivo: a biophysical study of mechanical

1.2 Organization of the tissue and underlying mechanical elements

The typical organization of the eukaryotic cell is illustrated in Fig.1.1a. Briefly it is enclosed in a bilipid

membrane with various organelles embedded in an viscous cytoplasm. The nucleus and organelles carry

out their functions and maintain their organization by means of a cytoskeleton. Different filaments

constitute to the cytoskeleton, mainly actin filaments, microtubules and intermediate filaments which are

distributed throughout the cell in an organized manner (Fig. 1.2a). The actin filaments form a meshwork

(cortex) below the plasma membrane, and they are cross-linked by a variety of proteins, including motor

proteins, which are capable of generating forces and performing mechanical work. The filaments connect

and interact with the plasma membrane via trans-membrane proteins embedded in the lipid bilayer. The

microtubules originate from the centrosome, close to the nucleus, extending to the periphery, at the actin

cortex. Intermediate filaments are concentrated around the nucleus and extend away from it. The actin

network is the most important cytoskeletal entity for mechanical functions like control of cell shape and

cell locomotion. They are capable of rapid reorganization through de-polymerization and polymerization

cycles and can generate active contractile forces with the help of motor proteins (primarily different types

of myosin). The hydrolysis of Adenosine triphosphate (ATP) to Adenosine diphosphate (ADP) provides

continuous energy for the system. A single reaction liberates about 10 kT of free energy per molecule.

(Alberts et al. 2002, Bray et al. 2000, Pullarkat et al. 2007)

Acto-myosin complexes are formed by the association of myosin motors with actin filaments. Myosin-II

self-assembles into short bipolar chains, which act on neighbouring actin filaments to produce relative

motion and active stresses within the actin network. Acto-myosin complexes can be highly organized, for

example in stress fibers or myofibrils, which consist of contractile bundles of actin filaments and myosin

motors. They also exist as a random network of highly cross-linked “active gel” with mesh sizes of ~100

nm. A good example is the actin cortex, which forms a thin layer (∼ 1 µm) attached to the plasma

membrane in eukaryotic cells. The cortex is the main cytoskeletal component responsible for a range of

functions including control of cell shape, generation of active stresses and cell locomotion. (Alberts et al.

2002, Bray et al. 2000) In all of these cases, contractility results from the activity of myosin II motors

associated with bundles of actin filaments. Myosin motors attach and detach from actin filaments in a

cyclic manner and while attached they undergo a conformational change tightly coupled to ATP hydrolysis

that moves the motor along the filament and generates a displacement of the filament. Because of their

non-processivity and low duty ratio, single Myosin II motors spend an appreciable fraction of their time

detached from the filament. As a consequence, a single motor is not sufficient for motility but their

assembly into mini-filaments converts them into a highly processive motor complex. (Gorfinkiel et al.

2011)5

Page 16: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 1.2 Mechanical elements in cell and tissue organization. a) A schematic showing the coupling

between the cytoskeleton and the extracellular matrix. Cell adhesion is mediated by specialised adhesion

proteins which, typically, form complexes called focal adhesions. The actin cytoskeleton is mechanically

coupled to the substrate at the focal adhesions. (Pullarkat et al. 2007) b) Schematic representation of how

the protein composition of focal adhesions (FAs) is re-organized in response to mechanical force. Within

immature FAs, force-insensitive proteins (grey squares), force-sensitive proteins (blue shapes) and force-

responsive proteins (green shapes) coordinately transmit the specific integrin-mediated signals. In response

to mechanical force, focal adhesion abundance of force sensitive proteins (blue shapes) and force-

responsive proteins (green shapes) are decreased, while the abundance of force-sensitive proteins (orange

shapes) and force-responsive proteins (yellow shapes) are increased. (Kuo 2013). c) Schematic of E-

cadherin adherens junction in epithelial cells. The extracellular domains of E-cadherin homodimers enter

homotypic Ca2+-dependent-binding interactions with those of dimers on adjacent cells. The intracellular

domain of E-cadherin interacts with the actin cytoskeleton via α-catenin and either β- or γ-catenin.

Cadherin–catenin complexes constitute the adherens junction. (Perry et al. 2010) d) Cadherin–actin

interaction through a-catenin .The structure and functional domains of a-catenin in an ‘open’ form,

showing b-catenin, vinculin and F-actin (actin filament) binding regions. The central domain can mask the

vinculin-binding site and can unmask this site when the C-terminus is pulled by acto-myosin forces. Below

- Simplified classical model of F-actin linkage at the AJ. (Yonemura 2011)

6

Page 17: applying optical tweezers in vivo: a biophysical study of mechanical

Cells organize together to form tissues in vivo (Fig. 1.1b, illustrates an example of tissue (endothelium)

with various organizational elements.). Within tissues (as well as in monolayers cultured on artificial

substrates) cells exist in close association with an extracellular matrix (Fig. 1.2a,b). Several cell types

cannot proceed through normal cell division cycle if not attached to this matrix by trans-membrane

proteins belonging to the class of integrins which form mechanical connections with bundles of actin

filaments to form stress fibers. Focal adhesions are sensitive to cell surface chemistry and are promoted

by proteins like fibronectin, collagen etc. Focal adhesions anchor the cell and its actin skeleton to

the substrate and also perform important signaling roles. Interactions between focal adhesion proteins

and extracellular matrix can regulate the cytoskeletal structure as well as active contractile state of the

cell. (Burridge et al. 1996, Zhu et al. 2000)

In addition cells associate with each other through different types of cell- cell junctions. I will focus here

on cell-adherens junctions which regulate adhesion and force equilibrium between cells in a tissue

(Ladoux et al. 2000). Adherens junctions consist of cadherin-class trans-membrane proteins that associate

through homophilic adhesions to form clusters that adhere cells together (Fig. 1.2c, Perry et al. 2010).

Based on ultrastructural observations, the AJs are characterized as a region at the interface of two adjacent

cells with opposing membranes typically ~20 nm apart, with an intercellular space spanned by molecular

strands, and with a dense undercoat associated with actin filaments at the cytoplasmic surface. There are

several types of AJs. Typically, in highly polarized epithelial cells, the AJ encircles the cell completely at

the apical/ basolateral border like a belt and is called the zonula adherens (ZA). Close to the ZA, the tight

junction (TJ) forms apically. Punctate forms of AJs are called punctum adherens (PA). Homophilic

binding results in recruitment of ARP2/3 complex (responsible for branched actin

polymerization) to the adhesive contact (Yonemura 2011). In cadherin junctions (zonula adherens), the

intracellular domain of cadherin associates with the actin cortex through accessory proteins (mainly

catenins). Cadherin junctions are crucial for cells to adhere, migrate, segregate and differentiate in a

selective and coordinated fashion. Moreover they play an important mechanical role in regulating and

responding to cell –cell interface tension (Fig. 1.2d). At the cellular level, the behavior of the cadherin–

catenin complex depends on a-catenin and acto-myosin (Fig. 1.2d). Binding of cadherin associated a-

catenin to actin can occur through interaction with actin-binding proteins, but also directly through its

actin-binding site in a stretch dependent manner. (Yonemura 2011).

The mechanical and active behaviour of cells are surprisingly different when probed in suspension and in

contact with specifically treated substrates which will be reviewed in the following sections.

7

Page 18: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 1.3 Terminology and methods to understand cell and tissue mechanics. a) Basic mechanical

terms (i) Diagrams illustrating definitions of tension, compression and shear forces on a block of

material. Open arrows indicate directions of forces on the block before (left) and after (right) deformation.

(ii-iii) Graphs showing the relationship between stress and strain after application of step changes in stress

(ii) or after rapid changes in strain (iii). The responses of four hypothetical materials are shown: elastic

(an ideal spring), viscous (a fluid), and two examples of viscoelastic materials. Also shown in ii are

simple network models in which the four materials are represented by combinations of springs and

dashpots. (Davidson et al. 2009). b) Viscous phase lag in sinusoidal stress. c) A schematic showing the

main rheometry techniques by which the viscoelastic propertied can be probed at different length and

time scales. The double arrows indicate the sense of applied deformations or force. Laser light is shown in

red. (Pullarkat et al. 2007)

8

Page 19: applying optical tweezers in vivo: a biophysical study of mechanical

1.3 Current mechanical understanding of single cells and tissues in vitro

Typically three schemes are followed to probe the viscoelastic properties of soft materials. These are (i)

creep experiments, where a step force is applied and held constant ,the resulting strain is measured as a

function of time, (ii) relaxation experiments, where step strain is applied and the resulting relaxation of

force is measured, and (iii) oscillatory experiments, where a sinusoidal strain is imposed on the cell and

the phase as well as the amplitude of the force response is measured (Fig. 1.3a and b). In practice, the

coupling between mechanical stress and strain is empirically determined by experiments on how materials

deform in response to applied forces. “Stiffness” in such experiments refers to the elastic response of a

body to an applied force and depends on the its geometry and material properties. (Davidson et al. 2009).

Various techniques have been developed to probe cell mechanics (Fig. 1.3c).Traction force microscopy is

a mostly passive technique aimed at measuring the active force generated by cells in response to different

mechanical or chemical perturbations or in mapping the traction forces generated by locomoting cells. It is

an approach with minimal perturbation to the cell, which is allowed to take any shape it would like.

Initially elastic substrates were used and the stress field was studied by looking at substrate wrinkles

caused by the active contraction of the cells. A major improvement in the methodology are non-wrinkling

silicon substrates where the deformation field can be mapped by measuring the displacement of embedded

beads. A much more sophisticated procedure are arrays of discrete, sub-micrometer-sized elastic pillars,

resembling a fakir bed, on which the cells can adhere and crawl. Micropipette aspiration is a relatively

simple technique where a micropipette with a tip diameter of a few micrometers is used to aspiratea

portion of a cell by applying a known negative pressure. Micropipettes have diameters of about 1–8 µm,

suction pressures are about 0.1–105 Pa. This technique has been successfully applied to probe the

mechanical properties of cells like the red blood cells which have a simple thin spectrin cytoskeleton

tightly adhering to the membrane. Atomic force microscopy (AFM, Fig. 1.3c ii) can probe cell mechanical

properties at a submicrometer scale. Typically, forces in the range of 0.1–1 nN and indentations of about

50 nm or less are applied to the cell to probe its viscoelastic response. Frequency responses of both storage

and loss moduli in the range of 0.1–300 Hz can be measured if corrections to hydrodynamic artifacts are

applied. Modified AFM tips, where a bead with a diameter of a few microns is glued to the normally sharp

tip, allows the modification of the stress range by changing the bead radius and diminishes the

indentations. Magnetic bead microrheology (Fig 1.3c iii) generates well defined forces on tiny

paramagnetic beads with an electromagnet to probe the mechanical response of cells. The beads can be

attached to the outside of the cell using specific adhesion promoting proteins to form transmembrane

9

Page 20: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 1.4 Mechanical models for cells. a) Overview of different mechanical models (Lim et al. 2006).

b) Steady-state contractile force as a function of support stiffness predicted with the constrained mixture

model. The steady-state morphology and corresponding stress fiber distributions are shown for select

values of substrate stiffness. For comparison, experimental results from Ghibaudo et al. 2009 are

also reported. (Rodriguez et al. 2013). c) Model of cadherin contact formation and strengthening in

response to mechanical changes in the cell-cell contacts. The substrate is coated with N-cadherin and

represents a neighbouring cell. The close-ups of cadherin contacts show the balance of external and

internal forces (Fext and Fcell, respectively). (Ladoux et al. 2010)

10

Page 21: applying optical tweezers in vivo: a biophysical study of mechanical

By combining experiments and modeling detailed understanding of cell mechanics in various in

vitro systems has been deduced. Fig. 1.4b shows comparison of results obtained by traction force

microscopy and then constrained mixture model. This model is able to accurately capture the relationship

between cell contraction and substrate stiffness and predicts correctly, the formation and orientation of

stress fibers in cells stretched on substrates with different stiffness. (Rodriguez et al. 2013, Fig. 1.4b).

Fig. 1.4c illustrates an experimental model of cadherin contact formation and force response, deduced

from traction force microscopy experiments. As the cell pulls on the substrate via cadherin adhesions, it

induces an increase of its internal tension by recruiting adhesion proteins and upregulation of acto-

myosin contractility. On a stiff substrate (large K), the internal tension (Kint) is supported by the

formation of large clusters of cadherin complexes. When the cellular environment provides less

resistance to deformation (small k), small forces are observed with limited cadherin links and

therefore, smaller internal rigidity (Kint). (Ladoux et al. 2010 , Fig. 1.4c).

11

linkages with the cytoskeleton (Pullarkat et al. 2007). Techniques like micropipette aspiration can effectively probe the creep response of cells, while AFM, magnetic twisting cytometry, microplates, etc, can perform oscillatory probing at different frequencies. AFM and magnetic twisting cytometry offers the widest range of frequencies, from 0.01 Hz up to 1 kHz. The microplate technique (iv), although limited in frequency, can perform variety of measurements not easily achievable in other methods. On the other hand, optical techniques like laser tweezers (i) and optical stretcher (v) can make rheological measurements on non-adhering cells. (Pullarkat et al. 2007).

Mechanical models for biological materials are derived using either the micro/nanostructural approach or the continuum approach. The micro-structural approach focuses on the cytoskeleton as the major structural component and is applied especially for understanding cytoskeletal mechanics in adherent as well as floating cells.. The continuum approach attributes continuum material properties to cell components which is more straightforward for understanding the biomechanical response at the cell level. It also provides a distribution of stress and strain on the cell which can be used to calculate transmission and distribution of forces/tension. This can furthermore assist in the development of more accurate micro and nanostructural models. (Lim et al. 2006, Fig. 1.4a).

Page 22: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 1.5 a) Tissue movements and rearrangements during development. (i) Mesoderm invagination

on the ventral side of the Drosophila embryo starts 3 h after fertilization and is completed within 5 min.

Presumptive mesoderm cells undergo pulsatile apical contraction as they start the process of gastrulation.

Right; graph showing the stepwise cell area reduction of these cells and the fluorescence intensity levels of

a MyosinGFP reporter. (ii) Dorsal closure in Drosophila embryo. This process takes approximately

three hours from the end of germ-band retraction with slow AS contraction starting after around 45 min.

Right; anticorrelation between the cell area and MyosinGFP intensity levels in amnioserosa cells.

(Gorfinkiel et al. 2014). b) Cell reorganization and shape changes underlying tissue movements

i) convergence of single cell boundaries ii) Contraction of multicellular boundaries iii) Apical constriction

and iv) directional basal constriction (Baum et al. 2011).

12

Page 23: applying optical tweezers in vivo: a biophysical study of mechanical

1.4 Morphogenesis: Molecular events and dynamics during embryonic

development

Development of a single-celled embryo into a complete organism is a complex process involved many

changes and dynamics at the molecular, cellular and tissue level. Cells form tissues (epithelia) through

cell division. With the onset of gastrulation cells undergo variety of changes in their morphology/shape

and also remodel their contacts. This process of tissue morphogenesis has been studied for many years in

the context of biochemistry and molecular players involved. But with the advent of time lapse imaging,

lot of focus has been shed recently on the temporal and spatial dynamics of this process using

various model systems (Gorfinkiel et al. 2014). Initially reported in the C. elegans oocyte, thus far,

epithelial cells in at least five tissues undergoing morphogenesis have been shown to exhibit episodic

actomyosin behaviour, four in Drosophila and one in the Xenopus embryo. Mesoderm invagination and

dorsal closure during Drosophila gastrulation are 2 such examples. Mesoderm invagination is the first

step in Drosophila gastrulation and leads to the formation of the mesoderm. Dorsal closure occurs much

later wherein the dorsal amnioserosa undergoes contraction and disappears while the surrounding

tissues closes the gap (which has also been used as a model for wound healing). Both these

processes exhibit pulsatile behavior in cell changes (apical constriction for cells in mesoderm

invagination, contraction of cell area for amnioserosa) (Fig. 1.5a). Such pulsatile behavior for cell

reorganization and shape changes has been observed in other model systems as well. Fig. 1.5b shows a

schematic of the various shape changes that can be observed in tissue morphogenesis (Baum et al. 2014).

Cells in an epithelium can shrink or contract their boundaries leading to loss of junctions (cell

intercalation, results in formation of now boundaries/junction in a perpendicular direction, eg.

germband elongation) (i). Cells can also coordinate together to contract multicellular boundaries,

forming rosette-like structures (ii) as well as constrict their apical and basal ends (Fig. 1.5b iii and iv).

Cells actively produce these shape changes by generating forces at their surfaces, which are transmitted

through cell interfaces. These ‘cortical’ forces, are generated in the cell cortex. Cortical forces build up

from a variety of mechanisms, importantly from Myosin II and actin filaments assembly which can be

spatially and temporally controlled in the cell. Understanding how these forces emerge from the assembly

and contraction of acto-myosin networks coupled to adhesion structures is a central issue in cell and

developmental biology. Therefore the spatiotemporal behavior of acto-myosin has been studied in

many systems. Indeed in some of these systems pulsatile changes in actin and myosin levels are

observed with the onset of cell contractions (Fig. 1.6b). The frequency of actomyosin foci and of the

13

Page 24: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 1.6 Acto-myosin dynamics in morphogenesis. (a) Schematic representations of the acto-

myosin organization and ratchet mechanisms within cells of the model tissues (red; actin, green, myosin).

Cell views are apical. In the Drosophila mesoderm cells, high frequency myosin foci reinforce themselves

and contribute to the progressive development of an apical ratchet as the acto-myosin network becomes

connected across cells. Both junctional recruitment and the development of a multi-cellular network of

acto-myosin as the frequency of foci increases contribute to increasing stiffness of the amnioserosa. (b)

Time offsets of the maximal rate of change in cell area (or junction length for the germ-band cells)

relative to the maximal rate of increase in fluorescence intensity of myosin and/or actin. (c) Range of

fluctuation period lengths in all model tissues. (Ect, ectoderm; EM, early mesoderm; LM, late mesoderm).

(d) Tissue contraction rate versus fluctuation period length (DMI, Drosophila mesoderm

invagination; GB, germ-band; AS, amnioserosa; ND, not detectable). (Gorfinkiel et al. 2014)

14

Page 25: applying optical tweezers in vivo: a biophysical study of mechanical

resulting fluctuating contractility of cells in the model tissues is similar The pulses of acto-myosin

changes as well as well cell contraction are coordinated, and seem ot occur between intervals of a few

minutes (1-10 minutes, Fig. 1.6c). The amount of tissue contraction resulting from coordinated cell

contractions shows a correlation with the period of fluctuation/pulses (Fig. 1.6d). In the absence of

intrinsic cellular mechanisms that generate polarity in the plane of the epithelium, anisotropic tissue

tension could organise the spatial architecture of the actomyosin cytoskeleton and orient the flow of

foci. Both mesoderm and amnioserosa cells in Drosophila do not exhibit planar localization of junctional

proteins but experience anisotropic tension that is greatest along the A/P axis. Pulsatile actomyosin

contractility within cells raises the question of how tissues effectively contract. A ratchet

mechanism has been postulated for the stabilization of fluctuations and the generation of net contraction.

Both medial and junctional actomyosin populations contribute to tissue remodelling in various

tissues. In the amnioserosa, an increase in junctional actomyosin fluorescence is observed during dorsal

closure and is correlated with a straightening of initially wiggly membranes and productive tissue

contraction. In the germ-band the dynamic apical foci are thought to feed and reinforce the D/V oriented

junctions. In both these tissues it is possible that the junctional population acts as a ratchet that maintains

the contraction generated by the medial actomyosin network. Overall this suggests that cell contractions

are regulated by acto-myosin ratchets (Fig. 1.6a illustrates that the organization of acto-myosin and

the ratchets involved in mesoderm invagination and dorsal closure). (Gorfinkiel et al. 2014, Lecuit et al.

2011, Rauzi et al. 2011).

Elucidating the forces that form and reshape multicellular structures is thus integral to the understanding

of development. It is therefore important to understand the material properties of in vivo epithelia

to understand how dynamic acto-myosin behavior evolves the forces within the tissue.

15

Page 26: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 1.7 Measuring mechanical forces in morphogenesis. a) An example of laser microsurgery that cuts

a single cell-cell boundary . The cut (blue arrow in 0sec) locally releases tension, which creates an opening

in the tissue. The cells surrounding the opening (in red) and cells further away (in green) are analyzed in

80sec and 180sec. (Solon et al. 2009). b) Annular ablation experiment in the scutellum. The epithelial cell

apical junctions are marked by E-cadherin:GFP. The region between two concentric circles indicated by blue

lines defines the annular severed region (left). The circular domain retracts after the cutting as shown in the

images 1 s after (middle) and 30 s after (right) cutting. Yellow: fitted ellipse. The larger opening along y-

than x-axes indicates anisotropic stress in the tissue. (Ishihara et al. 2013). c) Estimated tensions and

pressures for a Drosophila pupal wing. (i) An image of a Drosophila wing at 23 h APF. D_catenin-TagRFP is used to highlight cell shape. Scale bar: 20 μm. (ii) Color maps of estimated cell-junction tension using ST.

A color scale is shown. (iii) Comparison of estimated tensions obtained using STP and ST. (Ishihara et al. 2013).

16

Page 27: applying optical tweezers in vivo: a biophysical study of mechanical

1.5 Mechanical changes underlying morphogenetic movements: Current

methods and theoretical understanding

Understanding the mechanical nature of morphogenetic movements requires force measurements in vivo.

However, while a variety of techniques have been developed for in vitro measurements of cells and

tissues, applying these techniques in vivo is challenging due to the direct mechanical contact required

between the probes and the cell/tissue required in most of these techniques. As cells and tissues within

embryos respond differently to external stimuli compared to isolated cells on artificial substrates, it

requires a unique approach to make these measurements.

Most information on mechanics in embryos comes from indirect methods. A well-studied example is laser

ablation. In this technique a highly focused and high power femtosecond laser beam is pulsed within the

cell or tissue to induce perturbation/breakage. In the context of mechanics, at the single cell level this

typically involves making a ‘cut’ within the acto-myosin cortex by focusing the laser on a spot of

few 100nms (Rauzi et al. 2011). As the cortex is under dynamic tension which is suddenly ‘released’

by the ablation, it tries to relax in response (Fig. 1.7a amnioserosa during dorsal closure, Solon et al.

