Pre-implantation mammalian development unites extreme plasticity with a robust outcome: the formation of a blastocyst, an organised multi-layered structure ready for implantation. The process of blastocyst formation is one of the best-known examples of self-organisation. The first three cell lineages in mammalian development specify and arrange themselves during the morphogenic process based on cell-cell interactions. Despite decades of research, the unifying principles driving early mammalian development are still not fully defined. Here, we discuss the role of physical forces, and molecular and cellular mechanisms, in driving self-organisation and lineage formation that are shared between eutherian mammals.

One of the most fundamental questions in biology is: how can one cell, a fertilised zygote, give rise to all of the different cell types found in the adult body? The sequential formation of different types of cells that arrange themselves into tissues and organs has been studied extensively in various model organisms. The mammalian embryo offers a unique perspective for tackling this question, as its early development is not ‘pre-programmed’ by the arranged deposition of maternal determinants (neither proteins nor mRNA, as in other organisms), nor does it fully depend on the position-related information linked to the sperm entry point or any specific landmarks related to the oocyte (Plusa and Hadjantonakis, 2018; Wennekamp et al., 2013). Instead, a combination of cell-cell and cell-environment interactions, geometrical constraints, changing physical forces and adaptive gene regulatory networks drive the formation of the first mammalian embryonic lineages and their robust organisation into a three-dimensional structure in preparation for implantation. Such a system relies on self-organisation; therefore, when the naïve architecture of the embryo is changed, for example by the death of some of its cells or external manipulation, the embryo can adapt and continue developing (reviewed by Klimczewska et al., 2018).

Although there is some variation, pre-implantation development of all eutherian mammals follows similar morphological stages (discussed below) and results in the formation of a fluid-filled blastocyst made of three distinct lineages. The outer layer of the blastocyst, the trophectoderm (TE), is the first to be specified from cells that occupy an ‘outside position’ in the embryo before blastocyst formation. The fully-specified TE forms an epithelium, and during implantation it facilitates contact between maternal and embryonic tissues, later giving rise to the embryonic part of the placenta (Aplin and Ruane, 2017; Hemberger et al., 2020). Concomitantly with the formation of the outer layer surrounding the embryo, the cells positioned inside give rise to the inner cell mass (ICM), a group of inner cells pushed to one side of the embryo by the growing blastocyst cavity (Johnson and McConnell, 2004; Johnson and Ziomek, 1981a). The ICM further differentiates into the epiblast (Epi), the founding lineage of the foetus, and the primitive endoderm (PrE, also known as the hypoblast), a supportive lineage that separates the Epi from the blastocyst cavity and contributes to the endoderm of the yolk sac (Dickson, 1966; Gardner and Rossant, 1979). Specification of these three lineages, as well as preparation for implantation, are the main objectives of mammalian pre-implantation development.

Despite considerable progress in our understanding of the genetic networks that direct the specification of the first three lineages in development, we still know very little about how these networks integrate and respond to the physical forces affecting cells in early embryos. The questions surrounding how the cell microenvironment modulates and transmits these forces have only recently begun to be answered.

Mounting evidence for differences in pre-implantation development among mammalian species has been recently presented and demands re-evaluation of information we have gleaned so far from the mouse model, as well as posing questions about conserved and fundamental features of pre-implantation mammalian development. In this Review, we explore the diversity of embryos from different mammalian species, looking for the common denominator of early eutherian development. The development of monotremes and marsupials is beyond the scope of this Review and is reviewed elsewhere (Frankenberg et al., 2016).

Cleavages

Similarities in early embryo architecture among different groups of eutherian mammals suggest common mechanisms driving pre-implantation development. After fertilisation, the mammalian embryo undergoes sequential morphological changes before it reaches the blastocyst stage (Fig. 1). The zygote divides into two, then four, then eight identical-looking cells called blastomeres. These early divisions are not fully synchronous; for example, 3-cell embryos may transiently exist during the 2- to 4-cell-stage transition. However, synchronicity might be important for proper embryo development, as a high synchronicity of divisions has been correlated with a higher probability of successful implantation of human embryos resulting from in vitro fertilisation (IVF) treatment (Cetinkaya et al., 2015; Wong et al., 2010; reviewed by Milewski and Ajduk, 2017). There is no growth phase during the cell cycle for these divisions. Therefore, each subsequent division produces smaller and smaller cells, which leads to changes in the surface area-to-volume ratio (Aiken et al., 2004). These types of divisions are called cleavages and they appear to persist through most of the pre-implantation period (reviewed by Rossant and Tam, 2009). During early cleavages, the embryonic genome remains transcriptionally inactive, and progression of early development depends on maternally stored proteins and mRNA. The timing of embryonic genome activation differs among mammalian species: major embryonic genome activation occurs as early as the 2-cell stage in rodents (Latham et al., 1992), between the 4- and 8-cell stages in humans (Braude et al., 1988; Dobson et al., 2004) and at the 8- to 16-cell stages in sheep (Crosby et al., 1988) and cattle (Meirelles et al., 2004). These temporal differences correlate with the timing of other developmental events such as lineage specification, which generally occurs much earlier in rodents than in other mammals (reviewed by Piliszek and Madeja, 2018).

Fig. 1.

Consecutive stages of preimplantation development of mouse, human, rabbit, sheep and cattle embryos. Bright-field images of in vivo (mouse and rabbit) and in vitro (human, cattle and sheep) derived embryos. Note that the length of time between morphological events depicted here may vary between species. Schematic representation of lineage formation: epiblast, white; trophectoderm, grey; primitive endoderm, black. Pictures not to scale.

Fig. 1.

Consecutive stages of preimplantation development of mouse, human, rabbit, sheep and cattle embryos. Bright-field images of in vivo (mouse and rabbit) and in vitro (human, cattle and sheep) derived embryos. Note that the length of time between morphological events depicted here may vary between species. Schematic representation of lineage formation: epiblast, white; trophectoderm, grey; primitive endoderm, black. Pictures not to scale.

Compaction and polarisation

After several rounds of division, mammalian embryos undergo the first major morphological change: compaction (Fig. 1; Fig. 2) (Ducibella and Anderson, 1975). During compaction, loosely attached blastomeres transform into a compact, ball-like structure, in which individual cells cannot be easily distinguished; the embryo at this stage is referred to as a morula (Ducibella et al., 1975; Ducibella and Anderson, 1975). The timing of compaction differs among mammals, varying between the 8-cell stage in mouse and up to the 32-cell stage in rabbit and cattle (Table 1) (Soom et al., 1997; Ziomek et al., 1990), whereas pig embryos initiate partial compaction at the 8-cell stage, with full compaction occurring only shortly before cavitation (Reima et al., 1993). Despite this relatively high variability in the timing of compaction among different mammalian species, the initiation of this event is tightly controlled, as premature compaction may lead to impaired developmental progress towards the blastocyst stage (Iwata et al., 2014). In mice, the timing of compaction can be modulated by treatment with conventional protein kinase C (PKC) activators or inhibitors (Bloom, 1989; Winkel et al., 1990). However, it is currently not clear which PKC subfamily is involved in this process or what influences the timing of PKC activity during the pre-implantation period (reviewed by Saini and Yamanaka, 2018).

Fig. 2.

Physical forces in pre-implantation development in mouse embryo. (A) The compaction process is driven by spatial rearrangement of E-cadherin, which in turn induces reorganisation of the acto-myosin cytoskeleton and increase of the cell-cell contact area. An ezrin-rich apical domain forms on the outside-facing surface. (B) During subsequent divisions, inside cells form either by asymmetric divisions, when the spindle is positioned perpendicular to the position of the apical domain (*), or due to internalisation of the apolar cells (**) because of the differences in contractility. Higher cortical tension in apolar cells leads to their internalisation (yellow arrow). Hippo pathway activity is restricted in polar cells due to the inheritance of the apical domain. In apolar cells the Hippo pathway is active. (C) A positive feedback loop between physical forces and tight junction maturation controls blastocyst cavity expansion. The increase in luminal pressure (blue arrows) leads to increased stiffness of the TE layer and stimulates tight junction maturation. This in turn strengthens TE integrity and allows for further cavity growth.

Fig. 2.

Physical forces in pre-implantation development in mouse embryo. (A) The compaction process is driven by spatial rearrangement of E-cadherin, which in turn induces reorganisation of the acto-myosin cytoskeleton and increase of the cell-cell contact area. An ezrin-rich apical domain forms on the outside-facing surface. (B) During subsequent divisions, inside cells form either by asymmetric divisions, when the spindle is positioned perpendicular to the position of the apical domain (*), or due to internalisation of the apolar cells (**) because of the differences in contractility. Higher cortical tension in apolar cells leads to their internalisation (yellow arrow). Hippo pathway activity is restricted in polar cells due to the inheritance of the apical domain. In apolar cells the Hippo pathway is active. (C) A positive feedback loop between physical forces and tight junction maturation controls blastocyst cavity expansion. The increase in luminal pressure (blue arrows) leads to increased stiffness of the TE layer and stimulates tight junction maturation. This in turn strengthens TE integrity and allows for further cavity growth.

Table 1.

Examples of developmental timing in different species of eutherian mammals

Examples of developmental timing in different species of eutherian mammals
Examples of developmental timing in different species of eutherian mammals

The initial model of compaction proposed that the whole process is driven by the Ca2+-dependent and cell-cell contact-positioned adhesion molecule E-cadherin (also known as uvomorulin, encoded by Cdh1) (Sefton et al., 1996, 1992; Vestweber et al., 1987). This hypothesis has been supported by the observation that culturing mouse embryos in calcium-free medium or in the presence of neutralising antibodies against E-cadherin in the medium prevents compaction, whereas placing already compacted embryos in such culture conditions results in complete de-compaction of the morula (Ducibella and Anderson, 1979, 1975; Shirayoshi et al., 1983). Similar observations have been made in human (De Paepe et al., 2013; Zakharova et al., 2014), pig (Matsunari et al., 2020) and hamster (Suzuki et al., 1999) embryos. Moreover, although zygotic Cdh1−/− mutant embryos are able to undergo compaction due to the presence of maternally deposited protein and maternal stores of Cdh1 mRNA (Larue et al., 1994), maternal-zygotic (MZ) Cdh1−/− mutants fail to compact (Stephenson et al., 2010; Vries et al., 2004).

Recent studies in mice have revealed that E-cadherin drives compaction by reorganising the acto-myosin cytoskeleton (Anani et al., 2014; Maître, 2017; Maître et al., 2015) (Fig. 2A). E-cadherin accumulation at cell-cell interfaces induces the redistribution of acto-myosin-dependent contractility away from the cell-cell contact and leads to an increase in contractility on contact-free surfaces. This process allows for the extension of cell-cell contacts, with a concomitant contraction of the contact-free surfaces, leading to the transformation of the loosely-attached blastomeres into the compact morula (Maître et al., 2016, 2015). Therefore, according to this model, compaction does not depend on the adhesion forces generated by the interlinked E-cadherin/β-catenin complex, but rather it is induced by a twofold increase in the tension at the cell-medium interface due to the intense remodelling of the cytoskeleton (Maître et al., 2016). Indeed, extensive remodelling of the actin cytoskeleton during compaction has also been reported in pigs (Albertini et al., 1987; Reima et al., 1993). The increase in cell-cell contact during compaction facilitates the formation of adherens and then tight junctions, and is accompanied by the de novo establishment of apico-basal polarity in early blastomeres (reviewed by Johnson and McConnell, 2004; Mihajlović and Bruce, 2017; Saini and Yamanaka, 2018).

The first sign of polarisation is the formation of the non-adhesive apical, ezrin-rich, microvilli domain that spans the contact-free, outside environment facing the surface of the blastomeres in mouse (Louvet et al., 1996; Louvet-Vallée et al., 2001; Reeve and Ziomek, 1981), human (Nikas et al., 1996) and rabbit (Ziomek et al., 1990). Ezrin is an ERM (ezrin/radixin/moesin) family protein that links the actin cytoskeleton and plasma membrane proteins (Gautreau et al., 1999), and is known to localise to microvilli in epithelial cells as well as early blastomeres (Bretscher, 1983; Dard et al., 2004; Hanzel et al., 1991; Louvet-Vallée et al., 2001). In mouse, known polarity proteins such as atypical protein kinase C (aPKC) (Eckert et al., 2004) and the partitioning defective (PAR) proteins PAR3 (Plusa et al., 2005; Vinot et al., 2005) and PARD6B (Alarcon, 2010) also localise to the apical domain (the outward-facing blastomere surface), whereas PAR1 and scribble are positioned basolaterally, i.e. in the cell-cell contact-facing area (Kono et al., 2014; Mihajlović and Bruce, 2016; reviewed by Yamanaka et al., 2006).

Although compaction and polarisation are initiated at similar developmental stages, often involve overlapping sets of proteins (Dard et al., 2009a; Fleming et al., 1986; Johnson and Maro, 1984; Pratt et al., 1982), and can occur in the complete absence of mRNA and protein synthesis (Kidder and McLachlin, 1985; Levy et al., 1986), they do not rely on each other and can occur independently (Dard et al., 2009a; Hirate et al., 2013; Pratt et al., 1982; Stephenson et al., 2010). After the initial compaction and polarisation, during each subsequent cell division blastomeres de-compact and subsequently re-compact after all cell divisions are completed until the formation of the blastocyst in cattle (Betteridge and Fléchon, 1988; Ducibella et al., 1977), pig (Reima et al., 1993) and mouse (Skrzecz and Karasiewicz, 1987; Watanabe et al., 2014). Compaction and polarisation are the first steps in blastomere differentiation, and they initiate the cascade that leads to lineage specification in the mammalian embryo. These processes also pave the way for the next morphological event in development, the formation of the blastocyst (reviewed by Chazaud and Yamanaka, 2016).

Cavitation: formation of the blastocyst

Polarised blastomeres can divide in a way that leads to both blastomeres inheriting the ezrin-rich apical domain and maintaining an outside position (symmetric division; Fig. 2B). Alternatively, if the division plane is more parallel to the outside-facing surface, asymmetric division can result in only one of the cells inheriting the apical domain (remaining polarised and maintaining contact with the outside environment), whereas the other, apolar, blastomere will be positioned inside the embryo (Johnson et al., 1988; Johnson and Ziomek, 1981a,b). Polarised blastomeres envelop the non-polar cells, gradually forming a fluid-tight seal and gaining an epithelial character, as reported in mouse (Anani et al., 2014; Ducibella et al., 1975; Johnson and Ziomek, 1983), human (Gualtieri et al., 1992; Nikas et al., 1996), pig (Reima et al., 1993) and cattle (Barcroft et al., 2003; Manejwala et al., 1989; Manejwala and Schultz, 1989; Van Soom et al., 1997a,b).

Maturation of the outside cells towards a fully functional TE epithelium requires the activity of Na+/K+ ATPases and aquaporins, which are important for the active transport of ions into the intracellular space and for the passive transport of water along the osmotic gradient (Barcroft et al., 2003; Manejwala et al., 1989; Manejwala and Schultz, 1989); reviewed by Marikawa and Alarcon, 2012). The maturation of gap and tight junctions ensures that fluid stays in the area surrounded by the mature epithelium, as reported in mouse (Eckert and Fleming, 2008; Moriwaki et al., 2007), human (Bloor et al., 2004; Gualtieri et al., 1992) and cattle (Barcroft et al., 1998; Moriwaki et al., 2007). In addition, recently it has been proposed that expansion of actin rings, called ‘actin ring zippering’, plays an important role in the formation of a water tight seal (Zenker et al., 2018). These processes culminate in the formation and expansion of the fluid-filled cavity, and the formation of the blastocyst, in a process known as cavitation (Fig. 1; Fig. 2C) (Smith and McLaren, 1977; reviewed by Johnson and McConnell, 2004). Mutation in zonula occludens (ZO-1; also known as TJP1), a tight junction protein, prevents cavity formation in the mouse (Wang et al., 2008). The expanding cavity forms adjacent to the ICM and the resulting asymmetry defines the first axis in mammalian development: the embryonic-abembryonic axis (reviewed by Rossant and Tam, 2009).

