Coordination of Morphogenesis and Cell-Fate

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Current Biology

Review

Coordination of Morphogenesis and Cell-Fate


Specification in Development
Chii J. Chan1,*, Carl-Philipp Heisenberg2, and Takashi Hiiragi1,*
1European Molecular Biology Laboratory, 69117 Heidelberg, Germany
2Instituteof Science and Technology Austria, 3400 Klosterneuburg, Austria
*Correspondence: chii.chan@embl.de (C.J.C.), hiiragi@embl.de (T.H.)
http://dx.doi.org/10.1016/j.cub.2017.07.010

During animal development, cell-fate-specific changes in gene expression can modify the material properties
of a tissue and drive tissue morphogenesis. While mechanistic insights into the genetic control of tissue-
shaping events are beginning to emerge, how tissue morphogenesis and mechanics can reciprocally impact
cell-fate specification remains relatively unexplored. Here we review recent findings reporting how multicel-
lular morphogenetic events and their underlying mechanical forces can feed back into gene regulatory path-
ways to specify cell fate. We further discuss emerging techniques that allow for the direct measurement and
manipulation of mechanical signals in vivo, offering unprecedented access to study mechanotransduction
during development. Examination of the mechanical control of cell fate during tissue morphogenesis will
pave the way to an integrated understanding of the design principles that underlie robust tissue patterning
in embryonic development.

Introduction signals in vivo and long-term cell-lineage tracking during


The fields of physical biology and mechanobiology have bur- development.
geoned in the past decade [1–3], revealing that mechanical
forces play an integral role in tissue morphogenesis. In recent Impact of Tissue Stress on Cell-Fate Specification
years, substantial progress has been made towards understand- Cells can ‘sense’ forces through mechanosensing and mecha-
ing how mechanical cues, either extrinsically induced by the notransduction, and subsequently control their differentiation.
cellular microenvironment or intracellularly generated, can be The terms ‘mechanosensation’ and ‘mechanotransduction’
transduced into biochemical signals that regulate cell prolifera- have been used extensively, and sometimes interchangeably,
tion, migration and differentiation [3–5]. Understanding the in different biological contexts. In this review we specifically
crosstalk between tissue-scale mechanics and cell-fate specifi- define mechanosensation as the physical mechanisms by which
cation is essential to uncover the key design principles that regu- a cell senses mechanical cues, such as forces or stiffness, from
late robust tissue patterning during development. Furthermore, its microenvironment. Meanwhile mechanotransduction involves
these principles may provide a mechanistic framework for future the conversion of these mechanical cues to biochemical signals
tissue engineering efforts, with clear implications for disease downstream of mechanosensation. While such transduction
modeling and regenerative medicine [6–8]. may affect various cellular components at a post-translational
Extensive studies in the past have focused on elucidating level, here we focus on how it triggers gene expression changes
how differential gene expression can modify the physical to impact fate specification.
properties of cells and influence tissue patterning [9,10]. While Cellular Mechanisms of Mechanotransduction
these studies provide insights into how differential adhesive or In general, cellular mechanotransduction can be classified into
contractile properties may arise in different cell types to chemical or physical transduction [13]. Chemical transduction
mediate cell sorting and tissue shaping [11,12], how genetic relies on a series of chemical signaling cascades bridging the
cascades link cell-fate specification to tissue morphogenesis plasma membrane and the nucleus. When the cells experience
remains unclear. Even less well understood is how morphogen- mechanical strain, focal adhesion proteins such as zyxin and
esis and tissue mechanics can conversely feed back into paxillin can alter their binding kinetics and shuttle between the
cell-fate decisions and regulate developmental programs. In cytoplasm and nucleus to modulate transcriptional activity.
this review, we specifically discuss the potential upstream Force-induced phosphorylation of E-cadherin-bound b-catenin
roles of tissue geometry and forces in controlling cell-fate at intercellular junctions results in its translocation to the nucleus
specification during embryonic development. We first summa- and activation of transcription of downstream effectors [14–16].
rize our current understanding of the molecular mechanisms Mechanotransduction can also involve the nuclear translocation
underlying cellular mechanotransduction in vitro and in vivo. of transcription factors from the cytoplasm, such as NF-kB or
We then describe how changes in 3D tissue architecture, MAL, in response to fluid shear [17] and mechanical stress
cellular rearrangements and cytoskeletal flow can shape [18,19], respectively. Recently, the stretch-activated channels
morphogen gradients and tissue polarity to critically influence Piezo1 and Piezo2 were identified as key mechanosensors regu-
cell-fate determination. Finally, we review new biophysical tools lating calcium signaling by sensing changes in membrane
that allow the measurement and manipulation of mechanical tension [20–22] and they can potentially play a role in neural

R1024 Current Biology 27, R1024–R1035, September 25, 2017 ª 2017 Elsevier Ltd.
Current Biology

Review

stem cell differentiation [23]. Another well-studied signaling asymmetrical inheritance of less-contractile apical domains re-
cascade involved in mechanotransduction is Yap/Taz signaling, sults in a polarized daughter cell that exhibits lower cortical
where mechanical tension arising from physical stretching of contractility than its apolar sister cell (Figure 1B) [43]. The
cells or increased stiffness of the extracellular matrix (ECM) ac- difference in contractility leads to subsequent internalization of
tivates F-actin remodeling and Yap/Taz nuclear localization the apolar cell. The outer polar cell adopts the trophectoderm
[24,25]. While ECM stiffness is known to direct differentiation fate with Yap localized to the nucleus, while the apolar cell
or self-renewal of embryonic or somatic stem cells [26–30], adopts the inner cell mass fate with Yap restricted to the cyto-
ECM viscosity can also influence cell-fate specification, for plasm [44]. Notably, reducing cortical contractility leads to a
example during the osteogenic differentiation of mesenchymal loss of Yap in the nuclei of externally positioned cells, pointing
stem cells [31]. Collectively, these studies highlight the diverse to the possibility that tension in the outer cells may trigger the
set of mechanical cues that cells can sense and biochemically Yap nuclear localization required for trophectoderm cell-fate
transduce to regulate differentiation. specification.
Physical mechanotransduction relies on direct force transmis- Fluid flow also plays a critical role in driving cell-fate specifica-
sion from the cell surface to the nucleus through physical tion in development. For example, during development of the
coupling between the nuclear membrane and the extracellular circulatory system, haemodynamic forces can induce an endo-
space by cytoskeletal components. The nucleus has been pro- thelial-to-hematopoietic cell-fate switch through Runx1, a mas-
posed to act as a mechanosensor, whereby changes in nuclear ter transcription factor for hematopoiesis (Figure 1C) [45,46]. In
shape can evoke transcriptional changes by locally altering the zebrafish heart valve morphogenesis, intracardiac and pericar-
spatial accessibility of chromatin to transcriptional regulators diac fluid flow influence cardiac chamber differentiation
[13]. Importantly, the physical links between the cytoskeleton [47,48], while klf2a, an endocardial flow-responsive gene, con-
and nuclear membrane proteins allow the entire cell to function trols fibronectin synthesis and cardiac valve development [49].
as a single mechanically coupled system [32–36]. A key molec- Recent work demonstrated that Notch signaling is sensitive to
ular player involved in such physical transduction is the nuclear shear stress and plays a role in arteriovenous differentiation
lamina component lamin A [37,38]. In stiff tissue, tension from [50,51]. Another morphogenetic force that shapes cell differenti-
the ECM transmitted via the cytoskeleton to the nucleus can ation is hydrostatic pressure. For example, the sustained stretch
enhance lamin A transcription and stability of lamin A/C, leading of undifferentiated pulmonary mesenchymal cells by fluid pres-
to activation of Yap, serum response factor (SRF) and retinoic sure leads to SRF activation and smooth muscle differentiation
acid signaling pathways [37]. A recent study on mouse epidermal in lung organ cultures (Figure 1D) [52]. During lymphatic vascula-
stem cells also revealed that stress transmitted through the actin ture development, increased interstitial fluid pressure leads to
cytoskeleton can induce enrichment of the nuclear membrane stretch and proliferation of endothelial cells [53]. Pressure me-
protein emerin on the outer face of the nuclear membrane, lead- chanotransduction in urothelial cells mediated by Piezo1 leads
ing to Polycomb-mediated gene silencing and precocious differ- to ATP release, activating afferent nerves of the bladder to indi-
entiation [39]. cate bladder fullness [54,55]. Interestingly, osmotic pressure can
Mechanotransduction in Development modulate plasma membrane tension, which can influence endo-
During embryonic development, forces are generated and trans- cytosis of signaling molecules involved in cell differentiation [56].
mitted across multiple scales. The interplay between tissue me- There is also growing evidence from plant studies that organ
chanics and biochemical signaling involves multiple feedback shape changes generate mechanical signals that directly impact
interactions, rendering the assessment of mechanotransduction cell-fate determination (Figure 1E). In Arabidopsis shoot meri-
in vivo at the organismal level non-trivial. To date, mechanotrans- stems, differential growth generates anisotropic mechanical
duction has been identified in only a few developmental contexts stress, leading to tissue folding [57]. It was shown that the
where the underlying genetics, cell-fate changes and mechanics expression of SHOOT MERISTEMLESS, a key regulator of meri-
are sufficiently well characterized. stematic identity, is correlated with tissue curvature, and its
In Drosophila embryos, tissue deformations caused by germ- expression can be induced by mechanical stress through an
band extension upregulate the expression of the transcription auxin-independent mechanotransduction pathway. Since tissue
factor Twist via nuclear translocation of b-catenin (Figure 1A) folding naturally occurs during morphogenesis, such ‘shape-to-
[14]. Furthermore, both static and pulsed mechanical deforma- gene’ feedback may serve to constrain the gene regulatory
tions of mesodermal cells have been shown to promote Folded network and ensure developmental robustness. In the following
gastrulation (Fog) signaling and apical myosin II accumulation, section we discuss the influence of tissue geometry on cell-fate
which is required for mesoderm invagination prior to germ- specification.
band extension [40,41]. Interestingly, mechanical strains occur-
ring during the early phases of gastrulation in zebrafish and fly Impact of Geometry and Cytoskeletal Forces on Tissue
embryos both lead to nuclear translocation of b-catenin, thereby Patterning
triggering the expression of transcription factors required for Establishment of well-defined forms and patterns during devel-
early mesoderm identity [15]. In addition, tissue compression reg- opment requires that cells recognize their position within the em-
ulates tooth formation in mouse embryos, where morphogen- bryo and differentiate accordingly. Two current pre-eminent
induced condensation of mesenchymal cells beneath the dental concepts for tissue patterning are based on the positional infor-
epithelium leads to odontogenetic cell-fate specification [42]. mation and reaction–diffusion models [58,59]. In brief, positional
Mechanical cues also play a key role during mouse embry- information involves the graded distribution of a morphogen
onic development. In pre-implantation stage mouse embryos, across the tissue, from which the cells interpret their position