2009). This can be measured with time-lapse imaging as a function of cortex movement away from

the point of ablation. A variant of this approach is to do a tissue level ablation (for eg. A circular/

annular cut) and measure the relaxation at the tissue level. An example in the Drosophila scutellum is

shown in Fig. 1.7b (note the anisotropic relaxation in the anterior-posterior and dorso-ventral

axes). The nature and anisotropy of tension in epithelia can be estimated. However since the

relaxation depends on the cytoplasmic friction coefficient (for which there are no direct measurement

in most systems), in addition to the tension; it only provides an indirect estimate of the forces involved.

Nevertheless due to its non-invasive nature, laser ablation has been used to understand epithelial

mechanics in various experimental models of morphogenesis including germband elongation, dorsal

closure and mesoderm invagination (Rauzi et al. 2008, Solon et al. 2009).

An alternative to perturbation is force inference from image analysis. The force inference method is non-

invasive and provides space-time maps of stress in a whole tissue, unlike existing methods (Fig. 1.8iii,

example in wing disc). Different force-inference methods differ in their approach of treating

indefiniteness between cell shapes and forces. Tests using artificial and experimental data sets

consistently indicate that the Bayesian force inference, by which cell-junction tensions and cell pressures

are simultaneously estimated, performs best in terms of accuracy and robustness. Moreover, by measuring

17

Page 28: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 1.8 Mechanical models of morphogenesis. a) Experiments and 3D model for Drosophila gastrulation. (i) Ventral (from Grumbling and Strelets, 2006) and cross-sectional (from Muñoz et al., 2007) views of ventral furrow formation in experiments. (ii) Same views from finite element model. b)

Two-step process for invagination in ascidian gastrulation. Step 1: Apical constriction results in wedge-

shaped cells in endoderm surrounded by ectodermal cells. Step 2: Subsequent apico-basal contraction of

endodermal cells results in invagination. c) Cell shape organization in the ommatidium of the Drosophila

retina. The top left panel shows the cell organization in a normal fly, and the bottom left panel is the cell

organization predicted by a simple mechanical model that considered only adhesion energies and

membrane elasticity. The right panels show the cell organization in a mutant fly in which the left cone

cell (black) indicated by the red lines (bottom panel) lacks N-cadherin. The effect of this deletion is

accurately predicted by the mechanical model. (a,b from Wyczalkowski et al. 2012 and c from Niessen

et al. 2011)

18

Page 29: applying optical tweezers in vivo: a biophysical study of mechanical

the stress anisotropy and relaxation, the force inference and the global annular ablation (Fig. 1.8b) of the

tissue can be cross-validated, as each of them relies on different prefactors. (Ishihara et al. 2013)

While experimental probing of in vivo mechanics has only developed recently, mathematical models

have been applied to morphogenesis for a few decades. In one of the earliest theoretical

investigations of morphogenesis, Odell (Odell et al. 1981) presented a 2-dimensional (2D) model for

an epithelium that treats each cell as a viscoelastic element with a contractile apex. In a ring of cells,

contraction in one cell apex (which simulated by shortening the stress-free length) stretches

neighbouring cells, which can also contract themselves if stretched beyond a limit. On selecting the right

parameters this response produces a wave of contraction that generates a local invagination. (similar to

observed in ventral furrow formation and neurulation). Since this significant study, lot of models

have been developed, extending into 3D. Particularly, continuum growth theory based models of

ventral furrow formation can simulate active changes in cell shapes (Fig. 1.8 a and b). In these

models, cell dimensions change by specifying positive or negative growth along particular directions,

while cell wedging occurs via an apico-basal growth gradient. The models were used to study the

effects of various combinations of these cell shape changes, as well as 3-dimensional (3D) ellipsoidal

geometry and constraints imposed by the surrounding vitelline membrane and internal fluid (Fig. 1.8b).

The 3D ellipsoidal model yielded global shape changes similar to those observed in experiments, and

confirmed the important role of the vitelline membrane and yolk in gastrulation. (Bonnet et al. 2012,

Fletcher et al. 2014, Hutson et al. 2003, Munoz et al. 2007, Wyczalkowski et al. 2012,)

Quantitative vertex models can also predict effect of biochemical perturbation on cell-shapes in vivo Fig.

1.8c shows such an example in the Drosophila ommatidium. The cell shapes in the ommatidia are

reminiscent of soap bubbles, whose geometry is determined entirely by surface tension. A simple

mechanical model was sufficient to predict cell geometries in vivo, for different cluster sizes and for

different mutants. But this correlation is only indicated for small cell clusters. In more complex tissues,

other forces might play a significantly greater role to determine cell morphologies than adhesion energies.

(Niessen et al. 2011)

19

Page 30: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 1.9 Methods for in vivo force measurements. a) Schematic of magnetic tweezer perturbation in Drosophila embryos: Tissue level laser ablation and application of external force (1.5 amp for 1.7 min)

using a magnetic tweezers on Drosophila embryo injected with 100 nm paramagnetic beads). (Kumar et al. 2012) b) AFM. (i,ii) Confocal zx profiles of a cell monolayer (green) grown on a soft collagen gel (black),

before (i) and during (ii) indentation with an AFM cantilever (dotted line). White arrowhead, an individual

cell; grey arrowhead, the tip of the cantilever. A fluorescent dye was added to the extracellular medium (red).

Scale bar: 20 mm. (Harris et al. 2014) c) Schematic drawing of the tension-sensing (TS) FRET sensor

module. Teal fluorescent protein (mTFP) is separated from Venus, a yellow fluorescent protein, by a

nanospring protein domain from spider silk. In the relaxed state, the two fluorophores are close enough to

allow FRET. The spider silk domain stretches in response to pico Newton forces, reducing FRET (Grashoff

et al., 2010). d) Force measurments with FRET sensor in Drosophila embryos. (i-iv) Rescue of Armadillo

expression (Arm, which is Drosophila b-catenin) in border cells after EcadRNAi (ii) by CadTS (iii) and

control (iv). Scale bar shows 10 μm. v) Histogram showing CadTS and control rescuing border cell

migration after border cell-specific (ii) EcadRNAi. (Cai et al. 2014).

D

20

Page 31: applying optical tweezers in vivo: a biophysical study of mechanical

1.6 In vivo tools for direct understanding of forces

Most of the information on mechanical forces in vivo comes from indirect approaches. While such

measurements from direct methods are lacking, approaches are nevertheless being developed. In

particular magnetic tweezers have been applied in Drosophila embryos in conjunction with laser ablation,

to probe the effect of mechanical perturbation on morphogenetic movements and the corresponding

gene expression patterns (Fig. 1.9a). Mechanical forces induced by magnetic tweezers in paramagnetic

beads injected in the embryo, altered nuclear morphology as well a induced movement (Kumar et al.

2012). Atomic force microscopy represents another approach to probe-less manipulation. Although it is

limited to probing forces at the surfaces of embryo, it provides the advantage of high-resolution

imaging combined with measurement and application of pN-nN forces. While imaging of live embryos

has been done with AFM, mechanical probing has been limited to tissue monolayers (Fig. 1.9b).

Nevertheless with the advent in technology, it will prove a useful tool for understanding in vivo

mechanics (Harris et al. 2014). FRET-based force sensing is another important approach developed in

recent years (Grashoff et al.). A force sensitive peptide is placed between a FRET donor-acceptor pair.

Depending on the force applied, the distance between the donor and acceptor changes, thus altering

the FRET efficiency (Fig. 1.9c schematic). By calibrating the force versus FRET efficiency in vitro,

the approach can then be applied in vivo. In fact such a measurement has already been done in

Drosophila embryos during border cell migration (Cai et al. 2014, Fig. 1.9d).

Despite these developments, optical tweezers and particle tracking rheology remain the most viable and

least-explored tools to understand morphogenesis. Optical forces can manipulate and apply forces to few

100nm-micron sized probes. The most classical application of this is the single beam gradient optical trap,

also known as optical tweezers (Ashkin 1987). In a tightly focused laser beam, the scattering force that

pushes a probe in the direction of laser propagation is balanced by the gradient force which

is proportional to the intensity gradient of the beam (Fig. 1.10a). When the probed object has a

positive refractive index mismatch with the surrounding medium, it gets ‘trapped’ in the optical

tweezers. Since refractive index variation can arise naturally in biological tissues, native objects such as

lipid vesicles and or organelles can be probed and manipulated. For small displacements up to 1 µm and

for sufficient index mismatch, the trap can be applied with minimal optical damage with a linear force-

dependent spring-like behavior. The stiffness of this spring like system needs to be calibrated for precise

21

Page 32: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 1.10 Overview of optical tweezers and particle tracking rheology a) Origin of Fscat and Fgrad in single beam gradient trap (optical tweezer) for high index sphere displaced from TEM00 beam axis

(Ashkin 1987). b) Manipulation of DNA attached to a glass surface with optical tweezer. (Hormeno et al. 2006). c) Artificial cytoplasmic filaments in a scallion cell. The laser trap is moved from A to B, pulling out the viscoelastic filament AB into the central vacuole (Ashkin et al. 1989). d) An example of in vivo

stall force measurement for motor proteins (kinesin I and cytoplasmic dynein). (left) Snapshots showing a

lipid droplet in an embryo trapped at multiple positions along its trajectory shown in right. (Leidel et al. 2012). e) Schematic for particle-tracking microrheology. Left:The random spontaneous movements

of injected beads within the cytoplasm are monitored with high spatial and temporal resolution.

Right: Typical trajectory of the beads in the cytoplasm. (Wu et al. 2012). f) Left: Inert nanoparticles

injected in C. elegans embryo uniformly dispersed after first cell divison. Right: The bulk diffusion

coefficient of the anterior and posterior cytoplasm. (Daniels et al. 2006)

22

Page 33: applying optical tweezers in vivo: a biophysical study of mechanical

force measurements. In fact, the ability of optical tweezers to achieve probe-less manipulation was

demonstrated long ago (Ashkin 1989, Fig. 1.10c). But classical probes such as silica or polystyrene

microspheres are easier to calibrate for force measurements (Fig.1.10b, Hormeno et al. 2006).

Hence such measurements are traditionally limited to in vitro biochemical systems. However

measurements have been done on motor transport in vivo by trapping endogenous lipid droplets

(Fig. 1.10d, Leidel et al. 2012). Such measurements rely on advances in trap calibration as well

as particle tracking rheology. In particle tracking rheology beads are injected in cells or embryos.

The time-dependent (x, y) coordinates of the beads are mathematically transformed into mean squared

displacements (MSDs). The time lag-dependent MSDs of the beads are subsequently transformed into

local values of either the frequency-dependent viscoelastic moduli or the creep compliance of the

cytoplasm. (Fig. 1.10e, Wu et al. 2012). The measured viscosity depends on the size of the probe, hence

the use of endogenous particles which can vary in size, has not been verified. Nevertheless this

approach has already been used in vivo using injected beads, for example to passively probe the

viscoelastic properties in the C. elegans embryo (Fig. 1.10f, Daniels et al. 2006). Combining passive

particle tracking with active force perturbation with optical tweezers could provide the missing gap

in extending mechanical understanding from in vitro to in vivo, since such active and passive

microrhelogy is well characterized in vitro.

23

Page 34: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 1.11 Germband elongation: Molecular and cellular events a) Cartoon depicting a Drosophilaembryo during gastrulation. The germband (GB) converges in one direction extending in the

perpendicular direction. GB convergence-extension is driven by a cell cell intercalation. Cell intercalation

is polarized along the anterior/posterior axis. (Rauzi et al. 2011) b) Enrichment of Sqh–GFP at type 1

junctions (arrowheads). On left, time-lapse sequence of Sqh–GFP (times in minutes), showing the

enrichment at type 1 junctions (0, pink and orange arrowheads), in type 2 junctions (20 min, pink

arrowheads) and the reduced localization at type 3 junctions (33 min, pink and orange arrowheads).

(Bertet et al. 2004). c) Two-tiered actin model to explain regulation of the stability and mobility of homo-

E-cad clusters by actin, in SAJs. (Cavey et al. 2010). d) Left; localization of Myo-II and E-cad before

and during intercalation. On right; respective distribution of medial (red) and junctional (green) Myo-II

along the apico-basal (z) axis (Rauzi et al. 2010).

24

Page 35: applying optical tweezers in vivo: a biophysical study of mechanical

1.7 Germband elongation in Drosophila as a tool for understanding the

mechanical forces in morphogenesis

The first 13 nuclear divisions of the Drosophila embryo occur in a syncytium, resulting in 6,000

peripheral nuclei located beneath the plasma membrane. During cellularization, the membrane surface

increases ~25-fold, invaginates between the nuclei, and ultimately yields 6,000 epithelial cells 30 μm tall

(Lecuit et al. 2000). The global contraction of the actomyosin network at the leading edge of the furrow

provides the force necessary to pull down the membrane. Myosin enrichment is mostly at the

cellularization front which proceeds basally to complete cellularization. At the onset of gastrulation the

membrane folds disappear and myosin enriches apically. At the onset of gastrulation, the resultant

epidermal cells have the hallmarks of polarized cells with adherens junctions separating the apical and

basal–lateral domains. It is still unknown how this polarity is established. However it has been observed

that cadherin clusters form at the end of cellularization and migrate apically by the time germband

extension occurs (TroungQuang et al. 2013). The Drosophila germband at the onset of gastrulation is an

excellent model system to study cell intercalation. (Fig. 1.11a). It has been suggested that

differential adhesion between groups of cells could drive cell rearrangement (Doubrovinski

et al. 2014), Cells in the germband rearrange their neighbours by remodeling their junction

in a polarized fashion so that junctions parallel to the dorsal/ventral axis (vertical

junctions) shrink bringing four cells in contact and then expand in a direction parallel to the

anterior/posterior axis so that more dorsal and ventral cells form new contacts. (Fig. 1.11b Bertet et

al. 2004) During this process, Myosin II enriches along vertical junctions. Myosin II is

necessary for junction remodeling as the contractile activity of Myosin II might create a local

tension that orients the disassembly of junctions. This hypothesis was tested by a quantitative

comparison between in vivo data and in silico predictions and laser subcellular dissection (Rauzi et

al. 2008). A two-tiered meshanism for regulation of cadherin cluster dynamics has been proposed. In

this model, Stable, small actin patches concentrate and stabilize homo-E-cad in SAJs.

a-Cat is not necessary for stability, and unknown linkers between actin and E-cad are

involved. A dynamic, contractile network regulated by Bitesize and Moesin tethers all SAJs and

limits their lateral mobility through a-Cat (right), thereby maintaining adhesion in a defined domain.

(Cavey et al. 2010).

25

Page 36: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 1.12 Mechanical forces during germband elongation. a) Forces during cell intercalation revealed

by laser nano-dissection. (i) Local ablation of the subcortical acto-myosin network (left) causes

redistribution of F-actin and E-cadherin and changes the force balance at cell junctions (middle). (ii, iii) Time-lapse sequence following nano-dissection, which was performed at t = 0 in a subwavelength

volume (red arrowhead). MoeABD::GFP to mark F-actin (ii), E-cadherin::GFP (iii). Three-photon

uncaging of fluorescein in a single cell ensures that the plasma membrane is not permeabilized.

Kymographs show the temporal evolution of the fluorescence intensity along the targeted junctions

where the nano-dissection was performed. The two vertices of targeted junctions move apart after

nanodissection. Scale bars, 5 μm. (white), 50 s (blue). (iv) νmax as a function of ϕ for v-junctions (vertical) and t-junctions (transverse). Black and red solid curves are fits of νmax for the v- and t-

junctions, respectively. The orange solid curve represents the expected speeds for t-junctions assuming

mechanical equilibrium between t- and v-junctions. (v) Expected tension in a line tension or a cortical

elasticity model for v- and t-junctions in mechanical equilibriumand for a simple geometric

transformation at constant area (inset). (Rauzi et al. 2008). b) Tension anisotropy is sufficient to drive

tissue elongation. (i) A group of cells observed in vivo at the onset and at 40 min of elongation (top).

Comparison with simulations, started from the same cell network, imposing a cortical elasticity anisotropy β = 1.8. Bottom right: in silico states at the same relative elongation (rel. elong.) as that observed in vivo.

Bottom left: in silico final state. (ii) The final relative elongation of the cell network is a function of the

tension anisotropy imposed at the onset of simulations (squares, cortical elasticity anisotropy β). (iii) Tissue elongation as a function of T1 transitions in vivo (red squares) and in silico for different values of

tension anisotropy. (Rauzi et al. 2008).

26

Page 37: applying optical tweezers in vivo: a biophysical study of mechanical

Rauzi et al. (2008) revealed an anisotropy of cortical forces along cell junctions controlled by Myosin II:

this was measured to be a factor 2 along vertical junctions (junctions with greater density of Myosin II)

compared to other junctions. Cortical forces were inferred by laser dissection experiments: disruption of

the acto-myosin network underlying a given junction modified the balance of forces and produced

junction relaxation, whose speed is indicative of cortical tension . The authors designed a model based on

the local (junctional) and global (cellular) natures of cortical forces. The cellular network configurations

during tissue shape changes were described as the succession of local minima of an energy Starting the

simulations with cell patterns observed at the onset of intercalation, we monitored the same cells in silico

and in vivo during elongation (Fig. 1.12a). We studied three embryos for 40 min. Tissue elongation after

complete energy minimization depends on cortical elasticity anisotropy (Fig. 1.12b). For tension

anisotropies below 1.4, the tissue failed to elongate significantly, but above 1.6, final elongation was

maximal, indicating that moderate tension anisotropy is sufficient to drive maximal tissue elongation.

(Rauzi et al. 2008)

27

Page 38: applying optical tweezers in vivo: a biophysical study of mechanical

1.8 Summary

We are at the stage where we continue to develop a better understanding of mechanical equilibrium in

tissues, particularly in the context of cell-cell and cell-ECM adhesion. However recent studies have made

it clear that tissue monolayers in vitro have different properties than in vivo epithelia. Thus it becomes

ever more important to probe tissue dynamics within embryos. Understanding the role of mechanical

forces in tissue morphogenesis is a longstanding problem in biology. Due to advances in

fluorescent labeling and real-time in vivo imaging in recent years, a lot of light has been shed on the

nature of forces in development. Particularly, the dynamic nature of acto-myosin organization in the

form of pulses over a seconds plays a crucial role in morphogenetic events in different model systems.

One such system is the germband elongation in Drosophila gastrulation. Mechanical properties

underlying embryo morphogenesis have been mainly done with through laser ablation and force

inference methods. However methods for direct measurements of forces in vivo are only being

developed now, particularly through advances in particle tracking rheology and optical tweezers.

28

Page 39: applying optical tweezers in vivo: a biophysical study of mechanical

CHAPTER 2

MATERIALS AND METHODS

29

Page 40: applying optical tweezers in vivo: a biophysical study of mechanical

2.1 Materials

Drosophila melanogaster strains with either Gap43::mcherry fluorescent marker ( [w ; UASt-

sqhE20E21/Cyo ; UAS-diaCA/TM6tb II,III], [w, upd/Fm7 ; endocad-GFP, sqh-cherry /Cyo X,II] or [w ;

sqh-GFP::utABD/Cyo ; UASp-Gap43::mCherry/TM6tb II,III]) were used for optical tweezer

experiments. Cadherin imaging was done with endocad::GFP (,[tl1-endoCad-TCSi::GFP-12 II], [w,

upd/Fm7 ; endocad-GFP, sqh-cherry /Cyo X,II]). Myosin imaging was done with squash::GFP (w ; sqh-

GFP::utABD/Cyo ; UASp-Gap43::mCherry/TM6tb II,III) and actin imaging was done with lifeact::GFP

(sqhp-lifeact::eGFP/MKRS II ). Flies were maintained at 22ºC in tubes with fly medium and in a cage at

25ºC for experiments. Fresh medium plates (with yeast) were left for 2:30 hours in the cage before

collecting embryos, to provide a variation of stages and thus enough embryos for experiment (to account

slow experiment with tweezers and rapid development of stages of interest). Alternatively for more

precision, plates were left for 1 hour in cage and then an additional hour at 25C after removing from the

cage. Embryos of interest were typically at the end of cellularization and could be probed till the early

germband elongation. For myosin perturbation 10mM concentration of Y27632 ROCK Inhibitor

(Invitrogen) was used. Actin perturbation was done either with 100 µM cytochalasin D or 1mM

latrunculin A.

2.2 Sample Preparation

To prepare samples for imaging/optical tweezer experiments, plates with embryos in suitable stage were

collected. After washing off yeast from the plate with deionized water, the embryos were washed with

100% bleach for 50 seconds to remove the vitelline membrane, then washed further with water to remove

the bleach and fragments of the vitelline. Embryos in the end of cellularization (stage 5 end) were then

selected under a Zeiss Stereo dissection scope. For deflection experiments, the embryos were glued on the

edge of a custom made SPIM coverslip which was then attached to a custom made chamber and

immersed in deionized water.. For spinning disk and phase measurements experiments, embryos were

aligned in the center of a rectangular cover-glass with a drop of halocarbon oil on top. In every case,

alignment was done with the germband visible in the imaging plane typically with anterior-posterior axis

vertically in reference to the camera image and the dorso-ventral axis horizontal. For perturbation,

embryos were placed in halocarbon oil and injected in a microinjection setup (Eppendorf Femtojet , 250-

400 psi,Ti -0.3, Tc -0) with water, ROCK inhibitor, latrunculin A or cytochalasin D at stage 5 end and

analysed during germband elongation (or the corresponding time point). Injection micro-needles were

custom made from glass capillaries with a glass pulling apparatus. Spilling embryo contents was carefully

30

Page 41: applying optical tweezers in vivo: a biophysical study of mechanical

avoided by adjusting the needle tip to less than cell size (~6 and injecting with a droplet few cell-sizes in

diameter. For perturbation followed by deflection experiments, the halocarbon oil was removed carefully

with a scalpel and rigorous washing under water before imaging.