Specification of ICM lineages

In the next phase of pre-implantation development, the ICM cells differentiate into Epi and PrE progenitors, which are first distributed in a seemingly random ‘salt-and-pepper’ pattern within the ICM, but then gradually segregate into two distinct layers (Chazaud et al., 2006; Plusa et al., 2008; Saiz et al., 2016). The PrE forms a second epithelium that separates the blastocyst cavity from the most inner-positioned Epi in mouse (Dickson, 1966; Gardner and Rossant, 1979) and human (Blakeley et al., 2015). The correct specification and positioning of all three lineages is crucial for successful implantation and further development in utero, whereas the number of cells at the time of blastocyst formation and the timing of implantation vary substantially among different mammalian species (Table 1; reviewed by Madeja et al., 2019).

As we have discussed, mammalian embryos pass through similar developmental stages and all form a blastocyst composed of the Epi, TE and PrE. Therefore, one could assume that the basic molecular mechanisms of early lineage specification are shared among different mammalian groups. However, several studies have revealed differences in the expression patterns of some lineage markers, as well as differences in the responses to modulation of various signalling pathways among different groups of mammals (Fig. 3). Despite these differences, mammals appear to share several general principles that govern lineage formation in pre-implantation development, on which we focus in this Review.

Fig. 3.

Schematicshowinglineage-specific transcription factor localization in mouse, human and cattle embryos at morula, early blastocyst and expanded blastocyst stage. CDX2 (green); OCT4 (red); NANOG (yellow); GATA6 (purple); SOX17 (blue).

Fig. 3.

Schematicshowinglineage-specific transcription factor localization in mouse, human and cattle embryos at morula, early blastocyst and expanded blastocyst stage. CDX2 (green); OCT4 (red); NANOG (yellow); GATA6 (purple); SOX17 (blue).

In all the studied mammalian embryos, lineage specification is preceded by a period of high molecular heterogeneity between cells, when transcription factors that drive the specification of different lineages display overlapping, and often highly variable, expression at both the mRNA and protein levels (Fig. 3; reviewed by Martinez-Arias et al., 2013). Indeed, this is the case in mouse (Dietrich and Hiiragi, 2007; Guo et al., 2010; Plusa et al., 2008), human (Blakeley et al., 2015; Niakan and Eggan, 2013; Petropoulos et al., 2016), non-human primates (Boroviak et al., 2015), pig (Ramos-Ibeas et al., 2019) and rabbit embryos (Piliszek et al., 2017b). Subsequent refinement of these expression patterns leads to the segregation of the TE and ICM transcriptional profiles, which is followed by the establishment of separate PrE and Epi lineages (Chazaud et al., 2006; Guo et al., 2010; Plusa et al., 2008; Saiz et al., 2016). The timing of the lineage separation events differs among mammals and, although in most mammals the specification of the TE versus the ICM appears to precede PrE versus Epi formation, in humans it has been reported that all lineages are specified simultaneously rather than consecutively (Petropoulos et al., 2016).

Trophectoderm versus inner cell mass specification

Interestingly, compaction and cavitation appear not to be prerequisites for the initiation of TE- and ICM-specific transcriptional programmes (Box 1). Recent studies have shown that the Hippo signalling pathway is responsible for interpreting polarisation signals in the mouse morula and for initiating the TE-specific programme in the outside cells (Anani et al., 2014; Hirate et al., 2012; Nishioka et al., 2009, 2008; Yagi et al., 2007). In the polarised (outside) cells, the junction-associated angiomotin proteins AMOT and AMOTL2 are sequestered from the cell junction to the apical domain, resulting in inactivation of the Hippo pathway (Hirate et al., 2013; Nishioka et al., 2008; Sasaki, 2017) and activation of TE-specific transcription factors, such as caudal type homeobox 2 (CDX2) and GATA binding protein 3 (GATA3) (Nishioka et al., 2009, 2008; Ralston et al., 2010; Yagi et al., 2007). In non-polarised (inside) cells, adherens junction-bound AMOT activates the Hippo pathway kinases LATS1/2, repressing the TE differentiation programme (Hirate et al., 2013) (Fig. 2B). The activated Hippo pathway also restricts SOX2 to the inside cells, independently of CDX2 activity (Frum et al., 2018), and blocks premature onset of pluripotency (Frum et al., 2019). Although the specific activity of the Hippo pathway in TE differentiation has only been shown in mouse embryos so far, it is likely to be a more general mechanism of TE differentiation, because the pathway activity is evolutionarily conserved. Indeed, modulation of Hippo signalling has been suggested as a key mechanism for the evolution of blastocyst formation in ancestral mammals (Frankenberg, 2018). The Hippo pathway components TEA domain transcription factor 4 (TEAD4) and yes associated protein (YAP) are found in the morula- and blastocyst-stage embryos of other mammals, such as pigs (Emura et al., 2019, 2016), cattle (Fujii et al., 2010; Ozawa et al., 2012; Sakurai et al., 2017) and horse (Iqbal et al., 2014). Furthermore, a recent study has shown that downregulation of TEAD4 inhibits pig embryo development beyond the morula stage (Emura et al., 2019).

Box 1. Compaction and cavitation appear not to be prerequisites for the initiation of TE- and ICM-specific transcriptional programmes

Maternal-zygotic E-cadherin mouse mutant embryos (MZ Cdh1−/−) fail to compact and form a blastocyst, yet single blastomeres refine their expression pattern, with most of the blastomeres expressing CDX2 and some expressing OCT4. Outside blastomeres in such mutants cannot form an epithelium despite the existence of some residual cell adhesion, which could permit tight junction formation. Moreover, Cdh1−/− blastomeres express CDX2 regardless of their position within the embryo. Similarly, OCT4 expression in such cells is not position-dependent (Stephenson et al., 2010). Moreover, despite the lack of extensive cell-cell contact, blastomeres in MZ Cdh1−/− embryos can polarise and form an apical domain, which in turn leads to activation of CDX2 expression.

WNT signalling might also play a conserved role in the early development of non-murine mammals. Several studies have pointed to the role of WNT signalling in TE versus ICM differentiation. Specifically, WNT signalling activation leads to upregulation of several pluripotency-associated genes in cattle (Madeja et al., 2015), and concomitantly reduces the number of TE cells in both cattle (Denicol et al., 2014, 2013) and pig (Lim et al., 2013) embryos. Accordingly, WNT inhibition increases TE cell number in pig and cattle embryos. In human embryos, WNT activation promotes the expression of the TE-associated marker T-box gene eomesodermin (EOMES) (Krivega et al., 2015). Interestingly, the activation of Wnt/β-catenin signalling and inhibition of ERK signalling has a positive effect on blastocyst development in vitro, i.e. it increases survival, hatching and cell number in cattle (Harris et al., 2013) and pig (Kwon et al., 2019). In addition, Wnt/β-catenin signalling and inhibition of ERK signalling improves stem cell line derivation and maintenance of cattle trophoblast stem cells (Wang et al., 2019) and pluripotent mouse embryonic stem cells (Nichols et al., 2009).

Several transcription factors originally identified as lineage-specific in mouse are also present in other mammalian species (Table 2). For example, CDX2 is restricted to the TE after blastocyst formation in mouse (Blij et al., 2012; Strumpf et al., 2005), human (Niakan and Eggan, 2013) and cattle (Goissis and Cibelli, 2014b). Although its function is not necessary for the initial steps of TE specification, mouse Cdx2 mutant blastocysts collapse soon after cavity formation because CDX2 is crucial for TE maintenance (Blij et al., 2012; Strumpf et al., 2005). TE-specific expression of CDX2 has been confirmed in human, pig, cattle, rabbit, dog and sheep embryos (Table 2); however, in most of these species, there is no evidence of CDX2 expression before cavitation. Conversely, in mouse (Dietrich and Hiiragi, 2007; Strumpf et al., 2005) and cattle morulae (Madeja et al., 2013), Cdx2 expression is detected as early as the 8-cell stage. A recent analysis of a TE-specific enhancer in mouse embryos indicates a specific role for Notch signalling in early Cdx2 expression, which is distinct from later Hippo-dependent activation (Menchero et al., 2019; Rayon et al., 2014). It is plausible that the Notch-dependent Cdx2 expression that might occur in some species has been lost in others, for example where rapid TE specification is not necessary. Species-specific enhancer activity is strongly supported by studies in cattle where, in contrast to mouse, the POU5F1 locus lacks a specific cis-acting regulatory region that is necessary to suppress TE-specific transcription (Berg et al., 2011).

Table 2.

Expression of lineage-specific markers in different mammalian species

Expression of lineage-specific markers in different mammalian species
Expression of lineage-specific markers in different mammalian species

Primitive endoderm versus epiblast specification

Of the mammals that have been studied so far, PrE versus Epi specification appears to depend on both stabilisation of the pluripotency network in Epi cells and activation of the endodermal programme downstream of GATA6 in the PrE (Table 2). A pluripotency network that consists of octamer-binding transcription factor 4 (OCT4; encoded by Pou5f1) (Frum et al., 2013; Nichols et al., 1998), SRY-related HMG box-containing transcription factor 2 (SOX2) (Avilion et al., 2003; Wicklow et al., 2014), and homeodomain-containing protein NANOG (Chambers et al., 2003; Mitsui et al., 2003) appears to be conserved among different mammals (Table 2; Fig. 3) (reviewed by Artus et al., 2020; Piliszek and Madeja, 2018). In the mouse, PrE versus Epi specification depends on FGF/ERK signalling; thus various methods of FGF or ERK pathway inhibition result in ICM cells adopting an Epi fate at the expense of PrE (Krawchuk et al., 2013; Nichols et al., 2009; Thamodaran and Bruce, 2016; Yamanaka et al., 2010). Conversely, overabundance of FGF2/4 results in biasing the specification of ICM cells towards PrE (Arman et al., 1998; Frankenberg et al., 2011; Kang et al., 2013; Krawchuk et al., 2013; Kuijk et al., 2012; Nichols et al., 2009; Roode et al., 2012; Yamanaka et al., 2010). FGF signalling has also been extensively studied in non-murine mammals, but its role is less clear. In human embryos, ERK signalling inhibition does not interfere with PrE-specific GATA6 or GATA4 expression in the ICM; therefore, unlike in the mouse embryo, it is not necessary for PrE specification (Kuijk et al., 2012; Roode et al., 2012). A similar treatment of cattle or pig embryos also does not eliminate GATA6 or GATA4-positive cells; however, it increases the proportion of NANOG-positive cells within the ICM, indicating that ERK signalling has a role in cattle and pig lineage specification (Kuijk et al., 2012; Rodríguez et al., 2012). In rabbit embryos, ERK signalling inhibition fully blocks formation of SOX17-positive PrE cells, but at the same time does not expand the Epi population within the ICM, suggesting that in some species pluripotency acquisition may require additional stimulus (Piliszek et al., 2017b). Increased levels of FGF in cattle embryos in in vitro culture expand the PrE compartment at the expense of the Epi, similar to what occurs in mouse (Kuijk et al., 2012). Similar results of FGF2/4 treatment have been observed in sheep (Moradi et al., 2015), rabbit (Piliszek et al., 2017b) and pig (Rodríguez et al., 2012) embryos. Interestingly, inhibition of FGF receptor signalling by small molecule inhibitors has no effect on human, cattle, sheep or pig lineage specification, suggesting differences in receptor specificity or different pathway components. In addition, the WNT pathway might play a role in PrE versus Epi specification in non-human primates, as WNT inhibition is able to increase NANOG expression in the marmoset Epi (Boroviak et al., 2015).

The specification of Epi and PrE progenitors within the ICM is asynchronous, and initially the ICM is a mixture of committed PrE and Epi progenitors, and non-committed double-positive cells, as evidenced in mouse (GATA6+/NANOG+ cells; Grabarek et al., 2012; Saiz et al., 2020 preprint, 2016), non-human primates (SOX17+/NANOG+; Boroviak et al., 2015), pig (SOX17+/NANOG+; Ramos-Ibeas et al., 2019) and rabbit (SOX2+/SOX17+ cells; Piliszek et al., 2017b). A combined mathematical modelling and experimental data analysis approach in mouse revealed that specification of a subset of Epi progenitors precedes specification of PrE progenitors, as the induction of the PrE fate depends on FGF4 production by Epi cells (Bessonnard et al., 2014; Grabarek et al., 2012; Kang et al., 2013; Krawchuk et al., 2013; Ohnishi et al., 2014; Saiz et al., 2016).

The cell sorting process is initiated concomitantly with PrE and Epi progenitor specification and the formation of the salt-and-pepper pattern, as shown in mouse (Chazaud et al., 2006; Gerbe et al., 2008; Plusa et al., 2008; Saiz et al., 2013) and rabbit (Piliszek et al., 2017b). Originally, differences in cell adhesion were proposed to be responsible for the Epi and PrE cell sorting process (Chazaud et al., 2006; Rossant et al., 2003); however, recent studies have found no evidence for the involvement of E-cadherin-mediated cell adhesion in the cell sorting process in mouse embryos (Filimonow et al., 2019). Whether other adhesion molecules are involved remains to be determined. On the other hand, it has been demonstrated that cell sorting can be largely explained by a selective mechanism involving cell displacement during cell division and cavity expansion, aPKC-dependent immobilisation of the PrE progenitors once they reach the blastocyst cavity, and selective apoptosis of mispositioned cells (Meilhac et al., 2009; Morris et al., 2010; Plusa et al., 2008; Saiz et al., 2013; Xenopoulos et al., 2015). Interestingly, in mouse embryos, the Hippo pathway has been implicated in the selective elimination of unspecified cells with low levels of pluripotency factors via a cell competition mechanism (Hashimoto and Sasaki, 2019).

The cell microenvironment is constantly changing during development because of cell divisions and morphogenic events such as compaction, cavitation and lineage segregation. Existence of the extensive crosstalk between biochemical and mechanical inputs during pre-implantation development is now well established (Maître, 2017; Plusa and Hadjantonakis, 2016).

The specification of the outside and inside cells in mouse embryos is one of the best-studied examples of how a redistribution of physical forces can influence the lineage specification process. Originally, it was proposed that the differences in cell position within the embryo determined future cell fate. Outside cells contribute to the TE, whereas inside cells contribute to the ICM – the so-called ‘inside-outside’ model (Tarkowski and Wróblewska, 1967).

The most recent theory of TE versus ICM specification combines the elements of the inside-outside model and changes in mechanical force distribution with polarity and Hippo pathway activity. In this model, the formation of the low-contractility apical domain during compaction and polarisation in early embryos creates differences in force distribution that influence cell position during subsequent divisions (Fig. 2B) (Anani et al., 2014; Maître et al., 2016; Niwayama et al., 2019). As a result, the cells that inherit the apical domain are less contractile than their non-polar counterparts, envelop more contractile cells and differentiate into the TE (Anani et al., 2014; Maître et al., 2016). Highly contractile non-polar cells compete for the inside position and become internalised due to the angle of division via asymmetric divisions (Fig. 2B; Johnson and Ziomek, 1981a; Watanabe et al., 2014) or become internalised shortly after division, giving rise to ICM cells – as is most frequently the case (Anani et al., 2014; Korotkevich et al., 2017; Watanabe et al., 2014). In mice, the apical domain plays an important role in orienting cell division (Anani et al., 2014; Dard et al., 2009a,b) via the recruitment of one of the poles of the mitotic spindle (Korotkevich et al., 2017). The presence of the polarised apical domain inhibits Hippo pathway activity in the outside polar cells (Fig. 2B) and allows for the initiation of expression of TE-specific genes, such as Cdx2 and Gata3 (Hirate et al., 2013; Nishioka et al., 2009, 2008; Ralston et al., 2010; Rayon et al., 2014; Yagi et al., 2007). In contrast, inside blastomeres with an activated Hippo pathway retain Pou5f1, Sox2 and Gata6 expression and downregulate Cdx2 (Frum et al., 2013; Schrode et al., 2014; Wicklow et al., 2014). Differences in contractility also facilitate cell sorting: if an apolar cell is positioned outside owing to an oblique division plane or is placed there by mechanical manipulation, it will ‘sink’ inside the embryo (Maître et al., 2016). Alternatively, in the rare case when such a cell inherits only a small portion of the apical domain, it can extend the polarity domain and contribute to TE (Anani et al., 2014). This elegant mechanism explains how morphogenic events during pre-implantation development are closely linked to changes in the forces and mechanical properties of the embryonic cells. Considering that all mammalian embryos appear to have the ability to reconstitute the whole embryo after disaggregation, and that different mammalian embryos can be reaggregated together, a model in which cells use the information provided by polarity and cell-cell contact to recognise their new position and adjust their developmental path accordingly is probably common to most, if not all, eutherian mammals.