Current Biology 27, R1024–R1035, September 25, 2017 R1025


Current Biology

Review

A B
Compression of Trophectoderm fate
stomodeal cells Germ band extension Apical domain

Nuclear
β-catenin

Twist Higher contractility


Inner cell mass fate
Anterior midgut

C D

Shear stress Hydrostatic pressure


Haemodynamic flow

Nitric oxide
Mechanical stretching
of mesenchymal cells
Fluid pressure
Runx1 in lung airway
Smooth muscle
differentiation

Endothelial–hematopoietic
cell fate transition

E
Tissue folding

Mechanical stress
Shoot
meristem
STM and
meristematic identity

STM is upregulated at curved regions of


boundary domains
Current Biology

Figure 1. Examples of mechanotransduction during embryonic and adult organ development.


(A) Tissue-scale compression in Drosophila embryos caused by germ-band extension triggers Src42A-dependent nuclear translocation of b-catenin in anterior
stomodeal cells, thereby activating the transcription of Twist, which is crucial for subsequent midgut differentiation. (B) In early mouse embryonic development,
asymmetrical inheritance of the apical domain results in a polarized daughter cell that exhibits lower cortical contractility than its apolar sister cell. The difference
in contractility leads to internalization of the apolar cell. The outer polar cell adopts the trophectoderm fate with nuclear Yap localization while the apolar cell
adopts the inner cell mass fate with cytosolic Yap localization. (C) During the formation of the hematopoietic system, shear stress can induce nitric oxide
production and upregulation of Runx1 transcriptional factor, thereby inducing endothelial-to-hematopoietic cell-fate switch. (D) Mechanical stretching of the lung
drives myogenesis in airway smooth muscle, which is crucial for bronchial development. This process is mediated by the activation of the transcription factor
SRF. (E) In Arabidopsis, strong mechanical stresses arising from tissue folding at the shoot apical meristem can induce the expression of STM, a master regulator
of meristematic identity.

to make fate choices. The morphogen concentration therefore based purely on biochemical signaling. Here we discuss the
provides positional coordinates along the embryonic axis. The new possibility that cytoskeletal forces, cell adhesion and polar-
reaction–diffusion model involves a self-amplifying local signal ity, and tissue-scale mechanical signals can confer positional in-
and a long-range inhibitor that is stimulated by the former. Inter- formation and control gene expression and fate at the cellular
actions between these two components lead to the generation of level, thereby allowing cells to coordinate embryonic patterning.
periodic patterns of diffusible morphogens and consequently Cytoskeletal Forces Drive Tissue Patterning
cell fates. While the significance of both models has long been Forces generated by motor proteins can couple with biochem-
recognized in developmental biology, these models are so far ical signals to generate morphogen gradients and patterns

R1026 Current Biology 27, R1024–R1035, September 25, 2017


Current Biology

Review

A Cytoskeletal force and advection

Advection of aPAR by
actomyosin-dependent cytoplasmic flow

Stably polarized
C. elegans zygote

pPAR binds to
posterior membrane

B Position sensing via polarity C Multicellular rearrangement

FGF trapped in lumen

Lateral line
primodium

Differential cell contact Apical domain Trophectoderm fate Luminal signaling leads to enhanced differentiation

D Geometric constraints

TGF-β localized at apical surface

Sonic
hedgehog

Edge Center
Mechanical
buckling
BMP4

Local increase in morphogen


induces villus cluster NOGGIN

Current Biology

Figure 2. Cytoskeletal forces and geometric cues can mechanically feed back to morphogen signaling.
(A) Mechanochemical patterning in C. elegans zygote is established through the advection of anterior PAR proteins (red) by actomyosin-driven cytoplasmic flow.
The binding of posterior PARs (blue) to the posterior membrane results in the amplification and stabilization of polarity by a reaction–diffusion mechanism. (B) In
8-cell early mouse embryos, contact asymmetry induces apical domain formation at the contact-free cell surface. Upon division, the daughter cells that inherit the
apical domain adopt the trophectoderm fate while the apolar daughter cells may or may not become trophectoderm, depending on their eventual position.
(C) Luminal signaling can impact cell-fate acquisition in a multicellular context. In migrating zebrafish lateral line primordium, concerted apical constriction leads
to the formation of a microlumen. Local trapping of FGF in these microlumens leads to enhanced signaling and restricted cellular differentiation in the neighboring
cells. (D) Left: In the mouse gut, mechanical buckling of the epithelium due to growth leads to a local build-up of Shh signals in the villi, which then restricts
progenitor cells to the base of these structures. Right: Geometric confinement of human embryonic stem cells leads to the spatial ordering of germ layers
recapitulating human gastrulation. In the presence of BMP4, differential localization of TGF-b receptors at the colony edge (apical) and center (basolateral)
triggers differential signaling and generates a reaction–diffusion mechanism that involves the BMP inhibitor Noggin.

[60,61]. In Caenorhabditis elegans zygotes, actomyosin contrac- such chiral actomyosin flows can also displace the midbody, a
tion induces large-scale, anterior-directed cortical flows, which remnant of the previous division, to the future ventral side of
in turn transport polarity and cell-fate-determining proteins to the embryo, thereby influencing spindle orientation and estab-
the anterior half of the cell [62]. Importantly, if contraction itself lishing dorsoventral axis patterning [65]. These studies clearly
causes flows that carry regulators of contraction, such as parti- highlight the role of differential cell-contact signaling in partition-
tioning defective (PAR) proteins, this will lead to positive feed- ing cell fates along the body axis.
back and subcellular patterning (Figure 2A) [63]. Following Microscopic fluid flow can also drive cell differentiation and
division, asymmetrical cortical contractility between sister cells embryonic patterning through chiral machineries in tissues. Dur-
in turn induces chiral cortical actomyosin flows, thereby driving ing gastrulation in several vertebrate species, posteriorly tilted
skewing of the spindle orientation at the 4-cell stage and estab- cilia in the ventral node rotate in a clockwise fashion to induce
lishing the left–right body axis [64]. From the 2- to 4-cell stage, leftward extracellular fluid flow [66]. This unidirectional flow has