2.3 Experimental setup

The light sheet setup is similar to the one in the seminal paper of Huisken (Huisken et al. 2004), and is

already well-described in a previous publication (Chardes et al. 2014). But briefly, 3 imaging lasers

(408, 488 and 561nm) are aligned along the same optical path using dichroic mirrors. An AOTF

allows switching the wavelength and power of the outgoing beam. This beam is passed through a beam

expander and cylindrical lens to create a light sheet, which is then focused on the sample using a Nikon

Plan Flour objective (0.3NA, 10x) with a resulting thickness of ~3 µm. Imaging is done with a

Zeiss Axiovert upright microscope. The light from the sample is collected using a 100x water immersion

Figure 2.1 Optical setup combining light sheet microscopy and optical tweezers.

lens (Nikon Plan 1.1NA), passed through dichroic lens and imaged onto a Andor IXON 3EMCCD

camera. Using this setup 3 channel acquisition is possible by alternating the AOTF output

wavelength and also simultaneously by an dual-view system before the camera path that can split the

emission from the GFP and cherry channels on 2 halves of the camera CCD (however only single

channel measurements were used in all the experiments). To enable optical tweezers on this setup, a

31

Page 42: applying optical tweezers in vivo: a biophysical study of mechanical

1071nm infra-red beam was introduced on the optical table using a continuous wave (CW),

Ytterbium Laser Module (YLM) (IPG photonics, maximum output 5W, 3A, 20% efficiency). All

the components for the trap setup were purchased from Thor labs. The beam was first passed

through a polarizer and lambda/4 plate. A system of 2 galvometric mirrors (galvo) was then

introduced in the path. A conjugating telescope (2 50mm lenses) was aligned between the 1st galvo

and 2nd galvo. While keeping the conjugation, the beam was then expanded with a long distance beam

expander to ensure that the objective back aperture was filled and to provide sufficient working

distance. The beam was then lifted to the objective plane using a periscope . Completing the

conjugation and alignment ensured that the laser trap was in focusing plane and the field of view of

the camera. The galvos could be triggered at the same time as the camera acquisition using a

National Instruments data acquisition card and custom software made in Qt creator. After building the

setup, the transmission of the entire setup was calculated with a reflecting Power meter and found to be

20% (fig 2.2 a). The galvo conjugation was confirmed for the experimental conditions by a linear

Figure 2.2 Calibration of trap power and displacement. a) Measurement of trap laser power at

different locations in the optical setup. b) Displacement of trap (as measured using trapped beads in

water) as a function of galvo voltage. c) Measured laser amplitude during interface experiments plotted

against expected amplitude from (b). Error bars represent range of values.

32

Page 43: applying optical tweezers in vivo: a biophysical study of mechanical

dependence of laser displacement against applied galvo voltage, calculated by trapping 500nm

fluorospheres in water. On average the conversion for a ~100nm excursion was 0.0027V and 0.0037V

for Y and X movements respectively (fig 2.2 b). In practice, the measured laser movement showed an

error of 30-40 nm compared to expected but the fluctuation of trap amplitude for any given position

calibration was 2-4% (~20±10 nm for 0.4-1 µm amplitudes) . (fig 2.2 c, standard deviation too small

to show on graph).

2.4 Optical tweezers experiments

Individual cell-cell interfaces in embryos were mechanically probed in a custom-built light sheet/selective

plane illumination (SPIM) setup coupled with a single beam gradient optical trap. A 100x water-

immersion lens (1.1NA) Nikon was used for imaging the sample on an EMCCD camera (supplementary

figure 1) as well as for introducing the optical trap in the imaging plane. Imaging was done using 488nm

and 561 nm excitation lasers (at 10mW and 60mW laser power). Laser position could be controlled

independently in X and Y with various functions using conjugated galvanometric mirrors (galvos). The

laser position in the camera image was calibrated using 500 nm carboxylate coated flurophores

(Invitrogen, excitation/emission peaks 580/605nm). In practice, the optical trap was sufficient to excite

the beads due to 2 photon absorption. The beads were trapped at ~50-100 mW laser power (after

objective) and given a circular oscillation of 10 µm at 0.07 Hz. All these parameters were standardized

after trial and error by checking the calibration accuracy with the beads. Nevertheless the calibration can

vary for larger excursions after just a few hours, therefore it was done for every experiment when

possible). The entire protocol for manipulating and recording the experiment was developed with custom

scripts and GUIs (Claire Chardes, Olivier Blanc) in Qt Creator (C++), and the triggering as well recording

of the voltage sent to the galvos was synchronized with the trigger for the camera recording (which could

be done in the range of 8-30 fps, depending on the experiment). The recordings were then analysed in

Matlab, using custom scripts, to calibrate the conversion of galvo voltage to trap position on the camera

image. For manipulating the cell-cell interface, initially, sinusoidal oscillations were given in X, Y or X-Y

axis, with different time periods ranging from 0.3- 5 seconds and amplitudes from 0.3-1.1 µm. Laser

power was varied from ~50-200 mW (after objective). These experiments had to be done during the end

of cellularization till the beginning of gastrulation (to ensure minimum variation in tissue properties).

Cycles recorded depended on period frequency but ranged from 2-20. After measuring the effect of these

parameters on deflection measurements, values were kept constant with time period of 2 seconds, laser

amplitude ~0.5 µm and laser power ~200 mW. Experiments were done in the cadherin plane (1-3 µm

33

Page 44: applying optical tweezers in vivo: a biophysical study of mechanical

from the apical cortex, depending on embryo stage) as well as in a plane that was 3 µm basal; at the end

of cellularization as well as during early germband elongation. In addition, pull-release experiments were

also done, where the laser was kept stationary and switched on at a distance of a few 100 nm-2 µm from

the cell-cell interface for 10s-1minute and the interface deformation as well as subsequent relaxation were

recorded.

2.5 Data Analysis

Kymographs of interface deflections were initially produced from the camera movies manually in Fiji

(multiple kymograph plugin, line thickness 6-9 pixels, with the resolution of the setup being 194 nm per

pixel). Using a gaussian fitting super-localization matlab script (developed earlier in the lab by Olivier

Blanc) on the kymograph, the cell-cell interface could be localized with sub-pixel accuracy. From this, a

minimum of 2 to a maximum of 10 oscillations (for fast frequencies >2hz) were used for measuring the

amplitude of deflection. However, the laser position calibration, along with additional matlab scripts

developed recently (Claire Chardes, Raphael Clement), can obtain a semi-automatic kymograph

(autokymo) and a sub-pixel localization of both laser and interface position directly from the recorded

images and voltage. In addition to measuring amplitude, this provides information on the spatiotemporal

relationship of laser-interface interaction. This position displacement data was therefore used to measure

the phase lag and position offset between the laser-interface. Data was stored in either Microsoft Excel

sheets or matlab figure and data files. Graphs were obtained using either Excel or matlab. Statistics was

done using the unpaired T-test (Graphpad).

2.6 Experimental limitations

There were variations in membrane intensity resulting from aberrations inherent in light sheet

microscopy (scattering and shadow effects) and 2 photon-excitation/bleaching that was observed in every

interface (fig 2.3 a) (excitation peak for cherry is close to 561nm which can be excited by the 1070nm

infra red trap). The excitation laser power was also varied to account for different acquisitions speeds at

different time periods of amplitude. Hence exact localization accuracy for the trap experiments could not

be calculated but amplitude of noise for stationary interface without any trap was consistent at different

stages between ~30±10 nm (fig 2.3 b). In addition the autokymo always optimizes a kymograph

34

Page 45: applying optical tweezers in vivo: a biophysical study of mechanical

perpendicular to laser movement axis, however in many cases this was not the best axis for measuring

interface deflection, due to stochastic variation in interface orientation as well as the rapidness with which

Figure 2.3 Noise in interface position detection. a) Interface deflection (black) and interface intensity

(green, AU) plotted as a function of time. Images correspond to interface image at the respective time

points (red=trap position). B) Amplitude of stationary interface noise (no trap) at different embryo stages.

Error bars are standard deviation.

the embryo could develop. Experiments in germband elongation were furthermore affected by tissue drift

and thus measurements were limited to 2-3 cycles and fast time period (2s or less). Also when measuring

offset between interface and laser positions, there is an uncertainty in laser position resulting from input

voltage noise (20±10 nm) and localization error of beads used for calibration (20-30 nm). Localization

accuracy for trap-interface offset could therefore be estimated to be ~ 70-100 nm. Indeed the offset of

expected to calibrated laser amplitude is also 20-40 nm (fig 2.2 c).

An obvious concern with a high power infra-red laser, is thermal damage to the tissue. While we did not

calculate the point spread function (PSF) of the trap, an objective NA of 1.1 and laser wavelength of

1070nm gives a PSF estimate of ~500nm (lambda/2NA). It has been shown under trapping conditions

35

Page 46: applying optical tweezers in vivo: a biophysical study of mechanical

similar to my experiments, that an instantaneous temperature increase of few degrees can be induced

locally (few 100 nm) on laser exposure, which decays within a second when the trap is switched off

(Ebbesen et al. 2012),. In my experiments, the major concern with this temperature increase is

of thermally induced breakage/perturbation of the cell-cell interface. Indeed at 400mW laser powers

we do see some damage (in terms of irreversible calcium uptake and interface shape changes). We do

not see this damage in 200mW power used for most experiments. Additionally, the interface

movement amplitude does not increase over several cycles in control embryos and the intial position of

the interface is restored after removing the trap in most cases. But to be on the safe side most

measurments were done for less than 30s per interface and only the first few cycles/oscillations were used

for analysis.

2.7 Quantification of E-cadherin, myosin II and lifeact

Imaging was done in a Perkin-Elmer spinning disk microscope using a 100x oil immersion lens. Z-stacks

from 6-12 µm thickness were acquired, starting from the apical cortex, with a slice thickness of 1 µm and

a sampling rate of 2 Hz (along with low laser power of 20% to minimize bleaching). Single channel

imaging was done with either 488nm excitation alone or sometimes (especially for myosin-II and ROCK

inhibitor where tissue could become largely invisible) squash-Gap43::cherry images were acquired with

the 561 laser for reference. Analysis of images was done with Fiji and the Perkin Elmer plugin. Maximum

intensity projections of 3 slices (with the plane of interest centered) were taken and then background

subtracted (ImageJ, 20 pixel rolling ball radius for E-cadherin and myosin-II, 60 pixel for lifeact). The

mean junctional intensity (AU-arbitrary units) of individual junctions in the resulting images was then

calculated and averaged.

36

Page 47: applying optical tweezers in vivo: a biophysical study of mechanical

CHAPTER 3

ESTABLISHING OPTICAL TWEEZERS

AS AN IN VIVO TOOL TO UNDERSTAND

CELL-CELL INTERFACE MECHANICS

37

Page 48: applying optical tweezers in vivo: a biophysical study of mechanical

38

Page 49: applying optical tweezers in vivo: a biophysical study of mechanical

3.1 Motivation

Optical tweezers and microrheology remain the most promising techniques in vivo due to their non-

invasive nature and precision. Previous work in our lab was directed towards developing and applying

these techniques to measure directly the forces involved in germband elongation. But these techniques

remain limited by the requirement of external probes (typically beads). One way to overcome this

limitation is to use natural probes like lipid granules or vesicles. Since germband elongation is

immediately preceded by cellularization, it is possible to inject classical microsphere probes at the end of

cellularization and study their dynamics during gastrulation. This however limits the success of the

measurements to the few cell-cell interfaces where the beads successfully localize without interacting

with the cellular cortex. I tried to overcome this limitation by developing a new approach to probe cell-

cell interfaces directly in Drosophila embryos.

39

Page 50: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 3.1 Characterizing the deflection of cell-cell interfaces imposed by optical tweezers

(A) Schematic of the setup: the embryo is optically sectioned by a light sheet and imaged, while a laser

trap (red) allows manipulation. Top image shows the epithelium labeled by a membrane marker

(GAP43::mcherry) and the laser trap position (marked by a yellow arrowhead). (B) Schematic of

deflection with distribution of forces and overlaid as well as separate images of the interface in mean and

the extreme positions of deflection. Scale bar: 5µm. (C) Representative plot of deflection versus time

showing both trap (red solid line) and interface positions (black solid line). (D) Interface deflection as a

function of laser trap power. (E) Interface deflection plotted against junction size (stage 6 end)

40

Page 51: applying optical tweezers in vivo: a biophysical study of mechanical

3.2 Results

3.2.1 Characterizing the deflection of cell-cell interfaces imposed by

optical tweezers

The mechanics of cell-cell interfaces was probed using a setup combining optical tweezers and light sheet

microscopy (Fig. 3.1A). This setup allows imaging of the embryonic epithelium at a high acquisition

rates, while manipulating objects in vivo. The laser trap was custom built by me and is produced by a

near-infrared laser light focused by the collection objective lens into the sample and is moved by

galvanometric mirrors in the plane of the epithelium (Fig. 3.1A and Fig. 2.1). Optical tweezers

experiments usually require a probe (glass or a polystyrene bead) to apply forces on attached molecules or

structures. This can be problematic for in vivo experiments which would require the probes to be

injected externally. I found that the cell-cell interfaces can be manipulated directly, without the need of an

external probe (Fig. 3.1B. Inset showing 3 snapshots of deflected membrane interfaces and trap

position.). Applying the laser directly, I imposed a sinusoidal movement to the trap perpendicular to a

cell interface and centered on it, and imaged the resulting deflection in the adherens junction plane

(Fig. 3.1C). The interface deflection followed the trap movement with lower amplitude, indicating that

the interface resists the mechanical load imposed by the laser trap (Fig. 1C).

The laser power was then varied while keeping trap sinusoidal amplitude �� (0.5µm) and the period of

oscillation (2s) constant. I found that �� also increases linearly with the laser power up to 300 mW (Fig.

3.1D). To explore the regime of deformation that the laser trap imposes to the interfaces, I then varied��, while keeping the period of oscillation constant (2s). The amplitude of the interface deflection ��, increases with ��, yet it deviates from a linear relationship for trap amplitude larger than 1 µm (Fig. 3.2b).

The results were similar even when the velocity was kept constant (2µm/s) by increasing the time period

with amplitude (Fig. 3.2a). Thus the variation is produced largely by the amplitude if the velocity is slow.

Together, these results imply that in the regime of small deformations (for an average interface length of

4.5 µm), the trap acts as a linear spring, whose stiffness �� is linearly proportional to the laser power.

Furthermore, �� does not vary with interface length at constant ��, time period and power (Fig. 3.1E).

Hence all further experiments were carried out within this range of deformation (<1 µm trap amplitude)

and with power 200 mW, unless otherwise stated.

41

Page 52: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 3.2 Interface deflection as a function of trap movement amplitude at, a) constant velocity (2µm/s)

and b) constant time period (2s). Interface deflection (c) Interface deflection as a function of the trap

oscillation period. (amplitude: 0.5µm, period: 2s). (fitting in b) and c) by Raphael Clement)

42

Page 53: applying optical tweezers in vivo: a biophysical study of mechanical

The resistance to deformation can arise not only from the mechanical properties of the interface and its

cortical elements, including the acto-myosin cytoskeleton, but also from the viscous drag force exerted by

the cytosol. To determine whether the resistance to deformation is time-dependent, we varied the period

of oscillation, while keeping the trap amplitude constant (amplitude At = 0.5 µm, Fig.3.2c). For periods

larger than or equal to 1 s, which correspond to mean speeds smaller than 2 µm.s-1 the amplitude was

constant. We observed that for periods smaller than 1 s (speed > 2 µm.s-1), the deflection was reduced

(Fig. 3.2c), which is the characteristic signature of viscous damping, presumably related to the viscous

drag in the cytosol.

43

Page 54: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 3.3 Offset between laser and interface positions

a) Schematic of trap-interface displacement curve, defining initial offset (ini_off), final offset (fin_off)

and interface movement (int_mov)

b) Final offset plotted as a function of initial offset

Interface deflection amplitude as a function of c) final position offset (data from 1,2,4 and 5 second

periods for 0.5 µm trap amplitude were pooled together, red: linear fit, blue: data points for individual

measurements),and d) initial offset.

(for c and d, measurements with initial offset greater than 600 nm were rejected. Negative values of final

offset in b) represent movement away from the laser. For c), all values were taken as positive)

ini_off

fin_off

int_mov

44

Page 55: applying optical tweezers in vivo: a biophysical study of mechanical

3.2.2 Dynamics of cell-cell interface interaction with optical tweezers as

a measure of position offset and phase lag

While the trap is positioned on the junction during the experiments there can be an uncertainty of few

100nm (pixel size of camera is 194nm). Therefore it was important to track the position of the trap with

respect to the interface during the experiments. It was observed that the interface often does not oscillate

at its rest position but instead acquires a new position to oscillate (Fig. 3.3a). Therefore I defined initial

(ini_off) and final offset (fin_off), as the distance between the trap and interface mean positions before

and during the experiments (Fig. 3.3a). Addition of ini_off and fin_off gives the interface movement

(int_mov) during the experiment. Fin_off varies linearly with ini_off (Fig. 3.3b). For small initial offset

values (<600nm) the final offset remained between 0-500nm. Importantly the final offset tended to be

positive in most cases indicating tendency of the interface to move towards the trap (negative final offset

indicates movement away from the trap, in the example in 3.3a, fin_off is positive). Assuming a trap size

and cell-cell interface thickness of few 100nm, this implies that the interface is always under the influence

of the trap during the experiments. Nevertheless final offset can have a small but measurable influence on

interface deflection amplitude �� (Fig. 3.3c). As fin_off increases, �� decreases linearly. For final offset

values within 500nm we see a variation in �� of ~500nm. It is possible that this variation might arise to

some extent due to variation in interface tension. Nevertheless we do not see any correlation between

ini_off and ��(Fig. 3.3d).

Since we observed a weak time dependence of �� it was interesting to look at the phase lag between trap

and interface oscillation. A purely elastic behavior would show 0 phase lag while a purely viscous

behavior can show a phase lag of 90°. Between 2-5 seconds of time period of oscillation, phase lag

remains consistently between ~5-10° (Fig. 3.4 a) implying minimal viscous drag. Experiments below 1

second were hampered by noise in the interface oscillation. But at 1 second of time period we do see a

larger scatter in phase lag (Fig. 3.4a). We then looked at the variation of phase lag with int_mov (Fig.

3.4b) and final offset (Fig. 3.4c). Interestingly, phase lag increases linearly with int_mov but seems to

decrease with increasing final offset for small values of fin_off (<300nm).

45

Page 56: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 3.4 Phase lag between laser and interface movement as a function of a) time period (a), b)

interface movement (Initial offset – final offset) and c) final offset. (The mean trap position is defined as

the zero axis. For b and c, data with net movement <100nm and/or final offset larger than 700nm were not

taken into consideration. All values were taken as positive for plotting) Error bars are standard deviation.

46

Page 57: applying optical tweezers in vivo: a biophysical study of mechanical

3.2.3 Relaxation of interface deformation

We performed pull-release experiments, which have been used in vitro on single cells with optical and

magnetic tweezers (Alenghat et al. 2000) but have never been applied in vivo. Pull-release

experiments consist of switching the laser trap on/off at a few hundred nanometers distance from the

junction and then monitoring the deflection of the cell-cell interface, both towards (trap on) and away

from the trap (trap off) (Fig. 3.5a). The pull-release curves obtained show that the dynamics are not

purely exponential, and exhibit at least two characteristic times, in the range of 1 s and 10 s (Fig. 3.5a).

3.2.4 Propagation of interface deformation

A challenging question in tissue morphogenesis is whether local forces produce long-range deformation,

and at what speed the mechanical information propagates. We imposed the local deflection of a cell

interface using sinusoidal oscillations, to see how single cell deformation propagates throughout the

tissue. We tracked the deflection of neighbouring interfaces away from this point. For that purpose, we

plotted kymographs along lines perpendicular to cell interfaces (Fig. 3.5b). We oscillated the target

interface using a deflection amplitude of 1 ± 0.1 µm (Fig. 3.5b, bottom panel). The neighbouring

interfaces within a distance of 1 to 2 cells also deflected periodically, but with much lower amplitudes and

a small phase shift (Fig. 3.5e), top panel). This indicates that the deformation typically decays over a

distance of 1-2 cells.

Having observed the local propagation of interface deformation, we then wondered if we could deform

the interface without trapping it directly. For this purpose we introduced the trap in the cytoplasm of the

cell (Fig. 3.6a) and oscillated it at a 2 second period. With a cell size of ~6-7 µm, it is quite likely that one

junction would be closer to the trap than the others. Moreover we used a trap displacement of 1-3 µm. In

these conditions, the junction which deformed the most was defined as i (assuming it was closest to the

trap). The displacement of this junction was used as reference to also track ii, iii and iv. All the junctions

oscillate in response to the trap movement with an amplitude range of 50-200nm although the amplitude

of iv is significantly reduced (Fig. 3.6c). As the trap amplitude is increased, the interface amplitude

increases reaching a value of 50% of a normally trapped interface at 1.5 µm (Fig. 3.6c). Interestingly the

interface movement is much larger compared to deflection amplitude (Fig. 2b, negative is movement

away from trap) but shows no net movement towards the trap on average. We also measured the phase lag

of junctions ii and iii in comparison to i (Fig. 3.6d). There is a significant phase lag at 0.5 µm trap

amplitude (20-30°) which decreases linearly with increase in amplitude.

47

Page 58: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 3.5 Mechanical model of the interface and tissue response (modified from Bambardekar et al. in

press. Analysis, schematic and model by Raphael Clement) a) Deflection of the interface in a pull-release

(trap on - trap off) experiment. The model (blue line) fits accurately the experimental data (black). Notably,

the relaxation is not a simple exponential. The simplest analogous visco-elastic model (inset): a maxwell arm in parallel with a spring, moving in a fluid of viscosity η (the cytosol). b) Deflection perpendicular to

the interfaces is tracked in time along lines perpendicular to cell interfaces (red), as a kymograph, which

allows measuring deformation away from the targeted interface. c) A viscosity decrease in the simulations

results in increased propagation, whereas a viscosity increase results in decreased propagation. d) Overlay

of the rest (purple) and deformed (green) tissue in the experimental (left) and simulated (right) tissue. e)

Spatial decay of interfaces deflection over the neighbouring cells. The model quantitatively predicts the

decay.

48

Page 59: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 3.6 Interface deflection induced by cytoplasmic trap. a) Schematic of experiment. Cell-cell

interfaces in violet. Trap (red) is oscillated in the apical cytoplasm (black arrows). The junctions

perpendicular to the trap movement get drawn towards the trap. The junction that deforms the most is

taken to be I as reference for identifying neighbouring junctions. b) Interface deflection amplitude at

different amplitudes of trap excursion in the cytoplasm. c) Interface movement at different trap

amplitudes. d) Phase lag between junctions (i and ii/iii) at different trap amplitudes. Error bars are

standard deviation.

49

Page 60: applying optical tweezers in vivo: a biophysical study of mechanical

3.3 Discussions

3.3.1 Origin and nature of in vivo optical forces

Figure 3.7 (A) Quantitative Phase Microscopy Image obtained on early Drosophila embryo (Stage 6).

