Open access peer-reviewed chapter

Control of Programmed Cell Death During Zebrafish Embryonic Development

Written By

Nikolay Popgeorgiev, Benjamin Bonneau, Julien Prudent and Germain Gillet

Submitted: 29 December 2017 Reviewed: 26 January 2018 Published: 05 November 2018

DOI: 10.5772/intechopen.74494

From the Edited Volume

Recent Advances in Zebrafish Researches

Edited by Yusuf Bozkurt

Chapter metrics overview

1,501 Chapter Downloads

View Full Metrics

Abstract

Programmed cell death (PCD) is a conserved cellular process, which is essential during embryonic development, morphogenesis and tissue homeostasis. PCD participates in the elimination of unwanted or potentially harmful cells, and contributes in this way to the precise shaping of the developing embryo. In this review, the current knowledge related to the role of PCD during zebrafish development is described and an overview is provided about the main actors that induce, control and execute the apoptotic pathways during zebrafish development. Finally, we point out some important issues regarding the regulation of apoptosis during the early stages of zebrafish development.

Keywords

  • Bcl-2 family
  • apoptosis
  • cell death
  • embryonic development
  • mitochondria

1. Introduction

What would today be called genuine apoptotic cells were first observed by German scientist Carl Vogt in 1842. He was studying the morphogenesis of the tadpole notochord of the midwife toad Alytes obstetricans when he observed the formation and subsequent disappearance of vesicular nuclei of the embryonic notochord cells. The idea that cell death could be a fundamental inherited process was first proposed more than century later by Lockshin and Williams. They proposed that rather than a sporadic event, cell death appears in defined spatiotemporal windows during development [1]. In 1972, Kerr et al. used the term “apoptosis” meaning “to fall away from” (apo = from, ptosis = a fall), previously used to describe the falling of leaves in autumn to describe a relatively conserved set of morphological features observed in a wide variety of cell types during physiological episodes of cell death [2]. About 12 different types of programmed cell death (PCD) have been described to date, depending on the morphological features and the molecular pathways that lead to the execution of the PCD. Apoptosis, also called programmed cell death type I, is an inherited metabolically active process during which the cell dies without induction of inflammatory response. Cells undergoing apoptosis exhibit typical morphological features. Apoptotic cells appear to be shrunken and rounded shaped, without any more pseudopodia like cytoplasmic extensions. At the cytoplasmic level, mitochondria undergo fragmentation with a concomitant loss of their transmembrane potential (ΔΨm) [3, 4]. At the nuclear level, apoptosis is characterized by typical chromatin condensation and fragmentation giving rise to pyknotic nuclei, which can be easily observed using specific dyes such as DAPI or Hoescht. Chromatin fragmentation appears to be induced by intracellular endonucleases such as caspase-activated deoxyribonuclease (CAD) and Endonuclease G (EndoG) which preferentially cut DNA strands between nucleosomes resulting in the typical “ladder pattern” observed by electrophoresis [5].

An important feature of apoptosis is the absence of inflammatory response. Indeed, the apoptotic cell maintains its plasma membrane integrity during the whole cell death process, thus preventing the intracellular proteins to interact with surrounding cells. However, phosphatidylserine (PS), an anionic phospholipid usually found at the inner leaflet of the bilayer, is exposed outwardly in apoptotic cells [6]. This morphological feature allows macrophages to detect these cells via specific PS receptors, which is then followed by rapid internalization and phagocytosis.

In vertebrates, there are two main molecular cascades for apoptosis induction [7]. The first one, called the extrinsic pathway, activates cell death by the transduction of external death signals through plasma membrane death receptors. The second one is called the intrinsic (mitochondrial) apoptosis pathway, which essentially leads to the mitochondrial outer membrane permeabilization (MOMP) and the release of apoptotic agents. Although presented at first as individual pathways, these cascades are actually interconnected. Here, we will describe the current knowledge related to the role of the apoptosis during zebrafish development.

Advertisement

2. The apoptotic machinery of the zebrafish embryo

2.1. Caspases

Caspases (for Cysteine ASPartate proteASE) are intracellular cysteine proteases belonging to the family of the interleukine-1β converting enzymes (ICE) family of proteases [8]. Members of the caspase family, share similar 3D conformation and are synthesized as inactive precursors called zymogens (or pro-caspases) containing a prodomain, composed of a p20 large subunit and a p10 small subunit. Caspase activation is achieved by proteolytic cleavage between the large and small subunits and removal of the N-terminus prodomain. This post-translational modification leads to a new conformational state in which caspase homodimers are fully active. The p20 subunit contains the active site of the enzyme harboring a ‘QACXG’ pentapeptide motif [9]. Although caspases are primarily cytosolic, they can also be found at mitochondrial and endoplasmic reticulum (ER) membranes. Caspases have been divided into three groups: interleukin activating caspases and two additional subgroups participating in the initiation and the execution of the apoptosis, respectively.

To date, 17 caspases family members have been identified in zebrafish, including initiator and effector caspases (Table 1). Initiator Caspases (2, 8, 9 and 10) are characterized by the presence of a long N-terminal prodomain. Zebrafish genome contains one ortholog for each of the three Caspases 2, 9 and 10, genes with conserved synteny with the human genome [10]. It also encodes for three Caspase 8 homologous genes. Caspases 8 and 10 possess two death effector domain (DED) domains in the N-terminus whereas Caspases 2 and 9 contain one caspase recruitment domain (CARD). These domains interact with adaptor proteins and are crucial for caspase activation. Importantly, Caspase 8 was identified as an actor of the death receptor pathway (cf. Section 2.3) whereas Caspase 9 belongs to the mitochondrial apoptosis pathway. In zebrafish, Caspases 2 and 8 have been identified as important regulators of vascular development. Indeed, silencing of tumor necrosis factor receptor superfamily member 1A (TNFRSF1A) expression using morpholinos led to the aberrant activation of Caspases 2 and 8 (but not of Caspase 3), resulting in apoptosis of endothelial cells [11]. Furthermore, Caspase 8 was found to be a downstream effector of the Yaf2 apoptosis regulator, a zinc finger-containing protein shown to interact with the DED domain of Caspase 8 and to inhibit apoptosis. Injection of zebrafish embryos with yaf2 targeting morpholino did not affect gastrulation but compromised somitogenesis, which could be rescued by inhibiting Caspase 8 [12].

