Next Article in Journal
Editorial for the Special Issue “Bioactive Compounds of Natural Products on Metabolic Disorders and Complications”
Next Article in Special Issue
Effects of Di-(2-Ethylhexyl) Phthalate (DEHP) on Gamete Quality Parameters of Male Koi Carp (Cyprinus carpio)
Previous Article in Journal
PCID2 Subunit of the Drosophila TREX-2 Complex Has Two RNA-Binding Regions
Previous Article in Special Issue
Sperm Cryopreservation Today: Approaches, Efficiency, and Pitfalls
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

RNA Helicase Vasa as a Multifunctional Conservative Regulator of Gametogenesis in Eukaryotes

by
Vladimir E. Adashev
,
Alexei A. Kotov
and
Ludmila V. Olenina
*
Department of Molecular Mechanisms for Realization of Genetic Information, Laboratory of Biochemical Genetics of Animals, National Research Center “Kurchatov Institute”, Kurchatov Sq. 1, 123182 Moscow, Russia
*
Author to whom correspondence should be addressed.
Curr. Issues Mol. Biol. 2023, 45(7), 5677-5705; https://doi.org/10.3390/cimb45070358
Submission received: 18 June 2023 / Revised: 30 June 2023 / Accepted: 3 July 2023 / Published: 5 July 2023
(This article belongs to the Special Issue Reproductive Biology and Germ Cell Development)

Abstract

:
Being a conservative marker of germ cells across metazoan species, DEAD box RNA helicase Vasa (DDX4) remains the subject of worldwide investigations thanks to its multiple functional manifestations. Vasa takes part in the preformation of primordial germ cells in a group of organisms and contributes to the maintenance of germline stem cells. Vasa is an essential player in the piRNA-mediated silencing of harmful genomic elements and in the translational regulation of selected mRNAs. Vasa is the top hierarchical protein of germ granules, liquid droplet organelles that compartmentalize RNA processing factors. Here, we survey current advances and problems in the understanding of the multifaceted functions of Vasa proteins in the gametogenesis of different eukaryotic organisms, from nematodes to humans.

1. Introduction

The DEAD-box-containing RNA helicase Vasa (DDX4) is a well-known marker of animal germ cells. Using model and non-model organisms, it has been shown that Vasa is involved in germline specification, gamete production, and the piRNA silencing pathway that controls the activity of transposons and other harmful genomic elements in a large number of Metazoan species [1,2,3,4,5,6,7]. The vasa gene was initially discovered in Drosophila by screening for recessive lethal maternal effect mutations [8,9]. Like all members of the DEAD box superfamily, Vasa has a highly conservative core helicase region of about 400 amino acids, which contains at least 12 characteristic motifs in conservative positions, forming two RecA-like domains, the N-terminal DEAD-like helicase domain (DEXDc) and the C-terminal helicase superfamily domain (HELICc) (Figure 1). These domains bind both ATP and RNA and provide RNA helicase and ATPase activities [10,11]. The DEAD box designation for the superfamily originated from the amino acid sequence of the conservative catalytic site (DEAD) which is a part of motif II, also referred to as the Walker B motif. The core helicase region of Vasa is able to bind a six-nucleotide fragment along the sugar–phosphate backbone of target RNAs [12]. During the unwinding of double-stranded RNA duplexes, DEAD box helicases undergo a series of conformational rearrangements. The co-binding of RNA and ATP results in a significant conformational change in the helicase, bringing the RecA-like domains together in a closed state. Closing the interdomain cleft bends the RNA duplex due to steric hindrance with the help of a wedge-shaped helix formed by motif Ib. It leads to local denaturation of RNA strands. After the release of the first strand of RNA, ATP hydrolysis occurs. The subsequent release of phosphate is the rate-limiting step in the RNA unwinding cycle [13,14]. The N- and C-terminal amino acid sequences of Vasa proteins in different organisms vary considerably. Helicases containing the DEAD box are involved at all stages of cellular RNA metabolism.
Understanding the molecular functions of Vasa is crucial for elucidating important aspects of developmental biology. Here, we survey the different Vasa functions that have been revealed in the germline of model and non-model organisms over the past years. Despite the observed diversity of functional manifestations in different organisms, Vasa is considered a basic conservative factor that is necessary for the process of gametogenesis and, thus, for the maintenance of multiple eukaryotic species from generation to generation.

2. Vasa in Germ Cell Specification

There are two main mechanisms of germ cell specification in eukaryotes: preformation and induction [15]. Preformation is defined as a mechanism for germ cell specification driven by the transfer of maternally inherited molecular determinants in the form of ribonucleoprotein (RNP) granules. The RNP granules segregate from the embryonic cytoplasm at the posterior end and ensure the subsequent germline fate of primordial germ cells (PGCs) [15,16]. Drosophila is one of those animals that inherit maternal germline determinants through the preformation of PGCs. The preformation of PGCs also occurs in many vertebrates, such as the teleost fish Danio rerio, the tailless amphibian Xenopus laevis, and birds, as well as in the invertebrate worm Caenorhabditis elegans [15,17,18,19].
Pole cells (PGC precursors) in Drosophila are specified at the syncytial stage of embryonic cleavage, and their specialization is determined by the set of maternal proteins and transcripts that accumulate in the pole granules of the developing oocyte [20]. Mutations in the genes encoding the components of the pole granules disrupt the development of gonads and make the progeny of mutant female flies infertile. This phenomenon is known as the grandchildless effect. The products of many grandchildless genes are necessary for the correct assembly of the pole plasm. Among them, proteins Oskar (its short isoform) and Vasa are considered to be the most significant factors because, in their absence, pole granules are not organized, and developing embryos are not able to form posterior structures and specify PGCs [21,22,23,24]. The maternal protein Vasa accumulates in the pole granules at the posterior end of the developing oocyte starting from stage 10 of Drosophila oogenesis (Figure 2a). The recruitment of Vasa is dependent on the advanced location of short Oskar at the posterior pole of the oocyte after stage 9 [21,25].
Mammalian germ cell determination appears to be dependent on inducible cellular interactions during embryonic gastrulation rather than the transmission of cytoplasmic maternal factors. It is known as the induction process of germ cell specification [15,27]. For example, in mice, PGCs are formed from the proximal area of the epiblast near the extraembryonic ectoderm approximately 6.25 days after fertilization at the stage of early gastrulation [28,29]. The first factors of PGC specification found in mice are the transcription factors BLIMP1, PRDM14, and AP2-gamma. It is thought that a variety of signaling events combine to trigger the expression of these factors. Among them is the secretion of BMP4 signaling protein from nearby extraembryonic ectoderm cells and the simultaneous expression of WNT3 factor in the rear visceral endoderm of the epiblast [30,31]. Together, BLIMP1, PRDM14, and AP2-gamma promote the expression of certain germline-specific genes, such as Nanos3, and downregulate genes involved in somatic differentiation, such as those of the Hox family [30]. By day 7.5 of mouse embryonic development, at the stage of late gastrulation, complementary processes of repression of the somatic program and induction of germline genes lead to the establishment of 30–40 PGCs, which subsequently migrate to the genital ridge [32]. In mice, the zygotic Vasa/MVH protein begins to be expressed in migrating PGCs and exhibits sexual dimorphism in germ cells during gametogenesis [33,34,35,36]. The onset of MVH expression in mammals is comparable to the timing of zygotic expression of Vasa in Drosophila [37]. MVH and another DEAD box RNA helicase, GRTH/DDX25, are associated with perinuclear RNP granules in early germ cells of mouse males and are required for spermatogenesis and male fertility [15,35,38,39]. No matter what mechanism of germ cell specification is defined, germ cells in gonads can be identified by their specific expression of Vasa proteins.

3. The Role of Vasa in the Structure and Dynamics of Germ Granules

According to recent pioneering studies, germ granules, or nuage, are thought to form as biomolecular liquid-like condensates through phase separation from the remainder of the cytosol [40,41,42]. They lack a membrane-like physical barrier enclosing their content from the outside environment. Phase separation can occur when interactions between macromolecules (proteins and nucleic acids) are thermodynamically more favorable than interactions between these molecules and their solvent (e.g., cytosol). As a result, the molecules and the solvent separate into a condensed phase with a smaller volume and a dilute phase with a bigger volume. Thus, biomolecular condensates partially exclude the aqueous phase [43,44]. The phase separation of proteins is mainly due to weak electrostatic, hydrophobic, π–cationic, cation–cationic, and π–π interactions between neighboring molecules and can be achieved through the interaction of certain domains and the presence of extended regions of intrinsic disorders [43,45]. Intrinsically disordered regions (IDRs) have no defined tertiary structure but contain low-complexity amino acid sequences that can participate in the formation of dynamic multivalent networks that appear in vivo and in vitro as phase-separated droplets [42,43,44]. In addition, the phase separation of proteins can be modulated by post-translational modifications of proteins [46,47], interactions with nucleic acids, and changes in their physicochemical environment. The constituents of the liquid-like condensates are mobile and capable of rapid exchange with the cytoplasm [44].
Recent studies have shown that both prokaryotic and eukaryotic cells use phase separation mediated by DEAD-box-containing helicases with intrinsically disordered terminal domains to compartmentalize various RNA processing reactions. This represents an example of a conservative intracellular organization that has been maintained throughout evolution, from bacteria to humans [48]. A spontaneous condensation of monomers of human homologue Vasa, DDX4, into protein aggregates occurs upon transfection of the culture of somatic cells by DDX4 carrying construct. The number of aggregates and their sizes are regulated by the available concentration of the free protein monomers, whereas methylation of arginine residues in DDX4 RGG motifs, a reduced ionic strength, and an increase in temperature can lead to reversible dissociation of these structures [42]. It has been shown that the N-terminal IDR and RNA helicase domains of Bombyx mori Vasa are responsible for self-association and RNA binding, and all of them are required for liquid-like droplet formation in vitro and for nuage Vasa-body assembly in vivo [49]. DEAD box RNA helicases are thought to broadly control the formation and dynamics of phase-separated RNP granules. In the ATP- and RNA-bound state, these helicases form multivalent interactions with RNAs, which promote the phase separation of these complexes from the cytosol. ATP hydrolysis causes the release of target RNAs from complexes with helicases, thereby disrupting multivalence, and, as a result, provokes the translocation of components of the RNA-containing liquid-like condensates [42,48].
The germ granules of nematode C. elegans represent a set of individual condensates. These granules contain subdomains with different electron densities, including an electron-dense layer closest to the nuclear envelope and protrusions facing the cytoplasm [50]. There are at least four classes of condensates, each defined by two or more unique components: SIMR foci, Mutator foci, Z granules, and P granules [51,52,53,54,55,56] (Figure 3). The spatial organization of the granules changes dynamically during the development of the nematode. The P granule and Z granule components overlap significantly in PGCs; these granules are completely fused in early embryos and then separate again when reassembled near the PGC nuclei in 100-cell embryos and remain separated in adult germ cells [53,56]. The majority of research has focused on P granule activity as a liquid drop. P granule proteins are rapidly exchanged between the condensate and the cytoplasm and released into the cytoplasm under high temperatures [40]. P granules are formed by GLH helicases (homologues of Drosophila Vasa) and PGL proteins containing arginine-glycine-glycine (RGG) repeats [55,57]. Phenylalanine-glycine-glycine (FGG) repeats are abundant in GLH helicases and are also present in nucleoporins surrounding the central channel of nuclear pores [58,59]. FGG-repeat domains of nucleoporins form a reversible hydrogel matrix that works as the permeability barrier of nuclear pores [60]. Similar permeability characteristics are shared by P granules, and it has been proposed that they serve as extensions of the environment of nuclear pores [58].
Numerous nuage granules that form a network around the nuclei of nurse cells in Drosophila ovaries (Figure 2a) also contain components of the piRNA pathway, foremost endonucleases of the ARGONAUTE/PIWI family, Aubergine (Aub) and AGO3, and several members of the TUDOR family proteins. Vasa is a basic architectural factor of these granules [6,61]. piNG-body, a large nuage-associated organelle, is found in the perinuclear area of Drosophila spermatocytes together with small nuage granules [26] (Figure 2b–d). The piNG-body is considered the center of post-transcriptional piRNA silencing in spermatocytes owing to the compartmentalization and accumulation of components of the piRNA machinery. Vasa, Aub, Tudor, and some other proteins occupy the periphery of the piNG-body, whereas AGO3 is found in the core lobe [26,62] (Figure 2b–d). The piNG-bodies and small nuage granules are able to migrate along the external nuclear surface and rapidly exchange their components with the cytoplasm [63,64,65,66]. Arginine methylation of the amino-terminal domains of Vasa and Aub provided by arginine methyltransferase Capsuleen/PRMT5 is essential for the assembly of the piNG-body [26]. Disruption of Aub methylation leads to its delocalization into the cytoplasm, the absence of the piNG-bodies, and impaired piRNA biogenesis, while a perinuclear localization of Vasa in small nuage granules is preserved [26,46,47].
Early electron microscopic studies uncovered that the pole granules form round, membraneless, dense bodies 0.2–0.5 μm in diameter that contain fibrous material, ribosomes, and nucleic acids at the posterior pole of the developing oocyte in Drosophila (Figure 2a). These granules accumulate maternally stored mRNAs and proteins, which are essential for the establishment of PGCs during embryonic development [21,22,67,68]. The short isoform of Oskar plays a central role in the assembly of pole granules. Oskar contains IDR between its LOTUS and RNA-binding domains and a short region of low-complexity sequences. Owing to the presence of these domains, the short Oskar is able to form granules in the absence of other components in Drosophila and human cell cultures [69]. While Oskar is not conservative beyond Diptera, other germ granule protein components are present throughout the animal kingdom. Among them are the RNA helicase Vasa, the ARGONAUTE/PIWI family member endonuclease Aub, the founder of the TUDOR domain family Tudor (Figure 2a), factors of translational regulation Nanos, Pumilio, Dazl, and others [20]. A physical interaction of the HELICc domain of Vasa with the LOTUS domain of Oskar is shown [11,23]. The complex Oskar-Vasa provides instructive functions in the enrichment of several hundred specific mRNAs at the posterior pole [70,71]. The Tudor protein, possessing 11 Tudor domains, plays a structural role in the pole granule assembly by binding proteins containing dimethylated arginine residues, including Vasa and Aub [46,47]. Methyltransferase Capsuleen function is required for the assembly of pole granules [72,73]. A lack of arginine methylation of Vasa and Aub results in the grandchildless phenotype of the offspring, which does not form pole cells and develops into sterile adults. Drosophila pole granules appear to exhibit both liquid-like and hydrogel-like properties [20,69].
Two types of nuage granules, intermitochondrial cement (IMC) and piP-bodies, were found in mouse embryonic gonocytes, which separately contain key endonucleases of the piRNA machinery and proteins of the ARGONAUTE/PIWI family, MILI and MIWI2, respectively [74,75,76] (Figure 4). Both types of granules contain MVH, a murine homologue of Vasa. After meiosis in the cytoplasm of round spermatids in adult mice, MVH is concentrated in a single, large, perinuclear, electron-dense granule known as the chromatoid body (CB) (Figure 4). It has been found that MVH and the ARGONAUTE/PIWI family endonuclease MIWI are the two major components of murine CB [77]. These findings confirm the crucial role of Vasa in the development and maintenance of mammalian germ granules.
The results obtained on vertebrate and invertebrate organisms suggest that the formation of biomolecular condensates of liquid-like and hydrogel-like types is a conservative principle of the organization of germ granules throughout the animal kingdom, and RNA helicase Vasa is the essential factor for their development and maintenance.

