You are on page 1of 14

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/325835600

Instructions for Assembling the Early Mammalian Embryo

Article  in  Developmental Cell · June 2018


DOI: 10.1016/j.devcel.2018.05.013

CITATIONS READS

33 1,676

4 authors, including:

Melanie White Jennifer Zenker


The University of Queensland Monash University (Australia)
31 PUBLICATIONS   1,400 CITATIONS    25 PUBLICATIONS   244 CITATIONS   

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Melanie White on 20 June 2018.

The user has requested enhancement of the downloaded file.


Developmental Cell

Perspective

Instructions for Assembling


the Early Mammalian Embryo
Melanie D. White,1 Jennifer Zenker,1 Stephanie Bissiere,1 and Nicolas Plachta1,2,*
1Instituteof Molecular and Cell Biology, A*STAR, 61 Biopolis Drive, Singapore 138673
2Department of Biochemistry, NUS, Singapore
*Correspondence: plachtan@imcb.a-star.edu.sg
https://doi.org/10.1016/j.devcel.2018.05.013

The preimplantation mouse embryo is a simple self-contained system, making it an excellent model to
discover how mammalian cells function in real time and in vivo. Work over the last decade has revealed
some key morphogenetic mechanisms that drive early development, yielding rudimentary instructions for
the generation of a mammalian embryo. Here, we review the instructions revealed thus far, and then discuss
remaining challenges to discover upstream factors controlling cell fate determination, test the role of mech-
anisms based on biological noise, and take advantage of recent technological developments to advance the
spatial and temporal resolution of our current understanding.

The Preimplantation Embryo Is a Simple, Self-Contained possible to dissect mechanisms of mammalian meiosis. When
System isolated after fertilization, the one-cell embryo undergoes
A central question in developmental biology is how complex cleavage divisions and develops as a self-contained group of
forms and functions arise from the simple starting point of a sin- blastomeres enclosed by the zona pellucida (ZP) and indepen-
gle cell. During morphogenesis, spatiotemporal order and dent of interactions with extracellular matrices or tissues. This
patterning spontaneously emerge through local interactions at developing embryo can be imaged in such ex utero conditions
cellular and subcellular scales (Davies, 2017; Sasai, 2013). This until the blastocyst stage when it is ready for implantation into
process of self-organization occurs in the absence of external the uterus (Behringer, 2014; Brinster, 1963). The accessibility
control and results in the emergence of higher order properties of the cultured mouse embryo to imaging and genetic manipula-
that cannot be accounted for by the properties of the individual tions while retaining the advantage of a physiological in vivo
components. In this way, large numbers of cells can organize system has placed this system at the center of research in
into complex tissues by obeying simple rules of local interaction mammalian genetics, physiology, and development.
without reference to a global pattern. Such self-organization has Although there is some variability in timing, the major events of
the advantage that it is relatively robust to error, highly scalable, preimplantation development are well conserved across
and retains the capacity for self-repair. mammalian species (Reijo Pera and Prezzoto, 2016). From
The preimplantation mouse embryo is an excellent system for mouse to human, most mammals follow a general sequence of
studying how simple cells organize dynamically into increasingly events consisting of fertilization followed by early development
complex structures. The mouse embryo executes a develop- without embryonic transcription; cleavage divisions in the
mental program from zygote to hatched blastocyst without the absence of growth; compaction to form a morula; and, finally,
requirement for external input and both cellular and subcellular generation of a blastocyst. Early studies manipulating cells of
elements of the embryo interact dynamically to generate emer- the mouse embryo demonstrated that this developmental pro-
gent patterned structures. These elements consist of physical gram is not predetermined, as in the case of lower organisms
properties such as cell shape, adhesion and polarity, biochem- such as Drosophila, Caenorhabditis elegans, and Xenopus, but
ical signaling and feedback loops, genetically encoded informa- consists of varying division patterns and flexibility of cell fate
tion, and stochastically generated variation (Nissen et al., 2017; that can compensate for cell loss (Hoppe and Whitten, 1972;
White et al., 2016c; Wennekamp et al., 2013). Adaptive self-orga- Tarkowski and Wroblewska, 1967; Tarkowski, 1959) and even
nization in response to the changing interactions between these artificial rearrangements (Ziomek and Johnson, 1982; Ziomek
elements enables robust regulative development of the embryo et al., 1982). These initial findings hinted at what is now thought
without the external imposition of order. In this way, the embryo to be a complex interaction of regulatory mechanisms working in
progresses from a single cell into a complex multi-layered struc- concert to gradually drive cells toward separate fates. Intrinsic to
ture consisting of multiple cell lineages prior to implantation. this regulative process are not only subcellular processes but
In practical terms, the preimplantation mouse embryo repre- also intercellular interactions that lead to a progressive refine-
sents an excellent midway point between complex tissues and ment of cellular identity. This plasticity, and the consequent abil-
organs, and simplistic cell culture systems (Figure 1). Progres- ity to compensate in the face of internal or external disturbances,
sion through the entire preimplantation stage can proceed may be especially important in the context of mammalian devel-
in vitro, in the absence of maternal tissues, without any discern- opment where significant energy and resources are invested into
ible consequences on subsequent development (Whitten and each embryo.
Biggers, 1968). Even before fertilization, the mouse egg can be Recent advances in live-cell imaging, computational cell
isolated and studied directly under the microscope, making it tracking, and optogenetics are beginning to yield the details

Developmental Cell 45, June 18, 2018 ª 2018 Elsevier Inc. 667
Developmental Cell

Perspective
Figure 1. Comparison of Model Systems for
Studying Mammalian Development
(A) A mouse embryonic stem cell culture system.
This system has the advantage of being simple
and highly accessible for manipulations but is not
in a physiological context. Cells are labeled for the
Oct4 transcription factor (green) and a nuclear
marker (red).
(B) A 16-cell stage preimplantation mouse embryo
labeled for actin (green), microtubules (red), and
DNA (blue). The preimplantation embryo com-
bines the advantages of simplicity and accessi-
bility with the physiological relevance of an in vivo
system.
(C) A post-implantation gastrulation stage embryo
labeled for the cell membrane (orange) and DNA
(blue). By this stage the embryo has complex
architecture and is relatively inaccessible for im-
aging and manipulation. Graphs in right panels are
schematic representations of the relative advan-
tages and disadvantages of each model system.
Scale bars represent 10 mm in (A) and (B) and
30 mm in (C).

with maternal mRNAs and proteins that


will be required to sustain development
through the first stages of life (Li and
Albertini, 2013). In response to a surge
of luteinizing hormone released from the
pituitary gland, the oocyte resumes
meiosis, breaks down its nucleus, and
assembles a microtubule spindle around
its chromosomes (Bury et al., 2017). In
a critical first step of symmetry breaking,
the cortex of the oocyte softens and
thickens, promoting migration of the
spindle toward the nearest cortex
(Chaigne et al., 2013). This off-center
positioning of the spindle ensures the
asymmetry of the first meiotic division.
Unlike spindle positioning in somatic
cells, which typically depends on astral
underlying these regulatory mechanisms and reveal how cells microtubules, oocytes lack centrosomes, and therefore migra-
interact with their neighbors to resolve their fates within the intact tion of the meiotic spindle is driven by actin polymerization (Yi
developing embryo. Although some of the elements driving the et al., 2013; Dumont et al., 2007). After spindle migration, the
self-organization of the preimplantation embryo have been iden- close proximity of the chromosomes to the cortex induces the
tified, extensive work is still required to understand how multiple formation of a cortical actomyosin cap (Deng et al., 2007). An
mechanisms operating at the level of the cell nucleus, cytoskel- actin flow is established that drives cytoplasmic streaming to
eton, and cell-cell interactions are integrated to fully assemble maintain spindle positioning and the first polarization events
the embryo. (Yi et al., 2011). The oocyte then undergoes a division that is
extremely asymmetric in size and discards half of its chromo-
Preimplantation Development from Oocyte to somes into a small cell, called the polar body, which will later
Blastocyst degenerate. Recent high-resolution imaging studies have re-
Cytoskeletal Changes Mediate the Transition from vealed that actin filaments also infiltrate the meiotic spindle
Oocyte to Fertilized Embryo and are crucial for correct alignment and segregation of the
The transformation from an egg to an embryo requires precise chromosomes (Mogessie and Schuh, 2017). Immediately after
coordination of a series of intricate processes that unite the this division, the meiosis II spindle assembles around the chro-
haploid genomes of each parent into a single diploid genome mosomes just under the cortex and the oocyte remains arrested
of a new living organism. However, preparation for these events at this stage, awaiting fertilization. Completion of the maturation
begins long before fertilization with the maturation of the egg process can occur spontaneously in isolated oocytes in vitro
precursor cell, or oocyte (Figure 2). Within the ovary, oocytes (Edwards, 1965), indicating that it proceeds by self-organiza-
are stored in a quiescent state, arrested in prophase of the first tion. Fusion of a sperm with the mature oocyte triggers the
meiotic division. Surrounding support cells provide the oocyte completion of meiosis II with another asymmetric division that