TypeMammalian proteinZebrafish homologsAccession number
Death receptor ligandsTNF (TNFSF2)Tnfa (Tnf1)NM_212859
Tnfb (Tnf2)NM_00102444
CD95/FasL (TNFSF6)Faslg (Fas ligand)NM_001042701
Apo2L/TRAIL (TNFSF10)Tnfsf10 (Tnfsf10l2)NM_001002593
Tnfsf10l (DL1a)NM_131843
Tnfsf10l3 (DL1b)NM_001042713
Tnfsf10l4 (DL3)NM_001013283
APPAppaNM_131564
TL1Asi:ch211-158d24.4
(tnfsf15)
NM_001123259
Death receptorTNFR1 (TNFRSF1A)Tnfrsf1aNM_213190
CD95/Fas (TNFRSF6)FasXM_021467407
TNFRSF10A and B (DR4 and DR5)Hdr (ZH-DR)NM_194391
TNFRSF10A and B (DR4 and DR5)Tnfrsfa (OTR)NM_131840
TNFRSF21 (DR6)Tnfrsf21 (DR6)NM_001042688
Adaptor proteinFADDFaddXM_001923858
Initiator caspasesCaspase 2NM_001042695
Caspase 4CaspyNM_131505
Caspy 2NM_152884
Caspase b, likeNM_001145592
zgc:171731NM_001109712
Caspase 8Caspase 8aNM_131510
Caspase 8 l1NM_001098619
Caspase 8 l2XM_680338
Effector caspasesCaspase 3Caspase 3aNM_131877
Caspase 3bNM_001048066
Caspase 6Caspase 6aNM_001020497
Caspase 6bNM_001005973
Caspase 6cNM_001039980
Caspase 7Caspase 7aNM_001020607
Inhibitor of caspasesc-FLIP (c-FLAR)Cflara (Cflar)NM_001313772
Birc 2Birc 2, IAP 1NM_194395
Birc 4Birc 4, XIAPNM_194396
Birc 5Birc 5a, SurvivinNM_194397
Birc 5b, Survivin 2NM_145195
Birc 6Birc 6XM_009293036.3
Birc 7Birc 7NM_001098768
Hsp 70Hsp70AB062116.1
BCL-2 family
(BCL-2 like)
Bcl-xLzBlp1NM_131807
Bcl-2zBlp2NM_001030253
Mcl-1Mcl-1aNM_131599
Mcl-1bNM_194394
Nrh, Bcl-2 l10NrzNM_194398
(BAX-like)BaxBaxa, zBax1NM_131562
Baxb, Zbax2NM_001013296
BokBokaNM_001003612
BokbNM_201185
Bcl-wav, Bcl-2 l16NM_001172402
Bcl-2 l13, Bcl-ramboBcl-2 l13NM_001044891
(BH3-only)BadBadaXM_005161364
BadbNM_001270595
Bbc3, PumazPuma, Bbc3NM_001045472
Noxa, Pmaip1zNoxa, Pmaip1NM_001045474
BimzBim, Bcl2-l11NM_001135791
BidzBidNM_001079826
BikzBikNM_001045038
BmfzBmf1NM_001045224
zBmf2NM_001045473
Bnip1Bnip1aXM_684156
Bnip1bXM_001333689
Bnip2Bnip2NM_201218
Bnip3lBnip3la, NixNM_001012242
Bnip3lb, Nip3aNM_205571
Bnip4lBnip4l, Nip3bNM_212693
BniplNM_001128394
MIRAF*Cytochrome CCytochorome C, CycsbNM_001002068
EndoGEndoGNM_001024214.1
AIFAIF, Aifm1NM_200102.2
Smac/DiabloDiabloaNM_200346
DiablobNM_001243034
HtrA2/OmiLOC110437853XM_021472675.1
OthersApaf-1Apaf-1NM_001045243
P53p53, TP53NM_131327

Table 1.

Summary table of apoptosis-associated genes found in the zebrafish genome.

MIRAF, mitochondria released apoptotic factors.


Accession numbers from NCBI database were presented on the left.

Effector caspases are characterized by a short N-terminus end devoid of a recruitment domain. They are activated via proteolytic cleavage by initiator caspases. Among effector caspases, Caspase 3 is critical for the execution of apoptosis, being activated by both, Caspases 8 and 9. In addition, Caspase 3 is capable of feed-back self-activation, thus accelerating apoptosis. Caspase 3, by targeting a wide range of vital cellular components, behaves actually as a genuine apoptosis executor. Activation of Caspase 3 is often considered as a no-return point of apoptosis. Among the substrates cleaved by Caspase 3, are found cytoskeleton proteins, anti-apoptotic factors, metabolic enzymes and several nucleases. For example, proteolytic cleavage of the inhibitory domain of CAD endonuclease (ICAD) leads to CAD activation and subsequent DNA fragmentation, a typical feature of the apoptotic process. The zebrafish genome encodes two Caspase 3 homologs, namely Caspases 3a and 3b [13, 14]. In fact, Caspase 3 activity has been first assessed in stressful conditions following cycloheximide or staurosporine treatments at mid-gastrula stage. In these conditions embryonic development is rapidly blocked, with Caspase 3 being activated within 8 h [15]. Furthermore, Yamashita and colleagues generated a transgenic zebrafish strain expressing full length pro-Caspase 3 to study the role of Caspase 3 in embryonic development. Indeed, these transgenics exhibited a marked increase in the number of apoptotic cells specifically in the retina, the notochord, the heart and the yolk sac, suggesting an essential role of Caspase 3 in numerous morphogenetic processes. Interestingly, silencing of caspase 3 using specific morpholinos did not lead to any significant developmental defects, suggesting some redundancy with other effector caspases, such as Caspases 6 and 7 [16].

In addition, the zebrafish genome encodes for two caspases belonging to the interleukin activating caspases. These caspases known as Caspy (Caspy and Caspy2) contain N-terminal pyrin domains [17]. In the case of Caspy, the pyrin domain was found to be essential for its interaction with the apoptosis-associated speck-like protein containing a CARD (zAsc). In effect, zAsc binding to Caspy led to its activation and apoptosis execution in cellulo. In zebrafish, both genes are specifically expressed in the pharyngeal arches, caspy silencing resulting in developmental defects in this particular region.

2.2. Caspase inhibitors

Since the discovery of viral caspase inhibitors, it became clear that multicellular organisms were also able to make their own caspase inhibitors. These proteins called inhibitors of apoptosis proteins (IAPs) are characterized by one or more baculoviral IAP repeats (BIR), allowing them to prevent caspase activation and apoptosis. The IAPs can also possess a RING domain with an ubiquitin-ligase activity at their C-terminus end. This feature allows the IAPs not only to block caspase activity but also to promote their degradation by the proteasome [18]. It should be noted that six IAP proteins are found in zebrafish [19]. Zebrafish IAPs including survivins (BIRC5a and BIRC5b) appear to play a role in embryonic development. Indeed, the knockdown of birc5a leads to multiple defects including in the nervous system, the cardiovascular system and the hematopoietic system. Interestingly, this phenotype was rescued by ectopic expression of both birc5 paralogs, suggesting the existence of functional redundancy during embryogenesis [20, 21].

Another well characterized caspase inhibitor is the Hsp70 chaperone. Actually, under stressful conditions, the cell can protect itself from the uncontrolled activation of the apoptosis by increasing Hsp70 levels. This protein can bind to and block the recruitment of initiator Caspase 9 into the apoptosome complex [22]. In zebrafish, injection of hsp70 targeting morpholinos resulted in “small eye” phenotype. Close analysis of this phenotype identified a significant increase in apoptotic cells specifically in the developing lens of the zebrafish embryo [23, 24].