4. Sexual Dimorphism of Vasa Functions and Expression in Different Animals

It has been shown that sexual dimorphism in Vasa functions or expression occurs in a variety of model animals and other species. In Drosophila, Vasa is the basic component of the pole plasma (or pole granules) accumulated at the posterior pole of developing oocytes. Females homozygous for the vasa null allele lay very few embryos, but they do not form germ cells, and adult offspring have abdominal defects [25,81]. However, adult males with the vasa loss-of-function mutations remain fertile for a short time after eclosion (Adashev et al., unpublished data). In Danio rerio, both males and females need DDX4 for the formation of germ cells and meiosis; however, vasa mutants only produce sterile males [4]. Sex-related differences in DDX4 expression during oogenesis and spermatogenesis have also been found for other fish, including medaka, gibel carp, tilapia, catfish, and European sea bass [82,83,84,85,86]. The Asian yellow pond turtle, Mauremys mutica, shows a dimorphism in Vasa expression depending on sex. In adult turtles, the level of vasa mRNA is much higher in the testes than in the ovaries [87]. In chickens, Gallus gallus domesticus, however, male and female embryonic gonads with DDX4 knockdown contain a reduced amount of PGCs, which confirms the need for Vasa functions to maintain PGCs in embryos of both sexes [88].
The mouse Vasa homologue, MVH, is detected in the cytoplasm of premeiotic germ cells in both males and females [89]. In male mice, loss of MVH function causes defects in the proliferation and differentiation of early germ cells into pachytene spermatocytes, and spermatogenesis is arrested at the premeiotic stages from leptotene to zygotene, subsequently leading to germ cell apoptosis [35,89,90]. However, in female mice, impaired expression of the Mvh gene does not affect the development of female PGCs, and females homozygous for the Mvh mutations are fertile at the same level as wild-type females [35]. MVH, together with MILI, TDRD9, and GASZ proteins, are found in nuage-like aggregates in the cytoplasm of primordial ovarian follicles shortly after birth. However, these aggregates disintegrate during growing and maturing oocytes. In females mutated in Mvh, Mili, or Gasz, there is a decrease in the piRNA production in primordial ovarian follicles, and derepression of retrotransposons is detected. Nevertheless, folliculogenesis appears to be normal in the mutant females, and disruption of the piRNA pathway does not result in sterility. Moreover, wild-type oogenesis is observed even in triple mutants for the Mvh, Mili, and Gasz genes [91]. Compared to oogenesis, the Mvh mutations have a more severe impact on mouse spermatogenesis [89,90]. The stronger regulation of oogenesis can be explained by the obvious differences in the impact of the Mvh mutations on spermatogenesis and oogenesis. Thus, in the oogenesis of mammals, as a result of strict selection, only some of the oocytes (10–20%) survive and develop into primordial follicles [92]. In addition, oocytes undergo prolonged arrest at prophase I of meiosis, during which effective repair of damaged sites can occur. Another possible explanation for germline independence from the Mvh expression and the piRNA pathway in female mice could be the absence of the fourth member of the PIWI/ARGONAUTE protein subfamily, PIWIL3. PIWIL3 is found in placental mammals but is absent in the Muridae family, which includes mice and rats [93]. PIWIL3 may perform some functions specific to female germ cells, as indicated by its high level of expression in bovine oocytes and in human ovaries at the prenatal stage [93,94]. In the absence of PIWIL3, expression of the oocyte-specific isoform of nuclease Dicer, which arose due to the insertion of a retrotransposon into the Dicer gene, was detected in female mice. This insertion ensures the expression of a truncated Dicer variant, which is more efficient in processing double-stranded RNAs into endogenous siRNAs [95]. The functions of MVH and the piRNA pathway in regulating the activity of transposable elements appear to be replaced by Dicer and the siRNA pathway for the passage of mouse oocytes during meiosis [96]. Other undiscovered mechanisms for maintaining oogenesis in the absence of Vasa may also exist, but further study is necessary to clarify them.

5. Vasa Functions in Drosophila

5.1. Multiple Functions of Vasa during Oogenesis

The multiplicity of Vasa functions is studied mostly in Drosophila melanogaster. Vasa expression in the germline is essential for embryonic development and PGC specification, as well as for ensuring the process of fly oogenesis [21,97,98]. The RNA helicase activity of Vasa in the ovarian germarium is critical for the continuous development of Drosophila germ cells during oogenesis [98,99]. The germarium is the apical region of the Drosophila ovary where germline stem cells (GSCs), somatic niche cap cells, and escort cells are found. GSCs divide in a self-renewal mode with the generation of one daughter cell as a stem cell and another cell that leaves the niche and becomes a cystoblast. The cystoblast enters the differentiation niche consisting of somatic escort cells, where it mitotically divides four times, forming a cyst of 16 cystocytes interconnected due to incomplete cytokinesis. The 16 cystocytes are separated into the egg chamber; an oocyte is specified at the basal end of the cyst; and the remaining 15 cells become nurse cells (Figure 2a). The molecular mechanisms of ovarian pathogenesis associated with a lack of Vasa function are the subject of intensive research [97,98,100,101,102]. Atrophy of the germarium, loss of GSCs, cystoblasts, and dividing cysts often occur in females with the vasa null mutation. The functions of Vasa in early oogenesis seem to be associated with the maintenance of self-renewing GSC divisions and do not depend on its helicase activity [103]. Only a small number of oocytes in vasa null-mutant females complete their development, while most of them degenerate at earlier stages [21,99]. vasa mutations also lead to premature reduction in egg chambers in ovarioles with aging [98]. However, the precise molecular role of Vasa in early oogenesis remains poorly understood.
Vasa is enriched in the perinuclear nuage granules of ovarian nurse cells and all premeiotic germ cells of the testes [6,26,61,63] (Figure 2a,b). Nuage granules are centers for post-transcriptional processing of piRNAs and piRNA-mediated silencing of harmful genomic elements, including transposons, genomic repeats, and some protein-coding genes [6,26,61,104]. Being the top hierarchical structural component of nuage, Vasa also facilitates the transfer of long piRNA precursors from the nucleus to nuage [105] and participates in the piRNA amplification cycle (also known as the ping-pong cycle) carried out by Aub and AGO3 proteins for efficient silencing of target transcripts [6,65,66,106,107]. Loss-of-function vasa mutations cause the disruption of nuage formation, leading to the dispersion of Aub and AGO3 as well as other components of the piRNA machinery in the cytoplasm and the violation of piRNA-mediated silencing in both the ovaries and testes [6,26,61,104,108]. In the ovaries of vasa mutants, a loss of the piRNA biogenesis in nurse cells leads to the activation of multiple mobile elements [6,109], whereas in the testes, transposon mobilization is less pronounced, but protein-coding Stellates genes become derepressed in spermatocytes of vasa mutant males [26,110,111], causing subsequent meiotic disorders and compromised male fertility [112,113]. Even a transient loss of Vasa expression only between oogenesis stages 2 and 6 leads to transposon derepression in ovarian nurse cells [97]. Mobilization of transposons in female germ cells of vasa mutant flies leads to transportation of their transcripts in the oocyte and integration into various loci of the genome. This induces DNA double-strand genomic breaks, leading to severe defects of nuclear chromatin integrity and replication stress, and this can also cause arrest of oogenesis through activation of the checkpoint kinase 2 (Chk2) signaling cascade and subsequent sterility [97,114,115]. It can also result in transposon transfer from mother to offspring embryos, where it can cause DNA breaks, chromatin damage, and the arrest of early embryogenesis [97,116]. The helicase activity of Vasa is shown to be essential for oogenesis and embryogenesis [97].
The localization of Vasa in nuage, as well as the recruitment of Aub and AGO3 in nuage, are independent of Vasa helicase activity, but this activity is required for the piRNA-mediated silencing of target transcripts [98]. The HELICc domain of Vasa is essential for its nuage and pole plasm localization, whereas the N-terminal nonconservative sequence of Vasa is critical for proper pole cell and abdomen formation [11]. In the case of a point mutation in the catalytic core site (DEAD → DQAD), resulting in a permanently closed Vasa conformation, preventing ATP hydrolysis and RNA release, the perinuclear localization of nuage granules is not maintained, they increase in size, and post-transcriptional piRNA silencing is impaired [66,98].

5.2. Vasa as a Translational Activator

Early studies have shown the involvement of Vasa in the translational activation of nanos, gurken, and mei-P26 mRNAs during oogenesis [99,100,101,117,118], although no direct physical interaction of nanos or gurken transcripts with Vasa has been shown. Vasa directly interacts with the translation initiation factor eIF5B (also known as dIF2) to positively regulate the translation of mRNAs required for oogenesis [100,101,102]. Disruption of Vasa interaction with eIF5B in functional vasa mutants results in reduced translation of Gurken, a component of the EGFR signaling pathway, leading to female sterility and phenotypes identical to vasa null mutations [100]. Vasa functions as a positive translational regulator of Mei-P26 in vivo by interacting with the U-rich element in the 3′-UTR of mei-P26 mRNA. Mei-P26 is essential to restrict the expansion and proliferation of germ cells during the early stages of oogenesis. While its expression is modest in GSCs, it is highly expressed in nurse cells of 16-cell cysts. The absence of Mei-P26 expression promotes early germ cell proliferation and aberrant development, which results in a tumor-like ovarian phenotype [119,120].
A cohort of translation initiation factors involved in the formation of the pre-initiation complex, eIF2, eIF3, and the eIF4E-4G cap-binding complex, is co-immunoprecipitated with Vasa from Drosophila ovaries [24,98]. In addition, it has been shown that Vasa genetically interacts with the translation initiation factor eIF4A. eIF4A is essential for the formation of PGCs in embryos [121]. However, the exact mechanism and many possible targets of Vasa as a translational regulator remain unidentified to date.
It appears that various intracellular Vasa populations are engaged in the regulation of translation and the piRNA silencing. Although it is unclear how this functional separation is governed in germ cells, post-translational modifications of Vasa may be important. Vasa could be a component of a previously discovered network of protein–protein interactions, including Aub, Rm62, PABP, and eIF3 proteins, that have been shown to control germ cell development [122,123].

5.3. The Role of Vasa in Pole Granule Assembly

The products of a large number of grandchildless genes are required for pole granule assembly in Drosophila (Figure 2a). However, the most significant components of the cascade are the products of the oskar, staufen, and vasa genes [21,124]. Vasa is attracted to the pole granules through the interaction of the Vasa HELICc domain (residues 463–621) with the C-terminal extended LOTUS domain (eLOTUS) of the short Oskar isoform [11,23]. Correct localization of Vasa at the posterior pole of the oocyte also requires the presence of an invariant tryptophan residue in the C-terminal position (W660 in Drosophila melanogaster), as well as the C-terminal motif 636–646 aa, which is highly conserved among many species and consists of acidic amino acid residues. Deletion of this motif results in an insufficient concentration of pole granules at the posterior pole and the subsequent absence of germ cells in developing embryos [103,125]. The replacement of the C-terminal tryptophan residue with glutamic acid (W660 → E) also leads to a loss of the ability of Vasa to maintain the formation of PGCs and embryonic pattern, disruption of gurken translation activation, and significantly reduced activity of Vasa in the piRNA biogenesis [125]. Both Vasa and Oskar are required for the localization and translation of specific mRNAs in the pole granules [20,126,127,128,129,130]. Nevertheless, a direct involvement of Vasa in the activation of their translation is not shown, since translation can be suppressed due to a violation of the assembly of the pole granules themselves. The proteins encoded by these mRNAs are essential for the patterning of embryos and the specification of future germ cells. Translation of nanos at the posterior pole of the embryo normally results in the creation of a gradient of morphogen Nanos, while the absence of maternal nanos mRNA in pole granules leads to defects in abdominal segmentation of the developing embryo and, in addition, disrupts the migration of pole cells into the developing gonads [128].

5.4. Vasa in Drosophila Testes

Despite the fact that Vasa is expressed in germ cells both in the ovaries and testes, the data on its functions in spermatogenesis are rather contradictory. The fertility of young males with vasa null mutations have allowed researchers to propose that this gene is not essential for spermatogenesis, which indicates in favor of the sexual dimorphism of Vasa functions in fruit flies [81,131]. However, data about the critical role of Vasa in the formation of nuage granules, piRNA biogenesis, and Stellate gene silencing clearly contradict this point of view [26,111,132]. Derepression of Stellate genes in the case of piRNA pathway disruption leads to the formation of needle-like Stellate protein aggregates, meiotic disorders, and a decrease in male fertility up to complete sterility [112,113]. Our data also suggest a premature loss of GSCs in the testes of vasa mutants (Adashev et al., unpublished data). It remains unclear whether the functions of Vasa in the piRNA pathway and in the maintenance of GSCs are mutually independent. Vasa participates in a wide range of cellular processes, which makes it challenging to learn and comprehend how its competencies are distributed in germ cells.

5.5. Vasa Is a Target of the piRNA Pathway in the Gonads of Interspecific Hybrids

Data from a recent study point to the potential participation of Vasa in interspecies reproductive isolation [133]. Earlier, piRNAs that are complementary to vasa transcripts were found in the testes of D. melanogaster males [134]. However, the biological significance of putative piRNA-mediated downregulation of Vasa, a key germ cell marker, a regulator of gametogenesis and gonad production, has remained unclear for a long time. Multiple repeated genomic regions, AT-chX loci, residing in the X chromosome of D. melanogaster were later discovered. They are found to be unique for the genome of D. melanogaster and to share a homology with two exons of the vasa gene [133]. AT-chX loci generate a significant amount of piRNAs, being one of the major piRNA clusters in the testes [133,135,136]. However, piRNA-mediated repression of vasa via decreasing mRNA or protein product levels is not observed. The identity between the AT-chX repeats and the vasa sequence (about 76%) does not seem to be enough for effective piRNA silencing. However, the homology between the AT-chX loci of D. melanogaster and the vasa sequences of closely related species Drosophila mauritiana (and other fruit flies of the clade simulans) is found to be much higher (more than 90%). piRNA-dependent repression in this case should be effective, i.e., the vasa allele of D. mauritiana potentially could be silenced by piRNAs produced from the AT-chX loci. This hypothesis was tested by analyzing testes of D. melanogaster/D. mauritiana interspecies hybrids. It is found that only mRNA and protein products of the vasa allele of D. melanogaster, but not those of D. mauritiana, are expressed in the testes of the hybrids (Figure 5a). Presumably owing to the repression of parental allele of vasa that are essential for germline development, most hybrid males have reduced testes containing a small number of germ cells and no mature gametes. In the testes of hybrids that possess perfect germline content and normal morphology, strong derepression of Stellate genes in spermatocytes is detected due to the absence of the Y-linked Su(Ste) piRNA cluster in the hybrid genome (Figure 5b) [133,136]. As previously shown, it disrupts germ cell passage throughout meiosis and subsequent sperm production [112]. Thus, male sterility of hybrid progeny from crossing D. melanogaster with closely related species can be caused by at least two circumstances: piRNA-mediated repression of the alien vasa allele and derepression of Stellate genes (Figure 5). These findings indicate the involvement of the piRNA pathway in reproductive isolation and speciation. Preserving the functionality of its protein product, the vasa gene of D. melanogaster apparently was subjected to selection pressure and accumulated a sufficient number of nucleotide substitutions to avoid silencing by AT-chX piRNAs.

6. Functional Homologues of Vasa in C. elegans

6.1. General Overview of GLH Proteins

The homologue of Vasa GLH-1 (Germ Line Helicase 1) protein is the first identified constitutional component of germ granules, also known as P granules, in the nematode C. elegans [1]. Similar to Vasa in Drosophila, GLH-1 is part of the protein core of germinal granules, which is conservative in metazoans. Other conservative P granule core proteins include members of the TUDOR and ARGONAUTE families that bind small RNAs. GLH-1 has three paralogs (GLH-2, GLH-3, and GLH-4), but only GLH-1 and GLH-2 contain all domains specific to Vasa proteins [59]. Vasa-defining domains include glycine-rich repeats at the amino end, the helicase core domain, and a carboxy-terminal negatively charged domain. Intrinsic disorder regions, or IDRs, in GLH proteins contain phenylalanine-glycine-glycine (FGG) repeats that facilitate P granule association with nuclear pores [58,59]. In addition to the IDRs and the helicase core, GLH proteins also contain several copies of the retroviral zinc finger (CCHC) type domain [137,138].

