• No results found

The maternal to zygotic transition revisited

N/A
N/A
Protected

Academic year: 2020

Share "The maternal to zygotic transition revisited"

Copied!
20
0
0

Loading.... (view fulltext now)

Full text

(1)

REVIEW

The maternal-to-zygotic transition revisited

Nadine L. Vastenhouw1,*, Wen Xi Cao2and Howard D. Lipshitz2,*

ABSTRACT

The development of animal embryos is initially directed by maternal gene products. Then, during the maternal-to-zygotic transition (MZT), developmental control is handed to the zygotic genome. Extensive research in both vertebrate and invertebrate model organisms has revealed that the MZT can be subdivided into two phases, during which very different modes of gene regulation are implemented: initially, regulation is exclusively post-transcriptional and post-translational, following which gradual activation of the zygotic genome leads to predominance of transcriptional regulation. These changes in the gene expression program of embryos are precisely controlled and highly interconnected. Here, we review current understanding of the mechanisms that underlie handover of developmental control during the MZT.

KEY WORDS: MZT, Genome activation, Maternal mRNAs

Introduction

Early animal development is controlled by maternal mRNAs and proteins that are loaded into the egg during oogenesis. These implement fundamental molecular and cellular processes, as well as the specification of initial cell fates and pattern. As development proceeds, control is handed over from maternally provided gene products to those synthesized from the zygotic genome. This maternal-to-zygotic transition (MZT) is characterized by multiple events, the most striking among which are the elimination of a large subset of maternal gene products and the onset of transcription from the zygote’s genome (Box 1).

The MZT has been referred to previously as a‘play in two acts’ (Tadros and Lipshitz, 2009). Since then, technological advances have dramatically increased our understanding of the MZT. Although it remains a two-act play, many scenes and players have been added, and the plot has become more complex and better understood. Here, we describe how the gene expression program of embryos is shaped during the MZT. First, we describe the scale and dynamics of maternal control, as well as the processes that influence the maternal gene expression program. These include the regulation of maternal mRNA stability and translation, as well as the modulation of protein stability and functions by post-translational modifications. Then, we discuss how the zygotic gene expression program is activated. This includes roles for cell cycle length, transcriptional repressors and activators, and changes in chromatin organization. Finally, we describe how and why the maternal and zygotic gene expression programs are intertwined with each other to ensure successful handover of developmental control from one generation to the next.

Regulation of maternal gene products Regulation of mRNA stability and clearance

All animal oocytes are loaded with mRNA species representing a large fraction of the protein-coding capacity of their genome (summarized in Table 1). Estimates from global analyses suggest that the maternal transcripts represent a range of the protein-coding transcriptome, from about one-third in the mouse (Wang et al., 2004) andCaenorhabditis elegans(Baugh et al., 2003; Stoeckius et al., 2014), to three-quarters in Drosophila melanogaster (De Renzis et al., 2007; Lécuyer et al., 2007; Tadros et al., 2007; Thomsen et al., 2010), the echinoderm Strongylocentrotus purpuratus (Wei et al., 2006) and the zebrafish, Danio rerio

(Aanes et al., 2011; Harvey et al., 2013). During the MZT, a subset of the maternal transcripts is cleared from the embryo (Boxes 2 and 3). Estimates range from about one-quarter of the maternal transcripts in zebrafish (Aanes et al., 2011; Bazzini et al., 2012; Harvey et al., 2013; Mishima and Tomari, 2016), one-third in

C. elegansand mouse (Baugh et al., 2003; Hamatani et al., 2004), to about two-thirds inDrosophila(Thomsen et al., 2010).

All animal embryos undergo phased maternal mRNA clearance; the initial phase is directed by maternally provided gene products whereas the later phases require zygotically synthesized products. The scale and dynamics of these phases varies across species, with respect to both time and developmental stage (Fig. 1; Table 1). Furthermore, technical differences among studies affect quantification (Boxes 2 and 3). Finally, in animals that set aside primordial germ cells in the early embryo, the MZT differs in soma versus germline (Box 4).

In Drosophila, an early maternally directed wave of decay is triggered upon egg activation and does not require fertilization; this is followed by one or more waves that require transcription of the zygotic genome and are therefore zygotically directed (Bashirullah et al., 2001; Bashirullah et al., 1999; Tadros et al., 2007; Tadros et al., 2003). About 25% of the cleared transcripts are degraded strictly by the maternal machinery, 35% strictly through the zygotic machinery, while 40% show mixed decay effected by both maternal and zygotic mechanisms (Fig. 1A) (De Renzis et al., 2007; Tadros et al., 2007; Thomsen et al., 2010). InC. elegans, about 30% of the maternal mRNAs are targeted for degradation by the one-cell stage, whereas another 30% are degraded starting at the four-cell stage, coinciding with the activation of zygotic transcription (Fig. 1B) (Baugh et al., 2003; Stoeckius et al., 2014).

In the echinodermS. purpuratus, there appear to be three waves of maternal transcript degradation, but the role of maternal versus zygotic gene products in directing clearance is unknown (Fig. 1C) (Tu et al., 2014). In zebrafish, of the cleared maternal transcripts, about 60% degrade between the one-cell and 16-cell stages, while the remaining 40% degrade subsequently, coinciding with zygotic transcription (Aanes et al., 2011; Mathavan et al., 2005; Mishima and Tomari, 2016). Expression profiling in zebrafish lacking TATA-binding protein (Ferg et al., 2007) or treated with α-amanitin (Mishima and Tomari, 2016), and therefore unable to undergo zygotic genome activation, have shown that the latter wave of 1Max Planck Institute of Molecular Cell Biology and Genetics, Pfotenhauerstraße

108, 01307 Dresden, Germany.2Department of Molecular Genetics, University of Toronto, 661 University Avenue, Toronto, Ontario M5G 1M1, Canada.

*Authors for correspondence (vastenhouw@mpi-cbg.de; howard.lipshitz@ utoronto.ca)

N.L.V., 0000-0001-8782-9775; H.D.L., 0000-0002-7372-4419

DEVEL

O

(2)

clearance requires zygotic transcription (Fig. 1D). In Xenopus tropicalis, about one-third of the maternal transcripts are cleared before the major wave of zygotic genome activation (ZGA) (Tan et al., 2013), while 15% are cleared after ZGA (Fig. 1E) (Graindorge et al., 2006).

In the mouse, the onset of the earliest wave of maternal transcript degradation is triggered by completion of meiosis (i.e. before fertilization and thus not pictured in Fig. 1F); a second wave of degradation is triggered by fertilization; and a third, zygotically directed, wave follows the major activation of the zygotic genome after the two-cell stage (Hamatani et al., 2004; Svoboda et al., 2015). Gene expression profiling in human embryos suggests that they too undergo at least two waves of maternal transcript clearance (Fig. 1G) (Dobson et al., 2004; Zhang et al., 2009b).

As maternal mRNAs can be stable for days, weeks, months or years in oocytes but are then degraded rapidly in a matter of hours or, at most, days, understanding this component of the MZT requires elucidation of the mechanisms underlying both long-term transcript stabilization and rapid transcript destabilization. These processes are regulated by RNA-binding proteins, small non-coding RNAs, RNA modifications and‘codon optimality’(Fig. 2).

RNA-binding proteins

RNA-binding proteins (RBPs) regulate either stabilization of stored maternal mRNAs in oocytes or their destabilization during the MZT. The particular RBPs that mediate these processes appear to vary across species (Table 1). However, in general, there have not been any systematic studies aimed at assessing whether RBPs identified as key regulators in one species exert similar effects in others.

In X. laevis, the Y-box RBP FRGY2 binds without sequence specificity to maternal mRNAs and functions to translationally repress and stabilize them (Bouvet and Wolffe, 1994; Matsumoto et al., 1996). Similarly, the zebrafish Y-box protein, Ybx1, globally represses maternal mRNA translation in oocytes (Sun et al., 2018), and in mouse oocytes, the Y-box protein MSY2 stabilizes maternal transcripts (Medvedev et al., 2011; Yu et al., 2001; Yu et al., 2002). Here, upon oocyte maturation, MSY2 is phosphorylated by the CDC2A (CDK1) kinase and MSY-bound mRNAs are released (Medvedev et al., 2008). In C. elegans and Drosophila, Y-box proteins are expressed during oogenesis but a role in transcript stabilization has not yet been identified (Arnold et al., 2014; Boag et al., 2005; Mansfield et al., 2002). Instead, homologs of the

dead-box helicase Dhh1 (CGH-1 inC. elegans, ME31B inDrosophila) have been implicated in stabilization and translational repression of maternal mRNAs (Arnold et al., 2014; Boag et al., 2005; Wang et al., 2017).