The method uses a transmission light microscope and quadriwave lateral shearing interferometry as

described in (Bon et al. 2009) . The epithelial cells are observed in cross-section (cartoon, top) Scale bar, 5

µm (white). The calibration bar shows the optical path difference in nm. (B) A plot profile along a line

(red line in (A)) shows that the optical path difference is larger at cell interfaces than inside the cells.

Given that the line defines positions where the geometrical thickness of the embryo is constant, this

indicates a refraction index increase at cell interfaces. Arrows of different colors mark the position of 3

interfaces.

Optical trapping results from a positive refractive index mismatch between the trapped object and the

surrounding medium (Ashkin 1997). Indeed, trapping of particles within the cytoplasm has

been previously reported (Ashkin 1989, Welte et al. 1998). Furthermore the lipid membrane can be

trapped and manipulated directly, however the forces produced are very weak. Thus the membrane itself

is unlikely to be responsible for the deformation observed in the interface. To check if the interface

has a positive refractive index compared to the cytoplasm, we performed quantitative phase imaging

of the epithelial cells (Fig.3.7). Indeed there is a positive mismatch of refractive index. Yet, the value of

this mismatch is difficult to determine because of the geometry of the tissue. Therefore, to estimate the

trap stiffness on the interfaces, the deformation produced by direct application of the focused laser was

compared with that induced by 0.46 µm diameter beads pushed against the cell-cell interfaces (Claire

Chardes, Fig. 3.8). The former was only by 2- to 3-fold larger than the latter (2.5 ± 0.4, mean±SD, 4

measurements), indicating that the trap stiffness on the interfaces was 2- to 3-fold smaller than that on50

Page 61: applying optical tweezers in vivo: a biophysical study of mechanical

beads. The trap stiffness on beads was 120 ± 50 pN.µm-1 at 200 mW laser excitation (20 measurements),

thus the trap stiffness on interfaces was estimated to be 50 ± 30 pN.µm-1, in the regime of small

deformations.

Figure 3.8 Interface deformation with and without beads. (A) Snapshots of an interface deformation

induced by a 500 nm diameter bead, moved by the laser trap. The red and green channels correspond to

two different positions of the trap separated by approximately 0.5 µm. (B) Positions of laser trap, bead

and interface in an oscillatory experiment at 100 mW laser power. (C) Positions of laser trap and interface

in the same conditions as in (B) in the absence of bead. (Claire Chardes)

Other than sinusoidal amplitude, another way to understand the optical forces produced in our

experiments, is the interface movement. Given that the trap-interface distance tends to have a positive

final offset is a clear indication that the forces produced are attractive (Fig. 3.3a,b). If such was the case

the interface should deform less with increase final offset during the oscillations, whereas the initial offset

should have no bearing by itself on interface amplitude. Indeed this seems to be the case (Fig. 3.3b and c).

Interestingly, placing the trap ~3 µm away from the interface still produced interface movement as well as

oscillation which also seems to propagate to the neighbouring junctions (Fig. 3.6b and c). This oscillatory

movement is only 2-4 fold smaller than when the interface is trapped directly (trap-interface distance 0-

600nm). For small deformations (trap excursion 1 µm) this might indicate a linear relationship between

interface amplitude and final offset (for a range of 0-3 µm), implying that the force produced is linear

51

Page 62: applying optical tweezers in vivo: a biophysical study of mechanical

akin to a spring. Furthermore, the propagation in this case is due to optical attraction rather than tissue

mechanics, as the interface further away from the trap but at a similar distance from interface i compared

to interface ii and iii shows a very weak oscillation at all excursions (Fig. 3.6c). Additionally the phase

lag between the interfaces decreases with increase in trap excursion (Fig. 3.6d) which one would expect in

this scenario since the trap varies its proximity to interfaces i and ii at different time points in the

oscillation (Fig. 3.6a). Large fluctuations in final offset for all interfaces (Fig. 3.6b) probably imply

stochastic mechanical changes produced in the cell by the cytoplasmic trap and need to be further

investigated to be understood properly.

Figure 3.9 Viscosity measurements obtained from trajectories of individual beads. (A) Image showing

100 nm diameter beads (red) injected in the embryo. Cell contours are labeled by E-cadherin::GFP. (B)

Single particle trajectories superimposed to an image of the cells. (C) Fraction of beads exhibiting

diffusive, subdiffusive and superdiffusive behaviours from analysis of the mean-square displacement

(using criteria as described in (32)). Trajectories were acquired at 38 Hz over a time from 13 to 26 s. (D)

Histogram of the viscosity coefficient deduced from the analysis of 1348 particles, exhibiting free-like

diffusion. (Olivier Blanc)

52

Page 63: applying optical tweezers in vivo: a biophysical study of mechanical

3.3.2 Towards a mechanistic understanding of the experiments

The constancy of the deflection amplitude at speeds below 2 µm/s indicates that the force applied by the

trap and that produced by the interface are in quasi-static equilibrium. Interestingly, this condition of low

speeds (< 2 µm/s) is always fulfilled during cell shape changes driven by Myo-II contractility during

tissue morphogenesis of Drosophila, indicating that short-time viscous damping should play no role in the

process.

In the quasi-static regime (deformation speeds < 2 µm/s), we can thus assume that the shape of the

interface mainly results from the balance of forces between the trapping force �� and the tension of the

interface �: �� = �� ��s �., where � is the angle that the interface makes with respect to the trapping

force (Fig. 3.1B). As the vertices of the cell-cell contact did not move significantly during the

deformation (Fig.3.11), we could neglect the contribution of other cells and use this simple local

equilibrium formula.

For small deformations (that is for maximal deflections much smaller than the initial junction length

l0), ��� � ≈ ���� and �� ≈ ��(�� − �), where � is the position of the interface and �� the position of the

trap. The tension of the interface thus approximates as: � ≈ ����� ���� − ��. We found that the ratio ��� remains almost constant during periodic oscillations (Fig. 3.2c), indicating that the pre-existing tension

of the interface is not significantly modified during small deformations. Importantly, this also implies that

tension measurements, while relying on geometrical and physical approximations, do not require a

mechanical model of cell contacts. Thus, tension values can be obtained as a simple linear function of the

ratio between the interface deflection and the trap position. From the estimated trap stiffness on the order

of 50 ± 30 pN.µm-1 (Fig. 3.8), and using the slope in Fig. 3.2c as a mean value for the ratio, tension at

cell-cell interfaces � is estimated on the order of 60 ± 40 pN. Given that a single molecular motor of

Myo-II at maximal load produces 5 pN of force, the range of tension measured here suggests that a net

imbalance of about a dozen of motors pulling on a cell interfaces could deflect it by a few hundred

nanometers. The tensions reported here are 2-3 orders of magnitude below cell-cell forces in cell

aggregates on adhesive substrates in vitro (Maruthamuthu et al. 2011)

We do see a measurable viscous damping effect at oscillation periods smaller than 1s (Fig. 3.2c). This is

further indicated from the increase in phase lag range at 1s (Fig. 3.4a, although measurements at smaller

periods could not be done). The increase in phase lag with interface movement (Fig. 3.4b) might be an

53

Page 64: applying optical tweezers in vivo: a biophysical study of mechanical

implication that the damping arises from the cytosol. To test whether damping was due to the viscosity of

the cytosol, we performed relaxation experiments after release of the trap, which provide an

Figure 3.10 Relaxation of interfaces after trap release. Initial speed of relaxation is related to the

friction coefficient. (Raphael Clement)

alternative measurement of tension. At the onset of relaxation, tension is balanced only by viscous

damping: �� ��� � = ����, where v0 is the initial relaxation velocity and �� is the friction coefficient.

Measuring v0 (Fig. 3.10) and using the mean tension value of 60 pN, this provided an indirect

measurement of ��, which is in the order of 1.2 ± 0.6 10-4 m.Pa.s. To determine viscosity, �� should be

rescaled by both a typical length scale (the junction length, 4.5 µm) and a geometric coefficient (=16 for a

circular disk, which we will use here to obtain an order of magnitude). The viscosity associated with the

observed damping is thus in the order of a few Pa.s. To confront this estimate to the actual viscosity of the

cytosol, beads of radius � = �� �� were injected in the cytosol (He B et al. 2014) and their mean square

displacement was measured (Olivier Blanc, Fig. 3.9). Relating the diffusion constant � to the viscosity �

by the Stokes-Einstein equation (��� = �����), where ��is the Boltzmann constant and � the

temperature, the viscosity of the cytosol is found to be 3.6 ± 0.1 Pa.s (mean ± SE, 1350 beads). This is

consistent with the order of magnitude found with the relaxation method, which confirms that damping

might indeed be caused by the viscous drag in the cytosol.

54

Page 65: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 3.11 Elongation of interfaces adjacent to the optically deformed interface. (Raphael Clement)

3.3.3 Developing a mechanical model (Raphael Clement)

Different types of visco-elastic models were considered to fit the experimental data and it was found that a

so-called standard linear solid (SLS) model, composed of a Maxwell arm (a spring and a dashpot) in

parallel with a spring and embedded in a viscous medium of viscosity � (associated with the cytosol), is

the best and simplest model to correctly account for the observed behavior (Fig. 3.5a; fit: solid blue curve,

model: inset). Importantly, this simple model also fits well the experiments using a sinusoidal movement

described above (Fig. 3.2b & c, solid blue curve). At these short time scales, the mechanical response of

an epithelial tissue should result both from the constitutive mechanics of its acto-myosin cortex and from

the viscosity of the cytosol. The timescale of 1 s is consistent with the drop of deflection amplitude

observed below 1 s in the oscillatory experiments (Fig. 3.2c), and can be attributed to the viscous drag in

the cytosol. Consequently, the existence of another time scale (10 s) suggests that the cortex itself is not

purely elastic but visco-elastic (hence the SLS model). Importantly, the visco-elastic dynamics

characterized here are on a short time scale (under one minute). Neither the experiments nor the model

concern the long-term dynamics (minutes to hours), which presumably implies creep and therefore fluid-

like behavior (He B et al. 2014).

55

Page 66: applying optical tweezers in vivo: a biophysical study of mechanical

Single junction

The mechanical model for a single junction is derived from the constitutive equation of the standard linear

solid model and from a force balance equation at the interface. The constitutive equation relates the

horizontal force � pulling back the interface, to the horizontal displacement � of the interface:

�̇ +��� � = (�� + ��) �̇ +

��+��� �,

where �� and �� are elastic parameters (N/m), � a viscous parameter (m.Pa.s) for the cortex and the dot

denotes a temporal derivative. At these very low Reynolds number, inertia can be neglected, and the

balance of forces at the interface then simply reads: � = ��( �� − �) − ���̇,

where �� and �� are the stiffness and position of the optical trap, respectively, and �� is the friction

coefficient of the interface in the cytosol. The first term on the right-hand side thus corresponds to the

force exerted by the optical trap, while the second corresponds to the viscous drag in the cytosol – and is

therefore proportional to the velocity �̇. This linear system can then be solved for any trap trajectory ��(�); in particular for our experimental conditions: a sinusoidal oscillation or a pull-release experiment.

Tissue scale

In the tissue scale simulations, the epithelium is considered as a network of bonds – the cell contacts –

between vertices. Each bond is considered as a visco-elastic segment. The constitutive equation of each

segment, similar to the first equation, is:

�̇ +�2� � = (�1 + �2) �̇ +

�1+�2� �,

where � is the tension increase and � is the length increase, � = � − �0. The motion of each vertex is then

computed using the force balance equation (Fig. 3.11). For the vertex i:

�������⃗̇ = � �������⃗�=���(�)

56

Page 67: applying optical tweezers in vivo: a biophysical study of mechanical

Notably, the midpoint of the target interface is treated as a vertex in the simulations. Its position is

imposed as to mimic the considered experiment. The rest of the vertices move according to the force

balance equation, therefore, their movement ultimately results from the deflection movement of the target

interface.

To evaluate the ability of this simple mechanical model to reproduce these data, the single junction

model was first transposed to a network of contacts, as observed in vivo. Each contact is

considered as a viscoelastic (SLS) element pulling on vertices, so each vertex motion results from the

balance between tensile forces and fluid friction (Fig. 3.5d). It was then possible to simulate the

mechanical response of the tissue to periodic manipulations in the extracted experimental geometry,

imposing only the sinusoidal displacement at the midpoint of the target interface. The kymographs in the

simulated tissues were plotted along the same lines as in the experiments (Fig. 3.5c, bottom panel).

The typical parameter values that was estimated from the single interface experiments faithfully

reproduced the tissue-scale observation: one or two neighbouring cells deform away from the source

point of deflection (Fig. 3.5). In contrast, taking higher or lower viscosity for the cytosol into

account leads to shorter and longer propagation distance, respectively, which does not correctly

reproduce the observed behavior (Fig. 3.5). The speed of propagation can be experimentally

estimated by the phase delay between the trapped interface’s deflection and that of its neighbors.

At a 1-cell distance (~7 µm), a time delay of 375 ± 125 ms was measured, which corresponds to a

propagation speed with a typical phase velocity of 20 µm/s. At a 2-cell distance, a mechanical

signal would propagate in less than 1 s. This is much faster than any diffusible chemical; for instance,

a protein diffusing at 1 µm2.s

-1 would explore the same distance in 100 s. This speed is also much

larger than the speed of acto-myosin flows, which are observed in different systems, including

Drosophila and C. elegans (0.1 µm/s) (Mayer 2010).

57

Page 68: applying optical tweezers in vivo: a biophysical study of mechanical

3.4 Summary

To summarize, cell-cell interfaces in Drosophila embryo can be directly manipulated with a laser trap.

This trapping arises from a positive refractive index mismatch between the interface and the cytosol. The

temporal and spatial kinetics of this manipulation were analyzed by sinusoidal excursions and pull-release

experiments. For small excursions and final offsets the force produced is linear and quasistatic. Both

manipulations reveal the role of viscous damping at time scales below 1 second.

Interestingly the sinusoidal excursion can propagate to neighbouring junctions. The propagation decays

within a distance of one cell-length and shows only a small phase lag. Cytoplasmic trap experiments

showed very different behaviour by comparison, indicating that the propagation seen after trapping

interface was mechanical in origin.

Using my experimental data a passive viscoelastic model was developed which successfully replicates the

findings from the experiments. Having standardized the method for a passive tissue (apical epithelium at

the end of cellularization), this sets the stage to understand mechanical forces in epithelial rearrangements

during gastrulation.

58

Page 69: applying optical tweezers in vivo: a biophysical study of mechanical

CHAPTER 4

PROBING MECHANICAL ANISOTROPY

DURING DROSOPHILA GERMBAND

ELONGATION

59

Page 70: applying optical tweezers in vivo: a biophysical study of mechanical

60

Page 71: applying optical tweezers in vivo: a biophysical study of mechanical

4.1 Motivation

Having established methodology and a viscoelastic model for passive epithelium (Chapter 3) we decided

to probe epithelial mechanics in early gastrulation of the Drosophila embryo. The early epithelium

consists of a simple sheet of cells that spread over the yolk and are in contact with each other through E-

cadherin-based adhesion. During early embryogenesis at the blastula stage just after the end of

cellularization, epithelial cells have very similar hexagonal shapes, suggesting that cell junctions have

similar mechanical properties and the internal pressure of these cells is homogeneous. At the later gastrula

stage (germband elongation), cells undergo shape changes at distinct regions in the embryo. On the

ventral side of the embryo, apical cell constriction of a few rows of cells drives tissue invagination

(Kolsch et al. 2007), while on the ventro-lateral side of the embryo, cell intercalation, a process

whereby cells exchange neighbors by polarized remodeling of their junctions, drives tissue

extension. Laser dissection of cortical acto-myosin networks at cell junctions in the ventro-lateral

tissue has shown that anisotropic distribution of Myo-II causes an anisotropic cortical tension (Rauzi et

al. 2008). However, the absolute values of tensile forces have not yet been measured, and more

generally, the mechanics of cell-cell interfaces in vivo is largely unknown. These questions are addressed

in the following section by local mechanical measurements at cell junctions during germband elongation.

61

Page 72: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 4.1 Tension at cell contacts before and during germband elongation in wild type embryos.

a) Schematic of tissue elongation and images at three different stages: before elongation (stage 5 end), at

the onset of Myosin-II accumulation at cell junctions (stage 6 end) and during tissue extension (stage 7)

showing the cell interface (pink, GAP43::mcherry), E-cadherin (green) and Myosin-II (red). b) Interface

deflection amplitude at the adherens junction plane, for different stages t-Test (* p-value<0.05 ) c)

interface deflectionfor different stages normalized to junction size. d) apical E-cadherin levels

(endocad::GFP) in different stages. e) apical myosin-II levels in different stages. t-test ( *** P<0.001)

Scale bar 10 µm. Error bars are standard deviation.

62

Page 73: applying optical tweezers in vivo: a biophysical study of mechanical

4.2 Results

4.2.1 Probing the anisotropy of mechanical forces in early gastrulation

Interface deflection amplitude was measured at different stages of Drosophila embryo. At stage end the

epithelium is finishing cellularization. At stage 6 end mesoderm invagination ends and germband

extension begins. Stage 7 is early germband extension (Fig. 4.1a). The amplitude of deflection of the

cell-cell interface is reduced by two-fold from the end of stage 5 to the onset of stage 7 (Fig. 4.1b). The

decrease in amplitude indicates an increase in tension which is accompanied by significant enrichment of

E-cadherin and Myosin-II (Fig. 4.1d and e). Comparing the tensions for different orientations at this stage

revealed that the cell-cell interfaces along the D/V axis with myosin enrichment (Fig.4.2d) are about 2.5

times more tensed than those along the horizontal axis which are enriched in cadherin and have less

myosin (Fig. 4.2 c and d), which corresponds to mean absolute tensions of 275 ± 137 pN and 108 ± 54

pN, respectively (Fig. 4.2a). We probed the tension of interfaces of lengths varying from 4 to 8 µm, but

normalizing the deflection amplitude to junction length leaves our conclusion unchanged (Fig. 4.1c, Fig.

4.2b).

Measurements of interface deflection by laser trap at different positions along the apico-basal axis showed

that during tissue morphogenesis, there is a gradual polarization of the tension along this axis (Fig 4.3d)

concomitant with development of apico-basal polarity of cadherin and myosin (Fig. 4.3a,b and c). By

comparison apical and basal deflection remains unchanged in stage 5 end. To see if cadherin clusters can

regulate tension in the absence of myosin we probed stage 5 end further in the lateral plane (Fig. 4.4a).

Cadherin is localized in the first ~0-7 µm at this stage (Fig. 4.4a). Hence all measurements in this range

were taken as apical. Basal measurements were made between 8-20 µm. Myosin activity is not seen in

this distance (at the end of cellularization myosin II was localized basally at the cellularization front, as

seen in SPIM as well as spinning disk imaging). We find that the interface deforms more basally

indicating a reduced tension (Fig. 4.4b).

Inhibition of Myo-II activity by injection of ROCK inhibitor resulted in a significant reduction of the

tension at cell junctions (stage 7, tensions reduced to 60 ± 40 pN) (Fig. 4.5c) concomitant with loss of

myosin (Fig.4.5a and b). By comparison, tension remains unchanged after ROCK injection in stage 5 end

(Fig. 4.5d). The value of tension in ROCK injected embryos at stage 7 is similar to tensions measured at

stage 5 end. This confirms that the typical 2-fold increase in tension measured between stage 5 (~60 pN)

and stage 7 (~108 (A/P) to 275 pN (D/V)) can be attributed to myosin-II activity.

63

Page 74: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 4.2 a) Interface deflection amplitude plotted at different stages for interface orientations along

the D-V axis. b) Interface deflection of junctions with different orientations normalized to junction size.

c) Quantification of apical E-cadherin intensity (endocad::GFP) and d) apical myosin-II intensity

(squash::GFP) as function of interface orientation in different stages. t-test (** P<0.01, *** P<0.001).

Error bars are standard deviation.

64

Page 75: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 4.3 a) Images of E-cadherin (endocad::GFP, green) and myosin-II (squash::GFP, red) in the

adherens junction plane and in a basal plane (3 µm below) at stage 7 b) Ratio of basal to apical interface

deflection amplitude at different stages c) Quantification of apico-basal E-cadherin intensity

(endocad::GFP) and e) apico-basal myosin-II intensity (squash::GFP) in different stages. t-test (* P<0.05,

*** P<0.001). Error bars are standard deviation. Scale bar, 10 µm.

65

Page 76: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 4.4 a) Images of E-cadherin (endocad::GFP, green) and interface (GAP43::cherry, pink) in the

lateral plane at stage 5 end. b) apical (0-7 µm from edge) and basal (8-20 µm) interface deflection along

the lateral plane in stage 5 end. c) Images of lateral interface at stationary and maximum deflection (red is

laser position). t-test (** P<0.01). Error bars are standard deviation. Scale bar, 10 µm.

66

Page 77: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 4.5 a) Images of E-cadherin (endocad::GFP, green) and interface (GAP43::cherry, red) in

control(WT) and ROCK inhibitor injected (ROCK) embryos in the adherens junction plane at stage 7. b)

Quantification of apical myosin-II intensity (squash::GFP) in WT, water and ROCK embryos at stage 7.

c) Interface deflection amplitude in the adherens plane in control (WT), water-injected (water) and ROCK

inhibitor injected embryos at stage 7. c) Interface deflection at stage 5 end in WT and ROCK embryos. t-

test (*** P<0.001). Error bars are standard deviation. Scale bar, 10 µm.

67

Page 78: applying optical tweezers in vivo: a biophysical study of mechanical

4.2.2 Effect of actin perturbation on interface tension

Since cadherin dynamics in germband elongation are influenced by actin activity (Cavey et al. 2008),

we looked at actin concentration in different stages. We observed that actin is strongly enriched

apically at the beginning of cellularization (stage 5 beginning), followed by a significant decrease at

stage 5 end (Fig.4.6a and b). Actin enriches again at stage 7 but not to the same levels as stage 5 (Fig.

4.6a and b). Comparing the interface deflection reveals a significantly higher tension in stage 5

beginning compared to stage 5 end and 7. To confirm if actin cortex is responsible for this difference

in tension, we perturbed actin levels by injecting embryos with cytochalasin D (which cuts actin

filaments) or latrunculin A (which blocks actin polymerization), both of which perturb the actin cortex.