Current Biology 27, R1024–R1035, September 25, 2017 R1027


Current Biology

Review

been proposed to trigger asymmetrical transport of morpho- cell mass of mouse blastocysts, cells undergo active cell sorting
gens, such as Sonic hedgehog (Shh) and retinoic acid, ultimately to form two distinct tissues, the epiblast and the primitive endo-
leading to left–right asymmetrical gene expression throughout derm [79,80]. Such sorting behavior was suggested to arise from
the entire body [67]. Alternatively, the unidirectional flow has the progressive polarization and epithelialization of the future
been suggested to trigger a mechanosensitive response in the primitive endoderm facing the blastocoel, leading to the spatial
cilia at the edge of the node, which in turn activates subsequent separation of primitive endoderm cells from the epiblast cluster
asymmetrical signaling [68,69]. [81]. Differential cortical tension and/or adhesion in the inner
Position Sensing at the Cellular Level cell mass have also been proposed to contribute to active cell
For early mouse embryogenesis, the inside–outside model pro- sorting during lineage segregation [80], although these mechan-
poses that cell fate depends on the position of a cell at the morula ical properties remain uncharacterized.
stage, with cells on the inside forming the inner cell mass and In the zebrafish neural tube, cell movements mediated by
cells on the outside taking up the trophectoderm fate [70]. How- changes in cell–cell adhesion appear to correct the initially noisy
ever, the underlying mechanism remained elusive. A recent study Shh signaling pattern to organize the neural progenitor cells into
shed new light in this direction by showing that, at the eight-cell strictly demarcated domains [82]. Changes in tissue architecture
stage, contact asymmetry induces apical domain formation on can also feed back directly onto the activity of signaling centers.
the contact-free cell surface (Figure 2B) [71]. Upon division, the During zebrafish lateral line formation, fibroblast growth factor
daughter cells that inherit the apical domain adopt the trophecto- (FGF) activity within the tissue controls the frequency at which
derm fate, while the apolar daughter cells may or may not rosette-like mechanosensory organs are deposited [83]. Here,
become trophectoderm, depending on whether they are pushed the lateral line primordium, consisting of about 100 cells, mi-
out to the surface due to mechanical constraint. This work there- grates along the length of the developing body. During this pro-
fore identified the presence of a contact-free cell surface and api- cess, concerted apical constriction of discrete groups of cells
cal polarization as crucial determinants distinguishing the leads to the formation of rosette-like structures and the assem-
‘outside’ from ‘inside’ positions, thereby directing cell fate. bly of a central microlumen that traps the secreted FGF
Tissue Tension Orients Cell Division and Polarity (Figure 2C). The local increase in FGF ligand concentration in
There is growing evidence that tissue-scale tension can orient turn creates a positive feedback loop by maintaining the apical
cell division patterns and polarity during development [72,73], constriction process and rosette formation [84]. Luminal
which can influence cell-fate determination. The mechanosens- signaling therefore represents an excellent mechanism by which
ing mechanism for Hertwig’s rule describes how astral microtu- cellular rearrangement can feed back to locally restrict and
bule length and spindle positioning are influenced by cell shape enhance molecular signaling and direct cell differentiation within
to result in division along the long axis of the cell. A recent study migrating tissues.
in Drosophila epithelia further revealed that the tricellular junc- Geometric Constraints Modulate Spatial Patterning of
tions (TCJs), to which the dynein-associated protein Mud Cell Fate
localizes and pulls on the astral microtubules, can translate infor- Recent studies have highlighted the extensive interplay between
mation about interphase cell shape anisotropy into a ‘stable’ tissue geometry and signaling. Physical deformation of the
biochemical cue, thereby determining spindle orientation during morphogenetic field in three dimensions can generate graded
mitotic rounding [74]. Mechanical stress has also been shown to signaling activities to pattern tissues. For example, during verte-
induce tissue-level planar cell polarity (PCP) in development. In brate gut formation proliferating intestinal stem cell progenitors,
the Drosophila wing disc, global anisotropic tension originating which are initially evenly distributed throughout the epithelium,
from the contraction of the wing hinge results in the reorientation become restricted to the base of the villi following tissue buckling
of intercellular junctions, cell elongation and tissue flow, which [85,86]. Physical buckling due to continued growth under me-
collectively establish long-range PCP along the proximal–distal chanical constraint causes the epithelial Shh signal to become
axis [75,76]. locally concentrated at the villi tips (Figure 2D, left) leading to vil-
Generation of cell-fate diversity from oriented cell division is lus cluster formation. Local maxima of Shh signals then induce
also observed in vertebrates. In Xenopus embryos, the deep reciprocal signaling by bone morphogenetic proteins (BMPs)
and superficial cells at the blastula stage show differential within the cluster, repressing intestinal stem cell identity in the
gene expression for primary neurogenesis. The distinct fate of overlying epithelium and restricting stem cell proliferation to
deep versus superficial cells was proposed to be generated by the base of the villi. Another example comes from post-implanta-
oriented cell divisions along the apicobasal axis at the surface tion stage mouse embryos, where mechanical constraints from
of the embryo, giving rise to a superficial cell with an inherited the maternal uterine tissues have been proposed to lead to the
apical domain marked by atypical protein kinase C (aPKC) and elongated cylindrical shape of the embryos and thinning of the
an apolar deep cell [77]. Similarly, in the early zebrafish gastrula, basement membrane at the distal tip [87]. This induces the for-
cell-division orientation is determined by cell-shape changes, mation of distal visceral endoderm cells that play a role in embry-
which are in turn regulated by tissue tension and geometrical onic anterior–posterior axis patterning. Interestingly, a recent
constraints imposed by the overall tissue surface area [78]. study demonstrated that the distal visceral endoderm can still
This leads to the robust formation and segregation of surface form in the absence of external mechanical constraints [88].
epithelial and deep cells. The discrepancy may result from differences in the culture con-
Cellular Rearrangements Shape Tissue Pattern ditions or methods of expression studies.
Active cell sorting mediated by changes in cell adhesion and Geometric cues can also trigger self-organized patterning dur-
cortical tension can also drive tissue patterning. In the inner ing embryonic development. Recent in vitro studies using human

R1028 Current Biology 27, R1024–R1035, September 25, 2017


Current Biology

Review

A Measuring forces and stiffness in vivo


Microdroplet
ΔP
Tension sensor
module F-actin

Pressure
sensor

FRET biosensor
Micropipette aspiration
Droplet-based sensor

High stiffness

Pressure

Tension
2D epithelium

Force inference
Brillouin microscopy

B Spatio-temporal control of forces in vivo

Optogenetics Magnetic tweezer


Blue laser

Inhibition of apical constriction


blocks mesoderm invagination Magnetic liposomes

C Ex vivo reduced systems

Organoid technology
Micropatterning

Intestinal organoid
Polarity inversion
controlled by
substrate stiffness

3D hydrogel with tunable


mechanical properties e.g.
Cell doublets on 2D micropattern
ligand density, matrix stiffness

Current Biology

Figure 3. Techniques to study morphogenetic feedback mechanisms in vivo.


(A) FRET biosensor: FRET tension sensor with known spring stiffness is inserted between F-actin molecules. The length of module increases with the tension
between the molecules, resulting in lower FRET efficiency. Quantifying the FRET efficiency generates a spatiotemporal map of endogenous forces in vivo.
Micropipette aspiration: Negative aspiration pressure is applied through a micropipette in contact with the mouse embryo blastomeres. Using the Young-Laplace
equation, the surface tension of the blastomeres can be measured during mouse compaction development. Droplet-based sensor: Microdroplets coated with
fluorescent surfactant molecules are inserted into the tissue. With the known surface tension and 3D deformation of the droplets, the anisotropic stress of the
(legend continued on next page)

Current Biology 27, R1024–R1035, September 25, 2017 R1029


Current Biology

Review

‘gastruloids’ (3D aggregates of embryonic stem cells) grown on (Figure 3A) and revealed the critical role of cortical contractility
micropatterned substrates demonstrated that spatial confine- in oocyte meiosis, compaction and lineage segregation
ment of human embryonic stem cells can trigger spatial [43,97,98]. The technique involves the local aspiration of a part
patterning similar to that observed during gastrulation [89,90]. of a cell or tissue by applying a constant negative pressure.
Mechanistically, transforming growth factor b (TGF-b) receptors The surface tension of the sample can then be calculated based
were found to be differentially localized to the apical and baso- on the Young-Laplace equation [97]. One major limitation of
lateral surfaces of the cells at the edges and center of the colony, micropipette aspiration is that it cannot be used to measure force
respectively, thereby triggering different signaling responses generation by cells within the interior of 3D tissues. To address
across the colony in the presence of the BMP4 ligand this issue, droplet-based sensors have been developed
(Figure 2D, right). BMP4 subsequently activates Noggin, a (Figure 3A): cell-sized oil microdroplets coated with adhesion
secreted BMP4 inhibitor, and generates self-organized gene ligands are injected into 3D tissues to serve as force transducers
expression patterning through a reaction–diffusion mechanism in vivo. Mechanical stresses exerted by the surrounding cells will
[90]. In this case, the geometric cues that allow the cells to deform the microdroplet. As a result, the tissue stress can be
‘sense’ the edge of the tissue come from the boundary condition deduced from the known interfacial tension and shape deforma-
that provides the initial receptor pre-pattern. tion of the microdroplet [99]. Additionally, pressure sensors made
of polyacrylamide microbeads have been developed to quantify
In Vivo Techniques to Study Morphogenetic Feedback mechanical stress in multicellular spheroids [100], offering the
Mechanisms exciting prospect of tracking tissue or luminal pressure in devel-
The study of how tissue morphogenesis and mechanics may in- oping embryos.
fluence cell-fate specification during development requires Perhaps the least invasive technique is the direct ‘imaging’ of
simultaneous imaging and/or perturbation of tissue forces, forces in vivo using the force inference method [101,102].
morphology and cell fate in vivo. In this section, we review Assuming that the tissue is in a quasi-static equilibrium state
recent technological progress in these areas and discuss where the deformation of cells is sufficiently slow, one can
complementary approaches of ex vivo assays that can disen- extract mechanical information about the tissue by analyzing
tangle the contribution of each parameter governing cell-fate the shape of the cells, which is governed by the local balance
specification. of the forces that the cells exert on one another (Figure 3A).
Non-Invasive Approaches to Measure Tissue Stress and While this method allows quantification of relative cell–cell junc-
Stiffness In Vivo tional tension and cellular pressure up to a prefactor, cross-
A number of approaches have been developed to quantify validation with other techniques can help provide an absolute
forces in living tissues and embryos [91,92]. Here, we limit measure of forces. So far, force inference methods have
our discussion to selected techniques that allow non-invasive, been limited to 2D systems, such as epithelial monolayers in
spatiotemporal mapping of tissue mechanics at develop- Drosophila embryos and pupae [76,102–104]. However,
mental timescales. One technique that offers high spatiotem- promising work extending the approach to 3D systems has
poral resolution of intracellular or intercellular tensions in living recently been reported for early mouse and zebrafish embryos
tissues is the use of Forster resonance energy transfer (FRET)- [105,106].
based tension sensors [93,94]. The FRET signal intensity de- With substantial evidence supporting the role of ECM stiffness
pends on the extension of a molecular spring between the in directing cell-fate specification in vitro, a label-free bioimaging
donor and acceptor fluorophore, which can be anchored to technique known as Brillouin microscopy has been developed to
molecular components of adherens junctions, such as a-cate- map tissue stiffness in vivo [107]. By analyzing the frequency shift
nin and E-cadherin (Figure 3A). Once the spring constant is of the scattered light when it interacts with sound waves in the
calibrated, the FRET signal provides an effective readout sample, one can obtain a spatiotemporal map of cell or tissue
about molecular tension at the cell–cell contact during devel- stiffness in situ (Figure 3A). While previous studies have demon-
opment. Other tools include a-actinin tension sensors, which strated its feasibility in measuring intracellular or cell mechanical
have been developed to reveal lineage-dependent intercellular properties with submicron resolution [108–110], recent studies
tension during Xenopus laevis embryonic development [95]. also provide proof-of-principle applicability in measuring the me-
Similarly, b-spectrin tension sensors allow visualization of chanical properties of ECMs in whole organisms [111]. Future
tissue stresses mediated via b-spectrin during C. elegans studies combining Brillouin microscopy with fluorescence imag-
embryogenesis [96]. ing of cell-fate reporters and mechanoreceptors will likely reveal
Micropipette aspiration has recently been used to quantify sur- new insights into the role of ECM stiffness and cell-fate specifi-
face tension of blastomeres during early mouse embryogenesis cation during development.