6.2. Peculiarities of the piRNA Pathway in Nematodes

C. elegans are hermaphrodites that undergo spermatogenesis in late larval stages and then switch to oogenesis as adults [139]. All stages of germ cell development are represented in a single animal, which makes C. elegans a unique model genetic system. In C. elegans, the piRNA pathway is carried out in a different way than it does in Drosophila and mammals; 21 nt-long piRNAs (21U-RNAs) enriched with uracil at the 5′ end interact with the PIWI subfamily protein PRG-1. These complexes function in the recognition of specific transcripts, ensuring their accumulation in P granules [140,141,142,143]. The sequences encoding more than 15,000 species of 21U-RNAs are located in two large genomic clusters on the chromosome IV, between protein-coding genes and in introns [143,144]. In contrast to long piRNA precursors needed for piRNA production in mice and flies, 21U-RNAs are transcribed as single transcription units about 25–26 nt long that are then shortened to 21 nt [142,144,145]. Transposons constitute only about 12% of the C. elegans genome, and most of them are non-mobile, making this model organism perfect for investigating alternative piRNA functions. While some 21U-RNAs are involved in transposon repression [141], the vast majority of them are complementary to transcripts of protein-coding genes [139,140,146,147,148]. It has been found that PRG-1 interacts with the transcripts of nearly all protein-coding genes expressed during oogenesis, requiring only a partial complementarity of piRNAs with target transcripts, with up to four mismatches being sufficient for the subsequent steps of regulation [146,148]. Although PRG-1 has a conservative catalytic motif DDH and can cut RNA in vitro, it usually does not cut its target transcripts in vivo [148].
The piRNA-dependent binding of PRG-1 to target mRNA triggers the recruitment of RNA-dependent RNA polymerase (RdRP) in association with the protein complex Mutator. This results in the production of endo-siRNAs, also known as 22G-RNAs, from nearby regions of target transcripts [139,147,148,149,150]. piRNA-dependent 22G-RNAs form complexes with WAGO proteins, a class of worm-specific proteins of the ARGONAUTE family [151]. These complexes can be translocated to the nucleus, where they bind to nascent transcripts and carry out epigenetic silencing across multiple generations, providing lysine methylation at position 9 of histone H3 (H3K9) and establishing heterochromatinization of corresponding genomic regions. They can also perform post-transcriptional silencing of mature mRNAs [139,147,148,149]. Only maternal, not paternal, 21U-RNAs are transmitted to the next generation and play an active role in 22G-RNA-mediated transgenerational inheritance. As a result, PRG-1 complexes with piRNAs function as a powerful control mechanism during oogenesis, ensuring transcript licensing for 22G-RNA-mediated silencing. Numerous aberrant 22G-RNAs derived from ribosomal RNA, histone genes, and some other annotated protein-coding genes are produced as a consequence of Prg-1 mutations. In Prg-1 mutants, the aberrant 22G-RNAs represent 45% of all WAGO class 22G-RNA reads, but only a very small portion in wild-type animals [152]. Aberrant silencing of histones, ribosomal RNA, and many protein-coding genes in the case of the disrupted piRNA pathway is associated with impaired worm fertility, not immediately but after multiple (up to 50) generations [152,153]. However, to date, it is unclear which molecular mechanisms lead to sterility [152]. Another ARGONAUTE protein, CSR-1, binds to a genetically different class of 22G-RNAs, and this complex is thought to work against the piRNA system and to support the stability of endogenous transcripts in germ cells [147,154,155,156,157].

6.3. Functional Features of Vasa-like GLH Proteins

Multiple studies of GLH protein functions in development have demonstrated that loss-of-function glh-1 mutant worms are viable at 20 °C but become sterile after multiple generations at 25 °C, a phenotype known as the “mortal germline” phenotype [158,159,160]. The loss of both maternal and zygotically encoded GLH-1 results in complete sterility at 26 °C. However, keeping at lower temperatures (16–24.5 °C) restores the fertility of most glh-1 mutant animals [159]. glh-4 null mutants show moderate fertility defects. glh-1; glh-2 and glh-1; glh-4 double mutants are practically sterile at any permissive temperature and demonstrate a significant loss or absence of germ cells, as well as an insignificant number of mature spermatozoa or their absence [59,158,159]. The ability for self-renewing division of GSCs is also impaired in glh-1 mutants [159].
RNA helicase GLH-1 has essential functions in the control of the dynamics and localization of the P granules during the development of C. elegans from embryos to adults, and it is implicated in the transgenerational inheritance of the RNAi interference system [160,161]. In early embryos, P granules are cytoplasmic and subsequently concentrated in PGCs. Prior to the first cell division, many P granule components are initially disseminated in the mature oocyte or fertilized embryo before being reassembled into the granules. P granules relocate to the nuclear periphery when zygotic transcription starts during embryogenesis and remain perinuclear almost all the time during germ cell development. In contrast to wild-type embryos, in which GLH-1-containing granules are disassembled during the pseudo-cleavage stage in early embryogenesis, many GLH-1-containing granules remain clearly visible in glh-1DQAD mutants with impaired ATP hydrolysis. Unlike in wild-type animals, GLH-1 granules in glh-1DQAD mutants are not sorted exclusively into PGCs but are randomly distributed among somatic and germ cells in early embryos [154]. glh-1 mutations impair the localization of other P granule components [59,158,159,162]. P granule components become almost completely dispersed in the case of the double mutants glh-1; glh-4 [159,162]. Binding of mRNAs by GLH-1 promotes localization of the transcripts and recruitment of proteins in the P granules, including members of the ARGONAUTE family. This requires the correct functioning of the GLH-1 core domain in the binding and subsequent hydrolysis of ATP in germ cells of both adults and embryos [59,162]. In the germ cells of glh-1DQAD adult worms, enlarged P granules containing GLH-1 are located in the cytoplasm, while their perinuclear localization is impaired. These enlarged granules also contain CSR-1 and PGL-1. It indicates that the formation of the P granules with the correct localization and component release dynamics requires the RNA-binding activity of GLH-1 [59,162].
glh mutant worms show no significant changes in piRNA expression level compared to wild-type animals. However, there is a noticeable reduction in the quantity of 22G-RNAs, both in glh-1; glh-4 double mutants and glh-1DQAD mutants. This implies that GLH-1 and GLH-4 are not required for piRNA biogenesis but rather promote the production of 22G-RNAs around piRNA-binding sites on target transcripts [55]. Both GLH-1 and GLH-4 facilitate piRNA-induced silencing and function redundantly in this process. The ARGONAUTE proteins PRG-1 and CSR-1 are localized in the P granule condensates on the outer surface of nuclear pores in adult worms (Figure 3) [59,162]. However, the biogenesis of 22G-RNAs, required for piRNA-mediated silencing, is associated with spatially distinct condensates known as the Mutator foci [53,150,163]. Earlier, it was unclear how piRNA targeting in the P granules causes the production of 22G-RNA in the Mutator foci to carry out transcriptional and post-transcriptional silencing. The Mutator foci are often located in close contact with the perinuclear P granules [163]. Recent work has shown that the Mutator foci are specifically incorporated into the P granules during spermatogenesis at the pachytene stage [157]. It can be proposed that GLH-1 is a factor that combines the initial targeting of transcripts by PRG-1 with the establishment of WAGO-dependent silencing since the physical interaction of GLH-1 with both PRG-1 and WAGO proteins has been confirmed [59,164]. P granules act as a control point for piRNA-mediated gene silencing, where harmful alien piRNA targets are recognized, while CSR-1 allows maintaining the expression of own essential genes. In loss-of-function glh-1 mutants, perinuclear P granules are not formed, resulting in the scattering of RNA interference machinery, including PRG-1, WAGO, and CSR-1 proteins, in the cytoplasm. In the absence of a compartmentalized environment achieved in the P granules, hundreds of normally repressed alien transcripts are derepressed, while hundreds of normally expressed own transcripts fail to bind with CSR-1, which results in their repression [55]. These results support the model according to which the binding and release of mRNAs from GLH-1 after ATP hydrolysis provide the influx and release of components that are located in the P granules. In this model, mRNA binding by GLH-1 facilitates multivalent interactions between mRNA, GLH-1, and other RNA-binding proteins that are attracted to the perinuclear P granules. Replacement of aspartic acid residue with glutamine in glh-1DQAD mutants, which prevent ATP hydrolysis, causes enhanced mRNA binding by GLH-1 in vitro and nonspecific association of ARGONAUTE proteins with 3′ ends of mRNA in vivo [164]. It should be noted that the DQAD mutation of vasa in germ cells of insects also leads to increased sizes of germ granules that accumulate Vasa and the PIWI protein AGO3 with impaired dynamics [66,98]. All of these results are consistent with the idea that Vasa-like helicases regulate the dynamics of germ granules in different species in a conservative manner. Thus, according to modern concepts, GLH-1 in worms and Vasa proteins in other species function as solvents for mRNAs in germ granules phase-separated from the cytoplasm [42], providing the accumulation of transcripts and maintenance of them in an unfolded state accessible for their scanning by complexes of ARGONAUTE protein with small RNAs. The functional similarity between GLH and Vasa proteins with respect to the assembly of germ granules is striking, given the features that distinguish the GLH proteins of C. elegans from most other Vasa homologues. The FG and FGG repeats are located in the glycine-rich amino-terminal regions of GLH proteins, in contrast to the RG and RGG repeats found in the Vasa protein of Drosophila and vertebrates. While RGG repeats presumably mediate the association of Vasa with mRNAs [165], FGG repeats may mediate the association of FGG-rich Vasa homologues with one another or with nuclear pore proteins that are also rich in FGG motifs [58,59]. In addition, all GLH proteins possess several (from two to six) CCHC-type zinc fingers not found in Vasa homologues of Drosophila and vertebrates. It is possible that zinc fingers are involved in RNA binding and compensate for the absence of RGG repeats in the GLH family. Note that the ability for self-renewing divisions of GSCs is impaired in glh-1 mutant worms [159]. This finding supports the conservative functions of Vasa and its orthologues in the maintenance and self-renewal of GSCs, as shown by Drosophila studies [21,97,98].

7. Vasa (MVH, Mouse Vasa Homologue) in Mice

7.1. Dynamics of MVH Expression in the Mouse Testes and the Structural Role of MVH in the Formation of Germ Granules

As previously mentioned, the specification of germ cells in mammals is not dependent on the inheritance of cytoplasmic maternal factors but occurs through the induction mechanism during embryonic gastrulation, when PGCs develop from pluripotent epiblast cells [15,27]. Spermatogenesis in mice is a sequence of events that tightly regulates the differentiation of germ cells in the epithelium of the seminiferous tubules. On the 12th day of embryonic development, during late gastrulation, about 30–40 PGCs migrate through embryonic tissues toward the genital ridge. After that, the following differentiation of PCGs into prospermatogonia (gonocytes) begins on the 12th day of embryonic development [32]. The fetal gonocytes preceding the formation of spermatogonial stem cells enter into a cell cycle arrest up to the second day of postnatal development. Then, on days 9–10 after birth, a part of the gonocytes develop into spermatogonial stem cells, while the rest are involved in the first rounds of spermatogenesis [166]. During the migration and differentiation of PGCs, germ RNP granules emerge in their cytoplasm [167]. There are two types of small germ granules abundant in mouse prenatal gonocytes that differ in their compositions: piP-bodies and intermitochondrial cement (IMC), while after birth the large chromatoid body (CB) starts to assemble in the late pachytene spermatocytes [75,76,78] (Figure 4). The dynamics of the appearance and disappearance of each type of granule have been studied. Thus, piP-bodies are only found in fetal gonocytes, where they are adjacent to IMC granules. IMC granules are found among mitochondrial clusters in fetal gonocytes, postnatal spermatogonia, and pachytene spermatocytes. In late pachytene spermatocytes, CB precursors appear and coexist with IMC granules. At first, they are small and concentrated in the perinuclear space. After meiotic division, IMC granules are no longer detected in haploid round spermatids, but CB is formed as a relatively large single structure adjacent to the nucleus and is detected during subsequent stages of spermatid differentiation, progressively decreasing in size and degrading [77,78]. Multiple protein components of germ granules are conservative across many species, among them DEAD-box-containing RNA helicase MVH (Mouse Vasa Homologue).
Both piP-bodies and IMC granules contain MVH, but the rest of their proteomic composition is distinct. piP-bodies contain components of somatic processing bodies and germ-cell-specific proteins of the piRNA pathway, MIWI2, and TDRD9. IMC granules contain other components of the piRNA pathway, MILI and TRDR1 [75,76]. MIWI2 and MILI are endonucleases of the ARGONAUTE/PIWI family and are key components of the piRNA system, ensuring silencing transposable elements in fetal gonocytes via the ping-pong amplification cycle and de novo DNA methylation of transposable elements in the genome [74,75,76] (Figure 4). TDRD9 and TRDR1 are Tudor-domain-containing proteins. MIWI2 loaded with guide piRNA translocates into the nucleus, where it participates in transcriptional silencing [74]. Thus, piP-bodies and IMC granules contain separate modules of the piRNA machinery, presumably interacting with each other through transient contacts. MVH appears to be an essential structural factor for both types of germ granules in fetal gonocytes.
Due to its relatively large size, CB was first identified under a light microscope in rat testes about 150 years ago [168], and since then it has been the subject of extensive studies. To date, mouse CB composition has been investigated by immunochemical and proteomic methods [77,167]. About 100 proteins have been identified as part of CB [77]. Most of them display RNA-binding activity, while others also participate in RNA metabolism [169]. MVH and MIWI, the third endonuclease of the piRNA pathway, are two major components of CB, together making up about 40% of its proteome. This indicates that CB is functionally involved in the piRNA pathway, and, indeed, transcripts of pachytene piRNA clusters were found in CB content. CB also contains components of the nonsense-mediated RNA decay pathway, RNA-binding proteins associated with translation, splicing complex proteins, proteins involved in microtubule-mediated transport, and other unclassified proteins [77]. Thus, in the testes, MVH is associated with germ granules of all three types, piP-bodies, IMC, and CB, apparently being the main architectural component for all of them. MVH disappears in elongated spermatids during CB degradation [89,90].
Two waves of active transcription are observed during spermatogenesis in mice, which coincide with the appearance of piP-bodies and CB (Figure 4). The first one occurs in the early stages of spermatogenesis, starting in fetal gonocytes and finishing in spermatogonia after birth. This is followed by a significant drop in the level of transcriptional activity during the entry of spermatogonia into meiosis and homologous recombination during meiotic entry (leptotenes and zygotenes) [170]. The next wave of active transcription begins during meiosis and gradually fades during the spermatid elongation stage, which includes chromatin compaction and the replacement of histones by protamines [171,172]. MIWI2 and MIWI expression patterns coincide with the corresponding waves of active transcription: MIWI2 is detected in fetal gonocytes and spermatogonia after birth, and MIWI is expressed at the last stages of meiosis and post-meiotically (Figure 4). Accordingly, prepachytene piRNAs are detected in the MIWI2 expression period, where they are involved mainly in suppressing the activity of mobile elements such as LINE1 and IAP [74], while pachytene piRNAs are detected in the MIWI expression period. The role of pachytene piRNAs in gametogenesis was mysterious until recently.

7.2. MVH Is Essential for Proper Spermatogenesis and the piRNA Pathway

Homozygous mutants for Mvh develop normally upon reaching adulthood and exhibit normal sexual behavior, but males become sterile [35]. Adult Mvh mutant males lack post-meiotic germ cells entirely, and their testes are underdeveloped (consisting of 4/5 of their wild-type size). At the same time, the testes of mice with the Mvh mutation at the age of two days do not differ morphologically from the heterozygous control, except for a slight decrease in the size of the seminiferous tubules. However, it has been shown that in the testes of Mvh mutant mice, apoptosis events in germ cells at the premeiotic stage occur 10 times more frequently compared to the heterozygous control. In addition, mutant males were found to have a low frequency of PGC proliferation during embryonic development [35]. There is much data suggesting that MVH is essential for spermatogenesis mainly due to its involvement in the piRNA pathway: first, as a direct participant in the piRNA biogenesis, and, second, as a factor that forms germ granules where this process takes place. Like Vasa in Drosophila gametogenesis, MVH acts as a basic component of germ granules. Male germ cells with the Mvh null mutation do not maintain the IMC granules, and MIWI2 localization in the piP-bodies is also impaired. The loss of MIWI2 from the piP-bodies leads to a significant drop in the amount of prepachytene piRNAs. Compared with the wild-type control, the total amount of piRNAs in male germ tissues decreases by five-fold in the background of the Mvh mutation. A decrease in the number of piRNAs mapped to repeated DNA sequences, including transposons, is also observed [173]. MVH ATPase activity is required for MIWI2-dependent transposon silencing. Catalytic mutant Mvh alleles (DEAD → DQAD) cause male infertility and derepression of the LINE1 retrotransposon, but not female sterility [174]. Activation of transposons in catalytically inactive Mvh mutants occurs due to a loss of prepachytene piRNAs associated with MIWI2. MILI is able to detect pre-piRNAs, bind to them, and perform slicing, but further processing with the generation of mature piRNAs is disrupted. The typical ping-pong signature in position 10 is entirely lost among MILI-linked piRNAs in catalytically inactive Mvh mutant males [174]. Since the nuclear localization of MIWI2 is licensed by its binding to piRNA [74], unloaded MIWI2 in fetal gonocytes of catalytically inactive Mvh mutants is not translocated into the nucleus [174]. This absence of MIWI2-associated piRNAs is phenocopied by that of Mvh null mutant males [173]. In norm, complexes of MIWI2 with piRNAs provide a control of de novo DNA methylation in genomic regions encoding transposable elements. In male mice, the methylation pattern is established in fetal gonocytes during cell cycle arrest. The DNA methylation process is disrupted by Mili, Miwi2, and Mvh mutations [74,173,174].
The pachytene piRNA subpopulation in the germline is the most abundant in the testes of adult males. It is known that pachytene piRNAs predominantly originate from long non-coding RNAs transcribed from intergenic regions [175]. As a rule, pachytene piRNAs are loaded into MIWI, which is expressed during the post-meiotic stages of spermatogenesis (Figure 4). According to recent studies, a significant number of mRNAs are targets of post-meiotic piRNA regulation, and the biogenesis of pachytene piRNAs occurs through the amplification ping-pong cycle [176]. There is a strict control over which mRNAs are the targets of pachytene piRNAs. When the human piRNA cluster was inserted into the mouse genome, this event caused a repression of the Dpy192 gene, which is required for spermatogenesis in mice, resulting in male sterility [176]. The deletion of the piRNA cluster pi18 also demonstrates that pachytene piRNAs function in the maintenance of mRNA homeostasis, which is required for the correct completion of spermatogenesis at the post-meiotic stage. Defects in the morphology of spermatozoa acrosomes and their motility are detected in such males [177]. Similar results were obtained in male mice lacking the pachytene piRNA cluster pi6 [178].
The requirement for MVH function at the pachytene stage has also been shown using the catalytically inactive Mvh mutation. Germ cells of mutant males are able to proceed through meiosis but then undergo an arrest of differentiation at the round spermatid stage, followed by apoptosis. This appears to be because the mutant population of MVH protein forms irreversible complexes with precursors of pachytene piRNAs, causing a disruption of the normal dynamics of the biogenesis and the functioning of pachytene piRNAs in spermiogenesis [174], similar to what has been shown earlier for Vasa in Bombyx mori [66]. Thus, MVH, as a key contributor to the piRNA biogenesis and piRNA-mediated silencing, is required for the repression of mobile elements with prepachytene piRNAs associated with MIWI2 and MILI, as well as for the correct regulation of genetic elements, represented in part by protein-coding genes, with pachytene piRNAs associated with MIWI and MILI.
In addition to its involvement in the piRNA pathway, MVH is also able to interact with a specific cohort of mRNAs. MVH in the mice testes is shown to associate with several hundred mRNAs, among which a large number are responsible for spermatogenesis directly or indirectly [179]. Interestingly, MVH targets included genes encoding various translational regulators, including eIF4B, and genes involved in the energy metabolism of germ cells. Presumably, MVH selectively interacts with multiple mRNAs, providing their storage as inactive RNP complexes in CB. A release of mRNAs from these complexes and their translocation to the cytoplasm for active translation can be triggered by acetylation of MVH and decreasing its RNA-binding activity [179]. Recently published mapping of MVH binding sites in single mouse oocytes reveals overrepresented GC-rich MVH-binding motifs in target transcripts [180].