668 Developmental Cell 45, June 18, 2018


Developmental Cell

Perspective

Figure 2. Instructions for Assembling a Mouse Embryo from Oocyte to Compacted Morula
An overview of the main mechanisms currently known to direct preimplantation development from an oocyte to a compacted morula. Scale bars represent 10 mm.

generates a second polar body, translation of maternal mRNAs, Early Heterogeneities among Cells of the Embryo
and the progression from meiosis to mitosis (Clift and Schuh, Upon fertilization, the maternal chromosomes and the sperm
2013). The first mitotic division in the newly formed zygote is chromatin decondense and form separate pronuclei. Within
also dependent on actin networks (Chaigne et al., 2016). each pronucleus there is a distinct program of epigenetic remod-
However unlike in the oocyte, the zygotic spindle is positioned eling of the parental DNA to remove gamete-specific modifica-
centrally and the subsequent division is symmetric, thereby pro- tions and transform the zygote to a totipotent state. Changes
ducing two daughter cells, or blastomeres, of similar sizes. in DNA methylation, histone modifications, and activation of
Thus, the transition from meiosis to mitosis is complete retrotransposons all contribute to nuclear remodeling (Habibi
(Figure 2). and Stunnenberg, 2017; Jachowicz et al., 2017; Burton and

Developmental Cell 45, June 18, 2018 669


Developmental Cell

Perspective

Torres-Padilla, 2014). It has been proposed that differences sion in a process known as a partitioning error (Shi et al., 2015;
between cells of the embryo do not arise until the 16-cell stage Huh and Paulsson, 2011). In the absence of transcription, parti-
when the cells first adopt inner and outer positions (Alarcon tioning errors likely account for the initial introduction of
and Marikawa, 2005; Motosugi et al., 2005; Hiiragi and Solter, variability between cells. One such variably distributed factor
2004). Indeed, initial studies revealed no differences between may relate to mitochondrial rRNAs, which have recently been
early blastomeres in gene expression; however, these studies shown to become progressively heterogeneous by the end of
were restricted to the analysis of selected candidate genes the two-cell stage (Zheng et al., 2016). Whether and how this
(Wicklow et al., 2014; Guo et al., 2010; Ralston and Rossant, might influence or translate into asymmetries arising at the
2008; Dietrich and Hiiragi, 2007). Furthermore, although blasto- four-cell stage remains to be determined.
meres can retain the capacity to differentiate into either inner Minor activation of zygotic transcription is first detected at the
cell mass (ICM) or trophectoderm until at least the 16-cell stage late one-cell stage with higher activity in the male pronucleus
(Suwinska et al., 2008; Ziomek et al., 1982; Rossant and Vijh, than the female pronucleus (Zhou and Dean, 2015). However,
1980; Tarkowski and Wroblewska, 1967), mounting evidence the majority of zygotic genome activation first occurs at the
suggests that heterogeneities that bias cells toward one fate or two-cell stage and contributes to the preparation of basic cellular
the other may arise earlier. Lineage tracing studies revealed a machinery involved in functions such as protein translation, cell
cell fate bias in some blastomeres at the four-cell stage suggest- metabolism, and RNA processing (Gao et al., 2017). From this
ing that early blastomeres may not contribute equally to all line- point on, the newly synthesized embryonic products gradually
ages (Tabansky et al., 2013; Piotrowska-Nitsche et al., 2005; replace the maternally supplied factors and become the main
Fujimori et al., 2003), although this effect may not be conserved regulators of development.
among all mouse strains (Alarcon and Marikawa, 2005). Analysis Polarization and Compaction
of epigenetic modifications at the four-cell stage has revealed an The first major morphological changes in the embryo begin
asymmetry between blastomeres in the levels of dimethylation of with the onset of polarization and compaction at the eight-
histone H3 at arginines 17 and 26 (H3R17 and H3R26) and in the cell stage. Until the early eight-cell stage, each cell within the
levels of the chromatin modifiers Carm1 (Torres-Padilla et al., embryo is morphologically similar. Within 1–3 hr of reaching
2007) and PRDM14 (Burton et al., 2013). The levels of H3 methyl- the eight-cell stage, blastomeres begin to display the first out-
ation are proposed to influence cell fate determination, as those ward signs of radial polarity with the establishment of distinct
four-cell blastomeres displaying higher levels of methylation at apical and basolateral domains (Figure 2) (Ziomek and John-
H3R17 and H3R26 tend to contribute more progeny to the ICM son, 1980). During this time, the actin cytoskeleton is reorgan-
of the embryo (Figure 2) (Burton et al., 2013; Torres-Padilla ized to form an apical pole rich in microvilli and there is a polar-
et al., 2007). Overexpression of CARM1 leads to increased ized redistribution of cytoplasmic organelles and cytoskeletal
H3R26 methylation and was associated with elevated expres- elements (Ducibella et al., 1977). The contact-free surface of
sion of the ICM fate markers Nanog and Sox2 and positioning each blastomere forms the apical domain, characterized by
of progeny within the ICM (Torres-Padilla et al., 2007). clustering of the microvilli and the progressive localization of
Recently, the development of photoactivatable fluorescence the actin-binding protein Ezrin (Louvet et al., 1996), and the
correlation spectroscopy (paFCS) techniques has made it Par3-Par6-aPKC complex (Vinot et al., 2005). The adhesion
possible to probe transcription factor dynamics in vivo (Zhao molecule E-cadherin (Vestweber et al., 1987) accompanied
et al., 2017; Kaur et al., 2013; Plachta et al., 2011). Application by Par-1 (Vinot et al., 2005), Jam-1 (Thomas et al., 2004),
of paFCS within the embryo revealed that the increases in and Na+/K+ ATPase (Watson and Kidder, 1988) become
Carm1-mediated histone H3 methylation promote DNA binding restricted to the basolateral cell-cell contacts and the first
of Sox2, a transcription factor important for maintaining pluripo- epithelial-like tissue architecture of the embryo is established
tency, likely by increasing the accessibility of Sox2 binding sites (Johnson, 2009). Although E-cadherin-mediated intercellular
(White et al., 2016a). This increased Sox2-DNA binding is adhesion is not required for the initiation of polarization in iso-
thought to result in upregulation of Sox2-dependent genes lated blastomeres (Ziomek and Johnson, 1980), it is necessary
linked to pluripotency (Goolam et al., 2016; White et al., 2016a) for correct segregation of the apical and basolateral domains
biasing cells toward an embryonic fate. How the variability in his- and therefore controls the timing and axis of polarization (White
tone H3 methylation arises de novo at the four-cell stage remains et al., 2016b; Stephenson et al., 2010). The process of polari-
unclear, but it may originate from interplay between cell cleavage zation does not require transcription or translation at the
orientations (Zernicka-Goetz et al., 2009), differences in gene eight-cell stage (Levy et al., 1986), suggesting it is instead
expression (Shi et al., 2015; Biase et al., 2014; Piras et al., regulated through post-translational mechanisms (Johnson,
2014) resulting from noise-excitable mechanisms (Eldar and 2009). The exact molecular trigger for polarization of the mouse
Elowitz, 2010), or the possible asymmetric distribution of uniden- embryo remains to be revealed, but recent work has elucidated
tified factors. some of the signaling pathways involved (Zhu et al., 2017).
The transition from oocyte to embryo occurs prior to zygotic Accumulation of actomyosin at the apical cortex is initiated
genome activation and proceeds without RNA transcription. In by phospholipase C-mediated PIP2 hydrolysis, which activates
fact, early development relies almost entirely on maternally pro- protein kinase C (PKC). This in turn activates RhoA, leading to
vided factors deposited in the egg, many of which execute func- polarization of the actin network. The Par3-Par6-aPKC com-
tions involved in protein transport, protein localization, and cell plex is then recruited to the apical surface in a manner that
cycle (Gao et al., 2017; Zhou and Dean, 2015). These factors likely requires additional unknown factors. Polarization of the
may be unevenly distributed into each daughter cell during divi- Par complex excludes actomyosin, forming a mature apical