tissue can be determined. Pressure sensors made of polyacrylamide microbeads yield the isotropic pressure distribution in 3D tissues. Force inference: Image
analysis of a 2D epithelium generates a segmented cell shape and angles between cell–cell junctions. Using force-balance equations at the vertex points, the
junctional tension and intracellular pressures of all cells can be inferred. Brillouin microscopy: Pressure waves are generated in the sample following illumination
with light of very high frequency. By measuring the frequency shift of the scattered light, the mechanical stiffness of the sample can be measured. (B) Left: Using
laser light to spatiotemporally modulate protein activities governing cell contractility, tissue tension and geometry were shown to influence mesoderm invagi-
nation in Drosophila. Right: Using a magnet, cells loaded with magnetic liposomes are mechanically towed and the impact of ectopic stress on gene expressions
can be monitored. (C) Left: 2D micropatterning of cell doublets revealed the role of matrix stiffness in polarity orientation and epithelial-to-mesenchymal cell-fate
transition. Right: Using a 3D hydrogel with tunable ECM parameters, high matrix stiffness is shown to enhance intestinal stem cell expansion while a soft matrix
leads to differentiation of these cells and organoid formation.

R1030 Current Biology 27, R1024–R1035, September 25, 2017


Current Biology

Review

Visualizing Cell Fate Specification and Tissue fish embryos [125]. It will be interesting to apply ectopic forces
Morphogenesis at such mesoscopic scales and study their downstream effects
To study the interplay between tissue morphogenesis and on signaling and cell-fate specification.
cell-fate specification in development, it is essential to monitor
cell-fate changes in live cells. Recent advances in imaging tech- Ex Vivo Reduced Systems
niques have allowed accurate reconstruction of cell lineages Given that tissue morphogenesis is inherently a non-linear, self-
and monitoring of gene expression dynamics in multicellular organizing process involving multiple feedback mechanisms,
organisms. The low phototoxicity associated with light-sheet mi- ex vivo and in vitro experimental systems have been engineered
croscopy ensures minimal photodamage when imaging living in which the function of each mechanical cue can be assessed
embryos [112]. In addition, the improved computational frame- independently. One such experimental model employed in early
work opens up the prospect of automated image segmentation mouse embryology is the analysis of single blastomeres isolated
and cell tracking in real time [113]. Meanwhile, label-free from 8-cell stage embryos (1/8-cell) [70]. Such a reduced system
nonlinear microscopy offers an alternative approach to trace is able to develop into mini-blastocysts and recapitulate the line-
cell lineages in deep tissues with high spatiotemporal resolution age segregation of inner cell mass and trophectoderm under
[114]. To further elucidate the decisive role of mechanical forces spatially simplified conditions [126]. This approach allows the
on cell-fate specification during development, it will be essential unambiguous identification of symmetry-breaking cues, such
to simultaneously monitor cell fate, cellular/tissue stress and as apical–basal polarity [71] and differential cortical contractility
morphology. This will require future development and integration [43], in driving the first lineage segregation in early mouse devel-
of advanced microscopy, biophysical tools and cell-fate re- opment. The use of minimal tissue models, such as cell doublets
porters, as demonstrated in recent studies [78,82,104,115,116]. on confined micropatterns, also reveals the role of centriole re-
Spatiotemporally Controlled Functional Perturbations positioning in driving polarity inversion during epithelial-to-
A powerful approach to decipher the causal links between tissue mesenchymal cell-fate transition (Figure 3C, left) [127].
stress and gene expression is functional perturbation. Recent Over the past decades, 3D organoids have served as excellent
technological advances have made it possible to manipulate tis- in vitro model systems to study organ development [6,7]. While
sue forces in a spatiotemporally controlled manner within the these studies employ primarily chemical cues, new findings
embryo. One such approach is optogenetics, which makes use have revealed the roles of biophysical cues in controlling cell
of pulses of laser light to inhibit or stimulate protein activity fate and tissue organization. For example, intestinal stem cells
with spatial and temporal precision [117–119]. By optically cultured on a stiff matrix with fibronectin-based adhesion display
manipulating the activity of plasma membrane phosphoinositi- enhanced expansion through a Yap-dependent mechanism,
des and cell contractility (Figure 3B, left), local inhibition of apical while those grown on a soft matrix with laminin-based adhesion
constriction was shown to be sufficient to globally arrest meso- differentiate and form organoids (Figure 3C, right) [30]. Studies of
derm invagination during ventral furrow formation in Drosophila neuroepithelial cyst formation using 3D hydrogels have identified
[120]. Using the same approach, local activation of the GTPase key mechanical parameters determining dorsoventral patterning
RhoA was found to be sufficient for inducing cleavage furrow during neural tube morphogenesis [128]. Further, the use of
formation in rounded interphase cells, in a manner uncoupled bone-marrow-like ECM has revealed the role of matrix stiffness
from usual spatiotemporal restrictions [121]. The possibility of in directing hematopoietic stem cell fate decisions [129]. Inter-
inducing de novo assembly or disassembly of actomyosin net- estingly, advances in microdevice fabrication technology have
works allows for the sufficiency of mechanical forces in regu- led to the concept of ‘development-on-chip’. For example, it
lating cell-fate specification to be tested in vivo. Given the recent was demonstrated that neural tube development can be recapit-
demonstration of compensatory networks that buffer against ulated in microfluidic systems with spatiotemporally tunable
deleterious mutations [122], it will be important to investigate chemical environments [130]. In the future, the combination
mechanisms governing development not only using permanent of microfluidic platforms with 3D hydrogels equipped with
gene deletion, but also using spatiotemporally and quantitatively tunable geometric/mechanical properties may represent an
controlled manipulations of gene expression and protein distri- integrated approach to study the interplay between mechanics
bution in vivo. An example of this approach is the use of GFP and biochemical signaling in cell-fate specification and tissue
nanobodies to abolish Dpp gradient formation in the Drosophila patterning.
wing imaginal disc, which has provided new insights into the
mechanisms of tissue size control [123]. Future Directions
Another approach to physically manipulate forces in vivo is the In this review, we discussed recent evidence for tissue-shaping
magnetic tweezer, a tool extensively used for cells in vitro events and tissue stress as important upstream regulators of
(Figure 3B, right). The application of magnetic forces on injected developmental signaling pathways. An emerging theme for
magnetic nanoparticles or liposomes in living tissues has been studies of mechanotransduction in development is the extension
shown to induce the expression of Twist and reveal the role of from 2D epithelial systems to 3D tissue morphogenetic pro-
b-catenin in mechanotransduction in Drosophila and zebrafish cesses. Mechanical constraints arising from tissue bending or
embryos [14,15]. The same technique was used to study tumor buckling can generate long-range tissue stress that propagates
progression in adult mice, where mechanically induced stress to individual cells and triggers potential mechanotransduction
can trigger the tumorigenic b-catenin pathway [124]. Recently, pathways. Dynamic cellular rearrangements and global shape
the magnetic tweezer has been successfully applied to ferrofluid changes of tissues can also direct the trapping and positioning
microdroplets to measure tissue viscosity and fluidity in zebra- of signaling centers, leading to spatial restriction of cell lineages.