Cellular FLICE-like inhibitory protein (c-FLIP) is another example of a cellular caspase inhibitor. This structural analog of Caspase 8 is devoided of proteolytic activity and, is able to bind to DD and prevents the activation of Caspase 8 downstream of the death receptors pathway [25]. c-FLIP is specific to the vertebrate lineage. The knockdown of c-flip in zebrafish results in important cardiovascular abnormalities, including cardiac edema and irregular blood flow consecutive to the formation of blood clots in the vessels [26].

2.3. Molecular actors linked to the death receptor pathway

The cell-extrinsic (or death receptor) pathway of apoptosis is activated by the binding of extracellular ligand proteins belonging to the tumor necrosis factor (TNF) superfamily to specialized receptors, the death receptors called TNF receptors [27]. In mammals, six death receptors have been characterized together with five “death-inducing” ligands. The death receptors are characterized by at least one extracellular cystein-rich domain (CRD) allowing the recognition between the ligand and the receptor and by an intracellular conserved domain called the death domain (DD). Activation of death receptor pathway induces clustering of the receptors through their pre ligand-binding assembly domain (PLAD) [28]. This clustering triggers the recruitment of adaptor proteins such as Fas-associated protein with death domain (FADD) which interact with the DD of the receptor but also with initiator caspases (Caspases 8 and 10) thus forming the death-inducing signaling complex (DISC) [29]. DISC formation then induces activation of initiator caspases leading to the activation of effector caspases and apoptosis execution [30]. DISC-dependent caspases activation can be inhibited by c-FLIP [25].

In zebrafish, on the basis of phylogenetic analysis, orthologs of the five death receptors ligands have been identified (Table 1). CD95/FasL, APP and TL1A each possess one zebrafish ortholog (faslg, appa and 158d24.4, respectively) while TNF possesses two orthologs (TNFa and TNFb) and Apo2L/TRAIL five of them (TNFSF10L, TNFSF10L2, TNFSF10L3 and TNFSF10L4) [31, 32].

The genes encoding death receptors have also been identified in zebrafish. Based on the presence and organization of their CRD and DD, clear orthologs of TNFR1 (TNFRSF1A), CD95/Fas (Fas) and DR6 (TNFRSF21) were characterized. Notably, a selective interaction between TNFa and TNFRSF1 were confirmed by immunoprecipitations [31]. Two other DD-containing receptors, HDR and TNFRSFA, have been described in zebrafish [33, 34]. Extracellular domains of these receptors are close to the one of CD95/Fas but their DD are more similar to DR4 and DR5 DD. However, the observation that HDR and TNFRSFA both bind three orthologs of Apo2L/TRAIL (TNFSF10L, TTNFSF10L2 and TTNFSF10L3) and that they are required for apoptosis induced by these ligands in zebrafish embryos strongly suggests that HDR and TNFRSFA are in fact orthologs of DR4 and DR5 [31]. Finally, so far, no zebrafish ortholog of DR3 have been characterized but the existence of an ortholog of TL1A suggests that a zebrafish DR3 will be identified shortly.

In addition to death receptors and their ligands, the components of DISC are also conserved in zebrafish. A clear ortholog of FADD, containing a DD and a death effector domain (DED), has been identified as well as an ortholog of Caspase 8, casp8a [31, 35]. The latter possesses an N-terminal DED which allow association with FADD and a QACQG active-site motif that is characteristic of Caspases 8 and 10 in mammals. Caspase 8a and FADD are both required for apoptosis induced by Apo2L/TRAIL orthologs in zebrafish embryos. Moreover, Caspase 8a is functionally conserved as it restores death receptor-induced apoptosis in mouse cells lacking endogenous Caspase 8 [35]. Two other genes related to casp8a exist (casp8l1 and casp8l2) but their possible involvement in the cell-extrinsic pathway is unclear. Indeed, Caspase 8l1 has a QACQG active-site motif but no DED whereas Caspase 8l2 possesses an N-terminal DED but its active site is similar to the one of Caspase 2. Finally, a zebrafish ortholog of c-FLIP, referred to as Cflar, able to inhibit apoptosis induced by Apo2L/TRAIL orthologs have also been described [31].

In mammals, the cell-extrinsic apoptosis pathway is essential for the functioning of the immune system. However, this has not been clearly established in zebrafish as yet. One study revealed that apoptosis is important for T and B cells homeostasis as overexpression of Bcl-2 in these cells increased their number, but this work did not show a role for the extrinsic pathway in lymphocyte homeostasis [36]. In contrast, the death receptor HDR appears to be involved in red blood cells homeostasis. Indeed, HDR is specifically expressed in hematopoietic lineage and its inhibition, using either a dominant negative mutant or antisense morpholinos, leads to abnormal accumulation of erythroid cells [34, 37].

During early development, hdr, tnfrsfa and fadd are expressed in the notochord suggesting that they might play a role in this structure [31]. Fas and FasLg are also present in developing notochord and their knockdown via morpholinos leads to prolonged expression of notochord specific genes and to an abnormally enlarged notochord at 4 days post-fertilization (dpf). As apoptosis plays a role in notochord regression, this suggests that extrinsic pathway may be involved in this process [38].

During development, TNFRSFA and its ligands TNFSF10L2 and TNFSF10L3 are expressed in particular in neuromasts that contain hair cells which turnover is regulated by apoptosis [31, 39]. Finally, the death receptor pathway seems to be involved in zebrafish eye development as FADD is required for slowing cell growth during this process [40].

2.4. Molecular actors linked to the intrinsic pathway

The intrinsic pathway of apoptosis also called the mitochondrial pathway can be induced by various signals including DNA damage, chemotherapy, viral infection or growth factors deprivation. The mitochondrial pathway of apoptosis is mainly controlled by the Bcl-2 family of proteins, which is described in more detail in the next chapter. The mitochondrion is a membrane-enclosed organelle found in most eukaryotic cells [41]. The diameter of these organelles falls under the 0.5–10 μm range. Mitochondria are often referred as the “powerhouse” of the cell because they generate most of the cell’s supply of adenosine triphosphate (ATP). In addition to their bioenergetic function, mitochondria are involved in a number of other processes, such as signal transduction, cell differentiation, cell cycle and cell growth [42]. Mitochondria are also major “decision centers” for the execution or prevention of apoptosis. Indeed, a number of pro-apoptotic molecules appear to be stored in the existing space between the inner mitochondrial membrane (IMM) and the OMM. At the mitochondrial level, the induction of apoptosis leads to the OMM permeabilization, which leads to the irreversible release into the cytosol of pro-apoptotic factors that promote caspase activation, DNA fragmentation and ultimately the death of the cell. Belong to this toxic molecular “cocktail” among others: Cytochrome C, AIF, EndoG, Smac/Diablo and HtrA2/Omi.