The role of RBPs as trans-acting factors that regulate and coordinate maternal transcript degradation has been particularly well studied inDrosophilaembryos (Fig. 2A). Two RBPs, Smaug, a SAM-domain RBP (Chen et al., 2014; Tadros et al., 2007) and Brain tumor (BRAT), a TRIM-NHL RBP (Laver et al., 2015a), regulate the maternally directed wave of decay. Subsequent, zygotically directed decay can be separated into two phases. The first is driven by miRNAs (Benoit et al., 2009; Bushati et al., 2008) (discussed below). The second is mediated by Brain tumor (BRAT) and possibly also Pumilio (PUM), a PUF-domain RBP (Gerber et al., 2006; Laver et al., 2015a,b; Thomsen et al., 2010). Thesmaug

mRNA itself is translationally repressed during oogenesis by PUM and additional unknown factors; repression is relieved upon egg activation by the Pan gu (PNG) serine/threonine kinase (Fig. 3A) (Tadros et al., 2007). Smaug binds to hairpin structures in its target mRNAs (Aviv et al., 2003; Chen et al., 2014; Dahanukar et al., 1999; Semotok et al., 2008; Smibert et al., 1999) and can trigger their degradation by recruiting the CCR4-NOT-deadenylase complex (Fig. 2A) (Semotok et al., 2005). Smaug also represses target mRNA translation via recruitment of Argonaute 1 and the eIF4E-binding protein Cup (Nelson et al., 2004; Pinder and Smibert, 2013). The majority of the targets of Smaug are both repressed and degraded, while smaller subsets are degraded but not repressed, or vice versa (Chen et al., 2014; Nelson et al., 2007; Semotok et al., 2005; Semotok et al., 2008). PUM and BRAT bind to single-stranded motifs in their target transcripts (Laver et al., 2015a; Loedige et al., 2014). BRAT directs degradation of hundreds of maternal transcripts via both the maternal and the zygotic pathways (Laver et al., 2015a,b), whereas PUM appears to participate predominantly in zygotically directed clearance (Gerber et al., 2006; Laver et al., 2015a,b). In the maternal wave of decay, the targets of BRAT do not significantly overlap with those of Smaug (Laver et al., 2015a), suggesting that these two RBPs act independently of each other on distinct subsets of the maternal transcriptome. Interestingly, the smaug mRNA is itself targeted for degradation at the end of the MZT by BRAT, and possibly other factors, including PUM (Fig. 3A) (Laver et al., 2015a,b). Smaug, BRAT and PUM all promote transcript deadenylation (Newton et al., 2015; Semotok et al., 2005; Temme et al., 2010; Weidmann et al., 2014). Thus, these RBPs act as specificity factors that recruit a conserved degradation machinery to clear specific subsets of maternal mRNAs.

Another class of RBPs implicated in maternal mRNA clearance comprises the AU-rich-element binding proteins (ARE-BPs). ARE-BPs have a role in transcript destabilization during the MZT inC. elegans(D’Agostino et al., 2006; Gallo et al., 2008; Schubert et al., 2000), zebrafish (Rabani et al., 2017) andX. laevis(Detivaud et al., 2003; Graindorge et al., 2006; Paillard et al., 1998; Voeltz and Steitz, 1998), and possibly also mouse (Fig. 2A) (Ramos et al., 2004). Although a motif resembling AREs is enriched in

Drosophila mRNAs that are destabilized during the MZT (De Renzis et al., 2007), a role for ARE-BPs has not yet been identified inDrosophila. InXenopus, ARE-BPs appear to function together with the embryonic deadenylation element binding protein (EDEN-BP, also known as CUGBP1) (Graindorge et al., 2008), which directs maternal transcript clearance via the recruitment of the poly(A) ribonuclease (PARN) (Moraes et al., 2006). In zebrafish, a recent study has implicated hnRNPA1, an RBP that binds to the 3′UTR of

Box 1. The maternal-to-zygotic transition versus the mid-blastula transition

The terms ‘maternal-to-zygotic transition’ (MZT) and ‘mid-blastula transition’ (MBT) are often used interchangeably in the literature, leading to confusion. Here, we define the MZT and MBT as two distinct concepts in early embryo development. The MZT describes a coordinated series of molecular events, starting with the degradation of maternally deposited transcripts and ending with global activation of the zygotic genome. The MZT can last from a few hours (as inDrosophila

andC. elegans) to days (as in mouse and human). In contrast, the MBT describes a specific stage during the development of the embryo, which is marked by lengthening and desynchronization of the cell cycles. For example, following rapid and synchronous cleavage cycles, the MBT occurs at cycle 10 in zebrafish, at cycle 12 inXenopusand at cycle 14 in

Drosophila. There are several cases where the MBT does not coincide with completion of the MZT. As examples, inAscaris suumthe MZT is completed long before its MBT (Wang et al., 2014), whereas, in

Drosophila, the MBT is completed before the final wave of maternal mRNA decay (Laver et al., 2015a,b).

DEVEL

O

(3)

0 0

0 12 24

1

36 2 Cleavage cycle

Time (h) F Mouse

Mus musculus

0 3

0 2 4

6

6 13 Cleavage cycle

Time (h) E Frog

Xenopus tropicalis Cleavage cycle Time (h)

0

0 1

8

2 13

3 14 A Insect

Drosophila melanogaster

1.6 2.4

0.8 0

Cleavage cycle

Time (h) 0

1 3 5

B Nematode

Caenorhabditis elegans

D Fish Danio rerio

8

3.7 12 2.3

2 0

5.3 14

0 1

Cleavage cycle Time (h)

0 0

0 16 32

1

48 2 Cleavage cycle

Time (h) G Human

Homo sapiens

16 24

8 1

Cleavage cycle

Time (h) 0

~7 ~8 ~10

[image:3.612.75.534.54.679.2]

C Echinoderm Strongylocentrotus purpuratus

Fig. 1. The maternal-to-zygotic transition (MZT).Red curves depict destabilized maternal transcripts; red gradients represent timing of the multiple phases of

transcript degradation. Blue curve depicts transcription from the zygotic genome; blue gradient represents a gradual increase in gene numbers transcribed during ZGA. Schematics depicting embryo developmental stages for each organism are shown above the corresponding cleavage cycle and time, after fertilization. (A)Drosophila melanogaster, (B)Caenorhabditis elegans, (C)Strongylocentrotus purpuratus, (D)Danio rerio, (E)Xenopus tropicalis, (F)Mus musculusand (G)Homo sapiens. The first wave of decay in mammals occurs prior to fertilization and is, therefore, not depicted. Adapted, with permission, from Tadros and Lipshitz (2009).

DEVEL

O

(4)

T able 1. Scale and regula tion of ma ternal tr anscripts Species Ma ternally loaded mRNA co v er age (% of genome) Stabiliza tion and/or repr ession of mRNAs in the oocyte mRNA clear ance (% of ma ternal ) Mechanism of ma ternal mRNA clear ance RNA-bi n di n g and/ o r associ a te d p ro te in s Small RNAs C odon use RNA modifica tions C. elegans ∼ 35% (Bau gh et al., 2003; Stoe ckius et al., 2014) CGH-1 (A rn ol d e t a l., 2 01 4; Bo ag e t al ., 20 05 ) ∼

30% (Bau

gh e t a l., 20 03 ; S toe c k ius e t al., 2 014 ) MEX-5/6 (ARE-BP) (D ’ Agos tino et al., 2006 ; Gallo et al., 2008; Schubert et al., 2000) Endo-siRN A* (Sto ec kius e t al., 20 14) Not s tudied Not s tudied D. melanogas ter ∼

75% (Ta

[image:4.612.141.480.58.741.2]
(5)

maternal transcripts through an AGGGA motif, in clearance of a subset of maternal mRNAs after ZGA (Despic et al., 2017).

In the mouse, the oocyte-expressed protein BTG4, although not itself an RBP, links eIF4E to the CCR4-NOT-deadenylase complex to trigger degradation of maternal mRNAs that are actively translated (Fig. 2A) (Yu et al., 2016). Embryos from female Btg4−/− mice

produce oocytes that complete meiosis but fail to clear maternal transcripts. Reminiscent of the case forDrosophila smaugmutants, zygotically transcribed genes are downregulated and the embryos fail to develop to the two-cell stage (Yu et al., 2016). Furthermore, also

reminiscent of smaug mRNA translation at egg activation being dependent on the PNG kinase (Tadros et al., 2007), translation of BTG4 protein is triggered by activation of the MAP kinases ERK1 and ERK2 during meiotic maturation (Yu et al., 2016).

Small non-coding RNAs

Small non-coding RNAs that have been implicated in the MZT include microRNAs (miRNAs), which are processed from primary transcripts produced by miRNA-encoding genes (reviewed by Gebert and MacRae, 2019), PIWI-interacting RNAs (piRNAs; PIWI, P-element A RNA-binding and/or RNA-associated proteins

B Small RNAs

C RNA modifications

D Codon use

D. melanogaster D. rerio M. musculus

N6-methyladenosine m7G

AAAAA

m6A YTHDF2

?