7.3. Post-Translational Regulation of MVH in Mice

Post-translational regulation of MVH activity is mediated by the acetylation of the lysine residue K405. The acetylated form of MVH is found at the post-meiotic stage in the CB. Histone acetyltransferase-1 (Hat1), which is found in the CB along with its cofactor p46, is responsible for MVH acetylation. This post-translational modification reduces the RNA-binding activity of MVH, while its ATP binding and ATPase activities remain unchanged [179]. How MVH acetylation modulates its functions at the molecular level remains unknown. The methylation of arginine residues within the RGG motifs in the amino-terminal domain of MVH, which is conservative in Vasa orthologues from Drosophila to humans, allows binding to Tudor domains of a number of protein components of germ granules and appears to regulate their structure and dynamic exchange with the cytosol [42,46,47]. Symmetric and asymmetric dimethylation of two arginine residues, R62 and R105, have been shown for MVH [47].
In recent years, studies using a mouse model have resulted in significant advances in understanding both the regulatory mechanisms of the mammalian piRNA pathway and the critical function of RNA helicase Vasa (MVH) in the piRNA silencing associated with the maintenance of spermatogenesis at different stages. The clear manifestation of sexual dimorphic functions of MVH makes the mouse the most valuable model object for studying this phenomenon. The functions of MVH as a regulator of translation of certain mRNAs are also widely investigated, with the assumption that this function is conservative in higher eukaryotes, including humans.

8. Human VASA (DDX4)

Germ cells of human gonads develop from a homogeneous population of PGCs by the induction mechanism. In both male and female embryonic gonads, PGCs initially proliferate, but subsequently their fates differ: female germ cells enter meiotic prophase I at the 11th week of pregnancy, and then bind to somatic cells, forming primary follicles. In comparison, the entry of male germ cells into meiosis does not occur until puberty [181,182]. Like other mammals, humans undergo a massive loss of potential oocytes before folliculogenesis by programmed cell death [92]. It was previously thought that girls are born with a limited number of resting primordial follicles, but recent research has shown that many mammalian species, including humans, have postnatal ovaries that contain mitotically active germ cells [183].
The human genome encodes a single vasa orthologue, Ddx4. Human PGCs become Vasa-positive only after colonization of the gonadal ridge of the embryo. The germline markers DAZL and Vasa (DDX4) are expressed by human fetal germ cells, but their expression patterns differ over time. In fetal ovaries, a strong increase in the expression of Vasa and DAZL mRNAs and proteins are detected in the second trimester of pregnancy, while they are not expressed in gonocytes or oogonia in the first trimester. The marked increase in Vasa expression in fetal ovaries between 9 and 14 weeks of gestation coincides with the entry of female germ cells into meiosis. In prenatal ovaries, Vasa is found in the cytoplasm of larger and more mature germ cells than DAZL and is also present in oocytes within primordial follicles [34,182,184]. In fetal oocytes at the age of 35 weeks, Vasa demonstrates a heterogeneous intracellular distribution: in addition to diffuse staining of the cytoplasm, it is also detected in the granular material located around the nucleus. This arrangement corresponds to Balbiani bodies, which were originally described as germinal complexes consisting of mitochondria, endoplasmic reticulum, and lamellar aggregates and are now considered nuage-related germinal structures [45,185,186]. In some oocytes, Vasa is not associated with the Balbiani bodies, despite having a perinuclear localization in granules. The association of Vasa with the Balbiani bodies was previously thought to be a hallmark of fetal oocytes [34,187], but this was later shown in neonatal and pubertal ovaries [188]. As germ cells progress toward meiosis entry and primordial follicle formation, the absolute number of Vasa-positive cells decreases through programmed cell death, but Vasa is easily detected in the remaining cells [187]. In neonatal and pubertal human ovaries, the Vasa protein is also detected in the cytoplasm of oocytes in early primordial follicles in a perinuclear localization corresponding to the location of the Balbiani bodies. Primordial follicles also show a subnuclear localization of Vasa, but with a broader distribution pattern compared to primordial follicles. The Vasa protein is not detected in secondary follicles or at subsequent stages of folliculogenesis [188]. The testes of a human fetus at 17 weeks of age contain one to six Vasa-positive cells per cross-section of the seminiferous tubules. A higher proportion of germ cells in fetal ovaries compared to testes corresponds to a higher level of Vasa mRNA in ovaries at the same stage [34].
The roles of Vasa and its necessity for human oogenesis and spermatogenesis remain unknown, despite the availability of data on the expression of Vasa in human gonads. Considering that Vasa-containing germ granules are present in human germ cells, it would be reasonable to suppose that Vasa functions in human gametogenesis similarly to those of other mammals. Analysis of the expression of key components of the piRNA pathway has shown that human fetal oocytes express PIWIL2 and piRNAs that are mapped to transposons, but PIWIL1 or PIWIL4 are not expressed or are below the detection threshold. While PIWI proteins are not detected in the fetal ovaries during the first trimester, PIWIL2, the mouse MILI homologue, is highly expressed later in the fetal ovaries in Vasa-positive germ cells, with PIWIL2 and Vasa both being localized to the perinuclear granules [93]. This pattern is similar to that observed in mouse gonocytes. PIWIL1 (a murine MIWI homologue), PIWIL2, and Vasa are also expressed in oocytes of adult females, but they are not found in shared granular structures [93]. Therefore, the role of Vasa in the biogenesis of human piRNAs has not been directly demonstrated by experimentation. In different organisms, vasa/Ddx4 genetic ablation or mutations that destroy its helicase activity cause sexually dimorphic germline abnormalities, leading to either male or female sterility. Currently, there is no evidence to suggest the existence of sexual dimorphism in Vasa function in human gametogenesis. Interestingly, Vasa and its transcripts are found in the testes of adult patients with a number of disorders of spermatogenesis [189,190]. The Vasa signal in testis immunostaining preparations was found to be lower in classical seminoma and dysgerminoma than in spermatocyte seminoma. However, Vasa expression is not detected in the cases of non-seminomatous germ cell tumors, which allows for the use of Vasa as a marker for these diseases [191]. Some cases of idiopathic azoospermia or severe oligospermia are associated with hypermethylation of the Vasa promoter, which presumably reduces the expression level of its mRNA [192]. In general, the study of the functions of Vasa in human gametogenesis seems to require different approaches compared to classical model organisms such as mouse and Drosophila. The data obtained in mice on germ cell maintenance and spermatogenesis are not always reproducible in humans. Research on apes, such as macaques, appears to be promising, although it is very expensive [193,194].

9. Conclusions

A number of questions regarding the functions of Vasa and its homologues in the gametogenesis of many animals remain unanswered. Thus, the involvement of Vasa proteins in the maintenance of GSCs has been shown for D. melanogaster and C. elegans [21,97,98,159]. In all cases, however, it remains unknown whether this involvement is mediated by the piRNA pathway or different subpopulations of Vasa are required to maintain GSCs and piRNA-mediated silencing. The first assumption is supported by the need for other participants of the piRNA pathway to maintain GSCs in Drosophila presumably via action in the translation activation of specific mRNAs in the piRNA-dependent mode [122,123,195]. However, the involvement of Vasa in piRNA-independent regulation of translation of a number of morphogens and its direct interaction with translation initiation factors in Drosophila [24,98,100,101] can rather suggest the existence of functionally separated intracellular populations of Vasa. Another unresolved issue concerns the specificity of Vasa interaction with mRNA targets. It is known that conservative RecA-like domains of DEAD-box-containing helicases unwind RNA duplexes by interacting nonspecifically with the sugar–phosphate backbone of the targets [12]. Thus, the terminal domains of Vasa, which are not conservative in different animals, may be proposed to be responsible for the recognition of specific targets. The specificity of the interaction of Vasa and homologues with a number of germline transcripts has been shown for Drosophila, C. elegans, and mice [101,164,179,180], and some MVH-binding motifs in mRNAs have been found in mice [180]. CLIP-seq or RIP-seq libraries, with sufficient read depth, can serve as valuable material for identifying conservative and species-specific transcripts that are targets of Vasa regulation. Obtaining such data can be a difficult task, given the transient time of the processive interaction of DEAD box RNA helicases with RNAs, but the development of CLIP methods makes it possible to circumvent this problem by in vivo cross-linking of RNA-protein complexes. In some animals, specific interactions of Vasa homologues with RNA targets have not been found experimentally. It is assumed that in these cases, Vasa-like proteins in germ granules create a hub structure that receives transcripts from nucleus for unwinding, subsequent sorting, and direction into various processing pathways [196].
The greatest advances in the study of Vasa in recent years have occurred due to the discovery of the principles of organization of germ granules as liquid biocondensates phase separated from the cytosol and the basic role of Vasa proteins in their architecture and functional dynamics in a variety of metazoan species [42,48,49]. Emerging Vasa functions and mechanisms of action for ensuring gametogenesis remains to be unraveled in future studies.

Author Contributions

Conceptualization, L.V.O. and V.E.A.; writing—original draft preparation, L.V.O. and V.E.A.; writing—review and editing, L.V.O., V.E.A. and A.A.K.; project administration and supervision, L.V.O. All authors have read and agreed to the published version of the manuscript.

