The Interactive Fly
Genes involved in tissue and organ development
INDEX
Stem cells are found in specialized microenvironments, or 'niches', which regulate stem cell identity and behavior. The adult testis and ovary in Drosophila contain germline stem cells (GSCs) with well-defined niches, and are excellent models for studying niche development. This study investigates the formation of the testis GSC niche, or 'hub', during the late stages of embryogenesis. By morphological and molecular criteria, the development of an embryonic hub that forms from a subset of anterior somatic gonadal precursors (SGPs) were identified and followed in the male gonad. Embryonic hub cells form a discrete cluster apart from other SGPs, express several molecular markers in common with the adult hub and organize anterior-most germ cells in a rosette pattern characteristic of GSCs in the adult. The sex determination genes transformer and doublesex ensure that hub formation occurs only in males. Interestingly, hub formation occurs in both XX and XY gonads mutant for doublesex, indicating that doublesex is required to repress hub formation in females. This work establishes the Drosophila male GSC niche as a model for understanding the mechanisms controlling niche formation and initial stem cell recruitment, as well as the development of sexual dimorphism in the gonad (Le Bras, 2006).
The evidence indicates that an embryonic hub, which appears to give rise to the adult hub and create the male GSC niche, forms during the late stages of embryogenesis. A subset of anterior SGPs initiates expression of several molecular markers that are also expressed in the adult hub. These SGPs segregate into a tight cluster in a distinct region of the gonad, and a subset of germ cells organizes around these SGPs in a manner similar to the organization of GSCs around the adult hub. Since spermatogenesis begins by early larval stages, it is possible that the embryonic hub already forms a functional GSC niche. The formation of the hub, or indeed any stem cell niche, can be divided into the distinct issues of niche cell identity, niche morphogenesis, and stem cell recruitment (Le Bras, 2006).
The data indicate that the specification of hub cell identity occurs in two stages. During the first stage, some SGPs acquire an anterior identity that is sexually dimorphic, as indicated by the male-specific expression of esg and upd. Anterior SGP identity is positively regulated by abd-A, and is repressed by Abd-B, while sexual identity is regulated by tra and dsx. During the second stage of hub cell specification, a subset of these anterior SGPs acquires hub cell identity during stage 17 of embryogenesis. Only some anterior SGPs maintain esg expression, and the control of late gene expression in the hub appears to be distinct from early expression in anterior SGPs, since some esg and upd enhancer traps only exhibit gonad expression in the hub at this later stage. Furthermore, cells that maintain esg expression during stage 17 also express every other marker of adult hub identity tested, including Fasciclin 3, cdi, DN-cadherin and DE-cadherin. It is concluded that these cells are specified as hub cells at this time. The fate of the anterior SGPs that lose esg expression and do not form part of the hub is unknown. An intriguing possibility is that these cells could form another important somatic cell type: the cyst progenitor cells (somatic stem cells) that associate with the hub along with the GSCs (Le Bras, 2006).
Based on its expression pattern, the transcription factor esg would seem to be an excellent candidate for specifying hub cell identity. However, no changes were observed in the expression of other hub markers in esg null mutants; this includes expression of DE-cadherin, which is known to be regulated by esg in other tissues. It has been reported, however, that esg is required for hub maintenance, and that the hub is severely defective at later stages in esg mutants that survive embryogenesis. Thus, esg is critical for the male GSC niche, but is either not important for the initial formation of this structure, or acts redundantly with another factor (Le Bras, 2006).
It has been possible to follow the morphogenesis of the hub from the time of gonad formation until the embryonic hub is fully formed. At the time of gonad coalescence, anterior SGPs interact with other SGPs, and with the germ cells, in a manner that is indistinguishable from posterior SGPs. However, during stage 17, the hub cells undergo dramatic changes in their relationship to other SGPs and germ cells. Hub cells segregate away from other SGPs to one pole of the gonad, and coalesce tightly with one another. In addition, hub cells do not ensheath the germ cells at this stage. Instead, a defined interface between hub cells and germ cells forms which is labeled by DE- and DN-cadherin, but not Fasciclin 3. Thus, hub cells appear to maximize their interactions with one another, and minimize their interactions with other cells in the gonad, although they clearly still contact a subset of germ cells (Le Bras, 2006).
It is apparent that the changes in cell–cell contact and morphology that occur during hub formation require changes in cell adhesion. Indeed, characteristic changes have been found in expression of the homophilic adhesion molecules Fasciclin 3, DN-cadherin and DE-cadherin occur during hub formation; all three are significantly upregulated in the embryonic and adult hub. Increased homophilic adhesion among hub cells could account for their ability to maximize their contacts with one another, and sort away from other SGPs. However, no changes were observed in embryonic hub formation in mutants for these cell adhesion molecules. Thus, these proteins, and possibly others, may act redundantly in this process (Le Bras, 2006).
It is clear that a subset of germ cells organizes specifically with the developing hub as it forms. During the last stage of hub formation, germ cells become oriented in a rosette distribution around the developing hub in a manner characteristic of GSCs in the adult. These may represent the subset of germ cells that will become GSCs. The presence of DE- and DN-cadherin at sites of hub–germ cell contact suggests that cadherin-mediated adhesion may be important for niche–GSC interaction in the testis, as has been observed in the ovary. Interestingly, germ cells are not required for hub formation. Analysis of a number of hub identity markers indicates that these cell form normally from a subset of anterior SGPs in embryos that lack germ cells. The hub does not appear as well compacted in these embryos, consistent with observations of the adult hub, indicating that hub–germ cell contact (or hub–germ cell signaling) affects the final shape of the hub. Nevertheless, the GSC niche can form in the absence of one of its stem cell populations (somatic stem cells may still be present). It will be of great interest in the future to determine if the subset of germ cells organized around the male embryonic hub are, indeed, developing GSCs, and to study how their transition to stem cell identity might be regulated by the niche (Le Bras, 2006).
The formation of the male GSC niche is a sex-specific characteristic of anterior SGPs. Male-specific expression of esg and hub formation both require the sex determination genes tra and dsx. In some tissues, DSXM is required to promote male development and repress female development, while the opposite is true for DSXF. Interestingly, it was found that embryonic hub development is entirely masculinized in dsx null mutants; XX and XY individuals appear identical when mutant for dsx and both resemble wild type males. Thus, no role is seen for DSXM in promoting embryonic hub formation, while DSXF is required in females to repress hub formation. Since esg is expressed male-specifically, it is one candidate for being directly regulated by DSX (Le Bras, 2006).
We can compare the development of the anterior SGPs and hub with the development of another sexually dimorphic cell type, the msSGPs that join the posterior of the male gonad. First of all, these two cell types are distinct and do not depend on one another for their proper development. The hub still forms in Abd-B mutants that lack msSGPs, while msSGPs are still found in the gonad in Pc mutants, in which no anterior SGPs or hub cells form. Second, the mechanism for how sexual dimorphism is created differs between the two cell types. msSGPs are present only in males because they have undergone sex-specific apoptosis in females. In contrast, no apoptosis was observed in anterior SGPs. These cells appear to remain present in both sexes, but only form a hub in males. Thus, although the sex determination genes tra and dsx regulate sex-specific development of both cell types, the cellular mechanisms employed are different. Finally, as was observed for the hub, development of the msSGPs is completely masculinized in dsx mutant embryos. Thus, for both of these cell types, the male pattern of development in the embryonic gonad is the default state in the absence of dsx function, and it is the role of DSXF to repress male development in females. However, DSXM may well play a role in development of one or both of these gonad cell types at later stages, since proper testis development in males clearly requires dsx (Le Bras, 2006).
The sex determination pathway must also ensure that GSC niches form in females and are different from those in males. Recently, it has been shown that germ cells populating the anterior of the gonad in female embryos are predisposed to become GSCs in the adult ovary, while germ cells populating the posterior rarely become GSCs. This suggests that anterior SGPs in the female embryonic gonad may promote GSC identity, similar to what is proposed to happen in the male during hub formation. One possibility is that anterior SGPs give rise to GSC niches in both sexes, while genes such as tra and dsx control whether these niches will be male or female (Le Bras, 2006).
In conclusion, the development has been followed of the embryonic hub, which may represent the nascent GSC niche for the testis. This work provides a basis for further understanding the mechanisms controlling niche formation and GSC recruitment in Drosophila, and determining if these mechanisms are conserved in other
stem cell systems, including the GSC niche of the mammalian testis (Le Bras, 2006).
In the Drosophila testis, germline stem cells (GSCs) and somatic cyst stem cells (CySCs) are arranged around a group of postmitotic somatic cells, termed the hub, which produce a variety of growth factors contributing to the niche microenvironment that regulates both stem cell pools. This study shows that CySC but not GSC maintenance requires Hedgehog (Hh) signalling in addition to Jak/Stat pathway activation. CySC clones unable to transduce the Hh signal are lost by differentiation, whereas pathway overactivation leads to an increase in proliferation. However, unlike cells ectopically overexpressing Jak/Stat targets, the additional cells generated by excessive Hh signalling remain confined to the testis tip and retain the ability to differentiate. Interestingly, Hh signalling also controls somatic cell populations in the fly ovary and the mammalian testis. These observations might therefore point towards a higher degree of organisational homology between the somatic components of gonads across the sexes and phyla than previously appreciated (Michel, 2012).
Hh thus provides a niche signal for the maintenance and proliferation of the somatic stem cells of the testis. CySCs that are unable to transduce the Hh signal are lost through differentiation, whereas pathway overactivation causes overproliferation. Hh signalling thereby resembles Jak/Stat signalling via Upd. Partial redundancy between these pathways might explain why neither depletion of Stat activity nor loss of Hh signalling causes complete CySC loss (Michel, 2012).
This study has shown that loss of Hh signalling in smo mutant cells blocks expression of the Jak/Stat target Zfh1, whereas mutation of ptc expands the Zfh1-positive pool. Overexpression of Zfh1 or another Jak/Stat target, Chinmo, is sufficient to induce CySC-like behaviour in somatic cells irrespective of their distance. By contrast, Hh overexpression in the hub using the hh::Gal4 driver only caused a moderate increase in the number of Zfh1-positive cells relative to a GFP control. Ectopic Hh overexpression in somatic cells under c587::Gal4 control increased this number further. However, unlike in somatic cells with constitutively active Jak/Stat signalling, the additional Zfh1-positive cells remained largely confined to the testis tip, although their average range was increased threefold. Thus, Hh appears to promote stem cell proliferation, in part, also independently of competition (Michel, 2012).
It is tempting to speculate that further stem cell expansion is limited by Upd range. Consistently, cells with an ectopically activated Jak/Stat pathway remain undifferentiated, whereas ptc cells can still differentiate. Future experiments will need to formally address the epistasis between these pathways. However, the observations already show that Hh signalling influences expression of the bona fide Upd target gene zfh1, and therefore presumably acts upstream, or in parallel to, Upd in maintaining CySC fate (Michel, 2012).
In addition, the reduction in GSC number following somatic stem cell loss implies cross-regulation between the different stem cell populations that presumably involves additional signalling cascades, such as the EGF pathway (Michel, 2012).
In recent years, research has focused on the differences between the male and female gonadal niches. This paper instead emphasizes the similarities: in both cases, Jak/Stat signalling is responsible for the maintenance and activity of cells that contribute to the GSC niche, and Hh signalling promotes the proliferation of stem cells that provide somatic cells ensheathing germline cysts. In the testis, both functions are fulfilled by the CySCs, whereas in the ovary the former task is fulfilled by the postmitotic escort stem cells/escort cells and the latter by the FSCs. Finally, male desert hedgehog (Dhh) knockout mice are sterile. Dhh is expressed in the Sertoli cells and is thought to primarily act on the somatic Leydig cells. However, the signalling microenvironment of the vertebrate spermatogonial niche is, as yet, not fully defined. Future experiments will need to clarify whether these similarities reflect convergence or an ancestral Hh function in the metazoan gonad (Michel, 2012).
The Drosophila testis harbors two types of stem cells: germ line stem cells (GSCs) and cyst stem cells (CySCs). Both stem cell types share a physical niche called the hub, located at the apical tip of the testis. The niche produces the JAK/STAT ligand Unpaired (Upd) and BMPs to maintain CySCs and GSCs, respectively. However, GSCs also require BMPs produced by CySCs, and as such CySCs are part of the niche for GSCs. This study describes a role for another secreted ligand, Hedgehog (Hh), produced by niche cells, in the self-renewal of CySCs. Hh signaling cell-autonomously regulates CySC number and maintenance. The Hh and JAK/STAT pathways act independently and non-redundantly in CySC self-renewal. Finally, Hh signaling does not contribute to the niche function of CySCs, as Hh-sustained CySCs are unable to maintain GSCs in the absence of Stat92E. Therefore, the extended niche function of CySCs is solely attributable to JAK/STAT pathway function (Amoyel, 2013).
This study has shown that Hh from the Drosophila testis niche is a self-renewal
factor for CySCs and that Hh signaling does not contribute
to the role of CySCs as a niche for GSCs. This supports the model
that the Hh and JAK/STAT pathways act independently within
CySCs. The results therefore confirm those recently
reported by another group (Michel, 2012), who showed that
Hh regulates CySC self-renewal, and extend their results by
demonstrating the genetic independence of Hh and the other
pathway (i.e. JAK/STAT) that is crucial in CySC function (Amoyel, 2013).
It is notable that two signals regulate CySC self-renewal but only
JAK/STAT signaling contributes to the GSC niche. Moreover,
despite the drastic reduction in CySCs in hhts2 testes (from ~36 in
controls to ~8), GSCs do remain in hh mutant animals albeit at
reduced numbers. The reduction in GSCs in hh mutants is not due
to changes in the size of the hub. These data suggest that most
CySCs are dispensable for their niche function and that only a few
BMP-producing CySCs are needed to maintain GSC self-renewal.
This raises the question as to whether, in a wild-type animal, there
are distinct populations of CySCs, some with activated Stat92E that
produce BMPs and act as a niche for GSCs, and others with
activated Hh signaling that participate only in self-renewal and the
production of cyst progeny. This is consistent with the fact that,
despite the presence of ~36 Zfh1-positive CySCs, elevated Stat92E
is only seen in a few CySCs. However, it is also
conceivable that all Zfh1-positive CySCs are equivalent and that
high Stat92E correlates, for instance, with a specific phase of the
cell cycle, such as the repositioning of the spindle during anaphase
that brings the nucleus of the CySC closer to the hub interface and might expose that CySC to more Upd ligand. This possibility implies a much more dynamic stem cell
niche for the GSCs than has been previously appreciated (Amoyel, 2013).
The results indicate that the Hh and JAK/STAT pathways act
mostly in parallel, although activating Hh may delay the
differentiation of CySCs that are deficient for JAK/STAT pathway
components. It is unclear why the CySC would require both
signaling inputs to be maintained. However, it should be noted that
these inputs contribute different information, as JAK/STAT
signaling imparts niche potential, and Hh
signaling additionally ensures that the right number of CySCs are
present and provide cyst cells for normal spermatogonial
development. Future work will establish whether self-renewal in CySCs depends on two sets of
genes controlled separately by the Hh and JAK/STAT pathways or
whether they converge on the same targets. The first possibility is
supported by the fact that Hh does not contribute to the niche
function of STAT in CySCs, indicating that different targets
(presumably BMPs) are regulated differently (Amoyel, 2013).
One consequence of this work is to lead to a reevaluation of the
differences between male and female gonad development in
Drosophila. Indeed, Hh signaling is an
essential regulator of the self-renewal and the number of follicle
stem cells, the offspring of which carry out a comparable function
to cyst cells by ensheathing germ line cysts. In the
ovary, as in the testis, JAK/STAT signaling in somatic cells is
required for the maintenance of GSCs via BMP production. However, in
the ovary, the escort cells and cap cells are the JAK/STAT-responsive
niche cells, implying that CySCs in the male gonad fulfill the function
of two cell types in the female gonad and require both the signals
used in the female to do so. Finally, the data evoke the interesting
possibility that Hh has a conserved ancestral role in male gonads.
Mutation in one of the three mammalian hh homologs, desert
hedgehog (Dhh), causes male sterility and a loss of somatic support cells called Leydig cells. However, the cellular niche for spermatogenesis in mammals is less well understood than in Drosophila and it remains to be established whether the Hh pathway orchestrates similar cellular functions (Amoyel, 2013).
Stem cell niches provide resident stem cells with signals that specify
their identity. Niche signals act over a short range such that only stem
cells but not their differentiating progeny receive the self-renewing
signals. However, the cellular mechanisms that limit niche signalling to
stem cells remain poorly understood. This study shows that the Drosophila
male germline stem cells form
previously unrecognized structures, microtubule-based nanotubes, which
extend into the hub, a major niche component. Microtubule-based nanotubes
are observed specifically within germline stem cell populations, and
require intraflagellar transport proteins for their formation. The bone
morphogenetic protein (BMP) receptor Tkv
localizes to microtubule-based nanotubes. Perturbation of
microtubule-based nanotubes compromises activation of Dpp
signalling within germline stem cells, leading to germline stem cell
loss. Moreover, Dpp ligand and Tkv receptor interaction is necessary and
sufficient for microtubule-based nanotube formation. The study proposes
that microtubule-based nanotubes provide a novel mechanism for selective
receptor-ligand interaction, contributing to the short-range nature of
niche-stem-cell signalling (Inaba, 2015).
The Drosophila testis represents an excellent model system to study niche-stem-cell interactions because of its well-defined anatomy: eight to ten germline stem cells (GSCs) are attached to a cluster of somatic hub cells, which serve as a major component of the stem cell niche. The hub secretes at least two ligands: the cytokine-like ligand Unpaired (Upd), and a BMP ligand Decapentaplegic (Dpp), both of which regulate GSC maintenance. GSCs typically divide asymmetrically, so that one daughter of the stem cell division remains attached to the hub and retains stem cell identity, while the other daughter, called a gonialblast, is displaced away from the hub and initiates differentiatio. Given the close proximity of GSCs and gonialblasts, the ligands (Upd and Dpp) must act over a short range so that signalling is only active in stem cells, but not in differentiating germ cells. The basis for this sharp boundary of pathway activation remains poorly understood (Inaba, 2015).
Using green fluorescent protein (GFP)-α1-tubulin84B expressed in germ cells (nos-gal4>UAS-GFP-αtub), this study found that GSCs form protrusions, referred to as microtubule-based (MT)-nanotubes hereafter, that extend into the hub. MT-nanotubes are sensitive to fixation similar to other thin protrusions reported so far, such as tunnelling nanotubes, explaining why they have escaped detection in previous studies. MT-nanotubes appear to be specific to GSCs: 6.67 MT-nanotubes were observed per testis in the GSC population (or 0.82 per cell). The average thickness and length of MT-nanotubes are 0.43 ± 0.29 µm (at the base of MT-nanotube) and 3.32 ± 1.6 µm, respectively. These GSC MT-nanotubes are uniformly oriented towards the hub area. By contrast, differentiating germ cells showed only 0.44 MT-nanotubes per testis (or <0.002 per cell), without any particular orientation when present. MT-nanotubes were sensitive to colcemid, the microtubule-depolymerizing drug, but not to the actin polymerization inhibitor cytochalasin B, suggesting that MT-nanotubes are microtubule-based structures. MT-nanotubes were not observed in mitotic GSCs, and GSCs form new MT-nanotubes as they exit from mitosis. By contrast, MT-nanotubes in interphase GSCs were stably maintained for up to 1 h of time-lapse live imaging. Although cell-cycle-dependent formation of MT-nanotube resembles that of primary cilia, MT-nanotubes are distinct structures, in that they lack acetylated microtubules and are sensitive to fixation. Furthermore, a considerable fraction of GSCs form multiple MT-nanotubes per cell (54% of GSCs with MT-nanotubes), and MT-nanotubes are not always associated with the centrosome/basal body, as is the case for the primary cilia (Inaba, 2015).
To examine the geometric relationship between MT-nanotubes and hub cells further, MT-nanotubes were imaged in combination with various cell membrane markers, followed by three-dimensional rendering. Although the MT-nanotubes are best visualized in unfixed testes that express GFP-αTub in germ cells, adding a low concentration (1 μM) of taxol to the fixative preserves MT-nanotubes, allowing immunofluorescence staining. First, Armadillo (Arm, β-catenin) staining, which marks adherens junctions formed at hub cell/hub cell as well as hub cell/GSC boundaries, revealed that adherens junctions do not form on the surface of MT-nanotubes. Using FM4-64 styryl dye, it was found that the MT-nanotubes are ensheathed by membrane lipids. Furthermore, myristoylation/palmitoylation site GFP (myrGFP), a membrane marker, expressed in either the germline or hub cells illuminated MT-nanotubes, suggesting that the surface membrane of a MT-nanotube is juxtaposed to hub-cell plasma membrane (Inaba, 2015).
Genes were examined that regulate primary cilia and cytonemes for their possible involvement in MT-nanotube formation. RNA interference (RNAi)-mediated knockdown of oseg2 (IFT172), osm6 (IFT52) and che-13 (IFT57), components of the intraflagellar transport (IFT)-B complex that are required for primary cilium anterograde transport and assembly, significantly reduced the length and the frequency of MT-nanotubes. Knockdown of Dlic, a dynein intermediate chain required for retrograde transport in primary cilia<, also reduced the MT-nanotube length and frequency. Knockdown of klp10A, a Drosophila homologue of mammalian kif24 (a MT-depolymerizing kinesin of the kinesin-13 family, which suppresses precocious cilia formation), resulted in abnormally thick/bulged MT-nanotubes. No significant changes were observed in MT-nanotube morphology upon knockdown of IFT-A retrograde transport genes, such as oseg1 and oseg3 (Inaba, 2015).
Endogenous Klp10A localized to MT-nanotubes both in wild-type testes and in GFP-αTub-expressing testes. GFP-Oseg2 (IFT-B), GFP-Oseg1, GFP-Oseg3 (IFT-A) and Dlic also localized to the MT-nanotubes when expressed in germ cells. The localization of IFT-A components to MT-nanotubes, without detectable morphological abnormality upon mutation/knockdown, is reminiscent of the observation that most of the genes for IFT-A are not required for primary cilia assembly. Expression of a dominant negative form of Dia (DiaDN) or a temperature-sensitive form of Shi (Shits) in germ cells (nos-gal4>UAS-diaDN or UAS-shits), which perturb cytoneme formation, did not influence the morphology or frequency of MT-nanotubes in GSCs. Taken together, these results show that primary cilia proteins localize to MT-nanotubes and regulate their formation (Inaba, 2015).
In search of the possible involvement of MT-nanotubes in hub-GSC signalling, it was found that the Dpp receptor, Thickveins (Tkv), expressed in germ cells (nos-gal4>tkv-GFP) was observed within the hub region, in contrast to GFP alone, which remained within the germ cells. A GFP protein trap of Tkv (in which GFP tags Tkv at the endogenous locus) also showed the same localization pattern as Tkv-GFP expressed by nos-gal4. By inducing GSC clones that co-express Tkv-mCherry and GFP-αTub, it was found that Tkv-mCherry localizes along the MT-nanotubes as puncta. Furthermore, using live observation, Tkv-mCherry puncta were observed to move along the MT-nanotubes marked with GFP-αTub, suggesting that Tkv is transported towards the hub along the MT-nanotubes. It should be noted that, in the course of this study, it was noticed that mCherry itself localized to the hub when expressed in germ cells, similar to Tkv-GFP and Tkv-mCherry. Importantly, the receptor for Upd, Domeless (Dome), predominantly stayed in the cell body of GSCs, demonstrating the specificity/selectivity of MT-nanotubes in trafficking specific components of the niche signalling pathways. A reporter of ligand-bound Tkv, TIPF localized to the hub region together with Tkv-mCherry, in addition to its reported localization at the hub-GSC interface. Furthermore, Dpp-GFP expressed by hub cells co-localized with Tkv-mCherry expressed in germline. These results suggest that ligand (Dpp)-receptor (Tkv) engagement and activation occurs at the interface of the MT-nanotube surface and the hub cell plasma membrane. Knockdown of IFT-B components (oseg2RNAi, che-13RNAi or osm6RNAi), which reduces MT-nanotube formation, resulted in reduction of the number of Tkv-GFP puncta in the hub area, concomitant with increased membrane localization of Tkv-GFP. A similar trend was observed upon treatment of the testes with colcemid, suggesting that MT-nanotubes are required for trafficking of Tkv into the hub area. By contrast, knockdown of Klp10A, which causes thickening of MT-nanotubes, led to an increase in the number of Tkv-GFP puncta in the hub area. Taken together, these data suggest that Tkv is trafficked into the hub via MT-nanotubes, where it interacts with Dpp secreted from the hub (Inaba, 2015).
Knockdown of klp10A (klp10ARNAi) led to elevated phosphorylated Mad (pMad) levels, a readout of Dpp pathway activation, in GSCs. By contrast, RNAi-mediated knockdown of oseg2, osm6 and che-13 (IFT-B components), which causes shortening of MT-nanotubes, reduced the levels of pMad in GSCs. Dad-LacZ, another readout of Dpp signalling activation, exhibited clear upregulation upon knockdown of klp10A. GSC clones of che-13RNAi, osm6RNAi or oseg2452 were lost rapidly compared with control clones, consistent with the idea that MT-nanotubes help to promote Dpp signal transduction. Knockdown of oseg2, che-13 and osm6 did not visibly affect cytoplasmic microtubules, suggesting that GSC maintenance defects upon knockdown of these genes are probably mediated by their role in MT-nanotube formation. Global RNAi knockdown of these genes in all GSCs using nos-gal4 did not cause a significant decrease in GSC numbers , indicating that compromised Dpp signalling due to MT-nanotube reduction leads to a competitive disadvantage in regards to GSC maintenance only when surrounded by wild-type GSCs (Inaba, 2015).
When klp10ARNAi GSC clones were induced, pMad levels specifically increased in those GSC clones, indicating that Klp10A acts cell-autonomously in GSCs to influence Dpp signal transduction. Importantly, klp10ARNAi spermatogonia did not show a significant elevation in pMad level compared with control spermatogonia, demonstrating that the role of Klp10A in regulation of Dpp pathway is specific to GSCs. pMad levels did not change in spermatogonia upon manipulation of MT-nanotube formation. GSC clones of klp10ARNAi or klp10A null mutant (klp10A24) did not dominate in the niche, despite upregulation of pMad, possibly because of its known role in mitosis. Importantly, these conditions did not significantly change STAT92E levels, which reflect Upd-JAK-STAT signalling in GSCs, revealing the selective requirement of MT-nanotubes in Dpp signalling. Together, these results demonstrate that MT-nanotubes specifically promote Dpp signalling and their role in enhancing the Dpp pathway is GSC specific (Inaba, 2015).
Since cytonemes are induced/stabilized by the signalling molecules themselves, the possible involvement of Dpp in MT-nanotube formation was explored First, it was found that a temperature-sensitive dpp mutant (dpphr56/dpphr4) exhibited a dramatic decrease in the frequency of MT-nanotubes (0.067 MT-nanotubes per GSC) and the remaining MT-nanotubes were significantly thinner. Knockdown of tkv (tkvRNAi) in GSCs also resulted in reduced length and frequency of MT-nanotubes. Conversely, overexpression of Tkv (tkvOE) in germ cells led to significantly longer MT-nanotubes. Interestingly, expression of a dominant negative Tkv (tkvDN), which has intact ligand-binding domain but lacks its intracellular GS domain and kinase domain, resulted in thickening of MT-nanotubes, rather than reducing the thickness/length. This indicates that ligand-receptor interaction, but not downstream signalling events, is sufficient to induce MT-nanotube formation. Strikingly, upon ectopic expression of Dpp in somatic cyst cells (tj-lexA>dpp), spermatogonia/spermatocytes were observed to have numerous MT-nanotubes, suggesting that Dpp is necessary and sufficient to induce or stabilize MT-nanotubes in the neighbouring germ cells. In turn, MT-nanotubes may promote selective ligand-receptor interaction between hub and GSCs, leading to spatially confined self-renewal (Inaba, 2015).
This study shows that previously unrecognized structures, MT-nanotubes, extend into the hub to mediate Dpp signalling. It is proposed that MT-nanotubes form a specialized cell surface area, where productive ligand-receptor interaction occurs. In this manner, only GSCs can access the source of highest ligand concentration in the niche via MT-nanotubes, whereas gonialblasts do not experience the threshold of signal transduction necessary for self-renewal, contributing to the short-range nature of niche signalling. In summary, the results reported here illuminate a novel mechanism by which the niche specifies stem cell identity in a highly selective manner (Inaba, 2015).
Neutral competition, an emerging feature of stem cell homeostasis, posits that individual stem cells can be lost and replaced by their neighbors stochastically, resulting in chance dominance of a clone at the niche. A single stem cell with an oncogenic mutation could bias this process and clonally spread the mutation throughout the stem cell pool. The Drosophila testis provides an ideal system for testing this model. The niche supports two stem cell populations that compete for niche occupancy. This study shows that cyst stem cells (CySCs) conform to the paradigm of neutral competition and that clonal deregulation of either the Hedgehog (Hh) or Hippo (Hpo) pathway allows a single CySC to colonize the niche. The driving force behind such behavior is accelerated proliferation. These results demonstrate that a single stem cell colonizes its niche through oncogenic mutation by co-opting an underlying homeostatic process (Amoyel, 2014).
To function properly, tissue-specific stem cells must reside in a niche. The Drosophila testis niche is one of few niches studied in vivo. Here, a single niche, comprising ten hub cells, maintains both germline stem cells (GSC) and somatic stem cells (cyst stem cells, CySC). This study shows that lines is an essential CySC factor. Surprisingly, lines-depleted CySCs adopted several characteristics of hub cells, including the recruitment of new CySCs. This led to an examination of the developmental relationship between CySCs and hub cells. In contrast to a previous report, no significant conversion was seen of steady-state CySC progeny to hub fate. However, it was found that these two cell types derive from a common precursor pool during gonadogenesis. Furthermore, embryos mutant for lines, an obligate antagonist of bowl function (Hatini, 2005), exhibited gonads containing excess hub cells, indicating that lines represses hub cell fate during gonadogenesis. In many tissues, lines acts antagonistically to bowl, and it was found that this is true for hub specification, establishing bowl as a positively acting factor in the development of the testis niche (Dinardo, 2011).
This analysis together with previous lineage-tracing shows that some hub cells and some CySCs are derived from the SGPs of PS11. The remaining CySCs could in principle derive from either PS10 or PS12. Currently, neither of those mesodermal parasegments can be uniquely lineage traced. However, the remaining hub cells probably derive from PS10 SGPs, as that would fit with the identification of receptor tyrosine kinase signaling as an antagonist of hub fate among posterior SGPs (Dinardo, 2011).
Aside from pathways known to repress hub fate, work is also beginning to identify positive functions necessary to specify these cells. This study found that bowl is one factor, as mutants had fewer hub cells, and those present appeared compromised for hub cell function. Still, the existence of residual hub cells suggests that Bowl is not the only factor required for hub cell specification, and, indeed, Notch signaling is a second positively acting component (Dinardo, 2011).
It is of interest that both Notch and bowl are positively required for hub cell specification, since these two genes act together in several other tissues. However, the exact epistatic relationship between bowl and the Notch pathway can be complex. There is some evidence that Notch activation leads to Bowl accumulation. Since it was found that Notch and also the relief-of-repression hierarchy consisting of drm/lines/bowl acts during hub cell specification, a simple model would be that Notch activation induces an antagonist of lines, for example, drm. This allows Bowl protein to accumulate in a subset of SGPs and to promote hub fate, while SGPs that retain functional Lines would adopt CySC fate. Attractive as this model is, testing some of its predictions was difficult. Attempts to visualize endogenous protein accumulation for Bowl and for Lines in the gonad has been frustrating. In addition, although drm mutants had reduced hub cell number, drm-expressing cells have not been identified within the forming gonad (Dinardo, 2011).
Thus, the relationship between Notch and the drm/lines/bowl cassette may be indirect, an outcome of the fact that both systems use the co-repressor Groucho. It has been suggested that conditions which alter the levels of available Bowl, such as in drm (down) or lines (up) mutants, could reciprocally affect the amount of Groucho available to Suppressor of Hairless, which requires this co-repressor to maintain repression of Notch target genes. Whether or not the relationship between Notch and Bowl for hub cell specification is direct, loss of Notch has a stronger phenotype than loss of bowl. Thus, the Notch pathway must also engage a separate pathway that specifies some hub cells (Dinardo, 2011).
During gonadogenesis, the current model suggests that Lines represses hub fate and promotes CySC fate. It is intriguing that a requirement for lines persists in CySCs during the steady-state operation of the testis. Analysis at this later stage suggests that lines plays a similar, though not identical, role. Although cells in gonads from lines mutant embryos fully adopt hub cell fate, in the testis the lines-depleted CySCs only partially adopt hub fate, as they do not recruit new GSCs. Thus, at steady-state, some additional regulation over the distinction between CySC and hub cell fate has been added on. Such a factor(s) remain to be identified (Dinardo, 2011).
Even the partial conversion of lines mutant CySCs into hub cells is an intriguing phenotype. Recently, a lineage relationship has been described for several stem cell-niche pairs, where stem cells can generate cells of their niche. These include production of Paneth cells in the mammalian intestine, the production of transient niche cells in the fruitfly intestine, and the repair of ependymal cells by neural progenitors of the sub-ventricular zone. In the steady-state testis, it was recently suggested that CySCs can efficiently generate new hub cells. Thus, it is considered whether lines might be deployed at steady state to govern this transition, but no increase was detected in conversion in flies with decreased lines gene dose. In fact, in wild type it was found that the conversion of cells into hub fate was insignificant compared with what has been reported. As one method used in this study was essentially identical to one used in the original report, the reason for the discrepancy is uncertain. Lineage-marking was very efficient. For example, two days after delivery of FLP by one heat-shock, 85% of testes possessed a labeled CySCs, with an average of 1.5 CySCs per testis. In the previous report, a similar regimen produced only 13% of testes with labeled CySCs. Still, it is not clear how an increase in marking efficiency could account for a decrease in apparent frequency of conversion of CySC progeny into hub cells (Dinardo, 2011).
Thus, since CySCs do not normally generate hub cells, why might lines function be maintained in CySCs so long after its embryonic requirement? The favored model is that lines is deployed during steady-state for a distinct purpose. For example, recent work on the lines/bowl cassette suggests that it assists in signal integration. This idea is appealing as the niche cells and their local environment are subjected to the action of a number of signaling pathways, such as Hh, Wnt, BMP, Jak/STAT and EGFR. Currently, it is not fully understood how these pathways function in the steady-state operation of the niche, nor how signals from distinct pathways integrate to produce a single outcome. Even the dogma of the heavily studied Jak/STAT pathway continues be challenged and refined by recent data. Perhaps as newer data uncovers the nuanced roles of several of these pathways, the lines/bowl cassette will figure into the integration of those signals (Dinardo, 2011).