2.4.1. Mitochondria released apoptotic factors

Cytochrome C is a small heme protein (approximately 12 kDa), exhibiting a positive net charge, located in the intermembrane space (IMS) where it can be loosely attached to the IMM. Cytochrome C is synthesized inside the cytosol, and subsequently transported into the IMS. Bioenergetically, Cytochrome C is a component of the mitochondrial electron transport chain. The heme molecule of cytochrome C accepts electrons originating from complex III and transfers them to the cytochrome oxydase complex, thus cytochrome C is indispensable for the oxidative phosphorylation and the maintenance of cellular energy fluxes. Cytochrome C plays an additional role in the context of apoptosis, as it is now well established by a large number of in vitro and in vivo studies [43, 44, 45, 46]. The release of Cytochrome C is a rapid and complete process, irrespective of the intensity of the death-inducing signal or the temperature, indicating that this is a non-enzymatic phenomenon obeying an “all or nothing” law. When released in the cytosol, Cytochrome C interacts with the adaptor protein apoptosis protease activating factor (Apaf-1) in presence of ATP. Apaf-1 was first characterized by Wang and collaborators [47]. It is a 130 kDa multidomain protein, containing a CARD domain at its N-terminus end, sharing homology with the CARD domain of Caspase 9, an ATPase domain as well as two WD-40 repeats at the C-terminus end. The Wang laboratory discovered the existence of a multiprotein complex called the apoptosome, comprising Cytochrome C, Apaf-1 and Caspase 9, which was found to be able to activate Caspase 3 [48]. The apoptosome was crystallized by Acehan and collaborators in 2002 [49]. Although the existence of the zebrafish apoptosome has not been directly demonstrated, its genome encodes for all functional components of this complex. Furthermore, treatment with drugs converging toward the intrinsic pathway was found to lead to Caspase 3 activation [50, 51].

Apoptosis-inducing factor (AIF) is a 57 kDa flavoprotein with NADH oxidase activity which is located in the mitochondrial IMS [52, 53]. As Cytochrome C, AIF is encoded by a nuclear gene, which is imported into the mitochondria after being synthesized in the cytosol. In response to diverse death stimuli, AIF is released into the cytosol and transferred to the nucleus where it binds to chromatin [54]. The binding of AIF to DNA induces chromatin peripheral condensation and subsequent fragmentation. Wang and colleagues showed that this process is due to AIF-dependent activation of endogenous nucleases such as endonuclease G and CAD. AIF is unable to induce the fragmentation of the DNA on its own. To this end, AIF needs the help of endonuclease G. EndoG is a mitochondrial nuclease of 30 kDa which is required for the replication of the mitochondrial chromosome. During apoptosis, EndoG is released from the mitochondria into the cytosol and subsequently enters into the nucleus. In this compartment, EndoG cleaves the DNA into nucleosomal fragments [55]. AIF and EndoG orthologs are both expressed during zebrafish embryonic development but their functional implications remain to be analyzed. Second mitochondria-derived activator of caspase/direct IAP-binding protein with low pI (Smac/Diablo) is a mitochondrial protein, which resides in the intermembrane space [56, 57]. The human gene is called smac whereas the mouse ortholog is called diablo. Smac/Diablo is synthesized as a precursor protein containing a 55 residue N-terminal mitochondrial targetting sequence. During the translocation of Smac/Diablo to the mitochondria, this sequence is cleaved which uncovers an IAP-binding motif (IBM) required for apoptotic activity. Indeed, in the presence of an apoptotic stress, the mitochondria release Smac/Diablo in the cytosol where it participates indirectly in the activation of caspases by binding IAP proteins. The binding of Smac/Diablo to IAP disrupts IAP-Caspase interactions. Thus, released caspases can then be activated and execute the cell death program. The Smac/Diablo-IAP complex formation appears to be a regulated process since IAP can ubiquitinate Smac/Diablo and drive it for proteasomal degradation [18]. In zebrafish, Smac/Diablo gene is mainly expressed in the late developmental stages with the most prominent expression in the heart, the lens and the liver. However, the possible implication of this IMS factor in the morphogenesis of these organs remains unknown.

High temperature requirement A2 (HtrA2/Omi) is a heat shock protein first identified in Escherichia coli [58]. Its ortholog in mammals, called Omi was initially described as an ER protein [59, 60]. However, it seems that this protein is mainly mitochondrial in non-apoptotic cells [61]. When the OMM is permeabilized, Omi is released into the cytosol where it binds to IAP and irreversibly inhibits its activity by proteolytic cleavage [57]. Thus, it seems that Omi is a more potent inhibitor of IAP than Smac/Diablo which blocks IAP reversibly. Currently, there is no data about the possible implication of HtrA2/Omi functional implication in zebrafish have not been assessed.

2.4.2. Bcl-2 family of proteins

The Bcl-2 family of proteins is a group of intracellular eukaryotic proteins best known for their implication in MOMP. The founding member of this family, the bcl2 (B-cell lymphoma/leukemia 2) gene was discovered in a study on a chromosome translocation frequently observed in human B-cell follicular lymphomas. Tsujimoto and colleagues showed that the translocation between chromosome 18 (q21) chromosome 14 (q32), t (14;18), results in the relocation of the bcl2 ORF downstream of the enhancer promoter region of the igh heavy chain immunoglobulin gene [62, 63, 64]. This translocation results in transcriptional upregulation of bcl2 gene expression [65]. The Bcl-2 family comprises proteins with antagonistic functions with respect to apoptosis regulation. Structurally, all Bcl-2 proteins contain in their primary structure one or more conserved Bcl-2 homology (BH) domains. Based on this criterion, three subgroups have been identified: (1) the anti-apoptotic multidomain members containing all four BH domains (BH1–4); (2) the pro-apoptotic members containing three BH domain (lacking BH4) and (3) the pro-apoptotic BH3-only members containing only the sole BH3 domain. In addition, Bcl-2 proteins may contain a transmembrane anchoring domain (TM domain) at the C-terminus end. Bcl-2 multidomain pro-apoptotic members (Bax and Bak) are the effectors for MOMP. Through their oligomerization, they form pores at the MOM and promote the release of apoptotic factors including Cytochome C, AIF and EndoG [66]. Anti-apoptotic Bcl-2 members block their activity and promote cell survival, whereas BH3-only proteins play the role as intracellular judges as they can inhibit anti-apoptotic members and/or promote Bax/Bak oligomerization.

Bcl-2 homologs of all three subgroup of Bcl-2 family have been identified and molecularly characterized in zebrafish [67, 68, 69]. Due to genomic duplication, the zebrafish genome possesses several Bcl-2-related paralogs including Mcl-1a, Mcl-1b, Bax1, Bax2, Boka, Bokb and Bmf1 and Bmf2. Interestingly, the ortholog of the bak gene has been lost in the teleost lineage. Instead the zebrafish genome harbors bcl-wav (an acronym for Bcl-2 homolog found in water living anamniotes), a bcl-2 homolog only found in fishes and anurans [70]. With the exception of the BH3-only protein Bik, all Bcl-2 related genes in zebrafish are maternally inherited since their corresponding mRNAs were detected before the mid-blastula transition [67].