Wildtype embryos in stage 7 do not show any apicobasal or D-V polarization of actin distribution (Fig.

4.7a and c). Perturbed embryos show a decrease in actin that is much more significant in latrunculin

treatment (Fig. 4.7b) than cytochalasin (Fig. 4.7c). Interestingly, cytochalsin induces a much stronger

depletion basally than apically (Fig. 4.7c). Deflection amplitude increases after cytochalasin

treatment but perplexingly remains unchanged in latrunculin treatment (Fig. 4.7d). Examining the

interface movement between initial and final offset, however, shows significantly larger movements

after actin perturbation (Fig. 4.7e, fig. 4.9d snapshots).

68

Page 79: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 4.6 a) Images of lifeact (lifeact::GFP) in apical plane at stage 7, beginning (stage 5 beg) and

end of cellularization (stage 5 end). b) Quantification of apical lifeact intensity (lifeact::GFP) in different

stages. c) Interface deflection amplitude in the apical plane at different stages. t-test (*** P<0.001, **

P<0.01, * P<0.05). Error bars are standard deviation. Scale bar, 10 µm.

69

Page 80: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 4.7 a) Apico-basal images of lifeact (lifeact::GFP) in control (WT), cytochalasin D (cyto D)

and latrunculin A (lat A, apical only) injected embryos in apical plane at stage 7. b) Quantification of

apical lifeact intensity (lifeact::GFP) in WT and lat A stage 7 embryos. c) Quantification of apico-basal

lifeact intensity in WT and cyto D stage 7 embryos. d) Interface deflection amplitude in the apical plane

of different stage 7 embryos. e) Interface movement (intial-final offset) in WT, cyto D and lat A stage 7

embryos. t-test (*** P<0.001, ** P<0.01, * P<0.05). Error bars are standard deviation. Scale bar, 10 µm.

70

Page 81: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 4.8 Phase lag and offset in different embryos a) Phase lag between trap and interface

movement in stage 5 end, stage 6 end, stage 7, ROCK (stage 5 end and stage 7), cyto D and lat A (stage

7) b) Interface movement (intial-final offset) in WT, cyto D and lat A stage 7 embryos. c) Final offset

between interface and trap position in different embryos. Error bars are standard deviation.

71

Page 82: applying optical tweezers in vivo: a biophysical study of mechanical

4.2.3 Temporal and spatial dynamics: Position offset and phase lag

measurements in different stages and perturbations

To ensure that all experiments were done under similar conditions, we measured the final offset for

different stages (Fig. 4.8c). We find that the offset does not vary significantly in different stages and

treatments (except ROCK injected at stage 5 end). To verify if interface movement in Fig. 4.7e is due to

large initial offset values, we measured interface movement in stage 5 end, stage 6 end and stage 7 (as

well as ROCK injected embryos) and found no difference (Fig. 4.8b). To look at influence of viscous

damping on dynamic mechanical behavior we measured phase lag in different stages (Fig. 4.8a). Phase

lag showed a much larger range in stage 7 while the rest were all within the same range. Since we

observed massive interface movement after cytochalasin treatment, we looked at this further by

successively increasing the deflection period on the same interface, in wildtype and cytochalasin

perturbed junctions stage 7 (Fig. 4.9 a,b,c). Time dependence in wildtype stage 7 embryos is similar

to stage 5 end (Fig. 4.9d and Chapter 3). In wildtype, the interface typically moves towards the trap but

tries to relax to its initial position when the trap is switched off between successively increasing

periods of deflection (Fig. 4.9a). For cytochalasin, we observe 2 distinct behavior, with some

junctions deforming (cyto D large) much larger than others (cyto D small). Cyto D large shows increase

in interface amplitude with increasing periods (Fig. 4.9d). Moreover it shows massive movements of

the interface towards trap position which subsequently fail to recover to the initial position. (Fig. 4.9c).

For cytochalasin, regions of low actin levels and patches/clusters with higher actin levels are observed

(Fig. 4.7a). Cyto D small also shows significant interface movement which does not recover but

exhibits a much smaller interface deflection amplitude (Fig. 4.9b). (Note in 4.9 b and c, initial offset

is >1 µm, at this offset interface in wildtype embryos show very little interface movement and

decreased amplitude as described in Chapter 3).

72

Page 83: applying optical tweezers in vivo: a biophysical study of mechanical

Figure 4.9 Creep and time dependence after cytochalsin D treatment a) Examples of interface

movement with increasing time periods in cyto D large (drastic movement), water and cyto D small

(small interface movement) stage 7 embryos. (blue=junction, red=trap. Images: interface marked with Gap43::cherry at corresponding time points) b) Interface deflection amplitude in stage 7 water (blue) and

cyto D large (black, only interfaces with large difference > 0.6 µm of initial and final offset analyzed)

injected embryos as a function of increasing time period (s) applied to individual interfaces.

73

Page 84: applying optical tweezers in vivo: a biophysical study of mechanical

4.3 Discussions

4.3.1 Interface tension is regulated by myosin II dynamics

Using direct optical manipulation, I probed how cell-cell tensions change during tissue morphogenesis, in

early germband of the Drosophila. Before gastrulation movements, at the end of stage 5, cells form a

regular lattice with isotropic shapes and isotropic distribution of Myo-II (Fig. 4.1a, 4.2d). Later, at the end

of stage 6 and the onset of stage 7, the total concentration of Myo-II increases at adherens junctions, and

its distribution becomes anisotropic, with higher levels along interfaces parallel to the D/V axis (Fig. 4.1a,

4.2d) (Bertet et al. 2004, Rauzi et al. 2008). This anisotropic distribution of Myo-II has been shown to

drive polarized junction shrinkage and cell intercalation The anisotropy of tension in different stages that

is observed is consistent with the estimates of the relative values of tension from laser nanodissection

(Rauzi et al. 2008). My method thus provides a direct measure of tension with low perturbation.

Moreover we establish further the development of tension polarity regulated by cadherin and myosin

polarity (Fig. 4.2). While normalizing with interface length does not change the trend, it enhances the

difference in tension from stage 5 to 7. This is probably because the analysis was restricted to junctions

between ~4-8 μm. In germband shrinking junctions are stiff while elongating junctions are more

deformable. Therefore the increase in tension difference up on normalization implies that our

measurements were somewhat biased towards elongating junctions.

The E-cadherin junctions, where the experiments were performed, begin to localize apically

during gastrulation. At the end of cellularization junctions are spread over a few microns below the

apical cortex (data not shown) but become restricted to a thin 1 µm section and are localized about 1-2

µm below the apical surface during germband (Harris et al. 2004, TruongQuang et al. 2013) (Fig. 4.3a,

left panel, 4.3b). As myosin-II accumulates at the adherens junction plane at the end of stage 6 and

onset of stage 7 (Fig. 4.3a, right, 4.3c), this might translate into different mechanical properties at the

adherens junction plane as compared to more basal positions. Indeed we do observe an apico-basal

polarity of tension. (Fi.g 4.3). Interestingly, apico-basal polarity precedes gastrulation and resulting

myosin enrichment, when I take the apical displacement of cadherin into account (Fig. 4.4). In this

experiment the apical measurements were extended upto the first 7 µm, to take into account the spread

of cadherin clusters at the end of cellularization, whereas basal measurements were at 8-20 µm. Myosin

is only present at the basal cellularization front at this stage (~25-30 µm). This implies that interaction

of E-cadherin clusters with the actin cortex might be sufficient to influence tension across the entire

columnar cell. However the mechanism involved could be different from the apico-basal anisotropy

74

Page 85: applying optical tweezers in vivo: a biophysical study of mechanical

4.3.2 Actin cortex is responsible for interface tension

Experiments at stage 5 beginning, end and stage 7, clearly show that interface tension correlates in

average with actin intensity at the interface (Fig. 4.6a, b and c). This is in accordance with the general

view that that the cell surface tension is controlled by the actin cortex (Bendix et al. 2008). Notably I

did not observe actin polarization in stage 7 (Fig. 4.6 and 4.7), which is in contrast with the previous

results (Blankenship et al. 2006). Since the marker I used (lifeact) is not as reliable as phalloidin (used

by Blankenship et al.) it could simply be an experimental error. Therefore, it is essential to repeat the

experiments, possibly with phalloidin for comparison. It has been shown that actin exists in 2

different pools in germband epithelia in Drosophila, a smaller pool associating with cadherin

junctions, regulating its stability and a larger pool regulating its mobility (Cavey et al. 2008).

Notably the smaller pool is resistant to latrunculin treatment, but its response to cytochalasin in

unknown. It might be crucial to look at the dynamics of these pools during the experiments as an

alternative explanation for the observed discrepancy.

The large scatter in cyto D interface amplitude as well as the smaller amplitude in latrunculin A is

perplexing (Fig. 4.7d). However actin perturbation experiments are complicated by massive interface

movement (fig. 4.7e, fig. 4.9c initially as well as during the experiments, this might represent the 2 time-

scales shown in Chapter 3, Fig. 3.5a). Branched actin networks can be stiff enough to sustain forces

produced during leading edge extension of motile cells or in vitro actin-based propulsion of beads (Marcy

et al. 2004, Chaudhuri et al. 2007). But they can rupture or tear when stretched beyond a certain limit

(Paluch et al. 2006). Particularly for initial offsets of >1μm (cyto D, fig. 4.9 b and c, latrunculin A

observed but not shown), interface movement can exceed 1 μm. This initial movement clearly implies

that the trap does produce considerably more force on the cell-cell interface than in water-injected

embryos. However interface amplitude was not dependent on interface movement (data not shown). The

existence of 2 responses to cytochalasin (cyto D small and large, fig. 4.9 b and c) might shed some light

on it. Particularly for large initial offsets in fig. 4.9 c, the interface moves massively in both scenarios but

75

seen in germband extension. While cadherin intensity in germband extension correlates inversely with

interface stiffness, dissecting the role of cadherin from that of myosin is not straightforward. To

investigate any apparent causal relation between cadherin organization and interface tension would

require further experiments that perturb the cadherin levels directly (for example alpha-catenin RNAi).

Page 86: applying optical tweezers in vivo: a biophysical study of mechanical

Given that time dependence of interface amplitude does not change from stage 5 to stage 7

(Chapter 3 and Fig. 4.9 d), this implies that viscous damping remains unchanged. Therefore the

apparent increase in phase lag at stage 7 (Fig. 4.8a) arises from another mechanism, possibly

from the tissue movement induced by germband elongation. This is further indicated from the

observation that injection with ROCK, cytochalasin D or latrunculin A blocked germband extension

(data not shown) in my experiments and exhibit similar phase lag between trap and interface to stage 5

(fig. 4.8a). The massive increase in interface movement, amplitude and failure to recover intial

position, observed in cyto D large and small (Fig. 4.7e, Fig. 4.9 b and c) imply a creep-like behavior.

Most of the movement is localized at the point of trap-interface contact (Fi.g 4.9 b and c snapshots)

compared to wildtype where movement spreads across the interface length (Fig. 4.9a snapshot). This

might be due to clusters of higher actin density produced by cytochalasin D along with cortex

disruption (Fig. 4.7a). Large movements are also observed in latrunculin A despite small average

amplitude, though not to the extent seen in cytochalasin D (Fig. 4.7e, Fig. 4.7d). Since the interface is far

from its initial position in these measurements, the behavior may not be quasistatic and direct

interpretation of tension from deflection is not possible. In this scenario, the pull and release experiment

(Fig. 3.5a) would be more suitable, both to understand interface tension and creep (especially over

longer time-scales). We do see some change in interface amplitude for successive cycles at the same

time period in cyto D large (Fig. 4.9c). Whether this indicates sinusoidal cyclic loading or is a

consequence of the constantly moving interface (Fig. 4.9c) is unclear.

76

4.3.3 Creep and sinusoidal loading

the response is faster in cyto D large. After this point the interface shows either large or small deflection

amplitudes. It might be that in cyto D large, the cortex is stretched beyond its limit, whereas in cyto D

small it continues to resist deformation.

Page 87: applying optical tweezers in vivo: a biophysical study of mechanical

4.4 Summary

Here I showed that optical tweezers can be robustly used to measure variations in tension at cell contacts

arising from tissue movements in vivo, and that this method can reveal planar-polarized as well as

apicobasal anisotropies of tension in a developing organism. The role of myosin dynamics in establishing

this polarity is established further from my results. I also observed an apico-basal polarization of tension

with changing cadherin levels in the absence of myosin dynamics, although the causality of this

correlation was not established. The experiments here, furthermore, confirmed that the actin cortex is

responsible for interface tension, especially in the absence of cadherin junctions and myosin

activity. Actin disruption can induce creep-like behavior in interface mechanics. Further work in

this direction, particularly with push and pull experiments will help in establishing a dynamic

viscoelastic model.

77

Page 88: applying optical tweezers in vivo: a biophysical study of mechanical

78

Page 89: applying optical tweezers in vivo: a biophysical study of mechanical

CHAPTER 5

CONCLUSIONS AND PERSPECTIVES

79

Page 90: applying optical tweezers in vivo: a biophysical study of mechanical

In this thesis, I have studied the mechanical properties of cell-cell contacts during tissue morphogenesis

and how local deformation propagates within a tissue. This work provides absolute values of tensions at

cell interfaces, while previous work estimated relative values based on assumptions of the viscous

properties at cell junctions. Surprisingly, the small forces produced by optical tweezers are sufficient to

produce significant deflection of cell interfaces and the tension at cell interfaces was estimated to be in

the range of tens to hundreds of pN during early stages of tissue morphogenesis. This suggests that the

forces that remodel cell-cell contacts during tissue morphogenesis and drive the shrinkage or extension of

cell contacts can be powered by a small number of molecular motors. Fluctuations in cell shape, which

are observed during these events, might thus result from stochastic fluctuations in motor numbers.

My work allowed designing a predictive mechanical model of cell contacts. Mechanical modeling of

epithelia is crucial for understanding epithelial morphogenesis events; suggested as early as 1981 by

Odell’s pioneering work (Odell et al. 1981). Since then, a variety of mechanical descriptions have been

proposed; however, testing of the underlying hypotheses has been limited due to the lack of in

vivo experimental tools. Notably, so-called vertex models (Fletcher et al. 2014), usually based on

energy minimization, do not incorporate energy dissipation and thus cannot predict the tissue

dynamics. Here a model is proposed, which does incorporate a viscoelastic constitutive behavior.

Therefore, this model incorporates the vertex models and more recent continuum mechanics

approaches (Bonnet et al. 2012, Hutson et al. 2009). In addition, it captures the non-trivial two-

timescale relaxation dynamics evidenced by pull-release experiments. The possibility of absolute

tension measurements at cell contacts might be beneficial to force inference methods (Chiou et al. 2012,

Cranston et al. 2010, Ishihara et al. 2013), which provide relative tensions based on the geometry

of the contact network. It might indeed allow experimental validation of the inference, but also

be used to calibrate the inferred tensions. The analysis was intentionally restricted to time scales and

speeds faster than the changes in contractility in order to deal with steady shape patterns. The

approach established here could allow the exploration of additional time scales and to probe long-term

plastic deformation of cell contacts.

One important aspect that cannot be overlooked is that our model is passive (mainly incorporating stage 5

end, a tissue with minimal acto-myosin activity or morphogenetic movements) and does not incorporate

active tissue properties such as the role of acto-myosin. Extending mechanical understanding to, for

example, germband elongation, was the motivation behind experiments in chapter 4. Particularly the

creep-like, actin dependent behavior that I observed needs to be studied carefully using push-pull

experiments. Simultaneous imaging of acto-myosin as well as cadherin dynamics in such experiments

could provide the crucial data missing for developing an active model.

80

Page 91: applying optical tweezers in vivo: a biophysical study of mechanical

Lastly, the trapping behavior needs to be better understood, especially the biological origin of refractive

index mismatch at the cell-interface. Understanding this will help better to extend this approach as a

general system-independent tool for understanding the mechanics of morphogenesis. Such approaches are

being developed in the lab currently. Particularly we have replicated the trapping behavior in different

optical and imaging setups (inverted microscope with spinning disk, optical tweezers with SLM) as

well as in another model system (C. elegans embryo). Additionally, having characterized the behavior

for a single trap, this analysis can be extended to novel configurations, for example multiple traps as

well to a Spinning disk setup equipped with a laser ablation system will provide better mechanical

understanding of morphogenetic events. The availability of FRET-based force sensors in Drosophila,

as well as their previous usage in literature (Cai et al. 2014) in conjunction with the approaches being

developed, provide an exciting perspective on the advancements to come in this emerging field that I

have advanced and contributed to.

81

Page 92: applying optical tweezers in vivo: a biophysical study of mechanical

ACKNOWLEDGEMENTS

I

This thesis has been an important part of my journey in understanding life as well as my

directions and goals over the last few years. I arrived in France 3 years ago with no knowledge

of French or Europe, as well as a crippling mood disorder (which was later re-diagnosed as

Aspergers). I owe much to Pierre-François for helping me through all the initial paperwork as well

as being patient with my chaotic and sloppy method of working. The various meetings with him

have been inspiring and motivational even if I had my own way of doing things. I'm grateful to

Olivier Blanc, Claire and Raphael for all the help with the work and data analysis, particularly the

programming which I never developed an interest in, until recently. Sebastien, for all the help

with the fly strains as well as biochemical protocols. All the members of the lab for providing a

helpful and safe environment to work in as well as the occasional bursts of non-scientific

conversations which have inspired me much in my artistic pursuit. IBDML has been a very good

environment to work, nestled in the forest of Luminy and its captivating environment. All the

members of the Lecuit team have been very helpful during the last few years, discussions with

Manos, Loic and Claudio were in particular, very insightful in the understanding and progress of my

work. Willi, Diego, Pierluigi, Sabrina have been good friends and support systems in addition to

all the scientific discussions. Outside of the department, the LABEX initiative and the various

interactions that it provided , has been very positive. Discussions with Pierre-Henri Puesch on Linux,

mindfulness and art have been life-changing and certainly play an important role in my future

pursuits. This thesis is dedicated to all my family of soul in the South of France. All my friends

from CCL for the helping hand during my initial social awkwardness. Hugo, Jess and Nico

for the artistic collaborations and much needed support during thesis writing. Cortical Systematics

Crew for helping me heal my paranoia and mistrust of people through the various events they

organized and I had the good fortune to be part of. And to all the reflections that I crossed my

paths with either in my journeys or through discovery of their work.

Page 93: applying optical tweezers in vivo: a biophysical study of mechanical

REFERENCES

Chapter 1

Alberts B., Bray D., Lewis J., Raff M., Roberts K., Watson JD. (2002). Molecular Biology of

the Cell, Garland Science Publ., NewYork.

Ashkin A. (1997) Optical trapping and manipulation of neutral particles using lasers. Proc. Natl.

Acad. Sci. USA 94:4853–4860.

Ashkin A., Dziedzic JM. (1989). Internal cell manipulation using laser traps. Proc. Natl. Acad. Sci.

USA 86:7914–7918.

Baum B., Georgiou M. (2011) Dynamics of adherens junctions in epithelial establishment,

maintenance and remodeling. J. Cell Biol. 192(6):907–917.

Bertet C., Sulak L., Lecuit T. (2004) Myosin-dependent junction remodelling controls planar cell

intercalation and axis elongation. Nature 429:667–671.

Bonnet I., Marcq P., Bosveld F., Fetler L., Bellauche Y., Graner F. (2012). Mechanical state,

material properties and continuous description of an epithelial tissue. arXiv:1205.2454.

Bray D. (2000) Cell Movements–From Molecules to Motility, Garland Science Publ., NewYork.

Burke-Smith A. Cellular organization of tissues. MCD Tissues.

Burridge K., Chrzanowska-Wodnicka M. (1996) Focal adhesions, contractility, and signaling,

Annu. Rev. Cell Dev. Biol. 12 :463.

Cai D., Chen SC., Prasad M., He L., Wang X., Choesmel-Cadamuro V., Sawyer JK., Danuser G.,

Montell DJ. (2014) Mechanical feedback through E-cadherin promotes direction sensing

during collective cell migration. Cell 157:1146–1159.

Cavey M., Rauzi M., Lenne PF., Lecuit T. (2008) A two-tiered mechanism for stabilization

and immobilization of E-cadherin. Nature 453:751-757.

Daniels BR., Masi BC., Wirtz D. (2006) Probing single-cell nicromechanics in vivo: the

microrheology of C. elegans developing embryos Biophys. J. 90:4712-4719.

Davidson L., Dassow MV., Zhou J. (2009) Multi-scale mechanics from molecules to morphogenesis.

Int J Biochem Cell Biol. 41(11): 2147–2162.

Fletcher AG., Ostefield M., Baker RE., Shvartsman SY. (2014) Vertex models of

epithelial morphogenesis. Biophys. J. 106:2291-2304.

II

Page 94: applying optical tweezers in vivo: a biophysical study of mechanical

Ghibaudo M., Saez A., Trichet L., Xayaphoummine A., Browaeys J., Silberzan P., Buguin A., and

Ladoux B. (2008) Traction forces and rigidity sensing regulate cell functions. Soft Matter

4(9):1836–1843.

Gorfinkiel N., Blanchard GB. (2011) Dynamics of actomyosin contractile activity during epithelial

morphogenesis. Current Opinion in Cell Biology 23:531–539.

Grashoff C., Hoffman BD., Brenner MD., Zhou R., Parsons M., Yang MT., McLean MA., Sligar SG.,

Chen CS., Ha T. (2010) Measuring mechanical tension across vinculin reveals regulation of focal

adhesions dynamics. Nature 466:263-266.

Grumbling G, Strelets V. (2006) FlyBase: anatomical data, images and queries. Nucleic Acids Res.

34:D484–8.

Harris AR., Daeden A., Charras GT. (2014) Formation of adherens junctions leads to emergence of a

tissue-level tension in epithelial monolayers. Journal of Cell Science 127:2507-2517.

He B., Doubrovinski K., Polyakov O., Wieschaus E. (2014) Apical constriction drives tissue-

scale hydrodynamic flow to mediate cell elongation. Nature 508:392–396.

Hormeno S., Arias-Gonzalez JR. Exploring mechanchemical processes in the cell with optical

tweezers. (2006) Biol. Cell. 98:679–695.

Hutson MS., Tokutake Y., Chang MS., Bloor JW., Venakides S., Kiehardt DP., Edwards GS.

(2003) Forces for morphogenesis investigated with laser microsurgery and quantitative modeling.

Science 300:145–149.