N6-methyladenosine

m7G

AAAAA

m6A YTHDF2

? ?

Endo-siRNA RISC m7G

AAAAA SMG

AAAAA SRE

CCR4/NOT m7G

PUM AAAAA PUM motif

CCR4/NOT

m7G

AAAAA BRAT

BRAT motif CCR4/NOT m7G

*

hnRNPA1

AAAAA AGGGA

m7G eIF4G

BTG4 m7G

eIF4E

AAAAA CCR4-NOT

*

ARE-BP

AAAAA AU-rich element

m7G

Terminal uridylation m7G

AAAAAUU

TUT4/7

Terminal uridylation m7G

AAAAAUU

TUT4/7

Slow ribosomes m7G

AAAAA ME31b?

Slow ribosomes m7G

AAAAA DDX6?

Slow ribosomes m7G

AAAAA DDX6? miR-309

Other miRs? piRNAs? AGO1

RISC

m7G

AAAAA CCR4/NOT

miR-430 RISC m7G

AAAAA CCR4/NOT

*

ARE-BP

AAAAA UUA/poly-U

m7G

*

PUM

AAAAA PUM motif

m7G

*

ARE-BP

AAAAA AU-rich element

m7G

D. melanogaster D. rerio M. musculus

D. melanogaster D. rerio M. musculus

[image:5.612.53.417.57.612.2]

D. rerio M. musculus

Fig. 2. Mechanisms of maternal transcript

degradation in model organisms.Several

mechanisms directing maternal transcript degradation are compared across

D. melanogaster,D. rerioandM. musculus. (A) The best-studied RNA-binding proteins (RBPs) for each organism are shown. Smaug (SMG), Pumilio (PUM), Brain tumor protein (BRAT), AU-rich binding element protein (ARE-BP) and Heterogeneous nuclear ribonucleoprotein A1 (hnRNPA1). Their homologs across organisms have largely not been studied and are omitted from this figure. B-cell translocation gene 4 (BTG4) in mouse, though not an RBP, is shown because it triggers transcript degradation by recruiting the CCNR4-NOT-deadenylase complex.

(B) MicroRNAs direct maternal transcript decay via the Argonaute (AGO1)/RNA-induced silencing complex (RISC) inDrosophilaand zebrafish; they are present but may not play a role in the mouse (therefore, not shown) where endo-siRNAs may instead function. (C) RNA modifications, such asN6-methyladenosine (m6A) and terminal uridylation, promote decay

through the action of YTH N6-Methyladenosine RNA Binding Protein 2 (YTHDF2) and Terminal Uridylyl Transferase 4/7 (TUT4/7), respectively. Analogous modifications have been found in

D. melanogaster, but their functions during the MZT have not been investigated. (D) Codon optimality directs maternal transcript stability in

D. melanogaster,D. rerioandM. musculus, but the mechanism for this regulation has only been studied in yeast thus far; homologs of yeast Dhh1p (ME31B, DDX6) are shown. Asterisks indicate enrichment of cis-elements, whereas the RBPs have not yet been investigated. Question marks indicate where trans-factors are known but mechanisms have not been studied in detail (Table 1).

DEVEL

O

(6)

induced wimpy testis), which are derived from transposable elements resident in the genome (reviewed by Czech et al., 2018), and endogenous small interfering RNAs (endo-siRNAs), which are derived from double-stranded hybrids of mRNAs or transposable elements (reviewed by Piatek and Werner, 2014) (Table 1).

The most well studied are miRNAs, which play a crucial role in directing the zygotic wave of maternal transcript decay in multiple species (Fig. 2B). In zebrafish, themiR-430family of miRNAs is expressed during early ZGA and is required for translational repression, deadenylation and decay of about 40% of the maternal transcripts that are cleared during the MZT (Fig. 2B) (Bazzini et al., 2012; Giraldez et al., 2006). An orthologous mechanism exists during the X.laevis MZT through its miR-430ortholog, miR-427. As in zebrafish, miR-427 is expressed zygotically and mediates the deadenylation and clearance of maternal transcripts (Lund et al., 2009). An analogous pathway exists in Drosophila, mediated by zygotic transcription of miRNAs from themiR-309cluster, which are required for the degradation of 14% of transcripts undergoing the late wave of decay (Fig. 2B) (Bushati et al., 2008). Zygotic expression of more than 70 species of miRNAs, including the miR-309family, depends on the Smaug RBP (Benoit et al., 2009; Luo et al., 2016). In

smaug mutants, failure to produce these miRNAs results in stabilization of a large number of maternal transcripts that are normally cleared in the zygotic wave of decay (Benoit et al., 2009; Luo et al., 2016). Thus, Smaug functions directly to clear maternal mRNAs prior to ZGA and indirectly to clear maternal transcripts after ZGA (Fig. 3A). TheDrosophilazygotic miRNAs do not bear sequence homology to vertebratemiR-430/miR-427, suggesting that microRNA targeting of the maternal transcriptome during the MZT may have arisen independently in the vertebrate and invertebrate lineages.

piRNAs have also been implicated in maternal transcript decay in

Drosophila(Barckmann et al., 2015; Dufourt et al., 2017; Rouget et al., 2010) where hundreds of maternal mRNAs have been reported to co-purify with PIWI, Argonaute 3 and Aubergine, and to be targeted by the piRNA pathway for degradation (Barckmann et al., 2015).

Endo-siRNAs (‘22G-RNAs’) have been shown to clear a subset of maternal mRNAs from earlyC. elegansembryos (Stoeckius et al., 2014). Although miRNAs and endo-siRNAs are present, miRNA activity appears to be suppressed in mouse oocytes and early embryos, and removal of DGCR8, which is essential for miRNA biogenesis, has no effect on preimplantation development (Ohnishi et al., 2010; Suh et al., 2010; Tang et al., 2007). It is possible

that endo-siRNAs serve as regulators of maternal mRNA clearance (Fig. 2B) (Lykke-Andersen et al., 2008); however, conditional removal of AGO2 from mouse oocytes results in failure to complete meiosis, preventing analyses in early embryos (Stein et al., 2015).

RNA modifications

Over the past few years, it has become clear that chemical modifications of mRNAs either during or after their synthesis can play a role in determining their fate.N6-methylation of adenosine (m6A), for example, can cause a structural switch in RNA that

controls accessibility to protein-binding motifs (Liu et al., 2015). In zebrafish embryos, more than one-third of maternal mRNA species are methylated on adenosine (Zhao et al., 2017). The m6

A-binding protein YTHDF2 is required for clearance of almost half of the maternal mRNAs that are degraded early in the MZT, as well as over one-quarter of those that are degraded later (Fig. 2C) (Zhao et al., 2017). Maternal YTHDF2 is also required for clearance of a subset of the transcriptome of the mouse oocyte (Ivanova et al., 2017). However, at least in zebrafish, not all methylated RNAs are destabilized, and not all transcripts that are degraded in a YTHDF2-dependent manner are methylated (Zhao et al., 2017). The exact mechanisms of regulation and the interplay between m6A-directed decay, and other pathways of transcript clearance,

remain to be elucidated.

Another form of mRNA modification during the vertebrate MZT is terminal uridylation of the poly(A) tail of an mRNA (Chang et al., 2018). In zebrafish,Xenopusand mouse, terminal uridylation of the transcriptome increases greatly during the MZT and promotes rapid decay of a subset of transcripts with short poly(A) tails (Fig. 2C) (Chang et al., 2018).

Codon optimality

In unicellular organisms,‘optimal’codons are defined as those with high cognate tRNA abundance, whereas non-optimal codons for the same amino acid have low cognate tRNA levels (Hanson and Coller, 2018). In yeast, the DEAD-box protein Dhh1p preferentially targets transcripts enriched for non-optimal codons by associating with

Box 2. Gene expression analysis

To precisely measure changes in RNA levels during the MZT, several gene expression analysis methods can be used, each of which has its their own benefits and caveats. RNA sequencing provides genome-wide quantitative information. However, as regulation of poly(A) tail length is an important component of the MZT, the most accurate quantification of RNA levels requires total RNA-sequencing after depletion of rRNAs rather than poly(A) RNA-sequencing, which biases reads towards transcripts with long poly(A) tails and, therefore, underestimates the number of transcript species present (Collart et al., 2014). An alternative to RNA sequencing is the Nanostring approach, which provides information on fewer transcripts but allows for precise quantification without the need to fragment, amplify or reverse transcribe the RNA (Geiss et al., 2008). Nanostring analysis has proven especially useful in detecting low-abundance zygotic transcripts against a background of enormous amounts of maternal RNA (Joseph et al., 2017; Sandler and Stathopoulos, 2016), which is often hard to achieve using (non-targeted) RNA-sequencing methods.