Finally, the fact that lines-depleted CySCs recruited neighboring wild-type somatic cells to adopt CySC fate is striking. Although the imaging tools necessary to reveal which somatic cells are recruited to CySC fate are unavailable, the fact of their recruitment suggests that under these mutant conditions cyst cells can de-differentiate into CySCs. It has been elegantly shown that maturing germ cells can de-differentiate, creating new GSCs. As those maturing germ cells are encysted by the somatic cyst cells, during de-differentiation this grouping must break apart to release individual germline cells that repopulate the niche. Whether cyst cells de-differentiate to CySCs in these cases has not been directly assessed. If this happens under physiological conditions, it would be of great interest to study how cyst cells de-differentiation occurs, and testes harboring lines-deficient clones might aid in such studies (Dinardo, 2011).
Drosophila is widely used to analyse functions of different genes. The phosphatidylcholine lysophospholipase gene swiss cheese was initially shown to be important in the fruit fly nervous system. However, the role of this gene in non-nervous cell types has not been elucidated yet, and the evolutional explanation for the conservation of its function remains elusive. This study analyses the expression pattern and some aspects of the role of the swiss cheese gene in the fitness of Drosophila melanogaster. The spatiotemporal expression of swiss cheese throughout the fly development is described and the survival and productivity of swiss cheese mutants were analyzed. swiss cheese was found to be expressed in salivary glands, midgut, Malpighian tubes, adipocytes, and male reproductive system. Dysfunction of swiss cheese results in severe pupae and imago lethality and decline of fertility, which is impressive in males. The latter is accompanied with abnormalities of male locomotor activity and courtship behavior, accumulation of lipid droplets in testis cyst cells and decrease in spermatozoa motility. These results suggest that normal swiss cheese is important for Drosophila melanogaster fitness due to its necessity for both specimen survival and their reproductive success (Melentev, 2021).
Stem cells can be controlled by their local microenvironment, known as the stem cell niche. The Drosophila testes contain a morphologically distinct niche called the hub, composed of a cluster of between 8 and 20 cells known as hub cells, which contact and regulate germline stem cells (GSCs) and somatic cyst stem cells (CySCs). Both hub cells and CySCs originate from somatic gonadal precursor cells during embryogenesis, but whereas hub cells, once specified, cease all mitotic activity, CySCs remain mitotic into adulthood. Cyst cells, derived from the CySCs, first encapsulate the germline and then, using occluding junctions, form an isolating permeability barrier. This barrier promotes germline differentiation by excluding niche-derived stem cell maintenance factors. This study shows that the somatic permeability barrier is also required to regulate stem cell niche homeostasis. Loss of occluding junction components in the somatic cells results in hub overgrowth. Enlarged hubs are active and recruit more GSCs and CySCs to the niche. Surprisingly, hub growth results from depletion of occluding junction components in cyst cells, not from depletion in the hub cells themselves. Moreover, hub growth is caused by incorporation of cells that previously express markers for cyst cells and not by hub cell proliferation. Importantly, depletion of occluding junctions disrupts Notch and mitogen-activated protein kinase (MAPK) signaling, and hub overgrowth defects are partially rescued by modulation of either signaling pathway. Overall, these data show that occluding junctions shape the signaling environment between the soma and the germline in order to maintain niche homeostasis (Fairchild, 2016).
The hub regulates stem cell behavior in multiple ways. First, the hub physically anchors the stem cells by forming an adhesive contact with both germline stem cells (GSCs) and cyst stem cells (CySCs). The hub thus provides a physical cue that orients centrosomes such that stem cells predominantly divide asymmetrically, perpendicular to the hub. Following asymmetric stem cell division, one daughter cell remains attached to the hub and retains stem cell identity while the other is displaced from the hub and differentiates. Second, hub cells produce signals, including the STAT ligand Unpaired-1 (Upd), Hedgehog (Hh), and the BMP ligands Decapentaplegic (Dpp) and Glass-bottomed boat (Gbb), that signal to the adjacent stem cells to maintain their identity. As germ cells leave the stem cell niche, two somatic cyst cells surround and encapsulate them to form a spermatocyst. As spermatocysts move from the apical to the basal end of the testis, both somatic cyst cells and germ cells undergo a coordinated program of differentiation. Previous studies have shown that differentiation of encapsulated germ cells requires their isolation behind a somatic occluding junction-based permeability barrier. Specifically, a role was identified for septate junctions, which are functionally equivalent to vertebrate tight junctions, in establishing and maintaining a permeability barrier for each individual spermatocyst (Fairchild, 2016).
During analysis of septate junction protein localization, it was observed that some, notably Coracle, were expressed in both the hub and the differentiating cyst cells. Moreover, knockdown of septate junction components in the somatic cells of the gonad resulted in enlarged hubs. Based on these results, the role of septate junction components in regulating the number of hub cells was explored in detail. To this end, RNAi was used to knock down the expression of the core septate junction components Neurexin-IV (Nrx-IV) and Coracle (Cora) in both the hub and cyst cell populations and the number of hub cells counted in testes from newly eclosed and 7-day-old adults. RNAi was expressed using three tissue-specific drivers: upd-Gal4, expressed in hub cells; tj-Gal4, expressed weakly in hub cells and strongly in both CySCs and early differentiating cyst cells; and eyaA3-Gal4, expressed strongly in all differentiating cyst cells, weakly in CySCs, and at negligible levels in the hub. To visualize hub cells, multiple established hub markers, including upd-Gal4, upd-lacZ, Fasciclin-III (FasIII), and DN-cadherin (DNcad) were used. Surprisingly, it was found that knockdown of Nrx-IV or cora driven by upd-Gal4 gave rise to normal hubs. In comparison, knockdown of Nrx-IV or cora using tj-Gal4 or eyaA3-Gal4 led to large increases in the number of the hub cells. Hub growth was not uniform and varied between testes, but median hub cells numbers in Nrx-IV and cora knockdown testes grew by 30% and 55%, respectively, between 1 and 7 days post-eclosion (DPEs). However, in extreme cases, hubs contained up to five times the number of cells found in age-matched control testes. This result was confirmed using a series of controls that discounted the possibility that hub overgrowth was due to temperature or leaky expression of the RNAi lines. These results suggested that hub growth occurred as a result of knockdown of septate junction proteins in cyst cells rather than the hub. This was further supported using another somatic driver that is not thought to be expressed in the hub, c587-Gal4. However, analysis of c587-Gal4 was complicated by the fact this driver severely impacted fly viability when combined with Nrx-IV or cora RNAi lines (Fairchild, 2016).
Intriguingly, hub growth largely occurred after adult flies eclosed and not in earlier developmental stages. For example, when the driver eyaA3-Gal4 was used to knock down Nrx-IV or cora, hubs from 1-day-old adults were not larger than controls, but hubs from 7-day-old adults were significantly larger. Moreover, overgrowth phenotypes were recapitulated when temperature-sensitive Gal80 was used to delay induction of eyaA3-Gal4-mediated Nrx-IV knockdown until after eclosion. Hub growth manifested both in a higher mean number of hub cells per testis and by a shift in the distribution of hub cells per testis upward, toward larger hubs sizes. This distribution suggested a gradual, stochastic process of hub growth, resulting in a population of testes containing a range of hub sizes. These results reveal progressive hub growth in adults upon knockdown of septate junction components in cyst cells and suggest that this growth is not driven by events occurring in the hub itself but rather by events occurring in cyst cells (Fairchild, 2016).
Niche size has been shown in various tissues, including vertebrate hematopoietic stem cells and somatic stem cells in the fly ovary, to be an important factor in regulating the number of stem cells that the niche can support. In the fly testes, it has been shown that mutants having few hub cells could nonetheless maintain a large population of GSCs. To determine how a larger hub, containing more cells, affects niche function, the number of GSCs and CySCs was monitored after knockdown of septate junction components in cyst cells. Overall, the average number of germ cells contacting the hub grew substantially in Nrx-IV or cora knockdown testes between 1 and 7 DPEs. To confirm that the germ cells contacting the hub were indeed GSCs, spectrosome morphology was studied, and it was found to be to be consistent with that seen in wild-type GSCs. Moreover, in individual testes, there was a positive correlation between the number of hub cells and the number of GSCs. Similar growth was also observed in the number of CySCs, defined as cyst cells expressing Zfh1, but not the hub cell marker DNcad. Control testes (from tj-Gal4 x w1118 progeny) had on average 34.3 CySCs whereas Nrx-IV and cora knockdown testes had 53.4 and 50.2 CySCs, respectively. These results show the importance of maintaining a stable stem cell niche size, as enlarged hubs were active and could support additional stem cells, which may result in the excess production of both germ cells and cyst cells (Fairchild, 2016).
Next, it was of interest to determine the mechanism driving hub growth in adult flies upon knockdown of septate junction components in cyst cells. One possible mechanism that can explain this growth is hub cell proliferation. However, a defining feature of hub cells is that they are not mitotically active. Consistent with this, a large number of testes were stained for the mitotic marker phospho-histone H3 (pH3), and cells were never observed where upd-LacZ and pH3 were detected simultaneously. These results argue that division of hub cells is unlikely to explain hub growth in the adult Nrx-IV and cora knockdown testes. To determine the origin of the extra hub cells, the lineage of eyaA3-expressing cells was traced using G-TRACE (Evans, 2009). eyaA3 was chosed as both the expression pattern of septate junctions, and Nrx-IV or cora knockdown results suggested that hub growth involved differentiating eyaA3-positive cyst cells. The eyaA3-Gal4 driver utilizes a promoter region of the eya gene, which is required for somatic cyst cell differentiation and is expressed at very low levels in CySCs and at high levels in differentiating cyst cells. Using G-TRACE allows identification of both cells that previously expressed eyaA3-Gal4 (marked with GFP) and cells currently expressing eyaA3-Gal4 (marked with a red fluorescent protein [RFP]); additionally, the hub was identified using expression of upd-LacZ and FasIII. In control experiments at both 1 and 7 DPEs, few GFP-positive cells were observed in the hub. Those few GFP-positive cells could be explained by the transient expression of eya in the embryonic somatic gonadal precursor cells that form both hub and cyst cell lineages or extremely low levels of expression in adult hub cells. When G-TRACE was combined with knockdown of Nrx-IV, the results were strikingly different. Initially, 1 DPE, hubs were only slightly larger than controls and few GFP-positive hub cells were observed. In comparison, 7-DPE hubs contained on average more than twice as many cells compared to controls. Importantly, hub growth in Nrx-IV knockdowns was largely attributable to the incorporation of GFP-positive cells. Moreover, a population of upd-LacZ-labeled cells that were also RFP-positive was observed consistent with ongoing or recent expression of eyaA3-Gal4 in hub cells. These results suggest that knockdown of Nrx-IV or cora leads cyst cells to adopt hallmarks of hub cell identity and express hub-cell-specific genes (Fairchild, 2016).
To learn more about the differentiation state of non-endogenous hub cells in Nrx-IV and cora knockdown testes, various markers were used to label the stem cell niche. This analysis showed normal expression of hub cell markers, such as Upd, FasIII, DNcad, as well as Hedgehog (hh-LacZ), Armadillo (Arm), and DE-Cadherin (DEcad). It was asked how cells that were previously, and in some instances were still, eyaA3 positive could express multiple hub-cell fate markers. To answer this question, the signaling mechanisms that determine hub fate were investigated in Nrx-IV and cora knockdown testes. Hub growth phenotypes similar to those produced by Nrx-IV and cora knockdown have been described previously, most notably in agametic testes that lack germ cells, suggesting that the germline regulates the formation of hub cells. One specific germline-derived signal shown to regulate hub fate is the epidermal growth factor (EGF) ligand Spitz. In embryonic testes, somatic cells express the EGF receptor (EGFR), which, when activated, represses hub formation. EGFR-induced mitogen-activated protein kinase (MAPK) signaling, visualized by staining for di-phosphorylated-ERK (dpERK), was active in CySCs and spermatogonial-stage cyst cells. Quantifying dpERK-staining intensity in cyst cell nuclei showed that MAPK activity was lower in CySCs following knockdown of Nrx-IV or cora, suggesting reduced EGFR signaling. Moreover, the effect of Nrx-IV or cora knockdown on MAPK signaling was not restricted to CySCs, as lower dpERK staining was observed at a distance from the hub. To see whether disruption of EGFR signaling could underlie hub defects in Nrx-IV and cora knockdown testes, attempts were made to rescue these phenotypes by increasing EGF signaling. When a constitutively activated EGF receptor (EGFR-CA) was co-expressed in cyst cells along with Nrx-IV RNAi, hub growth was attenuated, resulting in a reduction in the average number of hub cells compared to expressing only Nrx-IV RNAi. Similar results were also observed in the growth of the GSC population, suggesting that reduced EGFR activation in cyst cells contributes to the overall growth of the stem cell niche caused by the knockdown of Nrx-IV or cora. Surprisingly, analysis of testes with loss-of-function mutations in the EGFR/MAPK pathway reveals different phenotypes than those observed: encapsulation is disrupted and CySCs are lost, but hub size is largely unaffected. This result shows that the partial reduction in EGFR/MAPK signaling seen in Nrx-IV and cora knockdown testes results in distinct phenotypes and highlights the complexity of EGFR signaling in the fly testis (Fairchild, 2016).
Another pathway that is documented to regulate hub cell fate is Notch signaling. Notch plays important roles in hub specification in embryos. The Notch ligand Delta is produced by the embryonic endoderm and acts to promote hub cell specification in the anterior-most somatic gonadal precursor cells. Whereas it has been suggested that Notch acts in the adult to regulate hub fate, such a role has not been clearly demonstrated. A reporter for the Notch ligand Delta (Dl-lacZ) was observed in hub cells of both control and Nrx-IV knockdown testes. Intriguingly, reducing Notch signaling efficiently rescued the hub overgrowth seen in adult Nrx-IV knockdown testes. When a dominant-negative Notch (Notch-DN) was co-expressed in the somatic cells, along with Nrx-IV RNAi, the growth of the hub was reduced compared to the expression of Nrx-IV RNAi alone. Growth in the GSC population was not significantly reduced by co-expression of Notch-DN, suggesting that the Notch pathways may modulate hub growth through a different mechanism compared to the EGFR pathway. Because Notch is well established to regulate hub growth in the embryo, temperature-sensitive Gal80 was used to delay expression of Notch-DN and confirm that the reduction in hub cells was due to disruption of post-embryonic Notch signaling. These results suggest that Notch signaling in cyst cells may contribute to the hub overgrowth phenotypes caused by septate junction knockdown in the adult testes (Fairchild, 2016).
In addition to Notch and EGFR, other signaling pathways that regulate hub size may contribute to the hub growth seen upon somatic knockdown of septate junction components. For example, it has been previously shown that the range of BMP signaling is expanded following Nrx-IV or cora knockdown in cyst cells. Constitutive activation of BMP signaling in the germline was shown to increase the size of the hub and the number of GSCs. Additionally, the relative expression levels of the genes drm, lines, and bowl regulate hub size in the adult. In particular, it is known that lines maintains a “steady state” in the testes by repressing expression of a subset of hub genes in the cyst cell population. Unlike lines mutants, Nrx-IV or cora knockdowns generally lack ectopic hubs. This may reflect the more gradual hub growth seen in septate junction knockdowns or, alternatively, highlight key mechanistic differences in how hub growth is achieved in each respective genetic background. The current work is consistent with the model whereby occluding junctions are required for proper soma-germline signaling in the fly testes. This signaling maintains stem cell niche homeostasis by preventing somatic cyst cells from adopting hub cell fate, which would lead to niche overgrowth. It is well established that, in embryonic testes, hub fate is both positively and negatively regulated by signals from the germline and the endoderm.The results, and recent findings about the genes lines and traffic jam, argue that, in the adult testes, hub fate is actively repressed in the cyst cell lineage. Failure to repress hub fate allows cyst cells to exhibit features of hub cells and act as a functional stem cell niche. However, these cyst-cell-derived hub cells are distinct from the true endogenous hub cells in that they show non-hub-cell features, including expression of the differentiating cyst cell markers eyaA3-Gal4 and β3-tubulin. The data suggest that, following disruption of septate junctions proteins, the signaling environment surrounding the somatic cells is altered such that cyst cells gradually begin expressing hub cell markers (Fairchild, 2016).
One major outstanding question is how eyaA3-Gal4-expressing cyst cells become incorporated into the endogenous hub. Previously, it was shown that a septate-junction-mediated permeability barrier forms by the four-cell spermatogonial-stage spermatocyst. The hub growth phenotypes induced by Nrx-IV and cora knockdowns may occur due to defects in cell-cell signaling, possibly involving EGFR and Notch, that manifest in these later spermatocysts. However, this model requires an explanation for how these cyst cells translocate back to and join the hub. Alternatively, signaling defects in these later spermatocysts are somehow instructing earlier cyst cells, such as CySCs, to join the hub. It is easier to envisage the latter model, as early cyst cells are spatially much closer to the hub, but the sequence of signaling events in such a case will be complex and require further elucidation. The ability of CySCs to convert into hub cells in wild-type testes is a controversial subject. However, the incorporation of CySCs into the hub does not necessitate complete conversion into hub cells but could rather involve simple de-repression or activation of genes that confer hub cell function, including regulators of the cell-cycle- and hub-cell-specific signaling ligands. Notably, the transition between CySC and hub cell fate is linked to the cell cycle (Fairchild, 2016).
Why would loss of the septate-junction-mediated somatic permeability barrier result in disruption of signaling between the soma and germline? There are many possible answers, but it is possible to speculate about two such mechanisms that explain hub overgrowth. One possibility is that germline differentiation, which is dependent on the permeability barrier, is required for the release of signals that maintain stem cell niche homeostasis. Another possibility is that the permeability barrier locally concentrates germline-derived signals that repress hub cell fate by trapping them in the luminal space between the encapsulating cyst cells and the germline. The latter scenario could explain the observation that activated EGFR signaling partially rescues hub overgrowth. In this model, septate junctions allow localized buildup of the EGF ligand Spitz, ensuring that sufficient signaling is available to repress hub fate. It is more difficult to draw strong conclusions about how Notch signaling is altered when septate junctions are disrupted, particularly as the Notch ligand Delta appears restricted to the hub. Overall, an unexpected role was found for an occluding-junction-based permeability barrier in mediating stem cell niche homeostasis. This work highlights how the architecture of the stem cell niche system in the fly testes, which is highly regular and contains a reproducible number of stem cells and niche cells, is in fact the result of an active and dynamic signaling environment (Fairchild, 2016).
Tight junctions prevent paracellular flow and maintain cell polarity in an epithelium. These junctions are also required for maintaining the blood-testis-barrier, which is essential for sperm differentiation. Septate junctions in insects are orthologous to the tight junctions. In Drosophila testis, major septate junction components co-localize at the interface of germline and somatic cells initially and then condense between the two somatic cells in a cyst after germline meiosis. Their localization is extensively remodeled in subsequent stages. Characteristic septate junctions are formed between the somatic cyst cells at the elongated spermatid stage. Consistent with previous reports, knockdown of essential junctional components- Discs-large-1 and Neurexin-IV- during the early stages disrupted sperm differentiation beyond the spermatocyte stage. Knockdown of these proteins during the final stages of spermatid maturation caused premature release of spermatids inside the testes, resulting in partial loss of male fertility. These results indicate the importance of maintaining the integrity of the somatic enclosure during spermatid coiling and release in Drosophila testis. It also highlights the functional similarity with the tight junction proteins during mammalian spermatogenesis (Dubey, 2019).
Spermatogenesis is a dynamic developmental process requiring precisely timed transitions between discrete stages. Specifically, the germline undergoes three transitions: from mitotic spermatogonia to spermatocytes, from meiotic spermatocytes to spermatids, and from morphogenetic spermatids to spermatozoa. The somatic cells of the testis provide essential support to the germline throughout spermatogenesis, but their precise role during these developmental transitions has not been comprehensively explored. This study describes the identification and characterization of genes that are required in the somatic cells of the Drosophila melanogaster testis for progress through spermatogenesis. Phenotypic analysis of candidate genes pinpointed the stage of germline development disrupted. Bioinformatic analysis revealed that particular gene classes were associated with specific developmental transitions. Requirement for genes associated with endocytosis, cell polarity, and microtubule-based transport corresponded with the development of spermatogonia, spermatocytes, and spermatids, respectively. Overall, this study identified mechanisms that act specifically in the somatic cells of the testis to regulate spermatogenesis (Fairchild, 2017).
The stem cell niche normally prevents aberrant stem cell behaviors that lead to cancer formation. Recent studies suggest that some cancers are derived from endogenous populations of adult stem cells that have somehow escaped from normal control by the niche. However, the molecular mechanisms by which the niche retains stem cells locally and tightly controls their divisions are poorly understood. This study demonstrates that the presence of heparan sulfate (HS), a class glygosaminoglycan chains, in the Drosophila germline stem cell niche prevents tumor formation in the testis. Loss of HS in the niche, called the hub, led to gross changes in the morphology of testes as well as the formation of both somatic and germline tumors. This loss of hub HS resulted in ectopic signaling events in the Jak/Stat pathway outside the niche. This ectopic Jak/Stat signaling disrupted normal somatic cell differentiation, leading to the formation of tumors. This finding indicates a novel non-autonomous role for niche HS in ensuring the integrity of the niche and preventing tumor formation (Levings, 2017).
Stem cell regulation by local signals is intensely studied, but less is known about the effects of hormonal signals on stem cells. In Drosophila, the primary steroid twenty-hydroxyecdysone (20E) regulates ovarian germline stem cells (GSCs) but was considered dispensable for testis GSC maintenance. Male GSCs reside in a microenvironment (niche) generated by somatic hub cells and adjacent cyst stem cells (CySCs). This study shows that depletion of 20E from adult males by overexpressing a dominant negative form of the Ecdysone receptor (EcR) or its heterodimeric partner ultraspiracle (usp) causes GSC and CySC loss that is rescued by 20E feeding, uncovering a requirement for 20E in stem cell maintenance. EcR and USP are expressed, activated and autonomously required in the CySC lineage to promote CySC maintenance, as are downstream genes ftz-f1 and E75. In contrast, GSCs non-autonomously require ecdysone signaling. Global inactivation of EcR increases cell death in the testis that is rescued by expression of EcR-B2 in the CySC lineage, indicating that ecdysone signaling supports stem cell viability primarily through a specific receptor isoform. Finally, EcR genetically interacts with the NURF chromatin-remodeling complex, which has been shown to maintain CySCs. Thus, although 20E levels are lower in males than females, ecdysone signaling acts through distinct cell types and effectors to ensure both ovarian and testis stem cell maintenance (Li, 2014).
This work shows that the steroid hormone 20E plays an important role in maintaining stem cells in theDrosophila testis: 20E, receptors of ecdysone signaling, and downstream targets are required directly in CySCs for their maintenance. When ecdysone signaling is lost in CySCs, GSCs are also lost, but it is unclear if their maintenance requires an ecdysone-dependent or independent signal from the CySCs. The requirement for EcR in the testis is isoform-specific: expression of EcR-B2 in the CySC lineage is sufficient to rescue loss of GSCs and CySCs and increased cell death in EcR mutant testes, suggesting that there might be a temporal and spatial control of ecdysone signaling in the adult testis. In addition, evidence is provided that ecdysone signaling, as in the ovary, is able to interact with an intrinsic chromatin-remodeling factor, Nurf301, to promote stem cell maintenance. Therefore, these studies have revealed a novel role for ecdysone signaling in Drosophila male reproduction (Li, 2014).
Although ecdysone signaling is required in both ovaries and testes for stem cell maintenance, the responses in each tissue are likely to be sex-specific. In the ovary, 20E controls GSCs directly, by modulating their proliferation and self-renewal, and it acts predominantly through the downstream target gene E74. In contrast, male GSCs require ecdysone signaling only indirectly: ecdysone signaling was found to be required in the CySC lineage to maintain both CySCs and GSCs. In a previous study, RNAi-mediated knockdown of EcR, usp or E75 in the CySC lineage did not result in a significant loss of GSCs; however, the number of CySCs was not determined, and the phenotype was examined after 4 or 8 days, not 14 days as in this study. It is suspected that the earlier time points used in that study may not have allowed enough time for a significant number of GSCs to be lost (Li, 2014).
During development, 20E is produced in the prothoracic gland (PG) and further metabolized to 20E in target tissues, but the PG does not persist into adulthood. In adult female Drosophila, the ovary is a source of 20E. In contrast, the identification of steroidogenic tissues in adult male Drosophila remains the subject of active investigation. The level of 20E in adult males is significantly lower than in adult females, but it can be detected in the testis. Furthermore, RNA-seq data show that shade, which encodes the enzyme that metabolizes the prohomone ecdysone to 20E, is expressed in the adult testis, suggesting that the adult testis may produce 20E. However, the sources of 20E production in adult Drosophila males remain to be determined experimentally (Li, 2014).
20E, like other systemic hormones, can have tissue-specific effects or differential effects on the same cell type as development proceeds. These differences are mediated at least in part by the particular downstream target genes that are activated in each case. For example, in female 3rd instar larval ovaries, ecdysone signaling upregulates br expression to induce niche formation and PGC differentiation, but br is not required for GSC maintenance in the adult ovary; instead, E74 plays this role. Similarly, br is required for the establishment of intestinal stem cells (ISCs) in the larval and pupal stages but not for ISC function in adults. This study shows that ecdysone signaling in the adult testis is mediated by different target genes than in the ovary: E74, but not E75 or br, regulate stem cell function in the ovary, whereas E75 and ftz-f1 are important for stem cell maintenance in the testis. Since E75 is itself a nuclear hormone receptor that responds to the second messenger nitric oxide, it will be interesting to know whether E75's partner DHR3 also plays a role in CySCs. An intriguing question for future studies will be how different ecdysone target genes interact with the various signaling pathways that maintain stem cells in the ovary or testis (Li, 2014).
Since 20E levels can actively respond to physiological changes induced by environmental cues, it is possible that the effect of 20E on testis stem cell maintenance might reflect changes in diet, stress, or other environmental cues. For example, in Aedes aegypti, ecdysteroid production in the ovary is stimulated by blood feeding and this is an insulin-dependent process. In Drosophila, ecdysone signaling is known to interact with the insulin pathway in a complex way. Ovaries from females with hypomorphic mutations in the insulin-like receptor have reduced levels of 20E. Furthermore, ecdysone signaling can directly inhibit insulin signaling and control larval growth in the fat body. Thus, ecdysone signaling may interact with insulin signaling during testis stem cell maintenance. Previously, it was shown that GSCs in the ovary and testis can respond to diet through insulin signaling, which is required to promote stem cell maintenance in both sexes. It is possible that diet can affect 20E levels and thus regulate stem cell maintenance. In addition to diet, stress can also affect 20E levels, as is the case in Drosophila virilis, where 20E levels increase significantly under high temperature stress. A similar effect has been found in mammals, where the steroid hormone cortisol is released in response to psychological stressor. Finally, 20E levels are also influenced by mating. In Anopheles gambiae, males transfer 20E to blood-fed females during copulation, which is important for egg production. In female Drosophila, whole body ecdysteroid levels also increase after mating. Studying the roles of hormonal signaling in mediating stem cell responses to stress and other environmental cues will be an exciting topic for future studies. From this work it is now clear that, as in mammals, steroid signaling plays critical roles in adult stem cell function during both male and female gametogenesis (Li, 2014).
Local signals maintain adult stem cells in many tissues. Whether the sexual identity of adult stem cells must also be maintained was not known. In the adult Drosophila testis niche, local Jak-STAT signaling promotes somatic cyst stem cell (CySC) renewal through several effectors, including the putative transcription factor Chronologically inappropriate morphogenesis (Chinmo). This study found that Chinmo also prevents feminization of CySCs. Chinmo promotes expression of the canonical male sex determination factor DoublesexM (DsxM) within CySCs and their progeny, and ectopic expression of DsxM in the CySC lineage partially rescues the chinmo sex transformation phenotype, placing Chinmo upstream of DsxM. The Dsx homolog DMRT1 prevents the male-to-female conversion of differentiated somatic cells in the adult mammalian testis, but its regulation is not well understood. This work indicates that sex maintenance occurs in adult somatic stem cells and that this highly conserved process is governed by effectors of niche signals (Ma, 2014).
The development of stem cell daughters into the differentiated state normally requires a cascade of proliferation and differentiation steps that are typically regulated by external signals. The germline cells of most animals, in specific, are associated with somatic support cells and depend on them for normal development. In the male gonad of Drosophila melanogaster, germline cells are completely enclosed by cytoplasmic extensions of somatic cyst cells, and these cysts form a functional unit. Signaling from the germline to the cyst cells via the Epidermal Growth Factor Receptor (EGFR) is required for germline enclosure and has been proposed to provide a temporal signature promoting early steps of differentiation. A temperature-sensitive allele of the EGFR ligand Spitz (Spi) provides a powerful tool for probing the function of the EGRF pathway in this context and for identifying other pathways regulating cyst differentiation via genetic interaction studies. Using this tool, this study showed that signaling via the Ecdysone Receptor (EcR), a known regulator of developmental timing during larval and pupal development, opposes EGF signaling in testes. In spi mutant animals, reducing either Ecdysone synthesis or the expression of Ecdysone signal transducers or targets in the cyst cells resulted in a rescue of cyst formation and cyst differentiation. Despite of this striking effect in the spi mutant background and the expression of EcR signaling components within the cyst cells, activity of the EcR pathway appears to be dispensable in a wildtype background. It is proposed that EcR signaling modulates the effects of EGFR signaling by promoting an undifferentiated state in early stage cyst cells (Qian, 2014).
Eukaryotic gene expression is activated by factors that interact within complex machinery to initiate transcription. An important component of this machinery is the DNA repair/transcription factor TFIIH. Mutations in TFIIH result in three human syndromes: xeroderma pigmentosum, Cockayne syndrome and trichothiodystrophy. Transcription and DNA repair defects have been linked to some clinical features of these syndromes. However, how mutations in TFIIH affect specific developmental programmes, allowing organisms to develop with particular phenotypes, is not well understood. This study shows that mutations in the p52 and p8 subunits of TFIIH have a moderate effect on the gene expression programme in the Drosophila testis, causing germ cell differentiation arrest in meiosis, but no Polycomb enrichment at the promoter of the affected differentiation genes, supporting recent data that disagree with the current Polycomb-mediated repression model for regulating gene expression in the testis. Moreover, TFIIH stability was not compromised in p8 subunit-depleted testes that show transcriptional defects, highlighting the role of p8 in transcription. Therefore, this study reveals how defects in TFIIH affect a specific cell differentiation programme and contributes to understanding the specific syndrome manifestations in TFIIH-afflicted patients (Cruz-Becerra, 2016).
Adult stem cells reside in microenvironments called niches, where they are regulated by both extrinsic cues, such as signaling from neighboring cells, and intrinsic factors, such as chromatin structure. This study reports that in the Drosophila testis niche an H3K27me3-specific histone demethylase encoded by Ubiquitously transcribed tetratricopeptide repeat gene on the X chromosome (dUTX) maintains active transcription of the Suppressor of cytokine signaling at 36E (Socs36E) gene by removing the repressive H3K27me3 modification near its transcription start site. Socs36E encodes an inhibitor of the Janus kinase signal transducer and activator of transcription (JAK-STAT) signaling pathway. Whereas much is known about niche-to-stem cell signaling, such as the JAK-STAT signaling that is crucial for stem cell identity and activity, comparatively little is known about signaling from stem cells to the niche. The results reveal that stem cells send feedback to niche cells to maintain the proper gene expression and architecture of the niche. dUTX acts in cyst stem cells (CySCs) to maintain gene expression in hub cells through activating Socs36E transcription and preventing hyperactivation of JAK-STAT signaling. dUTX also acts in germline stem cells to maintain hub structure through regulating DE-Cadherin levels. Therefore, these findings provide new insights into how an epigenetic factor regulates crosstalk among different cell types within an endogenous stem cell niche, and shed light on the biological functions of a histone demethylase in vivo (Tarayrah, 2013).
This study identified a new epigenetic mechanism that
negatively regulates the JAK-STAT signaling pathway in the
Drosophila testis niche: the H3K27me3-specific
demethylase dUTX acts in CySCs to remove the repressive
H3K27me3 histone modification near the TSS of Socs36E to allow
its active transcription. Socs36E acts upstream to suppress Stat92E
activity and to restrict CySCs from overpopulating the testis niche.
In addition, dUTX acts in CySCs to prevent hyperactivation of
Stat92E in hub cells, which would otherwise ectopically turn on
Zfh1 expression. When zfh1 cDNA was ectopically driven in hub
cells using the upd-Gal4 driver, no obvious defect could be
identified. Therefore, the biological consequence of ectopic Zfh1
expression in hub cells remains unclear. However, ectopic Zfh1
expression in hub cells and the overpopulation of Zfh1-expressing
cells around the hub are two connected phenomena because both
phenotypes are caused by loss of dUTX in CySCs (Tarayrah, 2013).
UTX also acts in GSCs to regulate DE-Cadherin levels to
maintain proper GSC-hub interaction and normal morphology of
the hub. It has been reported that differential expression of
different cadherins causes cells with similar cadherin types and
levels to aggregate. In wt testes, hub cells express higher levels of
DE-Cadherin and therefore tightly associate with each other. Loss of dUTX in germ cells leads to
higher levels of DE-Cadherin in GSCs, which probably allows
them to intermingle with hub cells and causes disrupted hub
architecture. It has also been demonstrated that the major role of
JAK-STAT in GSCs is for GSC-hub adhesion, suggesting that the expression and/or activity of
cell-cell adhesion molecules, such as DE-Cadherin, depends on
JAK-STAT signaling. Therefore, the abnormal DE-Cadherin
activity in GSCs in dUTX testis could also result from misregulated JAK-STAT signaling in the testis niche. dUTX is a new negative epigenetic regulator of the JAK-STAT signaling pathway (Tarayrah, 2013).
The JAK-STAT signaling pathway plays crucial roles in stem cell
maintenance in many different stem cell types across a wide range
of species. These studies identify the histone demethylase dUTX
as a new upstream regulator of the JAK-STAT pathway, which
directly controls the transcription of Socs36E. In addition to acting
as an antagonist of JAK-STAT signaling, Socs36E has been reported
to be a direct target gene of the Stat92E transcription factor (Terry, 2006). Therefore, increased Stat92E would be expected to
upregulate Socs36E expression, but this was not observed in dUTX
mutant testes. Instead, the data revealed that Socs36E expression
decreased, whereas Stat92E expression increased, in dUTX testes,
consistent with the hypothesis that Socs36E is a direct target gene
of dUTX and acts upstream of Stat92E (Tarayrah, 2013).
Sustained activity of the JAK-STAT pathway in cyst cells has
been reported to activate BMP signaling, which leads to GSC self-renewal
outside the niche and gives rise to a tumor-like phenotype
in testis. To examine BMP pathway activity, immunostaining experiments were performed using antibodies against phospho-SMAD (pSMAD), a downstream target of BMP signaling. No
obvious difference was detected in the pSMAD signal between the dUTX testes
and wt control, nor were any germline tumors detected in dUTX testes. It is speculated that germline tumor formation upon activation of the JAK-STAT pathway is secondary to the overproliferation of Zfh1-expressing cells, which was not observed in dUTX testes (Tarayrah, 2013).