Ishihara S., Sugimara K., Cox SJ., Bonnet I., Bellauche Y., Graner F. (2013). Comparative study of

non-invasive force and stress inference methods in tissue. Eur. Phys. J. E Soft Matter 36:9859.

Kumar A., Shivshankar GV. (2012) Mechanical force alters morphogenetic movements and

segmental gene expression patterns during Drosophila embryogenesis. PLoS. one 7(3):e33089.

Kuo JC. (2013) Mechanotransduction at focal adhesions: integrating cytoskeletal mechanics in

migrating cells. J. Cell. Mol. Med. 17:704-712.

Ladoux B., Anon E., Lambert M., Rabodzey A., Hersen P., Buguin A., Silberzan P., Mege RM. (2010)

Strength dependence of cadherin-mediated adhesions. Biophys. J. 98:534-542.

Lecuit T., Lenne PF., Munro E. (2011) Force generation, transmission, and integration during cell and

tissue morphogenesis. Annu Rev Cell Dev Biol 27:157–184.

Lecuit T., Wieschaus E. (2000) Polarized insertion of new membrane from a cytoplasmic reservoir

during cleavage of the Drosophila embryo. J. Cell. Biol. 150(4):849-860.

Leidel, C., Longoria, R.A., Gutierrez, F.M., Shubeita, G.T. (2012) Measuring molecular motor forces

in vivo: Implications for tug-of-war models of bidirectional transport. Biophys. J. 103(3):492-500.

III

Page 95: applying optical tweezers in vivo: a biophysical study of mechanical

Lim CT., Zhoua EH., Quek ST. (2006) Mechanical models for living cells—a review. Journal of

Biomechanics 39:195-216.

Munoz JJ., Barrett K., Miodownik M. (2007) A deformation gradient decomposition method for

the analysis of the mechanics of morphogenesis. J Biomech 40: 1372–1380.

Odell GM., Oster G., Alberch P., Burnside B. (1981) The mechanical basis of

morphogenesis. Developmental Biology 85:446-462.

Panorchan P., Lee JSH., Daniels BR., Kole TP., Tseng Y., Wirtz D. (2007) Probing cellular

mechanical responses to stimuli using ballistic intracellular nanorheology. Methods in cell biology

8:115-140.

Perry JK., Lins RJ, Lobie PE and Mitchell MD. (2010) Regulation of invasive growth: similar

epigenetic mechanisms underpin tumour progression and implantation in human pregnancy. Clinical

Science (2010) 118:451–457.

Pullarkat PA., Fernandez PA., Albrecht O. (2007) Rheological properties of the eukaryotic cell

cytoskeleton. Physics Reports 449:29– 53.

Rauzi M., Lenne PF., Lecuit T. (2010) Planar polarized actomyosin contractile flows control epithelial

junction remodelling. Nature 468:1110–1114.

Rauzi M., Lenne PF. (2011) Cortical forces in cell shape changes and tissue morphogenesis. Current

Topics in Developmental Biology 95:93-144.

Rauzi M., Verant P., Lecuit T., Lenne PF. (2008) Nature and anisotropy of cortical forces orienting

Drosophila tissue morphogenesis. Nat. Cell. Biol. 10:1401–1410.

Rodriguez M., McGarry PJ., Sniadecki NJ. (2013) Review on cell mechanics: Experimental and

modeling approaches Applied Mechanics Reviews 65 (060801):1-41.

Solon J., Kaya-Copur A., Colombelli J., Brunner D. (2009) Pulsed forces timed by a ratchet-like

mechanism drive directed tissue movement during dorsal closure. Cell 137:1331–1342.

Truong Quang BA., Mani M., Markova O., Lecuit T., Lenne PF. (2013) Principles of e-cadherin

supramolecular organization in vivo. Current biology 23:2197–2207.

Wu PH., Hale CM., Chen WC., Lee JS., Tseng Y., Wirtz D. (2012) High-throughput ballistic injection

nanorheology to measure cell mechanics. Nat. Protoc. 5;7(1):155-70.

Wyczalkowski MA., Chen Z., Filas BA., Varner VD., Taber LA. (2012) Computational models for

mechanics of morphogenesis. J. Birth Defects Res. C Embryo Today 96(2):132–152.

Yonemura S. (2011) Cadherin-actin interactions at adherens junctions. Current opinions in cell

biology 23:515-522.

Zhu C., Bao G., Wang N. (2000) Cell mechanics: mechanical response, cell adhesion, and molecular

deformation. Annu. Rev. Biomed. Eng. 2:189–226.

IV

Page 96: applying optical tweezers in vivo: a biophysical study of mechanical

Chapter 2

Chardès C., Ménélec P., Bertrand V., Lenne PF. (2014) Setting-up a simple light sheet microscope

for in toto imaging of C. elegans development. J. Vis. Exp. 87:e 51342.

Ebbeson CL., Bruus H. (2012) Analysis of laser induced heating in optical neuronal guidance. J.

Neuroscience Methods 209:168– 177.

Huisken J., Swoger J., Del Bene F., Wittbrodt J., Stelzer EH. (2004) Optical sectioning deep inside

live embryos by selective plane illumination microscopy. Science 305:1007–1009.

Chapter 3

Alenghat FJ., Fabry B., Tsai KY., Goldmann WH., Ingber DE. (2000) Analysis of cell mechanics

in single vinculin-deficient cells using a magnetic tweezer. Biochemical and biophysical

research communications 277:93–99.

Ashkin A. (1997) Optical trapping and manipulation of neutral particles using lasers. Proc.

Natl. Acad. Sci. USA 94:4853–4860.

Ashkin A., Dziedzic JM. (1989) Internal cell manipulation using laser traps. Proc. Natl. Acad. Sci.

USA 86:7914–7918.

Bar-Ziv R., Moses E., Nelson P. (1998) Dynamic excitations in membranes induced by

optical tweezers. Biophys. J. 75:294-320.

Bon P., Maucort G., Wattellier B., Monneret S. (2009) Quadrilateral shearing interferometry

for quantitative phase microscopy of living cells. Optics Express 17:13080-13094.

Welte MA., Gross SP., Postner M., Block SM., Wieschaus EF. (1998) Developmental regulation

of vesicle transport in Drosophila embryos: Forces and kinetics. Cell 92:547-557.

He B., Doubrovinski K., Polyakov O., Wieschaus E. (2014) Apical constriction drives tissue-

scale hydrodynamic flow to mediate cell elongation. Nature 508:392–396.

Maruthamuthu V., Sabass B., Schwarz US., Gardel ML. (2011) Cell-ECM traction force

modulates endogenous tension at cell-cell contacts. Proc. Natl. Acad. Sci. USA 108:4708–4713.

Mayer M., Depken M., Bois JS., Julicher F., Grill SW. (2010) Anisotropies in cortical tension reveal

the physical basis of polarizing cortical flows. Nature 467:617-621.

V

Page 97: applying optical tweezers in vivo: a biophysical study of mechanical

Chapter 4

Bendix PM., Koenderink GH., Cuvelier D., Dogic Z., Koeleman BN. (2008). A quantitative

analysis of contractility in active cytoskeletal protein networks. Biophys. J. 94:3126–36.

Bertet C., Sulak L., Lecuit T. (2004) Myosin-dependent junction remodelling controls planar

cell intercalation and axis elongation. Nature 429:667–671.

Blankenship JT., Backovic ST., Sanny JS., Weitz O., Zallen JA. (2006) Multicellular rosette

formation links planar cell polarity to tissue morphogenesis. Dev Cell. 4:459-70.

Cavey M., Rauzi M., Lenne PF., Lecuit T. (2008) A two-tiered mechanism for stabilization

and immobilization of E-cadherin. Nature 453:751-757.

Chaudhuri O., Parekh SH., Fletcher DA. (2007) Reversible stress softening of actin networks.

Nature 445:295–98.

Chen H., Cohen DM., Choudhury DM., Kioka N., Craig SW. (2005) Spatial distribution and

functional significance of activated vinculin in living cells. J. Cell Biol. 169:459–70.

Harris TJC., Peifer M. (2004) Adherens junction-dependent and -independent steps in

the establishment of epithelial cell polarity in Drosophila. J. Cell. Biol. 167:135–147.

Kölsch V., Seher T., Fernandez-Ballester GJ., Serrano L., Leptin M. (2007) Control of

Drosophila gastrulation by apical localization of adherens junctions and RhoGEF2. Science

315:384–386.

Lecuit T., Lenne PF., Munro E. (2011) Force generation, transmission, and integration during cell

and tissue morphogenesis. Annu. Rev. Cell. Dev. Biol. 27:157–184.

Marcy Y., Prost J., Carlier MF., Sykes C. (2004) Forces generated during actin-based propulsion: a

direct measurement by micromanipulation. Proc. Natl. Acad. Sci. USA 101:5992–97.

Paluch E., Van der Gucht J., Sykes C. (2006) Cracking up: symmetry breaking in cellular systems. J.

Cell Biol. 175:687–92.

Rauzi M., Verant P., Lecuit T., Lenne PF (2008) Nature and anisotropy of cortical forces

orienting Drosophila tissue morphogenesis. Nat. Cell. Biol. 10:1401–1410.

Truong Quang BA., Mani M., Markova O., Lecuit T., Lenne PF. (2013) Principles of e-

cadherin supramolecular organization in vivo. Current biology 23:2197–2207.

VI

Page 98: applying optical tweezers in vivo: a biophysical study of mechanical

Chapter 5

Bonnet I., Marcq P., Bosveld F., Fetler L., Bellauche Y., Graner F. (2012) Mechanical state,

material properties and continuous description of an epithelial tissue. arXiv:1205.2454.

Cai D., Chen SC., Prasad M., He L., Wang X., Choesmel-Cadamuro V., Sawyer JK., Danuser G.,

Montell DJ. (2014) Mechanical feedback through E-cadherin promotes direction sensing

during collective cell migration. Cell 157:1146–1159.

Chiou KK., Hufnagel L., Shraiman BI. (2012) Mechanical stress inference for two dimensional

cell arrays. PLoS. computational biology 8:e1002512.

Cranston PG., Veldhuis JH., Narasimhan S., Brodland GW. (2010) Cinemechanometry (CMM):

A method to determine the forces that drive morphogenetic movements from time-lapse images.

Ann. Biomed. Eng. 38:2937–2947.

Fletcher AG., Ostefield M., Baker RE., Shvartsman SY. (2014) Vertex models of

epithelial morphogenesis. Biophys. J. 106:2291-2304.

Hutson MS., Tokutake Y, Chang MS, Bloor JW, Venakides S, Kiehart DP, Edwards GS.

(2003) Forces for morphogenesis investigated with laser microsurgery and quantitative

modeling. Science 300:145–149.

Ishihara S., Sugimara K., Cox SJ., Bonnet I., Bellauche Y., Graner F. (2013) Comparative study of

non-invasive force and stress inference methods in tissue. Eur. Phys. J. E Soft Matter 36:9859.

Odell GM., Oster G., Alberch P., Burnside B. (1981) The mechanical basis of

morphogenesis. Developmental Biology 85:446-462.

VII

Page 99: applying optical tweezers in vivo: a biophysical study of mechanical

1

Direct laser manipulation reveals the mechanics of cell contacts in vivo

Kapil Bambardekar1,*

, Raphaël Clément1,*

, Olivier Blanc1, Claire Chardès

1 and Pierre-François Lenne

1

1. Aix Marseille Université, CNRS, IBDM UMR7288, 13009, Marseille, France.

Corresponding author: [email protected]

* K.B. and R.C. contributed equally to this work

Abstract

Cell-generated forces produce a variety of tissue movements and tissue shape changes. The

cytoskeletal elements that underlie these dynamics act at cell-cell and cell-extracellular matrix contacts

to apply local forces on adhesive structures. In epithelia, force imbalance at cell contacts induces cell

shape changes, such as apical constriction or polarized junction remodeling, driving tissue

morphogenesis. The dynamics of these processes are well characterized; however, the mechanical

basis of cell shape changes is largely unknown due to a lack of mechanical measurements in vivo. We

have developed an approach combining optical tweezers with light sheet microscopy to probe the

mechanical properties of epithelial cell junctions in the early Drosophila embryo. We show that

optical trapping can efficiently deform cell-cell interfaces and measure tension at cell junctions, which

is on the order of picoNewtons (pN). We demonstrate that tension at cell junctions equilibrates

over few seconds; a short time scale compared to the contractile events that drive morphogenetic

movements. We further show that tension increases along cell interfaces during early tissue

morphogenesis and becomes anisotropic as cells intercalate during germ band extension. By

performing pull-and-release experiments, we identify time-dependent properties of junctional

mechanics consistent with a simple visco-elastic model. Integrating this constitutive law into a tissue-

scale model, we predict quantitatively how local deformations propagate throughout the tissue.

Significance statement

The shaping of tissues and organs relies on the ability of cells to adhere together and to deform in a

coordinated manner. It is therefore key to understand how cell-generated forces produce cell shape

changes, and how such forces transmit through a group of adhesive cells in vivo. In this context, we

have developed an approach using laser manipulation to impose local forces on cell contacts in the

early Drosophila embryo. Quantification of local and global shape changes using our approach can

provide both direct measurements of the forces acting at cell contacts, and delineate the time-

dependent visco-elastic properties of the tissue. The latter provides an explicit relationship, the so-

called constitutive law, between forces and deformations.

Page 100: applying optical tweezers in vivo: a biophysical study of mechanical

2

\body

Introduction

During the development of an organism, cells change their shape and remodel their contacts to give

rise to a variety of tissue shapes. Analysis of tissue kinematics has revealed that epithelial tissue

morphogenesis is partly controlled by actomyosin contractility. The spatio-temporal deployment and

coordination of actomyosin contractility produces shrinkage and extension of cell surfaces and

interfaces, which can drive tissue invagination, tissue folding, or tissue extension (1). Understanding

the mechanical nature of these processes requires force measurements in vivo; however, measurements

in developing epithelia are limited and most methods have been indirect. They rely either on force

inference from image analysis (2-4) or on laser dissection experiments at cell (5, 6) or tissue scales

(7, 8), which provide the relative magnitude and direction of stresses from cell or tissue shape

changes. In contrast, mechanical approaches have been developed in recent years to impose or

measure stresses of cells in contact, including cell monoloayer micromanipulation (9), pipette

microaspiration on cell doublets (10), and traction force microscopy on migrating epithelia (11) and

single cell doublets (12). Recently, an elegant method using deformable cell-sized oil microdroplets

has provided absolute values of stresses at the cell level in cell cultures and embryonic mesenchymes

(13), but not yet in live epithelia. In this context, we sought for a direct, in vivo method for tension

measurements and mechanical characterization at cell contacts, and developed an experimental

approach combining optical tweezers with light sheet microscopy.

To probe epithelial mechanics in a live organism, we chose the early epithelium of the Drosophila

embryo as a model system. It consists of a simple sheet of cells that spread over the yolk and are in

contact with each other through E-cadherin-based adhesion. During early embryogenesis at the

blastula stage just after the end of cellularization, epithelial cells have very similar hexagonal shapes,

suggesting that cell junctions have similar mechanical properties and the internal pressure of these

cells is homogeneous. At the later gastrula stage, cells undergo shape changes at distinct regions in the

embryo. On the ventral side of the embryo, apical cell constriction of a few rows of cells drives tissue

invagination (14), while on the ventrolateral side of the embryo, cell intercalation, a process whereby

cells exchange neighbors by polarized remodeling of their junctions, drives tissue extension. The latter

morphogenetic movement is driven by an anisotropic distribution of Myosin-II (Myo-II), which is

more concentrated along junctions aligned with the dorsal-ventral (D/V) axis (15). Laser dissection of

cortical actomyosin networks at cell junctions in the ventrolateral tissue has shown that such an

anisotropic distribution of Myo-II causes an anisotropic cortical tension (6). However, the absolute

values of tensile forces have not yet been measured, and more generally, the mechanics of cell-cell

interfaces in vivo is largely unknown. Here we addressed this issue by analyzing local mechanical

measurements at cell junctions during tissue morphogenesis and determining the contribution of Myo-

II to tension in this context. We determined the time-dependent response of cell-cell interfaces to

Page 101: applying optical tweezers in vivo: a biophysical study of mechanical

3

forced deflection and delineated a visco-elastic model of junctions. Finally, this led us to explore the

propagation of local forces within the epithelial tissue.

Results and Discussion

To probe the mechanics of cell-cell interfaces, we devised a setup combining optical tweezers and

light sheet microscopy (Fig. 1A and Fig. S1). This combination allows imaging of a whole tissue at a

high acquisition rate, while manipulating objects in vivo. Light sheet microscopy is also advantageous

to confocal microscopy because it reduces photobleaching (15). Our light sheet setup was designed

from an upright microscope: a light sheet sections the sample horizontally and the fluorescence light is

collected by a high numerical aperture objective lens pointing downwards (16) (Fig. 1A and Fig. S1).

The laser trap is produced by a near-infrared laser light focused by the collection objective lens into

the sample and is moved by galvanometric mirrors in the plane of the epithelium (Fig. 1A and Fig.

S1). Optical tweezers experiments usually require the use of a glass or a polystyrene bead to apply a

force onto an attached molecule or a cellular structure. We found that the cell-cell interfaces can be

manipulated directly, without the need of an external probe (Fig. 1B showing 3 snapshots of deflected

membrane interfaces and trap position; Fig. 1C; Movie 1). This is likely due to a positive refraction

index difference between the interface and the interior of the cells, as revealed by quantitative phase

imaging of the epithelial cells (Fig. S2). Yet, the value of this mismatch is difficult to determine

because of the geometry of the tissue.

Using the direct application of the laser, we imposed a sinusoidal movement to the trap perpendicular

to a cell interface and centered on it, and imaged the resulting deflection in the adherens junction plane

(Fig. 1C). The interface deflection followed the trap movement but with lower amplitude, suggesting

that the interface resists the mechanical load imposed by the laser trap (Fig. 1C).

To explore the regime of deformation that the laser trap imposes to the interfaces, we varied the

amplitude of the trap sinusoidal movement while keeping the period of oscillation constant. The

amplitude of the interface deflection increases with the trap amplitude, yet it deviates from a linear

relationship for trap amplitude larger than 1 µm (Fig. 1D). Then, we varied the laser power while

keeping the trap amplitude and the period of oscillation constant. We found that the interface

deflection amplitude also increases linearly with the laser power up to 300 mW (Fig. 1D, inset).

Together, these results confirm that in the case of small deformations (for comparison the average

length of an interface is 4.5 µm), the trap acts as a linear spring, whose stiffness �� is linearly

proportional to the laser power. Therefore, all the experiments are carried out within this range of

deformation (<1 µm trap amplitude) and with power 200 mW, unless otherwise stated.

To estimate the trap stiffness on the cell interfaces, and thus the forces directly applied by the

laser, we implemented a two-step procedure using beads. First, we determined the trap stiffness on

beads: single 0.46 µm diameter beads injected in the cytosol were trapped and moved in a stepwise

fashion between two trap positions separated by 0.5 µm (Fig. S3A). The resulting relaxation of the

Page 102: applying optical tweezers in vivo: a biophysical study of mechanical

4

bead towards the new trap position was exponential. The characteristic time is set by the ratio of the

drag coefficient, � �, over the trap stiffness on the bead, with the viscosity and R the bead radius.

An effective value of the viscosity was estimated independently by analyzing bead motion in the

cytosol in the absence of trap to measure its mean square displacement (Fig. S4) (17). Relating the

diffusion constant to viscosity by the Stokes-Einstein equation (� Θ = � �), where � is the

Boltzmann constant and Θ the temperature, we found that the effective viscosity of the cytosol is

3.6 ± 0.1 Pa.s (mean ± se, 1350 beads). Using this value, we could thus estimate the trap stiffness on

beads to be 120 ± 50 pN.µm-1

at 200 mW laser excitation (mean ± sd, 20 measurements). Second, we

compared the deformation produced by direct application of the focused laser on the interface with

that induced by beads pushed against the cell-cell interfaces (Fig. S3B-D and Movie 2). The former

was only 2- to 3-fold larger than the latter (2.5 ± 0.4, mean ± sd, 5 measurements), indicating that the

trap stiffness on the interfaces was 2- to 3-fold smaller than that on beads. Thus, the trap stiffness on

interfaces was estimated to be �� = 50 ± 30 pN.µm-1 at 200 mW laser excitation and in the regime of

small deformations.

The resistance to deformation can arise not only from the mechanical properties of the

interface and its apposed cortical elements, including the actomyosin cytoskeleton, but also from the

viscous cytosol. To determine whether the resistance to deformation is time-dependent, we varied the

period of oscillation while keeping the trap amplitude constant (amplitude At = 0.5 µm, Fig. 1E). For

periods larger than or equal to 1 s, which correspond to mean speeds smaller than 2 µm.s-1

, the

amplitude was constant. The constancy of the deflection amplitude at speeds below 2 µm.s-1

indicates

that the force applied by the trap and that produced by the interface are in quasi-static equilibrium. At

low speeds of deformation (speed < 2 µm.s-1

), we can thus assume that the shape of the interface

mainly results from the balance of forces between the trapping force �� and the tension of the interface�: �� = � cos , where is the angle that the interface makes with respect to the trapping force (Fig.

1B). As the vertices of the cell-cell contact did not move significantly during the deformation (Fig.

S5), we could neglect the contribution of other cells and use this simple local equilibrium formula.

For small deformations (that is for maximal deflections much smaller than the initial junction length

l0), cos ≈ �0 and �� ≈ �� �� − � , where � is the position of the interface and �� the position of the

trap. The tension of the interface thus approximates as: � ≈ � 04 ��� − . We found that within

experimental error the ratio ��� remains constant during periodic oscillations (Fig. 1F), indicating that

the pre-existing tension � of the interface is not significantly modified during small deformations.

Importantly, this also implies that tension measurements, while relying on geometrical and physical

approximations, do not require a mechanical model of cell contacts. Thus, tension values can be

obtained as a simple linear function of the ratio between the interface deflection and the trap position.

Using our estimate of �� = 50 pN.µm-1

, and that the ratio ��� = 1.88 ± 0.4 (mean ± sd, 16

Page 103: applying optical tweezers in vivo: a biophysical study of mechanical

5

measurements), we found that tension � at cell-cell interfaces is on the order of 44 ± 22 pN at the end

of cellularization. The tensions reported here are in the same range as cortical tensions measured on

single cells (18), but 2-3 orders of magnitude below cell-cell forces in cell aggregates on adhesive

substrates in vitro (12).