Funding

The study was supported by a grant from the Russian Science Foundation #22-74-00029, https://rscf.ru/en/project/22-74-00029/ accessed on 28 July 2022.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Roussell, D.L.; Bennett, K.L. glh-1, a germ-line putative RNA helicase from Caenorhabditis, has four zinc fingers. Proc. Natl. Acad. Sci. USA 1993, 90, 9300–9304. [Google Scholar] [CrossRef] [PubMed]
  2. Gustavson, E.A.; Wessel, G.M. Vasa genes: Emerging roles in the germ line and in multipotent cells. Bioessays 2010, 32, 626–637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Lasko, P. The DEAD-box helicase Vasa: Evidence for a multiplicity of functions in RNA processes and developmental biology. Biochim. Biophys. Acta 2013, 1829, 810–816. [Google Scholar] [CrossRef] [PubMed]
  4. Hartung, O.; Forbes, M.M.; Marlow, F.L. Zebrafish vasa is required for germ-cell differentiation and maintenance. Mol. Reprod. Dev. 2014, 81, 946–961. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Kotov, A.A.; Akulenko, N.V.; Kibanov, M.V.; Olenina, L.V. Dead-box RNA helicases in animal gametogenesis. Mol. Biol. 2014, 48, 16–28. [Google Scholar] [CrossRef]
  6. Malone, C.D.; Brennecke, J.; Dus, M.; Stark, A.; McCombie, W.R.; Sachidanandam, R.; Hannon, G.J. Specialized piRNA pathways act in germline and somatic tissues of the Drosophila ovary. Cell 2009, 137, 522–535. [Google Scholar] [CrossRef] [Green Version]
  7. Ramat, A.; Simonelig, M. Functions of PIWI Proteins in gene regulation: New arrows added to the piRNA quiver. Trends Genet. 2021, 37, 188–200. [Google Scholar] [CrossRef]
  8. Schupbach, T.; Wieschaus, E. Germline autonomy of maternal-effect mutations altering the embryonic body pattern of Drosophila. Dev. Biol. 1986, 113, 443–448. [Google Scholar] [CrossRef]
  9. Schupbach, T.; Wieschaus, E. Female sterile mutations on the second chromosome of Drosophila melanogaster. I. Maternal effect mutations. Genetics 1989, 121, 101–117. [Google Scholar] [CrossRef]
  10. Linder, P.; Jankowsky, E. From unwinding to clamping—The DEAD box RNA helicase family. Nat. Rev. Mol. Cell Biol. 2011, 12, 505–516. [Google Scholar] [CrossRef] [Green Version]
  11. Wang, S.C.; Hsu, H.J.; Lin, G.W.; Wang, T.F.; Chang, C.C.; Lin, M.D. Germ plasm localisation of the HELICc of Vasa in Drosophila: Analysis of domain sufficiency and amino acids critical for localisation. Sci. Rep. 2015, 5, 14703. [Google Scholar] [CrossRef] [Green Version]
  12. Andersen, C.B.; Ballut, L.; Johansen, J.S.; Chamieh, H.; Nielsen, K.H.; Oliveira, C.L.; Pedersen, J.S.; Séraphin, B.; Le Hir, H.; Andersen, G.R. Structure of the exon junction core complex with a trapped DEAD-box ATPase bound to RNA. Science 2006, 313, 1968–1972. [Google Scholar] [CrossRef] [PubMed]
  13. Sengoku, T.; Nureki, O.; Nakamura, A.; Kobayashi, S.; Yokoyama, S. Structural basis for RNA unwinding by the DEAD-box protein Drosophila Vasa. Cell 2006, 125, 287–300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Liu, F.; Putnam, A.; Jankowsky, E. ATP hydrolysis is required for DEAD-box protein recycling but not for duplex unwinding. Proc. Natl. Acad. Sci. USA 2008, 105, 20209–20214. [Google Scholar] [CrossRef] [PubMed]
  15. Hansen, C.L.; Pelegri, F. Primordial germ cell specification in vertebrate embryos: Phylogenetic distribution and conserved molecular features of preformation and induction. Front. Cell Dev. Biol. 2021, 9, 730332. [Google Scholar] [CrossRef]
  16. Houston, D.W.; King, M.L. A critical role for Xdazl, a germ plasmlocalized RNA, in the differentiation of primordial germ cells in Xenopus. Development 2000, 127, 447–456. [Google Scholar] [CrossRef]
  17. Karimi, K.; Fortriede, J.D.; Lotay, V.S.; Burns, K.A.; Wang, D.Z.; Fisher, M.E.; Pells, T.J.; James-Zorn, C.; Wang, Y.; Ponferrada, V.G.; et al. Xenbase: A genomic, epigenomic and transcriptomic model organism database. Nucleic Acids Res. 2018, 46, D861–D868. [Google Scholar] [CrossRef]
  18. Ruzicka, L.; Howe, D.G.; Ramachandran, S.; Toro, S.; Van Slyke, C.E.; Bradford, Y.M.; Eagle, A.; Fashena, D.; Frazer, K.; Kalita, P.; et al. The Zebrafish Information Network: New support for non-coding genes, richer Gene Ontology annotations and the Alliance of Genome Resources. Nucleic Acids Res. 2019, 47, D867–D873. [Google Scholar] [CrossRef] [Green Version]
  19. Harris, T.W.; Arnaboldi, V.; Cain, S.; Chan, J.; Chen, W.J.; Cho, J.; Davis, P.; Gao, S.; Grove, C.A.; Kishore, R.; et al. WormBase: A modern model organism information resource. Nucleic Acids Res. 2020, 48, D762–D767. [Google Scholar] [CrossRef]
  20. Trcek, T.; Lehmann, R. Germ granules in Drosophila. Traffic 2019, 20, 650–660. [Google Scholar] [CrossRef] [Green Version]
  21. Lasko, P.F.; Ashburner, M. Posterior localization of vasa protein correlates with, but is not sufficient for, pole cell development. Genes Dev. 1990, 4, 905–921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Hay, B.; Jan, L.Y.; Jan, Y.N. A protein component of Drosophila polar granules is encoded by vasa and has extensive sequence similarity to ATP-dependent helicases. Cell 1988, 55, 577–587. [Google Scholar] [CrossRef] [PubMed]
  23. Jeske, M.; Muller, C.W.; Ephrussi, A. The LOTUS domain is a conserved DEAD-box RNA helicase regulator essential for the recruitment of Vasa to the germ plasm and nuage. Genes Dev. 2017, 31, 939–952. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Bansal, P.; Madlung, J.; Schaaf, K.; Macek, B.; Bono, F. An interaction network of RNA-binding proteins involved in Drosophila oogenesis. Mol. Cell Proteom. 2020, 19, 1485–1502. [Google Scholar] [CrossRef]
  25. Hay, B.; Jan, L.Y.; Jan, Y.N. Localization of vasa, a component of Drosophila polar granules, in maternal-effect mutants that alter embryonic anteroposterior polarity. Development 1990, 109, 425–433. [Google Scholar] [CrossRef] [PubMed]
  26. Kibanov, M.V.; Egorova, K.S.; Ryazansky, S.S.; Sokolova, O.A.; Kotov, A.A.; Olenkina, O.M.; Stolyarenko, A.D.; Gvozdev, V.A.; Olenina, L.V. A novel organelle, the piNG-body, in the nuage of Drosophila male germ cells is associated with piRNA-mediated gene silencing. Mol. Biol. Cell 2011, 22, 3410–3419. [Google Scholar] [CrossRef]
  27. Lawson, K.A.; Dunn, N.R.; Roelen, B.A.; Zeinstra, L.M.; Davis, A.M.; Wright, C.V.; Korving, J.P.; Hogan, B.L. Bmp4 is required for the generation of primordial germ cells in the mouse embryo. Genes Dev. 1999, 13, 424–436. [Google Scholar] [CrossRef]
  28. Kojima, Y.; Kaufman-Francis, K.; Studdert, J.B.; Steiner, K.A.; Power, M.D.; Loebel, D.A.; Jones, V.; Hor, A.; de Alencastro, G.; Logan, G.J.; et al. The transcriptional and functional properties of mouse epiblast stem cells resemble the anterior primitive streak. Cell Stem Cell 2014, 14, 107–120. [Google Scholar] [CrossRef] [Green Version]
  29. Lawson, K.A.; Hage, W.J. Clonal analysis of the origin of primordial germ cells in the mouse. Ciba Found. Symp. 1994, 182, 68–84. [Google Scholar] [CrossRef]
  30. Magnúsdóttir, E.; Dietmann, S.; Murakami, K.; Günesdogan, U.; Tang, F.; Bao, S.; Diamanti, E.; Lao, K.; Gottgens, B.; Azim Surani, M. A tripartite transcription factor network regulates primordial germ cell specification in mice. Nat. Cell Biol. 2013, 15, 905–915. [Google Scholar] [CrossRef] [Green Version]
  31. Magnúsdóttir, E.; Surani, M.A. How to make a primordial germ cell. Development 2014, 141, 245–252. [Google Scholar] [CrossRef] [Green Version]
  32. Nicholls, P.K.; Schorle, H.; Naqvi, S.; Hu, Y.C.; Fan, Y.; Carmell, M.A.; Dobrinski, I.; Watson, A.L.; Carlson, D.F.; Fahrenkrug, S.C.; et al. Mammalian germ cells are determined after PGC colonization of the nascent gonad. Proc. Natl. Acad. Sci. USA 2019, 116, 25677–25687. [Google Scholar] [CrossRef] [PubMed]
  33. Fujiwara, Y.; Komiya, T.; Kawabata, H.; Sato, M.; Fujimoto, H.; Furusawa, M.; Noce, T. Isolation of a DEAD-family protein gene that encodes a murine homolog of Drosophila vasa and its specific expression in germ cell lineage. Proc. Natl. Acad. Sci. USA 1994, 91, 12258–12262. [Google Scholar] [CrossRef] [PubMed]
  34. Castrillon, D.H.; Quade, B.J.; Wang, T.Y.; Quigley, C.; Crum, C.P. The human VASA gene is specifically expressed in the germ cell lineage. Proc. Natl. Acad. Sci. USA 2000, 97, 9585–9590. [Google Scholar] [CrossRef] [PubMed]
  35. Tanaka, S.S.; Toyooka, Y.; Akasu, R.; Katoh-Fukui, Y.; Nakahara, Y.; Suzuki, R.; Yokoyama, M.; Noce, T. The mouse homolog of Drosophila Vasa is required for the development of male germ cells. Genes Dev. 2000, 14, 841–853. [Google Scholar] [CrossRef]
  36. Noce, T.; Okamoto-Ito, S.; Tsunekawa, N. Vasa homolog genes in mammalian germ cell development. Cell Struct. Fun. 2001, 26, 131–136. [Google Scholar] [CrossRef] [Green Version]
  37. Van Doren, M.; Williamson, A.L.; Lehmann, R. Regulation of zygotic gene expression in Drosophila primordial germ cells. Curr. Biol. 1998, 8, 243–246. [Google Scholar] [CrossRef] [Green Version]
  38. Tsai-Morris, C.-H.; Sheng, Y.; Gutti, R.K.; Tang, P.-Z.; Dufau, M.L. Gonadotropin-regulated testicular RNA helicase (GRTH/DDX25): A multifunctional protein essential for spermatogenesis. J. Androl. 2010, 31, 45–52. [Google Scholar] [CrossRef]
  39. Tutak, K.; Rozwadowska, N. Discrete roles of RNA helicases in human male germline and spermatogenesis. J. Appl. Genet. 2020, 61, 415–419. [Google Scholar] [CrossRef]
  40. Brangwynne, C.P.; Eckmann, C.R.; Courson, D.S.; Rybarska, A.; Hoege, C.; Gharakhani, J.; Jülicher, F.; Hyman, A.A. Germline P granules are liquid droplets that localize by controlled dissolution/condensation. Science 2009, 324, 1729–1732. [Google Scholar] [CrossRef]
  41. Hyman, A.A.; Brangwynne, C.P. Beyond stereospecificity: Liquids and mesoscale organization of cytoplasm. Dev. Cell 2011, 21, 14–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Nott, T.J.; Petsalaki, E.; Farber, P.; Jervis, D.; Fussner, E.; Plochowietz, A.; Craggs, T.D.; Bazett-Jones, D.P.; Pawson, T.; Forman-Kay, J.D.; et al. Phase transition of a disordered nuage protein generates environmentally responsive membraneless organelles. Mol. Cell. 2015, 57, 936–947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Banani, S.F.; Lee, H.O.; Hyman, A.A.; Rosen, M.K. Biomolecular condensates: Organizers of cellular biochemistry. Nat. Rev. Mol. Cell Biol. 2017, 18, 285–298. [Google Scholar] [CrossRef]
  44. So, C.; Cheng, S.; Schuh, M. Phase separation during germline development. Trends Cell Biol. 2021, 31, 254–268. [Google Scholar] [CrossRef] [PubMed]
  45. Dobrynin, M.A.; Bashendjieva, E.O.; Enukashvily, N.I. Germ granules in animal oogenesis. J. Dev. Biol. 2022, 10, 43. [Google Scholar] [CrossRef]
  46. Kirino, Y.; Vourekas, A.; Sayed, N.; de Lima Alves, F.; Thomson, T.; Lasko, P.; Rappsilber, J.; Jongens, T.A.; Mourelatos, Z. Arginine methylation of Aubergine mediates Tudor binding and germ plasm localization. Rna 2010, 16, 70–78. [Google Scholar] [CrossRef] [Green Version]
  47. Kirino, Y.; Vourekas, A.; Kim, N.; de Lima Alves, F.; Rappsilber, J.; Klein, P.S.; Jongens, T.A.; Mourelatos, Z. Arginine methylation of vasa protein is conserved across phyla. J. Biol. Chem. 2010, 285, 8148–8154. [Google Scholar] [CrossRef] [Green Version]
  48. Hondele, M.; Sachdev, R.; Heinrich, S.; Wang, J.; Vallotton, P.; Fontoura, B.M.A.; Weis, K. DEAD-box ATPases are global regulators of phase-separated organelles. Nature 2019, 573, 144–148. [Google Scholar] [CrossRef]
  49. Yamazaki, H.; Namba, Y.; Kuriyama, S.; Nishida, K.M.; Kajiya, A.; Siomi, M.C. Bombyx Vasa sequesters transposon mRNAs in nuage via phase separation requiring RNA binding and self-association. Nat. Commun. 2023, 14, 1942. [Google Scholar] [CrossRef]
  50. Sheth, U.; Pitt, J.; Dennis, S.; Priess, J.R. Perinuclear P granules are the principal sites of mRNA export in adult C. elegans germ cells. Development 2010, 137, 1305–1314. [Google Scholar] [CrossRef] [Green Version]
  51. Ouyang, J.P.T.; Seydoux, G. Nuage condensates: Accelerators or circuit breakers for sRNA silencing pathways? Rna 2022, 28, 58–66. [Google Scholar] [CrossRef] [PubMed]
  52. Manage, K.I.; Rogers, A.K.; Wallis, D.C.; Uebel, C.J.; Anderson, D.C.; Nguyen, D.A.H.; Arca, K.; Brown, K.C.; Cordeiro Rodrigues, R.J.; de Albuquerque, B.F.; et al. A tudor domain protein, SIMR-1, promotes siRNA production at piRNA-targeted mRNAs in C. elegans. eLife 2020, 9, e56731. [Google Scholar] [CrossRef] [PubMed]
  53. Wan, G.; Fields, B.D.; Spracklin, G.; Shukla, A.; Phillips, C.M.; Kennedy, S. Spatiotemporal regulation of liquid-like condensates in epigenetic inheritance. Nature 2018, 557, 679–683. [Google Scholar] [CrossRef] [PubMed]
  54. Sundby, A.E.; Molnar, R.I.; Claycomb, J.M. Connecting the dots: Linking Caenorhabditis elegans small RNA pathways and germ granules. Trends Cell Biol. 2021, 31, 387–401. [Google Scholar] [CrossRef]
  55. Chen, W.; Brown, J.S.; He, T.; Wu, W.S.; Tu, S.; Weng, Z.; Zhang, D.; Lee, H.C. GLH/VASA helicases promote germ granule formation to ensure the fidelity of piRNA-mediated transcriptome surveillance. Nat. Commun. 2022, 13, 5306. [Google Scholar] [CrossRef]
  56. Phillips, C.M.; Updike, D.L. Germ granules and gene regulation in the Caenorhabditis elegans germline. Genetics 2022, 220, iyab195. [Google Scholar] [CrossRef]
  57. Updike, D.; Strome, S. P granule assembly and function in Caenorhabditis elegans germ cells. J. Androl. 2010, 31, 53–60. [Google Scholar] [CrossRef]
  58. Updike, D.L.; Hachey, S.J.; Kreher, J.; Strome, S. P granules extend the nuclear pore complex environment in the C. elegans germ line. J. Cell Biol. 2011, 192, 939–948. [Google Scholar] [CrossRef] [Green Version]
  59. Marnik, E.A.; Fuqua, J.H.; Sharp, C.S.; Rochester, J.D.; Xu, E.L.; Holbrook, S.E.; Updike, D.L. Germline maintenance through the multifaceted activities of GLH/Vasa in Caenorhabditis elegans P granules. Genetics 2019, 213, 923–939. [Google Scholar] [CrossRef] [Green Version]
  60. Hülsmann, B.B.; Labokha, A.A.; Görlich, D. The permeability of reconstituted nuclear pores provides direct evidence for the selective phase model. Cell 2012, 150, 738–751. [Google Scholar] [CrossRef] [Green Version]
  61. Lim, A.K.; Kai, T. Unique germ-line organelle, nuage, functions to repress selfish genetic elements in Drosophila melanogaster. Proc. Natl. Acad. Sci. USA 2007, 104, 6714–6719. [Google Scholar] [CrossRef] [PubMed]
  62. Ryazansky, S.S.; Kotov, A.A.; Kibanov, M.V.; Akulenko, N.V.; Korbut, A.P.; Lavrov, S.A.; Gvozdev, V.A.; Olenina, L.V. RNA helicase Spn-E is required to maintain Aub and AGO3 protein levels for piRNA silencing in the germline of Drosophila. Eur. J. Cell Biol. 2016, 95, 311–322. [Google Scholar] [CrossRef] [PubMed]
  63. Snee, M.J.; Macdonald, P.M. Live imaging of nuage and polar granules: Evidence against a precursor-product relationship and a novel role for Oskar in stabilization of polar granule components. J. Cell Sci. 2004, 117, 2109–2120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Nosov, G.A.; Kibanov, M.V.; Olenina, L.V. Dynamic properties of germinal granule piNG-body in the testes of Drosophila melanogaster. Mol. Biol. 2014, 48, 805–813. [Google Scholar] [CrossRef]
  65. Webster, A.; Li, S.; Hur, J.K.; Wachsmuth, M.; Bois, J.S.; Perkins, E.M.; Patel, D.J.; Aravin, A.A. Aub and Ago3 are recruited to nuage through two mechanisms to form a ping-pong complex assembled by Krimper. Mol. Cell 2015, 59, 564–575. [Google Scholar] [CrossRef] [Green Version]