Current Biology 27, R1024–R1035, September 25, 2017 R1031


Current Biology

Review

Importantly, the interplay between mechanical signals and 4. Heisenberg, C.P., and Bellaı̈che, Y. (2013). Forces in tissue morphogen-
esis and patterning. Cell 153, 948–962.
cellular polarity can lead to the establishment of multicellular
patterning. 5. Miller, C.J., and Davidson, L.A. (2013). The interplay between cell signal-
Several outstanding questions and challenges remain ling and mechanics in developmental processes. Nat. Rev. Genet. 14,
733–744.
regarding the coordination of tissue morphogenesis, mechanics
and cell-fate specification during development. How do individ- 6. Sasai, Y. (2013). Cytosystems dynamics in self-organization of tissue ar-
chitecture. Nature 493, 318–326.
ual cells sense the global change in tissue shape and size and
transduce these physical signals into differential gene expres- 7. Lancaster, M.A., and Knoblich, J.A. (2014). Organogenesis in a dish:
sion in a highly robust manner? Usually there is considerable modeling development and disease using organoid technologies. Sci-
ence 345, 1247125.
delay between the primary events of mechanosensation,
involving the translation of mechanical signals into biochemical 8. Fatehullah, A., Tan, S.H., and Barker, N. (2016). Organoids as an in vitro
model of human development and disease. Nat. Cell Biol. 18, 246–254.
ones, and downstream transcriptional events. This implies that
cell-fate specification during development may arise from a 9. Heller, E., and Fuchs, E. (2015). Tissue patterning and cellular mechanics.
combination of physical force transduction and biochemical J. Cell Biol. 211, 219–231.
signaling induced by tissue morphogenesis. How can we 10. Gilmour, D., Rembold, M., and Leptin, M. (2017). From morphogen to
disentangle these two clearly distinct processes that act in morphogenesis and back. Nature 541, 311–320.
concert during development? For example, tissue strain also 11. Krieg, M., Arboleda-Estudillo, Y., Puech, P.-H., Ka €fer, J., Graner, F.,
drives cellular movement, which makes it difficult to distinguish Müller, D.J., and Heisenberg, C.-P. (2008). Tensile forces govern germ-
the direct impact of mechanical stress on differentiation layer organization in zebrafish. Nat. Cell Biol. 10, 429–436.
from secondary induction due to new contacts formed between 12. Maı̂tre, J.-L., Berthoumieux, H., Krens, S.F.G., Salbreux, G., Jülicher, F.,
the signaling and responding cells. Can we always devise Paluch, E., and Heisenberg, C.-P.C.-P. (2012). Adhesion functions in cell
sorting by mechanically coupling the cortices of adhering cells. Science
the necessary experimental approach to test not only for
338, 253–256.
requirement but also for sufficiency of a particular mechano-
transduction pathway, and in a quantitative manner? Here, 13. Shivashankar, G.V. (2011). Mechanosignaling to the cell nucleus and
gene regulation. Annu. Rev. Biophys. 40, 361–378.
computational simulation may offer one approach in identifying
key parameters sufficient for recapitulating developmental pro- 14. Desprat, N., Supatto, W., Pouille, P., Beaurepaire, E., and Farge, E.
(2008). Tissue deformation modulates Twist expression to determine
cesses in vivo [131].
anterior midgut differentiation in Drosophila Embryos. Dev. Cell 15,
Importantly, tissue morphogenesis characterized by feedback 470–477.
interactions between cell contact, mechanics, polarity and gene
15. Brunet, T., Bouclet, A., Ahmadi, P., Driquez, B., Brunet, A., Henry, L., Ser-
expression acts across multiple scales (organism, organ, multi- man, F., Gaelle, B., Menager, C., Dumas-Bouchiat, F., et al. (2013).
cellular, cellular and subcellular). It is therefore important to Evolutionary conservation of early mesoderm specification by mechano-
remember that the cell-fate switch itself can change cell/tissue transduction in Bilateria. Nat. Commun. 4, 1–15.
mechanical properties, such as cell adhesion and contractility, 16. Benham-Pyle, B.W., Pruitt, B.L., and Nelson, W.J. (2015). Mechanical
thereby modifying tissue-scale forces that eventually feed back strain induces E-cadherin-dependent Yap1 and -catenin activation to
drive cell cycle entry. Science 348, 1024–1027.
into cell-fate specification. How can we then specifically test
for the causal role of tissue mechanics in controlling develop- 17. Tzima, E., Irani-Tehrani, M., Kiosses, W.B., Dejana, E., Schultz, D.A., En-
mental fate changes, and vice versa? One approach is to gelhardt, B., Cao, G., DeLisser, H., and Schwartz, M.A. (2005). A mecha-
nosensory complex that mediates the endothelial cell response to fluid
spatiotemporally manipulate cell signaling, which might link shear stress. Nature 437, 426–431.
mechanics to cell-fate specification. Dissecting the specific
18. Somogyi, K., and Rørth, P. (2004). Evidence for tension-based regulation
instructive or permissive roles of mechanical forces in driving of Drosophila MAL and SRF during invasive cell migration. Dev. Cell 7,
cell-fate specification will continue to be an important and 85–93.
challenging goal in the field of developmental biology in the years
19. Vartiainen, M.K., Guettler, S., Larijani, B., and Treisman, R. (2007). Nu-
to come. clear actin regulates dynamic subcellular localization and activity of the
SRF cofactor MAL. Science 316, 1749–1753.
ACKNOWLEDGEMENTS 20. Wu, J., Lewis, A.H., and Grandl, J. (2017). Touch, tension, and transduc-
tion – the function and regulation of Piezo ion channels. Trends Biochem.
We thank François Graner and members of the Hiiragi lab for insightful discus- Sci. 42, 57–71.
sions. We sincerely apologize for not being able to include citations to all rele-
vant works in the field due to space limitations. C.J.C. is supported by the 21. Li, J., Hou, B., Tumova, S., Muraki, K., Bruns, A., Ludlow, M.J., Sedo, A.,
German Research Foundation (DFG) and Marie Curie Actions COFUND (EC Hyman, A.J., McKeown, L., Young, R.S., et al. (2014). Piezo1 integration
Grant Agreement No. 664726) EMBL Interdisciplinary Postdoc (EIPOD). of vascular architecture with physiological force. Nature 515, 279–282.

22. Ranade, S.S., Woo, S.-H., Dubin, A.E., Moshourab, R.A., Wetzel, C., Pet-
rus, M., Mathur, J., Begay, V., Coste, B., Mainquist, J., et al. (2014).
REFERENCES
Piezo2 is the major transducer of mechanical forces for touch sensation
in mice. Nature 516, 121–125.
1. Wozniak, M.A., and Chen, C.S. (2009). Mechanotransduction in develop-
ment: a growing role for contractility. Nat. Rev. Mol. Cell Biol. 10, 34–43. 23. Pathak, M.M., Nourse, J.L., Tran, T., Hwe, J., Arulmoli, J., Trang, D., and
Le, T. (2014). Stretch-activated ion channel Piezo1 directs lineage choice
2. Keller, R. (2012). Physical biology returns to morphogenesis. Science in human neural stem cells. Proc. Natl. Acad. Sci. USA 111, 16148–16153.
338, 201–203.
24. Dupont, S., Morsut, L., Aragona, M., Enzo, E., Giulitti, S., Cordenonsi, M.,
3. Mammoto, T., Mammoto, A., and Ingber, D.E. (2013). Mechanobiology Zanconato, F., Le Digabel, J., Forcato, M., Bicciato, S., et al. (2011). Role
and developmental control. Annu. Rev. Cell Dev. Biol. 29, 27–61. of YAP / TAZ in mechanotransduction. Nature 474, 179–183.