Of note during zebrafish early development physiological apoptosis is not observed suggesting that at least some of the Bcl-2 family members may have additional non-apoptotic roles during early embryogenesis. In this respect, Zhong and colleagues recently demonstrated that the zebrafish ortholog of the BH3-only protein Noxa (zNoxa) not only controls apoptosis during late stages of gastrulation but also plays role in cell cycle in the developing blastula [71]. Indeed znoxa knockdown led to a significant decrease in the number of mitotic cells. This phenotype seems to be dependent on the Wnt signaling pathway since znoxa knockdown led to increase of zwnt4b expression. In addition silencing of zwnt4b rescued zNoxa phenotype and restored a WT count of cells in G2/M phase.

Using the zebrafish model, we demonstrated that a Bcl-2 homolog, referred to as Nrz (Nr-13 ortholog in zebrafish), is critical during the early stages of zebrafish development [72]. In zebrafish, Nrz protein possesses a dual subcellular localization at the ER and the mitochondria [73]. Its functional invalidation causes embryo development arrest followed by detachment of the entire blastomeres from the yolk sac. By performing a series of time-lapse and confocal microscopy experiments, we demonstrated that this phenotype is due to the premature formation of the actin-myosin contractile ring, a supramolecular structure, which squeezes and halves the embryo at the level of the margin. Furthermore, by using single subcellular localization (SSL) Nrz mutants, we showed that the ER-resident Nrz but not mitochondrial-resident Nrz was critical for its physiological function. Indeed, at the ER membrane, Nrz interacts with the Ca2+ channel inositol trisphosphate receptor type 1 (IP3R1) via its BH4 domain [73, 74]. In this way, Nrz slows down the release of Ca2+ into the yolk sac, which consecutively allows controlling the formation of the contractile actin-myosin ring via the Calmodulin-MLCK pathway. Overall, our results highlighted for the first time that a Bcl-2 family member is able to orchestrate cellular migration events by controlling intracellular Ca2+ fluxes.

In addition, we identified the new Bcl-2 family member Bcl-wav [75]. The bclwav gene is expressed throughout zebrafish early development; it encodes a pro-apoptotic Bcl-2 family member with strict mitochondrial localization. However, bclwav silencing in zebrafish causes a specific apoptosis independent phenotype at 24 hpf. Macroscopically, this phenotype is characterized by an embryo anterioposterior axis reduction as well as notochord deviation. Using time-lapse microscopy, we demonstrated that this phenotype affects the convergence and extension movements which underlie the establishment of the embryonic axes. Indeed, in bclwav-silenced embryos, mesodermal cells migrated erratically compared to cells from control embryos, which moved in a coordinated fashion. In vivo analysis of the actin cytoskeleton revealed that these migration defaults correlated with randomization of F-actin protrusion dynamics. Interestingly, bclwav silencing was correlated with a decrease of mitochondrial Ca2+ levels and concomitant increase of cytosolic Ca2+. Together these results indicated that Bcl-wav controls the dynamics of the actin cytoskeleton by regulating intracellular Ca2+ homeostasis at the mitochondrial level. Indeed, at the mitochondria, Bcl-wav interacts with the voltage-dependent anion channel (VDAC) channel and enhances mitochondrial Ca2+ uptake, which in turn controls actin polymerization and cell migration. It is important to note that, mitochondrial calcium uniporter (mcu) knockdown phenocopies bclwav-silenced embryos. Indeed, MCU downregulation leads to decreased mitochondrial calcium uptake and impaired actin dynamics giving the first insights into the critical role of the mitochondrial Ca2+ oscillations in the vertebrate development [70].

Put into a broader context, the results demonstrate that members of the Bcl-2 family are able to control cell migration in a calcium-dependent manner via their direct interaction with intracellular Ca2+ channels independently of their involvement in the regulation of cell death [76].

2.4.3. P53

P53 is a transcription factor considered as the main tumor suppressor regulating cell fate decisions. Indeed, p53 is the most frequently mutated and/or inactivated gene in human cancer modulating cell responses to DNA damage, oncogenic signaling and hypoxia in order to preserve genome integrity. The zebrafish p53 ortholog is highly conserved with 48% of sequence identity compared to the human sequence [77]. P53 is highly and ubiquitously expressed during early zebrafish embryo development, then predominantly expressed in the brain during the pharyngula stage before to decrease in expression after 48 hpf [77, 78, 79]. P53 tumor suppressor activity has been confirmed in the zebrafish model by the characterization of a p53 mutant harboring a missense mutation in the DNA-binding domain, M214K, leading to the development of multiple organ tumors at around 8.5 months [80]. The mechanism of action of p53 has been widely studied in mammals and similar results have been obtained in the zebrafish model [81]. During different cell stresses, p53 expression is stabilized leading to its activation and accumulation in the nucleus, and subsequently to cell cycle arrest and the intrinsic pathway of apoptosis [82].

Its role in DNA damage and apoptosis has been well studied in the zebrafish model [82]. It has been shown that increased DNA damage leads to the stimulation of p53 transcription and an increase of p53 protein level [79]. In addition, knockdown of p53 decreases apoptosis induced by different stimuli including gamma and UV irradiation, camptothecin treatment or altered DNA replication [83, 84, 85].

Finally, an undesirable effect of p53 activation-induced cell death has been characterized and associated with off-target effects triggered by morpholino antisense oligonucleotides technologies. These off-target effects have been assimilated to p53 signaling pathway-dependent apoptosis [86, 87]. Indeed, the most commonly off-target reported phenotype is characterized by an increase of neural cell death at 22 hpf [86]. Even if the precise molecular mechanism is still unclear, it has been attributed to the activation of p53 leading to the apoptotic cell death. Indeed, the role of p53 in this phenotype has been shown by the characterization of smo and wnt5 morphant embryos, where simultaneous knockdown of p53 in these embryos rescued neuronal apoptosis [87]. These results highlight the extremely cautious, which has to be employed when potential actors of the apoptotic program are studied by the morpholino strategy. In addition to the use of different morpholino sequences targeted, the same gene and crucial rescue experiments, co-expression of p53 morpholino is now commonly used to discern between gene-specific and off-target effects [19]. Indeed, p53 is not required for proper early embryos development and p53 morpholino does not interfere with other gene-specific phenotypes [87, 88].