This study also provides an example of the multidimensional cell-cell
communication that takes place within a stem cell niche. Many
studies of the stem cell niche have focused on understanding niche-to-
stem cell signaling in controlling stem cell identity and activity.
For example, in the Drosophila female GSC niche, Upd secreted
from terminal filaments activates the JAK-STAT pathway in cap
cells and escort cells, which subsequently produce the BMP
pathway ligand Decapentaplegic (Dpp) to activate BMP signaling
and prevent transcription of the differentiation factor bag-of-marbles
(bam) in GSCs. In the Drosophila intestinal stem cell (ISC) niche, the
visceral muscle cells underlying the intestine secrete Wingless to
activate Wnt signaling and Upd to activate JAK-STAT signaling in
ISCs, which are required to maintain ISC identity and activity (Tarayrah, 2013).
More studies have now revealed the multidirectionality of
signaling within the stem cell niche. For example, in the Drosophila
female GSC niche, GSCs activate Epidermal growth factor receptor
(Egfr) signaling in the neighboring somatic cells, which
subsequently represses expression of the glypican Dally, a protein
required for the stabilization and mobilization of the BMP pathway
ligand Dpp. Through this communication between GSCs and the
surrounding somatic cells, only GSCs maintain high BMP signaling. The current studies establish another example of the multidimensional cell-cell communications that occur within
the testis stem cell niche, where CySCs and GSCs have distinct roles
in regulating hub cell identity and morphology (Tarayrah, 2013).
The data identified new roles of a histone demethylase in regulating
endogenous stem cell niche architecture and proper gene expression.
Previous studies have reported in vivo functions of histone
demethylases in several model organisms. For example, mammalian
UTX has been shown to associate with the H3K4me3 histone
methyltransferase MLL2, suggesting its potential antagonistic role to the PcG proteins. The PcG proteins
play a crucial role in Hox gene silencing in both Drosophila and
mammals. Consistently, mammalian UTX has been reported to directly bind
and activate the HOXB1 gene locus. In addition
to antagonizing PcG function, H3K27me3 demethylases play crucial
roles during development. For example, in zebrafish, inactivating
the UTX homolog (kdm6al) using morpholino oligonucleotides leads
to defects in posterior development, and in C. elegans the dUTX homolog (UTX-1) is required for embryonic and postembryonic development, including gonad development. Furthermore, loss of UTX
function in embryonic stem cells leads to defects in mesoderm
differentiation, and somatic cells derived from UTX loss-of-function human or mouse tissue fail to return to the ground state of pluripotency. These reports demonstrate that UTX is not only required for proper cellular differentiation but also for successful reprogramming. However,
despite multiple reports on the in vivo roles of H3K27me3-specific
demethylases, little is known about their functions in any
endogenous adult stem cell system (Tarayrah, 2013).
Whereas mammals have multiple H3K27me3 demethylases,
dUTX is the sole H3K27me3-specific demethylase in Drosophila.
This unique feature, plus the well-characterized nature of
Drosophila adult stem cell systems, make interpretation of the
endogenous functions of histone demethylases in Drosophila
unambiguous. Because mammalian UTX has been reported as a
tumor suppressor, understanding the endogenous functions of dUTX in an adult stem cell system might
facilitate the use of histone demethylases for cancer treatment.
In summary, these results demonstrate that stem cells send
feedback to the niche cells to maintain their proper gene expression
and morphology. Furthermore, this feedback is regulated through
the JAK-STAT signaling pathway, the activity of which is controlled
by a chromatin factor, providing an example of crosstalk between
these two regulatory pathways (Tarayrah, 2013).
Stem cells in tissues reside in and receive signals from local microenvironments called niches. Understanding how multiple signals within niches integrate to control stem cell function is challenging. The Drosophila testis stem cell niche consists of somatic hub cells that maintain both germline stem cells and somatic cyst stem cells (CySCs). This study shows a role for the axon guidance pathway Slit-Roundabout (Robo) in the testis niche. The ligand Slit is expressed specifically in hub cells while its receptor, Roundabout 2 (Robo2), is required in CySCs in order for them to compete for occupancy in the niche. CySCs also require the Slit-Robo effector Abelson tyrosine kinase (Abl) to prevent over-adhesion of CySCs to the niche, and CySCs mutant for Abl outcompete wild type CySCs for niche occupancy. Both Robo2 and Abl phenotypes can be rescued through modulation of adherens junction components, suggesting that the two work together to balance CySC adhesion levels. Interestingly, expression of Robo2 requires JAK-STAT signaling, an important maintenance pathway for both germline and cyst stem cells in the testis. This work indicates that Slit-Robo signaling affects stem cell function downstream of the JAK-STAT pathway by controlling the ability of stem cells to compete for occupancy in their niche (Stine, 2014: PubMed).
Socs36E, which encodes a negative feedback inhibitor of the JAK/STAT pathway, is the first identified regulator of niche competition in the Drosophila testis. The competitive behavior of (Suppressor of cytokine signaling at 36E) (Socs36E) mutant cyst stem cells (CySCs) has been attributed to increased JAK/STAT signaling. This study shows that competitive behavior of Socs36E mutant CySCs is due in large part to unbridled Mitogen-Activated Protein Kinase (MAPK) signaling. In Socs36E mutant clones, MAPK activity is elevated. Furthermore, it was found that clonal upregulation of MAPK in CySCs leads to their outcompetition of wild type CySCs and of germ line stem cells, recapitulating the Socs36E mutant phenotype. Indeed, when MAPK activity is removed from Socs36E mutant clones, they lose their competitiveness but maintain self-renewal, presumably due to increased JAK/STAT signaling in these cells. Consistently, loss of JAK/STAT activity in Socs36E mutant clones severely impairs their self-renewal. Thus, these results enable the genetic separation of two essential processes that occur in stem cells. While some niche signals specify the intrinsic property of self-renewal, which is absolutely required in all stem cells for niche residence, additional signals control the ability of stem cells to compete with their neighbors. Socs36E is the node through which these processes are linked, demonstrating that negative feedback inhibition integrates multiple aspects of stem cell behavior (Amoyel, 2016a).
Stem cell niches are complex environments that provide support for stem cells through molecular signals. Several well-characterized niches provide not just one but multiple signals which stem cells must integrate and interpret in order to remain at the niche and self-renew. How this integration is achieved is not well understood at present. Furthermore, in order to maintain the appropriate number of stem cells and the homeostatic balance between self-renewal and differentiation, it is necessary that self-renewal cues be present in limiting amounts or that their activity be dampened to prevent excessive accumulation of stem cells. One general feature of many signal transduction pathways is the presence of feedback inhibitors. These are dampeners of signaling, transcriptionally induced by the signaling itself, that prevent signal levels from being aberrantly high. One such family of feedback inhibitors is the Suppressor of Cytokine Signaling (SOCS) proteins, which were identified as inhibitors of JAK/STAT signal transduction, and are SH2- and E3-ligase domain-containing proteins. The SH2 domain binds phosphorylated (i.e., activated) signal transduction components and the E3-ligase targets them for degradation by Ubiquitin-dependent proteolysis. In mammals, SOCS proteins can thus inhibit several tyrosine kinase-dependent signaling pathways, including JAK/STAT and Mitogen-Activated Protein Kinase (MAPK) (Amoyel, 2016a).
The Drosophila testis is an ideal model system to study questions of signal regulation and integration in stem cells. The testis niche, called the hub, supports two stem cell populations. The first, germ line stem cells (GSCs), gives rise to sperm after several transit-amplifying divisions leading up to meiosis. The second, somatic cyst stem cells (CySCs), gives rise to cyst cells, the essential support cells for germ line development. Many ligands for signaling pathways are produced by the hub, including the JAK/STAT pathway agonist, Unpaired (Upd), the Hedgehog (Hh) pathway ligand Hh and the Bone Morphogenetic Protein (BMP) homologs Decapentaplegic (Dpp) and Glass Bottom Boat (Gbb). The latter two signals are also produced by CySCs and are required in GSCs for self-renewal, indicating that CySCs constitute part of the niche for GSCs along with the hub. CySCs require JAK/STAT and Hh activity for self-renewal (Amoyel, 2016a).
CySCs and GSCs compete for space at the niche, a phenomenon that was revealed by the analysis of testes lacking the JAK/STAT feedback inhibitor Socs36E. In these animals, excessive JAK/STAT activity was detected in CySCs, and Socs36E mutant CySCs displaced the resident wild type GSCs. Additionally, it has been shown that CySCs with sustained Hh signaling or sustained Yorkie (Yki) activity also outcompeted neighboring wild type GSCs, indicating that several signaling pathways can control niche competition. Moreover, prior to out-competing GSCs, mutant CySCs displaced neighboring wild type CySCs, indicating that both intra- (CySC-CySC) and inter-lineage (CySC-GSC) competition take place in the testis. While the two types of competition appear related, in that one precedes the other, there are instances in which only intra-lineage competition takes place. While the competitive phenotype of Socs36E mutant CySCs was ascribed to increased JAK/STAT signaling, it was surprising to find that clonal gain-of-function in JAK/STAT signaling in CySCs did not induce competitive behavior, and it was concluded that loss of Socs36E did not mimic increased JAK/STAT signaling in CySC (Amoyel, 2016a).
This study addressed whether other mechanisms could account for the competitive behavior of Socs36E mutant CySCs. Because SOCS proteins can inhibit MAPK signaling in cultured cells and in Drosophila epithelial tissues, this study examined if Socs36E repression of MAPK signaling underlied the Socs36E competitive phenotype. Indeed, it was found that Socs36E inhibits MAPK signaling in CySCs during self-renewal, and that gain of MAPK activity induces CySCs to outcompete wild type CySCs and GSCs at the niche. This study dissected the genetic relationship between Socs36E and the MAPK and JAK/STAT pathways and shows that loss of Socs36E can compensate for decreased self-renewal signaling within CySCs. Thus, CySCs integrate multiple self-renewal signals through the use of a feedback inhibitor that controls at least two signaling pathways regulating stem cell maintenance at the niche (Amoyel, 2016a).
The data presented in this study implicate MAPK signaling as a major regulator of CySC competition for niche access and establish that the competitiveness of CySCs lacking Socs36E is derived primarily from their increased MAPK activity. The ability of a stem cell to self-renew reflects not only intrinsic properties but also extrinsic relationships with its neighbors. For instance, if a cell is unable to compete for space at the niche then it will be no longer able to receive short-range niche signals and will be more likely to differentiate. Conversely, if a cell is more competitive for niche space, this cell and its offspring will replace wild type neighbors and colonize the entire niche. (Amoyel, 2016a).
These data show that CySCs with increased MAPK signaling out-compete neighboring stem cells in CySC-CySC as well as CySC-GSC competition and that CySCs with reduced MAPK activity are themselves out-competed. The interpretation is favored that MAPK regulates primarily competitiveness rather than self-renewal because while MAPK mutant clones are lost from the niche, lineage-wide inhibition of the pathway does not result in a complete loss of stem cells. This contrasts with the role of JAK/STAT signaling in CySCs. Stat92E mutant CySCs are lost and lineage-wide pathway inhibition results in pronounced and rapid stem cell loss. Based on these results, it is argued that JAK/STAT signaling in CySCs primarily controls their intrinsic self-renewal capability while MAPK signaling regulates their competitiveness. Interestingly, there are important similarities between Hh and MAPK function in CySCs in that CySCs lacking Hh signal transduction are out-competed and those with sustained Hh activity out-compete wild type neighbors. Lastly, it is noted that CySCs mutant for the tumor suppressor Hippo (Hpo) (which leads to sustained Yki activation) or Abelson kinase (Abl) also have increased competitiveness, suggesting the existence of multiple inputs controlling the ability of stem cells to stay in the niche at the expense of their neighbors. In the future, it would be interesting to determine if genetic hierarchies exist between competitive pathways or if they independently converge on similar targets. One outstanding question is how altering the competitiveness of CySCs affects the maintenance of the germ line. In the case of Socs36E, MAPK, Hh and Hpo, the competitive CySC displaces not only wild type CySCs but also wild type GSCs. While these observations suggest that out-competition of CySCs and GSCs is linked, the result that Abl mutant CySCs only compete with CySCs and not with GSCs indicates that these two competitive processes are separable genetically (Amoyel, 2016a).
It is well established that Egfr/MAPK signaling is required in somatic cells for their proper differentiation and for their encystment of the developing germ line. In this study, an additional function for Egfr/MAPK was identified in the somatic stem cells, specifically that this pathway regulates competitiveness of CySCs, with each other and with GSCs. Regarding the latter, it is possible that the loss of GSCs when somatic cells have high MAPK signaling is linked to their possibly increased encystment by these cells. Indeed, recent work has shown that Egfr activity in CySCs regulates cytokinesis and maintenance stem cell fate in GSCs. It is tempting to speculate that increased somatic Egfr activity leads to increased encystment of GSCs and loss of stem cell fate in GSCs (Amoyel, 2016a).
MAPK may play a conserved role in niche competitiveness as mouse intestinal stem cells that acquire activating mutations in Ras bias normal stem cell replacement dynamics and colonize the niche. Interestingly, the activating ligand Spi is produced by germ cells, suggesting that the germ line coordinates multiple behaviors in the somatic cell lineage. In addition to transducing signals from the germ line, CySCs also receive ligands from hub cells (including Hh and the JAK/STAT ligand Upd) and they have to integrate these various stimuli. If unmitigated, the combined effect of all of these signals could produce highly competitive CySCs, with overall negative effects on niche homeostasis. The data are consistent with a model in which the induction of Socs36E by the primary self-renewal pathway (JAK/STAT) results in the restraint of a competitive trigger (MAPK) in CySCs. In this way, Socs36E acts to integrate signals from different sources and maintain homeostatic balance between resident cell populations that share a common niche (Amoyel, 2016a).
Stem cell competition has emerged as a mechanism for selecting fit stem cells/progenitors and controlling tumourigenesis. However, little is known about the underlying molecular mechanism. This study identified Mlf1-adaptor molecule (Madm), a novel tumour suppressor that regulates the competition between germline stem cells (GSCs) and somatic cyst stem cells (CySCs) for niche occupancy. Madm knockdown results in overexpression of the EGF receptor ligand vein (vn), which further activates EGF receptor signalling and integrin expression non-cell autonomously in CySCs to promote their overproliferation and ability to outcompete GSCs for niche occupancy. Conversely, expressing a constitutively activated form of the Drosophila JAK kinase (hop(Tum-l)) promotes Madm nuclear translocation, and suppresses vn and integrin expression in CySCs that allows GSCs to outcompete CySCs for niche occupancy and promotes GSC tumour formation. Tumour suppressor-mediated stem cell competition presented in this study could be a mechanism of tumour initiation in mammals (Singh, 2016).
Cdc14 is an evolutionarily conserved serine/threonine phosphatase. Originally identified in S. cerevisiae as a cell cycle regulator, its role in other eukaryotic organisms remains unclear. In Drosophila melanogaster, Cdc14 is encoded by a single gene, thus facilitating its study. Cdc14 expression is highest in the testis of adult flies and cdc14 null flies are viable. cdc14 null female and male flies do not display altered fertility. cdc14 null males, however, exhibit decreased sperm competitiveness. Previous studies have shown that Cdc14 plays a role in ciliogenesis during zebrafish development. In Drosophila, sensory neurons are ciliated. The Drosophila cdc14 null mutants have defects in chemosensation and mechanosensation as indicated by decreased avoidance of repellant substances and decreased response to touch. In addition, it was shown that cdc14 null mutants have defects in lipid metabolism and resistance to starvation. These studies highlight the diversity of Cdc14 function in eukaryotes despite its structural conservation (Neitzel, 2018).
Multicellular organisms have evolved specialized mechanisms to control transcription in a spatial and temporal manner. Gene activation is tightly linked to histone acetylation on lysine residues that can be recognized by bromodomains. Previously, the testis-specifically expressed bromodomain protein tBRD-1 was identified in Drosophila. Expression of tBRD-1 is restricted to highly transcriptionally active primary spermatocytes. tBRD-1 is essential for male fertility and proposed to act as a co-factor of testis-specific TATA box binding protein-associated factors (tTAFs) for testis-specific transcription. This study performed microarray analyses to compare the transcriptomes of tbrd-1 mutant testes and wild-type testes. The data confirmed that tBRD-1 controls gene activity in male germ cells. Additionally, comparing the transcriptomes of tbrd-1 and tTAF mutant testes revealed a subset of common target genes. Two new members of the bromodomain and extra-terminal (BET) family, tBRD-2 and tBRD-3, were also characterized. In contrast to other members of the BET family in animals, both possess only a single bromodomain, a characteristic feature of plant BET family members. Immunohistology techniques not only revealed that tBRD-2 and tBRD-3 partially co-localize with tBRD-1 and tTAFs in primary spermatocytes, but also that their proper subcellular distribution was impaired in tbrd-1 and tTAF mutant testes. Treating cultured male germ cells with inhibitors showed that localization of tBRD-2 and tBRD-3 depends on the acetylation status within primary spermatocytes. Yeast two-hybrid assays and co-immunoprecipitations using fly testes protein extracts demonstrated that tBRD-1 is able to form homodimers as well as heterodimers with tBRD-2, tBRD-3, and tTAFs. These data reveal for the first time the existence of single bromodomain BET proteins in animals, as well as evidence for a complex containing tBRDs and tTAFs that regulates transcription of a subset of genes with relevance for spermiogenesis (Theofel, 2014).
Eukaryotic chromosomes are spatially segregated into topologically associating domains (TADs). Some TADs are attached to the nuclear lamina (NL) through lamina-associated domains (LADs). This study identified LADs and TADs at two stages of Drosophila spermatogenesis - in bamΔ86 mutant testes which is the commonly used model of spermatogonia (SpG) and in larval testes mainly filled with spermatocytes (SpCs). This study found that initiation of SpC-specific transcription correlates with promoters' detachment from the NL and with local spatial insulation of adjacent regions. However, this insulation does not result in the partitioning of inactive TADs into sub-TADs. It was also revealed an increased contact frequency between SpC-specific genes in SpCs implying their de novo gathering into transcription factories. In addition, the specific X chromosome organization was uncovered in the male germline. In SpG and SpCs, a single X chromosome is stronger associated with the NL than autosomes. Nevertheless, active chromatin regions in the X chromosome interact with each other more frequently than in autosomes. Moreover, despite the absence of dosage compensation complex in the male germline, randomly inserted SpG-specific reporter is expressed higher in the X chromosome than in autosomes, thus evidencing that non-canonical dosage compensation operates in SpG (Ilyin, 2022).
To selectively express cell type-specific transcripts during development, it is critical to maintain genes required for other lineages in a silent state. This study shows in the Drosophila male germline stem cell lineage that a spermatocyte-specific zinc finger protein, Kumgang (Kmg; CG5204), working with the chromatin remodeler dMi-2 prevents transcription of genes normally expressed only in somatic lineages. By blocking transcription from normally cryptic promoters, Kmg restricts activation by Aly, a component of the testis-meiotic arrest complex, to transcripts for male germ cell differentiation. These results suggest that as new regions of the genome become open for transcription during terminal differentiation, blocking the action of a promiscuous activator on cryptic promoters is a critical mechanism for specifying precise gene activation (Kim, 2017).
Highly specialized cell types such as red blood cells, intestinal epithelium, and spermatozoa are produced throughout life from adult stem cells. In such lineages, mitotically dividing precursors commonly stop proliferation and initiate a cell type-specific transcription program that sets up terminal differentiation of the specialized cell type. In the Drosophila male germ line, stem cells at the apical tip of the testis self-renew and produce daughter cells that each undergo four rounds of spermatogonial mitotic transit amplifying (TA) divisions, after which the germ cells execute a final round of DNA synthesis (premeiotic S-phase) and initiate terminal differentiation as spermatocytes. Transition to the spermatocyte state is accompanied by transcriptional activation of more than 1500 genes, many of which are expressed only in male germ cells. Expression of two-thirds of these depends both on a testis-specific version of the MMB (Myb-Muv B)/dREAM (Drosophila RBF, dE2F2, and dMyb-interacting proteins) complex termed the testis meiotic arrest complex (tMAC) and on testis-specific paralogs of TATA-binding protein-associated factors (tTAFs). Although this is one of the most dramatic changes in gene expression in Drosophila, it is not yet understood how the testis-specific transcripts are selectively activated during the 3-day spermatocyte period (Kim, 2017).
To identify the first transcripts up-regulated at onset of spermatocyte differentiation, germ cells were genetically manipulated to synchronously differentiate from spermatogonia to spermatocytes in vivo using bam-/- testes, which contain large numbers of overproliferating spermatogonia. Brief restoration of Bam expression under heat shock control in hs-bam;bam-/- flies induced synchronous differentiation of bam-/- spermatogonia, resulting in completion of a final mitosis, premeiotic DNA synthesis, and onset of spermatocyte differentiation by 24 hours after Bam expression, eventually leading to production of functional sperm. Comparison by means of microarray of transcripts expressed before versus 24 hours after heat shock of hs-bam;bam-/- testes identified 27 early transcripts that were significantly up-regulated more than twofold in testes from hs-bam;bam-/- but not from bam-/- flies subjected to the same heat shock regime. Among these was the early spermatocyte marker RNA binding protein 4 (Rbp4). At this early time point, the transcript for CG5204 - now named kumgang (kmg), from the Korean name of mythological guardians at the gate of Buddhist temples - had the greatest increase among all 754 Drosophila predicted transcription factors (Kim, 2017).
Kumgang (CG5204) encodes a 747-amino acid protein with six canonical C2H2-type zinc finger domains expressed in testes but not in ovary or carcass. Kmg protein was expressed independently from the tMAC component Always early (Aly) or the tTAF Spermatocyte Arrest (Sa), and both kmg mRNA and protein were up-regulated before Topi, another component of tMAC. Immunofluorescence staining of wild-type testes revealed Kmg protein expressed specifically in differentiating spermatocytes, where it was nuclear and enriched on the partially condensed bivalent chromosomes. Consistent with dramatic up-regulation of kmg mRNA after the switch from spermatogonia to spermatocyte, expression of Kmg was first detected with immunofluorescence staining after completion of premeiotic S-phase marked by down-regulation of Bam, coinciding with expression of Rbp4 protein (Kim, 2017).
Function of Kmg in spermatocytes was required for male germ cell differentiation. Reducing function of Kmg in spermatocytes-either by means of cell type-specific RNA interference (RNAi) knockdown (KD) or in flies trans-heterozygous for a CRISPR (clustered regularly interspaced short palindromic repeats)-induced kmg frameshift mutant and a chromosomal deficiency (kmgΔ7/Df)-resulted in accumulation of mature primary spermatocytes arrested just before the G2/M transition for meiosis I and lack of spermatid differentiation. A 4.3-kb genomic rescue transgene containing the 2.3-kb kmg open reading frame fully rescued the differentiation defects and sterility of kmgΔ7/Df flies, confirming that the meiotic arrest phenotype was due to loss of function of Kmg. In both kmg KD and kmgΔ7/Df, Kmg protein levels were less than 5% that of wild type. kmgΔ7/Df mutant animals were adult-viable and female-fertile but male-sterile, which is consistent with the testis-specific expression (Kim, 2017).
Function of Kmg was required in germ cells for repression of more than 400 genes not normally expressed in wild-type spermatocytes. Although the differentiation defects caused by loss of function of kmg appeared, by means of phase contrast microscopy, to be similar to the meiotic arrest phenotype of testis-specific tMAC component mutants, analysis of gene expression in kmg KD testes showed that many Aly (tMAC)-dependent spermatid differentiation genes were expressed, although some at a lower level than that in wild type. Among the 652 genes with more than 99% lower expression in aly-/- mutant as compared with wild-type testes, only four showed similar reduced expression in kmg KD as compared with that of sibling control (no Gal4 driver) testes. In contrast, transcripts from more than 500 genes were strongly up-regulated in kmg KD testes, with almost no detectable expression in testes from sibling control males. Hierarchical clustering identified 440 genes specifically up-regulated in kmg KD testes compared with testes from wild-type, bam-/-, aly-/-, or sa-/- mutant flies. These 440 genes were significantly associated with Gene Ontology terms such as 'substrate specific channel activity' or 'detection of visible light' that appeared more applicable to non-germ cell types, such as neurons. Analysis of published transcript expression data for a variety of Drosophila tissues revealed that the 440 were normally not expressed or extremely low in wild-type adult testes, but many were expressed in specific differentiated somatic tissues such as eye, brain, or gut. Confirming misexpression of neuronal genes at the protein level, immunofluorescence staining revealed that the neuronal transcription factor Prospero (Pros), normally not detected in male germ cells, was expressed in clones of spermatocytes that are homozygous mutant for kmg induced by Flp-FRT-mediated mitotic recombination. The misexpression of Pros was cell-autonomous, occurring only in mutant germ cells. Mid-stage to mature spermatocytes homozygous mutant for kmg misexpressed Pros, but mutant early spermatocytes did not, indicating that the abnormal up-regulation of Pros occurred only after spermatocytes had reached a specific stage in their differentiation program (Kim, 2017).
A small-scale cell type-specific RNAi screen of chromatin regulators revealed that KD of dMi-2 in late TA cells and spermatocytes resulted in meiotic arrest, similar to loss of function of kmg. Immunofluorescence analysis of testes from a protein trap line in which an endogenous allele of dMi-2 was tagged by green fluorescent protein (GFP) revealed that dMi-2-GFP, like the untagged endogenous protein, was expressed and nuclear in progenitor cells and spermatocytes, as well as in somatic hub and cyst cells. dMi-2-GFP colocalized to chromatin with Kmg in spermatocytes, and the level of dMi-2 protein appeared lower and less concentrated on chromatin in nuclei of kmg-/- spermatocytes than in neighboring kmg+/+ or kmg+/- spermatocytes, suggesting that Kmg may at least partially help recruit dMi-2 to chromatin in spermatocytes. Furthermore, in testis extracts Kmg coimmunoprecipitated with dMi-2 and vice versa, suggesting that Kmg and dMi-2 form a protein complex in spermatocytes. Comparison of microarray data revealed that most of the 440 transcripts up-regulated in testes upon loss of function of kmg were also abnormally up-regulated in dMi-2 KD testes, suggesting that Kmg and dMi-2 may function together to repress expression of the same set of normally somatic transcripts in spermatocytes (Kim, 2017).
Chromatin immunoprecipitation followed by sequencing (ChIP-seq) revealed that Kmg protein localized along the bodies of genes actively transcribed in the testis. ChIP-seq with antibody to Kmg identified 798 genomic regions strongly enriched by immunoprecipitation of Kmg from wild-type but not from kmg KD testes. Of the 798 robust Kmg ChIP-seq peaks, 698 overlapped with exonic regions of 680 different genes actively transcribed in testes. The enrichment was often strongest just downstream of the transcription start site (TSS), but with substantial enrichment along the gene body as well (Kim, 2017).
ChIP-seq with antibody to dMi-2 also showed enrichment along the gene bodies of the same 680 genes bound by Kmg, with a similar bias just downstream of the TSS. The dMi-2 ChIP signal along these genes was partially reduced in kmg KD testes, suggesting that Kmg may recruit dMi-2 to the bodies of genes actively transcribed in the testis (Kim, 2017).
RNA-seq analysis revealed that the 680 genes bound by Kmg were strongly expressed in testes and most strongly enriched in the GO term categories 'spermatogenesis' and 'male gamete generation'. One-third of the genes bound by Kmg were robustly activated as spermatogonia differentiate into spermatocytes and were much more highly expressed in the testes than in other tissues. The median levels of transcript expression of most of the 680 Kmg bound genes did not show appreciable change upon loss of Kmg (Kim, 2017).
Genes that are normally transcribed in somatic cells that became up-regulated upon loss of Kmg function in spermatocytes for the most part did not appear to be bound by Kmg. Only 3 of the 440 genes up-regulated in kmg KD overlapped with the 680 genes with robust Kmg peaks, suggesting that Kmg may prevent misexpression of normally somatic transcripts either indirectly or by acting at a distance (Kim, 2017).
Inspection of RNA-seq reads from kmg and dMi-2 KD testes mapped onto the genome showed that ~80% of the transcripts that were detected with microarray analysis as misexpressed in KD as compared with wild-type testes did not initiate from the promoters used in the somatic tissues in which the genes are normally expressed. Metagene analysis, as well as visualization of RNA expression centered on the TSSs annotated in the Ensembl database, showed that most of the 143 genes that are normally expressed in wild-type heads but not in wild-type testes were misexpressed in kmg or dMi-2 KD testes from a start site different from the annotated TSS used in heads. Transcript assembly from RNA-seq data by using Cufflinks for the 143 genes also showed that the transcripts that are misexpressed in kmg or dMi-2 KD testes most often initiate from different TSSs than the transcripts from the same gene assembled from wild-type heads (Kim, 2017).
Of the 440 genes scored via microarray as derepressed in kmg KD testes, 346 could be assigned with TSSs in kmg KD testes based on visual inspection of the RNA-seq data mapped onto the genome browser. Of these, only 67 produced transcripts in kmg KD testes that started within 100 base pairs (bp) of the TSS annotated in the Ensembl database, based on the tissue(s) in which the gene was normally expressed. In contrast, for the rest of the 346 genes, the transcripts expressed in kmg KD testes started from either a TSS upstream (131 of 346) or downstream (148 of 346) of the annotated TSSs. Of the 346 genes, 262 were misexpressed starting from nearly identical positions in dMi-2 KD as in kmg KD testes, suggesting that Kmg and dMi-2 function together to prevent misexpression from cryptic promoters (Kim, 2017).
Many of the ectopic promoters from which the misexpressed transcripts originated appeared to be bound by Aly, a component of tMAC, in kmg KD testes. ChIP for Aly was performed by using antibody to hemagglutinin (HA) on testis extracts from flies bearing an Aly-HA genomic transgene able to fully rescue the aly-/- phenotype. Of 346 genes with new TSSs assigned via visual inspection, 181 had a region of significant enrichment for Aly as detected with ChIP, with its peak summit located within 100 bp of the cryptic promoter. Motif analysis by means of MEME revealed that these regions were enriched for the DNA sequence motif (AGYWGGC). This motif was not significantly enriched in the set of 165 cryptic promoters at which Aly was not detected in kmg KD testes. Enrichment of Aly at the cryptic promoters was much stronger in kmg KD as compared with wild-type testes, suggesting that in the absence of Kmg, Aly may bind to and activate misexpression from cryptic promoters (Kim, 2017).
Genetic tests revealed that the misexpression of somatic transcripts in kmg KD spermatocytes indeed required function of Aly. The neuronal transcription factor Pros, abnormally up-regulated in kmg KD or mutant spermatocytes, was no longer misexpressed if the kmg KD spermatocytes were also mutant for aly, even though germ cells in kmg KD;aly-/- testes appear to reach the differentiation stage at which Pros turned on in the kmg KD germ cells. Assessment by means of quantitative reverse transcription polymerase chain reaction (RT-PCR) revealed that misexpression of five out of five transcripts in kmg KD testes also required function of Aly. Global transcriptome analysis via microarray of kmg KD versus kmg KD;aly-/- testes showed that the majority of the 440 genes that were derepressed because of loss of function of kmg in spermatocytes were no longer abnormally up-regulated in kmg KD;aly-/- testes. Even genes without noticeable binding of Aly at their cryptic promoters were suppressed in kmg KD;aly-/-, suggesting that Aly may regulate this group of genes indirectly (Kim, 2017).
Together, the ChIP and RNA-seq data show that Kmg and dMi-2 bind actively transcribed genes but are required to block expression of aberrant transcripts from other genes that are normally silent in testes. The mammalian ortholog of dMi-2, CHD4 (Mi-2β), has been shown to bind active genes in mouse embryonic stem cells or T lymphocyte precursors but also plays a role in ensuring lineage-specific gene expression in other contexts. It cannot be ruled out that Kmg and dMi-2 might also act directly at the cryptic promoter sites but that the ChIP conditions did not capture their transient or dynamic binding because several chromatin remodelers or transcription factors, such as the thyroid hormone receptor, have been difficult to detect with ChIP. Kmg and dMi-2 may repress misexpression from cryptic promoters indirectly by activating as-yet-unidentified repressor proteins. However, it is also possible that Kmg and dMi-2 act at a distance by modulating chromatin structure or confining transcriptional initiation or elongation licensing machinery to normally active genes (Kim, 2017).
Changes in the genomic localization of Aly protein in wild-type versus kmg KD testes raised the possibility that Kmg may in part prevent misexpression from cryptic promoters by concentrating Aly at active genes. Of the 1903 Aly peaks identified with ChIP from wild-type testes, the 248 Aly peaks that overlapped with strong Kmg peaks showed via ChIP an overall reduction in enrichment of Aly from kmg KD testes as compared with wild type. In contrast, the Aly peaks at cryptic promoters were more robust in kmg KD testes than in wild type. In general, over the genome 4129 new Aly peaks were identified by means of ChIP from kmg KD testes that were absent or did not pass the statistical cutoff in wild-type testes. More than 30% of the genomic regions with new Aly peaks in kmg KD showed elevated levels of RNA expression starting at or near the Aly peak in kmg KD but not in wild-type testes, suggesting that misexpression of transcripts from normally silent promoters in kmg KD testes is more widespread than initially assessed with microarray. Together, these findings raise the possibility that Kmg may prevent misexpression of aberrant transcript by concentrating Aly to active target genes in wild-type testes, preventing binding and action of Aly at cryptic promoter sites (Kim, 2017).
The results suggest that selective gene activation is not always mediated by a precise transcriptional activator but can instead be directed by combination of a promiscuous activator and a gene-selective licensing mechanism. Cryptic promoters may become accessible as chromatin organization is reshaped to allow expression of terminal differentiation transcripts that were tightly repressed in the progenitor state. It is posited that this chromatin organization makes a number of sites that are accessible for transcription dependent on the testis-specific tMAC complex component Aly. In this context, activity of Kmg and dMi-2 is required to prevent productive transcript formation from unwanted initiation sites, potentially by confining Aly to genes actively transcribed in the testis and limiting the amount of Aly protein acting at cryptic promoters (Kim, 2017).
The initiation of transcripts from cryptic promoters is reminiscent of loss of function of Ikaros, a critical regulator of T and B cell differentiation and a tumor suppressor in the lymphocyte lineage. Like Kmg, Ikaros is a multiple-zinc finger protein associated with Mi-2β, which binds to active genes in T and B cell precursors. In T cell lineage acute lymphoblastic leukemia (T-ALL) associated with loss of function of Ikaros, cryptic intragenic promoters were activated, leading to expression of ligand-independent Notch1 protein, contributing to leukemogenesis. Thus, in addition to being detrimental for proper differentiation, firing of abnormal transcripts from normally cryptic promoters because of defects in chromatin regulators may contribute to tumorigenesis through generation of oncogenic proteins (Kim, 2017).