670 Developmental Cell 45, June 18, 2018


Developmental Cell

Perspective

domain (Zhu et al., 2017), which is proposed to exhibit reduced direct lineage divergence. Consequently, regulation of cell repo-
cortical contractility (Maitre et al., 2016). sitioning during this stage is critical for the correct development
Concomitant with the establishment of basolateral domains in of the embryo.
the embryo is the process of compaction. This results in maximi- Traditionally, it was assumed that ‘‘asymmetric’’ cell divisions
zation of cell-cell contact area and flattening of the outer surface allocated inner cells by cleaving the parental cell perpendicular
of the blastomeres to form a tightly packed mass of cells with to the embryo’s surface and pushing one daughter cell into the
indistinct cell boundaries (White et al., 2016b; Ducibella and ICM as a direct result of the scission (Cockburn and Rossant,
Anderson, 1975). Embryo compaction is critical for blastocyst 2010; Johnson and Ziomek, 1981). However, division-indepen-
formation and ongoing development (Ducibella and Anderson, dent cell internalization events were also suggested to contribute
1975). Due to the requirement for E-cadherin in compaction, it varying numbers of inner cells (Anani et al., 2014; Watanabe
was widely believed that increases in intercellular adhesion et al., 2014; Yamanaka et al., 2010; Plusa et al., 2005).
were responsible for this morphological change. However, it Combining live imaging with computational membrane segmen-
has recently been demonstrated that there are no differences tation enabled quantitative tracking of changes in cell shape and
in the expression or mobility of E-cadherin during compaction position, and revealed that asymmetric divisions are relatively
(Samarage et al., 2015), and it is unclear whether E-cadherin rare (Samarage et al., 2015). Instead, most cells divide symmet-
trans-ligation could even generate sufficient forces to achieve rically, or at an oblique angle, on the surface of the embryo and
this cellular deformation (Maitre et al., 2012). Live imaging of then one of the daughter cells may internalize through a process
E-cadherin during compaction revealed long membrane protru- of apical constriction. Laser ablations of the cortical actomyosin
sions, classified as filopodia, that extend from the apical border network revealed the development of anisotropies in the magni-
of cells in the eight-cell embryo and adhere via E-cadherin to the tude and directionality of tensile forces acting at cell junctions
top of their neighboring cells (Fierro-Gonzalez et al., 2013). In and the apical cell cortex. Cells that will internalize to form the
addition to E-cadherin, the filopodia contain proteins that link it ICM display increased levels of myosin II relative to their neigh-
to the cytoskeleton and the unconventional myosin motor bors (Anani et al., 2014), resulting in higher tensile forces that
protein, myosin-10. Molecular manipulation of any of these com- drive constriction of their apical membrane (Samarage et al.,
ponents disrupts compaction and acute femtosecond laser 2015). Although many of these internalization events occur
ablation of filopodia results in rapid rounding and separation following cell rearrangements due to divisions in the embryo,
of neighboring cells. These findings indicate that filopodia the exact molecular mechanisms initiating the asymmetries in
contribute to the tensile force required to change cell shape dur- cortical tension between cells is still an open question. One
ing compaction (Figure 2). Additional support for a role of tensile classic model suggests that the asymmetric inheritance of the
force in cell shape change during compaction was provided us- apical domain during cell division determines cell fate. Recent
ing micropipette aspiration and measurement of blastomere work suggests that the apical domain recruits a mitotic spindle
deformation (Maitre et al., 2015). This demonstrated a progres- pole to ensure its asymmetric inheritance (Korotkevich et al.,
sive increase in cortical tension throughout compaction driven 2017). Presence of the apical domain is thought to minimize
by pulsed actomyosin contractions, and indicated that junctional cortical tension to prevent cell internalization and trigger tro-
E-cadherin acts to redistribute actomyosin contractility away phectoderm fate specification (Maitre et al., 2016). However,
from cell-cell contacts. How E-cadherin-dependent filopodia the inheritance of the apical domain during cell division has not
and pulsatile actomyosin contractility are each initiated and been demonstrated at sufficiently high temporal resolution by
regulated remains to be determined. However, the accessibility live imaging in whole embryos, and some have proposed
of the preimplantation embryo makes this an ideal system to that apical polarity may be reestablished following cytokinesis
study these basic biological processes, which are commonly (Watanabe et al., 2014; Zenker et al., 2018).
employed throughout morphogenesis in many systems (Lecuit In addition to remodeling of actin networks, dynamic changes
et al., 2011). in microtubule organization are essential for the formation of the
Formation of the Pluripotent ICM ICM (Zenker et al., 2017). Whereas in most animal cells the
Compaction and embryo polarization are followed by the first centrosome is the main microtubule organizing center (MTOC)
spatial segregation of cells into two separate populations that (Conduit et al., 2015), the cells of the mouse embryo do not
will establish the first distinct cell lineages (Mihajlovic and Bruce, form centrosomes until the blastocyst stage (Clift and Schuh,
2017) (Figure 2). During the eight- to 16-cell stage, the first cells 2015). This suggested a random microtubule organization during
will be positioned in the interior of the embryo. These inner cells preimplantation development. However, recent work revealed a
are completely surrounded by basolateral membrane and cell- different form of non-centrosomal microtubule organization
cell contacts, and are therefore apolar. They will subsequently within the early embryo (Zenker et al., 2017). Following cell divi-
differentiate to form the primitive endoderm, which gives rise sion, the cytokinetic bridge is not abscised but instead it is
to extra-embryonic membranes, and the pluripotent epiblast, retained throughout interphase and connects each pair of sister
from which the entire embryo proper is derived. The cells on cells within the embryo. This microtubule bridge accumulates
the outside of the embryo retain an apical surface, which is Nezha/CAMSAP3, a protein that stabilizes microtubule minus
devoid of cell-cell contacts and exposed to the exterior. These ends at non-centrosomal sites (Akhmanova and Hoogenraad,
cells are therefore polarized and most of them will differentiate 2015) and enables the bridge to direct the growth of microtu-
to form the trophectoderm, which gives rise to the placenta. bules into the cell.
This spatial segregation of cells creates the first discrete niche One function of this non-centrosomal MTOC is to direct the
inside the embryo and generates structural asymmetries that transport of endosomes carrying E-cadherin to the cell

Developmental Cell 45, June 18, 2018 671


Developmental Cell

Perspective

membrane. Moreover, as cells internalize to form the ICM, their permeability barrier that seals the embryo from the exterior and
bridge displays increased CAMSAP3 levels, higher microtubule resists the increasing hydrostatic pressure. This barrier is medi-
density, and more E-cadherin transport, which is proposed to ated by the establishment of tight junctions between the outer
facilitate the expansion of the basolateral cell membrane of the cells (Wang et al., 2008), forming the trophectoderm: the first
inner cells (Figure 2). Beyond E-cadherin, the microtubules of epithelium of the embryo. Recent work demonstrated that assem-
the interphase bridge may also transport other proteins essential bly of the permeability barrier begins with the formation of cortical
for cell polarity or cell fate determination. Moreover, the mainte- actin rings on the apical surface of outer cells at the 16-cell stage
nance of a shared cytoskeleton between sister cells likely en- (Zenker et al., 2018). Unlike stereotypical actin rings, which are
ables mechanical coupling between cells. contractile, these actin rings expand to the cell-cell junctions
The establishment of populations of inner and outer cells with where they recruit and stabilize components of adherens and tight
different patterns of cell-cell contacts creates structural asym- junctions. Coupling of the actin rings to the junctions triggers a
metries that are critical for the expression of fate-specifying tension-dependent zippering mechanism that extends along the
markers. Each cell relies on the balance between cell-cell adhe- junctions and seals the embryo, enabling blastocyst expansion.
sion and polarity to regulate the activity of the Hippo signaling As the blastocyst cavity expands, it breaks the radial symme-
pathway and downstream lineage-specific genes. The transcrip- try of the embryo and causes clustering of the inner cells at the
tion factors Gata2, Gata3, and Cdx2 promote and maintain tro- embryonic pole. This allows for a separation of cell layers that
phectoderm differentiation (Home et al., 2017; Ralston et al., may be critical for the further differentiation of cells of the ICM.
2010). Tead4 and its coactivator, Yap1, maintain the expression Until the early blastocyst stage, the ICM is a homogeneous pop-
of Gata3 and Cdx2 in outer cells in a Hippo-dependent manner ulation of cells co-expressing lineage markers including Nanog,
(Nishioka et al., 2009). Yap1 shuttles between the nucleus and Gata6, and Oct-4 (Ohnishi et al., 2014). However, as the blasto-
the cytoplasm and its phosphorylation state determines its sub- cyst expands, a reciprocal expression of Nanog and Gata6 leads
cellular localization (Sasaki, 2017). to the formation of two distinct populations, the primitive endo-
In the outer cells, Yap1 translocates into the nucleus to activate derm and the epiblast, in a heterogeneous ‘‘salt-and-pepper’’
expression of trophectoderm-specific genes, whereas, in the distribution (Figure 3) (Chazaud et al., 2006). The early variations
inner cells, Yap1 is phosphorylated and therefore excluded from in gene expression are known to be propagated by FGF4/FGFR2
the nucleus (Nishioka et al., 2009). Both cell-cell adhesion and signaling (Krawchuk et al., 2013; Guo et al., 2010); however, their
polarity act to regulate the junctional localization of Amot, a key temporal regulation is not yet fully understood (Bessonnard
member of the Hippo pathway required for the phosphorylation et al., 2017). Epiblast precursors expressing Nanog secrete
of Yap1 (Hirate et al., 2013). In apolar inner cells, Amot indirectly FGF4 to drive Gata6 expression and primitive endoderm fate
associates with E-cadherin and Nf2 at the junctions specification in their neighbor cells (Nissen et al., 2017; Lanner
encompassing the cell, where it is activated through phosphoryla- and Rossant, 2010). Recent evidence suggests that the hetero-
tion by Lats kinase (Cockburn et al., 2013; Hirate et al., 2013; Nish- geneities in FGF4/FGFR2 signaling may be initiated by variations
ioka et al., 2009). Phosphorylated Amot then acts together with in Sox2/Oct4-dependent Fgf4 transcription that arise from fluc-
Lats to enhance the phosphorylation and nuclear exclusion of tuating Sox2 expression levels (Mistri et al., 2018). In addition
Yap1, allowing the expression of pluripotency-associated genes. to FGFR2, FGFR1 has recently been shown to be required for
By contrast, in the polarized outer cells, the presence of an the differential transduction of the FGF4 signal within the ICM
apical domain sequesters Amot from the junctions and localizes (Kang et al., 2017; Molotkov et al., 2017).
it apically where it binds to cortical F-actin and remains unphos- Following cell fate specification, the precise spatial arrange-
phorylated (Hirate et al., 2013). Recent work demonstrated that ment of the two-cell populations occurs through cell migration,
Rho kinase also acts to stabilize Amot binding to F-actin and apoptosis, polarization, and adhesion. By the time of blastocyst
inhibit the interaction between Amot and Nf2 (Shi et al., 2017). implantation, the primitive endoderm is a morphologically
As Amot is absent from the junctional complexes, the Hippo distinct epithelial sheet of polarized cells facing the blastocyst
signaling pathway remains inactive and unphosphorylated cavity, while the cells of the pluripotent epiblast are enclosed
Yap1 is retained in the nucleus to drive a trophectoderm-specific by the primitive endoderm and trophectoderm (Saiz and Plusa,
gene expression program (Sasaki, 2017; Ralston et al., 2010). 2013; Lanner and Rossant, 2010).
Blastocyst Formation and Hatching To complete preimplantation development and implant into
Once the embryo has established separate populations of inner the uterus, the embryo must finally hatch out of the protective
and outer cells, it is ready to transition to a fluid-filled blastocyst. shell formed by the ZP (Figure 2). Blastocyst hatching has
The blastocyst cavity begins as microlumens that emerge in the been suggested to occur as a combined result of localized pro-
intercellular spaces within the embryo (Marikawa and Alarcon, teolytic lysis of the ZP and mechanical pressure exerted by the
2012). These microlumens are formed by exocytosis of vesicles expanding blastocyst (Seshagiri et al., 2009; Perona and Was-
or vacuoles from the basal membrane of the outer cells (Figure 3) sarman, 1986). The embryo then gradually emerges from the
(Aziz and Alexandre, 1991). Sodium ions are actively transported ZP in a poorly defined process that is regulated by dynamic
across the outer cell layer through transmembrane pumps (Wat- cellular trophectodermal projections, molecular signaling fac-
son and Barcroft, 2001; Benos et al., 1985). The differential tors, and proteases released by both the blastocyst and the
sodium ion concentration generates an osmotic gradient that en- maternal endometrium. From this point on, the preimplantation
ables fluid to be pumped into the embryo (Barcroft et al., 2003), stage is over and the embryo must establish interactions with
enlarging and coalescing the microlumens into a single cavity. the uterus and undergo implantation to support the increasingly
Expansion of this cavity requires the formation of a paracellular complex demands of its continuing development.