The influence of mechanical forces in early development is not restricted to specification of the TE and ICM. For example, it has been reported that hydraulic fracturing of the cell-cell contacts and uneven cortical tension between TE and ICM cells assure that blastocyst cavity forms on the TE and ICM interface (Dumortier et al., 2019) whereas increasing luminal pressure during cavity growth contributes to maturation of TE tight junctions (Box 2; Fig. 2C).

Box 2. Mechanical forces participate in blastocyst maturation

It has been recently demonstrated that increased luminal pressure related to blastocyst cavity growth result in increased cortical tension and stiffness of the TE layer surrounding the cavity (Fig. 2C) (Chan et al., 2019). The cellular response to the increased cortical tension (mediated by vinculin mechano-sensing) leads to rapid maturation of tight junctions, and this in turn strengthens the integrity of the epithelium and allows for further cavity growth, creating a positive feedback loop. Moreover, TE cells under increased pressure flatten, which ensures that there are no asymmetric divisions or additional ICM cells formed after the blastocyst cavity reaches a certain size (Chan et al., 2019). During cavity growth, cell-cell adhesion cannot be sustained above a critical threshold during subsequent TE cell divisions. This phenomenon leads to a series of TE ‘collapse and re-expansion’ cycles that have been linked to mitotic rounding during TE cell division (Chan et al., 2019; Leonavicius et al., 2018; Niimura, 2003). In addition, Chan and colleagues suggested a tissue size control mechanism in which hydraulically-generated tissue-yield stress may set up a threshold size for the mature blastocyst. Importantly, the increase in hydrostatic pressure within the blastocyst cavity has been implicated in the blastocyst hatching from the zona pellucida – a process that needs to occur in order to allow contact between maternal and embryonic tissues during implantation (Leonavicius et al., 2018).

Mammalian blastomeres retain high developmental potency for most of the pre-implantation period. Totipotency is commonly defined as the ability of the cell to give rise to the whole organism and, by definition, totipotency is a key feature of the mammalian zygote (Condic, 2014). It has been shown that even following one or two cleavage divisions of the zygote, the resulting individual cells can retain totipotency and can give rise to the whole organism. Isolated 2-cell stage mouse and rat blastomeres (1/2 blastomeres) are able to produce healthy offspring after transplantation to a foster mother (Tarkowski, 1959). In larger animals, the same is true for 1/4 blastomeres and even 1/8 blastomeres. In cattle, a live birth of genetically identical quadruplet calves has been reported after the transfer of separated 1/4 blastomeres originating from the same embryo (Johnson et al., 1995). In rabbit (Moore et al., 1968), sheep (Willadsen, 1981) and horse (Allen and Pashen, 1984), the transfer of embryos originating from single isolated 1/8 blastomeres can also result in live births. Similarly, individual human 1/4 blastomeres are able to form blastocysts in vitro (Van de Velde et al., 2008), and the birth of a child after transfer of an embryo with only one viable blastomere at the 4-cell stage has been reported. This confirms that in humans, at least one blastomere at the 4-cell stage can support development until birth, rendering it totipotent (Veiga et al., 1987). It is important to note that the prolonged totipotency period of larger mammalian species might be a result of larger embryo volume, longer pre-implantation period and higher cell number around time of lineage commitment but not greater plasticity per se. This concept is supported by the early observation that isolated 1/4 and 1/8 mouse blastomeres can form blastocyst-like structures in vitro (Tarkowski and Wróblewska, 1967) and, in some cases, can initiate implantation in vivo, but are not able to support development to term (Rossant, 1976). Moreover, single isolated 1/16 blastomeres (either outside or inside cells) can give rise to healthy mice when combined with tetraploid carrier blastomeres (that normally do not contribute to the embryonic lineages), suggesting that even 16-cell stage blastomeres might be totipotent (Tarkowski et al., 2010).

Totipotency can be also defined as the ability of a cell to contribute its progeny to all embryonic and extra-embryonic cell types. According to such a definition, mammalian blastomeres retain totipotency even after the process of cell lineage specification is initiated. It should be noted that the initiation of lineage differentiation does not necessarily equal or coincide with the final specification of cell fate. Shortly after blastocyst formation, mouse ICM cells retain the ability to re-form the TE layer, if the nascent TE is removed by immunosurgery, but this potential is quickly lost after cavitation (Gardner and Johnson, 1972; Posfai et al., 2017; Rossant, 1975a,b; Suwińska et al., 2008; reviewed by Klimczewska et al., 2018). This period of full ICM potency can be prolonged until the 64- to 128-cell stage in mouse when embryos are cultured in ERK inhibitors from the morula stage, suggesting that ERK activity is not only necessary for PrE formation, but it also plays a role in final TE and ICM specification (Wigger et al., 2017). A significantly extended period of ICM potency has been observed in non-murine mammals such as in cattle, where blastocysts regenerated from isolated ICMs and transferred to the recipient mother can result in live births (Kohri et al., 2019).

The plasticity of isolated embryonic cells can be also effectively tested by transplantation into different embryonic environments, i.e. by chimera complementation assay (Tam and Rossant, 2003). In this assay, single TE or ICM cells are transplanted into an early embryo (for example at the 8-cell or 16-cell stage) and the contribution of the progeny of the transplanted cells into different lineages can be assessed. Studies on human (De Paepe et al., 2013) and cattle (Berg et al., 2011) embryos have shown that a single cell isolated from a morphologically well-distinguished TE is able to contribute to both the TE and ICM after transplantation into morulae indicating that, even at that stage, these cells are not fully committed and retain a certain degree of plasticity. The length of the high plasticity period may be partially related to the timing of implantation. Mouse embryos implant after reaching about 128 cells, ∼4.5 days post coitum (dpc), whereas human and cattle embryos implant much later (7 and 22 dpc, respectively) (Bazer et al., 2009). This difference may explain why mouse TE cells isolated at the late 32-cell or 64-cell stage are unable to contribute to the ICM lineage if introduced into a morula-stage embryo (Posfai et al., 2017), whereas in human and cattle embryos TE cells from later stages can still contribute to all lineages (Berg et al., 2011; De Paepe et al., 2013). Even though TE versus ICM fate is determined early in mouse development, single ICM cells exhibit prolonged plasticity: following experimental transplantation to 8-cell embryos, they are able to contribute to all three cell lineages in the blastocyst (Grabarek et al., 2012), and embryos reconstructed entirely from isolated inner/ICM cells from 32-cell- or even 64-cell-stage embryos are able, in some cases, to re-form a whole blastocyst (Posfai et al., 2017; Suwińska et al., 2008).

The initial totipotency within the ICM must be lost eventually to allow the full specification of pluripotent Epi and extra-embryonic PrE cells (Boroviak et al., 2014). In the mouse embryo at the embryonic day (E)3.5 mid-blastocyst stage, during formation of the salt-and-pepper pattern, PrE and Epi precursors stained for respective lineage markers appear to be intermingled with some bipotent cells (Chazaud et al., 2006; Grabarek et al., 2012; Plusa et al., 2008; Saiz et al., 2016). The presence of bipotent cells that can still differentiate into the PrE or Epi lineage ensures the proper balance of cells in each lineage and explains why, even after both lineages start the segregation process, at least some of the ICM cells retain their ability to contribute to different lineages (Grabarek et al., 2012; Saiz et al., 2020 preprint, 2016).

The ability of mammalian embryos to self-organise does not depend on any pre-patterning or any particular cues originating from the oocyte, as cleavage-stage embryos can be disassociated into single blastomeres and then re-assembled again without any detrimental effect on further development, as demonstrated in mouse (Tarkowski, 1961), rat (Mayer and Fritz, 1974), sheep (Tucker et al., 1974), rabbit (Gardner and Munro, 1974), cattle (Brem et al., 1984), pig (Matsunari et al., 2013) and non-human primates (Tachibana et al., 2012). Moreover, the re-assembled embryo can contain blastomeres from several different embryos and even different developmental stages (Stern and Wilson, 1972). This remarkable feature crosses species boundaries, as shown by the aggregation of cleavage-stage embryos of different species, creating inter-specific chimeras between goat and sheep (Fehilly et al., 1984; Jaszczak et al., 1991; Meinecke-Tillmann and Meinecke, 1984), between Bos taurus and Bos indicus (Summers et al., 1983; Williams et al., 1990), and between mouse and rat (Bożyk et al., 2017) (reviewed by Zyzynska-Galenska et al., 2017).

How cell fate specification and lineage allocation occurs so robustly during pre-implantation development, in spite of high plasticity, is still debated. Although early blastomeres until the early 8-cell stage look seemingly identical, a few studies have reported that certain cleavage patterns between the 2- and 4-cell stage can produce ‘lineage-biased’ blastomeres, with a higher propensity to contribute to the TE or ICM lineage (Piotrowska et al., 2001; Piotrowska-Nitsche et al., 2005; Tabansky et al., 2013). Such variability in cell fate predispositions has been linked to differences in epigenetic modifications between blastomeres at the 4-cell stage (Torres-Padilla et al., 2007) and this in turn has been related to differences in the distribution of chromatin modifiers among early blastomeres (Burton et al., 2013; Wu et al., 2016). Subtle differences in the expression of certain key elements of the transcriptional machinery have been suggested to alter the lineage contribution probability of early blastomeres (Goolam et al., 2016; Shahbazi and Zernicka-Goetz, 2018). Moreover, differences in the levels of various lineage priming factors (such as transcription factors and polarity proteins) or in the different intracellular dynamics of pluripotency transcription factors, such as OCT4, NANOG and SOX2 (Burton et al., 2013; Frankenberg et al., 2011; Plachta et al., 2011; Plusa et al., 2005; Torres-Padilla et al., 2007; White et al., 2016), can lead to bias in the contribution of the progeny to a particular lineage. The authors who favour this hypothesis argue that in mammalian embryos, early cells are not homogenous/equivalent and that the compartmentalisation of subtle intracellular reactions is followed by amplification of the differences in subsequent development, leading eventually to the specification of different cell fates (Shahbazi and Zernicka-Goetz, 2018; Shi et al., 2015; White et al., 2018, 2017). Therefore, the gradual amplification of the original differences (Shahbazi and Zernicka-Goetz, 2018) or ‘partition errors’ (Shi et al., 2015) is at the heart of this model. However, the ‘cell-bias’ alone does not explain how separated blastomeres can re-assemble the whole embryo in a relatively short period of time or how cells recognise their new position and adjust their developmental paths. Moreover, it has also been suggested that developmental biases may originate from extrinsic constraints, such as the zona pellucida that encompasses the whole embryo (Kurotaki et al., 2007; Motosugi et al., 2005), and not necessarily be a manifestation of the differences in the intrinsic properties of the embryos.

In contrast to the ‘cell bias’ model, others have proposed that lineage allocation is stochastic and is achieved via cell-cell interactions and the existence of both positive and negative feedback loops, which integrate information about cell position with expression patterns (Dietrich and Hiiragi, 2007; Kurotaki et al., 2007; Motosugi et al., 2005). Moreover, the intra-embryonic variability of transcription factors and signalling molecules related to cell fate determination is well documented in the literature with regard to the pre-implantation development of several groups of mammals, including mouse (Dietrich and Hiiragi, 2007; Guo et al., 2010; Plusa et al., 2008), human (Blakeley et al., 2015) and rabbit (Piliszek et al., 2017b). In addition, single cell transcriptomic analyses have shown that, although slight differences in transcript abundance may be detectable between blastomeres of the early embryo, in most cases there is no consistent correlation – positive or negative – between lineage-specific transcription factors before the lineage specification process is initiated [mouse (Guo et al., 2010; Ohnishi et al., 2014), human (Blakeley et al., 2015; Boroviak et al., 2018; Petropoulos et al., 2016; Stirparo et al., 2018), non-human primates (Boroviak et al., 2018) and pig (Ramos-Ibeas et al., 2019)].

These observations favour the opinion that the observable heterogeneity of lineage-specific factors in either the TE versus ICM or the PrE versus Epi may be induced after the lineage specification process has begun, and is not inherited from previous events. In such a view, the initial variability in transcription levels is likely a representation of cell-unbound potential, not cell fate bias (Martinez-Arias et al., 2013; Dong and Liu, 2017).

Historically, most of our information about mammalian development originated from mouse studies, and for many years it was assumed that that knowledge could be easily applied to other mammalian species. Recently, however, an emerging body of evidence has pointed towards high variability in developmental control among different groups of mammals. Despite that variability, all eutherian mammals share several developmental features and all mammalian embryos seems to be able to assess and adjust to changes in embryo architecture, the feature closely linked to the way cells acquire and adjust their fate in unperturbed development.

All data collected thus far confirm that the observed variability in the intrinsic properties of early blastomeres is an important element of self-organisation in early mammals. Whatever their origin, these differences allow for the robust organisation of the early embryo, with an ‘assess and adjust’ mechanism that allows for the interpretation of the differences and correct positioning of the cells.

Many of the factors influencing cell behaviour and lineage contribution are interdependent, which leads to further complexity in the system. Moreover, cell micro-environment (outside or inside positions within the embryo, the shape and contractility of the neighbouring cells) have a relatively strong influence on cell fate at first; subsequently, the gradual establishment of the lineage-specific regulatory circuits increases the contribution of intrinsic factors to cell behaviour, whereas the contribution of extrinsic factors gradually weakens. Eventually, after cells become fully specified, the high levels of previous plasticity are lost. Studying different mammalian species is crucial to developing a deeper understanding of the processes governing pre-implantation development, including the acquisition and maintenance of pluripotency. This knowledge can then provide the foundation for further advances in pluripotent stem cell derivation, regenerative medicine, assisted reproduction techniques in humans and domestic animals, and conservation efforts for endangered species.

We thank the following researchers for providing images of embryos presented in Fig. 1: Ewelina Warzych-Plejer (cattle), Marta Czernik (sheep), Helen Hunter (human), Katarzyna Filimonow and Jessica Forsyth (mouse), and Anna Chołoniewska (rabbit). Images of IVF human embryos were collected under a licence granted by the Human Fertilisation and Embryology Authority and irreversibly anonymised before use. We thank Katarzyna Filimonow, Stephen Frankenberg and Nestor Saiz for comments on the manuscript, and Joanna Grabarek for help with Fig. 2.

Funding

The work of B.P. was funded by a Wellcome Trust Seed Award in Science (212372/Z/18/Z). The work of A.P. is funded by Narodowe Centrum Nauki (National Science Centre, Poland) grant number 2017/26/E/NZ3/01205.