Box 3. Temporal and spatial analysis of the MZT

To distinguish between maternal and zygotic transcripts that are often present at the same time, zygotic contributions have been removed by transcription inhibition (Lee et al., 2013), the use of chromosomal deficiencies (De Renzis et al., 2007) or the analysis of activated unfertilized eggs (Bashirullah et al., 2001; Bashirullah et al., 1999; Tadros et al., 2007; Tadros et al., 2003). Conversely, to analyze zygotically transcribed RNA, newly synthesized RNAs have been specifically labeled, isolated and sequenced (Chan et al., 2018 preprint; Heyn et al., 2014), single nucleotide variations between strains have been used to identify transcription from the paternal genome (Harvey et al., 2013; Lott et al., 2011) or intron signal has been used as a marker ofde novotranscription (Lee et al., 2013). The analysis of transcript levels in ( pools of ) whole embryos provides transcript information averaged over all cells of the embryo. Spatial information can be obtained by sequencing embryo sections (Combs and Eisen, 2013; Junker et al., 2014) or single-cell sequencing of dissociated cells (Briggs et al., 2018; Farrell et al., 2018), which can be mapped back onto the embryo based on markers with a known expression pattern. Alternatively, transcripts can be visualized either in fixed or live embryos (Campbell et al., 2015; Perez-Romero et al., 2018; Stapel et al., 2018; Stapel et al., 2016). Although low throughput, this provides the best spatial information and can be quantitative when single transcripts can be detected (Stapel et al., 2016).

DEVEL

O

(7)

slow-moving ribosomes and directing these transcripts for degradation (Presnyak et al., 2015; Radhakrishnan et al., 2016). Studies in zebrafish have found preferential degradation during the MZT of transcripts enriched for non-optimal codons (Bazzini et al., 2016; Mishima and Tomari, 2016), a process that may also occur in mouse,

XenopusandD. melanogaster(Bazzini et al., 2016) (Fig. 2D). It is not yet known whether proteins homologous to Dhh1p (DDX6 in mouse and fish, and ME31B inDrosophila) function in this process.

Regulation of maternal mRNA translation

DuringDrosophilaegg activation (i.e. at the beginning of its MZT), translational efficiency increases for up to 50% of maternal mRNAs and decreases for up to 20%, as measured by ribosome footprinting (Kronja et al., 2014b). The translation of maternal transcripts has also been profiled in mouse, where 20% of the maternal transcriptome is significantly depleted from polysomes during oocyte maturation, whereas 17% of the maternal transcriptome is significantly enriched (Chen et al., 2011). There is a direct correlation between poly(A) tail length and translational efficiency in early embryos of zebrafish,X. laevisandDrosophila(Eichhorn et al., 2016; Subtelny et al., 2014). These correlations diminish upon ZGA and are not found in non-embryonic tissue types across several species (Subtelny et al., 2014). Thus, the poly(A) tail may play a role in the regulation of translation primarily in transcriptionally silent stages.

The mechanisms of cytoplasmic polyadenylation and activation of maternal mRNA translation were first shown inX. laevisoocytes. Upon maturation, the Aurora kinase triggers binding of cytoplasmic polyadenylation element binding protein (CPEB) to cytoplasmic polyadenylation specificity factor (CPSF) and the mRNA. Together, they displace the PARN deadenylase and recruit the GLD2 poly(A) polymerase (Barnard et al., 1993; Cao and Richter, 2002; Groisman et al., 2002; Hake and Richter, 1994; Kim and Richter, 2006; Stebbins-Boaz et al., 1999). This results in poly(A) tail lengthening and translational activation. After fertilization, a subset of maternal transcripts undergoes CPEB-mediated polyadenylation in the embryo prior to zygotic transcription, likely regulating the first wave of translational activation (Collart et al., 2014). However, the mechanisms regulating polyadenylation in theX. tropicalisembryo

have yet to be determined. The CPEB protein family also coordinates cytoplasmic polyadenylation and translation in zebrafish, where more than 40% of maternal mRNAs undergo cytoplasmic polyadenylation in early embryos (Winata et al., 2018). The translation of these transcripts increases during the MZT, whereas there is a decrease in translation of transcripts lacking elongated poly(A) tails (Winata et al., 2018). CPEB1 also functions during mouse oocyte maturation; ERK1/2 phosphorylation of CPEB1 causes polyadenylation-mediated translation of BTG4 and other essential maternal proteins, such as the cell cycle regulator, Cyclin B and the DAZL RBP (Sha et al., 2017). DAZL, in turn, directs translational activation during oocyte maturation (Chen et al., 2011; Fukuda et al., 2018). Autoregulation of its own mRNA by DAZL establishes a positive feedback loop together with CPEB1, which ensures the progression of meiosis and proper translational activation (Sousa Martins et al., 2016).

In early Drosophila embryos, the Wispy (GLD2) cytoplasmic poly(A) polymerase directs polyadenylation of maternal mRNAs and is essential for the MZT (Benoit et al., 2008; Cui et al., 2008; Cui et al., 2013; Salles et al., 1994). However, relative changes in poly(A) tail length and translational efficiency are retained even in the absence of Wispy. Thus, the current hypothesis is that selective shortening, rather than lengthening of poly(A) tails serves as the major determinant of maternal mRNA translation at this stage (Eichhorn et al., 2016). Furthermore, inDrosophila, PNG triggers polyadenylation and translation at the beginning of the MZT (Fig. 3A) (Tadros et al., 2007; Vardy and Orr-Weaver, 2007). More than 60% of the maternal mRNAs that undergo either increases or decreases in translation depend on the PNG kinase for these processes (Kronja et al., 2014b). The Smaug RBP serves as a major mediator of PNG-dependent repression, with most of this repression attributable to poly(A) tail shortening (Eichhorn et al., 2016). However, cytoplasmic polyadenylation is not sufficient to rescue translation in png mutants, presumably because PNG must also relieve repression by specific RBPs such as PUM (Tadros et al., 2007; Vardy and Orr-Weaver, 2007).

Cytoplasmic polyadenylation and deadenylation also play key roles in regulation of mRNA stability and translation inC. elegans

(Fig. 3B) (Millonigg et al., 2014; Nousch et al., 2017; Nousch et al., 2013; Nousch et al., 2014). During oocyte maturation and the MZT, the translational status of hundreds of transcripts is coordinated by LIN-41, a TRIM-NHL RBP related toDrosophilaBRAT. Acting antagonistically to LIN-41 are the RBPs, OMA-1 and OMA-2, which are related to the Tristetraprolin/TIS-11 ARE-BP (Tsukamoto et al., 2017). LIN-41 and the OMAs also interact with both the GLD-2 poly(A) polymerase and the CCR4-NOT-deadenylase (Tsukamoto et al., 2017). Combinatorial action of LIN-41, OMAs and GLD-2 in distinct RNP complexes may direct a repression-to-activation switch for specific target transcripts, which in turn encode RBPs that function in early embryogenesis (Tsukamoto et al., 2017).

In summary, during the MZT, cytoplasmic polyadenylation and deadenylation affect both the stability and translation of mRNAs. While the machinery for these processes is conserved, the particular RBPs that confer specificity on subsets of transcripts vary across species. Likewise, whereas protein kinases play a key role in triggering translational derepression of maternal mRNAs, the identity of these kinases varies across species.

Post-translational regulation of maternal proteins

Despite the documented importance of post-transcriptional regulation of maternal transcripts at the level of mRNA translation and stability,

Box 4. The MZT in primordial germ cells

In animals that set aside primordial germ cells (PGCs) in their early embryos, there is additional spatial and temporal control of the MZT in the PGCs relative to the soma. InC. elegansthe smaller, posterior (P) product of the first cell division gives rise to the PGCs. ZGA is initiated by the four-cell stage in the soma, while transcription remains silenced in the P-cell lineage due to the presence of a global transcriptional repressor, PIE-1 (Seydoux et al., 1996). The germ plasm at the posterior of the

Drosophila embryo buds off to give rise to PGCs within which both maternal transcript degradation and ZGA are delayed relative to the soma (Siddiqui et al., 2012). In the PGCs, as in the soma, both maternal mRNA clearance and ZGA are abrogated insmaugmutants (Siddiqui et al., 2012). Transcriptional repression in theDrosophilaPGCs depends on the Polar Granule Component protein, PGC, which blocks transcription elongation (Nakamura et al., 2019; Hanyu-Nakamura et al., 2008). In echinoderms, the PGCs (small micromeres) are formed from asymmetric divisions by the 32-cell stage. Cleared maternal transcripts in the somatic cells are‘protected’ in the small micromeres due to degradation of theCnot6transcript, which encodes an essential component of the CCR4-NOT-deadenylase complex (Oulhen and Wessel, 2016; Wessel et al., 2014). In zebrafish, PGC-expressed nanos1 and tdrd7 are protected from miR-430-mediated transcript degradation by the germline-expressed RBP Dead end 1 (Dnd1) (Kedde et al., 2007; Mishima et al., 2006).