Reactive oxygen species (ROS) are byproducts generated during normal cellular metabolism, and redox states have been shown to influence stem cell self-renewal and lineage commitment across phyla. However, the downstream effectors of ROS signaling that control stem cell behavior remain largely unexplored. This study used the Drosophila testis as an in vivo model to identify ROS-induced effectors that are involved in the differentiation process of germline stem cells (GSCs). In the Affymetrix microarray analysis, 152 genes were either upregulated or downregulated during GSC differentiation induced by elevated levels of ROS, and a follow-up validation of the gene expression by qRT-PCR showed a Spearman's rho of 0.9173 (P<0.0001). Notably, 47 (31%) of the identified genes had no predicted molecular function or recognizable protein domain. These suggest the robustness of this microarray analysis, which identified many uncharacterized genes, possibly with an essential role in ROS-induced GSC differentiation. maf-S was shown to be transcriptionally downregulated by oxidative stress, and maf-S knockdown promotes GSC differentiation, but Maf-S overexpression conversely results in an over-growth of GSC-like cells by promoting the mitotic activity of germ cell lineage. Together with the facts that Maf-S regulates ROS levels and genetically interacts with Keap1/Nrf2 in GSC maintenance, this study suggests that Maf-S plays an important role in the Drosophila testis GSC maintenance by participating in the regulation of redox homeostasis (Tan, 2017).
Genotoxic stress such as irradiation causes a temporary halt in tissue regeneration. The ability to regain regeneration depends on the type of cells that survived the assault. Previous studies showed that this propensity is usually held by the tissue-specific stem cells. However, stem cells cannot maintain their unique properties without the support of their surrounding niche cells. This study shows that exposure of Drosophila melanogaster to extremely high levels of irradiation temporarily arrests spermatogenesis and kills half of the stem cells. In marked contrast, the hub cells that constitute a major component of the niche remain completely intact. It was further shown that this atypical resistance to cell death relies on the expression of certain antiapoptotic microRNAs (miRNAs) that are selectively expressed in the hub and keep the cells inert to apoptotic stress signals. It is proposed that at the tissue level, protection of a specific group of niche cells from apoptosis underlies ongoing stem cell turnover and tissue regeneration (Volin, 2018).
Cancer stem cells, in contrast to their more differentiated daughter cells, can endure genotoxic insults, escape apoptosis, and cause tumor recurrence. Understanding how normal adult stem cells survive and go to quiescence may help identify druggable pathways that cancer stem cells have co-opted. This study utilized a genetically tractable model for stem cell survival in the Drosophila gonad to screen drug candidates and probe chemical-genetic interactions. This study employs three levels of small molecule screening: a medium-throughput primary screen in male germline stem cells (GSCs), a secondary screen with irradiation and protein-constrained food in female GSCs, and a tertiary screen in breast cancer organoids in vitro. This study uncover a series of small molecule drug candidates that may sensitize cancer stem cells to apoptosis. This study assessed these small molecules for chemical-genetic interactions in the germline and identified the NF-kappaB pathway as an essential and druggable pathway in GSC quiescence and viability. This study demonstrates the power of the Drosophila stem cell niche as a model system for targeted drug discovery (Ishibashi, 2021).
In gene silencing, Hsp90 chaperone machinery assists Argonaute (Ago) Gene silencing mediated by non-coding small RNAs (sRNAs) underlies diverse biological processes, including development, homeostasis, immunity, and reproduction in eukaryotes. Lines of evidence indicate that the expression and steady-state accumulation of regulatory sRNAs themselves are controlled by multiple, intertwined mechanisms, the misregulation of which has been implicated in disease. However, understanding the regulation of sRNA expression in vivo is still incomplete (Iki, 2020).
The sRNAs cooperate with Argonaute (Ago) family proteins, forming the core effector, namely, RNA-induced silencing complex (RISC). Many sRNAs are generated through the processing of double-stranded RNAs (dsRNAs) by RNase III enzymes. Most microRNAs (miRNAs) are derived from imperfectly base-paired hairpin structures within non-coding transcripts or intronic fragments, whereas short interfering RNAs (siRNAs) originate from highly base-paired dsRNA of both endogenous and exogenous origins. Cleavages by RNase III excise ~20 to 24-nucleotide (nt) mi/siRNAs as duplex intermediates, which upon loading onto Ago proteins, become unwound during RISC formation. One strand is selectively stabilized as a 'guide strand' that recognizes the target RNA displaying sRNA-complementary sites, enabling gene silencing. The complementary so called 'passenger strand,' by contrast, is eliminated. Depending on various cell/tissue/organ-specific features, including but not restricted to sRNA alternative processing, 3'-end tailing, or even editing, a guide strand can become a passenger strand and vice versa. Hence, RISC formation represents one one of the key steps determining the steady-state accumulation of specific sRNA cohorts in vivo (Iki, 2020).
Previous studies in both animals and plants demonstrated the fundamental requirement of the Hsp70/90 chaperone machinery for RISC formation. One of the core factors, Hsp90, is of particular importance by retaining Ago in an 'apo' state, an open conformation competent for sRNA duplex incorporation. By an open-close cycle driven by ATP binding and its hydrolysis, Hsp90 chaperone homodimers transiently interact with selected client proteins as well as distinct co-chaperones at particular steps. Hsp90 then assists multiple reactions, including the binding of small ligands to client proteins. As one of Hsp90 co-chaperones, Cyclophilin 40 (CYP40)/SQUINT promotes miRNA activities in plants by facilitating the binding of duplexes (ligands) to Ago1 (client). In mammalian cells, other Hsp90 co-chaperones, namely, FK506-binding protein (Fkbp)4/Fkbp52 and the homolog Fkbp5/Fkbp51, interact with Ago2 to promote RISC formation. However, a broader range of ligand-client binding reactions can be controlled by these Hsp90 co-chaperones. Indeed, both Cyp40 and Fkbp52 were first isolated as cofactors of nuclear receptors (clients) binding to steroid hormones (ligands) in animals. These interactions were functionally characterized in various studies involving cultured mammalian cells, yeast systems, or purified proteins. Consistently, Fkbp52 knockout mice exhibit reproductive defects accompanied with attenuated steroid sensitivity. In contrast, Fkbp52 Fkbp51 double-knockout individuals die in the early embryonic stage, implying other important, yet uncharacterized, functions for both co-chaperones, possibly including a role in RISC formation. However, the biological significance of Fkbp52- and Fkbp51-regulated RISC formation has yet to be addressed in vivo, and likewise, Cyp40 knockout animals have been unavailable except in the case of a protist model. Consequently, physiological functions for animal Cyp40 remain largely elusive (Iki, 2020).
Drosophila melanogaster provides an excellent model to study the commonalities and specificities of the miRNA versus siRNA pathways. Among the five Ago protein members encoded in the fly genome, Ago1 and Ago2 incorporate RNase III Dicer 1 (Dcr1)-dependent miRNA and Dcr2-dependent siRNA duplexes. The other three Ago proteins are dedicated to the loading of Piwi-interacting RNA accumulating prominently in the gonads in a Dcr-independent manner. Both structural and nucleotide sequence features of mi/siRNA duplexes underpin their sorting into Ago1 or Ago2, as well as the subsequent guide-strand selection process. In this respect, fly Ago1 hosts a wide range of miRNAs and is essential for development from early embryogenesis. In contrast, fly Ago2 is primarily required for antiviral defense by interacting with virus-derived siRNAs. Independent of its defensive roles, Ago2 is also loaded with endogenous siRNA (endo-siRNA) in testes to support the male reproductive functions. Apart from viral and endo-siRNAs, a smaller and selective fraction of miRNAs is preferentially sorted into Ago2 rather than Ago1, although the biological roles of these rarer Ago2-bound miRNAs has remained largely unexplored (Iki, 2020).
Whether, in addition to their key role in activating the apo-Ago conformation, (co-)chaperones exhibit RISC regulatory activities influencing sRNA steady-state accumulation remains an open question, as does the potential tissue- or cell-type specificity of these putative regulatory functions. To explore these issues and, more generally, better understand the physiological roles of Hsp90 co-chaperones in animals, the loss-of-function alleles of cyp40, fkbp4, and fkbp5 orthologs were generated and characterized in D. melanogaster. Although the fkbp mutant died at the pupal stage, the cyp40 mutant was eclosed but the male exhibited infertility, a phenotype resembling that of the ago2 mutant. These detailed analyses of Ago-bound sRNAs in testes show that Cyp40 is selectively required for the accumulation of testis-unique and functional miRNAs sorted to Ago2. The significance of these results is discussed in a broader context of how regulation on specific sRNA repertoires might be enabled by tissue-specialized chaperone expression (Iki, 2020).
This study uncovered an essential function of Cyp40 as an Hsp90 co-chaperone during late spermatogenesis in Drosophila. The in vitro physical interaction analyses and in vivo Ago-bound sRNA profiling revealed a molecular activity of Cyp40 in regulating the steady-state accumulation of miRNAs uniquely forming RISC with Ago2 in testes. Although the Drosophila miRNA-sorting mechanism is now well established, the biological basis underlying the binding of a cohort of miRNAs to Ago2 had remained elusive. This study has uncovered one such mechanism and biological circumstance by showing how, under the control of germline Cyp40, Ago2 efficiently binds testis-unique miRNAs. Because the genetic ablation of one of these miRNAs suffice to noticeably impair proper spermatid differentiation, these findings also support the notion that the miRNA-Ago2 sorting pathway, which is conserved in Drosophila, is physiologically relevant (Iki, 2020).
Although miRNA hairpins are usually considered to spawn a single dominant guide strand, previous studies reported guide-strand alteration, or arm switching, in a tissue- and/or developmental-stage-specific manner. This study has identified several such miRNAs that are subjected not only to altered guide-strand selection but also to unconventional Ago sorting in a testis-unique manner. It was further demonstrated the functional relevance of these alternative events with the example of miR-316. What mechanisms might underlie the strand/Ago-switching phenomenon? The central unpaired nucleotides, terminal instabilities of sRNA duplexes, and 5' nucleotide identities largely govern, together, Ago sorting and strand selection in Drosophila. In both plants and metazoans, such duplex-intrinsic features are, however, modifiable through alternative precursor processing by RNase III enzymes. As a result, excised duplex isoforms acquire novel fates through RISC formation. In the case of miR-316, the majority of strands share an identical 5' end, reflecting a fixed, as opposed to flexible, processing mechanism. Thus, the switching strand-selection behavior of miR-316 is most likely attributable to other mechanisms. They could include RNA editing, as observed in mammals; although, testing this possibility in flies will require further work. Notably, loss of cyp40 function perturbed Ago2 binding to miR-316-3p, and yet, neither abrogated the 3'-over-5' strand preference nor affected Ago1 sorting. Hence, Cyp40 might in part influence the 5'/3'-strand use bias, but it is unlikely to regulate strand switching per se (Iki, 2020).
Besides these considerations of the strand selection of some miRNA species, Cyp40 had a profound, global effect on discriminating miRNAs from endo-siRNAs bound to Ago2. In flies, miRNA and siRNA duplexes are delivered to Ago2 by loading complexes constituted of Dcr-2 and co-partners R2D2 or Loqs-PD. By asymmetry sensing, the complexes preferentially bind near-perfectly base-paired siRNA duplexes to enable their efficient loading, whereas they simultaneously disfavor duplexes containing central instability, a feature of many miRNAs. The process is thought, among other possibilities, to avoid the saturation of Ago2 by miRNAs during virus infections, which would presumably compromise siRNA-mediated antiviral defenses. This circumstance added to the fact that many arthropod viruses encode silencing suppressor proteins probably explain why Ago2 and Dcr2, unlike Ago1 and Dcr1, are among the fastest evolving 3% of all Drosophila genes, owing to positive selection as a consequence of the never ending host-pathogen arms race. Nonetheless, these results are in line with the observation that the Drosophila argo2 gene has also evolved under the adaptive basis of testis specialization, most likely independently of antiviral roles. Indeed, as much as Ago2-bound hp-siRNAs have co-evolved with target genes and assist male fertility, the data suggest that miRNAs, such as miR-316, can also gain functional roles through unconventional Ago2 sorting in testes. However, as typical miRNAs, Cyp40-dependent duplexes displayed central bulges or mismatches, of which a likely consequence is their predicted low loading efficacy into Ago2. This might have constituted one of the driving forces for the deployment of a germline Cyp40, which, together with Hsp90, likely improves the affinity of Ago2 for the kinetically less favorable miRNA members (Iki, 2020).
Together with the TPR domain tethering Cyp40 to Hsp90 during RISC formation, the PPIase domain is required to form functional chaperone complexes in both flies and plants. The PPIase domain of human Cyp40/PPID maintains the activity for the cis-trans isomerization of peptidyl prolines. Further biochemical and structural analyses could help decipher how this, and perhaps other activities of the poorly characterized PPIase domain, enable fly and plant Cyp40 to regulate sRNA during RISC formation. In addition, Fkbp, among the Hsp90 co-chaperones regulating ligand-client binding, displays a similar domain architecture to Cyp40, but endows distinct physiological functions. The generated fkbp59-mutant-based study should help illuminate the molecular basis for such a difference (Iki, 2020).
Why is Cyp40 effective with Ago2 rather than Ago1? In testes, Ago2 appears to be relatively limited compared to Ago1, whereas its main endogenous cargoes, hairpin-siRNA duplexes, are actively produced. This likely creates a competition between miRNA and siRNA duplexes for a limited pool of Ago2. In the duplex-competing condition, Cyp40 activity could ensure Ago2 loading of suboptimal miRNA duplexes, which would otherwise be largely excluded from the effector phase of the testicular silencing pathway. However, Ago1 is inert for most endo-siRNAs and appears to be relatively abundant, compared to Ago2, in testes. This might allow Ago1 to naturally evade the testicular duplex-competing condition, explaining why Cyp40 interacts with Ago1 in vitro and yet is dispensable for steady-state accumulation of most Ago1-bound miRNAs in testes. Thus, the relative abundance of duplexes compared to that of the sRNA-free Ago proteins could be key for influencing the selectivity and effectiveness of Cyp40. The same rationale could also apply to other Hsp90 co-chaperones involved in RISC formation and sRNA regulation in other biological contexts and/or tissues (Iki, 2020).
The cyp40 mutant male showed a completely sterile phenotype, which appears to be exerted by the combined deficiencies of Ago2-sorted multiple miRNAs, including, but not restricted to, miR-316-3p. Similar to cyp40, argo2 mutant males are sterile in Drosophila simulans but are only semi-fertile in D. melanogaster, raising the question of how the loss of an Ago2 modulator, Cyp40, might cause a more severe phenotype than the loss of Ago2 itself. Perhaps the mis-accumulation of some sRNAs and the ensuring disparities among RISCs in the cyp40 mutant have deeper detrimental consequences than those caused by the mere depletion of the effector Ago2. For instance, it was shown that the reduction of selected miRNAs over-activates, in turn, other species, possibly engaging in a whole new boast of not primarily intended regulations. However, the possibility is not excluded that Cyp40 plays multiple roles in testes, in addition to the control of sRNA bound to Ago2. Also noteworthy as a caveat-inducing limitation of the approach taken in this study, Ago-bound sRNAs were isolated from whole testes, including somatic cells where cyp40 is dispensable. Refining this study using sRNA profiles from the purified germline cells, especially from the spermatocyte and spermatid stages, would likely confer more sensitivity to sRNA analyses and help more comprehensively decipher the sRNA-related molecular aberrations caused by loss of cyp40 function in the germline (Iki, 2020).
Producing mature spermatozoa is essential for sexual reproduction in metazoans. Spermiogenesis involves dramatic cell morphological changes going from sperm tail elongation, nuclear reshaping to cell membrane remodeling during sperm individualization and release. The sperm manchette, a temporary structure that assists in sperm elongation, plays a critical scaffolding function during nuclear remodeling by linking the nuclear lamina to the cytoskeleton. This study describes the role of an uncharacterized protein in Drosophila, Salto/CG13164, involved in nuclear shaping and spermatid individualization. Salto has a dynamic localization during spermatid differentiation, being progressively relocated from the sperm nuclear dense body, which is equivalent to the mammalian sperm manchette, to the centriolar adjunct and acrosomal cap during spermiogenesis. salto null male flies are sterile and exhibit complete spermatid individualization defects. salto deficient spermatids show coiled spermatid nuclei at late maturation stages and stalled individualization complexes. This work sheds light on a novel component involved in cytoskeleton based cell morphological changes during spermiogenesis (Augiere, 2019).
The niche critically controls stem cell behavior, but its regulatory input at the whole-genome level is poorly understood. This study elucidated transcriptional programs of the somatic and germline lineages in the Drosophila testis and genome-wide binding profiles of Zfh-1 and Abd-A expressed in somatic support cells and crucial for fate acquisition of both cell lineages. Key roles were identified of nucleoporins and V-ATPase proton pumps, and their importance was demonstrated in controlling germline development from the support side. To make the dataset publicly available, an interactive analysis tool was generated, that uncovered conserved core genes of adult stem cells across species boundaries. The functional relevance of these genes was tested in the Drosophila testis and intestine, and a high frequency of stem cell defects was found. In summary, this dataset and interactive platform represent versatile tools for identifying gene networks active in diverse stem cell types (Tamirisa, 2018).
Using cell-type-specific transcriptome profiling and in vivo TF binding site mapping together with an interactive data analysis tool, this study comprehensively identified genes involved in controlling proliferation and differentiation within a stem cell support system. Importantly, many candidates that were functionally tested not only were required within the soma, but also had non cell-autonomous functions in the adjacent (germline) stem cell lineage (Tamirisa, 2018).
An interconnected network of TFs was identified that plays an important role in the maintenance and differentiation of both germline and somatic cell populations, signal processing V-ATPase proton pumps, and nuclear-transport-engaged Nups as regulators in the Drosophila male stem cell system. V-ATPases have been implicated in the regulation of various cellular processes in not only invertebrates but also vertebrates. For example, the V-ATPase subunit V1e1 was previously shown to be essential for the maintenance of NBs in the developing mouse cortex, as loss of this subunit caused a reduction of endogenous Notch signaling and a depletion of NBs by promoting their differentiation into neurons. Furthermore, two independent studies revealed that V-ATPase subunits and their isoforms are required for proper spermatogenesis in mice, in particular for acrosome acidification and sperm maturation. Thus, it is tempting to speculate that these proton pumps also have important functions in the stem cell pool of the mammalian testis and very likely many other stem cell systems, and some evidence is provided for their crucial role also in ISCs (Tamirisa, 2018).
This work also uncovered nuclear transport associated proteins, the Nups, as important control hubs in the somatic lineage of the Drosophila testis. This is of particular interest, since cell-type-specific functions of Nups have been identified only recently and may represent a critical feature of different stem cell systems. Examples include Nup153, one of the Nups expressed in all four stem cell systems, which interacts with Sox2 neural progenitors and controls their maintenance as well as neuronal differentiation; Nup358, which plays a role at kinetochores, and Nup98, which regulates the anaphase promoting complex (APC) and mitotic microtubule dynamics to promote spindle assembly. Interestingly, it has been shown just recently that Nups play a critical role in regulating cell fate during early Drosophila embryogenesis, thereby contributing to the commitment of pluripotent somatic nuclei into distinct lineages. Current results suggest that they may play a similar role in controlling the transition of continuously active adult stem cells toward differentiation. The next challenge will be to unravel how variations in the composition of an essential and basic protein complex like the NPC causes differential responses of cells, in particular in stem cells and their progenies (Tamirisa, 2018).
The datasets in conjunction with the versatile and easy-to-use analysis tool allowed identification of a substantial number of stem cell regulators for detailed mechanistic characterization. Importantly, this analyses have shed first light on processes and genes shared between diverse invertebrate and vertebrate stem cell systems and uncovered functionally relevant differences. Owing to its flexibility and the option to include datasets from any species, the online tool represents a valuable resource for the entire stem cell community. It not only provides an open platform for data analysis but also leverages the power of comparative analysis to enable researchers mining genomic datasets from diverse origins in a meaningful and intuitive fashion (Tamirisa, 2018).
Adult stem cells are often found in specialized niches, where the constituent cells direct self-renewal of their stem cell pool. The niche is therefore crucial for both normal homeostasis and tissue regeneration. In many mammalian tissues, niche cells have classically been difficult to identify, which has hampered any understanding of how tissues first construct niches during development. Fortunately, the Drosophila germline stem cell (GSC) niche is well defined, allowing for unambiguous identification of both niche cells and resident stem cells. The testis niche: this study followed pro-niche cells as they assemble and assume their final form. After ex vivo culture the niche appears fully functional, as judged by enrichment of adhesion proteins, the ability to activate STAT in adjacent GSCs, and to direct GSCs to divide orthogonally to the niche, just as they would in situ. Collectively, imaging has generated several novel insights on niche morphogenesis that could not be inferred from fixed images alone. Dynamic processes were identified that constitute an assembly phase and a compaction phase during morphogenesis. The compaction phase correlates with cell neighbor exchange among the assembled pro-niche cells, as well as a burst of divisions among newly recruited stem cells. Before compaction, an assembly phase involves the movement of pro-niche cells along the outer periphery of the gonad, using the extracellular matrix (ECM) to assemble at the anterior of the gonad. Finally, live-imaging in integrin mutants allows definition of the role of pro-niche cell-ECM interaction with regard to the new assembly and compaction dynamics revealed in this study (Anllo, 2018).
This work is the first to characterize the dynamics of
testis niche morphogenesis. Through live-imaging both in vivo and ex
vivo, it was possible to follow pro-niche cells as they assemble and assume
the final compact form of this niche. After ex vivo culture, the niche
appears fully functional, as judged by enrichment of adhesion proteins,
the ability to activate STAT in adjacent germline cells, and to direct
the now GSCs to divide orthogonally to the niche, just as they would in
situ. Collectively, imaging has generated several novel insights on
niche morphogenesis that could not be inferred from fixed images alone.
The compaction phase correlates with cell neighbor exchange among the
assembled pro-niche cells, as well as a burst of divisions among newly
recruited stem cells. Before compaction occurs, an assembly phase
involves the movement of PS11 pro-niche cells along the outer periphery
of the gonad, using the ECM to assemble at the anterior of the gonad.
There they join PS10 pro-niche cells, forming an anterior and
internally-facing cap. Finally, live-imaging in integrin mutants allows
definition of the role of pro-niche cell-ECM interaction more precisely
with regard to the new assembly and compaction dynamics revealed in this study (Anllo, 2018).
Previously, analysis of fixed samples nicely inferred a change from an
initial cap of pro-hub cells to a compact niche. The current ex vivo imaging now visualizes compaction directly, and by
comparing cell numbers at assembly and after compaction, shows that the
initial assemblage of the elongate cap of cells indeed constitutes the
pro-hub. Scoring the same gonad before and after compaction showed
little, if any, shuffling in the number of pro-niche cells. This finding
is consistent with the fact that specification of niche cells occurs
much earlier in development, before the stages imaged in this study. Lastly, the current work suggests two cell biological phenomena that correlate with the process of
compaction (Anllo, 2018).
Firstly, while compaction is underway, there is a burst of GSC divisions
orthogonal to the hub. Interestingly, at late anaphase, the niche cells
adjacent to a dividing GSC occasionally appeared to become compressed
against their niche cell neighbors, subsequently occupying less space.
It is known that activation of the STAT pathway is necessary to induce
germline divisions in the (male) gonad. Since
the ligand for pathway activation is secreted by hub cells, this
suggests a possible scenario where niche cells orient GSC divisions,
which subsequently generate some of the force necessary for niche
compaction, pushing hub cells closer together. In principle, this scenario can now be tested by attempting to block GSC divisions during ex vivo culture (Anllo, 2018).
Secondly, hub compaction occurs concurrently with cell neighbor
exchanges. These exchanges are reminiscent of convergence and extension, where directed junctional re-organization changes the overall shape of the sheet of cells. For example, at assembly, an individual cell within
the elongate cap will shrink one of its boundaries, allowing cells that
were previously on either side of this cell to form new contacts
adhering to one another. Such events may impart an increasingly more
compact overall shape to the incipient niche. In addition to lateral
neighbor exchange, ingression was observed of some pro-niche cells during
compaction. This results in stratification of the niche, where a few
internal pro-hub cells now lie more internal to the superficial cap of
peripheral hub cells. These lateral and stratifying exchanges are
complex, but their comprehensive cataloging, now enabled by ex vivo
imaging, may well reveal patterns of cell movements. In turn, those
patterns will likely suggest specific circuits to test as regulators of
compaction (Anllo, 2018).
The neighbor exchange behavior associated with compaction is reminiscent
of cell sorting, where 'like' cells will maximize contacts among
themselves, while minimizing those with different neighbor types. Indeed, imaging describes a significant
smoothening of the boundary between the pro-niche and the tier of stem
cells, an outcome expected from sorting phenomena. Increased
accumulation of adhesion proteins such as ECadherin, NCadherin and
Fasciclin III, has been noted at the Assembly stage, although
manipulation of these did not appear to grossly affect compaction. While unidentified factors or significant
redundancy among these might drive sorting, the live-imaging parameters
established here now enable analysis of the dynamics of the process.
Future work might reveal changes in, for example, timing of compaction
and smoothening, rather than relying on endpoint analysis alone (Anllo, 2018).
The positioning of the niche at the gonad anterior could, in principle,
be driven by a gonad-intrinsic mechanism. For example, SGPs across the
gonad have different antero-posterior identities, as they originate from
three separate parasegments (ps10, ps11, and ps12). Thus, they accumulate Hox proteins to
differing levels, and AbdB mutant embryos can exhibit defects in hub
position. However, it remains possible
that these defects could be caused by broader patterning changes in
tissues outside the gonad, rather than reflecting an intrinsic
requirement for AbdB. Still, the fact that pro-hub cells are enriched
for adhesion proteins from the assembly stage onward is consistent with
the idea that sorting driven by homophilic adhesions could be a main
driver for niche formation. Nevertheless, two novel observations from live-imaging
strongly suggest the existence of an extra-gonadal influence (Anllo, 2018).
First, while the niche assembles at the anterior, it was found that it
exhibits a consistent tilt towards the interior of the embryo. This
biased placement of the hub towards internal organs suggests that an
extra-gonadal cue is influencing niche position. Live-imaging did
reveal tissues located near cells of the assembling pro-niche: a branch
of the trachea, and an alary muscle. The ablations that were conducted rule out
any major impact from these two tissues on niche formation. However,
other tissues, such as the nearby gut and its surrounding musculature,
remain to be tested (Anllo, 2018).
Second, the individual movements of PS11 cells to join the PS10 cohort
suggests that homotypic adhesions and sorting alone cannot explain the
position of assembly. During the stages leading up to those shown in this study,
SGPs exhibit robust cytoplasmic extensions that probe the internal
milieu of the newly formed gonad, tightly embracing each other, and
germline neighbors. Indeed,
imaging carried out in this study revealed PS11 cellular extensions around germ cells at the onset of assembly. If sorting were the main driver for assembly, these
extensions should be sufficient to cause PS11 cells to move through the
internal gonad milieu, seek out and coalesce with anteriorly located
PS10 SGPs. Instead, PS11 cells move out to the gonad periphery, and
retract their trailing extensions. They then move along the periphery
towards the anterior. In addition, if PS 11 cells were sorting together
one might have expected these pro-niche cells to move as one tight
collective, reminiscent of how the somatic border cell cluster migrates
through the internal milieu of the Drosophila egg chamber. In contrast, the PS11 cells often move individually or as a string of cells along the periphery. Taken
together, PS 11 cell behavior is much more in line with individual cell
migration directed by a cue (Anllo, 2018).
Previous work described an ECM visible at the gonad periphery at the end
of embryogenesis. Pigment cells ensheath the
gonad, and are the likely source of this ECM.
Indeed, live-imaging revealed that this ECM is present while PS11
cells are moving anteriorly. Thus, it is suspected that the matrix provides
traction for movement in response to the posited external guidance cue (Anllo, 2018).
In fact, previous work has suggested a role for ECM interaction, since
in the absence of Integrin function, the newly formed hub was located
inside the gonadal sphere (Tanentzapf, 2007). Since that work had
to rely on endpoint analysis, it was reasonable to assume that the hub
had fully formed at the gonad anterior and then was internalized whole
due to lack of Integrin-based adhesion. Instead, the current live-imaging of
Integrin mutants now reveals that Integrin-ECM interactions are
necessary well before compaction, as pro-niche cells can leave the
periphery of the gonad before the niche compacts. This is the case for
PS10 cells, which can be observed internalizing from the anterior of the
gonad before compaction, and for PS11 cells, either after they arrive at
the anterior or directly internalizing from their more posterior, PS11
location. Hub compaction occurs in Integrin mutants (Tanentzapf,
2007), but the current imaging reveals this to occur mostly after the niche
internalizes in the mutants. The observation that PS11 cells can
directly join an internalized cluster suggests that, left to their own,
pro-hub cells might sort to inappropriate locations due to preferential
adhesion. This implies that Integrin-ECM interactions at the periphery
antagonize the sorting tendency of pro-niche cells. If Integrin-ECM
interactions during migration are compromised, the resulting testis
would exhibit aberrant polarity, as the niche would not be localized at
the closed end of the tubule. Although some of the components making up the ECM
surrounding the gonad have been characterized, which of these engage Integrins on the pro-niche cells is not clear (Anllo, 2018).
The nature of the directional cue read by PS11
cells is not known. Imaging shows that some but not all PS11 cells successfully
arrive at the anterior in Integrin mutants. This suggests that pro-niche
cells have the capability to respond to the cue, at least partially,
even if they cannot interact with peripheral ECM. Normally, however,
since all the PS11 cells that were tracked in wild type gonads migrated to
the anterior along the ECM at the periphery of the gonad, these
Integrin-ECM interactions serve to ensure reproducible positioning of
the niche (Anllo, 2018).
The combination of in vivo and ex vivo imaging presented in this study overcomes
previous challenges to understanding how a stem cell niche is formed.
Visualizing the dynamics of niche formation has proven powerful in
suggesting unanticipated mechanisms for niche morphogenesis in this
work. This imaging should now allow for in-depth study of both the
process of assembly and compaction into a fully functioning stem cell
niche (Anllo, 2018).
Germ cell maturation is essential for spermatogenesis and testis homeostasis. ATP synthase serves significant roles in energy storage in germ cell survival and is catalyzed by alterations in the mitochondrial membrane proton concentration. The intrinsic cellular mechanisms governing stem cell maturation remain largely unknown. In the present study, in vivo RNA interference (RNAi) screening of major ATP synthase subunits was performed, and the function of ATP synthase for male fertility and spermatogenesis in Drosophila was explored. A Upstream Activation Sequence/Gal4 transcription factor system was used to knock down gene expression in specific cell types, and immunofluorescence staining was conducted to assess the roles of ATP synthase subunits in Drosophila testes. It was identified that knockdown of ATP synthase resulted in male infertility and abnormal spermatogenesis in Drosophila testes. In addition, knockdown of the ATP synthase beta subunit in germ cells resulted in defects in male infertility and germ cell maturation, while the hub and cyst cell populations were maintained. Other major ATP synthase subunits were also examined and similar phenotypes in Drosophila testes were identified. Taken together, the data from the present study revealed that ATP synthase serves important roles for male fertility during spermatogenesis by regulating germ cell maturation in Drosophila testes (Yu, 2019).
Stem cell niche is regulated by intrinsic and extrinsic factors. In the Drosophila testis, cyst stem cells (CySCs) support the differentiation of germline stem cells (GSCs). However, the underlying mechanisms remain unclear. This study found that somatic CG6015 is required for CySC maintenance and GSC differentiation in a Drosophila model. Knockdown of CG6015 in CySCs caused aberrant activation of dpERK in undifferentiated germ cells in the Drosophila testis, and disruption of key downstream targets of EGFR signaling (Dsor1 and rl) in CySCs results in a phenotype resembling that of CG6015 knockdown. CG6015, Dsor1, and rl are essential for the survival of Drosophila cell line Schneider 2 (S2) cells. The data showed that somatic CG6015 regulates CySC maintenance and GSC differentiation via EGFR signaling, and inhibits aberrant activation of germline dpERK signals. These findings indicate regulatory mechanisms of stem cell niche homeostasis in the Drosophila testis (Zheng, 2021).
Tissue-specific stem cells maintain tissue homeostasis by providing a continuous supply of differentiated cells throughout the life of organisms. Differentiated/differentiating cells can revert back to a stem cell identity via dedifferentiation to help maintain the stem cell pool beyond the lifetime of individual stem cells. Although dedifferentiation is important for maintaining the stem cell population, it is speculated that it underlies tumorigenesis. Therefore, this process must be tightly controlled. This study shows that a translational regulator, me31B, plays a critical role in preventing excess dedifferentiation in the Drosophila male germline: in the absence of me31B, spermatogonia dedifferentiate into germline stem cells (GSCs) at a dramatically elevated frequency. These results show that the excess dedifferentiation is likely due to misregulation of nos, a key regulator of germ cell identity and GSC maintenance. Taken together, the data reveal negative regulation of dedifferentiation to balance stem cell maintenance with differentiation (Jensen, 2021).
Stem cells have the potential to maintain undifferentiated state and differentiate into specialized cell types. Despite numerous progress has been achieved in understanding stem cell self-renewal and differentiation, many fundamental questions remain unanswered. In this study, dRTEL1, the Drosophila homolog of Regulator of Telomere Elongation Helicase 1 (CG4078), was identified as a novel regulator of male germline stem cells (GSCs). Genome-wide transcriptome analysis and ChIP-Seq results suggest that dRTEL1 affects a set of candidate genes required for GSC maintenance, likely independent of its role in DNA repair. Furthermore, dRTEL1 prevents DNA damage-induced checkpoint activation in GSCs. Finally, dRTEL1 functions to sustain Stat92E protein levels, the key player in GSC maintenance. Together, these findings reveal an intrinsic role of the DNA helicase dRTEL1 in maintaining male GSC and provide insight into the function of dRTEL1 (Yang, 2021).
In Drosophila, three types of UAS vectors (UASt, UASp, and UASz) are currently available for use with the Gal4-UAS system. They have been used successfully in somatic cells and germline cells from ovaries. However, it remains unclear whether they are functional in the germline cells of embryos, larvae, and adult testes. This study found that all three types of UAS vectors were functional in the germline cells of embryos and larvae and that the UASt and UASz vectors were active in the germline of the distal tip region in adult testes. Moreover, it was observed that protein expression from the UAS vectors was male-biased in germline cells of late embryos, whereas their respective mRNA expression levels were not. Furthermore, O-propargyl-puromycin (OPP) staining revealed that protein synthesis was male-biased in these germline cells. In addition, GO terms related to translation and ribosomal maturation were significantly enriched in the male germline. These observations show that translational activity is higher in male than in female germline cells. Therefore, it is proposed that male-biased protein synthesis may be responsible for the sex differences observed in the early germline (Masukawa, 2021).