672 Developmental Cell 45, June 18, 2018


Developmental Cell

Perspective

Figure 3. Instructions for Assembling a Mouse Embryo from Compacted Morula to Hatched Blastocyst
An overview of the main mechanisms currently known to direct preimplantation development from a compacted morula to a hatched blastocyst ready for
implantation. Scale bars represent 10 mm.

Future Challenges standing of how the embryo is formed. There are undoubtedly
Although research to date has yielded many insights into the additional mechanisms to discover, especially regarding the
myriad processes occurring during preimplantation develop- establishment and amplification of initial heterogeneities among
ment, substantial challenges remain to complete our under- blastomeres and the role of forces and changes in mechanical

Developmental Cell 45, June 18, 2018 673


Developmental Cell

Perspective

Figure 4. Progressive Refinement of Cell Fate


Biases in cell fate may arise at the four-cell stage due to a combination of biological noise and early heterogeneities between blastomeres in epigenetic
modifications (red triangles, green hexagons) and transcription factor binding to DNA. These early differences may be progressively amplified and refined until the
first two lineages diverge at the blastocyst stage.

properties in fate determination. A significant challenge ahead is and microtubule bridges (Zenker et al., 2017) between the cells
to identify crosstalk between the diverse processes identified in of the embryo, it would be interesting to determine if these
the early embryo and to understand how these mechanisms are cellular structures or organelles enable the transport of vesicles
integrated to progressively induce cells to transform into all the or signaling molecules as a mechanism of cell-cell communi-
requisite components of the embryo. Technological advance- cation.
ments and the implementation of approaches used in other sys- It will also be important to explore how rapid changes in cell
tems are likely to be critical to addressing these questions. shape in the mouse embryo correlate with redistributions of
How to Discover New Morphogenetic Processes signaling molecules, particularly those controlling the cytoskel-
Despite recent progress in revealing morphogenetic processes eton (Munjal and Lecuit, 2014). Pushing the application of fast-
controlling the assembly of the early mammalian embryo acquisition imaging techniques could enable discoveries of
(Figure 2), numerous additional cellular events will likely be changes in the distribution of second messengers, or the meta-
revealed as we broaden our technical spectrum of observation. bolic or mechanical state of cells at the whole-embryo level.
Advances in live-cell imaging, super-resolution and electron At the subcellular level, super-resolution microscopy has not
microscopy (EM) techniques, and computational analysis will yet been extensively applied to mouse embryos, yet these tech-
allow us to image faster events, occurring in the order of minutes niques will prove tremendously useful to reconstruct the spatial
or seconds, and the more microscopic organization of molecules arrangement of most organelles within early blastomeres. More-
and organelles within cells of intact embryos. over, the combination of EM with powerful computer segmenta-
Imaging methods providing faster time resolution could enable tion techniques has provided a previously inconceivable level of
the visualization of the movement of signaling vesicles, or the detail, uncovering the organization of brain synapses at the
reorganization of cytoskeletal elements, as cells change their nanometer scale (Hildebrand et al., 2017). This type of approach
fate, shape, and positions in vivo. For instance, imaging experi- is being applied in early mouse embryos to reveal the organiza-
ments in Drosophila embryos showed multiple endosomes car- tion of vesicular structures transported along microtubule tracks
rying signaling molecules shuttling rapidly between cells during in the intact embryo (Zenker et al., 2017). Together with ad-
asymmetric division (Derivery et al., 2015; Coumailleau et al., vances in histology methods to visualize fluorescently labeled
2009). Advances in light-sheet-based imaging technologies proteins in EM sections, these methods will significantly extend
may reveal if similar mechanisms of rapid endosome transport our understanding of subcellular organization within the embryo.
enable cell-cell communication in the mouse embryo (Whitehead Finally, it will also be valuable to explore the more macro-
et al., 2017). In particular, as recent studies demonstrated the scopic mechanisms regulating the assembly of the early mouse
presence of cellular protrusions (Fierro-Gonzalez et al., 2013) embryo. The timing of cell divisions may be tightly controlled

674 Developmental Cell 45, June 18, 2018


Developmental Cell

Perspective
Figure 5. Regulation of Mechanical Forces
Driving Embryo Formation
Signaling pathways and mechanical properties
within the embryo regulate actomyosin contrac-
tility and force production. Anisotropies in tensile
forces constrict the apical surface of cells, causing
them to internalize to form the inner cell mass.

patterning from homogeneous cell popu-


lations requires symmetry breaking,
which may be initiated by stochastic
processes. Recently, the application of
single-cell RNA sequencing technology
to the preimplantation embryo revealed
that differences in the transcriptome
initially arise between blastomeres as
early as the first cleavage division (Shi
et al., 2015; Biase et al., 2014; Piras
et al., 2014). This first variability is intro-
duced by partitioning errors unevenly
distributing substances into each
daughter cell after division. Although
some of these errors may be subse-
quently corrected, genes with a low
expression level are disproportionately
affected by partitioning errors (Shi
et al., 2015). This results in the establish-
ment of differential gene expression
between blastomeres. Early heterogene-
ities between blastomeres may be
further amplified by transcriptional noise
following zygotic gene activation during
the two-cell stage (Piras et al., 2014).
Random stochastic fluctuations, or
by structures providing intercellular communication within the noise, in the levels and activities of interacting genes and pro-
embryo. For example, cells extending filopodia on top of neigh- teins underlies the divergence of genetically identical cells in
boring cells appear to inhibit the division of their neighbors many biological systems (Eldar and Elowitz, 2010). Gene
(Fierro-Gonzalez et al., 2013). Moreover, cells connected via expression noise can be generated intrinsically by transcrip-
the microtubule interphase bridge undergo division in a tightly tional bursts, or extrinsically by the propagation of fluctuations
coordinated manner (Zenker et al., 2017). When a cell enters in the expression of one gene to cause fluctuations in expres-
mitosis, the depolymerization of its microtubule network is prop- sion of downstream genes. The introduction of biological noise
agated along the interphase bridge toward the connected sister into a system enables the probabilistic differentiation of genet-
cell, which only then enters mitosis. This suggests a mechanism ically identical cells while maintaining the flexibility to respond
whereby signals triggering mitosis could travel between sister to external factors (Balazsi et al., 2011). Within the embryo,
cells via the interphase bridge. positive and negative feedback loops fine-tune the initial differ-
These open problems can greatly benefit from the use of ences and affect the ratio of opposing lineage specifiers to bias
mathematical models predicting changes in embryo architecture cells toward different fates (Shi et al., 2015). In this way, cells
(Yu and Fernandez-Gonzalez, 2017). However, we should can slowly advance toward a specific fate but still maintain a
remain cautious about reductionist approaches that inadver- time window during which noise can act to correct any organi-
tently simplify the morphology and structural asymmetries that zational errors.
exist between cells in vivo. How inter-blastomere differences introduced at the two-cell
What Is Upstream of Changes in Epigenetics, stage by partitioning error and transcriptional noise relate to
Transcription Factor-DNA Interactions, and Gene the observed differences in epigenetic modifications (Torres-Pa-
Expression? dilla et al., 2007), transcription factor dynamics (White et al.,
Despite the identification of early epigenetic and transcriptional 2016a), and gene expression (Goolam et al., 2016; Shi et al.,
asymmetries, it remains an open question how these asymme- 2015; Biase et al., 2014) at the four-cell stage and beyond
tries initially develop and exactly how they are amplified to remains to be determined. Resolving this requires improvements
direct lineage divergence (Figure 4). The emergence of in the precise quantification of levels and dynamics of molecules