DEVEL

O

(8)

global transcriptome and proteome analyses prior to and during the MZT have revealed that post-translational regulation of protein stability and function also play a major role inDrosophila(Becker et al., 2018; Casas-Vila et al., 2017; Kronja et al., 2014a) andC. elegans(Stoeckius et al., 2014). For example, inC. elegansmore than one-quarter of the transcriptome is downregulated at least twofold in early embryos, whereas only 5% of the proteome shows a similar decrease (Stoeckius et al., 2014). These results highlight the importance of an additional level of regulation at the post-translational level, through modifications that affect maternal protein stability and function, which is essential for the progression of oocyte maturation and the MZT (Liu et al., 2018a).

One highly conserved mechanism of post-translational regulation that combines phosphorylation and ubiquitin-dependent proteolysis occurs at the beginning of the MZT (reviewed by Pesin and Orr-Weaver, 2008). Most animals are arrested during meiosis by the presence of Cyclin B. Release from meiotic arrest is achieved by the anaphase-promoting complex (APC), a cyclin-dependent E3 ubiquitin ligase complex that is phosphorylated and activated upon fertilization to target Cyclin B for degradation.

About 30% of the oocyte proteome is estimated to change in phosphorylation state duringDrosophilaegg activation (Krauchunas et al., 2012). The PNG kinase complex, which plays a key role in post-transcriptional regulation of maternal mRNAs, undergoes such a change (Fig. 3A) (Hara et al., 2017). InDrosophilaoocytes, the regulatory subunit, GNU, is phosphorylated by CYCB/CDK1, which blocks the binding of GNU to the kinase subunit. Degradation of CYCB by the APC during meiotic completion triggers dephosphorylation of GNU, which then binds to and activates the PNG kinase subunit. Subsequently, PNG phosphorylates GNU, leading to GNU degradation and the inactivation of the complex (Hara et al., 2017). This temporally coordinated activation and self-inactivation tightly restricts the activity of this kinase to a short window during the early MZT, ensuring precise post-translational control of its phosphorylation targets. The PNG kinase is essential for translation of the Smaug RBP in the early embryo (discussed above). However, Smaug is rapidly cleared a couple of hours later, at the end of the MZT (Benoit et al., 2009). Indeed, towards the end of the MZT, not only Smaug but also additional global post-transcriptional regulators, such as ME31B, Cup and TRAL, undergo among the largest decreases in abundance in the entire proteome (Sysoev et al., 2016; Wang et al., 2017). Although the mechanisms that direct clearance of these proteins are not yet known, they represent a striking example of how translational regulation has a major impact on post-transcriptional regulation during the MZT.

Another example of post-translational regulation comes from

C. elegans(Fig. 3B). In early embryos, maternally supplied OMA-1 and OMA-2 are phosphorylated by the DYRK kinase MBK-2, which allows binding to the transcription factor TAF-4, thus sequestering it in the cytoplasm and preventing premature ZGA (Guven-Ozkan et al., 2008). By the four-cell stage, phosphorylated OMA-1 and OMA-2 are ubiquitylated by accumulating levels of the Cullin-RING E3 ubiquitin ligase (CRL), leading to degradation of the OMAs and release of TAF-4, which is then imported into the nucleus where it directs zygotic transcription (Guven-Ozkan et al., 2008). Post-translational regulation is important for inactivation and clearance of additional post-transcriptional regulators during the

C. elegans MZT. The LIN-41, GLD-1 and CPB-3 RBPs are all cleared by the SCFSEL-10E3 ubiquitin ligase (Kisielnicka et al., 2018;

Spike et al., 2018). However, the signaling pathways that trigger inactivation and clearance of these RBPs differ: the CDK1 kinase

pathway clears LIN-41, while the MAP kinase pathway clears GLD-1 (Spike et al., 2018).

In the mouse pre-implantation embryo, massive degradation of maternal proteins by autophagy is triggered by fertilization and is required for progression of early embryogenesis (Tsukamoto et al., 2008). While proteomic studies have identified the ubiquitin-proteasome pathway to be important during pre-implantation development as a whole (Zhang et al., 2009a), the first specific role for this pathway during the MZT has only recently been characterized. The TAB1 kinase inhibits the action of the NFκB transcription factor by retaining it in the cytoplasm. In early embryos, the E3 ubiquitin ligase RnF114 directs degradation of TAB1, permitting translocation of NFκB into the nucleus and, thus, activation the NFκB pathway, which is essential for development (Yang et al., 2017). Indeed, either knockdown of RnF114 or persistence of TAB1 protein prevents development beyond the two-cell stage (Yang et al., 2017).

Regulation of zygotic transcription Scale and dynamics of zygotic transcription

The second act of the MZT is the onset of zygotic transcription. Historically, two transcriptional waves have been distinguished: a minor wave that occurs during the cleavage divisions; and a major wave that, in many species, coincides with the lengthening of the cell cycle. Genome-wide gene expression analysis, however, has revealed that ZGA does not consist of two distinct waves, but instead reflects a period over which transcription is gradually activated (Fig. 1) (Aanes et al., 2011; Collart et al., 2014; Harvey et al., 2013; Heyn et al., 2014; Lott et al., 2011; Owens et al., 2016; Pauli et al., 2012; Sandler and Stathopoulos, 2016; Tan et al., 2013; White et al., 2017).

In terms of the embryonic mitotic cell cycles, human, mouse and sea urchin start transcription the earliest, with the first zygotic transcripts being detected at the one-cell stage (Abe et al., 2015; Aoki et al., 1997; Gildor and Ben-Tabou de-Leon, 2015; Hamatani et al., 2004; Materna et al., 2010; Yan et al., 2013) (Table 2). InC. elegans,Xenopus, zebrafish andDrosophila, the first zygotic transcripts are detected in cell cycles 2, 3, 6 and 8, respectively (Chan et al., 2018 preprint; Collart et al., 2014; De Renzis et al., 2007; Edgar et al., 1994; Hadzhiev et al., 2019; Heyn et al., 2014; Hilbert et al., 2018 preprint; Kimelman et al., 1987; Kwasnieski et al., 2019 preprint; Lécuyer et al., 2007; Lott et al., 2011; Mathavan et al., 2005; Owens et al., 2016; Paranjpe et al., 2013; Seydoux and Fire, 1994; Skirkanich et al., 2011; Tan et al., 2013; Yanai et al., 2011; Yang et al., 2002), although there is some genetic evidence that transcription may begin earlier inDrosophila (Ali-Murthy et al., 2013). In terms of absolute time, the very rapid cleavage cycles of the early Drosophila,Xenopusand zebrafish embryos mean that, in these species, ZGA begins several hours earlier than in the mouse (Fig. 1).

During genome activation, a significant fraction of all genes in the genome is transcribed: 5% inX. tropicalis(Collart et al., 2014; Tan et al., 2013),∼10% inC. elegans(Baugh et al., 2003),∼20% in mouse (Hamatani et al., 2004), ∼25% in zebrafish (Aanes et al., 2011; Harvey et al., 2013; Lee et al., 2013) and∼35% inDrosophila

(De Renzis et al., 2007; Kwasnieski et al., 2019 preprint; Lécuyer et al., 2007; Lott et al., 2011) (Table 2). Although different sets of genes are activated in different species (Heyn et al., 2014), the proteins these genes encode are often enriched for transcription factors and other developmental regulators, such as microRNAs (Collart et al., 2014; De Renzis et al., 2007; Lee et al., 2013). Some of these are important for the degradation of maternally loaded

DEVEL

O

(9)

RNAs (Bushati et al., 2008; Giraldez et al., 2006; Lund et al., 2009), and others for the transcription of subsequently activated genes (Collart et al., 2014). For most genes, the onset of transcription during ZGA is stochastic, which means that not all cells start to activate that gene at the same time (Boettiger and Levine, 2009; Stapel et al., 2017). This initially results in large cell-to-cell differences in gene expression levels, which could–in principle–

hamper development (Lagha et al., 2013). Uniform patterns of gene expression, however, are reached by spatial averaging, such as in the

Drosophilasyncytium (Little et al., 2013), or temporal averaging, as in zebrafish (Stapel et al., 2017), thus overcoming this problem.

Although some of the genes that are activated during ZGA are strictly zygotic (i.e. not maternally provided), many others are loaded maternally and then re-expressed in the embryo. Developmental time

B Caenorhabditis elegans

A Drosophila melanogaster

Degradation

by SCFSel-10

Phosphorylation by MBK OMA1/2

TAF-4

TFIID OMA1/2

P Degradation

by CRL E3

TAF-4

TFIID

ZGA LIN-41

OMA1/2 LIN-41

Cytoplasm

Nucleus Oocyte transcripts

Maternal smg mRNA

PUM X

ZGA Cytoplasm

Nucleus Zygotic miRNAs

PNG

PLUGNU PNG

PLU GNUP

Activation by dephosphorylation

SMG

Inactivation by self-phosphorylation

PNG

PLU GNUPP

PLU GNU PNG

Maternal mRNAs

mRNA degradation

Translation of transcriptional repressors

Repressor

Maternal smg mRNA

Indirect SMG target mRNAs Direct SMG target mRNAs

mRNA degradation

Maternal smg mRNA

PUM? BRAT

Translational repression

mRNA degradation BRAT?