The Drosophila GAGA-factor encoded by the Trithorax-like (Trl) gene is DNA-binding protein with unusually wide range of applications in diverse cell contexts. In Drosophila spermatogenesis, reduced GAGA expression caused by Trl mutations induces mass autophagy leading to germ cell death. This work investigated the contribution of mitochondrial abnormalities to autophagic germ cell death in Trl gene mutants. Using a cytological approach, in combination with an analysis of high-throughput RNA sequencing (RNA-seq) data, it was demonstrated that the GAGA deficiency led to considerable defects in mitochondrial ultrastructure, by causing misregulation of GAGA target genes encoding essential components of mitochondrial molecular machinery. Mitochondrial anomalies induced excessive production of reactive oxygen species and their release into the cytoplasm, thereby provoking oxidative stress. Changes in transcription levels of some GAGA-independent genes in the Trl mutants indicated that testis cells experience ATP deficiency and metabolic aberrations, that may trigger extensive autophagy progressing to cell death (Dorogova, 2023).
This study provides a single nucleus and single cell RNA-seq resource covering all of spermatogenesis in Drosophila starting from in-depth analysis of adult testis single nucleus RNA-seq (snRNA-seq) data from the Fly Cell Atlas (FCA) study. With over 44,000 nuclei and 6000 cells analyzed, the data provide identification of rare cell types, mapping of intermediate steps in differentiation, and the potential to identify new factors impacting fertility or controlling differentiation of germline and supporting somatic cells. Assignment of key germline and somatic cell types was assigned using combinations of known markers, in situ hybridization, and analysis of extant protein traps. Comparison of single cell and single nucleus datasets proved particularly revealing of dynamic developmental transitions in germline differentiation. To complement the web-based portals for data analysis hosted by the FCA, datasets compatible with commonly used software such as Seurat and Monocle is provided. The foundation provided in this study will enable communities studying spermatogenesis to interrogate the datasets to identify candidate genes to test for function in vivo (Raz, 2023).
A unifying feature of polycystin-2 channels is their localization to both primary and motile cilia/flagella. In Drosophila melanogaster, the fly polycystin-2 homologue, Amo, is an ER protein early in sperm development but the protein must ultimately cluster at the flagellar tip in mature sperm to be fully functional. Male flies lacking appropriate Amo localization are sterile due to abnormal sperm motility and failure of sperm storage. A forward genetic screen was performed to identify additional proteins that mediate ciliary trafficking of Amo. Drosophila homologues of KPC1 and KPC2, which comprise the mammalian KIP1 ubiquitination-promoting complex (KPC), form a conserved unit that is required for the sperm tail tip localization of Amo. Male flies lacking either KPC1 or KPC2 phenocopy amo mutants and are sterile due to a failure of sperm storage. KPC is a heterodimer composed of KPC1, an E3 ligase, and KPC2 (or UBAC1), an adaptor protein. Like their mammalian counterparts Drosophila KPC1 and KPC2 physically interact and they stabilize one another at the protein level. In flies, KPC2 is monoubiquitinated and phosphorylated and this modified form of the protein is located in mature sperm. Neither KPC1 nor KPC2 directly interact with Amo but they are detected in proximity to Amo at the tip of the sperm flagellum. In summary this study has identified a new complex that is involved in male fertility in Drosophila melanogaster (Li 2020).
Tissue homeostasis and repair relies on proper communication of stem cells and their differentiating daughters with the local tissue microenvironment. In the Drosophila male germline adult stem cell lineage, germ cells proliferate and progressively differentiate enclosed in supportive somatic cyst cells, forming a small organoid, the functional unit of differentiation. This study show that cell polarity and vesicle trafficking influence signal transduction in cyst cells, with profound effects on the germ cells they enclose. The data suggest that the cortical components Dlg, Scrib, Lgl and the clathrin-mediated endocytic (CME) machinery downregulate epidermal growth factor receptor (EGFR) signaling. Knockdown of dlg, scrib, lgl, or CME components in cyst cells resulted in germ cell death, similar to increased signal transduction via the EGFR, while lowering EGFR or downstream signaling components rescued the defects. This work provides insights into how cell polarity and endocytosis cooperate to regulate signal transduction and sculpt developing tissues (Papagiannouli, 2019).
The capacity of stem cells to self-renew or differentiate has been attributed to distinct metabolic states. A genetic screen targeting regulators of mitochondrial dynamics revealed that mitochondrial fusion is required for the maintenance of male germline stem cells (GSCs) in Drosophila melanogaster. Depletion of Mitofusin (dMfn) or Opa1 led to dysfunctional mitochondria, activation of Target of rapamycin (TOR) and a marked accumulation of lipid droplets. Enhancement of lipid utilization by the mitochondria attenuated TOR activation and rescued the loss of GSCs that was caused by inhibition of mitochondrial fusion. Moreover, constitutive activation of the TOR-pathway target and lipogenesis factor Sterol regulatory element binding protein (SREBP) also resulted in GSC loss, whereas inhibition of SREBP rescued GSC loss triggered by depletion of dMfn. These findings highlight a critical role for mitochondrial fusion and lipid homeostasis in GSC maintenance, providing insight into the potential impact of mitochondrial and metabolic diseases on the function of stem and/or germ cells (Senos, 2019).
The Drosophila male stem cell niche is a well characterized structure in which a small cluster of somatic cells send self-renewal signals to neighbouring germ cells. Although the molecular information involved in the stem cell fate have been identified, much less is understand on the mechanisms driving their short-range specific release. Ultrastructural analysis reveals distinct protrusions of the stem cell plasma membrane that interdigitate with membrane protrusions of the facing hub cells. Some of these protrusions are very elongated and extend into the hub and could correspond to the Mt-Nanotubes. Therefore, the interface between the stem cells and the hub appears more complex than previously reported and the membrane protrusions of the stem cells might represent specialized surface areas involved in the niche-stem cell communication. The presence was noticed of clathrin-coated vesicles in the germline plasma membrane that might be also involved in delivering information from the hub (Persico, 2019).
Hub cells comprise a niche for germline stem cells and cyst stem cells in the Drosophila testis. Hub cells arise from common somatic gonadal precursors in embryos, but the mechanism of their specification is still poorly understood. This study found that RNA binding proteins Lin28 and Imp mediate transcript stability of Bowl, a known hub specification factor; Bowl transcripts were reduced in the testis of Lin28 and Imp mutants, and also when RNA-mediated interference against Lin28 or Imp was expressed in hub cells. In tissue culture Luciferase assays involving the Bowl 3'UTR, stability of Luc reporter transcripts depended on the Bowl 3'UTR and required Lin28 and Imp. These findings suggest that proper Bowl function during hub cell specification requires Lin28 and Imp in the testis hub cells (To, 2021).
The Drosophila male germline stem cell (GSC) lineage provides a great model to understand stem cell maintenance, proliferation, differentiation, and dedifferentiation. This study used Drosophila GSC lineage to systematically analyze transcriptome of discrete but continuous differentiating germline cysts. First single cysts were isolated at each recognizable stage from wild-type testes, which were subsequently applied for RNA-seq analyses. The data delineate a high-resolution transcriptome atlas in the entire male GSC lineage: The most dramatic switch occurs from early to late spermatocyte, followed by the change from the mitotic spermatogonia to early meiotic spermatocyte. By contrast, the transit-amplifying spermatogonia cysts display similar transcriptomes, suggesting common molecular features among these stages, which may underlie their similar behavior during both differentiation and dedifferentiation processes. Finally, distinct differentiating germ cell cyst samples do not exhibit obvious dosage compensation of X-chromosomal genes, even with consideration of the paucity of X-chromosomal gene expression during meiosis, which is different from somatic cells. Together, the single cyst-resolution, genome-wide transcriptional profile analyses provide an unprecedented resource to understand many questions in both germ cell biology and stem cell biology fields (Shi, 2020).
Maintenance of a multicellular organism during homeostasis and tissue repair requires active replenishment of depleted or injured cells during aging and regeneration. Adult stem cells fulfill this requirement owing to their unique ability to both self-renew and give rise to differentiated special cell types. In many adult stem cell lineages, progenitor cells often undergo a proliferative stage to expand their population before commitment for terminal differentiation. The switch from proliferation to differentiation must be tightly regulated, because mis-regulation of this transition might lead to tumorigenesis or tissue dystrophy. On the other hand, progenitor cells remain plastic and can dedifferentiate to become stem cell-like cells in multiple stem cell lineages. In order to properly differentiate adult stem cells or progenitor cells in vitro and/or to promote dedifferentiation in vivo for regenerative medicine, it is necessary to fully understand the molecular changes underlying the normal differentiation program of adult stem cells in vivo (Shi, 2020).
Drosophila spermatogenesis provides a great model system to study mechanisms that regulate the maintenance, proliferation and proper differentiation of adult stem cells. In adult testes of Drosophila melanogaster, germline stem cells (GSCs) can be precisely located by their proximity and attachment to a group of post-mitotic somatic cells called hub cells at the apical tip of testis. A GSC typically divides asymmetrically to self-renew and to give rise to a gonialblast (GB), the daughter cell that initiates differentiation. GBs first go through a transit-amplifying stage as spermatogonial cells, for which they undergo exactly four rounds of mitosis in D. melanogaster. Once this stage is complete, germ cells enter the spermatocyte stage, which is an elongated G2 phase of meiosis I. During this stage, each spermatocyte grows ~25-fold in volume, which involves a robust gene expression program allowing for meiotic divisions and spermatid terminal differentiation program. In parallel with GSC asymmetric cell division, the cyst stem cell (CySC), two of them surrounding each GSC, also divides asymmetrically, resulting in one daughter cell retaining CySC identity and the other daughter cell becoming a differentiated cyst cell. Two cyst cells encapsulate the differentiating germ cells throughout the entire spermatogenesis and they never divide again. It has been demonstrated that CySCs and cyst cells communicate with their accompanying germ cells via multiple signaling pathways for maintaining germ cell fate and regulating proper germline proliferation and differentiation throughout spermatogenesis. During this process, dynamic changes in the gene expression program are orchestrated by both extrinsic cues, such as paracrine factors that trigger signaling pathways, and intrinsic factors, such as chromatin regulators (Shi, 2020).
Previous studies have attempted to parse the transcriptional networks underlying GSC differentiation by comparing gene expression profiles of mutant testes that accumulate germ cells at distinct cellular differentiation stages with wild-type testes. Although these approaches have led to many interesting functional studies using a candidate gene approach, the information gleaned from intact tissues is limited by the inherently mixed population of cells, and the difficulty in extrapolating results obtained from mutant backgrounds to normal situation in wild-type tissue. This report systematically studied the gene expression profile of the Drosophila male GSC lineage at every recognizable and isolatable stage. Using this dataset, A the following questions were addressed: Do GSCs and GBs, the two daughter cells derived from GSC asymmetric division, have similar or distinct transcriptional profiles? How does the transcriptome change in continuously proliferating spermatogonial cells? Is the switch from mitosis to meiosis accompanied by a transcriptome change that leads to another transcriptome change during spermatocyte maturation? Does dosage compensation occur in germ cells? In summary, the single-cyst transcriptome profiles provide a comprehensive dataset at a resolution that has not been achieved before, which yields much-needed information on transcriptional status at each crucial stage from an endogenous stem cell system. Researchers from both germ cell biology and stem cell biology fields should benefit from using this resource to screen for genes with a particular expression pattern or examine genes from specific pathway(s), before designing detailed functional analyses (Shi, 2020).
The single cyst-resolution, genome-wide transcriptional profile analyses provide a data resource at each crucial stage of differentiation in an adult stem cell system under their wild-type situation, which will help in understanding many interesting issues in stem cell and germ cell biology fields. Previously, RNA-seq analyses were carried out using the entire testes with genetic mutations that arrest cellular differentiation pathways at distinct stages or dissected different regions of wild-type testes enriched with distinct stages of germ cells. However, caveats arise from mixed cell types within either intact or dissected tissues, as well as complications using genetic mutations. Furthermore, although single cell RNA-seq could be currently carried out using dissociated single cells, the relatively small number of early-stage cells, including germline stem cells, make it very challenging to obtain their transcriptomes. Single post-meiotic cysts have been isolated and applied for quantitative RT-PCR, which has revealed two dozen transcribed genes in spermatids. In this study, for the TASC experiments, single germline cysts were isolated from niche to late spermatocyte stages using cell type- and stage-specific markers, which allowed precise reconstruction of the single-cyst transcriptome atlas during the entire spermatogenesis process. Therefore, both the original technologies that were developed and the new knowledge gained in this study represent technical and conceptual advances, which will have a broad and long-term impact on studies in the stem cell and germ cell biology fields (Shi, 2020).
As a proof of principle, several new discoveries have already been made using the TASC dataset. For example, a candidate gene called slamdance (sda) was identified that has enriched transcription in the niche sample. The sda gene encodes an aminopeptidase, the function of which had never been reported in any adult stem cell systems. Extensive functional analyses was carried out and it was demonstrated that the Sda protein indeed acts as an aminopeptidase, the in vivo function of which depends on its enzymatic domain. Sda plays important roles in regulating GSC maintenance and progenitor germline dedifferentiation. In addition, it was found that the known germline differentiation factor bam has prolonged transcription in early-stage spermatocytes , even though Bam protein is downregulated. This difference between mRNA expression pattern and protein enrichment indicated a potential post-transcriptional regulation of the bam transcript, which was revealed later through recognition of the 3' untranslated region of bam by microRNAs. Therefore, the TASC dataset can be used to identify genes with interesting expression patterns for in-depth functional analysis, or combined with other data such as protein expression patterns for studying post-transcriptional regulation in the germline, etc. Together, these studies will contribute to better understanding of gene regulation during spermatogenesis, which will also significantly enhance knowledge of reproductive biology (Shi, 2020).
In summary, this single cyst-resolution, genome-wide transcriptional profile analysis provides a supreme and unprecedented data resource at each crucial stage of differentiation in an adult stem cell system under its physiological conditions, which will help in understanding many issues in germ cell biology and stem cell biology fields. The new technologies that were developed and the knowledge gained in this study will have broad and high impact on basic research as well as on regenerative medicine (Shi, 2020).
Adult stem cells divide to renew the stem cell pool and replenish specialized cells that are lost due to death or usage. However, little is known about the mechanisms regulating how stem cells adjust to a demand for specialized cells. A failure of the stem cells to respond to this demand can have serious consequences, such as tissue loss, or prolonged recovery post injury. This study challenged the male germline stem cells (GSCs) of Drosophila melanogaster for the production of specialized cells, sperm cells, using mating experiments. Repeated mating reduced the sperm pool and increased the percentage of GSCs in M- and S-phase of the cell cycle. The increase in dividing GSCs depended on the activity of the highly conserved G-proteins. Germline expression of RNA-Interference (RNA-i) constructs against G-proteins, or a dominant negative G-protein eliminated the increase in GSC division frequency in mated males. Consistent with a role for the G-proteins in regulating GSC division frequency, RNA-i against seven out of 35 G-protein coupled receptors (GPCRs) within the germline cells also eliminated the capability of males to increase the numbers of dividing GSCs in response to mating (Malpe, 2020).
This study shows that repeated mating reduced the sperm pool and increased GSC division frequency. Using highly controlled experiments, it was demonstrated that mated males had more GSCs in M-phase and S-phase of the cell cycle compared to non-mated males. Mated males also showed faster incorporation of EdU in feeding experiments, suggesting that the GSCs of mated males entered the S-phase of the cell cycle more frequently. Though the possibility that mated males ingested more EdU-supplemented food than their non-mated siblings cannot be excluded, the data suggest that the GSCs in mated males cycle faster. The response curve obtained in a time-course experiment is different from the response curves reported by other groups that used bromo-deoxy-uridine (BrDU) as the thymidine analog instead of EdU. For example, the non-mated males in the current experiment had about 70% of EdU-positive GSCs after 48 hours of feeding. A study that uses white (w) mutant animals and fed the same concentration of the thymidine homologue had a steeper response curve, in which 85% of the GSCs were BrDU-marked after 48 hours of feeding. Another study using y, v flies showed even steeper response curves where 100% of the GSCs were BrDU-labeled after 24 hours. However, in this study, animals were fed a 30 times higher concentration of the thymidine homologue than used in the current study. It is proposed that the different response curves are due to the different genetic backgrounds, chemicals, and doses (Malpe, 2020).
These findings demonstrate that GSCs can respond to a demand for sperm by increasing their mitotic activity. Based on RNA-i targeting G-proteins and a dominant negative construct against Gγ1, the increase in MIGSC of mated males is dependent on G-protein signaling. Furthermore, signal transducers predicted to act downstream of G-proteins and GPCRs predicted to act upstream of G-proteins also appeared to be required for the response to mating. Whether G-protein signaling directly affects GSC division frequency, or whether G-protein-dependent communication among the early stage germline cells impacts the MIGSC remains to be investigated (Malpe, 2020).
Due to the lack of mutants and a potential interference of whole animal knock-down in the behavior of the flies, tissue-specific expression of RNA-i-constructs was used. It is surprising that these studies revealed potential roles for seven instead of a single GPCR in the increase of MIGSC in response to mating. A possible explanation is that some of the RNA-i-lines have off-target effects. RNA-i-hairpins can cause the down-regulation of unintended targets due to stretches of sequence homologies, especially when long hairpins are used. However, with the exception of the RNA-i-line directed against 5-HT7, all lines that produced a phenotype contain second generation vectors with a short, 21 nucleotide hairpin predicted to have no off-target effects. Thus, it is hypothesized that multiple GPCRs regulate the increase in MIGSC in response to mating. Consistent with this, expression of other RNA-i-lines directed against Mth or 5-HT1A interfered with the increase in MIGSC in mated males (Malpe, 2020).
The finding that RNA-i against several GPCRs blocked the increase in MIGSC in mated males suggests a high level of complexity in the regulation of GSC division frequency. One simple explanation could be that the increase in GSC division frequency is dependent on ideal physiological conditions and that lack of any of the seven GPCRs somehow impairs the cell's normal metabolism. Alternatively, the GPCRs could act in concert to impact the MIGSC. In the literature, increasing evidence has emerged that GPCRs can form dimers and oligomers and that these have a variety of functional roles, ranging from GPCR trafficking to modification of G-protein mediated signaling. In C. elegans, two Octopamine receptors, SER-3 and SER-6, additively regulate the same signal transducers for food-deprived-mediated signaling. One possible explanation for the non-redundant function of the two receptors was the idea that they form a functional dimer. In mammalian cells, 5-HT receptors can form homo-dimers and hetero-dimers and, dependent on this, have different effects on G-protein signaling. In cultured fibroblast cells, for example, G-protein coupling is more efficient when both receptors within a 5-HT4 homo-dimer bind to agonist instead of only one. In cultured hippocampal neurons, hetero-dimerization of 5-HT1A with 5-HT7 reduces G-protein activation and decreases the opening of a potassium channel compared to 5-HT1A homo-dimers59. The formation of hetero-dimers of GPCRs with other types of receptors plays a role in depression and in the response to hallucinogens in rodents (Malpe, 2020).
Alternatively, or in addition to the possibility that some or all of the seven GPCRs form physical complexes, a role for several distinct GPCRs in regulating GSC division frequency could be explained by cross-talk among the downstream signaling cascades. One signaling cascade could, for example, lead to the expression of a kinase that is activated by another cascade. Similarly, one signaling cascade could open an ion channel necessary for the activity of a protein within another cascade. Unfortunately, the literature provides little information on Drosophila GPCR signal transduction cascades and only very few mutants have been identified that affect a process downstream of GPCR stimulation. Thus, it remains to be explored how stimulation of the GPCRs and G-proteins increase GSC divisions (Malpe, 2020).
The role for G-protein signaling in regulating the frequency of GSC divisions is novel. The data suggest that the increase in MIGSC in response to mating is regulated by external signals, potentially arising from the nervous system, that stimulate G-protein signaling in the germline. Based on the nature of the GPCRs, the activating signal could be Serotonin, the Mth ligand, Stunted, Octopamine, or two other, yet unknown, signals that activate Mth-l5, and CG12290. It will be interesting to address which of these ligands are sufficient to increase MIGSC, in what concentrations they act, by which tissues they are released, and whether they also affect other stem cell populations (Malpe, 2020).
Stem cell activity and cell differentiation is robustly influenced by the nutrient availability in the gonads. The signal that connects nutrient availability to gonadal stem cell activity remains largely unknown. This study shows that tumor necrosis factor Eiger (Egr) is upregulated in testicular smooth muscles as a response to prolonged protein starvation in Drosophila testis. While Egr is not essential for starvation-induced changes in germline and somatic stem cell numbers, Egr and its receptor Grindelwald influence the recovery dynamics of somatic cyst stem cells (CySCs) upon protein refeeding. Moreover, Egr is also involved in the refeeding-induced, ectopic expression of the CySC self-renewal protein and the accumulation of early germ cells. Egr primarily acts through the Jun N-terminal kinase (JNK) signaling in Drosophila. This study shows that inhibition of JNK signaling in cyst cells suppresses the refeeding-induced abnormality in both somatic and germ cells. In conclusion, this study reveals both beneficial and detrimental effects of Egr upregulation in the recovery of stem cells and spermatogenesis from prolonged protein starvation (Chang, 2020).
The stem cell niche regulates the renewal and differentiation of germline stem cells (GSCs) in Drosophila. Previous work has identified a series of genes encoding ribosomal proteins that may contribute to the self-renewal and differentiation of GSCs. However, the mechanisms that maintain and differentiate GSCs in their niches were not well understood. Flies were used to generate tissue-specific gene knockdown. Small interfering RNAs were used to knockdown genes in S2 cells. qRT-PCR was used to examine the relative mRNA expression level. TUNEL staining or flow cytometry assays were used to detect cell survival. Immunofluorescence was used to determine protein localization and expression pattern. Using a genetic manipulation approach, this study investigated the role of ribosomal protein S13 (RpS13) in testes and S2 cells. RpS13 was shown to be required for the self-renewal and differentiation of GSCs. RpS13 regulates cell proliferation and apoptosis. Mechanistically, RpS13 regulates the expression of ribosome subunits and could moderate the expression of the Rho1, DE-cad and Arm proteins. Notably, Rho1 imitated the phenotype of RpS13 in both Drosophila testes and S2 cells, and recruited cell adhesions, which was mediated by the DE-cad and Arm proteins. These findings uncover a novel mechanism of RpS13 that mediates Rho1 signals in the stem cell niche of the Drosophila testis (Wang, 2020).
CG6015 controls spermatogonia transit-amplifying divisions by epidermal growth factor receptor signaling in Drosophila testes
Spermatogonia transit-amplifying (TA) divisions are crucial for the differentiation of germline stem cell daughters. However, the underlying mechanism is largely unknown. The present study demonstrated that CG6015 was essential for spermatogonia TA-divisions and elongated spermatozoon development in Drosophila melanogaster. Spermatogonia deficient in CG6015 inhibited germline differentiation leading to the accumulation of undifferentiated cell populations. Transcriptome profiling using RNA sequencing indicated that CG6015 was involved in spermatogenesis, spermatid differentiation, and metabolic processes. Gene Set Enrichment Analysis (GSEA) revealed the relationship between CG6015 and the epidermal growth factor receptor (EGFR) signaling pathway. Unexpectedly, it was discovered that phosphorylated extracellular regulated kinase (dpERK) signals were activated in germline stem cell (GSC)-like cells after reduction of CG6015 in spermatogonia. Moreover, Downstream of raf1 (Dsor1), a key downstream target of EGFR, mimicked the phenotype of CG6015, and germline dpERK signals were activated in spermatogonia of Dsor1 RNAi testes. Together, these findings revealed a potential regulatory mechanism of CG6015 via EGFR signaling during spermatogonia TA-divisions in Drosophila testes (Yu, 2021).
Stem-cell niche signaling is short-range in nature, such that only stem cells but not their differentiating progeny receive self-renewing signals. At the apical tip of the Drosophila testis, 8 to 10 germline stem cells (GSCs) surround the hub, a cluster of somatic cells that organize the stem-cell niche. Previous work has shown that GSCs form microtubule-based nanotubes (MT-nanotubes) that project into the hub cells, serving as the platform for niche signal reception; this spatial arrangement ensures the reception of the niche signal specifically by stem cells but not by differentiating cells. The receptor Thickveins (Tkv) is expressed by GSCs and localizes to the surface of MT-nanotubes, where it receives the hub-derived ligand Decapentaplegic (Dpp). The fate of Tkv receptor after engaging in signaling on the MT-nanotubes has been unclear. This study demonstrates that the Tkv receptor is internalized into hub cells from the MT-nanotube surface and subsequently degraded in the hub cell lysosomes. Perturbation of MT-nanotube formation and Tkv internalization from MT-nanotubes into hub cells both resulted in an overabundance of Tkv protein in GSCs and hyperactivation of a downstream signal, suggesting that the MT-nanotubes also serve a second purpose to dampen the niche signaling. Together, these results demonstrate that MT-nanotubes play dual roles to ensure the short-range nature of niche signaling by (1) providing an exclusive interface for the niche ligand-receptor interaction; and (2) limiting the amount of stem cell receptors available for niche signal reception (Ladyzhets, 2020).
The homeostasis of the stem cell niche is regulated by both intrinsic and extrinsic factors, and the complex and ordered molecular and cellular regulatory mechanisms need to be further explored. In Drosophila testis, germline stem cells (GSCs) rely on hub cells for self-renewal and physical attachment. GSCs are also in contact with somatic cyst stem cells (CySCs). Utilizing genetic manipulation in Drosophila, this study investigated the role of Wnt6 in vivo and in vitro. In Drosophila testis, Wnt6 was found to be required for GSC differentiation and CySC self-renewal. In Schneider 2 (S2) cells, Wnt6 was found to regulate cell proliferation and apoptosis. Mechanistically, it was demonstrated that Wnt6 can downregulate the expression levels of Arm, Rac1 and Cdc42 in S2 cells. Notably, Rac1 and Cdc42, which act downstream of the noncanonical Wnt signalling pathway, imitated the phenotypes of Wnt6 in Drosophila testis. Thus, the newly discovered Wnt6-Rac1/Cdc42 signal axis is required for the homeostasis of the stem cell niche in the Drosophila testis (Wang, 2021).
The global rise in obesity has revitalized a search for genetic and epigenetic factors underlying the disease. This study presents a Drosophila model of paternal-diet-induced intergenerational metabolic reprogramming (IGMR) and identifies genes required for its encoding in offspring. Intriguingly, as little as 2 days of dietary intervention in fathers elicits obesity in offspring. Paternal sugar acts as a physiological suppressor of variegation, desilencing chromatin-state-defined domains in both mature sperm and in offspring embryos. Requirements were identified for H3K9/K27me3-dependent reprogramming of metabolic genes in two distinct germline and zygotic windows. Critically, evidence is found that a similar system may regulate obesity susceptibility and phenotype variation in mice and humans. The findings provide insight into the mechanisms underlying intergenerational metabolic reprogramming and carry profound implications for understanding of phenotypic variation and evolution (Ost, 2014).
This study shows that acute dietary interventions, as short as 24 hr, have the capacity to modify F1 offspring phenotype via the male germline. Reprogramming occurs in response to dietary manipulations over a physiological range, and phenotypic outcomes require polycomb- and H3K9me3-centric plasticity in spatially and chromatin-state-defined regions of the genome. The eye color shifts in wm4 h offspring and the reduced fat body H3K9me3 staining in adult IGMR offspring supports the conclusions, first, that there are chromatin state changes and, second, that these are stable lifelong. These data are corroborated by selective derepression of Su(var)3-9, SETBD1, Su(var)4-20, and polycomb-sensitive transcripts; chromatin-state-associated transcriptional rearrangements genome wide; selective reprogramming of highly dynamic histone-mark-defined regions; and the fact that intergenerational metabolic reprogramming (IGMR ) itself is sensitive to a string of distinct H3K9me3-centric and polycomb mutants. Although nontrivial, ChIP-seq comparisons of repressive chromatin architecture in mature sperm and multiple defined offspring tissues will be important to establishing the ubiquitousness of these regulatory events and the nature of intergenerational signal itself. These data highlight how acutely sensitive intergenerational control can be to even normal physiological changes, and they identify some of the first genes absolutely required for transmission evolution (Ost, 2014).
First categorized simply as heterochromatin versus euchromatin, multiple empirical models now divide the genome into 5 to 51 chromatin states, depending on the analysis. Paternal high sugar increases gene expression preferentially of heterochromatic-embedded genes in embryos. Specifically, these genes are characterized by active deposition of H3K9me3 and H3K27me3, by long distance from class I insulators, and by sensitivity to fully intact expression of Su(var)3-9, Su(var)4-20, SetDB1, Pc, and E(z) . The data support a model where phenotype has been evolutionarily encoded directly into the chromatin state of relevant loci. Specifically, an abundance of genes important to both cytosolic and mitochondrial metabolism appear to be embedded into H3K9me3- and distinct polycomb-dependent control regions. Indeed, GO analysis of the five chromatin colors indicate a largely mutually exclusive picture, in which functional pathways are not randomly distributed across chromatin states. The paternal IGMR data set revealed clear and strong overlaps with pathways of black (lamin-associated) and blue (polycomb) chromatin and included many key metabolic pathways, including glycolysis, TCA cycle, mitochondrial OxPhos, chitin, and polysaccharide metabolism, changes that could well prime the system for altered functionality given the appropriate stimulus. Indeed, paternal IGMR phenotype is a susceptibility to diet-induced obesity and is most readily observable upon high-sugar diet challenge evolution (Ost, 2014).
The data support a trans-acting mechanism. In the wm4h experiments, male offspring inherited their X chromosome and thus the reporter from their unchallenged mothers, i.e., the reporter allele never encounters the initial signal but is reproducibly reprogrammed. Further, the failure of Su(var) 4-20SP and Despite their genetic similarity, isogenic or congenic animals reared under controlled conditions exhibit measurable variation in essentially all phenotypes. Such variability in genome output is thought to arise largely from probabilistic or chance developmental events in early. This study mapped a mechanism that couples acute paternal feeding and zygotic chromatin state integrity directly to phenotypic output of the next generation. These same signatures predict obesity susceptibility in isogenic mouse and human obesity cohorts. Because acute circadian fluctuations in feeding are essentially constant over evolutionary timescales, they are the perfect mechanistic input upon which a system could evolve to ensure defined phenotypic variation within a given population evolution (Ost, 2014).
Sperm-packaged DNA must undergo extensive reorganization to ensure its timely participation in embryonic mitosis. Whereas maternal control over this remodeling is well described, paternal contributions are virtually unknown. This study shows that Drosophila melanogaster males lacking Heterochromatin Protein 1E (HP1E) sire inviable embryos that undergo catastrophic mitosis. In these embryos, the paternal genome fails to condense and resolve into sister chromatids in synchrony with the maternal genome. This delay leads to a failure of paternal chromosomes, particularly the heterochromatin-rich sex chromosomes, to separate on the first mitotic spindle. Remarkably, HP1E is not inherited on mature sperm chromatin. Instead, HP1E primes paternal chromosomes during spermatogenesis to ensure faithful segregation post-fertilization. This transgenerational effect suggests that maternal control is necessary but not sufficient for transforming sperm DNA into a mitotically competent pronucleus. Instead, paternal action during spermiogenesis exerts post-fertilization control to ensure faithful chromosome segregation in the embryo (Levine, 2015).
Faithful chromosome segregation requires careful orchestration of chromosomal condensation, alignment, and movement of mitotic chromosomes during every eukaryotic cell division. The very first embryonic mitosis in animals requires additional synchronization. Paternally and maternally inherited genomes undergo independent chromatin reorganization and replication prior to mitotic entry. For instance, maternal chromosomes must complete meiosis and then transition from a meiotic conformation to an interphase-like state in preparation for replication. The sperm-deposited, paternal chromosomes must undergo an even more radical transition from a highly compact, protamine-rich state to a decondensed, histone-rich state before DNA replication. Despite these divergent requirements to achieve replication- and mitotic-competency, maternal and paternal genomes synchronously enter the first mitosis. Failure to carry out paternal chromosome remodeling in a timely fashion results in paternal genome loss and embryonic inviability (Levine, 2015).
The transition from a protamine-rich sperm nucleus to a competent paternal pronucleus requires the action of numerous maternally deposited proteins in the egg. For instance, paternal genome decondensation post-fertilization requires the integration of histone H3.3, a histone variant deposited by the maternal proteins HIRA, CHD1, and Yemanuclein. Similarly, maternally-deposited MH/Spartan protein localizes exclusively to the replicating paternal genome and is required for faithful paternal chromosome segregation during the first embryonic division. These and other studies demonstrate the essential role of maternally-deposited machinery in rendering competent sperm-deposited DNA and ultimately, ensuring faithful paternal genome inheritance (Levine, 2015).
Is paternal control also necessary for the extensive decondensation and re-condensation of the post-fertilization paternal genome? If so, disruption of such control would manifest as paternal effect lethality (PEL). Unlike male sterility mutants that lack motile sperm, PEL mutants make abundant motile sperm that fertilize eggs efficiently. However, embryos 'fathered' by PEL mutants are inviable. Only a handful of PEL genes have been characterized in animals. These encode proteins that mediate sperm release of paternal DNA, sperm centriole inheritance, and paternal chromosome segregation. Only one of these PEL proteins directly localizes to paternal chromosomes; the sperm-inherited K81 protein localizes exclusively to paternal chromosome termini and ensures telomere integrity. The maintenance of telomeric epigenetic identity joins a growing list of examples of sperm-to-embryo information transmission via protein or RNA inheritance (e.g., diet, stress, embryonic patterning, transcriptional competency). Despite a new appreciation of paternal control over epigenetic information transfer, there are no reports of paternal control over the global chromatin reorganization required for synchronous mitosis across paternally and maternally inherited genomes. Indeed, in the absence of any known paternal protein-directed genome remodeling, a model has emerged that maternal proteins might be sufficient for transforming tightly packaged sperm DNA into a fully competent paternal pronucleus (Levine, 2015).
The notion that maternal control is sufficient to accomplish paternal genome remodeling is challenged by recent findings from the intracellular Wolbachia bacterium that infects more than 50% of insect species. Wolbachia-infected Drosophila males mated to uninfected females father embryos that arrest soon after the first zygotic mitosis. Embryonic arrest occurs because paternal genomes enter the first mitosis with unresolved sister chromatids that fail to separate on the mitotic spindle. Although the identity of the host factor(s) manipulated by Wolbachia to mediate this transgenerational effect is still unknown, what is clear is that pre-fertilization, Wolbachia subverts the paternal germline machinery that helps direct global genome remodeling of paternal chromosomes in the embryo. Wolbachia action during spermiogenesis leads to paternal-maternal genome asynchrony and ultimately, failure of paternal chromosomes to separate on the first mitotic spindle. Despite decades of interest, the molecular basis of paternal control has remained elusive (Levine, 2015).