Advertisement

3. Developmental control of apoptosis in the early zebrafish embryo

Following fertilization and during early stages of embryonic development, embryo relies entirely on the maternal inherited mRNAs and proteins which were accumulated during oogenesis. After several synchronous divisions, which lack G1 and G2 phases, the cell cycle slows down and divisions become asynchronous. This step, referred to as the mid-blastula transition (MBT), corresponds to the beginning of the expression of zygotic genes. Ikegami and colleagues first noticed that zebrafish embryos treated before the MBT with microtubule destabilizing agent nocodazole or DNA-damaging molecules such as camptothecine, hydroxyurea or aphidicoline did not result in direct apoptosis activation. Instead, the cell cycle was arrested with the apoptotic program being executed several hours later, during the mid-gastrula stage [50, 51]. This phenomenon is not restricted to zebrafish as a similar apoptotic control also operates in the Xenopus embryo [89]. These observations suggested that key molecular components of the apoptotic program were either missing or inactivated during early development. In this respect, one report using a proteomics approach identified that Apaf-1 was missing before the MBT suggesting that a functional apoptosome may be set up after the MBT transition [90]. However, these later data do not explain why inhibition of protein synthesis using cycloheximide is not able to slow the apoptotic program in the zebrafish gastrula [15]. Furthermore, our laboratory demonstrated that ectopic expression of zBax through recombinant mRNA injection in one cell stage embryos actually led to ΔΨm loss and Caspase 3 activation as early as the blastula stage [91]. Importantly, this apoptotic phenomenon specifically occurs in an extraembryonic structure referred to as the yolk syncytial layer (YSL). The YSL results from the fusion of blastomeres physically connected with the yolk cell by cytoplasmic bridges [92]. Fusion of margin blastomeres with the yolk leads to the release of cell nuclei and other cellular components including a dense network of active mitochondria interconnected with ER membranes. Purified YSL mitochondria can undergo MOMP and Cytochrome C release. This was demonstrated by performing in vitro Cytochrome C release assay using recombinant human truncated Bid protein. Indeed, Bid is a BH3-only protein which once cleaved by Caspase 8 translocates to the MOM and activates Bax oligomerization. Thus it is tempting to speculate that at least at the level of the YSL mitochondria harbor sufficient amount of Bax in order to initiate MOMP following BH3-only stimulus.

Altogether these results showed that zebrafish early embryo possesses a functional apoptotic machinery. Thus the tight apoptotic control observed by Ikegami et al. may be exerted at the post-translational level through protein-protein interactions. In this respect, Kratz and colleagues demonstrated that manipulation of the ratio between pro- or anti-apoptotic Bcl-2 proteins determines the capacity of early zebrafish embryo to undergo apoptosis. Notably, overexpression of BH3-only or zBax paralogs induced rapid Caspase 3-dependent cell death whereas co-expression of Bcl-2-related anti-apoptotic members effectively counteracted early embryo mortality [67].

Advertisement

4. Conclusion

Apoptosis represents a key cellular process that maintains tissue homeostasis and shapes the embryo. Impairment or a contrario overactivation of cell death often leads to severe developmental abnormalities and lethal phenotypes. Thus, the tight spatiotemporal control over apoptosis induction is critical for orchestrating embryonic development. The fact that zebrafish genome encodes for the majority of apoptosis actors found in the human genome makes zebrafish a valuable model for understanding the contribution of apoptosis regulators during embryonic development in vertebrates. The use of antisense chemically modified nucleotides, most notably morpholinos, allowed to assess the implication of many apoptosis regulators in the developmental process. However, the possible off targeting and unspecific activation of the p53 pathway can be a drawback in some instances. In this respect, the development of new genome editing approaches such as CRISPR/Cas9 will allow in the near future to assess the precise role of each and every member of the cell death machinery during embryogenesis.

Advertisement

Acknowledgments

This work is supported by AFM telethon, Ligue Nationale Contre le Cancer, Cancéropole Auvergne Rhônes- Alpes (CLARA—Oncostarter), Fondation ARC and Medical Research Council, UK (MC_UU_00015/7).

Advertisement

Conflict of interest

The authors declare that they have no conflict of interest.