R1032 Current Biology 27, R1024–R1035, September 25, 2017


Current Biology

Review
25. Aragona, M., Panciera, T., Manfrin, A., Giulitti, S., Michielin, F., Elvassore, 44. Nishioka, N., Inoue, K., ichi, Adachi, K., Kiyonari, H., Ota, M., Ralston, A.,
N., Dupont, S., and Piccolo, S. (2013). A mechanical checkpoint controls Yabuta, N., Hirahara, S., Stephenson, R.O., Ogonuki, N., et al. (2009). The
multicellular growth through YAP / TAZ regulation by actin-processing Hippo signaling pathway components Lats and Yap pattern Tead4 Activ-
factors. Cell 154, 1047–1059. ity to distinguish mouse trophectoderm from inner cell mass. Dev. Cell
16, 398–410.
26. Engler, A.J., Sen, S., Sweeney, H.L., and Discher, D.E. (2006). Matrix
elasticity directs stem cell lineage specification. Cell 126, 677–689. 45. Adamo, L., Naveiras, O., Wenzel, P.L., Mckinney-freeman, S., Mack,
P.J., Gracia-sancho, J., Suchy-dicey, A., Yoshimoto, M., Lensch,
27. Chowdhury, F., Li, Y., Poh, Y.C., Yokohama-Tamaki, T., Wang, N., and M.W., Yoder, M.C., et al. (2009). Biomechanical forces promote embry-
Tanaka, T.S. (2010). Soft substrates promote homogeneous self-renewal onic haematopoiesis. Nature 459, 1131–1135.
of embryonic stem cells via downregulating cell-matrix tractions. PLoS
One 5, 1–10. 46. North, T.E., Goessling, W., Peeters, M., Li, P., Ceol, C., Lord, A.M.,
Weber, G.J., Harris, J., Cutting, C.C., Huang, P., et al. (2009). Hematopoi-
28. Chowdhury, F., Na, S., Li, D., Poh, Y.C., Tanaka, T.S., Wang, F., and etic stem cell development is dependent on blood flow. Cell 137,
Wang, N. (2009). Material properties of the cell dictate stress-induced 736–748.
spreading and differentiation in embryonic stem cells. Nat. Mater 9,
82–88. 47. Hove, J.R., Köster, R.W., Forouhar, A.S., Acevedo-Bolton, G., Fraser,
S.E., and Gharib, M. (2003). Intracardiac fluid forces are an essential
29. Przybyla, L., Lakins, J.N., and Weaver, V.M. (2016). Tissue mechanics epigenetic factor for embryonic cardiogenesis. Nature 421, 172–177.
orchestrate Wnt-dependent human embryonic stem cell differentiation.
Cell Stem Cell 19, 462–475. 48. Peralta, M., Steed, E., Monduc, F., Rayo, T., Ariza-cosano, A., Corte, A.,
Rayon, T., Gomez-Skarmeta, J.-L., Zapata, A., Vermot, J., et al. (2013).
30. Gjorevski, N., Sachs, N., Manfrin, A., Giger, S., Bragina, M.E., Ordóñez- Heartbeat-driven pericardiac fluid forces contribute to epicardium
Morán, P., Clevers, H., and Lutolf, M.P. (2016). Designer matrices for in- morphogenesis. Curr. Biol. 23, 1726–1735.
testinal stem cell and organoid culture. Nature 539, 560–564.
49. Steed, E., Faggianelli, N., Roth, S., Ramspacher, C., Concordet, J.P., and
31. Chaudhuri, O., Gu, L., Klumpers, D., Darnell, M., Bencherif, S.A., Weaver, Vermot, J. (2016). Klf2a couples mechanotransduction and zebrafish
J.C., Huebsch, N., Lee, H.-P., Lippens, E., Duda, G.N., et al. (2015). Hy- valve morphogenesis through fibronectin synthesis. Nat Commun. 7,
drogels with tunable stress relaxation regulate stem cell fate and activity. 11646.
Nat. Mater 15, 326–333.
50. Gordon, W.R., Zimmerman, B., He, L., Miles, L.J., Huang, J., Tiyanont, K.,
32. Crisp, M., Liu, Q., Roux, K., Rattner, J.B., Shanahan, C., Burke, B., Stahl, McArthur, D.G., Aster, J.C., Perrimon, N., Loparo, J.J., et al. (2015). Me-
P.D., and Hodzic, D. (2006). Coupling of the nucleus and cytoplasm: role chanical allostery: evidence for a force requirement in the proteolytic
of the LINC complex. J. Cell Biol. 172, 41–53. activation of Notch. Dev. Cell 33, 729–736.

33. Na, S., Collin, O., Chowdhury, F., Tay, B., Ouyang, M., Wang, Y., and 51. Jahnsen, E.D., Trindade, A., Zaun, H.C., Lehoux, S., Duarte, A., and
Wang, N. (2008). Rapid signal transduction in living cells is a unique feature Jones, E.A.V. (2015). Notch1 is pan-endothelial at the onset of flow and
of mechanotransduction. Proc. Natl. Acad. Sci. USA 105, 6626–6631. regulated by flow. PLoS One 10, 1–14.

34. Wang, N., Tytell, J.D., and Ingber, D.E. (2009). Mechanotransduction at a 52. Yang, Y., Beqaj, S., Kemp, P., Ariel, I., and Schuger, L. (2000). Stretch-
distance: mechanically coupling the extracellular matrix with the nucleus. induced alternative splicing of serum response factor promotes bronchial
Nat. Rev. Mol. Cell Biol. 10, 75–82. myogenesis and is defective in lung hypoplasia. J. Clin. Invest. 106,
1321–1330.
35. Lombardi, M.L., Jaalouk, D.E., Shanahan, C.M., Burke, B., Roux, K.J.,
and Lammerding, J. (2011). The interaction between nesprins and sun 53. Planas-paz, L., Strilic, B., and Breier, G. (2012). Mechanoinduction of
proteins at the nuclear envelope is critical for force transmission between lymph vessel expansion. EMBO J. 31, 788–804.
the nucleus and cytoskeleton. J. Biol. Chem. 286, 26743–26753.
54. Olsen, S.M., Stover, J.D., and Nagatomi, J. (2011). Examining the role of
36. Isermann, P., and Lammerding, J. (2013). Nuclear mechanics and me- mechanosensitive ion channels in pressure mechanotransduction in rat
chanotransduction in health and disease. Curr. Biol. 23, R1113–R1121. bladder urothelial cells. Ann. Biomed. Eng. 39, 688–697.

37. Swift, J., Ivanovska, I.L., Buxboim, A., Harada, T., Dingal, P.C.D.P., 55. Miyamoto, T., Mochizuki, T., Nakagomi, H., Kira, S., Watanabe, M., Ta-
Pinter, J., Pajerowski, J.D., Spinler, K.R., Shin, J.-W., Tewari, M., et al. kayama, Y., Suzuki, Y., Koizumi, S., Takeda, M., and Tominaga, M.
(2013). Nuclear lamin-A scales with tissue stiffness and enhances ma- (2014). Functional role for Piezo1 in stretch-evoked Ca2+ influx and
trix-directed differentiation. Science 341, 1240104. ATP release in Urothelial cell cultures. J. Biol. Chem. 289, 16565–16575.

38. Cho, S., Irianto, J., and Discher, D.E. (2017). Mechanosensing by the nu- 56. Rauch, C., Brunet, A.-C., Deleule, J., and Farge, E. (2002). C2C12
cleus: From pathways to scaling relationships. J. Cell Biol. 216, 305–315. myoblast/osteoblast transdifferentiation steps enhanced by epigenetic
inhibition of BMP2 endocytosis. Am. J. Physiol. Cell Physiol. 283,
39. Le, H.Q., Ghatak, S., Yeung, C.C., Tellkamp, F., Günschmann, C., Diet- C235–C243.
erich, C., Yeroslaviz, A., Habermann, B., Pombo, A., Niessen, C.M.,
et al. (2016). Mechanical regulation of transcription controls Polycomb- 57. Landrein, B., Kiss, A., Sassi, M., Chauvet, A., Das, P., Cortizo, M., Laufs,
mediated gene silencing during lineage commitment. Nat. Cell Biol. 18, P., Takeda, S., Aida, M., Traas, J., et al. (2015). Mechanical stress con-
864–875. tributes to the expression of the STM homeobox gene in Arabidopsis
shoot meristems. Elife 4, 1–27.
40. Pouille, P.-A., Ahmadi, P., Brunet, A.-C., and Farge, E. (2009). Mechani-
cal signals trigger Myosin II redistribution and mesoderm invagination in 58. Green, J.B.A., and Sharpe, J. (2015). Positional information and reaction-
Drosophila embryos. Sci. Signal 2, 1–8. diffusion: two big ideas in developmental biology combine. Development
142, 1203–1211.
nager,
41. Mitrossilis, D., Röper, J.-C., Le Roy, D., Driquez, B., Michel, A., Me
C., Shaw, G., Le Denmat, S., Ranno, L., Dumas-Bouchiat, F., et al. (2017). 59. Wolpert, L. (2016). Positional Information and Pattern Formation, 1st ed
Mechanotransductive cascade of Myo-II-dependent mesoderm and (Elsevier Inc.).
endoderm invaginations in embryo gastrulation. Nat. Commun. 8, 13883.
60. Goehring, N.W., and Grill, S.W. (2013). Cell polarity: Mechanochemical
42. Mammoto, T., Mammoto, A., Torisawa, Y.S., Tat, T., Gibbs, A., Derda, R., patterning. Trends Cell Biol. 23, 72–80.
Mannix, R., de Bruijn, M., Yung, C.W., Huh, D., et al. (2011). Mechano-
chemical control of mesenchymal condensation and embryonic tooth 61. Howard, J., Grill, S.W., and Bois, J.S. (2011). Turing’s next steps: the
organ formation. Dev. Cell 21, 758–769. mechanochemical basis of morphogenesis. Nat. Rev. Mol. Cell Biol.
12, 392–398.
43. Maı̂tre, J.-L., Turlier, H., Illukkumbura, R., Eismann, B., Niwayama, R.,
 de
Ne lec, F., and Hiiragi, T. (2016). Asymmetric division of contractile do- 62. Mayer, M., Depken, M., Bois, J.S., Jülicher, F., and Grill, S.W. (2010).
mains couples cell positioning and fate specification. Nature 536, Anisotropies in cortical tension reveal the physical basis of polarizing
344–348. cortical flows. Nature 467, 617–621.