Developmental Cell 45, June 18, 2018 675


Developmental Cell

Perspective

at single-cell resolution without amplification and the capacity to Conclusion


track these molecules over time in living embryos (Zhao Decades of research have revealed basic mechanisms underly-
et al., 2016). ing key events during preimplantation development and yielded
Tilting the Balance of Mechanical Forces some preliminary instructions on how to assemble a mammalian
Most morphogenetic processes are driven by changes in the embryo. These instructions should be viewed as an initial guide
balance of mechanical forces within and between cells (Munjal and are likely to be refined and replaced over time as emerging
and Lecuit, 2014). Forces driving changes in cell shape are technologies facilitate new discoveries.
generated within the cell by actomyosin networks and trans- The high plasticity and regulative nature of the mammalian
mitted to neighboring cells and the extracellular environment embryo ensures that the identification of deterministic processes
via adhesion-mediated connections. In the preimplantation em- is more complicated than in non-mammalian systems. Although
bryo, the main forces at play are cortical tension and cell-cell blastomeres isolated from the embryo can reproduce key
adhesion mediated by adhesion molecules like E-cadherin. morphogenetic milestones such as polarization and compac-
Consistent with many other developmental systems, changes tion, they may do so via regulative processes. Thus results
in adhesion and cortical tension also drive morphological from such experiments should be interpreted with caution due
changes required to assemble the mouse embryo. For example, to the disruption of the normal geometry of cell-cell contacts
changes in both of these forces are important for embryo and mechanical forces within the embryo (Humiecka
compaction, and anisotropies in tensile forces drive the internal- et al., 2017).
ization of the cells that form the ICM (Maitre et al., 2015; Samar- Recent work also advises restraint in translating findings in
age et al., 2015). mouse embryos directly to humans. Although mouse and human
However, the key open challenge is to understand what mech- embryos share many similarities, there are important differences
anisms tilt the balance between these mechanical forces to in the timing of key events (Jukam et al., 2017) and the
direct morphogenesis. Two main mechanisms are likely in place: functioning of key lineages specifiers (Fogarty et al., 2017).
signaling pathways controlling the function or localization of key Future work adapting murine stem cell models of embryogenesis
molecules mediating cell-cell adhesion and cortical tension, and (Harrison et al., 2017) to human cells, and utilizing human
changes in mechanical properties within the embryo (Figure 5). embryos where possible (Shahbazi et al., 2016), will help to
For example, in cells and other systems, myosin-based contrac- define such species-specific permutations of common develop-
tility can be activated by signaling through the myosin light-chain mental processes.
kinase pathway, which commonly responds to intracellular cal-
cium, or via signaling through the Rho kinase pathway (Priya ACKNOWLEDGMENTS
and Yap, 2015). E-cadherin is recruited to junctions in response
to myosin activation and Rho signaling stabilizes junctional This work was supported by funds from the DFG (ZE99711), SNF
integrity (Hoffman and Yap, 2015). Actin regulators such as (PBLAP3145877), and HFSP (LT000164/2015) to J.Z. and from A*STAR,
EMBO, and HHMI-Wellcome Trust to N.P. We apologize to those colleagues
WAVE and Arp2/3 can be recruited to junctions to differentially
whose work we are unable to review here due to space limitations.
regulate actin organization and connect adhesion to contractility
(Brieher and Yap, 2013). While many of these pathways have
REFERENCES
been described in detail in other systems, their regulation is rela-
tively unexplored in the preimplantation embryo. The application Akhmanova, A., and Hoogenraad, C.C. (2015). Microtubule minus-end-target-
of optogenetic techniques for controlling protein localization has ing proteins. Curr. Biol. 25, R162–R171.
recently demonstrated how emerging technologies may enable
Alarcon, V.B., and Marikawa, Y. (2005). Unbiased contribution of the first two
further dissection of these pathways in vivo (Zhu et al., 2017). blastomeres to mouse blastocyst development. Mol. Reprod. Dev. 72,
Less is known, however, about changes in mechanical proper- 354–361.
ties in vivo, since few techniques enable their measurement in
Anani, S., Bhat, S., Honma-Yamanaka, N., Krawchuk, D., and Yamanaka, Y.
complex systems like embryos and tissues. For instance, it is (2014). Initiation of Hippo signaling is linked to polarity rather than to cell posi-
possible that, as blastomeres undergo cleavage divisions, they tion in the pre-implantation mouse embryo. Development 141, 2813–2824.
may be subject to changes in intracellular pressure, cytoplasmic Aziz, M., and Alexandre, H. (1991). The origin of the nascent blastocoele in
viscosity, or stiffness. Dynamic changes in these mechanical preimplantation mouse embryos ultrastructural cytochemistry and effect of
properties would affect the embryo’s response to internal forces. chloroquine. Roux Arch. Dev. Biol. 200, 77–85.
It will therefore be important to apply new methods to measure Balazsi, G., van Oudenaarden, A., and Collins, J.J. (2011). Cellular decision
mechanical properties in the embryo, which could include new making and biological noise: from microbes to mammals. Cell 144, 910–925.
tools for measuring intracellular pressure (Gomez-Martinez Barcroft, L.C., Offenberg, H., Thomsen, P., and Watson, A.J. (2003). Aquaporin
et al., 2013); optical tweezers, microindenters, or cantilevers to proteins in murine trophectoderm mediate transepithelial water movements
apply forces and measure responses; and magnetically respon- during cavitation. Dev. Biol. 256, 342–354.
sive ferrofluid microdroplets that enable highly precise measure- Behringer, R. (2014). Manipulating the Mouse Embryo: A Laboratory Manual
ments of mechanical properties such as viscosity in tissues and (Cold Spring Harbor Laboratory Press).
embryos (Serwane et al., 2017; Campas, 2016). It should be
Benos, D.J., Biggers, J.D., Balaban, R.S., Mills, J.W., and Overstrom, E.W.
emphasized that mechanical forces remain the main executers (1985). Developmental aspects of sodium-dependent transport processes of
of embryo patterning, yet future experiments would enable us preimplantation rabbit embryos. Soc. Gen. Physiol. Ser. 39, 211–235.
to examine how signaling pathways and changes in mechanical Bessonnard, S., Coqueran, S., Vandormael-Pournin, S., Dufour, A., Artus, J.,
properties are intertwined to affect these forces in vivo. and Cohen-Tannoudji, M. (2017). ICM conversion to epiblast by FGF/ERK

676 Developmental Cell 45, June 18, 2018


Developmental Cell

Perspective
inhibition is limited in time and requires transcription and protein degradation. Ducibella, T., Ukena, T., Karnovsky, M., and Anderson, E. (1977). Changes in
Sci. Rep. 7, 12285. cell surface and cortical cytoplasmic organization during early embryogenesis
in the preimplantation mouse embryo. J. Cell Biol. 74, 153–167.
Biase, F.H., Cao, X., and Zhong, S. (2014). Cell fate inclination within 2-cell and
4-cell mouse embryos revealed by single-cell RNA sequencing. Genome Res. Dumont, J., Million, K., Sunderland, K., Rassinier, P., Lim, H., Leader, B., and
24, 1787–1796. Verlhac, M.H. (2007). Formin-2 is required for spindle migration and for the late
steps of cytokinesis in mouse oocytes. Dev. Biol. 301, 254–265.
Brieher, W.M., and Yap, A.S. (2013). Cadherin junctions and their cytoskele-
ton(s). Curr. Opin. Cell Biol. 25, 39–46. Edwards, R.G. (1965). Maturation in vitro of mouse, sheep, cow, pig, rhesus
monkey and human ovarian oocytes. Nature 208, 349–351.
Brinster, R.L. (1963). A method for in vitro cultivation of mouse ova from two-
cell to blastocyst. Exp. Cell Res. 32, 205–208. Eldar, A., and Elowitz, M.B. (2010). Functional roles for noise in genetic circuits.
Nature 467, 167–173.
Burton, A., Muller, J., Tu, S., Padilla-Longoria, P., Guccione, E., and Torres-Pa-
dilla, M.E. (2013). Single-cell profiling of epigenetic modifiers identifies Fierro-Gonzalez, J.C., White, M.D., Silva, J.C., and Plachta, N. (2013). Cad-
PRDM14 as an inducer of cell fate in the mammalian embryo. Cell Rep. 5, herin-dependent filopodia control preimplantation embryo compaction. Nat.
687–701. Cell Biol. 15, 1424–1433.