References

  1. 1. Lockshin RA, Williams CM. Programmed cell death--I. Cytology of degeneration in the intersegmental muscles of the Pernyi Silkmoth. Journal of Insect Physiology. 1965;11:123-133
  2. 2. Kerr JF, Wyllie AH, Currie AR. Apoptosis: A basic biological phenomenon with wide-ranging implications in tissue kinetics. British Journal of Cancer. 1972;26(4):239-257
  3. 3. Adams JW et al. Cardiomyocyte apoptosis induced by Galphaq signaling is mediated by permeability transition pore formation and activation of the mitochondrial death pathway. Circulation Research. 2000;87(12):1180-1187
  4. 4. Zamzami N et al. Inhibitors of permeability transition interfere with the disruption of the mitochondrial transmembrane potential during apoptosis. FEBS Letters. 1996;384(1):53-57
  5. 5. Wyllie AH. Glucocorticoid-induced thymocyte apoptosis is associated with endogenous endonuclease activation. Nature. 1980;284(5756):555-556
  6. 6. Fadok VA et al. Regulation of macrophage cytokine production by phagocytosis of apoptotic and post-apoptotic cells. Biochemical Society Transactions. 1998;26(4):653-656
  7. 7. Elmore S. Apoptosis: A review of programmed cell death. Toxicologic Pathology. 2007;35(4):495-516
  8. 8. Alnemri ES et al. Human ICE/CED-3 protease nomenclature. Cell. 1996;87(2):171
  9. 9. Fuentes-Prior P, Salvesen GS. The protein structures that shape caspase activity, specificity, activation and inhibition. The Biochemical Journal. 2004;384(Pt 2):201-232
  10. 10. Sidi S et al. Chk1 suppresses a caspase-2 apoptotic response to DNA damage that bypasses p53, Bcl-2, and caspase-3. Cell. 2008;133(5):864-877
  11. 11. Espin R et al. TNF receptors regulate vascular homeostasis in zebrafish through a caspase-8, caspase-2 and P53 apoptotic program that bypasses caspase-3. Disease Models & Mechanisms. 2013;6(2):383-396
  12. 12. Stanton SE et al. Yaf2 inhibits caspase 8-mediated apoptosis and regulates cell survival during zebrafish embryogenesis. The Journal of Biological Chemistry. 2006;281(39):28782-28793
  13. 13. Yabu T et al. Characterization of zebrafish caspase-3 and induction of apoptosis through ceramide generation in fish fathead minnow tailbud cells and zebrafish embryo. The Biochemical Journal. 2001;360(Pt 1):39-47
  14. 14. Tucker MB et al. Phage display and structural studies reveal plasticity in substrate specificity of caspase-3a from zebrafish. Protein Science. 2016;25(11):2076-2088
  15. 15. Negron JF, Lockshin RA. Activation of apoptosis and caspase-3 in zebrafish early gastrulae. Developmental Dynamics. 2004;231(1):161-170
  16. 16. Yamashita M et al. Extensive apoptosis and abnormal morphogenesis in pro-caspase-3 transgenic zebrafish during development. The Journal of Experimental Biology. 2008;211(Pt 12):1874-1881
  17. 17. Masumoto J et al. Caspy, a zebrafish caspase, activated by ASC oligomerization is required for pharyngeal arch development. The Journal of Biological Chemistry. 2003;278(6):4268-4276
  18. 18. MacFarlane M et al. Proteasome-mediated degradation of Smac during apoptosis: XIAP promotes Smac ubiquitination in vitro. The Journal of Biological Chemistry. 2002;277(39):36611-36616
  19. 19. Eimon PM, Ashkenazi A. The zebrafish as a model organism for the study of apoptosis. Apoptosis. 2010;15(3):331-349
  20. 20. Delvaeye M et al. Role of the 2 zebrafish survivin genes in vasculo-angiogenesis, neurogenesis, cardiogenesis and hematopoiesis. BMC Developmental Biology. 2009;9:25
  21. 21. Ma A et al. The role of survivin in angiogenesis during zebrafish embryonic development. BMC Developmental Biology. 2007;7:50
  22. 22. Beere HM et al. Heat-shock protein 70 inhibits apoptosis by preventing recruitment of procaspase-9 to the Apaf-1 apoptosome. Nature Cell Biology. 2000;2(8):469-475
  23. 23. Evans TG et al. Heat shock factor 1 is required for constitutive Hsp70 expression and normal lens development in embryonic zebrafish. Comparative Biochemistry and Physiology. Part A, Molecular & Integrative Physiology. 2007;146(1):131-140
  24. 24. Evans TG et al. Zebrafish Hsp70 is required for embryonic lens formation. Cell Stress & Chaperones. 2005;10(1):66-78
  25. 25. Krueger A et al. Cellular FLICE-inhibitory protein splice variants inhibit different steps of caspase-8 activation at the CD95 death-inducing signaling complex. The Journal of Biological Chemistry. 2001;276(23):20633-20640
  26. 26. Sakamaki K et al. Conservation of structure and function in vertebrate c-FLIP proteins despite rapid evolutionary change. Biochemistry and Biophysics Reports. 2015;3:175-189
  27. 27. Aggarwal BB. Signalling pathways of the TNF superfamily: A double-edged sword. Nature Reviews. Immunology. 2003;3(9):745-756
  28. 28. Chan FK et al. A domain in TNF receptors that mediates ligand-independent receptor assembly and signaling. Science. 2000;288(5475):2351-2354
  29. 29. Berglund H et al. The three-dimensional solution structure and dynamic properties of the human FADD death domain. Journal of Molecular Biology. 2000;302(1):171-188
  30. 30. Micheau O, Tschopp J. Induction of TNF receptor I-mediated apoptosis via two sequential signaling complexes. Cell. 2003;114(2):181-190
  31. 31. Eimon PM et al. Delineation of the cell-extrinsic apoptosis pathway in the zebrafish. Cell Death and Differentiation. 2006;13(10):1619-1630
  32. 32. Glenney GW, Wiens GD. Early diversification of the TNF superfamily in teleosts: Genomic characterization and expression analysis. Journal of Immunology. 2007;178(12):7955-7973
  33. 33. Bobe J, Goetz FW. Molecular cloning and expression of a TNF receptor and two TNF ligands in the fish ovary. Comparative Biochemistry and Physiology. Part B, Biochemistry & Molecular Biology. 2001;129(2-3):475-481
  34. 34. Long Q et al. Stimulation of erythropoiesis by inhibiting a new hematopoietic death receptor in transgenic zebrafish. Nature Cell Biology. 2000;2(8):549-552
  35. 35. Sakata S et al. Conserved function of caspase-8 in apoptosis during bony fish evolution. Gene. 2007;396(1):134-148
  36. 36. Langenau DM et al. Suppression of apoptosis by bcl-2 overexpression in lymphoid cells of transgenic zebrafish. Blood. 2005;105(8):3278-3285
  37. 37. Kwan TT et al. Regulation of primitive hematopoiesis in zebrafish embryos by the death receptor gene. Experimental Hematology. 2006;34(1):27-34
  38. 38. Ferrari L et al. FAS/FASL are dysregulated in chordoma and their loss-of-function impairs zebrafish notochord formation. Oncotarget. 2014;5(14):5712-5724
  39. 39. Williams JA, Holder N. Cell turnover in neuromasts of zebrafish larvae. Hearing Research. 2000;143(1-2):171-181
  40. 40. Viringipurampeer IA et al. Pax2 regulates a fadd-dependent molecular switch that drives tissue fusion during eye development. Human Molecular Genetics. 2012;21(10):2357-2369
  41. 41. Henze K, Martin W. Evolutionary biology: Essence of mitochondria. Nature. 2003;426(6963):127-128
  42. 42. McBride HM, Neuspiel M, Wasiak S. Mitochondria: More than just a powerhouse. Current Biology. 2006;16(14):R551-R560
  43. 43. Brustugun OT et al. Apoptosis induced by microinjection of cytochrome c is caspase-dependent and is inhibited by Bcl-2. Cell Death and Differentiation. 1998;5(8):660-668
  44. 44. Kroemer G. Cytochrome c. Current Biology. 1999;9(13):R468
  45. 45. Liu X et al. Induction of apoptotic program in cell-free extracts: Requirement for dATP and cytochrome c. Cell. 1996;86(1):147-157
  46. 46. Pan Z, Voehringer DW, Meyn RE. Analysis of redox regulation of cytochrome c-induced apoptosis in a cell-free system. Cell Death and Differentiation. 1999;6(7):683-688
  47. 47. Zou H et al. Apaf-1, a human protein homologous to C. Elegans CED-4, participates in cytochrome c-dependent activation of caspase-3. Cell. 1997;90(3):405-413