To investigate the potential for paternal control over sperm genome remodeling post-fertilization, a candidate gene approach was taken, focusing on the Heterochromatin Protein 1 (HP1) proteins that orchestrate genome-wide chromosomal organization in plants, animals, fungi, and some protists. HP1 proteins are defined as such by a combination of two domains - a chromodomain that mediates protein-histone interactions and a chromoshadow domain that mediates protein-protein interactions. The biochemical properties of HP1 members support a diversity of chromatin-dependent processes in the soma, including DNA replication, telomere integrity, and chromosome condensation (Levine, 2015).
Recently, a detailed phylogenomic analysis of the HP1 gene family was carried out in Drosophila that revealed numerous testis-restricted HP1 proteins. Given the established roles of HP1 proteins, it was posited that these newly discovered male-specific HP1 genes might represent excellent candidates for encoding chromatin functions specialized for paternal genome organization and remodeling in the early embryo. Using detailed genetic and cytological analyses, this study shows that one of these testis-specific HP1 proteins, Heterochromatin Protein 1E (HP1E), is essential for priming the paternal genome to enter embryonic mitosis in synchrony with the maternal genome in D. melanogaster. Intriguingly, HP1E is able to mediate this priming function transgenerationally i.e., the HP1E protein itself is not epigenetically inherited. It was further shown that absence of HP1E especially imperils mitotic fidelity of the heterochromatin-rich, paternal sex chromosomes. Thus, this study firmly establishes that both maternal and paternal control are necessary for paternal genome remodeling in the early Drosophila embryo (Levine, 2015).
Properly coordinated chromosome segregation during virtually all mitotic divisions relies on the function of multiple cell cycle checkpoint proteins. No such cell cycle checkpoint proteins have been identified to act in the very first embryonic mitotic cycle, which must nevertheless accomplish the difficult task of synchronizing maternal and paternal chromosomes that were inherited in very different chromatin states. To investigate the paternal contributions that ensure timely participation of the paternal genome in early embryogenesis, a detailed functional analysis was carried out of the testis-restricted HP1E gene in D. melanogaster. It was found that HP1E encodes a novel function that ensures paternal genome stability in the embryo. Cytological and transcriptome analysis revealed that HP1E is developmentally restricted within the male germline, where it contributes to heterochromatin integrity. HP1E depletion during sperm development results in a highly penetrant PEL phenotype in which paternal chromosomes, especially the paternal sex chromosomes, fail to condense in synchrony with the maternal chromosomes and ultimately cause mitotic catastrophe. It was further shown that the PEL embryonic phenotype could not be rescued by egg-supplied HP1E but could be rescued if the paternal DNA was excluded from participating in embryonic mitosis. These observations support a model under which HP1E acts pre-fertilization to ensure proper chromosome condensation and segregation of paternal chromosomes post-fertilization (Levine, 2015).
The 'hit and run' priming function clearly distinguishes HP1E from all other previously characterized paternal effect lethal genes, which encode proteins that are transmitted to the embryo via sperm. These include the Drosophila paternal chromatin-associated PEL, k81, which encodes a protein that persists on paternal telomeres from late spermatogenesis to the first embryonic mitosis. The HP1E-depletion phenotype is instead reminiscent of Drosophila fathers infected with Wolbachia bacteria crossed to uninfected females. Embryonic lethality induced by Wolbachia testis infection is also caused by a pre-fertilization modification to the paternal genome that results in paternal-maternal chromatin asynchrony and mis-segregation at the very first zygotic mitosis. However, Wolbachia-associated PEL results in mis-segregation of the entire paternal genome rather than just the heterochromatin-rich chromosomes observed in HP1E PEL . Moreover, the HP1E PEL defect is completely independent of Wolbachia (PEL phenotype persists for Wolbachia-free males and females). It is therefore concluded that HP1E supports a novel chromatin requirement to prime paternally inherited genomes for synchronous and successful embryonic mitosis (Levine, 2015).
How does HP1E ensure timely mitotic entry? It is formally possible that the PEL phenotype is the consequence of a dysregulated spermatid transcriptome that is, up- or down-regulation of a downstream gene. However, the finding that HP1E depletion results in the global up-regulation of heterochromatin-embedded genes, together with the observation that the heterochromatin-rich paternal sex chromosomes are most vulnerable to HP1E depletion, lead to favoring the alternate model that HP1E functions as a canonical HP1 protein during spermiogenesis. Based on antibody localization and chromatin bridge morphology, no evidence was found for defects in kinetochore assembly or replication machinery engagement in PEL embryos. Instead, the observation that the lethality phenotype first manifests as decondensed paternal chromosomes relative to maternal chromosomes implicates condensation delay of the heterochromatin-rich sex chromosomes. This delay could be the consequence of incomplete replication. Indeed, large stretches of uninterrupted heterochromatic DNA, as found on the Drosophila sex chromosomes, pose a unique challenge to replication. Alternatively, the mitotic delay may be the result of inadequate condensin protein recruitment, which is required for timely resolution of sister chromatids post-replication. Previous studies have shown that heterochromatin can also impair chromosome condensation. Timely completion of replication and condensation requires the action of HP1E's closest relative, HP1A, in somatic cells. However, in developing spermatids, HP1A localizes to telomeres rather than broadly to heterochromatin as observed in virtually all other cell types. It is posited that HP1E adopts a global, HP1A-like chromatin function during this highly specialized developmental stage and ensures the recruitment or retention of either replication or condensin proteins that are required post-fertilization (Levine, 2015).
Previous studies have shown that HP1A is essential for embryo viability. This study shows that paternally-acting HP1E is also essential for embryogenesis. Both HP1A and HP1E evolve under purifying selection. However, unlike HP1A (encoded by Su(var)205), HP1E has an unusually dynamic evolutionary history. Despite ancient origins, HP1E has been recurrently lost over evolutionary time. HP1E has been apparently replaced by younger, testis restricted HP1 paralogs on at least two occasions during Drosophila evolution. Curiously, Drosophila pseudoobscura and related species encode neither HP1E nor a putative replacement testis-specific HP1 gene. How is the paradox of HP1E essentiality in D. melanogaster reconciled with its loss in D. pseudoobscura? It was previously found that HP1E loss along in D. pseudoobscura-related species occurred during the same 7-million evolutionary period as a major sex chromosome rearrangement event, in which the ancestral Y was lost, a neo-Y chromosome was born, and the ancestral X fused to an autosome. The finding that the D. melanogaster sex chromosomes are especially vulnerable to HP1E depletion, combined with the emergence of novel sex chromosome arrangements along the same narrow branch as HP1E pseudogenization, suggests a model under which rearrangements of heterochromatin-rich sex chromosomes in the obscuragroup rendered HP1E non-essential. Such karyotypic changes can bring distal heterochromatin into closer proximity to euchromatin and be sufficient to alter heterochromatin packaging, replication timing or even delete blocks of satellite repeats. Thus, heterochromatin evolution via chromosomal rearrangements may have obviated maintenance of HP1E's essential heterochromatin function, leading to its degeneration in D. pseudoobscura (Levine, 2015).
The finding that HP1E is essential in D. melanogaster yet lost in the obscura group highlights the lineage-restricted essential requirements of chromatin genes. Intriguingly, the only other characterized PEL gene that supports paternal chromatin function in Drosophila embryos, k81, is similarly lineage-restricted despite being essential for paternal telomere function. In contrast, maternally deposited proteins required for paternal chromatin reorganization following fertilization are generally conserved from fly to human. This dichotomy is striking. It specifically suggests that even though the essential functions of paternal control of DNA deposition and chromatin remodeling for embryonic mitosis are likely to be conserved in most animals, whereas the identity of those genes is not. PEL chromatin genes like HP1E and k81 thus challenge the dogma that ancient, conserved genes always encode essential conserved functions. Not only can young genes rapidly acquire essential chromatin functions due to dynamic chromatin evolution, but chromatin changes, such as those driven by karyotype evolution, may also drive the extinction of ancient genes encoding once-essential functions (Levine, 2015).
Adult stem cells often divide asymmetrically to produce one self-renewed stem cell and one differentiating cell, thus maintaining both populations. The asymmetric outcome of stem cell divisions can be specified by an oriented spindle and local self-renewal signals from the stem cell niche. This study shows developmentally programmed asymmetric behavior and inheritance of mother and daughter centrosomes underlies the stereotyped spindle orientation and asymmetric outcome of stem cell divisions in the Drosophila male germ line. The mother centrosome remains anchored near the niche while the daughter centrosome migrates to the opposite side of the cell before spindle formation (Yamashita, 2007).
Adult stem cells maintain populations of highly differentiated but short-lived cells throughout the life of the organism. To maintain the critical balance between stem cell and differentiating cell populations, stem cells have a potential to divide asymmetrically, producing one stem and one differentiating cell. The asymmetric outcome of stem cell divisions can be specified by regulated spindle orientation, such that the two daughter cells are placed in different microenvironments that either specify stem cell identity (stem cell niche) or allow differentiation. Drosophila male germline stem cells (GSCs) are maintained through attachment to somatic hub cells, which constitute the stem cell niche. Hub cells secrete the signaling ligand Upd, which activates the Janus kinase-signal transducer and activator of transcription (JAK-STAT) pathway in the neighboring germ cells to specify stem cell identity. Drosophila male GSCs normally divide asymmetrically, producing one stem cell, which remains attached to the hub, and one gonialblast, which initiates differentiation. This stereotyped asymmetric outcome is controlled by the orientation of the mitotic spindle in GSCs: The spindle lies perpendicular to the hub so that one daughter cell inherits the attachment to the hub, whereas the other is displaced away (Yamashita, 2007).
The stereotyped orientation of the mitotic spindle is set up by the positioning of centrosomes during interphase. GSCs remain oriented toward the niche throughout the cell cycle. In G1 phase, the single centrosome is located near the interface with the hub. When the duplicated centrosomes separate in G2 phase, one stays next to the hub, whereas the other migrates to the opposite side of the cell. Centrosomes in the GSCs separate unusually early in interphase, rather than at the G2-prophase transition, so it is common to see GSCs with fully separated centrosomes without a spindle (Yamashita, 2007).
Differences between the mother and daughter centrosomes underlie the stereotyped behavior of the centrosomes in Drosophila male GSCs. The mother centrosome normally remains anchored to the hub-GSC interface and is inherited by the GSC, whereas the daughter centrosome moves away from the hub and is inherited by the cell that commits to differentiation. Mother and daughter centrosomes were differentially labeled by transient expression of green fluorescent protein-pericentrin/AKAP450 C-terminus (GFP-PACT) from the Drosophila pericentrin-like protein under heat shock-Gal4 control. The PACT domain, which is necessary and sufficient for centriolar localization, is incorporated into centrioles only during centrosome duplication and does not exchange with the cytoplasmic pool. Both the mother and daughter centrosomes are labeled by GFP-PACT in the first cell cycle after heat shock. In the second cell cycle, the daughter centrosome retains GFP-PACT, whereas the mother centrosome is not labeled, thus distinguishing the mother and daughter centrosomes. After a short burst of GFP-PACT expression induced by a 2.5-hour heat shock, 20% - 30% of the GSCs had GFP-labeled centrosomes, indicating the duplication of centrosomes during the window of GFP-PACT expression. By 12 hours after heat shock, >90% of the labeled GSCs had two GFP-positive centrosomes, indicating that they had progressed to the G2 phase of the first cell cycle after GFP-PACT incorporation (Yamashita, 2007).
By 18 to 24 hours after heat shock, the number of GSCs with two GFP-positive centrosomes had decreased, whereas the number of GSCs with one GFP-positive and one GFP-negative centrosome had increased, suggesting progression into the second cell cycle. Generally, the centrosome distal to the hub was labeled, whereas the centrosome proximal to the hub was GFP-negative, indicating that the daughter centrosomes migrate away from the hub-GSC interface during asymmetric GSC divisions (Yamashita, 2007).
Labeling the mother rather than the daughter centrosomes confirmed that the male GSCs in the niche preferentially retain mother centrosomes over time. Centrioles assembled during early embryogenesis were labeled using the NGT40 Gal4 driver to drive the expression of GFP-PACT in blastoderm-stage embryos, shutting off after germband extension. In the first cell cycle after the depletion of the cytoplasmic pool of GFP-PACT in the GSCs, both the mother and daughter centrosomes should be labeled. In subsequent cell cycles, only the mother centrosomes should be labeled (Yamashita, 2007).
In most GSCs in the second or later cell cycle after the depletion of cytoplasmic GFP-PACT, the labeled centrosome was positioned next to the hub-GSC interface, and the unlabeled centrosome had moved away from the hub. The frequency of GSCs that had the proximal, but not distal, centrosome labeled remained constant over time for 10 days (L3 larvae to day-3 adults), suggesting that the mother centrosomes are reliably retained by the GSCs, even through multiple rounds of GSC divisions. Some GSCs maintained cytoplasmic GFP-PACT, especially in L3 larvae, suggesting that the GFP-PACT had not yet been diluted out. Some GSCs with two labeled centrosomes were observed, suggesting that they are in the first cell cycle after the depletion of cytoplasmic GFP-PACT (Yamashita, 2007).
The mother centrosomes in GSCs appeared to maintain robust interphase microtubule arrays. Ultrastructural analysis of the GSCs revealed that the centrosome proximal to the hub was commonly associated with many microtubules throughout the cell cycle. Nineteen centrosomes in GSCs were scored in serial sections of the apical tips of five wild-type testes. Eleven centrosomes were localized close to the adherens junctions between the hub and the GSCs. Nine of these proximal centrosomes appeared to be in interphase cells, based on nuclear morphology and microtubule arrangement. Typically, these interphase centrosomes proximal to the hub were associated with numerous microtubules. In some samples, microtubules appeared to extend from the centrosome toward the adherens junctions. The other two proximal centrosomes appeared to be in cells in mitotic prophase, based on their robust microtubule arrays containing bundled microtubules running parallel to or piercing the nuclear surface (Yamashita, 2007).
In contrast, of the five distal centrosomes in the apparently interphase cells that were scored, four had few associated microtubules. The remaining three distal centrosomes appeared to be in cells in mitotic prophase, based on microtubule arrays containing bundled microtubules. Thus, the mother centrosomes may maintain interphase microtubule arrays that anchor them to the hub-GSC interface, whereas the daughter centrosomes may initially have few associated microtubules and be free to move, establishing a robust microtubule array only later in the cell cycle (Yamashita, 2007).
Consistent with the idea that astral microtubules anchor the mother centrosomes to the hub-GSC interface, mother- versus daughter centrosome positioning was randomized in GSCs that were homozygous mutant for centrosomin (cnn), an integral centrosomal protein required to anchor astral microtubules to centrosomes. Analysis of mother and daughter centrosomes after transient expression of GFP-PACT revealed that, for cnn homozygous mutant GSCs where one of the two centrosomes was positioned next to the hub, it was essentially random whether the mother or the daughter centrosome stayed next to the hub. In addition, in >25% of total labeled GSCs, neither of the two centrosomes was next to the hub (Yamashita, 2007).
These results indicate that the two centrosomes in Drosophila male GSCs have different characters and fates. The mother centrosome stays next to the junction with the niche and is inherited by the cell that self-renews stem cell fate. Thus, GSCs can maintain an old centriole assembled many cell generations earlier. In contrast, the daughter centrosome migrates away from the niche and is inherited by the cell that will initiate differentiation. It is postulated that the mother centrosomes in male GSCs may remain anchored to the GSC-niche interface throughout the cell cycle by attachment to astral microtubules connected to the adherens junction, whereas the daughter centrosomes may initially have few associated microtubules and thus can move away from the niche. Microtubule-dependent differential segregation of mother and daughter spindle-pole bodies (equivalent to centrosomes in higher organisms) is observed in budding yeast. In cultured vertebrate cells, the centrioles mature slowly over the cell cycle, and the mother centrosomes (containing a mature centriole) attach astral microtubules more effectively and are more stationary than daughter centrosomes in interphase. The unusually early separation of centrosomes in interphase male GSCs may provide a way to move the daughter centrosome out of range of the stabilizing influence of the adherens junction complex before it becomes competent to hold a robust microtubule array (Yamashita, 2007).
Developmentally programmed anchoring of the mother centrosome may provide a key mechanism to ensure the stereotyped orientation of the mitotic spindle and thus the reliably asymmetric outcome of the male GSC divisions. Although it is tempting to speculate that determinants associated with the mother or daughter centrosome may play a role in specifying stem cell or differentiating-cell fates, such determinants are yet to be identified. Rather, the asymmetric inheritance of mother and daughter centrosomes in male GSCs may be a consequence of the cytoskeletal mechanisms that are imposed as part of the stem cell program to anchor one centrosome next to the niche throughout the interphase, ensuring a properly oriented spindle (Yamashita, 2007).
Stem cell division is tightly controlled via secreted signaling factors and cell adhesion molecules provided from local niche structures. Molecular mechanisms by which each niche component regulates stem cell behaviors remain to be elucidated. This study shows that heparan sulfate (HS), a class of glycosaminoglycan chains, regulates the number and asymmetric division of germline stem cells (GSCs) in the Drosophila testis. GSC number is sensitive to the levels of 6-O sulfate groups on HS. Loss of 6-O sulfation also disrupted normal positioning of centrosomes, a process required for asymmetric division of GSCs. Blocking HS sulfation specifically in the hub led to increased GSC numbers and mispositioning of centrosomes. The same treatment also perturbed the enrichment of Apc2, a component of the centrosome anchoring machinery, at the hub-GSC interface. This perturbation of the centrosome anchoring process ultimately led to an increase in the rate of spindle misorientation and symmetric GSC division. This study shows that specific HS modifications provide a novel regulatory mechanism for stem cell asymmetric division. The results also suggest that HS-mediated niche signaling acts upstream of GSC division orientation control (Levings, 2016).
A long-standing question concerns how stem cells maintain their identity
through multiple divisions. It has been reported that pre-existing and
newly synthesized histone H3 are
asymmetrically distributed during Drosophila male
germline stem cell (GSC) asymmetric division. This study shows that
phosphorylation at threonine 3 of H3 (H3T3P) distinguishes pre-existing
versus newly synthesized H3. Converting T3 to the unphosphorylatable
residue alanine (H3T3A) or to the phosphomimetic aspartate (H3T3D)
disrupts asymmetric H3 inheritance. Expression of H3T3A or H3T3D
specifically in early-stage germline also leads to cellular defects,
including GSC loss and germline tumors. Finally, compromising the activity
of the H3T3 kinase Haspin
enhances the H3T3A but suppresses the H3T3D phenotypes. These studies
demonstrate that H3T3P distinguishes sister chromatids enriched with
distinct pools of H3 in order to coordinate asymmetric segregation of
"old" H3 into GSCs and that tight regulation of H3T3 phosphorylation is
required for male germline activity (Xie, 2015).
Epigenetic phenomena are heritable changes in gene expression or function that can persist throughout many cell divisions without alterations in primary DNA sequences. By regulating differential gene expression, epigenetic processes are able to direct cells with identical genomes to become distinct cell types in humans and other multicellular organisms. However, with the exception of DNA methylation, little is known about the molecular pathways leading to epigenetic inheritance (Xie, 2015).
Prior research has shown that epigenetic events play particularly important roles in ensuring both proper maintenance and differentiation of several stem cell populations. Many types of adult stem cells undergo asymmetric cell division to generate a self-renewed stem cell and a daughter cell that will subsequently differentiate. Mis-regulation of this balance leads to many human diseases, ranging from cancer to tissue dystrophy to infertility. However, the mechanisms of stem cell epigenetic memory maintenance as well as how loss of this memory contributes to disease remain unknown (Xie, 2015).
During the asymmetric division of the Drosophila male germline stem cell (GSC), the pre-existing histone 3 (H3) is selectively segregated to the self-renewed GSC daughter cell whereas newly synthesized H3 is enriched in the differentiating daughter cell known as a gonialblast (GB) (Tran, 2012). In contrast, the histone variant H3.3, which is incorporated in a replication-independent manner, does not exhibit such an asymmetric pattern. Furthermore, asymmetric H3 inheritance occurs specifically in asymmetrically dividing GSCs, but not in the symmetrically dividing progenitor cells. These findings demonstrate that global asymmetric H3 histone inheritance possesses both molecular and cellular specificity. The following model is proposed to explain these findings (Xie, 2015).
First, the cellular specificity exhibited by the H3 histone suggests that global asymmetric histone inheritance occurs uniquely in a cell-type (GSC) where the mother cell must divide to produce two daughter cells each with a unique cell fate. Because this asymmetry is not observed in symmetrically dividing GB cells, asymmetric histone inheritance is proposed to be a phenomenon specifically employed by GSCs to establish unique epigenetic identities in each of the two daughter cells. Second, as stated previously, a major difference between H3 and H3.3 is that H3 is incorporated to chromatin during DNA replication, while H3.3 variant is incorporated in a replication-independent manner. Because this asymmetric inheritance mode is specific to H3, a two-step model is proposed to explain asymmetric H3 inheritance: (1) prior to mitosis, pre-existing and newly synthesized H3 are differentially distributed on the two sets of sister chromatids, and (2) during mitosis, the set of sister chromatids containing pre-existing H3 is segregated to GSCs, while the set of sister chromatids enriched with newly synthesized H3 is segregated to the GB that differentiates (Tran, 2012; Tran, 2013; Xie, 2015 and references therein)
This study reports that a mitosis-enriched H3T3P mark acts as a transient landmark that distinguishes sister chromatids with identical genetic code but different epigenetic information, shown as pre-existing H3-GFP and newly synthesized H3-mKO. By distinguishing sister chromatids containing different epigenetic information, H3T3P functions to allow these molecularly distinct sisters to be segregated and inherited differentially to the two daughter cells derived from one asymmetric cell division. The selective segregation of different populations of histones likely allows these two cells to assume distinct fates: self-renewal versus differentiation. Consequently, loss of proper epigenetic inheritance might lead to defects in both GSC maintenance and GB differentiation, suggesting that both cells need this active partitioning process to either 'remember' or 'reset' their molecular properties (Xie, 2015).
The temporal and spatial specificities of H3T3P make it a great candidate to regulate asymmetric sister chromatid segregation. First, H3T3P is only detectable from prophase to metaphase, the window of time during which the mitotic spindle actively tries to attach to chromatids through microtubule-kinetochore interactions. Second, the H3T3P signal is enriched at the peri-centromeric region, where kinetochore components robustly crosstalk with chromatin-associate factors. Third, H3T3 shows a sequential order of phosphorylation, first appearing primarily on sister chromatids enriched with pre-existing H3 and then subsequently appearing on sister chromatids enriched with newly synthesized H3 as the GSC nears metaphase. The distinct temporal patterns shown by H3T3P are unique to GSCs and would allow the mitotic machinery to differentially recognize sister chromatids bearing distinct epigenetic information; an essential step necessary for proper segregation during asymmetric GSC division. Furthermore, the tight temporal control of H3T3 phosphorylation suggests that rather than serving as an inherited epigenetic signature, H3T3P may act as transient signaling mark to allow for the proper partitioning of H3. It is hypothesized that H3T3P needs to be under tight temporal control in order to ensure proper H3 inheritance and germline activity (Xie, 2015).
These studies have shown that H3T3P is indeed subject to stringent temporal controls during mitosis. The H3T3P mark is undetectable during G2 phase. Upon entry to mitosis, sister chromatids enriched with pre-existing H3-GFP histone begin to show H3T3 phosphorylation prior to sister chromatids enriched with newly synthesized H3-mKO. As the cell continues to progress toward metaphase, H3T3P signal begins to appear on sister chromatids enriched with newly synthesized H3-mKO. Such a tight regulation of H3T3P is compromised when levels of H3T3P are altered due to the incorporation of mutant H3T3A or H3T3D. Incorporation of the H3T3A mutant results in a significant decrease in the levels of H3T3P on sister chromatids throughout mitosis, such that neither sister becomes enriched with H3T3P as the GSC progresses toward metaphase. Conversely, incorporation of the H3T3D mutant would result in seemingly elevated levels of H3T3P early in mitosis. Although H3T3A and H3T3D act in different ways, both mutations significantly disrupt the highly regulated temporal patterns associated with H3T3 phosphorylation, the result of which is randomized H3 inheritance patterns and germ cell defects in testes expressing either H3T3A or H3T3D (Xie, 2015).
To further evaluate the extent of H3T3A and H3T3D roles in the segregation of sister chromatids enriched with different populations of H3 during mitosis, all possible segregation patterns were modeled in male GSCs, and these estimates were compared to the experimental results. To simplify the calculations, two important assumptions were made: first, nucleosomal density was assumed to be even throughout the genome. This assumption allows the inference that the overall fluorescent signal contributed by each chromosome is proportional to their respective number of DNA base pairs. Second, by quantifying pre-existing H3-GFP asymmetry in anaphase and telophase GSCs, it was estimated that the establishment of H3-GFP asymmetry is ∼4-fold biased, i.e., 80% on one set of sister chromatids and 20% on the other set of sister chromatids, based on quantification of GFP signal in anaphase and telophase GSCs (Tran, 2012). With these two simplifying assumptions, both GFP and mKO ratios were caculated among all 64 possible combinations. If asymmetry is designed as a greater than 1.5-fold difference in fluorescence intensity, then based on a model of randomized sister chromatid segregation, it is estimated that a symmetric pattern should appear for 53.1% (34/64) of GSC-GB pairs whereas both conventional and inverted asymmetric patterns should occur with equal frequencies and account for 18.7% (12/64) of total GSC-GB pairs. The remaining 9.4% (6/64) of GSC-GB pairs should produce histone inheritance patterns with a 1.45- to 1.55-fold difference in signal intensity (Xie, 2015).
This estimation is close to the experimental data in both H3T3A- and H3T3D-expressing testes. Of the 64 quantified post-mitotic GSC-GB pairs in nos>H3T3A testes, ∼71.9% showed symmetric inheritance pattern. Conventional and inverted asymmetric patterns were detected at 9.4% and 12.5%, respectively, and 6.3% at the borderline. Similarly, of the 57 quantified post-mitotic GSC-GB pairs in nos>H3T3D testes, ∼79.0% showed symmetric inheritance pattern. Conventional and inverted asymmetric patterns were detected at 7.0% and 10.5%, respectively with 3.5% of pairs at the borderline. Some differences between predicted ratios and the experimental data could be due to the simplified assumptions, the limited sensitivity of the measurement, and/or some coordinated chromatid segregation modes that bias the eventual read-out. In summary, comparison between the modeling ratios and the experimental data suggest that loss of the tight control of H3T3 phosphorylation in GSCs randomizes segregation of sister chromatids enriched with different populations of H3 (Xie, 2015).
If the temporal separation in the phosphorylation of H3T3 on epigenetically distinct sister chromatids facilitates their proper segregation and inheritance during asymmetric cell division, it is likely that mutations of the Haspin kinase will also affect the temporal control of H3T3 phosphorylation. In the context of H3T3A, where the levels of H3T3P are already reduced, a further decrease in H3T3P by reducing Haspin levels should limit the GSC's ability to distinguish between sister chromatids enriched with distinct H3. Indeed, haspin mutants enhance the phenotypes in nos>H3T3A testes. A different situation appears in the context of H3T3D where sister chromatids experience seemingly elevated levels of H3T3P at the start of mitosis. These elevated H3T3P levels may be exacerbated by the phosphorylation activity of the Haspin kinase. Therefore, it is conceivable that by halving the levels of the Haspin kinase, H3T3 phosphorylation should be reduced to a level more closely resembling wild-type. In this way, some of the temporal specificity that is lost in the H3T3D mutant is restored, resulting in suppression of the phenotypes observed in nos>H3T3D testes. An exciting topic for future study would be to further explore how exactly Haspin phosphorylates H3T3 in the context of chromatin and whether H3T3A and H3T3D mutations act synergistically or antagonistically in regulating asymmetric sister chromatids segregation through differential phosphorylation of a key histone residue (Xie, 2015).
It would also be interesting to understand the potential connection between asymmetric histone inheritance and another phenomenon reported by several investigators: selective DNA strand segregation. Recent development of the chromosome orientation fluorescence in situ hybridization (CO-FISH) technique allows study of selective chromatid segregation at single-chromosome resolution. Using this technique in mouse satellite cells, it has been demonstrated that all chromosomes are segregated in a biased manner, such that pre-existing template DNA strands are preferentially retained in the daughter cell that retains stem cell identity. Interestingly, this biased segregation becomes randomized in progenitor non-stem cells. Using CO-FISH in Drosophila male GSCs, sex chromosomes have been shown to segregate in a biased manner. Remarkably, sister chromatids from homologous autosomes have been shown to co-segregate independent of any specific strand preference. Such findings hint at a possible epigenetic source guiding the coordinated inheritance of Drosophila homologous autosomes. In many cases of biased inheritance, researchers have speculated about the existence of a molecular signature that would allow the cell to recognize and segregate sister chromatids bearing differential epigenetic information. However, the identity of such a signature has remained elusive. The work represented in this paper provides experimental evidence demonstrating that a tightly-controlled histone modification, H3T3P, is able to distinguish sister chromatids and coordinate their segregation (Xie, 2015).
Epigenetic processes play important roles in regulating stem cell identity and activity. Failure to appropriately regulate epigenetic information may lead to abnormalities in stem cell behaviors, which underlie early progress toward diseases such as cancer and tissue degeneration. Due to the crucial role that such processes play in regulating cell identity and behavior, the field has long sought to understand whether and how stem cells maintain their epigenetic memory through many cell divisions. Yhe results of this study suggest that the asymmetric segregation of pre-existing and newly synthesized H3-enriched chromosomes may function to determine distinct cell fates of GSCs versus differentiating daughter cells (Xie, 2015).
Adult stem cells modulate their output by varying between symmetric and asymmetric divisions, but have rarely been observed in living intact tissues. Germline stem cells (GSCs) in the Drosophila testis are anchored to somatic hub cells and were thought to exclusively undergo oriented asymmetric divisions, producing one stem cell that remains hub-anchored and one daughter cell displaced out of the stem cell-maintaining micro-environment (niche). Extended live imaging of the Drosophila testis niche was developed, allowing the tracking of individual germline cells. Surprisingly, new wild-type GSCs are generated in the niche during steady-state tissue maintenance by a previously undetected event termed 'symmetric renewal', where interconnected GSC-daughter cell pairs swivel such that both cells contact the hub. GSCs were captured undergoing direct differentiation by detaching from the hub. Following starvation-induced GSC loss, GSC numbers are restored by symmetric renewals. Furthermore, upon more severe (genetically induced) GSC loss, both symmetric renewal and de-differentiation (where interconnected spermatogonia fragment into pairs while moving towards then establishing contact with the hub) occur simultaneously to replenish the GSC pool. Thus, stereotypically oriented stem cell divisions are not always correlated with an asymmetric outcome in cell fate, and changes in stem cell output are governed by altered signals in response to tissue requirements (Sheng, 2011).
Live imaging of the Drosophila germline stem cell niche has directly demonstrated many aspects of GSC behavior that were impossible to observe in fixed tissues. Asymmetrically oriented divisions do not necessarily determine asymmetric cell fate, but can occasionally result in the production of two GSCs. This is the primary mechanism by which GSCs are replenished in healthy tissues to compensate for GSC loss. As GSC-daughter pairs are adjacent to the hub and are enriched in the maintenance factor STAT92E, the process of symmetric renewal is probably distinct from de-differentiation of spermatogonia (which are non hub-adherent and express the differentiation factor Bam). The frequency of symmetric renewal increases during GSC recovery after protein starvation, and during GSC regeneration after genetically induced stem cell depletion. In the latter case, where the rate of GSC regeneration is higher, GSCs are concurrently derived from de-differentiating spermatogonia, a process characterized by movement, fragmentation and adhesion to the hub by spermatogonial cells. Together, these data demonstrate that lost GSCs can be regenerated by multiple mechanisms, some or all of which may be similar to events occurring in other stem cell systems (Sheng, 2011).
As changes in stem cell output are observed during regeneration, signaling from support cells or from systemic factors may underlie these effects. Niche-generating cells, transit amplifying daughter cells or even differentiated daughter cells may potentially signal to stem cells and modulate their division output. In the Drosophila testis, GSC maintenance depends on Jak-STAT signaling initiated from the hub, but it is not known whether this same pathway regulates division outcome. As STAT-null GSCs are rapidly lost from the niche, low levels of Jak-STAT signaling due to fluctuations in gene expression may be sufficient to cause GSC loss. In support of this hypothesis, three out of 556 GSCs examined for STAT92E expression had low levels of this protein. However, the mRNA expression pattern of the Jak-STAT pathway ligand Upd is unchanged during de-differentiation, suggesting that genes other than Upd may affect symmetric renewals. BMP signaling, which is required for GSC maintenance, is a good candidate. Combining live imaging with genetic tools for monitoring levels of signaling pathway activation in the Drosophila testis will provide a powerful platform for understanding how cell signaling affects the outcome of stem cell divisions in real time (Sheng, 2011).
The observation that both symmetric renewal and GSC loss occur when the GSC is attached to a daughter cell suggests that there may be a cell cycle-specific gene expression profile that primes the cells for these events to occur during S or early G2 in the cell cycle. It is speculated that the abscission accompanying symmetric renewal is similar to that occurring in GSC-GB pairs, another G2 event. Cell cycle regulation, which is characterized by a short G1 phase and relatively long S phase, maintains pluripotency in many types of cultured stem cells. As GSCs in the Drosophila testis have short G1 phases, and Drosophila GSCs require distinct cell cycle regulators, investigation of cell cycle regulation of Drosophila GSC division outcome may be informative (Sheng, 2011).
It was shown that GSCs in both centrosomin mutants and starved wild-type flies have increased frequencies of symmetric renewal, but surprisingly, there is no corresponding rise in GSC numbers. These results suggest that increased symmetric renewal is counterbalanced by increased GSC loss. Cnn mutant GSC are reported to have abnormal cell morphology and often appear to be detaching from the hub, suggesting an overall maintenance defect. During starvation, lowered insulin signaling results in GSCs loss, and this effect can be rescued by overactivation of insulin signaling. The results indicate that symmetric renewals of GSCs undergoing oriented divisions are the source of new GSCs. Starved flies initially have low insulin signaling, but when returned to normal food for a day have higher insulin signaling. However, this study found that both timepoints exhibited increased symmetric renewals, leading to the idea that activation of insulin signaling does not directly modulate division outcome. Perhaps during starvation, lowered insulin signaling causes GSC loss, which in turn triggers a compensatory increase in symmetric renewal. However, symmetric renewals are not able to fully compensate for the loss, yielding an overall decrease in GSC number. When flies are re-fed and insulin signaling returns to normal, GSCs are no longer rapidly lost, and the same rate of symmetric renewal is now able to increase overall GSC number. Together, these results suggest that the behavior of stem cells within the niche is much more dynamic than previously expected, and indicate that GSC number is controlled by the relative rates of symmetric renewal versus loss, not by the orientation of the division plane (Sheng, 2011).