  66. Xiol, J.; Spinelli, P.; Laussmann, M.A.; Homolka, D.; Yang, Z.; Cora, E.; Couté, Y.; Conn, S.; Kadlec, J.; Sachidanandam, R.; et al. RNA clamping by Vasa assembles a piRNA amplifier complex on transposon transcripts. Cell 2014, 157, 1698–1711. [Google Scholar] [CrossRef] [Green Version]
  67. Mahowald, A.P. Fine structure of pole cells and polar granules in Drosophila melanogaster. J. Exp. Zool. 1962, 151, 201. [Google Scholar] [CrossRef]
  68. Mahowald, A.P. Polar granules of Drosophila. II. Ultrastructural changes during early embryogenesis. J. Exp. Zool. 1968, 167, 237–261. [Google Scholar] [CrossRef]
  69. Kistler, K.E.; Trcek, T.; Hurd, T.R.; Chen, R.; Liang, F.X.; Sall, J.; Kato, M.; Lehmann, R. Phase transitioned nuclear Oskar promotes cell division of Drosophila primordial germ cells. eLife 2018, 7, e37949. [Google Scholar] [CrossRef]
  70. Frise, E.; Hammonds, A.S.; Celniker, S.E. Systematic image-driven analysis of the spatial Drosophila embryonic expression landscape. Mol. Syst. Biol. 2010, 6, 345. [Google Scholar] [CrossRef]
  71. Siddiqui, N.U.; Li, X.; Luo, H.; Karaiskakis, A.; Hou, H.; Kislinger, T.; Westwood, J.T.; Morris, Q.; Lipshitz, H.D. Genome-wide analysis of the maternal-to-zygotic transition in Drosophila primordial germ cells. Genome Biol. 2012, 13, R11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Gonsalvez, G.B.; Rajendra, T.K.; Tian, L.; Matera, A.G. The Sm-protein methyltransferase, dart5, is essential for germ-cell specification and maintenance. Curr. Biol. 2006, 16, 1077–1089. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Anne, J. Arginine methylation of SmB is required for Drosophila germ cell development. Development 2010, 137, 2819–2828. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Aravin, A.A.; Sachidanandam, R.; Bourc’his, D.; Schaefer, C.; Pezic, D.; Toth, K.F.; Bestor, T.; Hannon, G.J. A piRNA pathway primed by individual transposons is linked to de novo DNA methylation in mice. Mol. Cell. 2008, 31, 785–799. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Aravin, A.A.; van der Heijden, G.W.; Castañeda, J.; Vagin, V.V.; Hannon, G.J.; Bortvin, A. Cytoplasmic compartmentalization of the fetal piRNA pathway in mice. PLoS Genet. 2009, 5, e1000764. [Google Scholar] [CrossRef] [Green Version]
  76. Shoji, M.; Tanaka, T.; Hosokawa, M.; Reuter, M.; Stark, A.; Kato, Y.; Kondoh, G.; Okawa, K.; Chujo, T.; Suzuki, T.; et al. The TDRD9-MIWI2 complex is essential for piRNA-mediated retrotransposon silencing in the mouse male germline. Dev. Cell 2009, 17, 775–787. [Google Scholar] [CrossRef] [Green Version]
  77. Meikar, O.; Vagin, V.V.; Chalmel, F.; Sõstar, K.; Lardenois, A.; Hammell, M.; Jin, Y.; Da Ros, M.; Wasik, K.A.; Toppari, J.; et al. An atlas of chromatoid body components. Rna 2014, 20, 483–495. [Google Scholar] [CrossRef] [Green Version]
  78. Meikar, O.; Da Ros, M.; Kotaja, N. Epigenetic regulation of male germ cell differentiation. Subcell. Biochem. 2013, 61, 119–138. [Google Scholar] [CrossRef]
  79. Werner, A.; Piatek, M.J.; Mattick, J.S. Transpositional shuffling and quality control in male germ cells to enhance evolution of complex organisms. Ann. N. Y. Acad Sci. 2015, 1341, 156–163. [Google Scholar] [CrossRef] [Green Version]
  80. Ernst, C.; Odom, D.T.; Kutter, C. The emergence of piRNAs against transposon invasion to preserve mammalian genome integrity. Nat. Commun. 2017, 8, 1411. [Google Scholar] [CrossRef] [Green Version]
  81. Lasko, P.F.; Ashburner, M. The product of the Drosophila gene vasa is very similar to eukaryotic initiation factor-4A. Nature 1988, 335, 611–617. [Google Scholar] [CrossRef] [PubMed]
  82. Shinomiya, A.; Tanaka, M.; Kobayashi, T.; Nagahama, Y.; Hamaguchi, S. The vasa-like gene, olvas, identifies the migration path of primordial germ cells during embryonic body formation stage in the medaka, Oryzias latipes. Dev. Growth Differ. 2000, 42, 317–326. [Google Scholar] [CrossRef] [PubMed]
  83. Xu, H.; Gui, J.; Hong, Y. Differential expression of vasa RNA and protein during spermatogenesis and oogenesis in the gibel carp (Carassius auratus gibelio), a bisexually and gynogenetically reproducing vertebrate. Dev. Dyn. 2005, 233, 872–882. [Google Scholar] [CrossRef] [PubMed]
  84. Kobayashi, T.; Kajiura-Kobayashi, H.; Nagahama, Y. Differential expression of vasa homologue gene in the germ cells during oogenesis and spermatogenesis in a teleost fish, tilapia, Oreochromis niloticus. Mech. Dev. 2000, 99, 139–142. [Google Scholar] [CrossRef] [PubMed]
  85. Raghuveer, K.; Senthilkumaran, B. Cloning and differential expression pattern of vasa in the developing and recrudescing gonads of catfish, Clarias gariepinus. Comp. Biochem. Physiol. A Mol. Integr. Physiol. 2010, 157, 79–85. [Google Scholar] [CrossRef] [PubMed]
  86. Blázquez, M.; González, A.; Mylonas, C.C.; Piferrer, F. Cloning and sequence analysis of a vasa homolog in the European sea bass (Dicentrarchus labrax): Tissue distribution and mRNA expression levels during early development and sex differentiation. Gen. Comp. Endocrinol. 2011, 170, 322–333. [Google Scholar] [CrossRef] [Green Version]
  87. Liu, X.; Zhu, Y.; Zhao, Y.; Wang, Y.; Li, W.; Hong, X.; Yu, L.; Chen, C.; Xu, H.; Zhu, X. Vasa expression is associated with sex differentiation in the Asian yellow pond turtle, Mauremys mutica. J. Exp. Zool. B Mol. Dev. Evol. 2021, 336, 431–442. [Google Scholar] [CrossRef]
  88. Aduma, N.; Izumi, H.; Mizushima, S.; Kuroiwa, A. Knockdown of DEAD-box helicase 4 (DDX4) decreases the number of germ cells in male and female chicken embryonic gonads. Reprod. Fertil. Dev. 2019, 31, 847–854. [Google Scholar] [CrossRef]
  89. Toyooka, Y.; Tsunekawa, N.; Takahashi, Y.; Matsui, Y.; Satoh, M.; Noce, T. Expression and intracellular localization of mouse Vasa-homologue protein during germ cell development. Mech. Dev. 2000, 93, 139–149. [Google Scholar] [CrossRef]
  90. Kuramochi-Miyagawa, S.; Kimura, T.; Ijiri, T.W.; Isobe, T.; Asada, N.; Fujita, Y.; Ikawa, M.; Iwai, N.; Okabe, M.; Deng, W.; et al. Mili, a mammalian member of piwi family gene, is essential for spermatogenesis. Development 2004, 131, 839–849. [Google Scholar] [CrossRef] [Green Version]
  91. Lim, A.K.; Lorthongpanich, C.; Chew, T.G.; Tan, C.W.; Shue, Y.T.; Balu, S.; Gounko, N.; Kuramochi-Miyagawa, S.; Matzuk, M.M.; Chuma, S.; et al. The nuage mediates retrotransposon silencing in mouse primordial ovarian follicles. Development 2013, 140, 3819–3825. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Tilly, J.L. Commuting the death sentence: How oocytes strive to survive. Nat. Rev. Mol. Cell Biol. 2001, 2, 838–848. [Google Scholar] [CrossRef] [PubMed]
  93. Roovers, E.F.; Rosenkranz, D.; Mahdipour, M.; Han, C.T.; He, N.; Chuva de Sousa Lopes, S.M.; van der Westerlaken, L.A.; Zischler, H.; Butter, F.; Roelen, B.A.; et al. Piwi proteins and piRNAs in mammalian oocytes and early embryos. Cell Rep. 2015, 10, 2069–2082. [Google Scholar] [CrossRef] [Green Version]
  94. Williams, Z.; Morozov, P.; Mihailovic, A.; Lin, C.; Puvvula, P.K.; Juranek, S.; Rosenwaks, Z.; Tuschl, T. Discovery and characterization of piRNAs in the human fetal ovary. Cell Rep. 2015, 13, 854–863. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Flemr, M.; Malik, R.; Franke, V.; Nejepinska, J.; Sedlacek, R.; Vlahovicek, K.; Svoboda, P. A retrotransposon-driven Dicer isoform directs endogenous small interfering RNA production in mouse oocytes. Cell 2013, 155, 807–816. [Google Scholar] [CrossRef] [Green Version]
  96. Murchison, E.P.; Stein, P.; Xuan, Z.; Pan, H.; Zhang, M.Q.; Schultz, R.M.; Hannon, G.J. Critical roles for Dicer in the female germline. Genes Dev. 2007, 21, 682–693. [Google Scholar] [CrossRef] [Green Version]
  97. Durdevic, Z.; Pillai, R.S.; Ephrussi, A. Transposon silencing in the Drosophila female germline is essential for genome stability in progeny embryos. Life Sci. Alliance 2018, 1, e201800179. [Google Scholar] [CrossRef] [Green Version]
  98. Durdevic, Z.; Ephrussi, A. Germ cell lineage homeostasis in Drosophila requires the Vasa RNA helicase. Genetics 2019, 213, 911–922. [Google Scholar] [CrossRef] [Green Version]
  99. Styhler, S.; Nakamura, A.; Swan, A.; Suter, B.; Lasko, P. Vasa is required for GURKEN accumulation in the oocyte, and is involved in oocyte differentiation and germline cyst development. Development 1998, 125, 1569–1578. [Google Scholar] [CrossRef]
  100. Johnstone, O.; Lasko, P. Interaction with eIF5B is essential for Vasa function during development. Development 2004, 131, 4167–4178. [Google Scholar] [CrossRef] [Green Version]
  101. Liu, N.; Han, H.; Lasko, P. Vasa promotes Drosophila germline stem cell differentiation by activating mei-P26 translation by directly interacting with a (U)-rich motif in its 3′ UTR. Genes Dev. 2009, 23, 2742–2752. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Carrera, P.; Johnstone, O.; Nakamura, A.; Casanova, J.; Jäckle, H.; Lasko, P. VASA mediates translation through interaction with a Drosophila yIF2 homolog. Mol. Cell 2000, 5, 181–187. [Google Scholar] [CrossRef] [PubMed]
  103. Dehghani, M.; Lasko, P. In vivo mapping of the functional regions of the DEAD-box helicase Vasa. Biol. Open 2015, 4, 450–462. [Google Scholar] [CrossRef] [Green Version]
  104. Li, C.; Vagin, V.V.; Lee, S.; Xu, J.; Ma, S.; Xi, H.; Seitz, H.; Horwich, M.D.; Syrzycka, M.; Honda, B.M.; et al. Collapse of germline piRNAs in the absence of Argonaute3 reveals somatic piRNAs in flies. Cell 2009, 137, 509–521. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Zhang, F.; Wang, J.; Xu, J.; Zhang, Z.; Koppetsch, B.S.; Schultz, N.; Vreven, T.; Meignin, C.; Davis, I.; Zamore, P.D.; et al. UAP56 couples piRNA clusters to the perinuclear transposon silencing machinery. Cell 2012, 151, 871–884. [Google Scholar] [CrossRef] [Green Version]
  106. Brennecke, J.; Aravin, A.A.; Stark, A.; Dus, M.; Kellis, M.; Sachidanandam, R.; Hannon, G.J. Discrete small RNA-generating loci as master regulators of transposon activity in Drosophila. Cell 2007, 128, 1089–1103. [Google Scholar] [CrossRef] [Green Version]
  107. Gunawardane, L.S.; Saito, K.; Nishida, K.M.; Miyoshi, K.; Kawamura, Y.; Nagami, T.; Siomi, H.; Siomi, M.C. A slicer-mediated mechanism for repeat-associated siRNA 5′ end formation in Drosophila. Science 2007, 315, 1587–1590. [Google Scholar] [CrossRef] [Green Version]
  108. Patil, V.S.; Kai, T. Repression of retroelements in Drosophila germline via piRNA pathway by the Tudor domain protein Tejas. Curr. Biol. 2010, 20, 724–730. [Google Scholar] [CrossRef] [Green Version]
  109. Czech, B.; Preall, J.B.; McGinn, J.; Hannon, G.J. A transcriptome-wide RNAi screen in the Drosophila ovary reveals factors of the germline piRNA pathway. Mol. Cell 2013, 50, 749–761. [Google Scholar] [CrossRef] [Green Version]
  110. Aravin, A.A.; Naumova, N.M.; Tulin, A.V.; Vagin, V.V.; Rozovsky, Y.M.; Gvozdev, V.A. Double-stranded RNA-mediated silencing of genomic tandem repeats and transposable elements in the D. melanogaster germline. Curr. Biol. 2001, 11, 1017–1027. [Google Scholar] [CrossRef] [Green Version]
  111. Vagin, V.V.; Sigova, A.; Li, C.; Seitz, H.; Gvozdev, V.; Zamore, P.D. A distinct small RNA pathway silences selfish genetic elements in the germline. Science 2006, 313, 320–324. [Google Scholar] [CrossRef] [PubMed]
  112. Palumbo, G.; Bonaccorsi, S.; Robbins, L.G.; Pimpinelli, S. Genetic analysis of Stellate elements of Drosophila melanogaster. Genetics 1994, 138, 1181–1197. [Google Scholar] [CrossRef] [PubMed]
  113. Bozzetti, M.P.; Massari, S.; Finelli, P.; Meggio, F.; Pinna, L.A.; Boldyreff, B.; Issinger, O.G.; Palumbo, G.; Ciriaco, C.; Bonaccorsi, S.; et al. The Ste locus, a component of the parasitic cry-Ste system of Drosophila melanogaster, encodes a protein that forms crystals in primary spermatocytes and mimics properties of the beta subunit of casein kinase 2. Proc. Natl. Acad. Sci. USA 1995, 92, 6067–6071. [Google Scholar] [CrossRef] [PubMed]
  114. Chen, Y.; Pane, A.; Schüpbach, T. Cutoff and aubergine mutations result in retrotransposon upregulation and checkpoint activation in Drosophila. Curr. Biol. 2007, 17, 637–642. [Google Scholar] [CrossRef] [Green Version]
  115. Klattenhoff, C.; Bratu, D.P.; McGinnis-Schultz, N.; Koppetsch, B.S.; Cook, H.A.; Theurkauf, W.E. Drosophila rasiRNA pathway mutations disrupt embryonic axis specification through activation of an ATR/Chk2 DNA damage response. Dev. Cell 2007, 12, 45–55. [Google Scholar] [CrossRef] [Green Version]
  116. Wang, L.; Dou, K.; Moon, S.; Tan, F.J.; Zhang, Z.Z. Hijacking oogenesis enables massive propagation of LINE and retroviral transposons. Cell 2018, 174, 1082–1094.e12. [Google Scholar] [CrossRef] [Green Version]
  117. Tomancak, P.; Guichet, A.; Zavorszky, P.; Ephrussi, A. Oocyte polarity depends on regulation of gurken by Vasa. Development 1998, 125, 1723–1732. [Google Scholar] [CrossRef]
  118. Gavis, E.R.; Lunsford, L.; Bergsten, S.E.; Lehmann, R. A conserved 90 nucleotide element mediates translational repression of nanos RNA. Development 1996, 122, 2791–2800. [Google Scholar] [CrossRef]
  119. Neumüller, R.A.; Betschinger, J.; Fischer, A.; Bushati, N.; Poernbacher, I.; Mechtler, K.; Cohen, S.M.; Knoblich, J.A. Mei-P26 regulates microRNAs and cell growth in the Drosophila ovarian stem cell lineage. Nature 2008, 454, 241–245. [Google Scholar] [CrossRef] [Green Version]
  120. Li, Y.; Maines, J.Z.; Tastan, O.Y.; McKearin, D.M.; Buszczak, M. Mei-P26 regulates the maintenance of ovarian germline stem cells by promoting BMP signaling. Development 2012, 139, 1547–1556. [Google Scholar] [CrossRef] [Green Version]
  121. Thomson, T.; Liu, N.; Arkov, A.; Lehmann, R.; Lasko, P. Isolation of new polar granule components in Drosophila reveals P body and ER associated proteins. Mech. Dev. 2008, 125, 865–873. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Ma, X.; Zhu, X.; Han, Y.; Story, B.; Do, T.; Song, X.; Wang, S.; Zhang, Y.; Blanchette, M.; Gogol, M.; et al. Aubergine controls germline stem cell self-renewal and progeny differentiation via distinct mechanisms. Dev. Cell. 2017, 41, 157–169.e5. [Google Scholar] [CrossRef] [Green Version]
  123. Ramat, A.; Garcia-Silva, M.R.; Jahan, C.; Naït-Saïdi, R.; Dufourt, J.; Garret, C.; Chartier, A.; Cremaschi, J.; Patel, V.; Decourcelle, M.; et al. The PIWI protein Aubergine recruits eIF3 to activate translation in the germ plasm. Cell Res. 2020, 30, 421–435. [Google Scholar] [CrossRef] [Green Version]
  124. Nakamura, A.; Amikura, R.; Mukai, M.; Kobayashi, S.; Lasko, P.F. Requirement for a noncoding RNA in Drosophila polar granules for germ cell establishment. Science 1996, 274, 2075–2079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Dehghani, M.; Lasko, P. C-terminal residues specific to Vasa among DEAD-box helicases are required for its functions in piRNA biogenesis and embryonic patterning. Dev. Genes Evol. 2016, 226, 401–412. [Google Scholar] [CrossRef] [PubMed]
  126. Lehmann, R.; Nüsslein-Volhard, C. The maternal gene nanos has a central role in posterior pattern formation of the Drosophila embryo. Development 1991, 112, 679–691. [Google Scholar] [CrossRef]
  127. Mahowald, A.P. Assembly of the Drosophila germ plasm. Int. Rev. Cytol. 2001, 203, 187–213. [Google Scholar] [CrossRef]
  128. Santos, A.C.; Lehmann, R. Germ cell specification and migration in Drosophila and beyond. Curr. Biol. 2004, 14, R578–R589. [Google Scholar] [CrossRef] [Green Version]
  129. Rangan, P.; DeGennaro, M.; Jaime-Bustamante, K.; Coux, R.X.; Martinho, R.G.; Lehmann, R. Temporal and spatial control of germ-plasm RNAs. Curr. Biol. 2009, 19, 72–77. [Google Scholar] [CrossRef] [Green Version]