PUM

[image:9.612.83.528.61.522.2]

X BRAT? Y

Fig. 3. Interplay of post-transcriptional and post-translational regulation during the maternal-to-zygotic transition (MZT).(A) InD. melanogaster,

the RNA-binding protein (RBP), Pumilio (PUM) and additional factors, possibly including Brain tumor protein (BRAT) and an unknown factor, X, repress translation of thesmaugmRNA in the oocyte. The PNG-PLU kinase complex is inactive due to phosphorylation-mediated dissociation of its regulatory subunit, GNU. Upon egg activation, dephosphorylation of GNU activates PNG, directing derepression ofsmaugtranslation. The PNG kinase phosphorylates GNU, inactivating the complex and targeting it for degradation. Accumulating Smaug (SMG) protein directs the degradation and translational repression of a large number of maternal transcripts, allowing zygotic genome activation (ZGA). Early ZGA includes transcription of zygotic miRNAs, which direct the degradation of additional maternal transcripts. Thesmaugtranscript itself is targeted for degradation near the end of the MZT by BRAT and possibly additional factors, such as PUM and unknown factor Y. (B) InC. elegans, LIN-41, OMA-1 and OMA-2 act antagonistically to regulate transcripts in the oocyte. LIN-41 is degraded during oocyte maturation by the E3 ubiquitin ligase complex SCFSel-10. Additionally, phosphorylation of OMA-1 and OMA-2 by the MBK-2 kinase during

oocyte maturation results in binding and sequestering of the TAF-4 transcription factor in the cytoplasm, inhibiting ZGA in the early embryo. By the four-cell stage, degradation of phosphorylated OMA-1 and OMA-2, by accumulating levels of the CRL E3-ubiquitin ligase, releases TAF-4, which is then imported into the nucleus where it binds the transcriptional factor TFIID to mediate ZGA. Adapted, with permission, from Tadros and Lipshitz, 2009.

DEVEL

O

(10)

T able 2. Scale and regula tion of zygotic tr anscripts Species Firs t zygotic tr anscripts detected (cell/nuclear cycle number) Zygotically e xpr essed mRNA co v er age (% of genes)*

Strictly zygotically expr

essed mRNA co v er age (% of zygotic) F a ctors inv olv ed in ZGA C ell cycle length Gener al tr anscriptional repr essors

Specific transcriptional repr

essors T ranscriptional a ctiva tors Chr oma tin s tructur e ‡ C. elegans C ell cycle 2 (Edgar et al., 1994; Se ydoux and Fir e, 1994) ∼ 10% (Baugh et al., 2003) Not s tudied Not s tudied Not s tudied Not s tudied T AF-4 (Guv en-Ozkan et al., 2008) Not s tudied D. melanogas ter Nuclear cycle 8 § (De R enzi s et al., 2007; Lécuy er et al., 2007; Lott et al., 2011 ) ∼ 35% (DeR enzi s et al., 2007; Lécuy er et al., 2007; Lott et al., 2011; Ka wsnieski et al., 2019 pr eprin t) ∼ 33% (De R enzi s e t al., 2007) No (F arr ell and O ’ F arr ell 2013; McCleland and O ’ F ar ell, 2008; Blythe and W ieschaus, 2015; Pritchard and Schubiger, 1996) Not s tudied TTK (Br o wn and Wu, 1993; Pritchard and Sch ubiger, 1996) Zelda (Harrison et al., 2011 ; Liang et al., 2008; Nien et al., 2011) Not s tudied D. rerio C ell cycle 6 (Chan et al., 2018 pr eprin t; Hadzhiev et al., 2019; He yn et al., 2014; Hilbert et al., 2018 pr eprint; Ma tha van et al., 2005 ) ∼ 25% (Aanes et al., 2011; Harv e y et al., 2013; Lee et al., 2013 ) ∼ 25% (Lee et al., 2013) No (Dalle Nogar e et al., 2009 ; Zhang et al., 2014a) His tones (Joseph et al., 2017) Not s tudied P OU5F3, SOX19B and NANOG (Lee et al., 2013; Leichsen ring et al., 2013; Joseph et al., 2017) H3K27Ac (Chan et al., 2018 pr eprint; Zhang et al., 2018; Sa to et al., 2019 pr eprint) Xenopus C ell cycle 3 ( X. laevis ) (Y anai et al., 2011 ; Kimelman et al., 1987; Skirkanich et al., 2011 ; Y ang et al., 2002 ) C ell cycle 3 ( X. tr opicalis ) (Y anai et al., 2011 ; P ar anjpe et al., 2013 ; Ow ens et al., 2016; C ollart et al., 2014; T an et al., 2013) ∼

5% (X.

tr opicalis ) (C ollart et al., 2014; T an et al., 2013) Not s tudied Y

es (X.

[image:10.612.113.504.47.741.2]
(11)

In Drosophila, for example, only about one-third of the zygotic transcripts are purely zygotic (De Renzis et al., 2007) and in zebrafish 25% (Lee et al., 2013) (Table 2). In its simplest form, zygotic transcription of genes that are also maternally loaded provides an opportunity to reinforce the expression of specific genes. In Drosophila and zebrafish, however, many genes encode distinct maternal and zygotic mRNA isoforms through differential promoter usage, splicing and/or polyadenylation site usage (Aanes et al., 2013; Atallah and Lott, 2018; Haberle et al., 2014). In these cases, transcription does not just reinforce gene expression, it also changes the characteristics of transcripts and therefore their regulation. Functionally, maternal mRNA degradation and zygotic transcription are used to generate

localized patterns of expression (Combs and Eisen, 2013; De Renzis et al., 2007; Lécuyer et al., 2007; Vopalensky et al., 2018).

Mechanisms of zygotic genome activation

Next, we explore the effect of cell cycle length, transcriptional repressors, transcriptional activators, and chromatin accessibility on the change from transcriptional repression to transcriptional activation (Table 2). We end by proposing a mechanism for the onset of transcription that takes all of these aspects into account (Fig. 4).

Cell cycle length

The early stages of development in insects, amphibians and fish are characterized by rapid cleavage divisions. Because DNA replication generally interferes with transcription (Rothe et al., 1992; Shermoen and O’Farrell, 1991), rapid cell cycles may reduce transcriptional output during the early stages of development. Indeed, lengthening the cell cycle causes a premature onset of transcription in X. laevis (Collart et al., 2013; Kimelman et al., 1987), suggesting that genome activation may be a consequence of cell cycle lengthening. However, similar experiments in zebrafish andDrosophilahave not caused premature transcription (Farrell and O’Farrell, 2013; McCleland and O’Farrell, 2008; Zhang et al., 2014a). In fact, in Drosophila embryos, cell cycle lengthening depends on the onset of zygotic transcription (Blythe and Wieschaus, 2015; Pritchard and Schubiger, 1996) arguing against a dependence of transcription on cell cycle lengthening in these species. Thus, it remains unclear to what extent the lengthening of the cell cycle influences the onset of transcription during embryogenesis. What is clear, is that cell cycle length affects the length and number of transcripts that are produced in

Drosophila and zebrafish embryos (Edgar and Schubiger, 1986; Hadzhiev et al., 2019; Rothe et al., 1992; Dalle Nogare et al., 2009). Indeed, in these species, the first genes that are expressed during ZGA are often short and lack introns (Heyn et al., 2014; Kwasnieski et al., 2019; Rothe et al., 1992). Thus, cell cycle length has an effect on the length and number of transcripts that are produced, but it remains enigmatic whether it has a direct effect on the onset of transcription.

It should be noted that a recent study inDrosophilahas shown that the presence of ‘aborted’ truncated transcripts with intronic sequences from long genes is extensive during ZGA (Kwasnieski et al., 2019 preprint). Moreover, truncation of long transcripts can be precisely regulated and, if translated, their gene products could have developmental functions (Sandler et al., 2018). For example, the Sex lethal (SXL) RNA-binding protein directs formation of a shortened mRNA isoform of short gastrulation (sog), which encodes a dominant-negative SOG protein that may prevent

TGF-βsignaling during this stage of the MZT (Sandler et al., 2018). The generality of regulated and functional truncation of transcripts encoded by long genes remains to be determined.