We observed that for periods smaller than 1 s (speed > 2 µm.s-1

), the deflection was reduced

(Fig. 1E). This is the characteristic signature of viscous damping, presumably related to the viscous

drag in the cytosol. To test whether damping was indeed due to the viscosity of the cytosol, we

performed relaxation experiments after instantaneous release of the trap (Fig. S6A). At the onset of

relaxation, tension is balanced only by viscous damping: � cos = �� , where � is the initial

relaxation velocity and �is the damping coefficient. Measuring � (Fig. S6A) and using the mean

tension value of 44 pN, this provided an indirect measurement of �, which is in the order of 2 ±

1×10-5

m.Pa.s. To determine viscosity, � should be rescaled by both a typical length scale L of

deformation and a geometric coefficient g: = ���. In the plane of junctions, the deformation extends

to the whole contact line (4-5 µm). We found a similar value for the deformation along the apico-basal

direction (Fig. S6B). The deformation is thus likely to be akin to a two-dimensional Gaussian, with a

typical width of 4-5 µm. Therefore we took � = . µm and = , which corresponds to disk

approximation. The viscosity associated with the observed damping is thus on the order of 1 Pa.s,

which is consistent with our previous measurements of cytosol viscosity using beads. This is also

consistent with microrheological measurements made in the cytosol of C. elegans embryo (19).

During tissue morphogenesis, cells undergo cell shape changes driven by Myo-II contractile events,

which induce interface deformation at various speeds up to about 0.1 µm.s (20, 21). The condition of

low speeds (speed < 2 µm.s-1

) is thus always fulfilled during cell shape changes driven by Myo-II

contractility during tissue morphogenesis of Drosophila, indicating that short-time viscous damping

should play no role in the process.

Using direct optical manipulation, we then probed how cell-cell tensions change during tissue

morphogenesis of the early germband of the Drosophila. Before gastrulation movements, at the end of

stage 5, cells form a regular lattice with isotropic shapes (Fig. 2A, top panels) (6, 22, 23). Later, at the

end of stage 6 and the onset of stage 7, the total concentration of Myo-II increases at adherens

junctions, and its distribution becomes anisotropic, with higher levels along interfaces parallel to the

D/V axis (Fig. 2A, middle and bottom panels) (6, 22, 23). This anisotropic distribution of Myo-II has

been shown to drive polarized junction shrinkage and cell intercalation (6, 22, 23). We found that the

deflection amplitude of cell-cell interfaces caused by the optical trap is reduced by two-fold from the

end of stage 5 to the onset of stage 7, indicating a tension increase (Fig. 2B). Moreover, at stage 7,

cell-cell interfaces with a direction close to the D/V axis are about 2.5 times more tense than those

along the horizontal axis (Fig. 2C and Fig. S7 showing tension normalized to junction size). The

Page 104: applying optical tweezers in vivo: a biophysical study of mechanical

6

anisotropy of tension is consistent with a previous report, which estimated the relative values of

tension from laser nanodissection (6).

Inhibition of Myo-II activity by injection of ROCK inhibitor resulted in a significant reduction of the

tension at cell junctions at stage 7 (Fig. 2D) and a loss of tension anisotropy (Fig. 2C). This confirms

that the significant increase in tension measured between stage 5 and stage 7 can be attributed to Myo-

II activity. Given that a single molecular motor of Myo-II produces a few pN of force (24), the range

of forces measured here suggests that the increase in tensions from stage 5 to stage 7 could be powered

by only a few dozens of motors.

The E-cadherin junctions, where we performed the experiments presented above, are restricted to a

thin 1 µm section and are localized about 1-2 µm below the apical surface (25, 26) (Fig. 2D, left

panel). As Myo-II accumulates at the adherens junction plane at the end of stage 6 and onset of stage

7, we wondered if this might translate into different mechanical properties at the adherens junction

plane as compared to more basal positions. Measurements of interface deflection by laser trap at

different positions along the apicobasal axis showed that during tissue morphogenesis, there is a

gradual polarization of the tension along this axis (Fig 2E, lower panel). While at stage 5, tension at

adherens junctions is the same as in a more basal plane, it becomes larger at stage 7. The fact that we

measure the same deformation at the adherens junction plane and 3 µm more basally, at stage 5 end,

when Myosin-II is very apical and not junctional, also suggests that the apical cortex has not a

significant contribution to the restoring force. Altogether these results show that optical tweezers can

be robustly used to measure tension at cell contacts in vivo, and that this method can reveal planar-

polarized as well as apicobasal anisotropies of tension in a developing organism.

This led us to explore in more detail the mechanical response of cell-cell contacts to forced deflection,

at different temporal and spatial scales. First we performed pull-release experiments, which have been

used in vitro on single cells with optical and magnetic tweezers (27), but have never been applied in

vivo. Pull-release experiments consist of switching the laser trap on/off at a few hundred nanometers

distance from the junction and then monitoring the deflection of the cell-cell interface, both towards

(trap on) and away from the trap (trap off) (Fig. 3A). At these short time scales, the mechanical

response of an epithelial tissue should result both from the constitutive mechanics of its actomyosin

cortex and from viscous friction exerted by the cytosol. In other words, modeling deflection dynamics

requires both a cortical constitutive equation, and a force balance equation between the cortical

restoring force, the trapping force, and the viscous friction. The pull-release curves obtained show that

the dynamics is not purely exponential, and exhibits at least two characteristic times, in the range of

1 s and 10 s (Fig. 3A and Fig. S8A). We considered different types of visco-elastic constitutive models

coupled to the force balance equation to fit the experimental data, and found that a so-called standard

linear solid (SLS) model, composed of a Maxwell arm (a spring and a dashpot) in parallel with a

spring, is the best and simplest model to correctly account for the observed behavior (Fig. 3A; fit:

Page 105: applying optical tweezers in vivo: a biophysical study of mechanical

7

solid blue curve, SLS model: inset). Indeed, we can rule out Kelvin-Voigt and Maxwell models, which

predict simple exponential relaxation (Fig. S8B). Notably, the SLS model also fits well the periodic

experiments described above (Fig. 1E, solid blue curve). In particular, the 1 s time scale is consistent

with the drop of deflection amplitude observed for periods of oscillation below 1 s, which can

therefore be attributed to the viscous drag in the cytosol. The existence of another time scale (10 s)

denotes the fact that the cortex itself is not purely elastic but visco-elastic (hence the SLS model).

These two time scales can be derived analytically, and indeed, one is determined by the friction in the

cytosol, while the other is given by the viscous component of the cortex constitutive equation.

Importantly, both these visco-elastic time scales are under a minute; neither the experiments nor the

model deal with the long-term dynamics (minutes to hours), which presumably implies creep and

therefore fluid-like behavior (9, 28).

During tissue morphogenesis, the integration of local forces shapes the tissue (1). A

challenging question is whether local forces produce long-range deformation, and at what speed the

mechanical information propagates. Thus, having established a model for the mechanics of single cell

interfaces, we then asked how single cell deformation propagates throughout the tissue. We imposed

the local deflection of a cell interface using sinusoidal oscillations, and we tracked the deflection of

neighboring interfaces away from this point. For that purpose, we plotted kymographs along lines

perpendicular to cell interfaces (Fig. 3B). The target interface was oscillated using a deflection

amplitude of 1 ± 0.1 m (Fig. 3C, left panel), which is larger than in the experiments presented so far

in order to facilitate the detection of propagation. We observed that the neighboring interfaces within a

distance of 1 to 2 cells also deflected periodically, but with much lower amplitudes and a small phase

shift (Fig. 3C, left panel and 3D, left panel and Fig. S9A-D). This indicates that the deformation

typically decays over a distance of 1-2 cells.

To evaluate the ability of our simple mechanical model to reproduce these data, we first

transposed the single junction model to a network of contacts, as observed in vivo. Each contact is

considered as an SLS element (again with viscoelastic parameters � , � and ) pulling on its vertices

(Fig. 3A). The displacement of each vertex then results from the force balance between tensile forces

exerted by adjacent contacts, and external damping, ��̇, due to cytosol viscosity (Fig. 3E). We were

then able to simulate the mechanical response of the tissue to periodic manipulations in the extracted

experimental geometry, imposing only the sinusoidal displacement at the midpoint of the target

interface (Fig. 3F).

To quantitatively assess the accuracy of the model, we plotted kymographs in the simulated tissues

along the same lines as in the experiments (Fig. 3C, right panel, and Fig. S9). The parameter values

that we estimated from the single interface experiments faithfully reproduce the tissue-scale

observation: one or two neighboring cells deform away from the source point of deflection (Fig. 3C, D

and G, Movie 3). The propagation is a bit more efficient transverse (Fig. S9C) than perpendicular to

deformation (Fig. 3G). Taking higher or lower viscosity for the cytosol into account leads to shorter

Page 106: applying optical tweezers in vivo: a biophysical study of mechanical

8

and longer propagation distance, respectively, which does not correctly reproduce the observed

behavior (Fig. S9E). Note that our model underestimates the speed of propagation (Fig. 3D), which

suggests that constant volume constrains and/or transmission through the apical cortex may contribute

to propagation. The speed of propagation can be experimentally estimated by the phase delay between

the trapped interface’s deflection and that of its neighbors (Fig. 3D, time shift between black and

magenta curves). At a 1-cell distance (~7 m), we measured a time delay of 150 ± 85 ms (95%

confidence interval), which corresponds to a propagation speed with a typical phase velocity of 45

m.s-1

. This speed is much larger than the speed of actomyosin flows, which are observed in different

systems, including Drosophila and C. elegans (0.1 µm.s-1

) (21, 29).

Conclusion

Here we studied the mechanical properties of cell-cell contacts during tissue morphogenesis and how

local deformation propagates within a tissue. We provide absolute values of tensions at cell interfaces,

while previous work estimated relative values based on assumptions of the viscous properties at cell

junctions. Surprisingly, the small forces produced by optical tweezers are sufficient to produce

significant deflection of cell interfaces, and we could estimate that tension at cell interfaces is in the

100 pN range during early stages of tissue morphogenesis. This suggests that the forces that remodel

cell-cell contacts during tissue morphogenesis and drive the shrinkage or extension of cell contacts can

be powered by a small number of molecular motors. Fluctuations in cell shape, which are observed

during these events, might thus result from stochastic fluctuations in motor numbers.

The possibility of absolute tension measurements at cell contacts might be beneficial to force

inference methods (2-4), which provide relative tensions based on the geometry of the contact

network. It might indeed allow experimental validation of the inference, and can also be used to

calibrate the inferred tensions.

Our study has provided a predictive mechanical model of cell contacts. Modeling the

constitutive mechanics of epithelia by quantifying how forces dynamically cause deformations is

crucial for understanding epithelial morphogenesis events, which was suggested as early as 1981 by

Odell’s pioneering work (30). Since then, a variety of mechanical descriptions have been proposed;

however, testing of the underlying hypotheses has been limited due to the lack of in vivo experimental

tools. Notably, so-called vertex models (31), usually based on energy minimization, do not incorporate

energy dissipation and thus cannot predict the tissue dynamics. Here we propose a vertex-based

model, which bridges usual vertex models and continuum mechanics with finite elements approaches

that integrate visco-elastic constitutive behavior (32, 33). In addition, it captures the non-trivial two-

timescale relaxation dynamics evidenced by pull-release experiments. Finally, here we intentionally

restricted our analysis to time scales and speeds faster than the changes in contractility in order to deal

with steady shape patterns. We believe that the approach we have established here is now ready to

explore additional time scales and to probe long-term, irreversible deformation of cell contacts.

Page 107: applying optical tweezers in vivo: a biophysical study of mechanical

9

Materials and methods

Experiments and data analysis

Optical manipulation was done using a custom-built two-colors (488 and 516 nm) light sheet

microscope (16), coupled with a single beam gradient trap (1070 nm wavelength, ytterbium fiber laser,

IPG). A 100x water-immersion lens (1.1 NA, Nikon) was used for imaging as well as for introducing

the optical trap in the imaging plane. Galvanometric mirrors controlled laser trap position deflection to

produce sinusoidal oscillations or step movements. Prior to every in vivo experiment, the relationship

between galvanometer voltages and laser trap position was calibrated using fluorescent beads

(localization precision of 25 nm). During experiments both images and position of the galvanometers

were simultaneously recorded. Kymographs of interface deflection were extracted from movies along

lines perpendicular to the interfaces and were fitted at each time step by a gaussian to determine the

interface position with subpixel resolution (localization precision of 35 nm).

Quantification of E-cadherin::GFP and Squash::GFP was done in a spinning disk microscope (Perkin-

Elmer) using a 100x oil immersion lens (Nikon).

For details on sample preparation, see SI Materials and Methods.

Model

- In the single junction model (Fig. 3A and fit Fig. 1E), the horizontal restoring force is related to the

deflection � of the interface through the SLS constitutive mechanics of the cortex: ̇ + � = � + � �̇ + � � �.k1 and k2 are elastic parameters (N.m

-1), a viscous parameter (m.Pa.s), and the dot denotes a temporal

derivative. The force balance at the interface simply reads: = �� �� − � − ��̇,

where �� and �� are the stiffness and position of the optical trap, and � is the damping coefficient in

the cytosol. Combining these two equations yields two characteristic time scales, one related to

(viscous component of the cortex) and the other to � (damping coefficient of the cytosol).

- In the tissue scale model, each bond has a visco-elastic dynamics, given by the same model: �̇ + � � = � + � �̇ + � + � �,where � is the tension and � is the elongation. At vertex i, tensions of adjacent interfaces (j=adj(i)) are

balanced only by viscous damping. The force balance thus writes:

��i⃗⃗⃗ ̇ = ∑ �ij ⃗⃗ ⃗⃗ =�� , which provides direct access to vertices displacements. The midpoint of the target interface is treated

as a 2-way vertex in the simulations. Its movement is imposed as to mimic the considered experiment.

Page 108: applying optical tweezers in vivo: a biophysical study of mechanical

10

Acknowledgements

We thank Serge Monneret (Institut Fresnel Marseille) for quantitative phase imaging experiments. We

thank Sébastien Sénatore and Edith Laugier for help with bead injection. We thank all members of the

Lenne and Lecuit laboratories for discussions and comments on the manuscript. We also acknowledge

members of the Labex INFORM for discussions (ANR-11-LABX-0054) and the France-BioImaging

infrastructure (ANR-10-INSB-04-01, call "Investissements d'Avenir"). This work was supported by

the Fondation pour la Recherche Médicale (Equipe FRM DEQ20130326509) and an ANR-Blanc grant

(Morfor, ANR-11-BSV5-0008) to P-F.L. K.B. was supported by a PhD fellowship from the Région

PACA & NIKON, O.B. by a PhD MESR fellowship and a FRM individual grant.

Page 109: applying optical tweezers in vivo: a biophysical study of mechanical

11

References

1. Lecuit T, Lenne P-F, Munro E (2011) Force generation, transmission, and integration during

cell and tissue morphogenesis. Annu Rev Cell Dev Biol 27:157–184.

2. Cranston PG, Veldhuis JH, Narasimhan S, Brodland GW (2010) Cinemechanometry (CMM):

A method to determine the forces that drive morphogenetic movements from time-lapse

images. Ann Biomed Eng 38:2937–2947.

3. Chiou KK, Hufnagel L, Shraiman BI (2012) Mechanical stress inference for two dimensional

cell arrays. PLoS computational biology 8:e1002512.

4. Ishihara S et al. (2013) Comparative study of non-invasive force and stress inference methods

in tissue. Eur Phys J E Soft Matter 36:9859.

5. Farhadifar R, Roper JC, Aigouy B, Eaton S, Julicher F (2007) The influence of cell mechanics,

cell-cell interactions, and proliferation on epithelial packing. Curr Biol 17:2095–2104.

6. Rauzi M, Verant P, Lecuit T, Lenne P-F (2008) Nature and anisotropy of cortical forces

orienting Drosophila tissue morphogenesis. Nat Cell Biol 10:1401–1410.

7. Hutson MS et al. (2003) Forces for morphogenesis investigated with laser microsurgery and

quantitative modeling. Science 300:145–149.

8. Bosveld F et al. (2012) Mechanical control of morphogenesis by Fat/Dachsous/Four-jointed

planar cell polarity pathway. Science 336:724–727.

9. Harris AR et al. (2012) Characterizing the mechanics of cultured cell monolayers. Proc Natl

Acad Sci USA 109:16449–16454.

10. Maître J-L et al. (2012) Adhesion Functions in Cell Sorting by Mechanically Coupling the

Cortices of Adhering Cells. Science 338:253–256.

11. Trepat X et al. (2009) Physical forces during collective cell migration. Nature Physics 5:426–430.

12. Maruthamuthu V, Sabass B, Schwarz US, Gardel ML (2011) Cell-ECM traction force

modulates endogenous tension at cell-cell contacts. Proc Natl Acad Sci USA 108:4708–4713.

13. Campàs O et al. (2013) Quantifying cell-generated mechanical forces within living embryonic

tissues. Nature methods 11:183–189.

14. Kölsch V, Seher T, Fernandez-Ballester GJ, Serrano L, Leptin M (2007) Control of Drosophila

gastrulation by apical localization of adherens junctions and RhoGEF2. Science (New York,

NY) 315:384–386.

15. Huisken J, Swoger J, Del Bene F, Wittbrodt J, Stelzer EH (2004) Optical sectioning deep

inside live embryos by selective plane illumination microscopy. Science 305:1007–1009.

16. Chardès C, Ménélec P, Bertrand V, Lenne P-F (2014) Setting-up a simple light sheet microscope for in toto imaging of C. elegans development. Journal of visualized experiments: JoVE 87:e51342.

17. Kusumi A, Sako Y, Yamamoto M (1993) Confined lateral diffusion of membrane receptors as

studied by single particle tracking (nanovid microscopy). Effects of calcium-induced

Page 110: applying optical tweezers in vivo: a biophysical study of mechanical

12

differentiation in cultured epithelial cells. Biophys J 65:2021–2040.

18. Salbreux G, Charras G, Paluch E (2012) Actin cortex mechanics and cellular morphogenesis.

Trends in cell biology 22:536–545.

19. Daniels BR, Masi BC, Wirtz D (2006) Probing single-cell micromechanics in vivo: the

microrheology of C. elegans developing embryos. Bioph J 90:4712–4719.

20. Martin AC, Kaschube M, Wieschaus EF (2009) Pulsed contractions of an actin-myosin

network drive apical constriction. Nature 457:495–499.

21. Rauzi M, Lenne P-F, Lecuit T (2010) Planar polarized actomyosin contractile flows control

epithelial junction remodelling. Nature 468:1110–1114.

22. Bertet C, Sulak L, Lecuit T (2004) Myosin-dependent junction remodelling controls planar cell

intercalation and axis elongation. Nature 429:667–671.

23. Blankenship JT, Backovic ST, Sanny JSP, Weitz O, Zallen JA (2006) Multicellular rosette

formation links planar cell polarity to tissue morphogenesis. Dev Cell 11:459–470.

24. Howard J (2001) Mechanics of motor proteins and the cytoskeleton (Sinauer Associates).

25. Harris TJC, Peifer M (2004) Adherens junction-dependent and -independent steps in the

establishment of epithelial cell polarity in Drosophila. The Journal of Cell Biology 167:135–147.

26. Truong Quang B-A, Mani M, Markova O, Lecuit T, Lenne P-F (2013) Principles of e-cadherin

supramolecular organization in vivo. Current biology: CB 23:2197–2207.

27. Alenghat FJ, Fabry B, Tsai KY, Goldmann WH, Ingber DE (2000) Analysis of cell mechanics

in single vinculin-deficient cells using a magnetic tweezer. Biochemical and biophysical

research communications 277:93–99.

28. He B, Doubrovinski K, Polyakov O, Wieschaus E (2014) Apical constriction drives tissue-

scale hydrodynamic flow to mediate cell elongation. Nature 508:392–396.

29. Mayer M, Depken M, Bois JS, Jülicher F, Grill SW (2010) Anisotropies in cortical tension

reveal the physical basis of polarizing cortical flows. Nature 467:617–621.

30. Odell GM, Oster G, Alberch P, Burnside B (1981) The mechanical basis of morphogenesis. I.

Epithelial folding and invagination. Developmental biology 85:446–462.

31. Fletcher AG, Osterfield M, Baker RE, Shvartsman SY (2014) Vertex models of epithelial

morphogenesis. Biophys J 106:2291–2304.

32. Hutson MS et al. (2009) Combining laser microsurgery and finite element modeling to assess

cell-level epithelial mechanics. Biophys J 97:3075–3085.

33. Bonnet I et al. (2012) Mechanical state, material properties and continuous description of an

epithelial tissue. J R Soc Interface 9:2614–2623.

Page 111: applying optical tweezers in vivo: a biophysical study of mechanical

13

Figure Legends

Figure 1: Characterizing the deflection of cell-cell interfaces imposed by optical tweezers.

(A) Schematic of the setup: the embryo is optically sectioned by a light sheet and imaged while a laser

trap (red) allows manipulation. Top image shows the epithelium labeled by a membrane marker

(GAP43::mcherry) and the laser trap position (yellow arrowhead). (B) Separate images of the interface

in three different positions of deflection (yellow arrowheads, 3 top images). The bottom image is a

merge of the three upper images (position 1 in green, position 2 in red, position 3 in blue). Also a

schematic of deflection with distribution of forces labeled. Scale bar, 5 µm. (C) Representative plot of

deflection versus time showing both trap (red solid line) and interface positions (black solid line). (D)

Amplitude of interface deflection as a function of trap movement amplitude and laser trap power

(inset). 7-13 independent measurements per data point. Error bars represent one standard deviation.

(E) Interface deflection amplitude over the trap oscillation period. (F) Interface position as a function

of trap position during few cycles of laser oscillation (amplitude: 0.5 µm, period: 2 s). Successive

cycles are in different colours (black: first, red: second, green: third). The blue line represents a linear

fit.

Figure 2: Tension at cell contacts before and during germband elongation in wild type and

perturbed embryos.