  130. Lehmann, R. Germ plasm biogenesis—An Oskar-centric perspective. Curr. Top. Dev. Biol. 2016, 116, 679–707. [Google Scholar] [CrossRef] [Green Version]
  131. Dehghani, M.; Lasko, P. Multiple functions of the DEAD-box helicase Vasa in Drosophila oogenesis. Results Probl. Cell Differ. 2017, 63, 127–147. [Google Scholar] [CrossRef] [PubMed]
  132. Vagin, V.V.; Klenov, M.S.; Kalmykova, A.I.; Stolyarenko, A.D.; Kotelnikov, R.N.; Gvozdev, V.A. The RNA interference proteins and vasa locus are involved in the silencing of retrotransposons in the female germline of Drosophila melanogaster. RNA Biol. 2004, 1, 54–58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Kotov, A.A.; Adashev, V.E.; Godneeva, B.K.; Ninova, M.; Shatskikh, A.S.; Bazylev, S.S.; Aravin, A.A.; Olenina, L.V. piRNA silencing contributes to interspecies hybrid sterility and reproductive isolation in Drosophila melanogaster. Nucleic Acids Res. 2019, 47, 4255–4271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Nishida, K.M.; Saito, K.; Mori, T.; Kawamura, Y.; Nagami-Okada, T.; Inagaki, S.; Siomi, H.; Siomi, M.C. Gene silencing mechanisms mediated by Aubergine piRNA complexes in Drosophila male gonad. RNA 2007, 13, 1911–1922. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Chen, P.; Kotov, A.A.; Godneeva, B.K.; Bazylev, S.S.; Olenina, L.V.; Aravin, A.A. piRNA-mediated gene regulation and adaptation to sex-specific transposon expression in D. melanogaster male germline. Genes Dev. 2021, 35, 914–935. [Google Scholar] [CrossRef]
  136. Adashev, V.E.; Kotov, A.A.; Bazylev, S.S.; Shatskikh, A.S.; Aravin, A.A.; Olenina, L.V. Stellate genes and the piRNA pathway in speciation and reproductive isolation of Drosophila melanogaster. Front. Genet. 2021, 11, 610665. [Google Scholar] [CrossRef]
  137. Moss, E.G.; Lee, R.C.; Ambros, V. The cold shock domain protein LIN-28 controls developmental timing in C. elegans and is regulated by the lin-4 RNA. Cell 1997, 88, 637–646. [Google Scholar] [CrossRef] [Green Version]
  138. Wilbert, M.L.; Huelga, S.C.; Kapeli, K.; Stark, T.J.; Liang, T.Y.; Chen, S.X.; Yan, B.Y.; Nathanson, J.L.; Hutt, K.R.; Lovci, M.T.; et al. LIN28 binds messenger RNAs at GGAGA motifs and regulates splicing factor abundance. Mol. Cell 2012, 48, 195–206. [Google Scholar] [CrossRef] [Green Version]
  139. Lee, H.-C.; Gu, W.; Shirayama, M.; Youngman, E.; Conte, D.; Mello, C.C. C. elegans piRNAs mediate the genome-wide surveillance of germline transcripts. Cell 2012, 150, 78–87. [Google Scholar] [CrossRef] [Green Version]
  140. Batista, P.J.; Ruby, J.G.; Claycomb, J.M.; Chiang, R.; Fahlgren, N.; Kasschau, K.D.; Chaves, D.A.; Gu, W.; Vasale, J.J.; Duan, S.; et al. PRG-1 and 21U-RNAs interact to form the piRNA complex required for fertility in C. elegans. Mol. Cell 2008, 31, 67–78. [Google Scholar] [CrossRef] [Green Version]
  141. Das, P.P.; Bagijn, M.P.; Goldstein, L.D.; Woolford, J.R.; Lehrbach, N.J.; Sapetschnig, A.; Buhecha, H.R.; Gilchrist, M.J.; Howe, K.L.; Stark, R.; et al. Piwi and piRNAs act upstream of an endogenous siRNA pathway to suppress Tc3 transposon mobility in the Caenorhabditis elegans germline. Mol. Cell 2008, 31, 79–90. [Google Scholar] [CrossRef] [Green Version]
  142. Ruby, J.G.; Jan, C.; Player, C.; Axtell, M.J.; Lee, W.; Nusbaum, C.; Ge, H.; Bartel, D.P. Large-scale sequencing reveals 21U-RNAs and additional microRNAs and endogenous siRNAs in C. elegans. Cell 2006, 127, 1193–1207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Wang, G.; Reinke, V. A C. elegans Piwi, PRG-1, regulates 21U-RNAs during spermatogenesis. Curr. Biol. 2008, 18, 861–867. [Google Scholar] [CrossRef] [Green Version]
  144. Gu, W.; Lee, H.-C.; Chaves, D.; Youngman, E.M.; Pazour, G.J.; Conte, D.; Mello, C.C. CapSeq and CIP-TAP identify Pol II start sites and reveal capped small RNAs as C. elegans piRNA precursors. Cell 2012, 151, 1488–1500. [Google Scholar] [CrossRef] [Green Version]
  145. Cecere, G.; Zheng, G.X.Y.; Mansisidor, A.R.; Klymko, K.E.; Grishok, A. Promoters recognized by Forkhead proteins exist for individual 21URNAs. Mol. Cell 2012, 47, 734–745. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Shen, E.-Z.; Chen, H.; Ozturk, A.R.; Tu, S.; Shirayama, M.; Tang, W.; Ding, Y.-H.; Dai, S.-Y.; Weng, Z.; Mello, C.C. Identification of piRNA binding sites reveals the Argonaute regulatory landscape of the C. elegans germline. Cell 2018, 172, 937–951. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Shirayama, M.; Seth, M.; Lee, H.-C.; Gu, W.; Ishidate, T.; Conte, D.; Mello, C.C. piRNAs initiate an epigenetic memory of nonself RNA in the C. elegans germline. Cell 2012, 150, 65–77. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Bagijn, M.P.; Goldstein, L.D.; Sapetschnig, A.; Weick, E.-M.; Bouasker, S.; Lehrbach, N.J.; Simard, M.J.; Miska, E.A. Function, targets, and evolution of Caenorhabditis elegans piRNAs. Science 2012, 337, 574–578. [Google Scholar] [CrossRef] [Green Version]
  149. Ashe, A.; Sapetschnig, A.; Weick, E.M.; Mitchell, J.; Bagijn, M.P.; Cording, A.C.; Doebley, A.L.; Goldstein, L.D.; Lehrbach, N.J.; Le Pen, J.; et al. piRNAs can trigger a multigenerational epigenetic memory in the germline of C. elegans. Cell 2012, 150, 88–99. [Google Scholar] [CrossRef] [Green Version]
  150. Phillips, C.M.; Montgomery, T.A.; Breen, P.C.; Ruvkun, G. MUT-16 promotes formation of perinuclear mutator foci required for RNA silencing in the C. elegans germline. Genes Dev. 2012, 26, 1433–1444. [Google Scholar] [CrossRef] [Green Version]
  151. Gu, W.; Shirayama, M.; Conte, D., Jr.; Vasale, J.; Batista, P.J.; Claycomb, J.M.; Moresco, J.J.; Youngman, E.M.; Keys, J.; Stoltz, M.J.; et al. Distinct Argonaute-mediated 22G-RNA pathways direct genome surveillance in the C. elegans germline. Mol. Cell 2009, 36, 231–244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Montgomery, B.E.; Vijayasarathy, T.; Marks, T.N.; Cialek, C.A.; Reed, K.J.; Montgomery, T.A. Dual roles for piRNAs in promoting and preventing gene silencing in C. elegans. Cell Rep. 2021, 37, 110101. [Google Scholar] [CrossRef]
  153. Barucci, G.; Cornes, E.; Singh, M.; Li, B.; Ugolini, M.; Samolygo, A.; Didier, C.; Dingli, F.; Loew, D.; Quarato, P.; et al. Small-RNA-mediated transgenerational silencing of histone genes impairs fertility in piRNA mutants. Nat. Cell Biol. 2020, 22, 235–245. [Google Scholar] [CrossRef] [PubMed]
  154. Cecere, G.; Hoersch, S.; O’Keeffe, S.; Sachidanandam, R.; Grishok, A. Global effects of the CSR-1 RNA interference pathway on the transcriptional landscape. Nat. Struct. Mol. Biol. 2014, 21, 358–365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Claycomb, J.M.; Batista, P.J.; Pang, K.M.; Gu, W.; Vasale, J.J.; van Wolfswinkel, J.C.; Chaves, D.A.; Shirayama, M.; Mitani, S.; Ketting, R.F.; et al. The Argonaute CSR-1 and its 22G-RNA cofactors are required for holocentric chromosome segregation. Cell 2009, 139, 123–134. [Google Scholar] [CrossRef] [Green Version]
  156. Gerson-Gurwitz, A.; Wang, S.; Sathe, S.; Green, R.; Yeo, G.W.; Oegema, K.; Desai, A. A Small RNA-catalytic Argonaute pathway tunes germline transcript levels to ensure embryonic divisions. Cell 2016, 165, 396–409. [Google Scholar] [CrossRef] [Green Version]
  157. Cornes, E.; Bourdon, L.; Singh, M.; Mueller, F.; Quarato, P.; Wernersson, E.; Bienko, M.; Li, B.; Cecere, G. piRNAs initiate transcriptional silencing of spermatogenic genes during C. elegans germline development. Dev. Cell 2022, 57, 180–196.e7. [Google Scholar] [CrossRef]
  158. Kuznicki, K.A.; Smith, P.A.; Leung-Chiu, W.M.; Estevez, A.O.; Scott, H.C.; Bennett, K.L. Combinatorial RNA interference indicates GLH-4 can compensate for GLH-1; these two P granule components are critical for fertility in C. elegans. Development 2000, 127, 2907–2916. [Google Scholar] [CrossRef]
  159. Spike, C.; Meyer, N.; Racen, E.; Orsborn, A.; Kirchner, J.; Kuznicki, K.; Yee, C.; Bennett, K.; Strome, S. Genetic analysis of the Caenorhabditis elegans GLH family of P-granule proteins. Genetics 2008, 178, 1973–1987. [Google Scholar] [CrossRef] [Green Version]
  160. Spracklin, G.; Fields, B.; Wan, G.; Becker, D.; Wallig, A.; Shukla, A.; Kennedy, S. The RNAi inheritance machinery of Caenorhabditis elegans. Genetics 2017, 206, 1403–1416. [Google Scholar] [CrossRef] [Green Version]
  161. Seydoux, G. The P granules of C. elegans: A genetic model for the study of RNA–protein condensates. J. Mol. Biol. 2018, 430, 4702–4710. [Google Scholar] [CrossRef] [PubMed]
  162. Chen, W.; Hu, Y.; Lang, C.F.; Brown, J.S.; Schwabach, S.; Song, X.; Zhang, Y.; Munro, E.; Bennett, K.; Zhang, D.; et al. The dynamics of P granule liquid droplets are regulated by the Caenorhabditis elegans germline RNA helicase GLH-1 via its ATP hydrolysis cycle. Genetics 2020, 215, 421–434. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Uebel, C.J.; Agbede, D.; Wallis, D.C.; Phillips, C.M. Mutator foci are regulated by developmental stage, RNA, and the germline cell cycle in Caenorhabditis elegans. G3 2020, 10, 3719–3728. [Google Scholar] [CrossRef]
  164. Dai, S.; Tang, X.; Li, L.; Ishidate, T.; Ozturk, A.R.; Chen, H.; Dude, A.L.; Yan, Y.H.; Dong, M.Q.; Shen, E.Z.; et al. A family of C. elegans VASA homologs control Argonaute pathway specificity and promote transgenerational silencing. Cell Rep. 2022, 40, 111265. [Google Scholar] [CrossRef] [PubMed]
  165. Kiledjian, M.; Dreyfuss, G. Primary structure and binding activity of the hnRNP U protein: Binding RNA through RGG box. EMBO J. 1992, 11, 2655–2664. [Google Scholar] [CrossRef]
  166. Bortvin, A. PIWI-interacting RNAs (piRNAs)—A mouse testis perspective. Biochemistry 2013, 78, 592–602. [Google Scholar] [CrossRef]
  167. Lehtiniemi, T.; Kotaja, N. Germ granule-mediated RNA regulation in male germ cells. Reproduction 2018, 155, R77–R91. [Google Scholar] [CrossRef]
  168. Yokota, S. Historical survey on chromatoid body research. Acta Histochem. Cytochem. 2008, 41, 65–82. [Google Scholar] [CrossRef] [Green Version]
  169. Meikar, O.; Kotaja, N. Isolation of chromatoid bodies from mouse testis as a rich source of short RNAs. Methods Mol. Biol. 2014, 1173, 11–25. [Google Scholar] [CrossRef]
  170. Sassone-Corsi, P. Unique chromatin remodeling and transcriptional regulation in spermatogenesis. Science 2002, 296, 2176–2178. [Google Scholar] [CrossRef]
  171. Shima, J.E.; McLean, D.J.; McCarrey, J.R.; Griswold, M.D. The murine testicular transcriptome: Characterizing gene expression in the testis during the progression of spermatogenesis. Biol. Reprod. 2004, 71, 319–330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Vara, C.; Paytuví-Gallart, A.; Cuartero, Y.; Le Dily, F.; Garcia, F.; Salvà-Castro, J.; Gómez, H.L.; Julià, E.; Moutinho, C.; Aiese Cigliano, R.; et al. Three-dimensional genomic structure and cohesin occupancy correlate with transcriptional activity during spermatogenesis. Cell Rep. 2019, 28, 352–367.e9. [Google Scholar] [CrossRef]
  173. Kuramochi-Miyagawa, S.; Watanabe, T.; Gotoh, K.; Takamatsu, K.; Chuma, S.; Kojima-Kita, K.; Shiromoto, Y.; Asada, N.; Toyoda, A.; Fujiyama, A.; et al. MVH in piRNA processing and gene silencing of retrotransposons. Genes Dev. 2010, 24, 887–892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Wenda, J.M.; Homolka, D.; Yang, Z.; Spinelli, P.; Sachidanandam, R.; Pandey, R.R.; Pillai, R.S. Distinct Roles of RNA Helicases MVH and TDRD9 in PIWI slicing-triggered mammalian piRNA biogenesis and function. Dev. Cell 2017, 41, 623–637.e9. [Google Scholar] [CrossRef] [Green Version]
  175. Sun, Y.H.; Lee, B.; Li, X.Z. The birth of piRNAs: How mammalian piRNAs are produced, originated, and evolved. Mamm. Genome 2022, 33, 293–311. [Google Scholar] [CrossRef]
  176. Goh, W.S.; Falciatori, I.; Tam, O.H.; Burgess, R.; Meikar, O.; Kotaja, N.; Hammell, M.; Hannon, G.J. piRNA-directed cleavage of meiotic transcripts regulates spermatogenesis. Genes Dev. 2015, 29, 1032–1044. [Google Scholar] [CrossRef] [Green Version]
  177. Choi, H.; Wang, Z.; Dean, J. Sperm acrosome overgrowth and infertility in mice lacking chromosome 18 pachytene piRNA. PLoS Genet. 2021, 17, e1009485. [Google Scholar] [CrossRef]
  178. Wu, P.H.; Fu, Y.; Cecchini, K.; Özata, D.M.; Arif, A.; Yu, T.; Colpan, C.; Gainetdinov, I.; Weng, Z.; Zamore, P.D. The evolutionarily conserved piRNA-producing locus pi6 is required for male mouse fertility. Nat. Genet. 2020, 52, 728–739. [Google Scholar] [CrossRef]
  179. Nagamori, I.; Cruickshank, V.A.; Sassone-Corsi, P. Regulation of an RNA granule during spermatogenesis: Acetylation of MVH in the chromatoid body of germ cells. J. Cell Sci. 2011, 124, 4346–4355. [Google Scholar] [CrossRef] [Green Version]
  180. Su, R.; Fan, L.H.; Cao, C.; Wang, L.; Du, Z.; Cai, Z.; Ouyang, Y.C.; Wang, Y.; Zhou, Q.; Wu, L.; et al. Global profiling of RNA-binding protein target sites by LACE-seq. Nat. Cell Biol. 2021, 23, 664–675. [Google Scholar] [CrossRef] [PubMed]
  181. Bendsen, E.; Byskov, A.G.; Andersen, C.Y.; Westergaard, L.G. Number of germ cells and somatic cells in human fetal ovaries during the first weeks after sex differentiation. Hum. Reprod. 2006, 21, 30–35. [Google Scholar] [CrossRef] [Green Version]
  182. Anderson, R.A.; Fulton, N.; Cowan, G.; Coutts, S.; Saunders, P.T. Conserved and divergent patterns of expression of DAZL, VASA and OCT4 in the germ cells of the human fetal ovary and testis. BMC Dev. Biol. 2007, 7, 136. [Google Scholar] [CrossRef] [Green Version]
  183. Martin, J.J.; Woods, D.C.; Tilly, J.L. Implications and current limitations of oogenesis from female germline or oogonial stem cells in adult mammalian ovaries. Cells 2019, 8, 93. [Google Scholar] [CrossRef] [Green Version]
  184. Stoop, H.; Honecker, F.; Cools, M.; de Krijger, R.; Bokemeyer, C.; Looijenga, L.H. Differentiation and development of human female germ cells during prenatal gonadogenesis: An immunohistochemical study. Hum. Reprod. 2005, 20, 1466–1476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Balbiani, E.G. Sur la constitution du germne dans l’oeuf animal avant la fcondation. Compt. Rend. 1864, 58, 584–588. [Google Scholar]
  186. Hertig, A.T.; Adams, E.C. Studies on the human oocyte and its follicle. I. Ultrastructural and histochemical observations on the primordial follicle stage. J. Cell Biol. 1967, 34, 647–675. [Google Scholar] [CrossRef]
  187. Albamonte, M.S.; Willis, M.A.; Albamonte, M.I.; Jensen, F.; Espinosa, M.B.; Vitullo, A.D. The developing human ovary: Immunohistochemical analysis of germ-cell-specific VASA protein, BCL-2/BAX expression balance and apoptosis. Hum. Reprod. 2008, 23, 1895–1901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Albamonte, M.I.; Albamonte, M.S.; Stella, I.; Zuccardi, L.; Vitullo, A.D. The infant and pubertal human ovary: Balbiani’s body-associated VASA expression, immunohistochemical detection of apoptosis-related BCL2 and BAX proteins, and DNA fragmentation. Hum. Reprod. 2013, 28, 698–706. [Google Scholar] [CrossRef] [Green Version]
  189. Guo, X.; Gui, Y.T.; Tang, A.F.; Lu, L.H.; Gao, X.; Cai, Z.M. Differential expression of VASA gene in ejaculated spermatozoa from normozoospermic men and patients with oligozoospermia. Asian J. Androl. 2007, 9, 339–344. [Google Scholar] [CrossRef] [PubMed]
  190. Amirian, M.; Azizi, H.; Hashemi Karoii, D.; Skutella, T. VASA protein and gene expression analysis of human non-obstructive azoospermia and normal by immunohistochemistry, immunocytochemistry, and bioinformatics analysis. Sci. Rep. 2022, 12, 17259. [Google Scholar] [CrossRef]