Transcriptional repressors

Many proteins are maternally loaded to support embryonic development during the early cleavage stages. Among these are proteins that repress transcription (Table 2). In X. laevis and

Drosophila, for example, sequence-specific repressors have been identified that inhibit the expression of subsets of genes (Brown and Wu, 1993; Dunican et al., 2008; Pritchard and Schubiger, 1996; Ruzov, 2004; Ruzov et al., 2009). InDrosophila, the transcription factor Tramtrack (TTK) is maternally loaded and represses the transcription of the segmentation genefushi tarazu(ftz) (Brown and Wu, 1993; Pritchard and Schubiger, 1996). Reducing the amount of

Transcription activity

Developmental time

Chromatin accessibility

Local Global

Histone

Specific repressor Specific transcription factor

General transcription factor H3K27ac H3K4me3 H3R8me

Balance of repressors and

activators

[image:11.612.51.298.219.578.2]

Key

Fig. 4. Mechanisms of transcriptional activation during the second act.

Chromatin structure is relatively open during the early stages of development, which likely supports the large-scale transcriptional reprogramming that takes place during ZGA. This suggests that the absence of transcriptional activity is not due to a repressive chromatin structure. Approaching ZGA, the genome becomes more compacted overall, while local accessibility increases. At the same time, at least in zebrafish andX. laevis, the concentration of non-DNA-bound histones drops, which allows increasing levels of (general and gene-specific) transcription factors to successfully compete for DNA binding. Similarly, the loss of specific repressors allows for transcription of specific genes. While the increase in local accessibility facilitates the binding of transcription factors, accessibility, in turn, often depends on the binding of specific transcription factors. Thus, the loss of repressors, the accumulation of activators and local changes in chromatin accessibility together prime the genome for activation. The chromatin accessibility and the balance of

repressors and activators both influence transcription activity.

DEVEL

O

(12)

TTK or the number of TTK-binding sites results in prematureftz

transcription, while increasing the amount of TTK has the opposite effect. Interestingly, the Smaug RBP is required for destruction of maternal ttk mRNA, thus providing a link between clearance of maternal mRNAs and ZGA (Benoit et al., 2009).

In addition to specific repressors, histones have been identified as more general repressors of transcription in X. laevisand zebrafish (Almouzni and Wolffe, 1995; Amodeo et al., 2015; Joseph et al., 2017). Histones are present in large excess in early embryos (Adamson and Woodland, 1974; Anderson and Lengyel, 1980); they bind with high affinity and little sequence specificity to DNA (Campos and Reinberg, 2009), and, when bound in the form of nucleosomes, they block access of the transcriptional machinery to DNA (Lorch et al., 1987; Workman and Kingston, 1998). In zebrafish and Drosophila embryos, the concentration of soluble histones in the nucleus drops in the approach to genome activation (Joseph et al., 2017; Shindo and Amodeo, 2019). Moreover, experiments in zebrafish have revealed that the drop in the concentration of soluble histones in the nucleus provides an opportunity for the transcriptional machinery to successfully compete for DNA access and activate transcription (Joseph et al., 2017; Pálfy et al., 2017). Developmental changes in the early embryo provide at least two mechanisms by which the concentration of histones may be reduced during embryogenesis. First, the exponential increase in DNA during the rapid cleavage divisions might titrate out histones. A role for DNA content in regulating the onset of transcription is supported by experiments in which increasing DNA content results in premature onset of transcription in

X. laevis, zebrafish andDrosophila(Dekens et al., 2003; Jevtićand Levy, 2017; Lu et al., 2009; Newport and Kirschner, 1982; Prioleau et al., 1994). However, the quantification of both histones and DNA content has shown that, at least in zebrafish, the increase in DNA content is not sufficient to titrate the levels of soluble histones significantly (Joseph et al., 2017). Another possible explanation for the decrease in nuclear histone concentration involves a change in the nuclear import dynamics of histones. This could be a consequence of the dilution of import machinery caused by the increasing number of nuclei (Shindo and Amodeo, 2019) or of the marked increase in the ratio of nuclear over cytoplasmic volume during the cleavage stages (Joseph et al., 2017). The latter has been suggested to limit the capacity of the nucleus to concentrate proteins (Kim and Elbaum, 2013a,b; Kopito and Elbaum, 2007; Kopito and Elbaum, 2009), changing the distribution of histones between nucleus and cytoplasm. A role for import dynamics is supported by the observation that an experimentally induced increase in nuclear volume results in earlier genome activation inX. laevis(Jevtićand Levy, 2015; Jevtićand Levy, 2017). We conclude that both specific and general repressors play a role in changing the balance from transcriptional repression to transcriptional activation during ZGA.

Transcriptional activators

The general absence of transcription during early embryogenesis, as well as the gene-specific onset of transcription that follows, are affected by the availability of the transcriptional machinery. As described above, in C. elegans, the general TAF-4 transcription factor is unavailable for transcription during the early stages of development (Guven-Ozkan et al., 2008), because the protein is sequestered by the binding of OMA-1 and OMA-2. Not until their phosphorylation at the four-cell stage, is TAF-4 released and able to translocate into the nucleus, permitting the onset of transcription (Guven-Ozkan et al., 2008). Similarly, in X. laevis, the concentration of the general TBP transcription factor is limiting

before ZGA (Veenstra et al., 1999). Protein expression profiling has revealed that TBP levels increase and reach sufficiently high levels for genome activation as a result of translation of maternally stored TBP transcripts (Veenstra et al., 1999). Thus, the lack of general transcription factors in the early stages of development contributes to the general absence of transcriptional activity.

Following this period, the gene-specific onset of transcription requires gene-specific transcription factors. The transcription factors that activate zygotically expressed genes have been identified in several species, including Drosophila (Zelda), zebrafish (Pou5f3, Sox19b, Nanog), human (OCT4, DUX4) and mouse (Dppa2, Dppa4, Nfy, Dux) (De Iaco et al., 2019; De Iaco et al., 2017; Eckersley-Maslin et al., 2019; Gao et al., 2018; Harrison et al., 2011; Hendrickson et al., 2017; Lee et al., 2013; Leichsenring et al., 2013; Liang et al., 2008; Lu et al., 2016; Nien et al., 2011). A recent review provides details of their mechanisms of action (Schulz and Harrison, 2018). InDrosophilaand zebrafish, RNA encoding these factors is maternally loaded and their protein levels increase during the early cell cycles due to translation (Harrison et al., 2010; Harrison et al., 2011; Lee et al., 2013; Nien et al., 2011). In zebrafish, it has been shown that the levels of such gene-specific transcription factors affect the time at which transcription starts (Joseph et al., 2017). In this context, it is interesting to note that, in Drosophilaand zebrafish, zygotic transcription begins in two distinct areas in the nucleus (Blythe and Wieschaus, 2016; Chan et al., 2018 preprint; Chen et al., 2013; Hadzhiev et al., 2019; Hilbert et al., 2018 preprint; Hug et al., 2017), which may cause a local increase in the concentration of transcription factors, thereby facilitating transcription.

Chromatin accessibility

Chromatin regulates the accessibility of the genome for DNA-binding proteins. Thus, chromatin accessibility is key to transcriptional regulation. In its simplest form, chromatin consists of DNA wrapped around octamers of the core histone proteins joined by a histone linker protein. DNA accessibility is regulated through DNA methylation, nucleosome positioning and the stability of nucleosomes, which in turn is regulated by histone modifications and the presence of histone variants. Here, we discuss changes in global and local chromatin accessibility that accompany genome activation, and their role in activating transcription.

Overall, chromatin structure is relatively open during the transcriptionally inactive period of the MZT. Human and mouse genomes, for example, undergo global DNA demethylation upon fertilization (Guo et al., 2014; Li et al., 2018; Peat et al., 2014; Santos et al., 2002; Shen et al., 2014), but this has not been observed in other species. More generally, the open chromatin structure of early embryos is characterized by highly dispersed chromatin (Ahmed et al., 2010; Popken et al., 2014), the absence of heterochromatin domains (Ancelin et al., 2016; Laue et al., 2019; Mutlu et al., 2018), the absence of topologically associated domains (‘TADs’) (Du et al., 2017; Hug et al., 2017; Kaaij et al., 2018; Ke et al., 2017; Ogiyama et al., 2018), high levels of histone acetylation (Adenot et al., 1997; Li et al., 2014) and high chromatin mobility (Boškovićet al., 2014). During ZGA, the genome becomes more compact, which may be related to the replacement of embryonic linker histone variants with somatic variants around this time (Pérez-Montero et al., 2013; Saeki et al., 2005).