Burton, A., and Torres-Padilla, M.E. (2014). Chromatin dynamics in the regula- Fogarty, N.M.E., McCarthy, A., Snijders, K.E., Powell, B.E., Kubikova, N., Bla-
tion of cell fate allocation during early embryogenesis. Nat. Rev. Mol. Cell Biol. keley, P., Lea, R., Elder, K., Wamaitha, S.E., Kim, D., et al. (2017). Genome
15, 723–734. editing reveals a role for OCT4 in human embryogenesis. Nature 550, 67–73.

Bury, L., Coelho, P.A., Simeone, A., Ferries, S., Eyers, C.E., Eyers, P.A., Zer- Fujimori, T., Kurotaki, Y., Miyazaki, J., and Nabeshima, Y. (2003). Analysis of
nicka-Goetz, M., and Glover, D.M. (2017). Plk4 and Aurora A cooperate in cell lineage in two- and four-cell mouse embryos. Development 130,
the initiation of acentriolar spindle assembly in mammalian oocytes. J. Cell 5113–5122.
Biol. 216, 3571–3590.
Gao, Y., Liu, X., Tang, B., Li, C., Kou, Z., Li, L., Liu, W., Wu, Y., Kou, X., Li, J.,
Campas, O. (2016). A toolbox to explore the mechanics of living embryonic et al. (2017). Protein expression landscape of mouse embryos during pre-im-
tissues. Semin. Cell Dev. Biol. 55, 119–130. plantation development. Cell Rep. 21, 3957–3969.

Chaigne, A., Campillo, C., Gov, N.S., Voituriez, R., Azoury, J., Umana-Diaz, C., Gomez-Martinez, R., Hernandez-Pinto, A.M., Duch, M., Vazquez, P., Zinoviev,
Almonacid, M., Queguiner, I., Nassoy, P., Sykes, C., et al. (2013). A soft cortex K., de la Rosa, E.J., Esteve, J., Suarez, T., and Plaza, J.A. (2013). Silicon chips
is essential for asymmetric spindle positioning in mouse oocytes. Nat. Cell detect intracellular pressure changes in living cells. Nat. Nanotechnol. 8,
Biol. 15, 958–966. 517–521.

Chaigne, A., Campillo, C., Voituriez, R., Gov, N.S., Sykes, C., Verlhac, M.H., Goolam, M., Scialdone, A., Graham, S.J., Macaulay, I.C., Jedrusik, A., Hupa-
and Terret, M.E. (2016). F-actin mechanics control spindle centring in the lowska, A., Voet, T., Marioni, J.C., and Zernicka-Goetz, M. (2016). Heterogene-
mouse zygote. Nat. Commun. 7, 10253. ity in Oct4 and Sox2 targets biases cell fate in 4-cell mouse embryos. Cell
165, 61–74.
Chazaud, C., Yamanaka, Y., Pawson, T., and Rossant, J. (2006). Early lineage
Guo, G., Huss, M., Tong, G.Q., Wang, C., Li Sun, L., Clarke, N.D., and Robson,
segregation between epiblast and primitive endoderm in mouse blastocysts
P. (2010). Resolution of cell fate decisions revealed by single-cell gene expres-
through the Grb2-MAPK pathway. Dev. Cell 10, 615–624.
sion analysis from zygote to blastocyst. Dev. Cell 18, 675–685.
Clift, D., and Schuh, M. (2013). Restarting life: fertilization and the transition
Habibi, E., and Stunnenberg, H.G. (2017). Transcriptional and epigenetic con-
from meiosis to mitosis. Nat. Rev. Mol. Cell Biol. 14, 549–562.
trol in mouse pluripotency: lessons from in vivo and in vitro studies. Curr. Opin.
Genet. Dev. 46, 114–122.
Clift, D., and Schuh, M. (2015). A three-step MTOC fragmentation mechanism
facilitates bipolar spindle assembly in mouse oocytes. Nat. Commun. 6, 7217. Harrison, S.E., Sozen, B., Christodoulou, N., Kyprianou, C., and Zernicka-
Goetz, M. (2017). Assembly of embryonic and extraembryonic stem cells to
Cockburn, K., Biechele, S., Garner, J., and Rossant, J. (2013). The Hippo mimic embryogenesis in vitro. Science 356, https://doi.org/10.1126/science.
pathway member Nf2 is required for inner cell mass specification. Curr. Biol. aal1810.
23, 1195–1201.
Hiiragi, T., and Solter, D. (2004). First cleavage plane of the mouse egg is not
Cockburn, K., and Rossant, J. (2010). Making the blastocyst: lessons from the predetermined but defined by the topology of the two apposing pronuclei. Na-
mouse. J. Clin. Invest. 120, 995–1003. ture 430, 360–364.
Conduit, P.T., Wainman, A., and Raff, J.W. (2015). Centrosome function and Hildebrand, D.G.C., Cicconet, M., Torres, R.M., Choi, W., Quan, T.M., Moon,
assembly in animal cells. Nat. Rev. Mol. Cell Biol. 16, 611–624. J., Wetzel, A.W., Scott Champion, A., Graham, B.J., Randlett, O., et al.
(2017). Whole-brain serial-section electron microscopy in larval zebrafish. Na-
Coumailleau, F., Furthauer, M., Knoblich, J.A., and Gonzalez-Gaitan, M. ture 545, 345–349.
(2009). Directional Delta and Notch trafficking in Sara endosomes during
asymmetric cell division. Nature 458, 1051–1055. Hirate, Y., Hirahara, S., Inoue, K., Suzuki, A., Alarcon, V.B., Akimoto, K., Hirai,
T., Hara, T., Adachi, M., Chida, K., et al. (2013). Polarity-dependent distribution
Davies, J.A. (2017). Adaptive self-organization in the embryo: its importance to of angiomotin localizes Hippo signaling in preimplantation embryos. Curr. Biol.
adult anatomy and to tissue engineering. J. Anat. 232, 524–533. 23, 1181–1194.

Deng, M., Suraneni, P., Schultz, R.M., and Li, R. (2007). The Ran GTPase me- Hoffman, B.D., and Yap, A.S. (2015). Towards a dynamic understanding of
diates chromatin signaling to control cortical polarity during polar body extru- cadherin-based mechanobiology. Trends Cell Biol. 25, 803–814.
sion in mouse oocytes. Dev. Cell 12, 301–308.
Home, P., Kumar, R.P., Ganguly, A., Saha, B., Milano-Foster, J., Bhattacharya,
Derivery, E., Seum, C., Daeden, A., Loubery, S., Holtzer, L., Julicher, F., and B., Ray, S., Gunewardena, S., Paul, A., Camper, S.A., et al. (2017). Genetic
Gonzalez-Gaitan, M. (2015). Polarized endosome dynamics by spindle asym- redundancy of GATA factors in the extraembryonic trophoblast lineage
metry during asymmetric cell division. Nature 528, 280–285. ensures the progression of preimplantation and postimplantation mammalian
development. Development 144, 876–888.
Dietrich, J.E., and Hiiragi, T. (2007). Stochastic patterning in the mouse pre-im-
plantation embryo. Development 134, 4219–4231. Hoppe, P.C., and Whitten, W.K. (1972). Does X chromosome inactivation
occur during mitosis of first cleavage? Nature 239, 520.
Ducibella, T., and Anderson, E. (1975). Cell shape and membrane changes in
the eight-cell mouse embryo: prerequisites for morphogenesis of the blasto- Huh, D., and Paulsson, J. (2011). Random partitioning of molecules at cell
cyst. Dev. Biol. 47, 45–58. division. Proc. Natl. Acad. Sci. USA 108, 15004–15009.

Developmental Cell 45, June 18, 2018 677


Developmental Cell

Perspective
Humiecka, M., Szpila, M., Klos, P., Maleszewski, M., and Szczepanska, K. Molotkov, A., Mazot, P., Brewer, J.R., Cinalli, R.M., and Soriano, P. (2017).
(2017). Mouse blastomeres acquire ability to divide asymmetrically before Distinct requirements for FGFR1 and FGFR2 in primitive endoderm develop-
compaction. PLoS One 12, e0175032. ment and exit from pluripotency. Dev. Cell 41, 511–526.e4.

Jachowicz, J.W., Bing, X., Pontabry, J., Boskovic, A., Rando, O.J., and Torres- Motosugi, N., Bauer, T., Polanski, Z., Solter, D., and Hiiragi, T. (2005). Polarity
Padilla, M.E. (2017). LINE-1 activation after fertilization regulates global chro- of the mouse embryo is established at blastocyst and is not prepatterned.
matin accessibility in the early mouse embryo. Nat. Genet. 49, 1502–1510. Genes Dev. 19, 1081–1092.