Aiken
,
C. E. M.
,
Swoboda
,
P. P. L.
,
Skepper
,
J. N.
and
Johnson
,
M. H.
(
2004
).
The direct measurement of embryogenic volume and nucleo-cytoplasmic ratio during mouse pre-implantation development
.
Reprod. Camb. Engl.
128
,
527
-
535
.
Alarcon
,
V. B.
(
2010
).
Cell polarity regulator PARD6B is essential for trophectoderm formation in the preimplantation mouse embryo
.
Biol. Reprod
83
,
347
-
358
.
Albertini
,
D. F.
,
Overstrom
,
E. W.
and
Ebert
,
K. M.
(
1987
).
Changes in the organization of the actin cytoskeleton during preimplantation development of the pig embryo
.
Biol. Reprod
37
,
441
-
451
.
Allen
,
W. R.
and
Pashen
,
R. L.
(
1984
).
Production of monozygotic (identical) horse twins by embryo micromanipulation
.
J. Reprod. Fertil.
71
,
607
-
613
.
Anani
,
S.
,
Bhat
,
S.
,
Honma-Yamanaka
,
N.
,
Krawchuk
,
D.
and
Yamanaka
,
Y.
(
2014
).
Initiation of Hippo signaling is linked to polarity rather than to cell position in the pre-implantation mouse embryo
.
Dev. Camb. Engl.
141
,
2813
-
2824
.
Aplin
,
J. D.
and
Ruane
,
P. T.
(
2017
).
Embryo-epithelium interactions during implantation at a glance
.
J. Cell Sci.
130
,
15
-
22
.
Arman
,
E.
,
Haffner-Krausz
,
R.
,
Chen
,
Y.
,
Heath
,
J. K.
and
Lonai
,
P.
(
1998
).
Targeted disruption of fibroblast growth factor (FGF) receptor 2 suggests a role for FGF signaling in pregastrulation mammalian development
.
Proc. Natl. Acad. Sci. USA
95
,
5082
-
5087
.
Artus
,
J.
,
Piliszek
,
A.
and
Hadjantonakis
,
A.-K.
(
2011
).
The primitive endoderm lineage of the mouse blastocyst: sequential transcription factor activation and regulation of differentiation by Sox17
.
Dev. Biol.
350
,
393
-
404
.
Artus
,
J.
,
Hue
,
I.
and
Acloque
,
H.
(
2020
).
Preimplantation development in ungulates: a “ménage à quatre” scenario
.
Reprod. Camb. Engl.
159
,
R151
-
R172
.
Avilion
,
A. A.
,
Nicolis
,
S. K.
,
Pevny
,
L. H.
,
Perez
,
L.
,
Vivian
,
N.
and
Lovell-Badge
,
R.
(
2003
).
Multipotent cell lineages in early mouse development depend on SOX2 function
.
Genes Dev.
17
,
126
-
140
.
Barcroft
,
L. C.
,
Hay-Schmidt
,
A.
,
Caveney
,
A.
,
Gilfoyle
,
E.
,
Overstrom
,
E. W.
,
Hyttel
,
P.
and
Watson
,
A. J.
(
1998
).
Trophectoderm differentiation in the bovine embryo: characterization of a polarized epithelium
.
J. Reprod. Fertil.
114
,
327
-
339
.
Barcroft
,
L. C.
,
Offenberg
,
H.
,
Thomsen
,
P.
and
Watson
,
A. J.
(
2003
).
Aquaporin proteins in murine trophectoderm mediate transepithelial water movements during cavitation
.
Dev. Biol.
256
,
342
-
354
.
Bazer
,
F. W.
,
Spencer
,
T. E.
,
Johnson
,
G. A.
,
Burghardt
,
R. C.
and
Wu
,
G.
(
2009
).
Comparative aspects of implantation
.
Reproduction
138
,
195
-
209
.
Berg
,
D. K.
,
Smith
,
C. S.
,
Pearton
,
D. J.
,
Wells
,
D. N.
,
Broadhurst
,
R.
,
Donnison
,
M.
and
Pfeffer
,
P. L.
(
2011
).
Trophectoderm lineage determination in cattle
.
Dev. Cell
20
,
244
-
255
.
Bessonnard
,
S.
,
De Mot
,
L.
,
Gonze
,
D.
,
Barriol
,
M.
,
Dennis
,
C.
,
Goldbeter
,
A.
,
Dupont
,
G.
and
Chazaud
,
C.
(
2014
).
Gata6. Nanog and Erk signaling control cell fate in the inner cell mass through a tristable regulatory network
.
Dev. Camb. Engl.
141
,
3637
-
3648
.
Betteridge
,
K. J.
and
Fléchon
,
J.-E.
(
1988
).
The anatomy and physiology of pre-attachment bovine embryos. Theriogenology
.
Annual Conference of the International Embryo Transfer Society
29
,
155
-
187
.
Blakeley
,
P.
,
Fogarty
,
N. M. E.
,
del Valle
,
I.
,
Wamaitha
,
S. E.
,
Hu
,
T. X.
,
Elder
,
K.
,
Snell
,
P.
,
Christie
,
L.
,
Robson
,
P.
and
Niakan
,
K. K.
(
2015
).
Defining the three cell lineages of the human blastocyst by single-cell RNA-seq
.
Dev. Camb. Engl.
142
,
3151
-
3165
.
Blij
,
S.
,
Frum
,
T.
,
Akyol
,
A.
,
Fearon
,
E.
and
Ralston
,
A.
(
2012
).
Maternal Cdx2 is dispensable for mouse development
.
Development
139
,
3969
-
3972
.
Blomberg
,
L.
,
Hashizume
,
K.
and
Viebahn
,
C.
(
2008
).
Blastocyst elongation, trophoblastic differentiation, and embryonic pattern formation
.
Reprod. Camb. Engl.
135
,
181
-
195
.
Bloom
,
T. L.
(
1989
).
The effects of phorbol ester on mouse blastomeres: a role for protein kinase C in compaction?
Dev. Camb. Engl.
106
,
159
-
171
.
Bloor
,
D. J.
,
Wilson
,
Y.
,
Kibschull
,
M.
,
Traub
,
O.
,
Leese
,
H. J.
,
Winterhager
,
E.
and
Kimber
,
S. J.
(
2004
).
Expression of connexins in human preimplantation embryos in vitro
.
Reprod. Biol. Endocrinol.
2
,
25
.
Bontovics
,
B.
,
Maraghechi
,
P.
,
Lázár
,
B.
,
Anand
,
M.
,
Németh
,
K.
,
Fábián
,
R.
,
Vašíček
,
J.
,
Makarevich
,
A. V.
,
Gócza
,
E.
and
Chrenek
,
P.
(
2020
).
The effect of dual inhibition of Ras-MEK-ERK and GSK3β pathways on development of in vitro cultured rabbit embryos
.
Zygote Camb. Engl.
28
,
183
-
190
.
Boroviak
,
T.
,
Loos
,
R.
,
Bertone
,
P.
,
Smith
,
A.
and
Nichols
,
J.
(
2014
).
The ability of inner-cell-mass cells to self-renew as embryonic stem cells is acquired following epiblast specification
.
Nat. Cell Biol.
16
,
516
-
528
.
Boroviak
,
T.
,
Loos
,
R.
,
Lombard
,
P.
,
Okahara
,
J.
,
Behr
,
R.
,
Sasaki
,
E.
,
Nichols
,
J.
,
Smith
,
A.
and
Bertone
,
P.
(
2015
).
Lineage-specific profiling delineates the emergence and progression of naive pluripotency in mammalian embryogenesis
.
Dev. Cell
35
,
366
-
382
.
Boroviak
,
T.
,
Stirparo
,
G. G.
,
Dietmann
,
S.
,
Hernando-Herraez
,
I.
,
Mohammed
,
H.
,
Reik
,
W.
,
Smith
,
A.
,
Sasaki
,
E.
,
Nichols
,
J.
and
Bertone
,
P.
(
2018
).
Single cell transcriptome analysis of human, marmoset and mouse embryos reveals common and divergent features of preimplantation development
.
Dev. Camb. Engl.
145
,
dev167833
.
Bou
,
G.
,
Liu
,
S.
,
Sun
,
M.
,
Zhu
,
J.
,
Xue
,
B.
,
Guo
,
J.
,
Zhao
,
Y.
,
Qu
,
B.
,
Weng
,
X.
,
Wei
,
Y.
, et al. 
(
2017
).
CDX2 is essential for cell proliferation and polarity in porcine blastocysts
.
Dev. Camb. Engl.
144
,
1296
-
1306
.
Bożyk
,
K.
,
Gilecka
,
K.
,
Humięcka
,
M.
,
Szpila
,
M.
,
Suwińska
,
A.
and
Tarkowski
,
A. K.
(
2017
).
Mouse↔rat aggregation chimaeras can develop to adulthood
.
Dev. Biol.
427
,
106
-
120
.
Braude
,
P.
,
Bolton
,
V.
and
Moore
,
S.
(
1988
).
Human gene expression first occurs between the four- and eight-cell stages of preimplantation development
.
Nature
332
,
459
-
461
.
Brem
,
G.
,
Tenhumberg
,
H.
and
Kräußlich
,
H.
(
1984
).
Chimerism in cattle through microsurgical aggregation of morulae
.
Theriogenology
22
,
609
-
613
.
Bretscher
,
A.
(
1983
).
Purification of an 80,000-dalton protein that is a component of the isolated microvillus cytoskeleton, and its localization in nonmuscle cells
.
J. Cell Biol.
97
,
425
-
432
.
Burton
,
A.
,
Muller
,
J.
,
Tu
,
S.
,
Padilla-Longoria
,
P.
,
Guccione
,
E.
and
Torres-Padilla
,
M.-E.
(
2013
).
Single-cell profiling of epigenetic modifiers identifies PRDM14 as an inducer of cell fate in the mammalian embryo
.
Cell Rep.
5
,
687
-
701
.
Canizo
,
J. R.
,
Ynsaurralde Rivolta
,
A. E.
,
Vazquez Echegaray
,
C.
,
Suvá
,
M.
,
Alberio
,
V.
,
Aller
,
J. F.
,
Guberman
,
A. S.
,
Salamone
,
D. F.
,
Alberio
,
R. H.
and
Alberio
,
R.
(
2019
).
A dose-dependent response to MEK inhibition determines hypoblast fate in bovine embryos
.
BMC Dev. Biol.
19
,
13
.
Cao
,
S.
,
Han
,
J.
,
Wu
,
J.
,
Li
,
Q.
,
Liu
,
S.
,
Zhang
,
W.
,
Pei
,
Y.
,
Ruan
,
X.
,
Liu
,
Z.
,
Wang
,
X.
, et al. 
(
2014
).
Specific gene-regulation networks during the pre-implantation development of the pig embryo as revealed by deep sequencing
.
BMC Genomics
15
,
4
.
Cauffman
,
G.
,
De Rycke
,
M.
,
Sermon
,
K.
,
Liebaers
,
I.
and
Van de Velde
,
H.
(
2009
).
Markers that define stemness in ESC are unable to identify the totipotent cells in human preimplantation embryos
.
Hum. Reprod. Oxf. Engl.
24
,
63
-
70
.
Cetinkaya
,
M.
,
Pirkevi
,
C.
,
Yelke
,
H.
,
Colakoglu
,
Y. K.
,
Atayurt
,
Z.
and
Kahraman
,
S.
(
2015
).
Relative kinetic expressions defining cleavage synchronicity are better predictors of blastocyst formation and quality than absolute time points
.
J. Assist. Reprod. Genet.
32
,
27
-
35
.
Chambers
,
I.
,
Colby
,
D.
,
Robertson
,
M.
,
Nichols
,
J.
,
Lee
,
S.
,
Tweedie
,
S.
and
Smith
,
A.
(
2003
).
Functional expression cloning of Nanog, a pluripotency sustaining factor in embryonic stem cells
.
Cell
113
,
643
-
655
.
Chan
,
C. J.
,
Costanzo
,
M.
,
Ruiz-Herrero
,
T.
,
Mönke
,
G.
,
Petrie
,
R. J.
,
Bergert
,
M.
,
Diz-Muñoz
,
A.
,
Mahadevan
,
L.
and
Hiiragi
,
T.
(
2019
).
Hydraulic control of mammalian embryo size and cell fate
.
Nature
571
,
112
-
116
.
Chazaud
,
C.
and
Yamanaka
,
Y.
(
2016
).
Lineage specification in the mouse preimplantation embryo
.
Dev. Camb. Engl.
143
,
1063
-
1074
.
Chazaud
,
C.
,
Yamanaka
,
Y.
,
Pawson
,
T.
and
Rossant
,
J.
(
2006
).
Early lineage segregation between epiblast and primitive endoderm in mouse blastocysts through the Grb2-MAPK pathway
.
Dev. Cell
10
,
615
-
624
.
Chen
,
A. E.
,
Egli
,
D.
,
Niakan
,
K.
,
Deng
,
J.
,
Akutsu
,
H.
,
Yamaki
,
M.
,
Cowan
,
C.
,
Fitz-Gerald
,
C.
,
Zhang
,
K.
,
Melton
,
D. A.
, et al. 
(
2009
).
Optimal timing of inner cell mass isolation increases the efficiency of human embryonic stem cell derivation and allows generation of sibling cell lines
.
Cell Stem Cell
4
,
103
-
106
.
Condic
,
M. L.
(
2014
).
Totipotency: what it is and what it is not
.
Stem Cells Dev.
23
,
796
-
812
.
Crosby
,
I. M.
,
Gandolfi
,
F.
and
Moor
,
R. M.
(
1988
).
Control of protein synthesis during early cleavage of sheep embryos
.
J. Reprod. Fertil.
82
,
769
-
775
.
Dantzer
,
V.
(
1985
).
Electron microscopy of the initial stages of placentation in the pig
.
Anat. Embryol.
172
,
281
-
293
.
Dard
,
N.
,
Louvet-Vallée
,
S.
,
Santa-Maria
,
A.
and
Maro
,
B.
(
2004
).
Phosphorylation of ezrin on threonine T567 plays a crucial role during compaction in the mouse early embryo
.
Dev. Biol.
271
,
87
-
97
.
Dard
,
N.
,
Le
,
T.
,
Maro
,
B.
and
Louvet-Vallée
,
S.
(
2009a
).
Inactivation of aPKClambda reveals a context dependent allocation of cell lineages in preimplantation mouse embryos
.
PLoS ONE
4
,
e7117
.
Dard
,
N.
,
Louvet-Vallée
,
S.
and
Maro
,
B.
(
2009b
).
Orientation of mitotic spindles during the 8- to 16-cell stage transition in mouse embryos
.
PLoS ONE
4
,
e8171
.
De Paepe
,
C.
,
Cauffman
,
G.
,
Verloes
,
A.
,
Sterckx
,
J.
,
Devroey
,
P.
,
Tournaye
,
H.
,
Liebaers
,
I.
and
Van de Velde
,
H.
(
2013
).
Human trophectoderm cells are not yet committed
.
Hum. Reprod. Oxf. Engl.
28
,
740
-
749
.
Deglincerti
,
A.
,
Croft
,
G. F.
,
Pietila
,
L. N.
,
Zernicka-Goetz
,
M.
,
Siggia
,
E. D.
and
Brivanlou
,
A. H.
(
2016
).
Self-organization of the in vitro attached human embryo
.
Nature
533
,
251
-
254
.
Denicol
,
A. C.
,
Dobbs
,
K. B.
,
McLean
,
K. M.
,
Carambula
,
S. F.
,
Loureiro
,
B.
and
Hansen
,
P. J.
(
2013
).
Canonical WNT signaling regulates development of bovine embryos to the blastocyst stage
.
Sci. Rep.
3
,
1266
.
Denicol
,
A. C.
,
Block
,
J.
,
Kelley
,
D. E.
,
Pohler
,
K. G.
,
Dobbs
,
K. B.
,
Mortensen
,
C. J.
,
Ortega
,
M. S.
and
Hansen
,
P. J.
(
2014
).
The WNT signaling antagonist Dickkopf-1 directs lineage commitment and promotes survival of the preimplantation embryo
.
FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol.
28
,
3975
-
3986
.
Dickson
,
A. D.
(
1966
).
The form of the mouse blastocyst
.
J. Anat.
100
,
335
-
348
.
Dietrich
,
J.-E.
and
Hiiragi
,
T.
(
2007
).
Stochastic patterning in the mouse pre-implantation embryo
.
Dev. Camb. Engl.
134
,
4219
-
4231
.
Dobson
,
A. T.
,
Raja
,
R.
,
Abeyta
,
M. J.
,
Taylor
,
T.
,
Shen
,
S.
,
Haqq
,
C.
and
Pera
,
R. A. R.
(
2004
).