Current Biology 27, R1024–R1035, September 25, 2017 R1033


Current Biology

Review
63. Goehring, N., Philipp, K.T., Bois, J., D, C., E.M, N., Hyman, A., and Grill, 83. Durdu, S., Iskar, M., Revenu, C., Schieber, N., Kunze, A., Bork, P.,
S.W. (2011). Polarization of PAR proteins by advective triggering of a Schwab, Y., and Gilmour, D. (2014). Luminal signalling links cell commu-
pattern-forming system. Science 334, 1137–1141. nication to tissue architecture during organogenesis. Nature 515,
120–124.
64. Naganathan, S.R., Fürthauer, S., Nishikawa, M., Jülicher, F., and Grill,
S.W. (2014). Active torque generation by the actomyosin cell cortex 84. Harding, M.J., and Nechiporuk, A.V. (2012). Fgfr-Ras-MAPK signaling is
drives left-right symmetry breaking. Elife 3, 1–16. required for apical constriction via apical positioning of Rho-associated
kinase during mechanosensory organ formation. Development 139,
65. Singh, D., and Pohl, C. (2014). Coupling of rotational cortical flow, asym- 3467–3467.
metric midbody positioning, and spindle rotation mediates dorsoventral
axis formation in C. elegans. Dev. Cell 28, 253–267. 85. Shyer, A.E., Tallinen, T., Nerurkar, N.L., Wei, Z., Gil, E.S., Kaplan, D.L.,
Tabin, C.J., and Mahadevan, L. (2013). Villification: how the gut gets its
66. Nonaka, S., Shiratori, H., Saijoh, Y., and Hamada, H. (2002). Determina- villi. Science 342, 212–218.
tion of left–right patterning of the mouse embryo by artificial nodal flow.
Nature 418, 96–99. 86. Shyer, A.E., Huycke, T.R., Lee, C., Mahadevan, L., and Tabin, C.J. (2015).
Bending gradients: How the intestinal stem cell gets its home. Cell 161,
67. Tanaka, Y., Okada, Y., and Hirokawa, N. (2005). FGF-induced vesicular 569–580.
release of Sonic hedgehog and retinoic acid in leftward nodal flow is crit-
ical for left-right determination. Nature 435, 172–177. 87. Hiramatsu, R., Matsuoka, T., Kimura-Yoshida, C., Han, S.W., Mochida,
K., Adachi, T., Takayama, S., and Matsuo, I. (2013). External mechanical
68. Tabin, C.J., and Vogan, K.J. (2003). A two-cilia model for vertebrate left- cues trigger the establishment of the anterior-posterior axis in early
right axis specification. Genes Dev. 17, 1–6. mouse embryos. Dev. Cell 27, 131–144.
69. Yoshiba, S., Shiratori, H., Kuo, I.Y., Kawasumi, A., Shinohara, K., Non- 88. Bedzhov, I., and Zernicka-Goetz, M. (2014). Self-organizing properties of
aka, S., Asai, Y., Sasaki, G., Belo, J.A., Sasaki, H., et al. (2012). Cilia at mouse pluripotent cells initiate morphogenesis upon implantation. Cell
the node of mouse embryos sense fluid flow for left-right determination 156, 1032–1044.
via Pkd2. Science 338, 226–231.
89. Warmflash, A., Sorre, B., Etoc, F., Siggia, E.D., and Brivanlou, A.H.
70. Tarkowski, B.K. (1967). Development of blastomeres of mouse eggs iso- (2014). A method to recapitulate early embryonic spatial patterning in hu-
lated at the 4- and 8-cell stage. J. Embryol. Exp. Morph. 18, 155–180. man embryonic stem cells. Nat. Methods 11, 847–854.
71. Korotkevich, E., Niwayama, R., Courtois, A., Friese, S., Berger, N., Buch- 90. Etoc, F., Metzger, J., Ruzo, A., Kirst, C., Yoney, A., Ozair, M.Z., Brivanlou,
holz, F., and Hiiragi, T. (2017). The apical domain is required and suffi- A.H., and Siggia, E.D. (2016). A balance between secreted inhibitors and
cient for the first lineage segregation in the mouse embryo. Dev. Cell edge sensing controls gastruloid self-organization. Dev. Cell 39,
40, 235–247. 302–315.
72. Campinho, P., Behrndt, M., Ranft, J., Risler, T., and Minc, N. (2013). Ten-
91. Sugimura, K., Lenne, P.-F., and Graner, F. (2016). Measuring forces and
sion-oriented cell divisions limit anisotropic tissue tension in epithelial
stresses in situ in living tissues. Development 143, 186–196.
spreading during zebrafish epiboly. Nat. Cell Biol. 15, 1405–1414.
92. Campàs, O. (2016). A toolbox to explore the mechanics of living embry-
73. Mao, Q., and Lecuit, T. (2016). Mechanochemical Interplay Drives Polar-
onic tissues. Semin. Cell Dev. Biol. 55, 119–130.
ization in Cellular and Developmental Systems, 1st ed. (Elsevier Inc.).

74. Bosveld, F., Markova, O., Guirao, B., Martin, C., Wang, Z., Pierre, A., Ba- 93. Grashoff, C., Hoffman, B.D., Brenner, M.D., Zhou, R., Parsons, M., Yang,
lakireva, M., Gaugue, I., Ainslie, A., Christophorou, N., et al. (2016). M.T., McLean, M.A., Sligar, S.G., Chen, C.S., Ha, T., et al. (2010).
Epithelial tricellular junctions act as interphase cell shape sensors to Measuring mechanical tension across vinculin reveals regulation of focal
orient mitosis. Nature 530, 1–22. adhesion dynamics. Nature 466, 263–266.

75. Aigouy, B., Farhadifar, R., Staple, D.B., Sagner, A., Roper, J.-C., Julicher, 94. Borghi, N., Sorokina, M., Shcherbakova, O.G., Weis, W.I., Pruitt, B.L.,
F., and Eaton, S. (2010). Cell flow reorients the axis of planar polarity in Nelson, W.J., and Dunn, A.R. (2012). E-cadherin is under constitutive
the wing epithelium of Drosophila. Cell 142, 773–786. actomyosin-generated tension that is increased at cell–cell contacts
upon externally applied stretch. Proc. Natl. Acad. Sci. USA 109,
76. Sugimura, K., and Ishihara, S. (2013). The mechanical anisotropy in a tis- 913–916.
sue promotes ordering in hexagonal cell packing. Development 140,
4091–4101. 95. Yamashita, S., Tsuboi, T., Ishinabe, N., Kitaguchi, T., and Michiue, T.
(2016). Wide and high resolution tension measurement using FRET in em-
77. Chalmers, A.D., Strauss, B., and Papalopulu, N. (2003). Oriented cell di- bryo. Sci. Rep. 6, 1–8.
visions asymmetrically segregate aPKC and generate cell fate diversity in
the early Xenopus embryo. Development 130, 2657–2668. 96. Kelley, M., Yochem, J., Krieg, M., Calixto, A., Heiman, M.G., Kuzmanov,
A., Meli, V., Chalfie, M., Goodman, M.B., Shaham, S., et al. (2015).
78. Xiong, F., Ma, W., Hiscock, T.W., Mosaliganti, K.R., Tentner, A.R., FBN-1, a fibrillin-related protein, is required for resistance of the
Brakke, K.A., Rannou, N., Gelas, A., Souhait, L., Swinburne, I.A., et al. epidermis to mechanical deformation during C. elegans embryogenesis.
(2014). Interplay of cell shape and division orientation promotes robust Elife 4, 1–30.
morphogenesis of developing epithelia. Cell 159, 415–427.
 de
97. Maı̂tre, J.-L., Niwayama, R., Turlier, H., Ne lec, F., and Hiiragi, T. (2015).
79. Plusa, B., Piliszek, A., Frankenberg, S., Artus, J., and Hadjantonakis, A. Pulsatile cell-autonomous contractility drives compaction in the mouse
(2008). Distinct sequential cell behaviours direct primitive endoderm for- embryo. Nat. Cell Biol. 17, 849–855.
mation in the mouse blastocyst. Development 135, 3081–3091.
98. Chaigne, A., Campillo, C., Gov, N.S., Voituriez, R., Sykes, C., Verlhac,
80. Meilhac, S.M., Adams, R.J., Morris, S.A., Danckaert, A., Le Garrec, J.F., M.H., and Terret, M.E. (2015). A narrow window of cortical tension guides
and Zernicka-Goetz, M. (2009). Active cell movements coupled to posi- asymmetric spindle positioning in the mouse oocyte. Nat. Commun. 6,
tional induction are involved in lineage segregation in the mouse blasto- 6027.
cyst. Dev. Biol. 331, 210–221.
99. Campàs, O., Mammoto, T., Hasso, S., Sperling, R.A., O’Connell, D., Bis-
81. Moore, R., Cai, K.Q., Escudero, D.O., and Xu, X.X. (2009). Cell adhesive chof, A.G., Maas, R., Weitz, D.A., Mahadevan, L., and Ingber, D.E. (2014).
affinity does not dictate primitive endoderm segregation and positioning Quantifying cell-generated mechanical forces within living embryonic tis-
during murine embryoid body formation. Genesis 47, 579–589. sues. Nat. Methods 11, 183–189.