Why do the majority of Drosophila GSCs undergo asymmetric division if symmetric renewal plus symmetric differentiation produces the same output? As GSCs and CySCs function together within the niche during spermatogenesis, robust division orientation of both populations may enable differentiating germline cells to be generated at a rate that matches cyst cell production. Asymmetric divisions may also prevent clonal expansion of stem cells harboring harmful mutations within the niche, which can compete for niche occupancy. However, clonal expansion may not always be harmful; mammalian niches regularly progress towards mono-clonality with stem cells exhibiting neutral drift dynamics. Perhaps symmetrically renewing divisions are not detrimental to mammalian systems because mammalian niches are not as constrained spatially, and mammalian stem cells are often motile. So far, asymmetric division in Drosophila testes correlates with optimal GSC function, as it becomes less robust with aging. Whether symmetric divisions increase during aging has not been examined, but it might occur because GSCs are thought to be lost more frequently due to decreased maintenance cues. Interestingly, depleting STAT92E from GSCs displaces them from the hub, yet they are not lost from the tissue. Instead, they associate with BMP-producing CySCs, which probably promote GSC renewal. However, GSC division orientation is now randomized; suggesting that their output is composed of symmetric renewals and symmetric differentiation. Furthermore, APC2 mutants that affect centrosome position and E-cadherin mutants that have misoriented divisions still have wild-type GSC numbers. Together, these observations suggest that the Drosophila testis stem cell niche does not require invariant asymmetric GSC division outcomes (Sheng, 2011).
As mammalian stem cells are thought to undergo symmetric renewal in combination with stochastic differentiation, rather than strict asymmetric divisions, GSCs in Drosophila may share more aspects of stem cell behavior with mammalian systems than has been previously assumed. Wild-type GSCs were observed losing niche attachment and directly differentiating, which is consistent with reports that subsets of undifferentiated spermatogonia in the mouse testes can directly differentiate. Although a lost GSC being replaced by a neighboring GSC undergoing symmetric renewal was observed, this was only a single example where these events are coupled together. Thus, stem cell loss and symmetric renewal may occur stochastically in Drosophila GSCs, as in the mouse testis. It was also shown that differentiating spermatogonia revert into GSCs, which is consistent with findings that differentiating spermatogonia can contribute to the stem cell pool during reconstitution of spermatogenesis in the mouse testes. Therefore, this system provides an ideal platform for determining regulators of stem cell loss and replacement in vivo that may also be conserved in mammalian tissues (Sheng, 2011).
Adult stem cells reside in specialized regulatory microenvironments, or niches, where local signals ensure stem cell maintenance. The Drosophila testis contains a well-characterized niche wherein signals from postmitotic hub cells promote maintenance of adjacent germline stem cells and somatic cyst stem cells (CySCs). Hub cells were considered to be terminally differentiated; this study shows that they can give rise to CySCs. Genetic ablation of CySCs triggers hub cells to transiently exit quiescence, delaminate from the hub, and convert into functional CySCs. Ectopic Cyclin D-Cdk4 expression in hub cells is also sufficient to trigger their conversion into CySCs. In both cases, this conversion causes the formation of multiple ectopic niches over time. Therefore, this work provides a model for understanding how oncogenic mutations in quiescent niche cells could promote loss of quiescence, changes in cell fate, and aberrant niche expansion (Hetie, 2014).
Asymmetric division of adult stem cells generates one self-renewing stem cell and one differentiating cell, thereby maintaining tissue homeostasis. A decline in stem cell function has been proposed to contribute to tissue ageing, although the underlying mechanism is poorly understood. This study shows that changes in the stem cell orientation with respect to the niche during ageing contribute to the decline in spermatogenesis in the male germ line of Drosophila. Throughout the cell cycle, centrosomes in germline stem cells (GSCs) are oriented within their niche and this ensures asymmetric division. GSCs containing misoriented centrosomes accumulate with age, and these GSCs are arrested or delayed in the cell cycle. The cell cycle arrest is transient, and GSCs appear to re-enter the cell cycle on correction of centrosome orientation. On the basis of these findings, it is proposed that cell cycle arrest associated with centrosome misorientation functions as a mechanism to ensure asymmetric stem cell division, and that the inability of stem cells to maintain correct orientation during ageing contributes to the decline in spermatogenesis. It was also shown that some of the misoriented GSCs probably originate from dedifferentiation of spermatogonia (Cheng, 2008).
GSCs with misoriented centrosomes accumulate as flies age. Since such misoriented GSCs divide less frequently as compared to oriented GSCs, accumulation of misoriented GSCs contributes to the decline in spermatogenesis that occurs with age. Although misoriented GSCs rarely divide, they are not permanently arrested (or senescent) and are correctly oriented when they divide. Whether correction of GSC orientation is an active process that is part of the acquisition of stem cell identity remains to be determined. The low cell cycle activity of misoriented GSCs may also suggest that mechanisms are in place to detect misorientation and induce cell cycle arrest in response to this change, although the underlying mechanisms remain to be identified (Cheng, 2008).
It was also demonstrated that misoriented GSCs originate, at least in part, from dedifferentiation of spermatogonia. Although dedifferentiated GSCs have high frequency (>40%) of centrosome misorientation, they can function as stem cells by resuming the cell cycle, with correctly oriented mitotic spindles just like as constitutive GSCs. GSC numbers do not decrease as quickly as expected from the calculated GSC half-life, suggesting that a mechanism to compensate for the loss of GSCs exists. Since misoriented spindles, or symmetric stem cell division, was rarely observed, it is speculated that dedifferentiation is the major mechanism to replace stem cells over time in the Drosophila male germ line (Cheng, 2008).
A decline in GSC number in older males (day 50) was reported recently (Boyle, 2007) This decrease in stem cell number is likely due to failure of the niche function (via decreased signal from the niche as well as decreased E-cadherin-based attachment between the niche and GSCs. However, the decrease in the production of spermatogonia and testis involution precede the loss of GSCs such that decreasing GSC numbers cannot explain the testis involution that is observed at younger ages (Cheng, 2008).
The present results provide a novel mechanistic link between the control of stem cell polarity and the age-related decline in tissue regenerative capacity. Mechanisms responsible for monitoring stem cell orientation with respect to the niche not only prevent overproliferation of stem cells by ensuring the asymmetric outcome of the stem cell division, but they contribute to the decline in tissue regenerative capacity during aging. Many of the misoriented GSCs originate from the dedifferentiation of spermatogonia, a mechanism thought to be responsible for maintaining the stem cell population over extended periods of time. Therefore, although GSCs produce less progeny over time, the system appears to maximize the number of progeny produced throughout life, while maintaining asymmetric stem cell division (Cheng, 2008).
In summary, it is proposed that the GSCs with misoriented centrosome divide less frequently and that a combination of such a decreased stem cell division and a higher frequency of the GSC misorientation in aged testes leads to a decline in spermatogenesis with age (Cheng, 2008).
Spinal muscular atrophy is a severe neurogenic disease that is caused by mutations in the human survival motor neuron 1 (SMN1) gene. SMN protein is required for the assembly of small nuclear ribonucleoproteins and a dramatic reduction of the protein leads to cell death. It is currently unknown how the reduction of this ubiquitously essential protein can lead to tissue-specific abnormalities. In addition, it is still not known whether the disease is caused by developmental or degenerative defects. Using the Drosophila system, this study shows that SMN is enriched in postembryonic neuroblasts and forms a concentration gradient in the differentiating progeny. In addition to the developing Drosophila larval CNS, Drosophila larval and adult testes have a striking SMN gradient. When SMN is reduced in postembryonic neuroblasts using MARCM clonal analysis, cell proliferation and clone formation defects occur. These SMN mutant neuroblasts fail to correctly localise Miranda and have reduced levels of snRNAs. When SMN is removed, germline stem cells are lost more frequently. It was also shown that changes in SMN levels can disrupt the correct timing of cell differentiation. It is concluded that highly regulated SMN levels are essential to drive timely cell proliferation and cell differentiation (Grice, 2011).
This study shows a high demand for SMN in Drosophila stem cells. In addition, striking SMN concentration gradient, inversely proportional to the state of differentiation, has been identified in Drosophila larval CNS and testis. In Drosophila SMN mutant larvae, both the CNS and testis display growth defects which precede the previously reported motor defects and death. These larvae also fail to localise Miranda protein correctly at the basal membrane of the neuroblast. Clonal analysis indicates that SMN deficient stem cells have a reduced number of divisions and also generate cells with lower levels of U2 and U5 snRNPs. Overexpression of SMN alters the timing of CNS growth and disrupts the onset of pupariation and pupation. Using the male germline system, it was shown that prolonged SMN reduction leads to stem cell loss. Finally it was found that ectopic SMN expression in cells along the SMN gradient leads to changes in the timing of cell differentiation. It is therefore suggested that the fine-tuning of SMN levels throughout development can lead to complex developmental defects and reduce the capacity of stem cells to generate new cells in development (Grice, 2011).
SMN levels have been reported to be extremely high in early development. This study shows that SMN up-regulation occurs in neuroblasts prior to the initiation of their cell division, suggesting a distinct increase of SMN levels is required for new rounds of neurogenesis and local proliferation. Fewer immature neurons are generated in the thoracic ganglion of smn mutant MARCM clones. Provisional data has suggested there may be proliferation defects in the spinal cord of severe mouse models. In addition, a recent study using the severe SMA mouse model has shown proliferation defects in the mouse hippocampus, a region associated with higher SMN levels (Wishart, 2010). Together these data suggest that, in part, the pathology observed in more severe forms of SMA may be caused by defects in tissue growth (Grice, 2011).
Proteins involved in processes such as chromatin remodelling, histone generation and cell signalling have been identified as intrinsic factors for the maintenance of Drosophila stem cells. This is the first report of stem cell defects caused by the reduction of a protein involved in snRNP biogenesis. Although SMN is required in all cells, proper stem cell function requires a substantially higher level of SMN. This study also shows snRNP defects in Drosophila SMN mutant tissue. Previous studies in Drosophila have shown no gross changes in snRNP levels, including U2 and U5, in lysates from whole smnA and smnB mutant larvae. smnA MARCM neuroblast clones and male germline mitotic clones have reduced snRNP levels, suggesting snRNP assembly may be particularly sensitive to SMN reduction during CNS and germline development (Grice, 2011).
SMN mutant neuroblasts have abnormal Miranda localisation. Miranda, an adaptor protein, forms a complex with the RNA binding protein Staufen which binds to prospero mRNA. In addition to snRNPs, SMN protein has been implicated in the biogenesis of numerous RNP subclasses, including proteins involved in the transport and localisation of β-actin mRNA at the synapse. Whether Miranda mislocalisation is due to direct or indirect associations with SMN should be addressed (Grice, 2011).
SMN mutant larvae have been previously shown to have synaptic defects which include enlarged and fewer boutons and a reduction in the number of GluR-IIA clusters - the neurotransmitter receptor at the Drosophila neuromuscular junction. In addition, numerous developmental defects are observed including pupation and growth defects. Complementing this work, Drosophila Gemin5 a member of the Drosophila SMN-Gemin complex has been shown to interact with members of the ecdysone signalling pathway responsible for initiating pupation and growth. Drosophila Gemin5 is also enriched in pNBs, in a pattern comparable to SMN. There is increasing evidence that suggests the Drosophila SMN complex plays an important role in pupation. Ubiquitous overexpression of SMN using da-GAL4 advances CNS development and causes premature entry into pupation. The ecdysone pathway has been identified to play an important part in the regulation of neuroblast division and neuronal differentiation during development. How the Drosophila SMN complex plays a part in stem cell biology, and how the SMN complex interacts with specific signalling pathways should be the subject of further study (Grice, 2011).
Larval and adult testes exhibit the most distinct SMN gradients in Drosophila tissues. Drosophila testes have a constant population of germline stem cells that start to divide in the late larval stages and produce sperm throughout life. The removal of SMN from male germline stem cells results in stem cell loss. In the smnB mutant testis, the reduction of SMN causes a contraction of the SMN gradient towards the apical stem cells. As SMN is lost from the primary spermatoctyes, more mature sperm are observed. Increasing SMN levels leads to an increase in primary spermatocytes and a reduction in mature sperm in the adult. This result is the first to demonstrate that high SMN levels in undifferentiated cells can repress differentiation in sperm development. Interestingly, along with the CNS, Drosophila testes have the highest number of alternative splicing events and the most differentially expressed splicing factors during development. Understanding if differential expression of SMN in specific cell types controls a shift in splicing factors as cells switch from proliferation to differentiation will be the target of future study. A recent study has identified defects in gametogenesis and testis growth in mice lacking the Cajal body marker coilin, a binding partner of SMN. The authors speculated that coilin may facilitate the fidelity and timing of RNP assembly in the cell and coilin loss may limit rapid and dynamic RNA processing. It will be important to understand how SMN and coilin genetically interact in stem cells and developing tissues (Grice, 2011).
The Drosophila CNS and male germline offer two new tractable systems that can be used to study SMN biology in development and stem cells. It also offers a system to study how SMN, a protein associated with neuronal development, could cause SMA. Although SMA is classically a disease of the motor neuron, a severe reduction of SMN protein affects a wide spectrum of cells including stem cells. Consistent with this idea, symptoms in mild forms of SMA (type III or IV) are predominately limited to motor neurons. However, patients with the most severe type (type I), suffer from defects in multiple tissues including congenital heart defects, multiple contractures, bone fractures, respiratory insufficiency, or sensory neuronopathy. Elucidating the differential requirements of SMN in individual cell types, and how their sensitivity to SMN loss can mediate the disease, can contribute to the understanding of the selectivity of SMA (Grice, 2011).
Stem cells can self-renew and generate differentiating daughter cells. It is not known whether these cells maintain their epigenetic information during asymmetric division. Using a dual-color method to differentially label 'old' versus 'new' histones in Drosophila male germline stem cells (GSCs), it was shown that preexisting canonical H3, but not variant H3.3, histones are selectively segregated to the GSC, whereas newly synthesized histones incorporated during DNA replication are enriched in the differentiating daughter cell (see Experimental design and potential results). The asymmetric histone distribution occurs in GSCs but not in symmetrically dividing progenitor cells. Furthermore, if GSCs are genetically manipulated to divide symmetrically, this asymmetric mode is lost. This work suggests that stem cells retain preexisting canonical histones during asymmetric cell divisions, probably as a mechanism to maintain their unique molecular properties (Tran, 2013).
Although all cells in an organism contain the same genetic material, different genes are expressed in specific cell types, allowing them to differentiate along distinct pathways. Epigenetic mechanisms regulate gene expression and maintain a specific cell fate through many cell divisions. Stem cells have the remarkable ability to both self-renew and generate daughter cells that enter differentiation. Epigenetic mechanisms have been reported to regulate stem cell activity in multiple lineages. However, there has been little direct in vivo evidence demonstrating whether stem cells retain their epigenetic information (Tran, 2013).
The Drosophila male GSCs are well characterized in terms of their physiological location, microenvironment (i.e., niche), and cellular structures). Male GSCs can be identified precisely by their distinct anatomical positions and morphological features. A GSC usually divides asymmetrically to produce a self-renewed GSC and a daughter cell gonialblast (GB) that undergoes differentiation. Therefore, GSCs can be examined at single-cell resolution for a direct comparison (Tran, 2013).
In eukaryotes, the basic unit of chromatin called nucleosome contains histone octamer [2×(H3, H4, H2A, H2B)] and DNA wrapping around them. Indeed, histones are one of the major carriers of epigenetic information. To address how histones are distributed during the GSC asymmetric division, a switchable dual-color method was developed to differentially label 'old' versus 'new' histones that uses both spatial (by Gal4; UAS system) and temporal (by heat shock induction) controls to switch labeled histones from green [green fluorescent protein (GFP)] to red [monomeric Kusabira-Orange (mKO)]. Heat shock treatment induces an irreversible DNA recombination to shut down expression of GFP-labeled old histones and initiate expression of mKO-labeled new histones. If the old histones are partitioned nonselectively, the GFP will initially exhibit equal distribution in the GSC and GB, and will be gradually replaced by the mKO. However, if the old histones are preferentially retained in the GSCs to constitute potentially GSC-specific chromatin structure, the GFP will be detected specifically in the GSCs. During DNA replication-dependent canonical histone deposition, histones H3 and H4 are incorporated as a tetramer, and histones H2A and H2B are incorporated as dimers. Therefore, independent transgenic strains were generated for H3 and H2B, respectively. On the other hand, histone variants are incorporated into chromatin in a transcription-coupled but DNA replication-independent manner. Therefore, the histone variant H3.3 was used as a control for canonical histones (Tran, 2013).
To avoid potential complications caused by heat shock-induced DNA recombination on either one or both chromosomes in GSCs, each of the three transgenes (H3, H2B, and H3.3) was integrated as a single copy and analyzed in heterozygous flies. Examination of testes with the transgenes revealed nuclear GFP but little mKO signal before heat shock. After heat shock, mKO signals were detectable. Different GSCs undergo mitosis asynchronously, and an average cell cycle length of GSCs is approximately 12 to 16 hours. Among all GSCs, 75% to 77% are in G2 phase, 21% are in S phase, fewer than 2% are in mitosis, and G1-phase GSCs are almost negligible. Moreover, the GSC and GB arising from an asymmetric division remain connected after mitosis by a cellular structure known as the spectrosome, when they undergo the next G1 and S phases synchronously (Tran, 2013).
To examine the distribution of old versus new histones in GSC and GB after a round of DNA replication-dependent histone deposition, testes were studied 16 to 20 hours after heat shock. In particular, GSC-GB pairs connected by spectrosomes were examined. On the basis of cell cycle length of GSCs, these GSC-GB pairs were from GSCs that switched from histone-GFP to histone-mKO genetic code during their G2 phase and then underwent the first mitosis followed by G1, S, and G2 phase and the second mitosis. Within this time frame, both old histones and new histones were detectable in GSCs at the second G2 phase because new histones had been synthesized and incorporated during the first S phase. For histone H3, the GFP signal was detected primarily in the GSC but not in the GB. By contrast, the mKO signals were present in both the GSC and the GB, with a relatively higher level in the GB. The asymmetric distribution of histone H3 was specific for GSC divisions, because both the GFP and mKO signals were equally distributed in spermatogonial cells derived from a symmetric division of the GB in the same testis samples. Quantification of fluorescence intensity revealed that the old H3 (GFP-labeled) signal was more enriched in the GSC than in the GB by a factor of ~5.7, whereas new H3 (mKO-labeled) signal was more enriched in the GB than in the GSC by a factor of ~1.6. By contrast, this differential distribution of old versus new histone was not detected for symmetrically dividing spermatogonial cells (Tran, 2013).
In contrast to the asymmetric distribution pattern for the canonical histone H3, the histone variant H3.3 did not show this asymmetry during GSC divisions, by fluorescence images and by quantification. The symmetry of the histone variant H3.3 suggests that the asymmetric mode is specific for canonical histone H3 (Tran, 2013).
Fewer than 2% of all GSCs are undergoing mitosis; thus, all analyses above were based on postmitotic GSC-GB pairs. To further examine the histone segregation pattern during mitosis, a screen was carried out for mitotic GSCs. Indeed, old histones were mainly associated with the chromatids segregated to the GSC side at metaphase, anaphase, and telophase. By contrast, new histones were more enriched at the chromatids segregated to GB side). These results suggest that the sister chromatids preloaded with old histones are preferentially retained in GSCs and that the ones enriched with new histones are partitioned to GBs during GSC mitosis (Tran, 2013).
Next, the histone distribution pattern was examined during the first GSC division by recovering GSCs for 4 to 6 hours after heat shock. An asymmetric distribution pattern was also found in the GSC-GB pairs with the H3 transgene. By contrast, a symmetric distribution pattern was observed for both dividing spermatogonial cells with the H3 transgene and H3.3 during GSC division. Quantification of fluorescence intensity revealed that the old H3-GFP signal was enriched in the GSC by a factor of ~13 relative to the GB, whereas the new H3-mKO signal was enriched in the GB by a factor of ~2.4 relative to the GSC. By contrast, there was no differential distribution of the old versus new histone for the symmetrically dividing spermatogonial cells, or H3.3 during GSC division. Although an asymmetric histone distribution pattern was detected in postmitotic GSC-GB pairs, examination of the mitotic GSC at this stage did not show any asymmetry. These data suggest that the asymmetric segregation mode relies on replication-dependent histone incorporation prior to mitosis. However, the factor of >10 difference of GFP signal between GSC and GB could be contributed by faster turnover of old histones in GBs, probably as a mechanism to reset the chromatin for differentiation. By contrast, the difference of mKO in GSC and GB was less substantial, probably as a result of new histone synthesis in both cells. Furthermore, the H2B transgene showed a similar pattern to H3 after the first GSC division (Tran, 2013).
The consistent asymmetric cell divisions of GSCs could be lost under certain conditions, such as ectopic activation of the key JAK-STAT signaling pathway in the niche. It has been shown that overexpression of the JAK-STAT ligand unpaired (OE-upd) induces overpopulation of GSCs. Consistent with the loss of asymmetry in expanded GSCs, the asymmetric distribution pattern of the histone H3 was not observed in OE-upd testes 16 to 20 hours after heat shock. These results demonstrate that the asymmetric histone distribution pattern is dependent on GSC asymmetric divisions. A two-step process is proposed as the favored explanation: Old and newly synthesized histones are incorporated to different sister chromatids during S phase; then, during mitosis, the sister chromatid preloaded with old histones is preferentially segregated to GSC (Tran, 2013).
These data reveal that stem cells preserve preexisting histones through asymmetric cell divisions. The JAK-STAT signaling pathway required for the asymmetric GSC divisions contributes to the asymmetric histone distribution pattern. This work provides a critical first step toward identifying the detailed molecular mechanisms underlying old histone retention during GSC asymmetric division. These findings in the well-characterized GSC model system will facilitate understanding of how epigenetic information could be maintained by stem cells or reset in their sibling cells that undergo cellular differentiation (Tran, 2013).
Self-renewal and differentiation in germline stem cells (GSCs) are tightly regulated by the stem cell niche and via multiple approaches. In a previous study, the novel GSC regulatory gene Srlp (CG5844) was screened in Drosophila testes. However, the underlying mechanistic links between Srlp and the stem cell niche remain largely undetermined. Using genetic manipulation of the Drosophila model, this study systematically analyzed the function and mechanism of Srlp in vivo and in vitro. In Drosophila, Srlp is an essential gene that regulates the self-renewal and differentiation of GSCs in the testis. In the in vitro assay, Srlp is found to control the proliferation ability and cell death in S2 cells, which is consistent with the phenotype observed in Drosophila testis. Furthermore, results of the liquid chromatography-tandem mass spectrometry (LC-MS/MS) reveal that RpL6 binds to Srlp. Srlp also regulates the expression of spliceosome and ribosome subunits and controls spliceosome and ribosome function via RpL6 signals. Collectively, these findings uncover the genetic causes and molecular mechanisms underlying the stem cell niche. This study provides new insights for elucidating the pathogenic mechanism of male sterility and the formation of testicular germ cell tumor (Yu, 2019).
Many stem cells utilize asymmetric cell division (ACD) to produce a self-renewed stem cell and a differentiating daughter cell. How non-genic information could be inherited differentially to establish distinct cell fates is not well understood. This study reports a series of spatiotemporally regulated asymmetric components, which ensure biased sister chromatid attachment and segregation during ACD of Drosophila male germline stem cells (GSCs). First, sister centromeres are differentially enriched with proteins involved in centromere specification and kinetochore function. Second, temporally asymmetric microtubule activities and polarized nuclear envelope breakdown allow for the preferential recognition and attachment of microtubules to asymmetric sister kinetochores and sister centromeres. Abolishment of either the asymmetric sister centromeres or the asymmetric microtubule activities results in randomized sister chromatid segregation. Together, these results provide the cellular basis for partitioning epigenetically distinct sister chromatids during stem cell ACDs, which opens new directions to study these mechanisms in other biological contexts (Ranjan, 2019).
Although the evolutionary history of the X chromosome indicates its specialization in male fitness, its role in spermatogenesis has largely been unexplored. Currently only three X chromosome genes are considered of moderate-definitive diagnostic value. This study aimed to provide a comprehensive analysis of all X chromosome-linked protein-coding genes in 2,354 azoospermic/cryptozoospermic men from four independent cohorts. Genomic data were analyzed and compared with data in normozoospermic control individuals and gnomAD. While updating the clinical significance of known genes, it is proposed that 21 recurrently mutated genes strongly associated with and 34 moderately associated with azoospermia/cryptozoospermia not previously linked to male infertility (novel). The most frequently affected prioritized gene, RBBP7, was found mutated in ten men across all cohorts, and functional studies in Drosophila support its role in germ stem cell maintenance. Collectively, this study represents a significant step towards the definition of the missing genetic etiology in idiopathic severe spermatogenic failure and significantly reduces the knowledge gap of X-linked genetic causes of azoospermia/cryptozoospermia contributing to the development of future diagnostic gene panels (Riera-Escamilla, 2022).
Wolbachia, a vertically transmitted endosymbiont infecting many insects, spreads rapidly through uninfected populations by a mechanism known as cytoplasmic incompatibility (CI). In CI, a paternally delivered modification of the sperm leads to chromatin defects and lethality during and after the first mitosis of embryonic development in multiple species. However, whether CI-induced defects in later stage embryos are a consequence of the first division errors or caused by independent defects remains unresolved. To address this question, this study focused on ~1/3 of embryos from CI crosses in Drosophila simulans that develop apparently normally through the first and subsequent pre-blastoderm divisions before exhibiting mitotic errors during the mid-blastula transition and gastrulation. Single embryo PCR was developed
and whole genome sequencing to find a large percentage of these developed CI-derived embryos bypass the first division defect. Using fluorescence in situ hybridization, this study found increased chromosome segregation errors in gastrulating CI-derived embryos that had avoided the first division defect. Thus, Wolbachia action in the sperm induces developmentally deferred defects that are not a consequence of the first division errors. Like the immediate defect, the delayed defect is rescued through crosses to infected females. These studies inform current models on the molecular and cellular basis of CI (Warecki, 2022).
Wolbachia are common bacteria among terrestrial arthropods. These endosymbionts transmitted through the female germline manipulate their host reproduction through several mechanisms whose most prevalent form called Cytoplasmic Incompatibility -CI- is a conditional sterility syndrome eventually favoring the infected progeny. Upon fertilization, the sperm derived from an infected male is only compatible with an egg harboring a compatible Wolbachia strain, this sperm leading otherwise to embryonic death. The Wolbachia Cif factors CidA and CidB responsible for CI and its neutralization function as a Toxin-Antitoxin system in the mosquito host Culex pipiens. However, the mechanism of CidB toxicity and its neutralization by the CidA antitoxin remain unexplored. Using transfected insect cell lines to perform a structure-function analysis of these effectors, this study shows that both CidA and CidB are chromatin interactors and CidA anchors CidB to the chromatin in a cell-cycle dependent-manner. In absence of CidA, the CidB toxin localizes to its own chromatin microenvironment and acts by preventing S-phase completion, independently of its deubiquitylase -DUB- domain. Experiments with transgenic Drosophila show that CidB DUB domain is required together with CidA during spermatogenesis to stabilize the CidA-CidB complex. This study defines CidB functional regions and paves the way to elucidate the mechanism of its toxicity (Terretaz, 2023).
Wolbachia are the most widely distributed intracellular bacteria, and their most common effect on host phenotype is cytoplasmic incompatibility (CI). A variety of models have been proposed to decipher the molecular mechanism of CI, among which the host modification (HM) model predicts that Wolbachia effectors play an important role in sperm modification. However, owing to the complexity of spermatogenesis and testicular cell-type heterogeneity, whether Wolbachia have different effects on cells at different stages of spermatogenesis or whether these effects are linked with CI remains unknown. Therefore, this study used single-cell RNA sequencing to analyse gene expression profiles in adult male Drosophila testes that were infected or uninfected by Wolbachia. Wolbachia was found to be significantly affected the proportion of different types of germ cells and affected multiple metabolic pathways in germ cells. Most importantly, Wolbachia had the greatest impact on germline stem cells, resulting in dysregulated expression of genes related to DNA compaction, and Wolbachia infection also influenced the histone-to-protamine transition in the late stage of sperm development. These results support the HM model and suggest that future studies on Wolbachia-induced CI should focus on cells in the early stages of spermatogenesis (Dou, 2023).
Homologous recombination (HR) repair is uniquely regulated in the ericentromeric heterochromatin to enable 'safe' repair while preventing aberrant recombination. In Drosophila cells, DNA double-strand breaks (DSBs) relocalize to the nuclear periphery through nuclear actin-driven directed motions before recruiting the strand invasion protein Rad51 and completing HR repair. End-joining (EJ) repair also occurs with high frequency in heterochromatin of fly tissues, but how alternative EJ (alt-EJ) pathways operate in heterochromatin remains largely uncharacterized. This study induced DSBs in single euchromatic and heterochromatic sites using a new system that combines the DR- white reporter and I-SceI expression in spermatogonia of flies. Using this approach, higher frequency of HR repair is detected in heterochromatin, relative to euchromatin. Further, sequencing of mutagenic repair junctions reveals the preferential use of different EJ pathways across distinct euchromatic and heterochromatic sites. Interestingly, synthesis-dependent microhomology-mediated end joining (SD-MMEJ) appears differentially regulated in the two domains, with a preferential use of motifs close to the cut site in heterochromatin relative to euchromatin, resulting in smaller deletions. Together, these studies establish a new approach to study repair outcomes in fly tissues, and support the conclusion that heterochromatin uses more HR and less mutagenic EJ repair relative to euchromatin (Miller, 2023).
It has been established that UBR4 encodes E3 ubiquitin ligase, which determines the specificity of substrate binding during protein ubiquitination and has been associated with various functions of the nervous system but not the reproductive system. This study explored the role of UBR4 on fertility with a Drosophila model. Different Ubr4 knockdown flies were established using the UAS/GAL4 activating sequence system. Fertility, hatchability, and testis morphology were studied, and bioinformatics analyses were conducted. The results indicated that UBR4 deficiency could induce male sterility and influent egg hatchability in Drosophila. Ubr4 deficiency affected the testis during morphological analysis. Proteomics analysis indicated 188 upregulated proteins and 175 downregulated proteins in the testis of Ubr4 knockdown flies. Gene Ontology analysis revealed significant upregulation of CG11598 and Sfp65A, and downregulation of Pelota in Ubr4 knockdown flies. These proteins were involved in the biometabolic or reproductive process in Drosophila. These regulated proteins are important in testis generation and sperm storage promotion. Bioinformatics analysis verified that expression of UBR4 was low in cryptorchidism patients, which further supported the important role of UBR4 in male fertility. Overall, these findings suggest that UBR4 deficiency could promote male infertility and may be involved in the protein modification of UBR4 by upregulating Sfp65A and CG11598, whereas downregulating Pelota protein expression (Xie, 2023).
The origin and structural evolution of de novo genes in Drosophila
Although previously thought to be unlikely, recent studies have shown that de novo gene origination from previously non-genic sequences is a relatively common mechanism for gene innovation in many species and taxa. These young genes provide a unique set of candidates to study the structural and functional origination of proteins. However, understanding of their protein structures and how these structures originate and evolve are still limited, due to a lack of systematic studies. This study combined high-quality base-level whole genome alignments, bioinformatic analysis, and computational structure modeling to study the origination, evolution, and protein structure of lineage-specific de novo genes. 555 de novo gene candidates were identified in D. melanogaster that originated within the Drosophilinae lineage. A gradual shift was found in sequence composition, evolutionary rates, and expression patterns with their gene ages, which indicates possible gradual shifts or adaptations of their functions. Surprisingly, little overall protein structural changes were found for de novo genes in the Drosophilinae lineage. Using Alphafold2, ESMFold, and molecular dynamics, a number of de novo gene candidates were identified with protein products that are potentially well-folded, many of which are more likely to contain transmembrane and signal proteins compared to other annotated protein-coding genes. Using ancestral sequence reconstruction, it was found that most potentially well-folded proteins are often born folded. Interestingly, one case was observed where disordered ancestral proteins become ordered within a relatively short evolutionary time. Single-cell RNA-seq analysis in testis showed that although most de novo genes are enriched in spermatocytes, several young de novo genes are biased in the early spermatogenesis stage, indicating potentially important but less emphasized roles of early germline cells in the de novo gene origination in testis. This study provides a systematic overview of the origin, evolution, and structural changes of Drosophilinae -specific de novo genes (Peng,2023).
Quantitative proteomics and phosphoproteomics reveal insights into mechanisms of ocnus function in Drosophila testis development
Testis has the largest number of proteins and tissue-specific proteins in animals. Knockdown of ocnus (ocn), a testis-specific gene, has been shown to result in much smaller testis with no germ cells in Drosophila. Through iTRAQ quantitative proteomics sequencing, 606 proteins were identified as having a significant and at least a 1.5-fold change in expression after ocn knockdown in fly testes; 85 were up-regulated and 521 were down-regulated. Among the differential expressed proteins (DEPs) were proteins that affected generation of precursor metabolites and energy, metabolic process, and mitochondrial transport. Protein-protein interaction (PPI) analyses of DEPs showed that several kinases and/or phosphatases interacted with Ocn. qRT-PCR confirmed 12 genes appeared in both DEGs and DEPs were significantly down-regulated after ocn knockdown in testes. Furthermore, 153 differentially expressed phosphoproteins (DEPPs), including 72 up-regulated and 94 down-regulated ones were also identified. In addition to those DEPPs associated with spermatogenesis, the other DEPPs were enriched in actin filament-based process, protein folding, and mesoderm development. Some DEPs and DEPPs were involved in Notch, JAK/STAT, and cell death pathways. The differences in protein abundance in the ocn knockdown flies might not necessarily be the direct result of differential gene regulation due to the inactivation of ocn. Nevertheless, these results suggest that the expression of ocn is essential for Drosophila testis development and that its down-regulation disturbs key signaling pathways related to cell survival and differentiation (Zheng, 2023).
Ribosomal DNA (rDNA) loci contain hundreds of tandemly repeated copies of ribosomal RNA genes needed to support cellular viability. This repetitiveness makes it highly susceptible to copy number (CN) loss due to intrachromatid recombination between rDNA copies, threatening multigenerational maintenance of rDNA. How this threat is counteracted to avoid extinction of the lineage has remained unclear. This study shows that the rDNA-specific retrotransposon R2 is essential for restorative rDNA CN expansion to maintain rDNA loci in the Drosophila male germline. The depletion of R2 led to defective rDNA CN maintenance, causing a decline in fecundity over generations and eventual extinction. Double-stranded DNA breaks created by the R2 endonuclease, a feature of R2's rDNA-specific retrotransposition, initiate the process of rDNA CN recovery, which relies on homology-dependent repair of the DNA break at rDNA copies. This study reveals that an active retrotransposon provides an essential function for its host, contrary to transposable elements' reputation as entirely selfish. These findings suggest that benefiting host fitness can be an effective selective advantage for transposable elements to offset their threat to the host, which may contribute to retrotransposons' widespread success throughout taxa (Nelson, 2023).