The unique transcriptome through day 3 of human preimplantation development
.
Hum. Mol. Genet.
13
,
1461
-
1470
.
Dong
,
P.
and
Liu
,
Z.
(
2017
).
Shaping development by stochasticity and dynamics in gene regulation
.
Open Biol.
7
,
170030
.
Ducibella
,
T.
and
Anderson
,
E.
(
1975
).
Cell shape and membrane changes in the eight-cell mouse embryo: prerequisites for morphogenesis of the blastocyst
.
Dev. Biol.
47
,
45
-
58
.
Ducibella
,
T.
and
Anderson
,
E.
(
1979
).
The effects of calcium deficiency on the formation of the zonula occludens and blastocoel in the mouse embryo
.
Dev. Biol.
73
,
46
-
58
.
Ducibella
,
T.
,
Albertini
,
D. F.
,
Anderson
,
E.
and
Biggers
,
J. D.
(
1975
).
The preimplantation mammalian embryo: characterization of intercellular junctions and their appearance during development
.
Dev. Biol.
45
,
231
-
250
.
Ducibella
,
T.
,
Ukena
,
T.
,
Karnovsky
,
M.
and
Anderson
,
E.
(
1977
).
Changes in cell surface and cortical cytoplasmic organization during early embryogenesis in the preimplantation mouse embryo
.
J. Cell Biol.
74
,
153
-
167
.
Dumortier
,
J. G.
,
Le Verge-Serandour
,
M.
,
Tortorelli
,
A. F.
,
Mielke
,
A.
,
de Plater
,
L.
,
Turlier
,
H.
and
Maître
,
J.-L.
(
2019
).
Hydraulic fracturing and active coarsening position the lumen of the mouse blastocyst
.
Science
365
,
465
-
468
.
Eckert
,
J. J.
and
Fleming
,
T. P.
(
2008
).
Tight junction biogenesis during early development
.
Biochim. Biophys. Acta
1778
,
717
-
728
.
Eckert
,
J. J.
,
McCallum
,
A.
,
Mears
,
A.
,
Rumsby
,
M. G.
,
Cameron
,
I. T.
and
Fleming
,
T. P.
(
2004
).
Specific PKC isoforms regulate blastocoel formation during mouse preimplantation development
.
Dev. Biol.
274
,
384
-
401
.
Edwards
,
R. G.
,
Purdy
,
J. M.
,
Steptoe
,
P. C.
and
Walters
,
D. E.
(
1981
).
The growth of human preimplantation embryos in vitro
.
Am. J. Obstet. Gynecol.
141
,
408
-
416
.
Emura
,
N.
,
Sakurai
,
N.
,
Takahashi
,
K.
,
Hashizume
,
T.
and
Sawai
,
K.
(
2016
).
OCT-4 expression is essential for the segregation of trophectoderm lineages in porcine preimplantation embryos
.
J. Reprod. Dev.
62
,
401
-
408
.
Emura
,
N.
,
Takahashi
,
K.
,
Saito
,
Y.
and
Sawai
,
K.
(
2019
).
The necessity of TEAD4 for early development and gene expression involved in differentiation in porcine embryos
.
J. Reprod. Dev.
65
,
361
-
368
.
Enders
,
A. C.
and
Schlafke
,
S.
(
1971
).
Penetration of the uterine epithelium during implantation in the rabbit
.
Am. J. Anat.
132
,
219
-
230
.
Fehilly
,
C. B.
,
Willadsen
,
S. M.
and
Tucker
,
E. M.
(
1984
).
Interspecific chimaerism between sheep and goat
.
Nature
307
,
634
-
636
.
Filimonow
,
K.
,
Saiz
,
N.
,
Suwińska
,
A.
,
Wyszomirski
,
T.
,
Grabarek
,
J. B.
,
Ferretti
,
E.
,
Piliszek
,
A.
,
Plusa
,
B.
and
Maleszewski
,
M.
(
2019
).
No evidence of involvement of E-cadherin in cell fate specification or the segregation of Epi and PrE in mouse blastocysts
.
PLoS ONE
14
,
e0212109
.
Finn
,
C. A.
and
McLaren
,
A.
(
1967
).
A study of the early stages of implantation in mice
.
Reproduction
13
,
259
-
267
.
Fleming
,
T. P.
,
Cannon
,
P. M.
and
Pickering
,
S. J.
(
1986
).
The cytoskeleton, endocytosis and cell polarity in the mouse preimplantation embryo
.
Dev. Biol.
113
,
406
-
419
.
Frankenberg
,
S.
(
2018
).
Pre-gastrula Development of Non-eutherian Mammals
.
Curr. Top. Dev. Biol.
128
,
237
-
266
.
Frankenberg
,
S.
,
Gerbe
,
F.
,
Bessonnard
,
S.
,
Belville
,
C.
,
Pouchin
,
P.
,
Bardot
,
O.
and
Chazaud
,
C.
(
2011
).
Primitive endoderm differentiates via a three-step mechanism involving Nanog and RTK signaling
.
Dev. Cell
21
,
1005
-
1013
.
Frankenberg
,
S. R.
,
de Barros
,
F. R. O.
,
Rossant
,
J.
and
Renfree
,
M. B.
(
2016
).
The mammalian blastocyst
.
Wiley Interdiscip. Rev. Dev. Biol.
5
,
210
-
232
.
Frum
,
T.
,
Halbisen
,
M. A.
,
Wang
,
C.
,
Amiri
,
H.
,
Robson
,
P.
and
Ralston
,
A.
(
2013
).
Oct4 cell-autonomously promotes primitive endoderm development in the mouse blastocyst
.
Dev. Cell
25
,
610
-
622
.
Frum
,
T.
,
Murphy
,
T. M.
and
Ralston
,
A.
(
2018
).
HIPPO signaling resolves embryonic cell fate conflicts during establishment of pluripotency in vivo
.
eLife
7
,
e42298
.
Frum
,
T.
,
Watts
,
J. L.
and
Ralston
,
A.
(
2019
).
TEAD4, YAP1 and WWTR1 prevent the premature onset of pluripotency prior to the 16-cell stage
.
Dev. Camb. Engl.
146
,
dev179861
.
Fujii
,
T.
,
Moriyasu
,
S.
,
Hirayama
,
H.
,
Hashizume
,
T.
and
Sawai
,
K.
(
2010
).
Aberrant expression patterns of genes involved in segregation of inner cell mass and trophectoderm lineages in bovine embryos derived from somatic cell nuclear transfer
.
Cell. Reprogramming
12
,
617
-
625
.
Fujii
,
T.
,
Sakurai
,
N.
,
Osaki
,
T.
,
Iwagami
,
G.
,
Hirayama
,
H.
,
Minamihashi
,
A.
,
Hashizume
,
T.
and
Sawai
,
K.
(
2013
).
Changes in the expression patterns of the genes involved in the segregation and function of inner cell mass and trophectoderm lineages during porcine preimplantation development
.
J. Reprod. Dev.
59
,
151
-
158
.
Gamow
,
E.
and
Daniel
,
J. C.
(
1970
).
Fluid transport in the rabbit blastocyst
.
Wilhelm Roux Arch. Entwicklungsmechanik Org.
164
,
261
-
278
.
Gardner
,
R. L.
and
Johnson
,
M. H.
(
1972
).
An investigation of inner cell mass and trophoblast tissues following their isolation from the mouse blastocyst
.
J. Embryol. Exp. Morphol.
28
,
279
-
312
.
Gardner
,
R. L.
and
Munro
,
A. J.
(
1974
).
Successful construction of chimaeric rabbit
.
Nature
250
,
146
-
147
.
Gardner
,
R. L.
and
Rossant
,
J.
(
1979
).
Investigation of the fate of 4-5 day post-coitum mouse inner cell mass cells by blastocyst injection
.
J. Embryol. Exp. Morphol.
52
,
141
-
152
.
Gautreau
,
A.
,
Poullet
,
P.
,
Louvard
,
D.
and
Arpin
,
M.
(
1999
).
Ezrin, a plasma membrane–microfilament linker, signals cell survival through the phosphatidylinositol 3-kinase/Akt pathway
.
Proc. Natl. Acad. Sci. USA
96
,
7300
-
7305
.
Gerbe
,
F.
,
Cox
,
B.
,
Rossant
,
J.
and
Chazaud
,
C.
(
2008
).
Dynamic expression of Lrp2 pathway members reveals progressive epithelial differentiation of primitive endoderm in mouse blastocyst
.
Dev. Biol.
313
,
594
-
602
.
Goissis
,
M. D.
and
Cibelli
,
J. B.
(
2014a
).
Functional characterization of SOX2 in bovine preimplantation embryos
.
Biol. Reprod
90
,
30
.
Goissis
,
M. D.
and
Cibelli
,
J. B.
(
2014b
).
Functional characterization of CDX2 during bovine preimplantation development in vitro
.
Mol. Reprod. Dev.
81
,
962
-
970
.
Goolam
,
M.
,
Scialdone
,
A.
,
Graham
,
S. J. L.
,
Macaulay
,
I. C.
,
Jedrusik
,
A.
,
Hupalowska
,
A.
,
Voet
,
T.
,
Marioni
,
J. C.
and
Zernicka-Goetz
,
M.
(
2016
).
Heterogeneity in Oct4 and Sox2 targets biases cell fate in 4-cell mouse embryos
.
Cell
165
,
61
-
74
.
Grabarek
,
J. B.
,
Zyzyńska
,
K.
,
Saiz
,
N.
,
Piliszek
,
A.
,
Frankenberg
,
S.
,
Nichols
,
J.
,
Hadjantonakis
,
A.-K.
and
Plusa
,
B.
(
2012
).
Differential plasticity of epiblast and primitive endoderm precursors within the ICM of the early mouse embryo
.
Dev. Camb. Engl.
139
,
129
-
139
.
Gualtieri
,
R.
,
Santella
,
L.
and
Dale
,
B.
(
1992
).
Tight junctions and cavitation in the human pre-embryo
.
Mol. Reprod. Dev.
32
,
81
-
87
.
Guo
,
G.
,
Huss
,
M.
,
Tong
,
G. Q.
,
Wang
,
C.
,
Li Sun
,
L.
,
Clarke
,
N. D.
and
Robson
,
P.
(
2010
).
Resolution of cell fate decisions revealed by single-cell gene expression analysis from zygote to blastocyst
.
Dev. Cell
18
,
675
-
685
.
Hall
,
V. J.
,
Christensen
,
J.
,
Gao
,
Y.
,
Schmidt
,
M. H.
and
Hyttel
,
P.
(
2009
).
Porcine pluripotency cell signaling develops from the inner cell mass to the epiblast during early development
.
Dev. Dyn. Off. Publ. Am. Assoc. Anat.
238
,
2014
-
2024
.
Hanzel
,
D.
,
Reggio
,
H.
,
Bretscher
,
A.
,
Forte
,
J. G.
and
Mangeat
,
P.
(
1991
).
The secretion-stimulated 80K phosphoprotein of parietal cells is ezrin, and has properties of a membrane cytoskeletal linker in the induced apical microvilli
.
EMBO J.
10
,
2363
-
2373
.
Harris
,
D.
,
Huang
,
B.
and
Oback
,
B.
(
2013
).
Inhibition of MAP2K and GSK3 signaling promotes bovine blastocyst development and epiblast-associated expression of pluripotency factors
.
Biol. Reprod
88
.
Hashimoto
,
M.
and
Sasaki
,
H.
(
2019
).
Epiblast formation by TEAD-YAP-dependent expression of pluripotency factors and competitive elimination of unspecified cells
.
Dev. Cell
50
,
139
-
154.e5
.
Hemberger
,
M.
,
Hanna
,
C. W.
and
Dean
,
W.
(
2020
).
Mechanisms of early placental development in mouse and humans
.
Nat. Rev. Genet.
21
,
27
-
43
.
Hertig
,
A. T.
,
Rock
,
J
.,
Adams
,
E. C.
and
Mulligan
,
W. J.
(
1954
).
On the preimplantation stages of the human ovum: a description of four normal and four abnormal specimens ranging from the second to the fifth day of development
.
Contrib. Embryol.
,
35
,
199
-
220
.
Hertig
,
A. T.
,
Rock
,
J.
and
Adams
,
E. C.
(
1956
).
A description of 34 human ova within the first 17 days of development
.
Am. J. Anat.
98
,
435
-
493
.
Hirate
,
Y.
,
Cockburn
,
K.
,
Rossant
,
J.
and
Sasaki
,
H.
(
2012
).
Tead4 is constitutively nuclear, while nuclear vs. cytoplasmic Yap distribution is regulated in preimplantation mouse embryos
.
Proc. Natl. Acad. Sci. USA
109
,
E3389
-
E3390
.
; author reply E3391-3392
.
Hirate
,
Y.
,
Hirahara
,
S.
,
Inoue
,
K.-I.
,
Suzuki
,
A.
,
Alarcon
,
V. B.
,
Akimoto
,
K.
,
Hirai
,
T.
,
Hara
,
T.
,
Adachi
,
M.
,
Chida
,
K.
, et al. 
(
2013
).
Polarity-dependent distribution of angiomotin localizes Hippo signaling in preimplantation embryos
.
Curr. Biol. CB
23
,
1181
-
1194
.
Home
,
P.
,
Ray
,
S.
,
Dutta
,
D.
,
Bronshteyn
,
I.
,
Larson
,
M.
and
Paul
,
S.
(
2009
).
GATA3 is selectively expressed in the trophectoderm of peri-implantation embryo and directly regulates Cdx2 gene expression
.
J. Biol. Chem.
284
,
28729
-
28737
.
Hyslop
,
L.
,
Stojkovic
,
M.
,
Armstrong
,
L.
,
Walter
,
T.
,
Stojkovic
,
P.
,
Przyborski
,
S.
,
Herbert
,
M.
,
Murdoch
,
A.
,
Strachan
,
T.
and
Lako
,
M.
(
2005
).
Downregulation of NANOG induces differentiation of human embryonic stem cells to extraembryonic lineages
.
Stem Cells Dayt. Ohio
23
,
1035
-
1043
.
Iqbal
,
K.
,
Chitwood
,
J. L.
,
Meyers-Brown
,
G. A.
,
Roser
,
J. F.
and
Ross
,
P. J.
(
2014
).
RNA-seq transcriptome profiling of equine inner cell mass and trophectoderm
.
Biol. Reprod
90
,
61
.
Iwata
,
K.
,
Yumoto
,
K.
,
Sugishima
,
M.
,
Mizoguchi
,
C.
,
Kai
,
Y.
,
Iba
,
Y.
and
Mio
,
Y.
(
2014
).
Analysis of compaction initiation in human embryos by using time-lapse cinematography
.
J. Assist. Reprod. Genet
31
,
421
-
426
.
Jaszczak
,
K.
,
Członkowska
,
M.
,
Guszkiewicz
,
A.
and
Parada
,
R.
(
1991
).
Cytogenetic analysis of experimental interspecies goat-sheep chimera
.
J. Hered.
82
,
244
-
245
.
Johnson
,
M. H.
and
Maro
,
B.
(
1984
).
The distribution of cytoplasmic actin in mouse 8-cell blastomeres
.
J. Embryol. Exp. Morphol.
82
,
97
-
117
.
Johnson
,
M. H.
and
McConnell
,
J. M. L.
(
2004
).
Lineage allocation and cell polarity during mouse embryogenesis
.
Semin. Cell Dev. Biol.
15
,
583
-
597
.
Johnson
,
M. H.
and
Ziomek
,
C. A.
(
1981a
).
The foundation of two distinct cell lineages within the mouse morula
.
Cell
24
,
71
-
80
.
Johnson
,
M. H.
and
Ziomek
,
C. A.
(
1981b
).
Induction of polarity in mouse 8-cell blastomeres: specificity, geometry, and stability
.
J. Cell Biol.
91
,
303
-
308
.
Johnson
,
M. H.
and
Ziomek
,
C. A.
(
1983
).
Cell interactions influence the fate of mouse blastomeres undergoing the transition from the 16- to the 32-cell stage
.
Dev. Biol.
95
,
211
-
218
.
Johnson
,
M. H.
,
Pickering
,
S. J.
,
Dhiman
,
A.
,
Radcliffe
,
G. S.