82. Xiong, F., Tentner, A.R., Huang, P., Gelas, A., Mosaliganti, K.R., Souhait, 100. Dolega, M.E., Delarue, M., Ingremeau, F., Prost, J., Delon, A., and Cap-
L., Rannou, N., Swinburne, I.A., Obholzer, N.D., Cowgill, P.D., et al. pello, G. (2017). Cell-like pressure sensors reveal increase of mechanical
(2013). Specified neural progenitors sort to form sharp domains after stress towards the core of multicellular spheroids under compression.
noisy Shh signaling. Cell 153, 550–561. Nat. Commun. 8, 1–9.

R1034 Current Biology 27, R1024–R1035, September 25, 2017


Current Biology

Review
101. Brodland, G.W., Veldhuis, J.H., Kim, S., Perrone, M., Mashburn, D., and 116. Xenopoulos, P., Kang, M., Puliafito, A., Di Talia, S., and Hadjantonakis,
Hutson, M.S. (2014). CellFIT: A cellular force-inference toolkit using curvi- A.K. (2015). Heterogeneities in nanog expression drive stable commit-
linear cell boundaries. PLoS One 9, e99116. ment to pluripotency in the mouse blastocyst. Cell Rep. 10, 1508–1520.

102. Ishihara, S., Sugimura, K., Cox, S.J., Bonnet, I., Bellaı̈che, Y., and Graner, 117. Toettcher, J.E., Voigt, C. a, Weiner, O.D., and Lim, W.A. (2011). The
F. (2013). Comparative study of non-invasive force and stress inference promise of optogenetics in cell biology: interrogating molecular circuits
methods in tissue. Eur. Phys. J. E. 36, 45. in space and time. Nat. Methods 8, 35–38.

103. Brodland, G.W., Conte, V., Cranston, P.G., Veldhuis, J., Narasimhan, S., 118. Tischer, D., and Weiner, O.D. (2014). Illuminating cell signalling with op-
Hutson, M.S., Jacinto, A., Ulrich, F., Baum, B., and Miodownik, M. (2010). togenetic tools. Nat. Rev. Mol. Cell Biol. 15, 551–558.
Video force microscopy reveals the mechanics of ventral furrow invagi-
nation in Drosophila. Proc. Natl. Acad. Sci. USA 107, 22111–22116. 119. Guglielmi, G., Falk, H.J., and De Renzis, S. (2016). Optogenetic control of
protein function: from intracellular processes to tissue morphogenesis.
104. Guirao, B., Rigaud, S.U., Bosveld, F., Bailles, A., Lopez-Gay, J., Ishihara, Trends Cell Biol. 26, 864–874.
S., Sugimura, K., Graner, F., and Bellaiche, Y. (2015). Unified quantitative
characterization of epithelial tissue development. Elife 4, 1–52. 120. Guglielmi, G., Barry, J.D., Huber, W., and De Renzis, S. (2015). An opto-
genetic method to modulate cell contractility during tissue morphogen-
105. Veldhuis, J.H., Ehsandar, A., Maitre, J.-L., Hiiragi, T., Cox, S., and Brod- esis. Dev. Cell 35, 646–660.
land, G.W. (2017). Inferring cellular forces from image stacks. Phil. Trans.
R. Soc. B. 372, 20160261. 121. Wagner, E., and Glotzer, M. (2016). Local RhoA activation induces cyto-
kinetic furrows independent of spindle position and cell cycle stage.
106. Krens, S.F.G., Veldhuis, J.H., Barone, V., Capek, D., Maitre, J.-L., Brod- J. Cell Biol. 213, 641–649.
land, G.W., and Heisenberg, C.P. (2017). Interstitial fluid osmolarity mod-
ulates the action of differential tissue surface tension in progenitor cell 122. Rossi, A., Kontarakis, Z., Gerri, C., Nolte, H., Hölper, S., Krüger, M., and
segregation during gastrulation. Development. 1–9. Stainier, D.Y.R. (2015). Genetic compensation induced by deleterious
mutations but not gene knockdowns. Nature 13, 230–233.
107. Antonacci, G., and Braakman, S. (2016). Biomechanics of subcellular
structures by non-invasive Brillouin microscopy. Sci. Rep. 6, 37217. 123. Harmansa, S., Hamaratoglu, F., Affolter, M., and Caussinus, E. (2015).
Dpp spreading is required for medial but not for lateral wing disc growth.
108. Scarcelli, G., and Yun, S.H. (2007). Confocal Brillouin microscopy for
Nature 527, 317–322.
three-dimensional mechanical imaging. Nat. Photonics. 2, 39–43.

109. Scarcelli, G., Polacheck, W.J., Nia, H.T., Patel, K., Grodzinsky, A.J., 124. Fernandez-Sanchez, M.E., Barbier, S., Whitehead, J., Bealle, G., Michel,
Kamm, R.D., and Yun, S.H. (2015). Noncontact three-dimensional map- A., Latorre-ossa, H., Rey, C., Fouassier, L., Claperon, A., Brulle, L., et al.
ping of intracellular hydromechanical properties by Brillouin microscopy. (2015). Mechanical induction of the tumorigenic beta-catenin pathway by
Nat. Methods 12, 1132–1134. tumour growth pressure. Nature 523, 92–95.

110. Reiß, S., Burau, G., Stachs, O., Guthoff, R., and Stolz, H. (2011). Spatially 125. Serwane, F., Mongera, A., Rowghanian, P., Kealhofer, D.A., Lucio, A.A.,
resolved Brillouin spectroscopy to determine the rheological properties Hockenbery, Z.M., and Campàs, O. (2016). In vivo quantification of
of the eye lens. Biomed. Opt. Express. 2, 2144–2159. spatially varying mechanical properties in developing tissues. Nat.
Methods 14, 181–186.
111. Elsayad, K., Werner, S., Gallemı́, M., Kong, J., Guajardo, E.R.S., Zhang,
L., Jaillais, Y., Greb, T., and Belkhadir, Y. (2016). Mapping the subcellular 126. Dietrich, J.E., and Hiiragi, T. (2007). Stochastic patterning in the mouse
mechanical properties of live cells in tissues with fluorescence emission – pre-implantation embryo. Development 134, 4219–4231.
Brillouin imaging. Sci. Signal 9, 1–13.
127. Burute, M., Prioux, M., Blin, G., Young, J., Tseng, Q., Bessy, T., Lowell,
112. Strnad, P., Gunther, S., Reichmann, J., Krzic, U., Balazs, B., de Medeiros, S., Young, J., Filhol, O., and Thery, M. (2017). Polarity reversal by centro-
G., Norlin, N., Hiiragi, T., Hufnagel, L., and Ellenberg, J. (2015). Inverted some repositioning primes cell scattering during epithelial-to-mesen-
light-sheet microscope for imaging mouse pre-implantation develop- chymal transition. Dev. Cell 40, 168–184.
ment. Nat. Methods 13, 139–142.
128. Ranga, A., Girgin, M., Meinhardt, A., Eberle, D., Caiazzo, M., Tanaka,
113. Amat, F., Lemon, W., Mossing, D.P., McDole, K., Wan, Y., Branson, K., E.M., and Lutolf, M.P. (2016). Neural tube morphogenesis in synthetic
Myers, E.W., and Keller, P.J. (2014). Fast, accurate reconstruction of 3D microenvironments. Proc. Natl. Acad. Sci. USA 113, E6831–E6839.
cell lineages from large-scale fluorescence microscopy data. Nat.
Methods 11, 951–958. 129. Choi, J.S., and Harley, B. (2017). Marrow-inspired matrix cues rapidly
affect early fate decisions of hematopoietic stem and progenitor cells.
114. Olivier, N., Luengo-oroz, M.A., Duloquin, L., Faure, E., Savy, T., Veilleux, Sci. Adv. 3, e1600455.
barre, D., Bourgine, P., Santos, A., et al. (2010). Cell
I., Solinas, X., De
lineage reconstruction of early zebrafish embryos using label-free 130. Demers, C.J., Soundararajan, P., Chennampally, P., Cox, G.A., Briscoe,
nonlinear microscopy. Science 329, 967–971. J., Collins, S.D., and Smith, R.L. (2016). Development-on-chip: in vitro
neural tube patterning with a microfluidic device. Development 143,
115. Dietrich, J.-E., Panavaite, L., Gunther, S., Wennekamp, S., Groner, A.C., 1884–1892.
Pigge, A., Salvenmoser, S., Trono, D., Hufnagel, L., and Hiiragi, T. (2015).
Venus trap in the mouse embryo reveals distinct molecular dynamics un- ras, N., and Doursat, R. (2017). A cell-
131. Delile, J., Herrmann, M., Peyrie
derlying specification of first embryonic lineages. EMBO Rep. 16, based computational model of early embryogenesis coupling mechani-
e201540162. cal behaviour and gene regulation. Nat. Commun. 8, 1–10.

Current Biology 27, R1024–R1035, September 25, 2017 R1035

You might also like