  191. Zeeman, A.M.; Stoop, H.; Boter, M.; Gillis, A.J.; Castrillon, D.H.; Oosterhuis, J.W.; Looijenga, L.H. VASA is a specific marker for both normal and malignant human germ cells. Lab. Investig. 2002, 82, 159–166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Sugimoto, K.; Koh, E.; Sin, H.S.; Maeda, Y.; Narimoto, K.; Izumi, K.; Kobori, Y.; Kitamura, E.; Nagase, H.; Yoshida, A.; et al. Tissue-specific differentially methylated regions of the human VASA gene are potentially associated with maturation arrest phenotype in the testis. J. Hum. Genet. 2009, 54, 450–456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Hermann, B.P.; Sukhwani, M.; Lin, C.C.; Sheng, Y.; Tomko, J.; Rodriguez, M.; Shuttleworth, J.J.; McFarland, D.; Hobbs, R.M.; Pandolfi, P.P.; et al. Characterization, cryopreservation, and ablation of spermatogonial stem cells in adult rhesus macaques. Stem Cells 2007, 25, 2330–2338. [Google Scholar] [CrossRef] [Green Version]
  194. Fayomi, A.P.; Orwig, K.E. Spermatogonial stem cells and spermatogenesis in mice, monkeys and men. Stem Cell Res. 2018, 29, 207–214. [Google Scholar] [CrossRef] [PubMed]
  195. Rojas-Ríos, P.; Chartier, A.; Pierson, S.; Simonelig, M. Aubergine and piRNAs promote germline stem cell self-renewal by repressing the proto-oncogene Cbl. EMBO J. 2017, 36, 3194–3211. [Google Scholar] [CrossRef]
  196. Xu, C.; Cao, Y.; Bao, J. Building RNA-protein germ granules: Insights from the multifaceted functions of DEAD-box helicase Vasa/Ddx4 in germline development. Cell. Mol. Life Sci. 2021, 79, 4. [Google Scholar] [CrossRef]
Figure 1. General scheme of DEAD box helicase structure. The following regions are indicated: the N-terminal variable domain; the DEAD-like helicases superfamily (DEXDc) domain; the helicase superfamily C-terminal (HELICc) domain; the C-terminal nonconservative domain. Color bars in the bottom indicate conservative motifs involved in ATP-binding and hydrolysis (red), RNA binding (blue), and coordination between ATP and RNA binding (green). The scheme is modified from [10,11].
Figure 1. General scheme of DEAD box helicase structure. The following regions are indicated: the N-terminal variable domain; the DEAD-like helicases superfamily (DEXDc) domain; the helicase superfamily C-terminal (HELICc) domain; the C-terminal nonconservative domain. Color bars in the bottom indicate conservative motifs involved in ATP-binding and hydrolysis (red), RNA binding (blue), and coordination between ATP and RNA binding (green). The scheme is modified from [10,11].
Cimb 45 00358 g001
Figure 2. Vasa-positive germ granules in Drosophila gonads. (a) The scheme of Drosophila oogenesis and germ granule distribution. Each ovary contains 12–16 ovarioles. An individual ovariole is shown at the top. At the anterior end of the ovariole, the germarium structure maintains two or three germline stem cells (GSCs) (green ovals). The division of GSC produces a new GSC and a cystoblast. The germ cell cysts arise from cystoblasts that undergo four synchronous divisions. Each ovariole consists of developing egg chambers which comprise an oocyte (light brown) and 15 polyploid nurse cells, connected to each other and surrounded by a monolayer of somatic follicle cells (gray). The nuclei of nurse cells are surrounded by an irregular network of nuage germ granules (yellow dots around nuclei). At the 10/11 stages of oogenesis, pole plasm is assembled and pole granules (yellow dots) are concentrated at the posterior pole of the developing oocyte. RNA helicase Vasa is detected throughout oogenesis flies in high concentrations in the perinuclear nuage granules of nurse cells, as well as at the posterior end of the developing oocyte in pole granules. At the bottom: pole granule organization. Pole granules assemble through phase separation of core proteins and associated mRNAs at the posterior pole of the oocyte. Within pole granules, short isoforms of Oskar, Vasa, Tudor, and Aubergine (Aub) proteins are enriched and provide recruiting mRNAs in piRNA-mediated and piRNA-independent modes. Tudor consists of 11 Tudor domains, which are able to bind symmetrically methylated arginine residues found in Aub and Vasa (orange circles). (b) Left: general scheme of spermatogenesis in Drosophila. Hub cells at the apical tip of the testis function as a niche to support two populations of stem cells, germline stem cells (GSCs) and somatic cyst stem cells, through the signal molecule secretion. GSCs divide with production of a self-renewal GSC and goniablast; the latter is surrounded by two somatic cyst cells and additionally undergoes four mitotic divisions with incomplete cytokinesis. After that, the cysts of 16 germ cells switch to the spermatocyte program. Spermatocytes enter meiosis, forming 64 haploid round spermatids. At the end, the elongated spermatids move to the basal end of the testis. Mature individual spermatozoa enter the seminal vesicle. Right top: Spermatocyte (green oval) contains multiple small nuage granules (yellow dots) forming an irregular network round the nucleus (light gray) and a large nuage granule, the piNG-body, one per cell. Right bottom: Lobular structure of the piNG-body and mutual arrangement of Vasa, Aub, Tudor, and AGO3 within the piNG-body are shown. Vasa, Aub, and Tudor occupy the peripheral lobes of the piNG-body, whereas AGO3 protein is located in the piNG-body core. (c) Small nuage granules around the nuclear surface of spermatocyte. A confocal slice presenting the external surface of spermatocyte nucleus is shown. Significant colocalization of Vasa (green) and Aub (red) signals in the nuage granules producing yellow color is observed. (d) AGO3 (red) is located in the central lobe of the piNG-body and colocalizes with Vasa (green) only along the boundary line regions. The 3D immunofluorescent image of spermatocyte nucleus with the piNG-body on its surface is shown. (c,d) The images are reproduced from [26] by permission of MBoC under Creative Commons-Noncommercial-Share Alike 4.0 Unported license (https://creativecommons.org/licenses/by-nc/4.0/).
Figure 2. Vasa-positive germ granules in Drosophila gonads. (a) The scheme of Drosophila oogenesis and germ granule distribution. Each ovary contains 12–16 ovarioles. An individual ovariole is shown at the top. At the anterior end of the ovariole, the germarium structure maintains two or three germline stem cells (GSCs) (green ovals). The division of GSC produces a new GSC and a cystoblast. The germ cell cysts arise from cystoblasts that undergo four synchronous divisions. Each ovariole consists of developing egg chambers which comprise an oocyte (light brown) and 15 polyploid nurse cells, connected to each other and surrounded by a monolayer of somatic follicle cells (gray). The nuclei of nurse cells are surrounded by an irregular network of nuage germ granules (yellow dots around nuclei). At the 10/11 stages of oogenesis, pole plasm is assembled and pole granules (yellow dots) are concentrated at the posterior pole of the developing oocyte. RNA helicase Vasa is detected throughout oogenesis flies in high concentrations in the perinuclear nuage granules of nurse cells, as well as at the posterior end of the developing oocyte in pole granules. At the bottom: pole granule organization. Pole granules assemble through phase separation of core proteins and associated mRNAs at the posterior pole of the oocyte. Within pole granules, short isoforms of Oskar, Vasa, Tudor, and Aubergine (Aub) proteins are enriched and provide recruiting mRNAs in piRNA-mediated and piRNA-independent modes. Tudor consists of 11 Tudor domains, which are able to bind symmetrically methylated arginine residues found in Aub and Vasa (orange circles). (b) Left: general scheme of spermatogenesis in Drosophila. Hub cells at the apical tip of the testis function as a niche to support two populations of stem cells, germline stem cells (GSCs) and somatic cyst stem cells, through the signal molecule secretion. GSCs divide with production of a self-renewal GSC and goniablast; the latter is surrounded by two somatic cyst cells and additionally undergoes four mitotic divisions with incomplete cytokinesis. After that, the cysts of 16 germ cells switch to the spermatocyte program. Spermatocytes enter meiosis, forming 64 haploid round spermatids. At the end, the elongated spermatids move to the basal end of the testis. Mature individual spermatozoa enter the seminal vesicle. Right top: Spermatocyte (green oval) contains multiple small nuage granules (yellow dots) forming an irregular network round the nucleus (light gray) and a large nuage granule, the piNG-body, one per cell. Right bottom: Lobular structure of the piNG-body and mutual arrangement of Vasa, Aub, Tudor, and AGO3 within the piNG-body are shown. Vasa, Aub, and Tudor occupy the peripheral lobes of the piNG-body, whereas AGO3 protein is located in the piNG-body core. (c) Small nuage granules around the nuclear surface of spermatocyte. A confocal slice presenting the external surface of spermatocyte nucleus is shown. Significant colocalization of Vasa (green) and Aub (red) signals in the nuage granules producing yellow color is observed. (d) AGO3 (red) is located in the central lobe of the piNG-body and colocalizes with Vasa (green) only along the boundary line regions. The 3D immunofluorescent image of spermatocyte nucleus with the piNG-body on its surface is shown. (c,d) The images are reproduced from [26] by permission of MBoC under Creative Commons-Noncommercial-Share Alike 4.0 Unported license (https://creativecommons.org/licenses/by-nc/4.0/).
Cimb 45 00358 g002
Figure 3. Separation and compartmentalization of germ granules in the C. elegans germline. P granules scan nascent transcripts exiting from the nucleus via nuclear pores. P granules form with the aid of GLH helicases (homologues of Vasa) as liquid biocondensates containing CSR-1, PRG-1, and WAGO-1 ARGONAUTE proteins and their associated small RNAs, which can sort foreign transcripts from ones licensed for expression. Z granules provide the amplification of 22G-RNAs for memory storage and transgenerational silencing. Mutator foci are sites of WAGO-class 22G-RNA amplification. SIMR foci direct specific piRNA targets for 22G-RNA amplification. Two or more unique protein components for each type of granule are presented. The presumed flow of RNAs and proteins across the granules is indicated by white arrows. The scheme is created using data extracted from a number of papers [51,52,53,54,55,56].
Figure 3. Separation and compartmentalization of germ granules in the C. elegans germline. P granules scan nascent transcripts exiting from the nucleus via nuclear pores. P granules form with the aid of GLH helicases (homologues of Vasa) as liquid biocondensates containing CSR-1, PRG-1, and WAGO-1 ARGONAUTE proteins and their associated small RNAs, which can sort foreign transcripts from ones licensed for expression. Z granules provide the amplification of 22G-RNAs for memory storage and transgenerational silencing. Mutator foci are sites of WAGO-class 22G-RNA amplification. SIMR foci direct specific piRNA targets for 22G-RNA amplification. Two or more unique protein components for each type of granule are presented. The presumed flow of RNAs and proteins across the granules is indicated by white arrows. The scheme is created using data extracted from a number of papers [51,52,53,54,55,56].
Cimb 45 00358 g003
Figure 4. Dynamics of MVH-positive germ granules in mouse spermatogenesis. At the top: two types of germ granules, intermitochondrial cement (IMC) and piP-bodies, are generated in mouse embryonic gonocytes. Both types of granules contain MVH, a murine homologue of Vasa, and distinct components of piRNA machinery. After meiosis in the cytoplasm of round spermatids in adult mice, MVH is concentrated in a single, large, perinuclear, electron-dense granule known as the chromatoid body (CB). Graphs in the middle of the scheme characterize molecular processes in germ cells occurring during mouse spermatogenesis. At the bottom: expression patterns of key proteins of the piRNA pathway, MILI, MIWI2, and MIWI, are shown. The scheme is created using data extracted from a number of papers [74,75,76,78,79,80].
Figure 4. Dynamics of MVH-positive germ granules in mouse spermatogenesis. At the top: two types of germ granules, intermitochondrial cement (IMC) and piP-bodies, are generated in mouse embryonic gonocytes. Both types of granules contain MVH, a murine homologue of Vasa, and distinct components of piRNA machinery. After meiosis in the cytoplasm of round spermatids in adult mice, MVH is concentrated in a single, large, perinuclear, electron-dense granule known as the chromatoid body (CB). Graphs in the middle of the scheme characterize molecular processes in germ cells occurring during mouse spermatogenesis. At the bottom: expression patterns of key proteins of the piRNA pathway, MILI, MIWI2, and MIWI, are shown. The scheme is created using data extracted from a number of papers [74,75,76,78,79,80].
Cimb 45 00358 g004
Figure 5. piRNA-dependent misregulation of vasa and Stellate genes in the testes of interspecies Drosophila hybrids. (a) piRNA-dependent silencing of the D. mauritiana vasa allele in the testes of interspecies hybrids. Silencing of the alien allele of vasa by AT-chX piRNAs in diploid hybrids appears to cause an insufficiency of Vasa functions in germline maintenance in the early stages of spermatogenesis. (b) Derepession of Stellate genes in the testes of hybrids lacking the Y-linked Su(Ste) locus in the genome causes meiotic failures and male hybrid sterility. Chromosomes of hybrids inherited from D. melanogaster are shown in gray, and those from D. mauritiana are shown in beige. The schemes are adapted from [133] by permission of Oxford University Press (https://creativecommons.org/licenses/by/4.0/, accessed on 21 February 2019).
Figure 5. piRNA-dependent misregulation of vasa and Stellate genes in the testes of interspecies Drosophila hybrids. (a) piRNA-dependent silencing of the D. mauritiana vasa allele in the testes of interspecies hybrids. Silencing of the alien allele of vasa by AT-chX piRNAs in diploid hybrids appears to cause an insufficiency of Vasa functions in germline maintenance in the early stages of spermatogenesis. (b) Derepession of Stellate genes in the testes of hybrids lacking the Y-linked Su(Ste) locus in the genome causes meiotic failures and male hybrid sterility. Chromosomes of hybrids inherited from D. melanogaster are shown in gray, and those from D. mauritiana are shown in beige. The schemes are adapted from [133] by permission of Oxford University Press (https://creativecommons.org/licenses/by/4.0/, accessed on 21 February 2019).
Cimb 45 00358 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Adashev, V.E.; Kotov, A.A.; Olenina, L.V. RNA Helicase Vasa as a Multifunctional Conservative Regulator of Gametogenesis in Eukaryotes. Curr. Issues Mol. Biol. 2023, 45, 5677-5705. https://doi.org/10.3390/cimb45070358

AMA Style

Adashev VE, Kotov AA, Olenina LV. RNA Helicase Vasa as a Multifunctional Conservative Regulator of Gametogenesis in Eukaryotes. Current Issues in Molecular Biology. 2023; 45(7):5677-5705. https://doi.org/10.3390/cimb45070358

Chicago/Turabian Style

Adashev, Vladimir E., Alexei A. Kotov, and Ludmila V. Olenina. 2023. "RNA Helicase Vasa as a Multifunctional Conservative Regulator of Gametogenesis in Eukaryotes" Current Issues in Molecular Biology 45, no. 7: 5677-5705. https://doi.org/10.3390/cimb45070358

Article Metrics

Back to TopTop