  48. 48. Zou H et al. An APAF-1.Cytochrome c multimeric complex is a functional apoptosome that activates procaspase-9. The Journal of Biological Chemistry. 1999;274(17):11549-11556
  49. 49. Acehan D et al. Three-dimensional structure of the apoptosome: Implications for assembly, procaspase-9 binding, and activation. Molecular Cell. 2002;9(2):423-432
  50. 50. Ikegami R, Hunter P, Yager TD. Developmental activation of the capability to undergo checkpoint-induced apoptosis in the early zebrafish embryo. Developmental Biology. 1999;209(2):409-433
  51. 51. Ikegami R et al. Effect of inhibitors of DNA replication on early zebrafish embryos: Evidence for coordinate activation of multiple intrinsic cell-cycle checkpoints at the mid-blastula transition. Zygote. 1997;5(2):153-175
  52. 52. Miramar MD et al. NADH oxidase activity of mitochondrial apoptosis-inducing factor. The Journal of Biological Chemistry. 2001;276(19):16391-16398
  53. 53. Susin SA et al. Molecular characterization of mitochondrial apoptosis-inducing factor. Nature. 1999;397(6718):441-446
  54. 54. Ye H et al. DNA binding is required for the apoptogenic action of apoptosis inducing factor. Nature Structural Biology. 2002;9(9):680-684
  55. 55. Li LY, Luo X, Wang X. Endonuclease G is an apoptotic DNase when released from mitochondria. Nature. 2001;412(6842):95-99
  56. 56. Du C et al. Smac, a mitochondrial protein that promotes cytochrome c-dependent caspase activation by eliminating IAP inhibition. Cell. 2000;102(1):33-42
  57. 57. Verhagen AM et al. Identification of DIABLO, a mammalian protein that promotes apoptosis by binding to and antagonizing IAP proteins. Cell. 2000;102(1):43-53
  58. 58. Spiess C, Beil A, Ehrmann M. A temperature-dependent switch from chaperone to protease in a widely conserved heat shock protein. Cell. 1999;97(3):339-347
  59. 59. Faccio L et al. Characterization of a novel human serine protease that has extensive homology to bacterial heat shock endoprotease HtrA and is regulated by kidney ischemia. The Journal of Biological Chemistry. 2000;275(4):2581-2588
  60. 60. Gray CW et al. Characterization of human HtrA2, a novel serine protease involved in the mammalian cellular stress response. European Journal of Biochemistry. 2000;267(18):5699-5710
  61. 61. Suzuki Y et al. A serine protease, HtrA2, is released from the mitochondria and interacts with XIAP, inducing cell death. Molecular Cell. 2001;8(3):613-621
  62. 62. Tsujimoto Y et al. Involvement of the bcl-2 gene in human follicular lymphoma. Science. 1985;228(4706):1440-1443
  63. 63. Tsujimoto Y, Croce CM. Molecular cloning of a human immunoglobulin lambda chain variable sequence. Nucleic Acids Research. 1984;12(22):8407-8414
  64. 64. Tsujimoto Y et al. The t(14;18) chromosome translocations involved in B-cell neoplasms result from mistakes in VDJ joining. Science. 1985;229(4720):1390-1393
  65. 65. Cleary ML, Smith SD, Sklar J. Cloning and structural analysis of cDNAs for bcl-2 and a hybrid bcl-2/immunoglobulin transcript resulting from the t(14;18) translocation. Cell. 1986;47(1):19-28
  66. 66. Youle RJ, Strasser A. The BCL-2 protein family: Opposing activities that mediate cell death. Nature Reviews. Molecular Cell Biology. 2008;9(1):47-59
  67. 67. Kratz E et al. Functional characterization of the Bcl-2 gene family in the zebrafish. Cell Death and Differentiation. 2006;13(10):1631-1640
  68. 68. Inohara N, Nunez G. Genes with homology to mammalian apoptosis regulators identified in zebrafish. Cell Death and Differentiation. 2000;7(5):509-510
  69. 69. Chen MC et al. Cloning and characterization of a novel nuclear Bcl-2 family protein, zfMcl-1a, in zebrafish embryo. Biochemical and Biophysical Research Communications. 2000;279(2):725-731
  70. 70. Prudent J et al. Bcl-wav and the mitochondrial calcium uniporter drive gastrula morphogenesis in zebrafish. Nature Communications. 2013;4:2330
  71. 71. Zhong JX et al. Zebrafish Noxa promotes mitosis in early embryonic development and regulates apoptosis in subsequent embryogenesis. Cell Death and Differentiation. 2014;21(6):1013-1024
  72. 72. Arnaud E et al. The zebrafish bcl-2 homologue Nrz controls development during somitogenesis and gastrulation via apoptosis-dependent and -independent mechanisms. Cell Death and Differentiation. 2006;13(7):1128-1137
  73. 73. Popgeorgiev N et al. The apoptotic regulator Nrz controls cytoskeletal dynamics via the regulation of Ca2+ trafficking in the zebrafish blastula. Developmental Cell. 2011;20(5):663-676
  74. 74. Bonneau B et al. The Bcl-2 homolog Nrz inhibits binding of IP3 to its receptor to control calcium signaling during zebrafish epiboly. Science Signaling. 2014;7(312):ra14
  75. 75. Prudent J, Gillet G, Popgeorgiev N. Nrz but not zBcl-xL antagonizes Bcl-wav pro-apoptotic activity in zebrafish. Communicative & Integrative Biology. 2014;7(1):e28008
  76. 76. Prudent J et al. Bcl-2 proteins, cell migration and embryonic development: lessons from zebrafish. Cell Death & Disease. 2015;6:e1910
  77. 77. Cheng R et al. Zebrafish (Danio rerio) p53 tumor suppressor gene: cDNA sequence and expression during embryogenesis. Molecular Marine Biology and Biotechnology. 1997;6(2):88-97
  78. 78. Thisse C et al. The Mdm2 gene of zebrafish (Danio Rerio): Preferential expression during development of neural and muscular tissues, and absence of tumor formation after overexpression of its cDNA during early embryogenesis. Differentiation. 2000;66(2-3):61-70
  79. 79. Lee KC et al. Detection of the p53 response in zebrafish embryos using new monoclonal antibodies. Oncogene. 2008;27(5):629-640
  80. 80. Berghmans S et al. tp53 mutant zebrafish develop malignant peripheral nerve sheath tumors. Proceedings of the National Academy of Sciences of the United States of America. 2005;102(2):407-412
  81. 81. den Hertog J. Tumor suppressors in Zebrafish: From TP53 to PTEN and beyond. Advances in Experimental Medicine and Biology. 2016;916:87-101
  82. 82. Storer NY, Zon LI. Zebrafish models of p53 functions. Cold Spring Harbor Perspectives in Biology. 2010;2(8):a001123
  83. 83. Bladen CL et al. DNA damage response and Ku80 function in the vertebrate embryo. Nucleic Acids Research. 2005;33(9):3002-3010
  84. 84. Liu TX et al. Knockdown of zebrafish Fancd2 causes developmental abnormalities via p53-dependent apoptosis. Developmental Cell. 2003;5(6):903-914
  85. 85. Fischer S et al. Mutation of zebrafish caf-1b results in S phase arrest, defective differentiation, and p53-mediated apoptosis during organogenesis. Cell Cycle. 2007;6(23):2962-2969
  86. 86. Ekker SC, Larson JD. Morphant technology in model developmental systems. Genesis. 2001;30(3):89-93
  87. 87. Robu ME et al. p53 activation by knockdown technologies. PLoS Genetics. 2007;3(5):e78
  88. 88. Bill BR et al. A primer for morpholino use in zebrafish. Zebrafish. 2009;6(1):69-77
  89. 89. Hensey C, Gautier J. A developmental timer that regulates apoptosis at the onset of gastrulation. Mechanisms of Development. 1997;69(1-2):183-195
  90. 90. Alli Shaik A et al. Functional mapping of the zebrafish early embryo proteome and transcriptome. Journal of Proteome Research. 2014;13(12):5536-5550
  91. 91. Popgeorgiev N et al. The yolk cell of the zebrafish blastula harbors functional apoptosis machinery. Communicative & Integrative Biology. 2011;4(5):549-551
  92. 92. Kimmel CB, Law RD. Cell lineage of zebrafish blastomeres. II. Formation of the yolk syncytial layer. Developmental Biology. 1985;108(1):86-93

Written By

Nikolay Popgeorgiev, Benjamin Bonneau, Julien Prudent and Germain Gillet

Submitted: 29 December 2017 Reviewed: 26 January 2018 Published: 05 November 2018