(A) Left panels: Schematic of tissue elongation and images at three different stages: before elongation

(stage 5 end), at the onset of Myosin-II accumulation at cell junctions (stage 6 end) and during tissue

extension (stage 7). Right panels: images showing the cell interface (purple, GAP43::mcherry), E-

cadherin (green) and Myosin-II (red, MRLC: regulatory light chain of Myosin-II) at the different

stages. (B) Interface deflection amplitude (gray bars) and tension (red bars) at the adherens junction

plane, for different stages (16, 24 and 14 different interfaces measured, respectively). (C) Interface

deflection amplitude at different stages for different junction orientation, along and perpendicular to

the D/V axis. (D) Interface deflection amplitude (gray bars) and tension (red bars) in the adherens

plane in ROCK inhibitor-injected embryos and control embryos (WT: 14 interfaces, water injected; 11

interfaces, ROCK inhibitor: 15 interfaces. Error bars represent one standard deviation. (E) Ratio

between the interface deflection at the adherens junction plane and in a more basal plane (3 µm below

adherens junctions) at different stages. The red line is the median, the box edges are the lower and

upper quartiles, and the whiskers display the total range of measurements. T-test (* not significant,

**p-value <0.01, *** p-value <0.001). Scale bar, 10 µm.

Figure 3: Mechanical model of the interface and tissue response.

(A) Deflection of the interface in a pull-release (trap on - trap off) experiment. The model (blue line)

accurately fits the experimental data (black). The simplest analogous visco-elastic model is a maxwell

Page 112: applying optical tweezers in vivo: a biophysical study of mechanical

14

arm in parallel with a spring (inset). For a trap stiffness kt = 50 pN.m-1

, fit parameters values are:k = 15 pN.m-1

, k = 55 pN.m-1

, ζ = 1.5 × 10-4

m.Pa.s, and Cη = 1.5 × 10-5

m.Pa.s. Note that this

value of Cη is consistent with our previous estimate (2 ± 1 × 10-5

m.Pa.s). The same parameter values

are used in the simulations. (B) Deflection perpendicular to the interfaces is tracked over time along

lines perpendicular to cell interfaces (red), which allows measuring deformation away from the

targeted interface (yellow arrowhead). Scale bar, 10 µm. (C) Kymograph of interface deflections in the

experimental (left) and simulated (right) tissues. Only the interfaces adjacent to the target interface

display significant deflection. (D) Deflection of the target (black) and neighbor interfaces (at 1-cell

and 2-cell distance, magenta and green, respectively) in the experimental (left) and simulated (right)

tissues. (E) In the model, the movement of a vertex results from a balance between tension from

adjacent bonds and viscous friction. (F) Overlay of the undeformed (purple) and deformed (green)

tissue in the experimental (left) and simulated (right) tissue. (G) Spatial decay of interface deflections

over the neighboring cells. Comparison between experiments and simulations.

Page 113: applying optical tweezers in vivo: a biophysical study of mechanical

Position 1 Position 2 Position 3

0 0.2 0.4 0.6 0.8 1 1.20

0.1

0.2

0.3

0.4

0.5

0.6

A C

Time (s)

Inte

rface d

eflection (μm

) Trap

Interface

0 2 4 6 8 10 12

0

1

E

Inte

rface d

eflection a

mplit

ude (μm

)

Period of oscillation (s)

F

Trap deflection amplitude (μm)

Inte

rface

deflection a

mplit

ude (μm

)

00.10.20.30.40.5

0 100 200 300 400

Deflection

(μm

)

Laser Power (mW)

B

D

Trap position (μm)

Inte

rface p

ositio

n (μm

)

Embryo

0 1 2 3 4 50

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0 0.5 1

0

0.5

Page 114: applying optical tweezers in vivo: a biophysical study of mechanical

A

EC D apical

basal

AJ

Stage 7

water inj. ROCK inh.

St. 5 end St. 7

E-cad MRLC

MR

LC

E-cadGAP43 MRLC

Stage 6 end

Stage 7

0

0.05

0.1

0.15

0.2

0.25

0.3

0

50

100

150

200

250

300

Tensio

n(p

N)

Inte

rface D

eflection a

mplit

ude (μm

) ******B

0

50

100

150

200

250

300

350

400

Tensio

n (

pN

)

St. 6 end

St. 5 end St. 7 St. 6 end

PA

D

VP

**

St.7 St.7

water inj.

St.7 ROCK Inh.0

0.05

0.1

0.15

0.2

0.25

0.3

0

50

100

150

200

250

300

Tensio

n(p

N)

Inte

rface D

eflection a

mplit

ude (μm

) ******

St.5 end St.6 end St.7

0.5

1

1.5

AJ/B

asal

deflection r

atio

****

Stage 5 end

0-45°

45-90°

PA

D

V

St. 7

ROCK inh.

Page 115: applying optical tweezers in vivo: a biophysical study of mechanical

A

0 5 10 15 20 25 300

0.2

0.4

0.6

0.8

1

Deflection (

μm

)

Time (s)

trap on

off

cell interface

De

fle

ctio

n a

mp

litu

de

m)

C

Time

Experiment Simulation D

experimentmodel

0 5 10����

��

����

0

0.5

1

1.5

Time (s)

Inte

rface d

eflection (

μm

)Experiment

0 5 10Time (s)

Simulation

Time

G

0 1 2 3

00.20.40.60.81

1.2

E

F

B

Distance (cells)

Page 116: applying optical tweezers in vivo: a biophysical study of mechanical

1

Supporting Information

SI Materials and Methods Sample preparation. To mark the cell-cell interface in Drosophila embryos, flies double-labeled with

E-cadherin::GFP (endogenous promoter) and GAP43::mcherry (squash promotor) were selected.

Alternatively, in some cases squash::GFP GAP43::mcherry flies were used. Flies were maintained at

25°C. To obtain embryos, a fresh plate was incubated for 2-2.5 hours. After removing yeast from the

plate, the embryos were washed with 100% bleach for 50 seconds to remove the vitelline membrane.

Embryos at the end of cellularization (stage 5 end) were then selected under a dissection microscope

and aligned on the edge of the coverslip. Alignment was done with the germband visible in the

imaging plane. For experiments with beads and myosin perturbation, embryos were placed in

halocarbon oil and injected using a microinjection setup with either polystyrene beads (1:1000 stock

dilution, Molecular Probes) or with ROCK inhibitor (Y-27632, 10 mM, Invitrogen), respectively.

Embryos were immersed in water for the light sheet microscope setup and in halocarbon oil for

spinning disk imaging.

Optical manipulation and imaging. Optical manipulation of the cell-cell interfaces in individual

embryos was done using a custom-built light sheet microscope (1), coupled with a single beam

gradient trap (1070 nm wavelength, ytterbium fiber laser, IPG Photonics). A 100x water-immersion

lens (1.1 NA, x40, Nikon) was used for imaging as well as for introducing the optical trap in the

imaging plane. Imaging was done using 488 nm and 561 nm excitation lasers. Images were acquired

by an EMCCD camera using a dualview, simultaneous imaging system. Prior to every in vivo

experiment, we calibrated the relationship between galvanometer voltages and laser trap position using

the following procedure: single 500 nm diameter fluorescent polystyrene beads (fluorescence

excitation at 561 nm) were trapped in water and moved slowly by imposing galvanometer voltages

(V1, V2) of the form (V0 cos (ωt), V0 sin (ωt)) with ω < 0.3 rad/s. Images were acquired synchronously

to the voltage commands and successive (X,Y) positions of the bead were localized by a two-

dimensional Gaussian fit. The subpixel localization precision was 25 nm. The measurements were

repeated for different voltage amplitudes (corresponding to trap amplitudes in the image plane < 10

µm). This provided the relationship between (V1, V2) and (X,Y), which was subsequently extrapolated

linearly and inverted to determine (X,Y) as a function of (V1, V2) (Matlab script). This information

was used to provide laser position during every interface deflection experiment. Two kinds of deflections were given to the cell-cell interface: periodic or pull-release. Sinusoidal

oscillations of the galvos were performed to produce linear movements of the laser trap, varying time

periods from 0.3-5 s, amplitudes from 0.3-1.1 µm, and laser power from ~50-300 mW (after objective

lens). For most experiments values were kept constant with a time period 2 s, laser amplitude of

0.5 µm and laser power at 200 mW. For pull-release experiments, the stationary laser trap was

Page 117: applying optical tweezers in vivo: a biophysical study of mechanical

2

switched on at 100 nm – 2 µm from the cell-cell interface for between 10 s - 1 minute. All experiment

recordings were done for galvo-voltage as well as the camera images (either at 561 nm, or both 488

nm and 561 nm excitation).

Quantification of E-cadherin::GFP and Squash::GFP at stage 5 end and stage 7 was done in a Perkin-

Elmer spinning disk microscope using a 100x oil immersion lens.

Quantitative phase imaging uses a transmission light microscope and quadriwave lateral shearing

interferometry, as described in (2).

Data analysis. Kymographs of interface deflections were produced from the movies either in Fiji

(Multiple Kymograph plugin) or using a custom Matlab script. To extract an actual position of the

interface out of the kymograph, a gaussian fit perpendicular to the interface (along the kymograph

line) was performed. At each time step, the peak of the gaussian fit determines the interface position

with subpixel resolution. To determine the localization error, we fixed embryos expressing

Gap43::mCherry and imaged them in the same conditions as in vivo. We then localized cell interfaces

over 100 images, and found that the standard deviation of localization is 35 nm (10 cell interfaces

measured). The interface position, together with the optical trap position recorded through the laser

voltage, was then analyzed in Matlab to measure the response amplitude by maxima detection and

averaging. All fits were performed with Matlab: numerical equations were solved repeatedly,

exploring the space of parameters starting from random values and using the gradient descent method

to minimize error. Statistical analyses were done using the unpaired t-test. For the propagation

analysis, experimental tissue geometries were extracted using the Tissue Analyzer toolbox, courtesy of

B. Aigouy (3), then exported to Matlab to perform the simulations.

The delay between deformations of successive cell-cell contacts in the propagation study was

estimated using a custom Matlab script of “time-sliding fit”. We shifted one signal in time, i.e. we

plotted x(t+∆T) as a function of the trap position xt(t). The time shift ∆T which provides the best linear

fit between x(t+∆T) and xt(t) provides an estimate of the time delay between the two signals. The

confidence intervals were obtained using the nlparci function of the Matlab statistics toolbox.

Model Single junction

The mechanical model for a single junction (Fits in Fig. 1E and 3A) is derived from the constitutive

mechanics of the cortex and from a force balance equation at the interface. The visco-elastic

constitutive equation is given by the so-called standard linear solid (SLS), and relates the horizontal

restoring force to the deflection of the interface:

Page 118: applying optical tweezers in vivo: a biophysical study of mechanical

3

and are elastic parameters (N.m-1), a viscous parameter (m.Pa.s), and the dot denotes a

temporal derivative. At these very low Reynolds number, inertia can be neglected, and the balance of

forces at the interface then simply reads:

where and are the stiffness and position of the optical trap, respectively, and is the damping

coefficient of the interface in the cytosol. The first term on the right-hand side thus corresponds to the

force exerted by the optical trap, while the second corresponds to the viscous drag in the cytosol – and

is therefore proportional to the velocity . This linear system can then be solved for any trap

trajectory ; in particular for our experimental conditions, a sinusoidal oscillation or a pull-release

experiment. Moreover, the relaxation time scales associated to this system can be derived analytically.

Combining the constitutive equation and the force balance in the absence of trap yields:

The solution is in the form / / . In the limit , which we find

is verified from fit values, the two time scales and simplify into:

In that limit, one time scale is related to the viscous component of the cortex, , while the other is

related to the damping coefficient in the cytosol, .

Tissue scale

In the tissue scale simulations, the epithelium is considered as a network of bonds – the cell contacts –

between vertices. Each bond is considered as a visco-elastic segment. The constitutive equation of

each segment, similar to the first equation, is:

where is the tension and is the elongation, . The displacement of each vertex is then

Page 119: applying optical tweezers in vivo: a biophysical study of mechanical

4

computed using the force balance equation between the tension at adjacent contact lines ( ) and damping in the cytosol ( ) (Fig. 3D). The force balance at vertex i thus reads:

i ij This provides direct access to vertices displacements through velocities i. Notably, the midpoint of

the target interface is treated as a 2-way vertex in the simulations. Its movement is imposed as to

mimic the considered experiment. The rest of the vertices move according to the force balance

equation; therefore, their movement ultimately results from the deflection movement of the target

interface. We use fixed (zero displacement) boundary conditions. The areas are not constrained as we

consider small deformations only. For larger deformations, the model would almost certainly require

area or “pressure” constraints.

1. Chardès C, Ménélec P, Bertrand V, Lenne P-F (2014) Setting-up a simple light sheet microscope for intoto imaging of C. elegans development. Journal of visualized experiments: JoVE 87:e51342.

2. Bon P, Maucort G, Wattellier B, Monneret S (2009) Quadriwave lateral shearing interferometry forquantitative phase microscopy of living cells. Opt Express 17:13080–13094.

3. Aigouy B et al. (2010) Cell flow reorients the axis of planar polarity in the wing epithelium of Drosophila.Cell 142:773–786.

Page 120: applying optical tweezers in vivo: a biophysical study of mechanical

5

Supplementary Figure Legends

Figure S1: Optical setup combining light sheet microscopy and optical tweezers. In the light sheet illumination unit, the lasers are mixed by the dichroic mirrors, and enter the AOTF,

which controls the power of each laser independently. Then, the telescope increases the size of the

beam by 5-fold and the periscope brings it to the height of the microscope. The cylindrical lens forms

the light sheet, which is refocused by the illumination objective. The detection unit is integrated in the

upright microscope and is mainly composed by the detection lens, the filter, the tube lens and the

EMCCD. The sample is positioned at the intersection between the illumination and detection paths. A

piezoelectric stage allows vertical (Z) displacements of the sample for 3D acquisition. In the optical

tweezers unit, a near-infrared (NIR) laser beam (1070 nm, continuous wave) is deflected by two

galvanometric mirrors and expanded by a 5-fold telescope. The expanded laser beam is reflected by a

hot dichroic mirror and is tightly focused by the collection objective of the light sheet microscope.

Figure S2: Phase mismatch at cell interfaces. (A) Quantitative Phase Microscopy Image obtained from an early Drosophila embryo (stage 6). The

method uses a transmission light microscope and quadriwave lateral shearing interferometry, as

described in (2). The epithelial cells are observed in cross-section (cartoon, top). Scale bar, 5 µm. The

calibration bar shows the optical path difference in nm. (B) A plot profile along a line (red dotted line

in A) shows that the optical path difference is larger at cell interfaces than inside the cells. Given that

the line defines positions where the geometrical thickness of the embryo is constant, this indicates

there is a refraction index increase at cell interfaces. Arrows of different colors mark the position of 3

interfaces.

Figure S3: Deformation induced by optically-tweezed beads. (A) Position of a 0.46 µm diameter bead (blue line) in the cytosol moving between two trap positions

separated by 0.5 µm (red). From one trap position to another, the bead relaxation is exponential with a

characteristic time given by the ratio of the drag coefficient over the trap stiffness. (B) Snapshots of an

interface deformation induced by a 0.46 µm diameter bead moved by the laser trap against the

interface. The red and green channels correspond to two different positions of the trap separated by

approximately 0.5 µm. (C) Positions of laser trap, bead and interface in an oscillatory experiment with

bead at 100 mW laser power. (D) Positions of laser trap and interface in the same conditions as in (C)

in the absence of bead.

Figure S4: Viscosity measurements obtained from trajectories of individual beads. (A) Image showing 100 nm diameter beads (red) injected in the embryo. Cell contours are labeled by

E-cadherin::GFP. (B) Single particle trajectories superimposed to an image of the cells. (C) Fraction

Page 121: applying optical tweezers in vivo: a biophysical study of mechanical

6

of beads exhibiting diffusive, subdiffusive and superdiffusive behaviors from analysis of the mean-

square displacement (using criteria as described in Reference 1)). Trajectories were acquired at 38 Hz

over a time of 13 to 26 s. Note that the contribution of active fluctuations have been shown to be

important below 10 Hz (Reference 2), and we cannot fully assert that they do not contribute to bead

fluctuations at 38 Hz. Our measurements thus provide only an estimate of the cytosol viscosity

(effective viscosity). (D) Histogram of the viscosity coefficient determined from the analysis of 1348

particles exhibiting free-like diffusion.

Reference 1: Kusumi A, Sako Y, Yamamoto M (1993) Confined lateral diffusion of membrane

receptors as studied by single particle tracking (nanovid microscopy). Effects of calcium-induced

differentiation in cultured epithelial cells. Biophys J 65:2021–2040.

Reference 2: Mizuno D, Tardin C, Schmidt CF, Mackintosh FC (2007) Nonequilibrium mechanics of

active cytoskeletal networks. Science 315:370–373.

Figure S5: Elongation of interfaces adjacent to the optically deformed interface. and denote the length of the optically-tweezed interface, prior to deformation and at maximal

deformation, respectively. Elongation of the interface n (n = 1, 2, 3 or 4), adjacent to the interface 0 is

given by . Left plot shows the ratio of over . The red line is the median, the box

edges are the lower and upper quartiles, and the whiskers display the total range of measurements.

Scale bar, 5 µm.

Figure S6: Relaxation of interfaces after trap release and deformation along the apico-basal axis. (A) After trap release, the interface relaxes with an initial velocity, which is dependent on tension and

onthe damping coefficient in the cytosol . (B) Deformation of a cell interface in regions of the

embryo where the apico-basal axis is in the plane of imaging. The epithelial cells are observed in

cross-section (left panel). Interface prior deflection (top right) and deflected interface (top left). The

deformation extends over a 4-5 µm width along the apico-basal direction (bottom right panel). The

blue lines are eye-guides. Scale bars, 10 µm (left panel), 5 µm (right panels).

Figure S7: Tension normalized to junction length along AP and DV directions.

Figure S8: Pull-release experiments and comparison to simple visco-elastic models. (A) Relaxation dynamics of the interface in linear-log representation. A simple exponential does not

fit the data (cyan solid line). The two characteristic times are visible (purple and green dashed lines as

eye-guides). (B) Alternative visco-elastic models for the pull-release experiments: the Kelvin-Voigt

model (top) is composed of a spring and a dashpot in parallel, while the Maxwell model (bottom) is

composed of a spring and a dashpot in series. In the presence of external viscosity, both predict

exponential relaxation.

Page 122: applying optical tweezers in vivo: a biophysical study of mechanical

7

Figure S9: Mechanical model of the interface and tissue response. (A) Deflection perpendicular to the interfaces is tracked over time along lines perpendicular to cell

interfaces (red). Scale bar, 10 µm. (B) Kymograph of interface deflections in the experimental (left)

and simulated (right) tissues. (C) Deflection of the target (black) and neighbor interfaces (at 1-cell, 2-

cell and 3-cell distance, magenta, green and blue, respectively) in the experimental (left) and simulated

(right) tissues. (D) Spatial decay of interface deflections over the neighboring cells. Comparison

between experiments (black circles) and simulations (red stars). (E) Propagation of deformation for

different values of cytosol viscosity (top: viscosity , middle: , bottom: . 5). A small

viscosity results in a more efficient propagation of the deformation, which becomes limited only by

the fixed boundary conditions. On the opposite, a high viscosity results in a less efficient propagation.

Supplementary Movies Movie 1: Interface deflection produced by a laser trap (red) following a sinusoidal movement of 0.5

µm amplitude and a 2 s time period.

Movie 2: Interface deformation imposed by a laser trap moving a bead against a cell-cell interface.

Movie 3: Propagation of local deformation – comparison between in vivo and in silico experiments.

Page 123: applying optical tweezers in vivo: a biophysical study of mechanical

Filter

Illumination

lens

Tube lens

Dete

ction

lens

Illumination

lens

561 nm

AOTF

Sample chamber

405 nm

488 nm

100x / 1.1 NA

10x / 0.3 NA

Laser beam (front view)

Laser beam (bottom view)

Fluorescence light

Front view

Bottom view

Detection

lens

Cylindrical

lensPeriscope

Beam

expander

Mirror

Dichroic

mirror

Dichroic

mirrorIllumination

Unit

(horizontal)

Detection

Unit

(vertical)EMCCD

Camera

Piezo stage

Z

Dichroic mirror

1070 nm

Beam

expander

Galvo

Galvo

Optical

tweezers

Fig. S1

Page 124: applying optical tweezers in vivo: a biophysical study of mechanical

Position (µm)

Optical path

diffe

rence (

nm

)

nm

QPD

A B

Fig. S2

Page 125: applying optical tweezers in vivo: a biophysical study of mechanical

B

01 2 3 4 5 6 7 8 90

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

Time (s)

Positio

n (μm

)

C

Bead

Trap

Interface

Trap

Interface

Time (s)

Positio

n (μm

)

D

01 2 3 4 5 6 7 8 90

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

A

Fig S3

Bead p

ositio

n

in t

he c

yto

sol (μ

m)

-0.2

0

0.2

Time (s)

0 1 2 3 4

Page 126: applying optical tweezers in vivo: a biophysical study of mechanical

BA

0 5 10 15 200

20

40

60

80

100

120

140

160D

Viscosity coefficient (Pa.s)

Fre

quency

C

Number of embryos 26

Number of beads 4650

Fraction of diffusive beads 0.29

Fraction of superdiffusive beads 0.705

Fraction of subdiffusive beads 0.005

Fig S4

Page 127: applying optical tweezers in vivo: a biophysical study of mechanical

Fig S5

Page 128: applying optical tweezers in vivo: a biophysical study of mechanical

0 2 4 6 8 10 12 14 161.95

2

2.05

2.1

2.15

2.2

2.25

Time(s)

V0

Trap switched off

Inte

rface p

ositio

n (μm

)

Fig S6

A

B

Page 129: applying optical tweezers in vivo: a biophysical study of mechanical

Figure S7

St.5 end St.6 end St.7 0

10

20

30

40

50

60

Tensio

n /

junction length

(pN

/μm

)

Page 130: applying optical tweezers in vivo: a biophysical study of mechanical

B

Mem

bra

ne a

nd T

rap P

ositio

ns (μm

)

Fig S8

A

25 26 27 28 29 30 31 32

106

105

104

103

102

101

Time (sec)

Mem

bra

ne P

ositio

n (μm

)

Page 131: applying optical tweezers in vivo: a biophysical study of mechanical

1 2 3

1

Deflection

(μm

)

Distance (cells)0 5 10

0

1

Experiment

0 5 10

0

1

Simulation

A

Time

Experiment Simulation

Time

Time (s)

Inte

rface d

eflection (μm

)

Time (s)

Inte

rface d

eflection (μm

)

B

C D

E

Fig. S9

experimentmodel