and
Maro
,
B.
(
1988
).
Cytocortical organization during natural and prolonged mitosis of mouse 8-cell blastomeres
.
Dev. Camb. Engl.
102
,
143
-
158
.
Johnson
,
W. H.
,
Loskutoff
,
N. M.
,
Plante
,
Y.
and
Betteridge
,
K. J.
(
1995
).
Production of four identical calves by the separation of blastomeres from an in vitro derived four-cell embryo
.
Vet. Rec.
137
,
15
-
16
.
Kang
,
M.
,
Piliszek
,
A.
,
Artus
,
J.
and
Hadjantonakis
,
A.-K.
(
2013
).
FGF4 is required for lineage restriction and salt-and-pepper distribution of primitive endoderm factors but not their initial expression in the mouse
.
Dev. Camb. Engl.
140
,
267
-
279
.
Khan
,
D. R.
,
Dubé
,
D.
,
Gall
,
L.
,
Peynot
,
N.
,
Ruffini
,
S.
,
Laffont
,
L.
,
Le Bourhis
,
D.
,
Degrelle
,
S.
,
Jouneau
,
A.
and
Duranthon
,
V.
(
2012
).
Expression of pluripotency master regulators during two key developmental transitions: EGA and early lineage specification in the bovine embryo
.
PLoS ONE
7
,
e34110
.
Kidder
,
G. M.
and
McLachlin
,
J. R.
(
1985
).
Timing of transcription and protein synthesis underlying morphogenesis in preimplantation mouse embryos
.
Dev. Biol.
112
,
265
-
275
.
Kirchhof
,
N.
,
Carnwath
,
J. W.
,
Lemme
,
E.
,
Anastassiadis
,
K.
,
Schöler
,
H.
and
Niemann
,
H.
(
2000
).
Expression pattern of Oct-4 in preimplantation embryos of different species
.
Biol. Reprod
63
,
1698
-
1705
.
Klimczewska
,
K.
,
Kasperczuk
,
A.
and
Suwińska
,
A.
(
2018
).
The regulative nature of mammalian embryos
.
Curr. Top. Dev. Biol.
128
,
105
-
149
.
Kobolak
,
J.
,
Kiss
,
K.
,
Polgar
,
Z.
,
Mamo
,
S.
,
Rogel-Gaillard
,
C.
,
Tancos
,
Z.
,
Bock
,
I.
,
Baji
,
A. G.
,
Tar
,
K.
,
Pirity
,
M. K.
, et al. 
(
2009
).
Promoter analysis of the rabbit POU5F1 gene and its expression in preimplantation stage embryos
.
BMC Mol. Biol.
10
,
88
.
Kohri
,
N.
,
Akizawa
,
H.
,
Iisaka
,
S.
,
Bai
,
H.
,
Yanagawa
,
Y.
,
Takahashi
,
M.
,
Komatsu
,
M.
,
Kawai
,
M.
,
Nagano
,
M.
and
Kawahara
,
M.
(
2019
).
Trophectoderm regeneration to support full-term development in the inner cell mass isolated from bovine blastocyst
.
J. Biol. Chem.
294
,
19209
-
19223
.
Kono
,
K.
,
Tamashiro
,
D. A. A.
and
Alarcon
,
V. B.
(
2014
).
Inhibition of RHO-ROCK signaling enhances ICM and suppresses TE characteristics through activation of Hippo signaling in the mouse blastocyst
.
Dev. Biol.
394
,
142
-
155
.
Korotkevich
,
E.
,
Niwayama
,
R.
,
Courtois
,
A.
,
Friese
,
S.
,
Berger
,
N.
,
Buchholz
,
F.
and
Hiiragi
,
T.
(
2017
).
The apical domain is required and sufficient for the first lineage segregation in the mouse embryo
.
Dev. Cell
40
,
235
-
247.e7
.
Krawchuk
,
D.
,
Honma-Yamanaka
,
N.
,
Anani
,
S.
and
Yamanaka
,
Y.
(
2013
).
FGF4 is a limiting factor controlling the proportions of primitive endoderm and epiblast in the ICM of the mouse blastocyst
.
Dev. Biol.
384
,
65
-
71
.
Krivega
,
M.
,
Essahib
,
W.
and
Van de Velde
,
H.
(
2015
).
WNT3 and membrane-associated β-catenin regulate trophectoderm lineage differentiation in human blastocysts
.
Mol. Hum. Reprod
21
,
711
-
722
.
Kuijk
,
E. W.
,
Du Puy
,
L.
,
Van Tol
,
H. T. A.
,
Oei
,
C. H. Y.
,
Haagsman
,
H. P.
,
Colenbrander
,
B.
and
Roelen
,
B. A. J.
(
2008
).
Differences in early lineage segregation between mammals
.
Dev. Dyn. Off. Publ. Am. Assoc. Anat.
237
,
918
-
927
.
Kuijk
,
E. W.
,
van Tol
,
L. T. A.
,
Van de Velde
,
H.
,
Wubbolts
,
R.
,
Welling
,
M.
,
Geijsen
,
N.
and
Roelen
,
B. A. J.
(
2012
).
The roles of FGF and MAP kinase signaling in the segregation of the epiblast and hypoblast cell lineages in bovine and human embryos
.
Dev. Camb. Engl.
139
,
871
-
882
.
Kurotaki
,
Y.
,
Hatta
,
K.
,
Nakao
,
K.
,
Nabeshima
,
Y.-I.
and
Fujimori
,
T.
(
2007
).
Blastocyst axis is specified independently of early cell lineage but aligns with the ZP shape
.
Science
316
,
719
-
723
.
Kwon
,
J.
,
Li
,
Y.-H.
,
Jo
,
Y.-J.
,
Oh
,
Y.
,
Namgoong
,
S.
and
Kim
,
N.-H.
(
2019
).
Inhibition of MEK1/2 and GSK3 (2i system) affects blastocyst quality and early differentiation of porcine parthenotes
.
PeerJ
6
,
e5840
.
Larue
,
L.
,
Ohsugi
,
M.
,
Hirchenhain
,
J.
and
Kemler
,
R.
(
1994
).
E-cadherin null mutant embryos fail to form a trophectoderm epithelium
.
Proc. Natl. Acad. Sci. USA
91
,
8263
-
8267
.
Latham
,
K. E.
,
Solter
,
D.
and
Schultz
,
R. M.
(
1992
).
Acquisition of a transcriptionally permissive state during the 1-cell stage of mouse embryogenesis
.
Dev. Biol.
149
,
457
-
462
.
Leonavicius
,
K.
,
Royer
,
C.
,
Preece
,
C.
,
Davies
,
B.
,
Biggins
,
J. S.
and
Srinivas
,
S.
(
2018
).
Mechanics of mouse blastocyst hatching revealed by a hydrogel-based microdeformation assay
.
Proc. Natl. Acad. Sci. USA
115
,
10375
-
10380
.
Levy
,
J. B.
,
Johnson
,
M. H.
,
Goodall
,
H.
and
Maro
,
B.
(
1986
).
The timing of compaction: control of a major developmental transition in mouse early embryogenesis
.
J. Embryol. Exp. Morphol.
95
,
213
-
237
.
Lim
,
K. T.
,
Gupta
,
M. K.
,
Lee
,
S. H.
,
Jung
,
Y. H.
,
Han
,
D. W.
and
Lee
,
H. T.
(
2013
).
Possible involvement of Wnt/β-catenin signaling pathway in hatching and trophectoderm differentiation of pig blastocysts
.
Theriogenology
79
,
284
-
290.e1–2
.
Liu
,
S.
,
Bou
,
G.
,
Sun
,
R.
,
Guo
,
S.
,
Xue
,
B.
,
Wei
,
R.
,
Cooney
,
A. J.
and
Liu
,
Z.
(
2015
).
Sox2 is the faithful marker for pluripotency in pig: evidence from embryonic studies
.
Dev. Dyn
244
,
619
-
627
.
Louvet
,
S.
,
Aghion
,
J.
,
Santa-Maria
,
A.
,
Mangeat
,
P.
and
Maro
,
B.
(
1996
).
Ezrin becomes restricted to outer cells following asymmetrical division in the preimplantation mouse embryo
.
Dev. Biol.
177
,
568
-
579
.
Louvet-Vallée
,
S.
,
Dard
,
N.
,
Santa-Maria
,
A.
,
Aghion
,
J.
and
Maro
,
B.
(
2001
).
A major posttranslational modification of ezrin takes place during epithelial differentiation in the early mouse embryo
.
Dev. Biol.
231
,
190
-
200
.
Madeja
,
Z. E.
,
Sosnowski
,
J.
,
Hryniewicz
,
K.
,
Warzych
,
E.
,
Pawlak
,
P.
,
Rozwadowska
,
N.
,
Plusa
,
B.
and
Lechniak
,
D.
(
2013
).
Changes in sub-cellular localisation of trophoblast and inner cell mass specific transcription factors during bovine preimplantation development
.
BMC Dev. Biol.
13
,
32
.
Madeja
,
Z. E.
,
Hryniewicz
,
K.
,
Orsztynowicz
,
M.
,
Pawlak
,
P.
and
Perkowska
,
A.
(
2015
).
WNT/β-catenin signaling affects cell lineage and pluripotency-specific gene expression in bovine blastocysts: prospects for bovine embryonic stem cell derivation
.
Stem Cells Dev.
24
,
2437
-
2454
.
Madeja
,
Z. E.
,
Pawlak
,
P.
and
Piliszek
,
A.
(
2019
).
Beyond the mouse: non-rodent animal models for study of early mammalian development and biomedical research
.
Int. J. Dev. Biol.
63
,
187
-
201
.
Maître
,
J.-L.
(
2017
).
Mechanics of blastocyst morphogenesis
.
Biol. Cell
109
,
323
-
338
.
Maître
,
J.-L.
,
Niwayama
,
R.
,
Turlier
,
H.
,
Nédélec
,
F.
and
Hiiragi
,
T.
(
2015
).
Pulsatile cell-autonomous contractility drives compaction in the mouse embryo
.
Nat. Cell Biol.
17
,
849
-
855
.
Maître
,
J.-L.
,
Turlier
,
H.
,
Illukkumbura
,
R.
,
Eismann
,
B.
,
Niwayama
,
R.
,
Nédélec
,
F.
and
Hiiragi
,
T.
(
2016
).
Asymmetric division of contractile domains couples cell positioning and fate specification
.
Nature
536
,
344
-
348
.
Manejwala
,
F. M.
and
Schultz
,
R. M.
(
1989
).
Blastocoel expansion in the preimplantation mouse embryo: stimulation of sodium uptake by cAMP and possible involvement of cAMP-dependent protein kinase
.
Dev. Biol.
136
,
560
-
563
.
Manejwala
,
F. M.
,
Cragoe
,
E. J.
and
Schultz
,
R. M.
(
1989
).
Blastocoel expansion in the preimplantation mouse embryo: role of extracellular sodium and chloride and possible apical routes of their entry
.
Dev. Biol.
133
,
210
-
220
.
Marikawa
,
Y.
and
Alarcon
,
V. B.
(
2012
).
Creation of trophectoderm, the first epithelium, in mouse preimplantation development
.
Results Probl. Cell Differ.
55
,
165
-
184
.
Martinez-Arias
,
A. M.
,
Nichols
,
J.
and
Schröter
,
C.
(
2013
).
A molecular basis for developmental plasticity in early mammalian embryos
.
Development
140
,
3499
-
3510
.
Matsunari
,
H.
,
Nagashima
,
H.
,
Watanabe
,
M.
,
Umeyama
,
K.
,
Nakano
,
K.
,
Nagaya
,
M.
,
Kobayashi
,
T.
,
Yamaguchi
,
T.
,
Sumazaki
,
R.
,
Herzenberg
,
L. A.
, et al. 
(
2013
).
Blastocyst complementation generates exogenic pancreas in vivo in apancreatic cloned pigs
.
Proc. Natl. Acad. Sci. USA
110
,
4557
-
4562
.
Matsunari
,
H.
,
Watanabe
,
M.
,
Hasegawa
,
K.
,
Uchikura
,
A.
,
Nakano
,
K.
,
Umeyama
,
K.
,
Masaki
,
H.
,
Hamanaka
,
S.
,
Yamaguchi
,
T.
,
Nagaya
,
M.
, et al. 
(
2020
).
Compensation of disabled organogeneses in genetically modified pig fetuses by blastocyst complementation
.
Stem Cell Rep.
14
,
21
-
33
.
Mayer
,
J. F.
and
Fritz
,
H. I.
(
1974
).
The culture of preimplantation rat embryos and the production of allophenic rats
.
J. Reprod. Fertil.
39
,
1
-
9
.
Meilhac
,
S. M.
,
Adams
,
R. J.
,
Morris
,
S. A.
,
Danckaert
,
A.
,
Le Garrec
,
J.-F.
and
Zernicka-Goetz
,
M.
(
2009
).
Active cell movements coupled to positional induction are involved in lineage segregation in the mouse blastocyst
.
Dev. Biol.
331
,
210
-
221
.
Meinecke-Tillmann
,
S.
and
Meinecke
,
B.
(
1984
).
Experimental chimaeras--removal of reproductive barrier between sheep and goat
.
Nature
307
,
637
-
638
.
Meirelles
,
F. V.
,
Caetano
,
A. R.
,
Watanabe
,
Y. F.
,
Ripamonte
,
P.
,
Carambula
,
S. F.
,
Merighe
,
G. K.
and
Garcia
,
S. M.
(
2004
).
Genome activation and developmental block in bovine embryos
.
Anim. Reprod. Sci.
82-83
,
13
-
20
.
Menchero
,
S.
,
Rollan
,
I.
,
Lopez-Izquierdo
,
A.
,
Andreu
,
M. J.
,
Sainz de Aja
,
J.
,
Kang
,
M.
,
Adan
,
J.
,
Benedito
,
R.
,
Rayon
,
T.
,
Hadjantonakis
,
A.-K.
, et al. 
(
2019
).
Transitions in cell potency during early mouse development are driven by Notch
.
eLife
8
,
e42930
.
Mihajlović
,
A. I.
and
Bruce
,
A. W.
(
2016
).
Rho-associated protein kinase regulates subcellular localisation of Angiomotin and Hippo-signalling during preimplantation mouse embryo development
.
Reprod. Biomed. Online
33
,
381
-
390
.
Mihajlović
,
A. I.
and
Bruce
,
A. W.
(
2017
).
The first cell-fate decision of mouse preimplantation embryo development: integrating cell position and polarity
.
Open Biol.
7
.
Milewski
,
R.
and
Ajduk
,
A.
(
2017
).
Time-lapse imaging of cleavage divisions in embryo quality assessment
.
Reproduction
154
,
R37
-
R53
.
Mitsui
,
K.
,
Tokuzawa
,
Y.
,
Itoh
,
H.
,
Segawa
,
K.
,
Murakami
,
M.
,
Takahashi
,
K.
,
Maruyama
,
M.
,
Maeda
,
M.
and
Yamanaka
,
S.
(
2003
).
The homeoprotein Nanog is required for maintenance of pluripotency in mouse epiblast and ES cells
.
Cell
113
,
631
-
642
.
Moore
,
N. W.
,
Adams
,
C. E.
and
Rowson
,
L. E.
(
1968
).
Developmental potential of single blastomeres of the rabbit egg
.
J. Reprod. Fertil
.
17
,
527
-
531
.
Moradi
,
M.
,
Riasi
,
A.
,
Ostadhosseini
,
S.
,
Hajian
,
M.
,
Hosseini
,
M.
,
Hosseinnia
,
P.
and
Nasr-Esfahani
,
M. H.
(
2015
).
Expression profile of FGF receptors in preimplantation ovine embryos and the effect of FGF2 and PD173074
.
Growth Factors
33
,
393
-
400
.
Moriwaki
,
K.
,
Tsukita
,
S.
and
Furuse
,
M.
(
2007
).
Tight junctions containing claudin 4 and 6 are essential for blastocyst formation in preimplantation mouse embryos
.
Dev. Biol.
312
,
509
-
522
.
Morris
,
S. A.
,
Teo
,
R. T. Y.
,
Li
,
H.
,
Robson
,
P.
,
Glover
,
D. M.
and
Zernicka-Goetz
,
M.
(
2010
).
Origin and formation of the first two distinct cell types of the inner cell mass in the mouse embryo
.
Proc. Natl. Acad. Sci. USA
107
,
6364
-
6369
.
Motosugi
,
N.
,
Bauer
,
T.
,
Polanski
,
Z.