Johnson, M.H. (2009). From mouse egg to mouse embryo: polarities, axes, Munjal, A., and Lecuit, T. (2014). Actomyosin networks and tissue morphogen-
and tissues. Annu. Rev. Cell Dev. Biol. 25, 483–512. esis. Development 141, 1789–1793.

Johnson, M.H., and Ziomek, C.A. (1981). The foundation of two distinct cell Nishioka, N., Inoue, K., Adachi, K., Kiyonari, H., Ota, M., Ralston, A., Yabuta,
lineages within the mouse morula. Cell 24, 71–80. N., Hirahara, S., Stephenson, R.O., Ogonuki, N., et al. (2009). The Hippo
signaling pathway components Lats and Yap pattern Tead4 activity to distin-
Jukam, D., Shariati, S.A.M., and Skotheim, J.M. (2017). Zygotic genome acti- guish mouse trophectoderm from inner cell mass. Dev. Cell 16, 398–410.
vation in vertebrates. Dev. Cell 42, 316–332.
Nissen, S.B., Perera, M., Gonzalez, J.M., Morgani, S.M., Jensen, M.H., Snep-
Kang, M., Garg, V., and Hadjantonakis, A.K. (2017). Lineage establishment and pen, K., Brickman, J.M., and Trusina, A. (2017). Four simple rules that are suf-
progression within the inner cell mass of the mouse blastocyst requires FGFR1 ficient to generate the mammalian blastocyst. PLoS Biol. 15, e2000737.
and FGFR2. Dev. Cell 41, 496–510.e5.
Ohnishi, Y., Huber, W., Tsumura, A., Kang, M., Xenopoulos, P., Kurimoto, K.,
Kaur, G., Costa, M.W., Nefzger, C.M., Silva, J., Fierro-Gonzalez, J.C., Polo, Oles, A.K., Arauzo-Bravo, M.J., Saitou, M., Hadjantonakis, A.K., et al. (2014).
J.M., Bell, T.D., and Plachta, N. (2013). Probing transcription factor diffusion Cell-to-cell expression variability followed by signal reinforcement progres-
dynamics in the living mammalian embryo with photoactivatable fluorescence sively segregates early mouse lineages. Nat. Cell Biol. 16, 27–37.
correlation spectroscopy. Nat. Commun. 4, 1637.
Perona, R.M., and Wassarman, P.M. (1986). Mouse blastocysts hatch in vitro
Korotkevich, E., Niwayama, R., Courtois, A., Friese, S., Berger, N., Buchholz, by using a trypsin-like proteinase associated with cells of mural trophecto-
F., and Hiiragi, T. (2017). The apical domain is required and sufficient for the derm. Dev. Biol. 114, 42–52.
first lineage segregation in the mouse embryo. Dev. Cell 40, 235–247.e7.
Piotrowska-Nitsche, K., Perea-Gomez, A., Haraguchi, S., and Zernicka-Goetz,
Krawchuk, D., Honma-Yamanaka, N., Anani, S., and Yamanaka, Y. (2013). M. (2005). Four-cell stage mouse blastomeres have different developmental
FGF4 is a limiting factor controlling the proportions of primitive endoderm properties. Development 132, 479–490.
and epiblast in the ICM of the mouse blastocyst. Dev. Biol. 384, 65–71.
Piras, V., Tomita, M., and Selvarajoo, K. (2014). Transcriptome-wide variability
Lanner, F., and Rossant, J. (2010). The role of FGF/Erk signaling in pluripotent in single embryonic development cells. Sci. Rep. 4, 7137.
cells. Development 137, 3351–3360.
Plachta, N., Bollenbach, T., Pease, S., Fraser, S.E., and Pantazis, P. (2011).
Lecuit, T., Lenne, P.F., and Munro, E. (2011). Force generation, transmission, Oct4 kinetics predict cell lineage patterning in the early mammalian embryo.
and integration during cell and tissue morphogenesis. Annu. Rev. Cell Dev. Nat. Cell Biol. 13, 117–123.
Biol. 27, 157–184.
Plusa, B., Frankenberg, S., Chalmers, A., Hadjantonakis, A.K., Moore, C.A.,
Levy, J.B., Johnson, M.H., Goodall, H., and Maro, B. (1986). The timing of Papalopulu, N., Papaioannou, V.E., Glover, D.M., and Zernicka-Goetz, M.
compaction: control of a major developmental transition in mouse early (2005). Downregulation of Par3 and aPKC function directs cells towards the
embryogenesis. J. Embryol. Exp. Morphol. 95, 213–237. ICM in the preimplantation mouse embryo. J. Cell Sci. 118, 505–515.

Li, R., and Albertini, D.F. (2013). The road to maturation: somatic cell interac- Priya, R., and Yap, A.S. (2015). Active tension: the role of cadherin adhesion
tion and self-organization of the mammalian oocyte. Nat. Rev. Mol. Cell Biol. and signaling in generating junctional contractility. Curr. Top. Dev. Biol.
14, 141–152. 112, 65–102.

Louvet, S., Aghion, J., Santa-Maria, A., Mangeat, P., and Maro, B. (1996). Ezrin Ralston, A., Cox, B.J., Nishioka, N., Sasaki, H., Chea, E., Rugg-Gunn, P., Guo,
becomes restricted to outer cells following asymmetrical division in the preim- G., Robson, P., Draper, J.S., and Rossant, J. (2010). Gata3 regulates tropho-
plantation mouse embryo. Dev. Biol. 177, 568–579. blast development downstream of Tead4 and in parallel to Cdx2. Development
137, 395–403.
Maitre, J.L., Berthoumieux, H., Krens, S.F., Salbreux, G., Julicher, F., Paluch,
E., and Heisenberg, C.P. (2012). Adhesion functions in cell sorting by mechan- Ralston, A., and Rossant, J. (2008). Cdx2 acts downstream of cell polarization
ically coupling the cortices of adhering cells. Science 338, 253–256. to cell-autonomously promote trophectoderm fate in the early mouse embryo.
Dev. Biol. 313, 614–629.
Maitre, J.L., Niwayama, R., Turlier, H., Nedelec, F., and Hiiragi, T. (2015). Pul-
satile cell-autonomous contractility drives compaction in the mouse embryo. Reijo Pera, R.A., and Prezzoto, L. (2016). Species-specific variation among
Nat. Cell Biol. 17, 849–855. mammals. Curr. Top. Dev. Biol. 120, 401–420.

Maitre, J.L., Turlier, H., Illukkumbura, R., Eismann, B., Niwayama, R., Nedelec, Rossant, J., and Vijh, K.M. (1980). Ability of outside cells from preimplantation
F., and Hiiragi, T. (2016). Asymmetric division of contractile domains couples mouse embryos to form inner cell mass derivatives. Dev. Biol. 76, 475–482.
cell positioning and fate specification. Nature 536, 344–348.
Saiz, N., and Plusa, B. (2013). Early cell fate decisions in the mouse embryo.
Marikawa, Y., and Alarcon, V.B. (2012). Creation of trophectoderm, the first Reproduction 145, R65–R80.
epithelium, in mouse preimplantation development. Results Probl. Cell Differ.
55, 165–184. Samarage, C.R., White, M.D., Alvarez, Y.D., Fierro-Gonzalez, J.C., Henon, Y.,
Jesudason, E.C., Bissiere, S., Fouras, A., and Plachta, N. (2015). Cortical ten-
Mihajlovic, A.I., and Bruce, A.W. (2017). The first cell-fate decision of mouse sion allocates the first inner cells of the mammalian embryo. Dev. Cell 34,
preimplantation embryo development: integrating cell position and polarity. 435–447.
Open Biol. 7, https://doi.org/10.1098/rsob.170210.
Sasai, Y. (2013). Cytosystems dynamics in self-organization of tissue architec-
Mistri, T.K., Arindrarto, W., Ng, W.P., Wang, C., Lim, L.H., Sun, L., Chambers, ture. Nature 493, 318–326.
I., Wohland, T., and Robson, P. (2018). Dynamic changes in Sox2 spatio-tem-
poral expression promote the second cell fate decision through Fgf4/Fgfr2 sig- Sasaki, H. (2017). Roles and regulations of Hippo signaling during preimplan-
nalling in preimplantation mouse embryos. Biochem. J. 475, 1075–1089. tation mouse development. Dev. Growth Differ. 59, 12–20.

Mogessie, B., and Schuh, M. (2017). Actin protects mammalian eggs against Serwane, F., Mongera, A., Rowghanian, P., Kealhofer, D.A., Lucio, A.A., Hock-
chromosome segregation errors. Science 357, https://doi.org/10.1126/sci- enbery, Z.M., and Campas, O. (2017). In vivo quantification of spatially varying
ence.aal1647. mechanical properties in developing tissues. Nat. Methods 14, 181–186.