In contrast, local genome accessibility increases during genome activation inDrosophila, zebrafish, mouse and human (Blythe and Wieschaus, 2016; Gao et al., 2018; Li et al., 2018; Liu et al., 2018b; Lu et al., 2016; Wu et al., 2016; Wu et al., 2018). This is accompanied by the appearance of specific histone modifications. InX. laevisembryos, for example, Histone 3 Arginine 8 (H3R8)

DEVEL

O

(13)

methylation poises genes for expression (Blythe et al., 2010). Similarly, Histone 3 Lysine 4 trimethylation (H3K4me3), a mark that is generally associated with active transcription, appears on promoters during genome activation in Drosophila (Chen et al., 2013; Li et al., 2014),X. tropicalis(Hontelez et al., 2015; Lindeman et al., 2011), zebrafish (Vastenhouw et al., 2010; Zhang et al., 2014b) and mouse (Dahl et al., 2016; Zhang et al., 2016) embryos, often prior to transcription. Although in most cases the relevance of H3K4me3 has not yet been investigated directly, studies in mouse have revealed that it is required for the onset of transcription (Aoshima et al., 2015; Dahl et al., 2016; Liu et al., 2016; Zhang et al., 2016), which supports the observation that Brg1, which is required for H3K4 methylation, is necessary for ZGA (Bultman et al., 2006). Finally, acetylation of Histone 3 Lysine 27 (H3K27Ac) precedes ZGA in zebrafish (Chan et al., 2018 preprint; Sato et al., 2019 preprint; Zhang et al., 2018), and is required for the transcription of–at least–miR-430, which is one of the first genes to be transcribed. Although there is clearly a role for local histone modifications in ZGA, no local DNA methylation changes have been observed that coincide with the onset of transcription (Jiang et al., 2013; Kaaij et al., 2016; Potok et al., 2013), arguing against a role for DNA methylation in regulating ZGA. DNA methylation patterns, however, do play a role in the regulation of transcription during embryogenesis. In zebrafish, for example, hypermethylation at enhancers predicts transcription factor binding and enhancer activity (Kaaij et al., 2016; Liu et al., 2018b), whereas low levels of DNA methylation at promoters predict H3K4me3 and promoter activity (Andersen et al., 2012; Liu et al., 2018b). Finally, ZGA coincides with a significant increase in both the repressive histone modification H3K27me3 (Akkers et al., 2009; Hontelez et al., 2015; Li et al., 2014; Lindeman et al., 2011; Liu et al., 2016; van Heeringen et al., 2014; Vastenhouw et al., 2010) and DNA methylation (Potok et al., 2013), which may help to ensure the onset of gene-specific transcription (Potok et al., 2013; Zenk et al., 2017).

Taken together, chromatin structure is relatively open during the early stages of embryogenesis, which likely supports the large-scale transcriptional reprogramming that takes place during ZGA (Fig. 4). Indeed, the removal of mouse LSD1, which results in premature heterochromatin formation, interferes with developmental progression (Ancelin et al., 2016). Thus, the absence of transcriptional activity is likely not due to a repressive chromatin structure. Approaching ZGA, the genome becomes more compacted overall, while local accessibility increases. At the same time, at least in zebrafish andX. laevis, the concentration of non-DNA-bound histones drops (Amodeo et al., 2015; Joseph et al., 2017), which allows increasing levels of transcription factors to successfully compete for DNA binding (Joseph et al., 2017; Pálfy et al., 2017). Similarly, the loss of specific repressors allows for transcription of specific genes. Although the increase in local accessibility facilitates the binding of transcription factors, accessibility, in turn, often depends on the binding of specific transcription factors (Gao et al., 2018; Liu et al., 2018b; Lu et al., 2016; Schulz et al., 2015; Sun et al., 2015; Veil et al., 2019). Thus, the loss of repressors, the accumulation of activators and local changes in chromatin accessibility together prime the genome for activation.

Concluding remarks

We now have a reasonably thorough descriptive and mechanistic understanding of the two‘acts’of the MZT. Maternally loaded gene products are required to control early development, while the regulated onset of transcription is required for developmental progression: if transcription is inhibited, embryos arrest prior to

gastrulation (Kane et al., 1996; Lee et al., 2013; Zamir et al., 1997). It remains mysterious, however, why such a large fraction of the gene products encoded by the maternal genome is loaded into the egg and then largely eliminated. One clue as to why maternal mRNAs and proteins are cleared may come from the fact that maternal and zygotic transcripts from the same gene often represent different mRNA isoforms (Aanes et al., 2013; Atallah and Lott, 2018; Haberle et al., 2014). Where the isoforms differ in untranslated regions, these would impose differential regulation of transcript stability, translation and localization. Where they differ in the open reading frame, there may also be differences in maternal versus zygotic protein structure and function. Another clue derives from the observation that zygotically re-expressed genes in

Drosophila are enriched for ones whose zygotic expression is patterned (De Renzis et al., 2007). This implies that degradation of ubiquitous maternal transcripts followed by patterned zygotic re-expression is a mechanism that imposes spatial regulation of processes that underlie differential cell behavior and fate in the developing embryo.

Another possible function of degradation of maternal mRNAs and proteins is to supply nutrients to the developing embryo. Evidence for this comes from studies in theDrosophilaembryo, where maternally supplied dNTPs are insufficient for completion of the DNA replication and nuclear divisions that occur prior to large-scale ZGA, thus requiringde novodNTP synthesis from NTPs (Song et al., 2017). Calculations using previous measurements in oocytes of

XenopusandS. purpuratushave revealed similar dNTP shortages for the cleavage divisions (Song et al., 2017). Thus, degradation of maternally loaded mRNAs may provide a source of nucleotides to support DNA replication. Consistent with a requirement for NTPs, rather than a specific set of mRNAs, the maternally loaded transcriptome shows more variability across species than that of most other developmental stages in Drosophila and zebrafish (Atallah and Lott, 2018; Domazet-Lošo and Tautz, 2010; Kalinka et al., 2010). Furthermore, inDrosophila, the maternal transcripts that are degraded during the MZT are much less conserved than those that remain after the onset of ZGA (Atallah and Lott, 2018).

Mechanistically, the regulation of maternal gene products and the activation of zygotic transcription are highly interrelated. For example, in zebrafish, transcription factors that increase in abundance during the early stages of the MZT, because of translation of maternal mRNAs, are required for ZGA (Lee et al., 2013). Similarly, in Drosophila, the Zelda transcription factor, which is essential for early ZGA, increases in abundance early in the MZT through translation of its cognate maternal mRNA (Nien et al., 2011). Additional proteins that increase in abundance show enrichment for functions in chromatin organization (Kronja et al., 2014b) and therefore might play a role in ZGA. Clearance and/or repression of maternal gene products appears also to be a prerequisite for ZGA. In

C. elegans, for example, degradation of OMA-1 and OMA-2 are required to permit nuclear translocation of TAF-4 and, thus, ZGA (Guven-Ozkan et al., 2008). Likewise, mutants that fail to clear and/or repress a significant fraction of the maternal transcriptome either fail to undergo, or suffer a severe delay in, ZGA (e.g.Drosophila smaug, mouseBtg4) (Benoit et al., 2009; Luo et al., 2016; Yu et al., 2016). Finally, genome activation is itself required for the clearance of a large subset of maternal RNAs. For example, in zebrafish,X. laevisand

Drosophila, ZGA produces sets of miRNAs that feed back to destabilize subsets of maternal mRNAs (Bushati et al., 2008; Giraldez et al., 2006; Lund et al., 2009). If ZGA does not occur (e.g. in

Drosophila smaug mutants or in zebrafish if transcription is chemically inhibited) then the miRNAs are not synthesized and

DEVEL

O

Figure

Fig. 1. The maternal-to-zygotic transition (MZT). Red curves depict destabilized maternal transcripts; red gradients represent timing of the multiple phases oftranscript degradation
Table 1. Scale and regulation of maternal transcripts
Fig. 2. Mechanisms of maternal transcriptdegradation in model organisms.mechanisms directing maternal transcriptdegradation are compared acrossD
Fig. 3. Interplay of post-transcriptional and post-translational regulation during the maternal-to-zygotic transition (MZT)
+3

References

Related documents

Contract city pair fare tickets may be issued three (3) business days prior to travel, provided ADTRAV has an approved travel authorization. Employees may request that the

Consider a 3-variate multivariate variance components estimation problem with an among and within classification for the random components and with E[Yi]=lPi for i=l, 2 and 3.

In square systems, individual lines appear on separate buttons at voice terminals. If you want a standard square system, you will give all your voice terminals access to the

With neighbourhood policing came the introduction of purposely built integrated neighbourhood policing teams (INPTs) and the establishment of a new community role;

For throughput AODV and DSR perform better, for packet drop DSR performs better, for end-to-end delay DSR and DSDV perform better and for routing overhead

Particle size and the zeta potential of the reconstituted AmB-SDCS, AmB-SDC, AmB-KDC, AmB-KC and AmB-SC dry powder in distilled water during 7 day at a conc.. 10 mg/mL of

WORLD JOURNAL OF SURGICAL ONCOLOGY Mei et al World Journal of Surgical Oncology 2014, 12 1 http //www wjso com/content/12/1/1 RESEARCH Open Access Femur performed better than tibia