,
Solter
,
D.
and
Hiiragi
,
T.
(
2005
).
Polarity of the mouse embryo is established at blastocyst and is not prepatterned
.
Genes Dev.
19
,
1081
-
1092
.
Negrón-Pérez
,
V. M.
,
Zhang
,
Y.
and
Hansen
,
P. J.
(
2017
).
Single-cell gene expression of the bovine blastocyst
.
Reprod. Camb. Engl
.
154
,
627
-
644
.
Niakan
,
K. K.
and
Eggan
,
K.
(
2013
).
Analysis of human embryos from zygote to blastocyst reveals distinct gene expression patterns relative to the mouse
.
Dev. Biol.
375
,
54
-
64
.
Niakan
,
K. K.
,
Ji
,
H.
,
Maehr
,
R.
,
Vokes
,
S. A.
,
Rodolfa
,
K. T.
,
Sherwood
,
R. I.
,
Yamaki
,
M.
,
Dimos
,
J. T.
,
Chen
,
A. E.
,
Melton
,
D. A.
, et al. 
(
2010
).
Sox17 promotes differentiation in mouse embryonic stem cells by directly regulating extraembryonic gene expression and indirectly antagonizing self-renewal
.
Genes Dev.
24
,
312
-
326
.
Nichols
,
J.
,
Zevnik
,
B.
,
Anastassiadis
,
K.
,
Niwa
,
H.
,
Klewe-Nebenius
,
D.
,
Chambers
,
I.
,
Schöler
,
H.
and
Smith
,
A.
(
1998
).
Formation of pluripotent stem cells in the mammalian embryo depends on the POU transcription factor Oct4
.
Cell
95
,
379
-
391
.
Nichols
,
J.
,
Silva
,
J.
,
Roode
,
M.
and
Smith
,
A.
(
2009
).
Suppression of Erk signalling promotes ground state pluripotency in the mouse embryo
.
Dev. Camb. Engl.
136
,
3215
-
3222
.
Niimura
,
S.
(
2003
).
Time-lapse videomicrographic analyses of contractions in mouse blastocysts
.
J. Reprod. Dev.
49
,
413
-
423
.
Nikas
,
G.
,
Ao
,
A.
,
Winston
,
R. M.
and
Handyside
,
A. H.
(
1996
).
Compaction and surface polarity in the human embryo in vitro
.
Biol. Reprod
55
,
32
-
37
.
Nishimura
,
M.
(
2001
).
Timing of implantation in New Zealand White rabbits
.
Congenit. Anom.
41
,
198
-
203
.
Nishioka
,
N.
,
Yamamoto
,
S.
,
Kiyonari
,
H.
,
Sato
,
H.
,
Sawada
,
A.
,
Ota
,
M.
,
Nakao
,
K.
and
Sasaki
,
H.
(
2008
).
Tead4 is required for specification of trophectoderm in pre-implantation mouse embryos
.
Mech. Dev.
125
,
270
-
283
.
Nishioka
,
N.
,
Inoue
,
K.
,
Adachi
,
K.
,
Kiyonari
,
H.
,
Ota
,
M.
,
Ralston
,
A.
,
Yabuta
,
N.
,
Hirahara
,
S.
,
Stephenson
,
R. O.
,
Ogonuki
,
N.
, et al. 
(
2009
).
The Hippo signaling pathway components Lats and Yap pattern Tead4 activity to distinguish mouse trophectoderm from inner cell mass
.
Dev. Cell
16
,
398
-
410
.
Niwayama
,
R.
,
Moghe
,
P.
,
Liu
,
Y.-J.
,
Fabrèges
,
D.
,
Buchholz
,
F.
,
Piel
,
M.
and
Hiiragi
,
T.
(
2019
).
A tug-of-war between cell shape and polarity controls division orientation to ensure robust patterning in the mouse blastocyst
.
Dev. Cell
51
,
564
-
574.e6
.
Ohnishi
,
Y.
,
Huber
,
W.
,
Tsumura
,
A.
,
Kang
,
M.
,
Xenopoulos
,
P.
,
Kurimoto
,
K.
,
Oleś
,
A. K.
,
Araúzo-Bravo
,
M. J.
,
Saitou
,
M.
,
Hadjantonakis
,
A.-K.
, et al. 
(
2014
).
Cell-to-cell expression variability followed by signal reinforcement progressively segregates early mouse lineages
.
Nat. Cell Biol.
16
,
27
-
37
.
Ozawa
,
M.
,
Sakatani
,
M.
,
Yao
,
J.
,
Shanker
,
S.
,
Yu
,
F.
,
Yamashita
,
R.
,
Wakabayashi
,
S.
,
Nakai
,
K.
,
Dobbs
,
K. B.
,
Sudano
,
M. J.
, et al. 
(
2012
).
Global gene expression of the inner cell mass and trophectoderm of the bovine blastocyst
.
BMC Dev. Biol.
12
,
33
.
Palmieri
,
S. L.
,
Peter
,
W.
,
Hess
,
H.
and
Schöler
,
H. R.
(
1994
).
Oct-4 transcription factor is differentially expressed in the mouse embryo during establishment of the first two extraembryonic cell lineages involved in implantation
.
Dev. Biol.
166
,
259
-
267
.
Petropoulos
,
S.
,
Edsgärd
,
D.
,
Reinius
,
B.
,
Deng
,
Q.
,
Panula
,
S. P.
,
Codeluppi
,
S.
,
Plaza Reyes
,
A.
,
Linnarsson
,
S.
,
Sandberg
,
R.
and
Lanner
,
F.
(
2016
).
Single-cell RNA-seq reveals lineage and X chromosome dynamics in human preimplantation embryos
.
Cell
165
,
1012
-
1026
.
Piliszek
,
A.
and
Madeja
,
Z. E.
(
2018
).
Chapter eleven - pre-implantation development of domestic animals
. In
Current Topics in Developmental Biology, Cell Fate in Mammalian Development
(ed.
B.
Plusa
and
A.-K.
Hadjantonakis
), pp.
267
-
294
.
Academic Press
.
Piliszek
,
A.
,
Barłowska
,
K.
,
Madeja
,
Z.
,
Pawlak
,
P.
,
Plusa
,
B
. (
2017a
).
Differentiation of trophectoderm in rabbit embryos is initiated in the absence of Gata3 and Cdx2
.
Mech. Dev.
145
,
S79
.
Piliszek
,
A.
,
Madeja
,
Z. E.
and
Plusa
,
B.
(
2017b
).
Suppression of ERK signalling abolishes primitive endoderm formation but does not promote pluripotency in rabbit embryo
.
Development
144
,
3719
.
Piotrowska
,
K.
,
Wianny
,
F.
,
Pedersen
,
R. A.
and
Zernicka-Goetz
,
M.
(
2001
).
Blastomeres arising from the first cleavage division have distinguishable fates in normal mouse development
.
Dev. Camb. Engl.
128
,
3739
-
3748
.
Piotrowska-Nitsche
,
K.
,
Perea-Gomez
,
A.
,
Haraguchi
,
S.
and
Zernicka-Goetz
,
M.
(
2005
).
Four-cell stage mouse blastomeres have different developmental properties
.
Dev. Camb. Engl.
132
,
479
-
490
.
Plachta
,
N.
,
Bollenbach
,
T.
,
Pease
,
S.
,
Fraser
,
S. E.
and
Pantazis
,
P.
(
2011
).
Oct4 kinetics predict cell lineage patterning in the early mammalian embryo
.
Nat. Cell Biol.
13
,
117
-
123
.
Plusa
,
B.
and
Hadjantonakis
,
A.-K.
(
2016
).
Mammalian development: mechanics drives cell differentiation
.
Nature
536
,
281
-
282
.
Plusa
,
B.
and
Hadjantonakis
,
A.-K.
(
2018
).
(De)constructing the blastocyst: lessons in self-organization from the mouse
.
Curr. Opin. Syst. Biol.
11
,
98
-
106
.
Plusa
,
B.
,
Frankenberg
,
S.
,
Chalmers
,
A.
,
Hadjantonakis
,
A.-K.
,
Moore
,
C. A.
,
Papalopulu
,
N.
,
Papaioannou
,
V. E.
,
Glover
,
D. M.
and
Zernicka-Goetz
,
M.
(
2005
).
Downregulation of Par3 and aPKC function directs cells towards the ICM in the preimplantation mouse embryo
.
J. Cell Sci.
118
,
505
-
515
.
Plusa
,
B.
,
Piliszek
,
A.
,
Frankenberg
,
S.
,
Artus
,
J.
and
Hadjantonakis
,
A.-K.
(
2008
).
Distinct sequential cell behaviours direct primitive endoderm formation in the mouse blastocyst
.
Dev. Camb. Engl.
135
,
3081
-
3091
.
Posfai
,
E.
,
Petropoulos
,
S.
,
de Barros
,
F. R. O.
,
Schell
,
J. P.
,
Jurisica
,
I.
,
Sandberg
,
R.
,
Lanner
,
F.
and
Rossant
,
J.
(
2017
).
Position- and Hippo signaling-dependent plasticity during lineage segregation in the early mouse embryo
.
eLife
6
,
e22906
.
Pratt
,
H. P.
,
Ziomek
,
C. A.
,
Reeve
,
W. J.
and
Johnson
,
M. H.
(
1982
).
Compaction of the mouse embryo: an analysis of its components
.
J. Embryol. Exp. Morphol.
70
,
113
-
132
.
Ralston
,
A.
,
Cox
,
B. J.
,
Nishioka
,
N.
,
Sasaki
,
H.
,
Chea
,
E.
,
Rugg-Gunn
,
P.
,
Guo
,
G.
,
Robson
,
P.
,
Draper
,
J. S.
and
Rossant
,
J.
(
2010
).
Gata3 regulates trophoblast development downstream of Tead4 and in parallel to Cdx2
.
Development
137
,
395
-
403
.
Ramos-Ibeas
,
P.
,
Sang
,
F.
,
Zhu
,
Q.
,
Tang
,
W. W. C.
,
Withey
,
S.
,
Klisch
,
D.
,
Wood
,
L.
,
Loose
,
M.
,
Surani
,
M. A.
and
Alberio
,
R.
(
2019
).
Pluripotency and X chromosome dynamics revealed in pig pre-gastrulating embryos by single cell analysis
.
Nat. Commun.
10
,
500
.
Rayon
,
T.
,
Menchero
,
S.
,
Nieto
,
A.
,
Xenopoulos
,
P.
,
Crespo
,
M.
,
Cockburn
,
K.
,
Cañon
,
S.
,
Sasaki
,
H.
,
Hadjantonakis
,
A.-K.
,
de la Pompa
,
J. L.
, et al. 
(
2014
).
Notch and Hippo converge on Cdx2 to specify the trophectoderm lineage in the mouse blastocyst
.
Dev. Cell
30
,
410
-
422
.
Reeve
,
W. J.
and
Ziomek
,
C. A.
(
1981
).
Distribution of microvilli on dissociated blastomeres from mouse embryos: evidence for surface polarization at compaction
.
J. Embryol. Exp. Morphol.
62
,
339
-
350
.
Reima
,
I.
,
Lehtonen
,
E.
,
Virtanen
,
I.
and
Fléchon
,
J.-E.
(
1993
).
The cytoskeleton and associated proteins during cleavage, compaction and blastocyst differentiation in the pig
.
Differentiation
54
,
35
-
45
.
Rodríguez
,
A.
,
Allegrucci
,
C.
and
Alberio
,
R.
(
2012
).
Modulation of pluripotency in the porcine embryo and iPS cells
.
PLoS ONE
7
,
e49079
.
Roode
,
M.
,
Blair
,
K.
,
Snell
,
P.
,
Elder
,
K.
,
Marchant
,
S.
,
Smith
,
A.
and
Nichols
,
J.
(
2012
).
Human hypoblast formation is not dependent on FGF signalling
.
Dev. Biol.
361
,
358
-
363
.
Rossant
,
J.
(
1975a
).
Investigation of the determinative state of the mouse inner cell mass. II. The fate of isolated inner cell masses transferred to the oviduct
.
J. Embryol. Exp. Morphol.
33
,
991
-
1001
.
Rossant
,
J.
(
1975b
).
Investigation of the determinative state of the mouse inner cell mass. I. Aggregation of isolated inner cell masses with morulae
.
J. Embryol. Exp. Morphol.
33
,
979
-
990
.
Rossant
,
J.
(
1976
).
Postimplantation development of blastomeres isolated from 4- and 8-cell mouse eggs
.
J. Embryol. Exp. Morphol.
36
,
283
-
290
.
Rossant
,
J.
and
Tam
,
P. P. L.
(
2009
).
Blastocyst lineage formation, early embryonic asymmetries and axis patterning in the mouse
.
Dev. Camb. Engl.
136
,
701
-
713
.
Rossant
,
J.
,
Chazaud
,
C.
and
Yamanaka
,
Y.
(
2003
).
Lineage allocation and asymmetries in the early mouse embryo
.
Philos. Trans. R. Soc. Lond. B. Biol. Sci.
358
,
1341
-
1348
.
discussion 1349
.
Saini
,
D.
and
Yamanaka
,
Y.
(
2018
).
Cell polarity-dependent regulation of cell allocation and the first lineage specification in the preimplantation mouse embryo
.
Curr. Top. Dev. Biol.
128
,
11
-
35
.
Saiz
,
N.
,
Grabarek
,
J. B.
,
Sabherwal
,
N.
,
Papalopulu
,
N.
and
Plusa
,
B.
(
2013
).
Atypical protein kinase C couples cell sorting with primitive endoderm maturation in the mouse blastocyst
.
Dev. Camb. Engl.
140
,
4311
-
4322
.
Saiz
,
N.
,
Williams
,
K. M.
,
Seshan
,
V. E.
and
Hadjantonakis
,
A.-K.
(
2016
).
Asynchronous fate decisions by single cells collectively ensure consistent lineage composition in the mouse blastocyst
.
Nat. Commun.
7
,
13463
.
Saiz
,
N.
,
Mora-Bitria
,
L.
,
Rahman
,
S.
,
George
,
H.
,
Herder
,
J. P.
,
García-Ojalvo
,
J.
and
Hadjantonakis
,
A.-K.
(
2020
).
Growth factor-mediated coupling between lineage size and cell fate choice underlies robustness of mammalian development
.
bioRxiv
,
2019.12.27.889006
.
Sakurai
,
N.
,
Takahashi
,
K.
,
Emura
,
N.
,
Hashizume
,
T.
and
Sawai
,
K.
(
2017
).
Effects of downregulating TEAD4 transcripts by RNA interference on early development of bovine embryos
.
J. Reprod. Dev.
63
,
135
-
142
.
Sasaki
,
H.
(
2017
).
Roles and regulations of Hippo signaling during preimplantation mouse development
.
Dev. Growth Differ.
59
,
12
-
20
.
Schrode
,
N.
,
Saiz
,
N.
,
Di Talia
,
S.
and
Hadjantonakis
,
A.-K.
(
2014
).
GATA6 levels modulate primitive endoderm cell fate choice and timing in the mouse blastocyst
.
Dev. Cell
29
,
454
-
467
.
Sefton
,
M.
,
Johnson
,
M. H.
and
Clayton
,
L.
(
1992
).
Synthesis and phosphorylation of uvomorulin during mouse early development
.
Dev. Camb. Engl.
115
,
313
-
318
.
Sefton
,
M.
,
Johnson
,
M. H.
,
Clayton
,
L.
and
McConnell
,
J. M.
(
1996
).
Experimental manipulations of compaction and their effects on the phosphorylation of uvomorulin
.
Mol. Reprod. Dev.
44
,
77
-
87
.
Shahbazi
,
M. N.
and
Zernicka-Goetz
,
M.
(
2018
).
Deconstructing and reconstructing the mouse and human early embryo
.
Nat. Cell Biol.
20
,
878
-
887
.
Shen
,
Q.-Y.
,
Yu
,
S.
,
Zhang
,
Y.
,
Zhou
,
Z.
,
Zhu
,
Z.-S.
,
Pan
,
Q.
,
Lv
,
S.
,
Niu
,
H.-M.
,
Li
,
N.
,
Peng
,
S.
, et al. 
(
2019
).
Characterization of porcine extraembryonic endoderm cells
.
Cell Prolif.
52
,
e12591
.
Shi
,
J.
,
Chen
,
Q.
,
Li
,
X.
,
Zheng
,
X.
,
Zhang
,
Y.
,
Qiao
,
J.
,
Tang
,
F.
,
Tao
,
Y.
,
Zhou
,
Q.
and
Duan
,
E.
(
2015
).
Dynamic transcriptional symmetry-breaking in pre-implantation mammalian embryo development revealed by single-cell RNA-seq
.
Dev. Camb. Engl.