678 Developmental Cell 45, June 18, 2018


Developmental Cell

Perspective
Seshagiri, P.B., Sen Roy, S., Sireesha, G., and Rao, R.P. (2009). Cellular and White, M.D., Bissiere, S., Alvarez, Y.D., and Plachta, N. (2016b). Mouse
molecular regulation of mammalian blastocyst hatching. J. Reprod. Immunol. embryo compaction. Curr. Top. Dev. Biol. 120, 235–258.
83, 79–84.
White, M.D., Zenker, J., Bissiere, S., and Plachta, N. (2016c). How cells change
Shahbazi, M.N., Jedrusik, A., Vuoristo, S., Recher, G., Hupalowska, A., Bolton, shape and position in the early mammalian embryo. Curr. Opin. Cell Biol.
V., Fogarty, N.M., Campbell, A., Devito, L.G., Ilic, D., et al. (2016). Self-organi- 44, 7–13.
zation of the human embryo in the absence of maternal tissues. Nat. Cell Biol.
18, 700–708. Whitehead, L.W., McArthur, K., Geoghegan, N.D., and Rogers, K.L. (2017).
The reinvention of twentieth century microscopy for three-dimensional imag-
Shi, J., Chen, Q., Li, X., Zheng, X., Zhang, Y., Qiao, J., Tang, F., Tao, Y., Zhou, ing. Immunol. Cell Biol. 95, 520–524.
Q., and Duan, E. (2015). Dynamic transcriptional symmetry-breaking in pre-im-
plantation mammalian embryo development revealed by single-cell RNA-seq. Whitten, W.K., and Biggers, J.D. (1968). Complete development in vitro of the
Development 142, 3468–3477. pre-implantation stages of the mouse in a simple chemically defined medium.
J. Reprod. Fertil. 17, 399–401.
Shi, X., Yin, Z., Ling, B., Wang, L., Liu, C., Ruan, X., Zhang, W., and Chen, L.
(2017). Rho differentially regulates the Hippo pathway by modulating the inter- Wicklow, E., Blij, S., Frum, T., Hirate, Y., Lang, R.A., Sasaki, H., and Ralston, A.
action between Amot and Nf2 in the blastocyst. Development 144, 3957–3967. (2014). HIPPO pathway members restrict SOX2 to the inner cell mass where it
promotes ICM fates in the mouse blastocyst. PLoS Genet. 10, e1004618.
Stephenson, R.O., Yamanaka, Y., and Rossant, J. (2010). Disorganized epithe-
lial polarity and excess trophectoderm cell fate in preimplantation embryos Yamanaka, Y., Lanner, F., and Rossant, J. (2010). FGF signal-dependent
lacking E-cadherin. Development 137, 3383–3391. segregation of primitive endoderm and epiblast in the mouse blastocyst.
Development 137, 715–724.
Suwinska, A., Czolowska, R., Ozdzenski, W., and Tarkowski, A.K. (2008). Blas-
tomeres of the mouse embryo lose totipotency after the fifth cleavage division: Yi, K., Rubinstein, B., Unruh, J.R., Guo, F., Slaughter, B.D., and Li, R. (2013).
expression of Cdx2 and Oct4 and developmental potential of inner and outer Sequential actin-based pushing forces drive meiosis I chromosome migration
blastomeres of 16- and 32-cell embryos. Dev. Biol. 322, 133–144. and symmetry breaking in oocytes. J. Cell Biol. 200, 567–576.
Tabansky, I., Lenarcic, A., Draft, R.W., Loulier, K., Keskin, D.B., Rosains, J., Yi, K., Unruh, J.R., Deng, M., Slaughter, B.D., Rubinstein, B., and Li, R. (2011).
Rivera-Feliciano, J., Lichtman, J.W., Livet, J., Stern, J.N., et al. (2013). Devel- Dynamic maintenance of asymmetric meiotic spindle position through Arp2/3-
opmental bias in cleavage-stage mouse blastomeres. Curr. Biol. 23, 21–31. complex-driven cytoplasmic streaming in mouse oocytes. Nat. Cell Biol. 13,
1252–1258.
Tarkowski, A.K. (1959). Experiments on the development of isolated blasto-
meres of mouse eggs. Nature 184, 1286–1287. Yu, J.C., and Fernandez-Gonzalez, R. (2017). Quantitative modelling of epithe-
lial morphogenesis: integrating cell mechanics and molecular dynamics.
Tarkowski, A.K., and Wroblewska, J. (1967). Development of blastomeres of
Semin. Cell Dev. Biol. 67, 153–160.
mouse eggs isolated at the 4- and 8-cell stage. J. Embryol. Exp. Morphol.
18, 155–180. Zenker, J., White, M.D., Templin, R.M., Parton, R.G., Thorn-Seshold, O., Bis-
siere, S., and Plachta, N. (2017). A microtubule-organizing center directing
Thomas, F.C., Sheth, B., Eckert, J.J., Bazzoni, G., Dejana, E., and Fleming,
intracellular transport in the early mouse embryo. Science 357, 925–928.
T.P. (2004). Contribution of JAM-1 to epithelial differentiation and tight-junc-
tion biogenesis in the mouse preimplantation embryo. J. Cell Sci. 117,
Zenker, J., White, M.D., Gasnier, M., Alvarez, Y.D., Lim, H.Y.G., Bissiere, S.,
5599–5608.
Biro, M., and Plachta, N. (2018). Expanding actin rings zipper the mouse
Torres-Padilla, M.E., Parfitt, D.E., Kouzarides, T., and Zernicka-Goetz, M. embryo for blastocyst formation. Cell 173, 776–791.
(2007). Histone arginine methylation regulates pluripotency in the early mouse
Zernicka-Goetz, M., Morris, S.A., and Bruce, A.W. (2009). Making a firm deci-
embryo. Nature 445, 214–218.
sion: multifaceted regulation of cell fate in the early mouse embryo. Nat. Rev.
Vestweber, D., Gossler, A., Boller, K., and Kemler, R. (1987). Expression and Genet. 10, 467–477.
distribution of cell adhesion molecule uvomorulin in mouse preimplantation
embryos. Dev. Biol. 124, 451–456. Zhao, Z.W., White, M.D., Alvarez, Y.D., Zenker, J., Bissiere, S., and Plachta, N.
(2017). Quantifying transcription factor-DNA binding in single cells in vivo with
Vinot, S., Le, T., Ohno, S., Pawson, T., Maro, B., and Louvet-Vallee, S. (2005). photoactivatable fluorescence correlation spectroscopy. Nat. Protoc. 12,
Asymmetric distribution of PAR proteins in the mouse embryo begins at the 1458–1471.
8-cell stage during compaction. Dev. Biol. 282, 307–319.
Zhao, Z.W., White, M.D., Bissiere, S., Levi, V., and Plachta, N. (2016). Quanti-
Wang, H., Ding, T., Brown, N., Yamamoto, Y., Prince, L.S., Reese, J., and Pa- tative imaging of mammalian transcriptional dynamics: from single cells to
ria, B.C. (2008). Zonula occludens-1 (ZO-1) is involved in morula to blastocyst whole embryos. BMC Biol. 14, 115.
transformation in the mouse. Dev. Biol. 318, 112–125.
Zheng, Z., Li, H., Zhang, Q., Yang, L., and Qi, H. (2016). Unequal distribution of
Watanabe, T., Biggins, J.S., Tannan, N.B., and Srinivas, S. (2014). Limited pre- 16S mtrRNA at the 2-cell stage regulates cell lineage allocations in mouse em-
dictive value of blastomere angle of division in trophectoderm and inner cell bryos. Reproduction 151, 351–367.
mass specification. Development 141, 2279–2288.
Zhou, L.Q., and Dean, J. (2015). Reprogramming the genome to totipotency in
Watson, A.J., and Barcroft, L.C. (2001). Regulation of blastocyst formation. mouse embryos. Trends Cell Biol. 25, 82–91.
Front. Biosci. 6, D708–D730.
Zhu, M., Leung, C.Y., Shahbazi, M.N., and Zernicka-Goetz, M. (2017). Actomy-
Watson, A.J., and Kidder, G.M. (1988). Immunofluorescence assessment of osin polarisation through PLC-PKC triggers symmetry breaking of the mouse
the timing of appearance and cellular distribution of Na/K-ATPase during embryo. Nat. Commun. 8, 921.
mouse embryogenesis. Dev. Biol. 126, 80–90.
Ziomek, C.A., and Johnson, M.H. (1980). Cell surface interaction induces
Wennekamp, S., Mesecke, S., Nedelec, F., and Hiiragi, T. (2013). A self-orga- polarization of mouse 8-cell blastomeres at compaction. Cell 21, 935–942.
nization framework for symmetry breaking in the mammalian embryo. Nat.
Rev. Mol. Cell Biol. 14, 454–461. Ziomek, C.A., and Johnson, M.H. (1982). The roles of phenotype and position
in guiding the fate of 16-cell mouse blastomeres. Dev. Biol. 91, 440–447.
White, M.D., Angiolini, J.F., Alvarez, Y.D., Kaur, G., Zhao, Z.W., Mocskos, E.,
Bruno, L., Bissiere, S., Levi, V., and Plachta, N. (2016a). Long-lived binding of Ziomek, C.A., Johnson, M.H., and Handyside, A.H. (1982). The developmental
Sox2 to DNA predicts cell fate in the four-cell mouse embryo. Cell 165, 75–87. potential of mouse 16-cell blastomeres. J. Exp. Zool 221, 345–355.

Developmental Cell 45, June 18, 2018 679


View publication stats

You might also like