142
,
3468
-
3477
.
Shirayoshi
,
Y.
,
Okada
,
T. S.
and
Takeichi
,
M.
(
1983
).
The calcium-dependent cell-cell adhesion system regulates inner cell mass formation and cell surface polarization in early mouse development
.
Cell
35
,
631
-
638
.
Skrzecz
,
I.
and
Karasiewicz
,
J.
(
1987
).
Decompaction and recompaction of mouse preimplantation embryos. Rouxs Arch
.
Dev. Biol.
196
,
397
-
400
.
Smith
,
R.
and
McLaren
,
A.
(
1977
).
Factors affecting the time of formation of the mouse blastocoele
.
J. Embryol. Exp. Morphol.
41
,
79
-
92
.
Smith
,
C. S.
,
Berg
,
D. K.
,
Berg
,
M.
and
Pfeffer
,
P. L.
(
2010
).
Nuclear transfer-specific defects are not apparent during the second week of embryogenesis in cattle
.
Cell. Reprogramming
12
,
699
-
707
.
Stephenson
,
R. O.
,
Yamanaka
,
Y.
and
Rossant
,
J.
(
2010
).
Disorganized epithelial polarity and excess trophectoderm cell fate in preimplantation embryos lacking E-cadherin
.
Dev. Camb. Engl.
137
,
3383
-
3391
.
Steptoe
,
P. C.
,
Edwards
,
R. G.
and
Purdy
,
J. M.
(
1971
).
Human Blastocysts grown in Culture
.
Nature
229
,
132
-
133
.
Stern
,
M. S.
and
Wilson
,
I. B.
(
1972
).
Experimental studies on the organization of the preimplantation mouse embryo. I. Fusion of asynchronously cleaving eggs
.
J. Embryol. Exp. Morphol.
28
,
247
-
254
.
Stirparo
,
G. G.
,
Boroviak
,
T.
,
Guo
,
G.
,
Nichols
,
J.
,
Smith
,
A.
and
Bertone
,
P.
(
2018
).
Integrated analysis of single-cell embryo data yields a unified transcriptome signature for the human pre-implantation epiblast
.
Dev. Camb. Engl.
145
,
dev158501
.
Strumpf
,
D.
,
Mao
,
C.-A.
,
Yamanaka
,
Y.
,
Ralston
,
A.
,
Chawengsaksophak
,
K.
,
Beck
,
F.
and
Rossant
,
J.
(
2005
).
Cdx2 is required for correct cell fate specification and differentiation of trophectoderm in the mouse blastocyst
.
Dev. Camb. Engl.
132
,
2093
-
2102
.
Summers
,
P. M.
,
Shelton
,
J. N.
and
Bell
,
K.
(
1983
).
Synthesis of primary Bos taurus - Bos indicus chimaeric calves
.
Anim. Reprod. Sci.
6
,
91
-
102
.
Suwińska
,
A.
,
Czołowska
,
R.
,
Ozdzeński
,
W.
and
Tarkowski
,
A. K.
(
2008
).
Blastomeres of the mouse embryo lose totipotency after the fifth cleavage division: expression of Cdx2 and Oct4 and developmental potential of inner and outer blastomeres of 16- and 32-cell embryos
.
Dev. Biol.
322
,
133
-
144
.
Suzuki
,
H.
,
Azuma
,
T.
,
Koyama
,
H.
and
Yang
,
X.
(
1999
).
Development of cellular polarity of hamster embryos during compaction
.
Biol. Reprod
61
,
521
-
526
.
Tabansky
,
I.
,
Lenarcic
,
A.
,
Draft
,
R. W.
,
Loulier
,
K.
,
Keskin
,
D. B.
,
Rosains
,
J.
,
Rivera-Feliciano
,
J.
,
Lichtman
,
J. W.
,
Livet
,
J.
,
Stern
,
J. N. H.
, et al. 
(
2013
).
Developmental bias in cleavage-stage mouse blastomeres
.
Curr. Biol. CB
23
,
21
-
31
.
Tachibana
,
M.
,
Sparman
,
M.
,
Ramsey
,
C.
,
Ma
,
H.
,
Lee
,
H.-S.
,
Penedo
,
M. C. T.
and
Mitalipov
,
S.
(
2012
).
Generation of chimeric rhesus monkeys
.
Cell
148
,
285
-
295
.
Tam
,
P. P. L.
and
Rossant
,
J.
(
2003
).
Mouse embryonic chimeras: tools for studying mammalian development
.
Dev. Camb. Engl.
130
,
6155
-
6163
.
Táncos
,
Z.
,
Bock
,
I.
,
Nemes
,
C.
,
Kobolák
,
J.
and
Dinnyés
,
A.
(
2015
).
Cloning and characterization of rabbit POU5F1, SOX2, KLF4, C-MYC and NANOG pluripotency-associated genes
.
Gene
566
,
148
-
157
.
Tarkowski
,
A. K.
(
1959
).
Experiments on the development of isolated blastomers of mouse eggs
.
Nature
184
,
1286
-
1287
.
Tarkowski
,
A. K.
(
1961
).
Mouse chimaeras developed from fused eggs
.
Nature
190
,
857
-
860
.
Tarkowski
,
A. K.
and
Wróblewska
,
J.
(
1967
).
Development of blastomeres of mouse eggs isolated at the 4- and 8-cell stage
.
J. Embryol. Exp. Morphol.
18
,
155
-
180
.
Tarkowski
,
A. K.
,
Suwińska
,
A.
,
Czołowska
,
R.
and
Ożdżeński
,
W.
(
2010
).
Individual blastomeres of 16- and 32-cell mouse embryos are able to develop into foetuses and mice
.
Dev. Biol.
348
,
190
-
198
.
Thamodaran
,
V.
and
Bruce
,
A. W.
(
2016
).
p38 (Mapk14/11) occupies a regulatory node governing entry into primitive endoderm differentiation during preimplantation mouse embryo development
.
Open Biol.
6
,
160190
.
Torres-Padilla
,
M.-E.
,
Parfitt
,
D.-E.
,
Kouzarides
,
T.
and
Zernicka-Goetz
,
M.
(
2007
).
Histone arginine methylation regulates pluripotency in the early mouse embryo
.
Nature
445
,
214
-
218
.
Tucker
,
E. M.
,
Moor
,
R. M.
and
Rowson
,
L. E.
(
1974
).
Tetraparental sheep chimaeras induced by blastomere transplantation. Changes in blood type with age
.
Immunology
26
,
613
-
621
.
Van de Velde
,
H.
,
Cauffman
,
G.
,
Tournaye
,
H.
,
Devroey
,
P.
and
Liebaers
,
I.
(
2008
).
The four blastomeres of a 4-cell stage human embryo are able to develop individually into blastocysts with inner cell mass and trophectoderm
.
Hum. Reprod. Oxf. Engl.
23
,
1742
-
1747
.
van Eijk
,
M. J. T.
,
van Rooijen
,
M. A.
,
Modina
,
S.
,
Scesi
,
L.
,
Folkers
,
G.
,
van Tol
,
H.T.A.
,
Bevers
,
M. M.
,
Fisher
,
S. R.
,
Lewin
,
H. A.
,
Rakacolli
,
D.
et al.
(
1999
).
Molecular cloning, genetic mapping, and developmental expression of bovine POU5F1
.
Biol. Reprod.
60
,
1093
-
1103
.
Van Soom
,
A. V.
,
Boerjan
,
M. L.
,
Bols
,
P. E. J.
,
Vanroose
,
G.
,
Lein
,
A.
,
Coryn
,
M.
and
de Kruif
,
A.
(
1997a
).
Timing of compaction and inner cell allocation in bovine embryos produced in vivo after superovulation
.
Biol. Reprod
57
,
1041
-
1049
.
Van Soom
,
A.
,
Ysebaert
,
M. T.
and
de Kruif
,
A.
(
1997b
).
Relationship between timing of development, morula morphology, and cell allocation to inner cell mass and trophectoderm in in vitro-produced bovine embryos
.
Mol. Reprod. Dev.
47
,
47
-
56
.
Veiga
,
A.
,
Calderon
,
G.
,
Barri
,
P. N.
and
Coroleu
,
B.
(
1987
).
Pregnancy after the replacement of a frozen-thawed embryo with less than 50% intact blastomeres
.
Hum. Reprod. Oxf. Engl.
2
,
321
-
323
.
Vestweber
,
D.
,
Gossler
,
A.
,
Boller
,
K.
and
Kemler
,
R.
(
1987
).
Expression and distribution of cell adhesion molecule uvomorulin in mouse preimplantation embryos
.
Dev. Biol.
124
,
451
-
456
.
Vinot
,
S.
,
Le
,
T.
,
Ohno
,
S.
,
Pawson
,
T.
,
Maro
,
B.
and
Louvet-Vallée
,
S.
(
2005
).
Asymmetric distribution of PAR proteins in the mouse embryo begins at the 8-cell stage during compaction
.
Dev. Biol.
282
,
307
-
319
.
Vries
,
W. N. d.
,
Evsikov
,
A. V.
,
Haac
,
B. E.
,
Fancher
,
K. S.
,
Holbrook
,
A. E.
,
Kemler
,
R.
,
Solter
,
D.
and
Knowles
,
B. B.
(
2004
).
Maternal β-catenin and E-cadherin in mouse development
.
Development
131
,
4435
-
4445
.
Wang
,
H.
,
Ding
,
T.
,
Brown
,
N.
,
Yamamoto
,
Y.
,
Prince
,
L. S.
,
Reese
,
J.
and
Paria
,
B. C.
(
2008
).
Zonula occludens-1 (ZO-1) is involved in morula to blastocyst transformation in the mouse
.
Dev. Biol.
318
,
112
-
125
.
Wang
,
C.
,
Han
,
X.
,
Zhou
,
Z.
,
Uyunbilig
,
B.
,
Huang
,
X.
,
Li
,
R.
and
Li
,
X.
(
2019
).
Wnt3a Activates the WNT-YAP/TAZ pathway to sustain CDX2 expression in bovine trophoblast stem cells
.
DNA Cell Biol.
38
,
410
-
422
.
Watanabe
,
T.
,
Biggins
,
J. S.
,
Tannan
,
N. B.
and
Srinivas
,
S.
(
2014
).
Limited predictive value of blastomere angle of division in trophectoderm and inner cell mass specification
.
Development
141
,
2279
-
2288
.
Wathes
,
D. C.
and
Wooding
,
F. B.
(
1980
).
An electron microscopic study of implantation in the cow
.
Am. J. Anat.
159
,
285
-
306
.
Wennekamp
,
S.
,
Mesecke
,
S.
,
Nédélec
,
F.
and
Hiiragi
,
T.
(
2013
).
A self-organization framework for symmetry breaking in the mammalian embryo
.
Nat. Rev. Mol. Cell Biol.
14
,
452
-
459
.
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.
(
2016
).
Long-lived binding of Sox2 to DNA predicts cell fate in the four-cell mouse embryo
.
Cell
165
,
75
-
87
.
White
,
M. D.
,
Zenker
,
J.
,
Bissiere
,
S.
and
Plachta
,
N.
(
2017
).
How cells change shape and position in the early mammalian embryo
.
Curr. Opin. Cell Biol.
44
,
7
-
13
.
White
,
M. D.
,
Zenker
,
J.
,
Bissiere
,
S.
and
Plachta
,
N.
(
2018
).
Instructions for Assembling the Early Mammalian Embryo
.
Dev. Cell
45
,
667
-
679
.
Wicklow
,
E.
,
Blij
,
S.
,
Frum
,
T.
,
Hirate
,
Y.
,
Lang
,
R. A.
,
Sasaki
,
H.
and
Ralston
,
A.
(
2014
).
HIPPO pathway members restrict SOX2 to the inner cell mass where it promotes ICM fates in the mouse blastocyst
.
PLoS Genet.
10
,
e1004618
.
Wigger
,
M.
,
Kisielewska
,
K.
,
Filimonow
,
K.
,
Plusa
,
B.
,
Maleszewski
,
M.
and
Suwińska
,
A.
(
2017
).
Plasticity of the inner cell mass in mouse blastocyst is restricted by the activity of FGF/MAPK pathway
.
Sci. Rep.
7
,
15136
.
Willadsen
,
S. M.
(
1981
).
The development capacity of blastomeres from 4- and 8-cell sheep embryos
.
J. Embryol. Exp. Morphol.
65
,
165
-
172
.
Williams
,
T. J.
,
Munro
,
R. K.
and
Shelton
,
J. N.
(
1990
).
Production of interspecies chimeric calves by aggregation of Bos indicus and Bos taurus demi-embryos
.
Reprod. Fertil. Dev.
2
,
385
-
394
.
Winkel
,
G. K.
,
Ferguson
,
J. E.
,
Takeichi
,
M.
and
Nuccitelli
,
R.
(
1990
).
Activation of protein kinase C triggers premature compaction in the four-cell stage mouse embryo
.
Dev. Biol.
138
,
1
-
15
.
Wolf
,
X. A.
,
Serup
,
P.
and
Hyttel
,
P.
(
2011
).
Three-dimensional localisation of NANOG, OCT4, and E-cadherin in porcine pre- and peri-implantation embryos
.
Dev. Dyn.
240
,
204
-
210
.
Wong
,
C. C.
,
Loewke
,
K. E.
,
Bossert
,
N. L.
,
Behr
,
B.
,
De Jonge
,
C. J.
,
Baer
,
T. M.
and
Reijo Pera
,
R. A.
(
2010
).
Non-invasive imaging of human embryos before embryonic genome activation predicts development to the blastocyst stage
.
Nat. Biotechnol.
28
,
1115
-
1121
.
Wu
,
J.
,
Huang
,
B.
,
Chen
,
H.
,
Yin
,
Q.
,
Liu
,
Y.
,
Xiang
,
Y.
,
Zhang
,
B.
,
Liu
,
B.
,
Wang
,
Q.
,
Xia
,
W.
, et al. 
(
2016
).
The landscape of accessible chromatin in mammalian preimplantation embryos
.
Nature
534
,
652
-
657
.
Xenopoulos
,
P.
,
Kang
,
M.
,
Puliafito
,
A.
,
Di Talia
,
S.
and
Hadjantonakis
,
A.-K.
(
2015
).
Heterogeneities in nanog expression drive stable commitment to pluripotency in the mouse blastocyst
.
Cell Rep.
10
,
1508
-
1520
.
Yagi
,
R.
,
Kohn
,
M. J.
,
Karavanova
,
I.
,
Kaneko
,
K. J.
,
Vullhorst
,
D.
,
DePamphilis
,
M. L.
and
Buonanno
,
A.
(
2007
).
Transcription factor TEAD4 specifies the trophectoderm lineage at the beginning of mammalian development
.
Dev. Camb. Engl.
134
,
3827
-
3836
.
Yamanaka
,
Y.
,
Ralston
,
A.
,
Stephenson
,
R. O.
and
Rossant
,
J.
(
2006
).
Cell and molecular regulation of the mouse blastocyst
.
Dev. Dyn. Off. Publ. Am. Assoc. Anat.
235
,
2301
-
2314
.
Yamanaka
,
Y.
,
Lanner
,
F.
and
Rossant
,
J.
(
2010
).
FGF signal-dependent segregation of primitive endoderm and epiblast in the mouse blastocyst
.
Dev. Camb. Engl.
137
,
715
-
724
.
Zakharova
,
E. E.
,
Zaletova
,
V. V.
and
Krivokharchenko
,
A. S.
(
2014
).
Biopsy of human morula-stage embryos: outcome of 215 IVF/ICSI cycles with PGS
.
PLoS ONE
9
,
e106433
.
Zenker
,
J.
,
White
,
M. D.
,
Gasnier
,
M.
,
Alvarez
,
Y. D.
,
Lim
,
H. Y. G.
,
Bissiere
,
S.
,
Biro
,
M.
and
Plachta
,
N.
(
2018
).
Expanding actin rings zipper the mouse embryo for blastocyst formation
.
Cell
173
,
776
-
791.e17
.
Ziomek
,
C. A.
,
Chatot
,
C. L.
and
Manes
,
C.
(
1990
).
Polarization of blastomeres in the cleaving rabbit embryo
.
J. Exp. Zool.
256
,
84
-
91
.
Zyzynska-Galenska
,
K.
,
Piliszek
,
A.
and
Modliński
,
J. A.
(
2017
).
From blastomeres to somatic cells: reflections on cell developmental potential in light of chimaera studies - a review
.
Anim. Sci. Paper. Rep.
35
,
225
-
240.

Competing interests

The authors declare no competing or financial interests.