Chronic exposure to warm temperature causes low sperm abundance and quality in Drosophila melanogaster Temperature influences male fertility across organisms; however, how suboptimal temperatures affect adult spermatogenesis remains understudied. In a recent study on Drosophila melanogaster oogenesis, a drastic reduction was observed in the fertility of adult males exposed to warm temperature (29s°C). This study showss that males become infertile at 29 °sC because of low sperm abundance and quality. The low sperm abundance at 29 °C does not stem from reduced germline stem cell or spermatid numbers, as those numbers remain comparable between 29 °C and control 25 °C. Notably, males at cold 18 °C and 29 °C had similarly increased frequencies of spermatid elongation and individualization defects which, considering the high sperm abundance and male fertility measured at 18 °C, indicate that spermatogenesis has a high tolerance for elongation and individualization defects. Interestingly, the abundance of sperm at 29 ° C decreases abruptly and with no evidence of apoptosis as they transition into the seminal vesicle near the end of spermatogenesis, pointing to sperm elimination through an unknown mechanism. Finally, sperm from males at 29 °C fertilize eggs less efficiently and do not support embryos past the first stage of embryogenesis, indicating that poor sperm quality is an additional cause of male infertility at 29 °C (O'Neill, 2023).
Branched actin networks are critical in many cellular processes, including cell motility and division. Arp2, a protein within the 7-membered Arp2/3 complex, is responsible for generating branched actin. Given its essential roles, Arp2 evolves under stringent sequence conservation throughout eukaryotic evolution. This study unexpectedly discovered recurrent evolutionary diversification of Arp2 in Drosophila, yielding independently arising paralogs Arp2D in obscura species and Arp2D2 in montium species. Both paralogs are unusually testis-enriched in expression relative to Arp2. Whether their sequence divergence from canonical Arp2 led to functional specialization was investigated by replacing Arp2 in D. melanogaster with either Arp2D or Arp2D2. Despite their divergence, it was surprisingly found that both complement Arp2's essential function in the soma, suggesting they have preserved the ability to polymerize branched actin even in a non-native species. However, it was found that Arp2D -expressing males are subfertile and display many defects throughout sperm development. Two highly diverged structural regions in Arp2D were pinpointed that contribute to these defects: subdomain 2 and the C-terminus. It was expected that germline function would be rescued by replacing Arp2D's long and charged C-terminus with Arp2's short C-terminus, yet surprisingly, the essential somatic function of Arp2D was lost. Therefore, while Arp2D's structural divergence is incompatible with D. melanogaster sperm development, its unique C-terminus has evolved a critical role in actin polymerization. These findings suggest canonical Arp2's function differs between somatic versus germline contexts, and Arp2 paralogs have recurrently evolved and specialized for actin branching in the testis (Stromberg, 2023).
Antimony (Sb), is thought to induce testicular toxicity, although this remains controversial. This study investigated the effects of Sb exposure during spermatogenesis in the Drosophila testis and the underlying transcriptional regulatory mechanism at single-cell resolution. Firstly, it was found that flies exposed to Sb for 10 days led to dose-dependent reproductive toxicity during spermatogenesis. Protein expression and RNA levels were measured by immunofluorescence and quantitative real-time PCR (qRT-PCR). Single-cell RNA sequencing (scRNA-seq) was performed to characterize testicular cell composition and identify the transcriptional regulatory network after Sb exposure in Drosophila testes. scRNA-seq analysis revealed that Sb exposure influenced various testicular cell populations, especially in GSCs_to_Early_Spermatogonia and Spermatids clusters. Importantly, carbon metabolism was involved in GSCs/early spermatogonia maintenance and positively related with SCP-Containing Proteins, S-LAPs, and Mst84D signatures. Moreover, Seminal Fluid Proteins, Mst57D, and Serpin signatures were highly positively correlated with spermatid maturation. Pseudotime trajectory analysis revealed three novel states for the complexity of germ cell differentiation, and many novel genes (e.g., Dup98B) were found to be expressed in state-biased manners during spermatogenesis. Collectively, this study indicates that Sb exposure negatively impacts GSC maintenance and spermatid elongation, damaging spermatogenesis homeostasis via multiple signatures in Drosophila testes and therefore supporting Sb-mediated testicular toxicity (Yu, 2023).
Transcriptome analysis of several animal clades suggests that male reproductive tract gene expression evolves quickly. However, the factors influencing the abundance and distribution of within-species variation, the ultimate source of interspecific divergence, are poorly known. Drosophila melanogaster, an ancestrally African species that has recently spread throughout the world and colonized the Americas in the last roughly 100 years, exhibits phenotypic and genetic latitudinal clines on multiple continents, consistent with a role for spatially varying selection in shaping its biology. Nevertheless, geographic expression variation in the Americas is poorly described, as is its relationship to African expression variation. This study investigate these issues through analysis of two male reproductive tissue transcriptomes (testis and accessory gland) in samples from Maine (USA), Panama, and Zambia. Dramatic differences were found between these tissues in differential expression between Maine and Panama, with the accessory glands exhibiting abundant expression differentiation and the testis exhibiting very little. Latitudinal expression differentiation appears to be influenced by selection on Panama expression phenotypes. While the testis shows little latitudinal expression differentiation, it exhibits much greater differentiation than the accessory gland in Zambia vs. American population comparisons. Expression differentiation for both tissues is non-randomly distributed across the genome on a chromosome arm scale. Interspecific expression divergence between D. melanogaster and D. simulans is discordant with rates of differentiation between D. melanogaster populations. Strongly heterogeneous expression differentiation across tissues and timescales suggests a complex evolutionary process involving major temporal changes in the way selection influences expression evolution in these organs (Cridland, 2023).
In lepidopteran insects, dichotomous spermatogenesis produces eupyrene spermatozoa, which are nucleated, and apyrene spermatozoa, which are anucleated. Both sperm morphs are essential for fertilization, as eupyrene sperm fertilize the egg, and apyrene sperm is necessary for the migration of eupyrene sperm. In Drosophila, Prmt5 acts as a type II arginine methyltransferase that catalyzes the symmetrical dimethylation of arginine residues in the RNA helicase Vasa. Prmt5 is critical for the regulation of spermatogenesis, but Vasa is not. To date, functional genetic studies of spermatogenesis in the lepidopteran model Bombyx mori has been limited. In this study, mutations were engineered in BmPrmt5 and BmVasa through CRISPR/Cas9-based gene editing. Both BmPrmt5 and BmVasa loss-of-function mutants had similar male and female sterility phenotypes. Through immunofluorescence staining analysis, it was found that the morphs of sperm from both BmPrmt5 and BmVasa mutants have severe defects, indicating essential roles for both BmPrmt5 and BmVasa in the regulation of spermatogenesis. Mass spectrometry results identified that R35, R54, and R56 of BmVasa were dimethylated in WT while unmethylated in BmPrmt5 mutants. RNA-seq analyses indicate that the defects in spermatogenesis in mutants resulted from reduced expression of the spermatogenesis-related genes, including BmSxl (see Drosophila Sxl), implying that BmSxl acts downstream of BmPrmt5 and BmVasa to regulate apyrene sperm development. These findings indicate that BmPrmt5 and BmVasa constitute an integral regulatory module essential for spermatogenesis in B. mori (Yang, 2023).
Ageing is a complex biological process that is accompanied by changes in gene expression and mutational load. In many species, including humans, older fathers pass on more paternally derived de novo mutations; however, the cellular basis and cell types driving this pattern are still unclear. To explore the root causes of this phenomenon, this study performed single-cell RNA sequencing on testes from young and old male Drosophila and genomic sequencing (DNA sequencing) on somatic tissues from the same flies. Early germ cells from old and young flies were found to enter spermatogenesis with similar mutational loads but older flies are less able to remove mutations during spermatogenesis. Mutations in old cells may also increase during spermatogenesis. These data reveal that old and young flies have distinct mutational biases. Many classes of genes show increased postmeiotic expression in the germlines of older flies. Late spermatogenesis-biased genes have higher dN/dS (ratio of non-synonymous to synonymous substitutions) than early spermatogenesis-biased genes, supporting the hypothesis that late spermatogenesis is a source of evolutionary innovation. Surprisingly, genes biased in young germ cells show higher dN/dS than genes biased in old germ cells. These results provide new insights into the role of the germline in de novo mutation (Witt, 2023).
Mitochondria are dynamic organelles that undergo frequent remodeling to accommodate developmental needs. This study describes a striking organization of mitochondria into a large ball-like structure adjacent to the nucleus in premeiotic Drosophila melanogaster spermatocytes, which we term "mitoball". Mitoballs are transient structures that colocalize with the endoplasmic reticulum, Golgi bodies, and the fusome. Similar premeiotic mitochondrial clusters were observed in a wide range of insect species, including mosquitos and cockroaches. Through a genetic screen, Milton, an adaptor protein that links mitochondria to microtubule-based motors, mediates mitoball formation. Flies lacking a 54 amino acid region in the C terminus of Milton completely lacked mitoballs, had swollen mitochondria in their spermatocytes, and showed reduced male fertility. It is suggested that the premeiotic mitochondrial clustering is a conserved feature of insect spermatogenesis that supports sperm development (Li, 2023).
Spermatogenesis is a critical part of reproduction in insects; however, its molecular mechanism is still largely unknown. This study identified a testis-specific gene CG3526 in Drosophila melanogaster. Bioinformatics analysis showed that CG3526 contains a zinc binding domain and 2 C(2) H(2) type zinc fingers, and it is clustered to the vertebrate really interesting new gene (RING) family E3 ubiquitin-protein ligases. When CG3526 was knocked down by RNA interference (RNAi), the testis became much smaller in size, and the apical tip exhibited a sharp and thin end instead of the blunt and round shape in the control testis. More importantly, compared to the control flies, only a few mature sperm were present in the seminal vesicle of C587-Gal4 > CG3526 RNAi flies. Immunofluorescence staining of the testis from CG3526 RNAi flies showed that the homeostasis of testis stem cell niche was disrupted, cell distribution in the apical tip was scattered, and the process of spermatogenesis was not completed. Furthermore, it was found that the phenotype of CG3526 RNAi flies' testis was similar to that of testis of Stat92E RNAi flies, the expression level of CG3526 was significantly downregulated in the Stat92EF06346 mutant flies, and the promoter activity of CG3526 was upregulated by STAT92E. Taken together, these results indicated that CG3526 is a downstream effector gene in the JAK-STAT signaling pathway that plays a key role in the spermatogenesis of Drosophila (Hu, 2023).
Centrosomes are multi-protein organelles that function as microtubule (MT) organizing centers (MTOCs), ensuring spindle formation and chromosome segregation during cell division. Centrosome structure includes core centrioles that recruit pericentriolar material (PCM) that anchors γ-tubulin to nucleate MTs. In Drosophila melanogaster, PCM organization depends on proper regulation of proteins like Spd-2, which dynamically localizes to centrosomes and is required for PCM, γ-tubulin, and MTOC activity in brain neuroblast (NB) mitosis and male spermatocyte (SC) meiosis. Some cells have distinct requirements for MTOC activity due to differences in characteristics like cell size or whether they are mitotic or meiotic. How centrosome proteins achieve cell-type-specific functional differences is poorly understood. Previous work identified alternative splicing and binding partners as contributors to cell-type-specific differences in centrosome function. Gene duplication, which can generate paralogs with specialized functions, is also implicated in centrosome gene evolution, including cell-type-specific centrosome genes. To gain insight into cell-type-specific differences in centrosome protein function and regulation, this study investigated a duplication of Spd-2 in Drosophila willistoni, which has Spd-2A (ancestral) and Spd-2B (derived). Spd-2A functions in NB mitosis, whereas Spd-2B functions in SC meiosis. Ectopically expressed Spd-2B accumulates and functions in mitotic NBs, but ectopically expressed Spd-2A failed to accumulate in meiotic SCs, suggesting cell-type-specific differences in translation or protein stability. This failure to accumulate and function in meiosis was mapped to the C-terminal tail domain of Spd-2A, revealing a novel regulatory mechanism that can potentially achieve differences in PCM function across cell types (Rusan 2023).
PIWI-interacting RNA (piRNA) pathways control transposable elements (TEs) and endogenous genes, playing important roles in animal gamete formation. However, the underlying piRNA biogenesis mechanisms remain elusive. This study shows that endogenous protein coding sequences (CDSs), which are normally used for translation, serve as origins of noncoding piRNA biogenesis in Drosophila melanogaster testes. The product, namely, CDS-piRNAs, formed silencing complexes with Aubergine (Aub) in germ cells. Proximity proteome and functional analyses show that CDS-piRNAs and cluster/TE-piRNAs are distinct species occupying Aub, the former loading selectively relies on chaperone Cyclophilin 40. Moreover, Argonaute 2 (Ago2) and Dicer-2 activities were found critical for CDS-piRNA production. We provide evidence that Ago2-bound short interfering RNAs (siRNAs) and microRNAs (miRNAs) specify precursors to be processed into piRNAs. This study further demonstrates that Aub is crucial in spermatid differentiation, regulating chromatins through mRNA cleavage. Collectively, these data illustrate a unique strategy used by male germ line, expanding piRNA repertoire for silencing of endogenous genes during spermatogenesis (Iki, 2023).
CP190 protein is one of the key components of Drosophila insulator complexes, and its study is important for understanding the mechanisms of gene regulation during cell differentiation. However, Cp190 mutants die before reaching adulthood, which significantly complicates the study of its functions in imago. To overcome this problem and to investigate the regulatory effects of CP190 in adult tissues development, a conditional rescue system was designed for Cp190 mutants. Using Cre/loxP-mediated recombination, the rescue construct containing Cp190 coding sequence is effectively eliminated specifically in spermatocytes, allowing study of the effect of the mutation in male germ cells. Using high-throughput transcriptome analysis i the function of CP190 on gene expression was determined in germline cells. Cp190 mutation was found to have opposite effects on tissue-specific genes, which expression is repressed by CP190, and housekeeping genes, that require CP190 for activation. Mutation of Cp190 also promoted expression of a set of spermatocyte differentiation genes that are regulated by tMAC transcriptional complex. These results indicate that the main function of CP190 in the process of spermatogenesis is the coordination of interactions between differentiation genes and their specific transcriptional activators (Romanov, 2023).
Unique patterns of inheritance and selection on Y chromosomes lead to the evolution of specialized gene functions. Yet characterizing the function of genes on Y chromosomes is notoriously difficult. This study reports CRISPR mutants in Drosophila of the Y-linked gene, WDY, which is required for male fertility. WDY mutants produce mature sperm with beating tails that can be transferred to females but fail to enter the female sperm storage organs. The sperm tails of WDY mutants beat approximately half as fast as wild-type sperm's and that the mutant sperm do not propel themselves within the male ejaculatory duct or female reproductive tract (RT). These specific motility defects likely cause the sperm storage defect and sterility of the mutants. Regional and genotype-dependent differences in sperm motility suggest that sperm tail beating and propulsion do not always correlate. Furthermore, significant differences were observed in the hydrophobicity of key residues of a putative calcium-binding domain between orthologs of WDY that are Y-linked and those that are autosomal. Given that WDY appears to be evolving under positive selection, these results suggest that WDY 's functional evolution coincides with its transition from autosomal to Y-linked in Drosophila melanogaster and its most closely related species. Finally, it was shown that mutants for another Y-linked gene, PRY, also show a sperm storage defect that may explain their subfertility. In contrast to WDY, PRY mutants do swim in the female RT, suggesting they are defective in yet another mode of motility, navigation, or a necessary interaction with the female RT. Overall, this study provides direct evidence for the long-held presumption that protein-coding genes on the Drosophila Y regulate sperm motility (Hafezi, 2023).
Maintenance of the Drosophila male germline stem cells (GSCs) requires activation of the Janus kinase/signal transducer and activators of transcription (JAK/STAT) pathway by niche signals. The precise role of JAK/STAT signaling in GSC maintenance, however, remains incompletely understood. This study shows that, GSC maintenance requires both canonical and non-canonical JAK/STAT signaling, in which unphosphorylated STAT (uSTAT) maintains heterochromatin stability by binding to heterochromatin protein 1 (HP1). GSC-specific overexpressing STAT, or even the transcriptionally inactive mutant STAT, increases GSC number and partially rescues the GSC-loss mutant phenotype due to reduced JAK activity. Furthermore, it was found that both HP1 and STAT are transcriptional targets of the canonical JAK/STAT pathway in GSCs, and that GSCs exhibit higher heterochromatin content. These results suggest that persistent JAK/STAT activation by niche signals leads to the accumulation of HP1 and uSTAT in GSCs, which promote heterochromatin formation important for maintaining GSC identity. Thus, the maintenance of Drosophila GSCs requires both canonical and non-canonical STAT functions within GSCs for heterochromatin regulation (Xing, 2023).
The GAGA protein (also known as GAF) is a transcription factor encoded by the Trl gene in D. melanogaster. GAGA is involved in the regulation of transcription of many genes at all stages of fly development and life. Recently, the participation of GAGA in spermatogenesis was studied and it was discovered that Trl mutants experience massive degradation of germline cells in the testes. Trl underexpression induces autophagic death of spermatocytes, thereby leading to reduced testis size. This study aimed to determine the role of the transcription factor GAGA in the regulation of ectopic germline cell death. How Trl underexpression affects gene expression in the testes was examined. 15,993 genes were identified in three biological replicates of an RNA-seq analysis, and transcript levels were compared between hypomorphic Trl (R85)/Trl (362) and Oregon testes. A total of 2,437 differentially expressed genes were found, including 1,686 upregulated and 751 downregulated genes. At the transcriptional level, the development of cellular stress was detected in the Trl-mutant testes: downregulation of the genes normally expressed in the testes (indicating slowed or abrogated spermatocyte differentiation) and increased expression of metabolic and proteolysis-related genes, including stress response long noncoding RNAs. Nonetheless, in the Flybase Gene Ontology lists of genes related to cell death, autophagy, or stress, there was no enrichment with GAGA-binding sites. Furthermore, no specific GAGA-dependent cell death pathway was identified that could regulate spermatocyte death. Thus, these data suggest that GAGA deficiency in male germline cells leads to an imbalance of metabolic processes, impaired mitochondrial function, and cell death due to cellular stress (Fedorova, 2023).
Spermatogenesis is an important process in reproduction and is conserved across species, but in Bombyx mori, it shows peculiarities, such as the maintenance of spermatogonia by apical cells and fertilization by dimorphic spermatozoa. This study attempted to characterize the genes expressed in the testis of B. mori, focusing on aspects of expression patterns and gene function by transcriptome comparisons between different tissues, internal testis regions, and Drosophila melanogaster. The transcriptome analysis of 12 tissues of B. mori, including those of testis, revealed the widespread gene expression of 20,962 genes and 1705 testis-specific genes. A comparative analysis of the stem region (SR) and differentiated regions (DR) of the testis revealed 4554 and 3980 specific-enriched genes, respectively. In addition, comparisons with D. melanogaster testis transcriptome revealed homologs of 1204 SR and 389 DR specific-enriched genes that were similarly expressed in equivalent regions of Drosophila testis. Moreover, gene ontology (GO) enrichment analysis was performed for SR-specific enriched genes and DR-specific enriched genes, and the GO terms of several biological processes were enriched, confirming previous findings. This study advances understanding of spermatogenesis in B. mori and provides an important basis for future research, filling a knowledge gap between fly and mammalian studies (Kakino, 2023).
Although the biological utilities of endogenous RNAi (endo-RNAi) have been largely elusive, recent studies reveal its critical role in the non-model fruitfly Drosophila simulans to suppress selfish genes, whose unchecked activities can severely impair spermatogenesis. In particular, hairpin RNA (hpRNA) loci generate endo-siRNAs that suppress evolutionary novel, X-linked, meiotic drive loci. The consequences of deleting even a single hpRNA (Nmy) in males are profound, as such individuals are nearly incapable of siring male progeny. In this study, comparative genomic analyses of D. simulans and D. melanogaster mutants of the core RNAi factor dcr-2 reveal a substantially expanded network of recently-emerged hpRNA-target interactions in the former species. The de novo hpRNA regulatory network in D. simulans provides insight into molecular strategies that underlie hpRNA emergence and their potential roles in sex chromosome conflict. In particular, these data support the existence of ongoing rapid evolution of Nmy/Dox-related networks, and recurrent targeting of testis HMG-box loci by hpRNAs. Importantly, the impact of the endo-RNAi network on gene expression flips the convention for regulatory networks, since strong derepression of targets of the youngest hpRNAs was observed, but only mild effects on the targets of the oldest hpRNAs. These data suggest that endo-RNAi are especially critical during incipient stages of intrinsic sex chromosome conflicts, and that continual cycles of distortion and resolution may contribute to speciation (Vedanayagam, 2023).
Meiotic drive loci distort the normally equal segregation of alleles, which benefits their own transmission even in the face of severe fitness costs to their host organism. However, relatively little is known about the molecular identity of meiotic drivers, their strategies of action, and mechanisms that can suppress their activity. This study presents data from the fruitfly Drosophila simulans that address these questions. A family of de novo, protamine-derived X-linked selfish genes (the Dox gene family: Distorter on the X-which is found on the X chromosome and kills Y chromosome-bearing sperm) was demonstrated to be silenced by a pair of newly emerged hairpin RNA (hpRNA) small interfering RNA (siRNA)-class loci, Nmy and Tmy. In the w[XD1] genetic background, knockout of nmy derepresses Dox and MDox in testes and depletes male progeny, whereas knockout of tmy causes misexpression of PDox genes and renders males sterile. Importantly, genetic interactions between nmy and tmy mutant alleles reveal that Tmy also specifically maintains male progeny for normal sex ratio. The Dox loci are functionally polymorphic within D. simulans, such that both nmy-associated sex ratio bias and tmy-associated sterility can be rescued by wild-type X chromosomes bearing natural deletions in different Dox family genes. Finally, using tagged transgenes of Dox and PDox2, the first experimental evidence is provided that Dox family genes encode proteins that are strongly derepressed in cognate hpRNA mutants. Altogether, these studies support a model in which protamine-derived drivers and hpRNA suppressors drive repeated cycles of sex chromosome conflict and resolution that shape genome evolution and the genetic control of male gametogenesis (Vedanayagam, 2023).
Although RNA is found in the seminal fluid of diverse organisms, it is unknown whether this RNA is functional within females. This study develop an experimental proteomic method called VESPA (Variant Enabled SILAC Proteomic Analysis) to test the hypothesis that Drosophila male seminal fluid RNA is translated by females. Strong evidence was found for 67 male-derived, female-translated proteins (mdFTPs) in female lower reproductive tracts at six hours postmating, many with predicted functions relevant to reproduction. Gene knockout experiments indicate that genes coding for mdFTPs play diverse roles in postmating interactions, with effects on fertilization efficiency, and the formation and persistence of the insemination reaction mass, a trait hypothesized to be involved in sexual conflict. These findings advance understanding of reproduction by revealing a novel mechanism of postmating molecular interactions between the sexes that strengthens and extends male influences on reproductive outcomes in previously unrecognized ways. Given the diverse species known to carry RNA in seminal fluid, this discovery has broad significance for understanding molecular mechanisms of cooperation and conflict during reproduction (Matzkin, 2023).
Ejaculate proteins are key mediators of post-mating sexual selection and sexual conflict, as they can influence both male fertilization success and female reproductive physiology. However, the extent and sources of genetic variation and condition dependence of the ejaculate proteome are largely unknown. Such knowledge could reveal the targets and mechanisms of post-mating selection and inform about the relative costs and allocation of different ejaculate components, each with its own potential fitness consequences. This study used liquid chromatography coupled with tandem mass spectrometry to characterize the whole-ejaculate protein composition across 12 isogenic lines of Drosophila melanogaster that were reared on a high- or low-quality diet. New proteins were discovered in the transferred ejaculate, and their origin in the male reproductive system was inferred. It was further found that the ejaculate composition was mainly determined by genotype identity and genotype-specific responses to larval diet, with no clear overall diet effect. Nutrient restriction increased proteolytic protein activity and shifted the balance between reproductive function and RNA metabolism. These results open new avenues for exploring the intricate role of genotypes and their environment in shaping ejaculate composition, or for studying the functional dynamics and evolutionary potential of the ejaculate in its multivariate complexity (Zeender, 2023).
In species with internal fertilization, sperm, and seminal fluid are transferred from male to female during mating. While both sperm and seminal fluid contain various types of molecules, including RNA, the role of most of these molecules in the coordination of fertilization or in other possible functions is poorly understood. In Drosophila, exosomes from the accessory gland, which produces seminal fluid, are transferred to females, but their potential cargoes have not been described. Moreover, while the RNA composition of sperm has been described in several mammalian species, little work on this problem has occurred in Drosophila. This study used single nucleotide polymorphism differences between males and females from a set of highly inbred lines of D. melanogaster, and transcriptome data from the female reproductive tract, sperm, testis, and accessory gland, to investigate the potential origin, male vs female, RNA molecules isolated from 3 female reproductive tract organs, the seminal receptacle and spermatheca, which store sperm, and the parovaria, which does not. Mated females were found to carry male-derived transcripts from many genes, including those that are markers of the accessory gland and known seminal fluid proteins. These observations also support the idea that intact sperm transcripts can be isolated from the female sperm storage organs (Cridland, 2023).
In Drosophila and other insects, the seminal fluid proteins (SFPs) and male sex pheromones that enter the female with sperm during mating are essential for fertility and induce profound post-mating effects on female physiology and behavior. The SFPs in D. melanogaster and other taxa include several members of the large gene family known as odorant binding proteins (Obps). This study used RNAi and CRISPR/Cas9 generated mutants to test the role of the seven seminal Obps in D. melanogaster fertility and the post-mating response (PMR). Obp56g was found to be required for male fertility and the induction of the PMR, whereas the other six genes had no effect on fertility when mutated individually. Obp56g is expressed in the male's ejaculatory bulb, an important tissue in the reproductive tract that synthesizes components of the mating plug. This study found males lacking Obp56g fail to form a mating plug in the mated female's reproductive tract, leading to ejaculate loss and reduced sperm storage. The evolutionary history of these seminal Obp genes was examined, as several studies have documented rapid evolution and turnover of SFP genes across taxa. Extensive lability in gene copy number and evidence were found of positive selection acting on two genes, Obp22a and Obp51a. Comparative RNAseq data from the male reproductive tract of multiple Drosophila species revealed that Obp56g shows high male reproductive tract expression only in species of the melanogaster and obscura groups, though conserved head expression in all species tested. Together, these functional and expression data suggest that Obp56g may have been co-opted for a reproductive function over evolutionary time (Brown, 2023).
Tissue homeostasis often requires a properly placed niche to support stem cells. Morphogenetic processes that position a niche are just being described. For the Drosophila testis, recent work showed that pro-niche cells, specified at disparate positions during early gonadogenesis, must assemble into one collective at the anterior of the gonad. Slit and FGF signals emanating from adjacent visceral mesoderm regulate assembly. In response to signaling, niche cells express islet, which was found to be also required for niche assembly. Without signaling, niche cells specified furthest from the anterior are unable to migrate, remaining dispersed. The function of such niches is severely disrupted, with niche cells evading cell cycle quiescence, compromised in their ability to signal the incipient stem cell pool, and failing to orient stem cell divisions properly. This work identifies both extrinsic signaling and intrinsic responses required for proper assembly and placement of the testis niche (Anllo, 2022).
Adult stem cells are maintained in niches, specialized microenvironments that regulate their self-renewal and differentiation. In the adult Drosophila testis stem cell niche, somatic hub cells produce signals that regulate adjacent germline stem cells (GSCs) and somatic cyst stem cells (CySCs). Hub cells are normally quiescent, but after complete genetic ablation of CySCs, they can proliferate and transdifferentiate into new CySCs. This study found that Epidermal growth factor receptor (EGFR) signaling is upregulated in hub cells after CySC ablation and that the ability of testes to recover from ablation is inhibited by reduced EGFR signaling. In addition, activation of the EGFR pathway in hub cells is sufficient to induce their proliferation and transdifferentiation into CySCs. It is proposed that EGFR signaling, which is normally required in adult cyst cells, is actively inhibited in adult hub cells to maintain their fate but is repurposed to drive stem cell regeneration after CySC ablation (Greenspan, 2022).
Whereas stem and progenitor cells proliferate to maintain tissue homeostasis, fully differentiated cells exit the cell cycle. How cell identity and cell-cycle state are coordinated during differentiation is still poorly understood. The Drosophila testis niche supports germline stem cells and somatic cyst stem cells (CySCs). CySCs give rise to post-mitotic cyst cells, providing a tractable model to study the links between stem cell identity and proliferation. While cell-cycle progression was shown to be required for CySC self-renewal, the E2f1/Dp transcription factor is dispensable for self-renewal but instead must be silenced by the Drosophila retinoblastoma homolog, Rbf, to permit differentiation. Continued E2f1/Dp activity inhibits the expression of genes important for mitochondrial activity. Furthermore, promoting mitochondrial biogenesis rescues the differentiation of CySCs with ectopic E2f1/Dp activity but not their cell-cycle exit. In sum, E2f1/Dp coordinates cell-cycle progression with stem cell identity by regulating the metabolic state of CySCs (Sainz de la Maza, 2022).
Specialized microenvironments, or niches, provide signaling cues that regulate stem cell behavior. In the Drosophila testis, the JAK-STAT signaling pathway regulates germline stem cell (GSC) attachment to the apical hub and somatic cyst stem cell (CySC) identity. This study demonstrates that chickadee, the Drosophila gene that encodes profilin, is required cell autonomously to maintain GSCs, possibly facilitating localization or maintenance of E-cadherin to the GSC-hub cell interface. Germline specific overexpression of Adenomatous Polyposis Coli 2 (APC2) rescued GSC loss in chic hypomorphs, suggesting an additive role of APC2 and F-actin in maintaining the adherens junctions that anchor GSCs to the niche. In addition, loss of chic function in the soma resulted in failure of somatic cyst cells to maintain germ cell enclosure and overproliferation of transit-amplifying spermatogonia (Shields, 2014).
Chickadee, the only Drosophila profilin homolog, is required cell intrinsically for GSC maintenance in the testis. As profilin is a regulator of actin filament polymerization and filamentous actin (F-actin) plays a crucial role in the development and stabilization of cadherin-catenin-mediated cell-cell adhesion, profilin likely maintains attachment of Drosophila male GSCs to the hub through its effect on F-actin, which concentrates at the hub-GSC interface where localized adherens junctions anchor GSCs to hub cells. It is proposed that profilin-dependent stabilization of F-actin at the GSC cortex next to the hub may help localize E-cadherin and APC2 to the junctional region. E-cadherin and APC2 in turn may recruit β-catenin/Armadillo, stabilizing the adherens junctions that attach GSCs to the hub. Chickadee may thus facilitate maintenance of GSCs through a cascade of interactions leading to localization and/or retention of both E-cadherin and β-catenin at the hub-GSC interface (Shields, 2014).
E-cadherin plays a crucial role in maintaining hub-GSC attachment. GSC clones mutant for E-cadherin are not maintained. In addition, germline overexpression of E-cadherin delayed GSC loss in stat-depleted GSCs. The results indicate that profilin function is required in GSCs for proper localization of E-cadherin to the hub-GSC interface. Several studies have shown that the actin cytoskeleton plays a crucial role in assembly and stability of adherens junctions. A favored model in the field is that actin filaments indirectly anchor and reinforce E-cadherin-mediated cell junctions by forming an intracellular scaffold for E-cadherin molecules. Indeed, binding to F-actin stabilized E-cadherin and promoted its clustering. Furthermore, the actin cytoskeleton participates in proper localization of E-cadherin molecules to cell-cell contacts. In chic/profilin mutant GSCs, disruption of actin polymerization at the cell cortex leading to local F-actin disorganization may destabilize E-cadherin and reduce its ability to localize to the GSC-hub junction, form clusters and build adequate adherens junctions (Shields, 2014).
Destabilization of E-cadherin may contribute to the mislocalization of APC2 seen in chic mutant GSCs, as E-cadherin recruits APC2 to cortical sites in GSCs. Raising possibilities of a more direct link, actin filaments have been shown to be required for association of APC2 with adherens junctions in the Drosophila embryo and ovary. Treatment of embryos with actin-depolymerizing drugs resulted in complete delocalization of APC2 from adhesive zones and diffuse APC2 staining throughout the cell. Moreover, in ovaries of chic1320/chic221 females, APC2 was substantially delocalized from the plasma membranes of nurse cells and their ring canals, and increased levels of cytoplasmic APC2 staining were observed. Similarly, this study found that APC2 was delocalized from the hub-GSC interface in larval testes of chic11/chic1320 hypomorphs (Shields, 2014).
In several studies, delocalization of APC2 from junctional membranes correlated with detachment of β-catenin/Armadillo from adherens junctions. APC2 co-localizes with Armadillo and E-cadherin at adherens junctions of Drosophila epithelial cells, nurse cells in Drosophila ovaries and at the hub-GSC interface in Drosophila testes. Disruption of APC2 function resulting in significant reduction of junctional APC2 was accompanied by delocalization of junctional Armadillo and increased levels of free cytoplasmic Armadillo in embryonic epithelial cells and ovaries. In a previous study, which used chic1320/chic221 strong loss-of-function mutants, the delocalizing effect on junctional Armadillo was variable, presumably due to incomplete penetrance of chic mutant effects. Although this study did not observe significant disruption in Armadillo staining along the hub-GSC interface of testes from chic hypomorphs, this may be due to incomplete penetrance. In addition, the Armadillo protein detected could be localized to the cortex of hub cells rather than GSCs (Shields, 2014).
The finding that germline specific overexpression of APC2 in chic11/chic1320 hypomorphs partially rescued GSC loss is consistent with a previously proposed model that actin filaments shuttle APC2 to adherens junctions and APC2 in turn recruits cytoplasmic Armadillo to junctional membranes, reinforcing the adherens junctions. It is possible that in chic11/chic1320 hypomorphs, residual actin filaments associated with adherens junctions between the hub and GSCs are sufficient to shuttle the increased amounts of cytoplasmic APC2 to adherens junctions. This APC2 may in turn recruit free cytoplasmic Armadillo to the hub-GSC interface, locally stabilizing the adherens junctions and anchoring GSCs to their niche. Notably, however, germline specific overexpression of APC2 in testes of strong loss-of-function chic1320/chic221 mutants failed to rescue GSC loss. Thus either, adequate levels of actin filament polymerization may be required for the proposed translocation of junctional proteins to the plasma membrane, or APC2 function/localization may not be the only or even the major cell-autonomous target of profilin function important for maintaining GSCs. Indeed, loss of APC2 function did not lead to GSC loss. It is suggested that the localized cortical F-actin underlying adherens junctions at the GSC-hub interface, best candidate for the most direct target of chic function, strongly stabilizes adherens junctions between GSCs and the hub, with high levels of cortical APC2 able to in part make up for weak chic function by also stabilizing adherens junctions (Shields, 2014).
Maintenance of hub-GSC attachment appears to be a key role of STAT in GSCs. The finding that STAT binds to a site near the upstream promoter of the chic gene raises the possibility that STAT might foster GSC attachment to the hub in part by ensuring high levels of transcription of profilin in GSCs. However, activation of STAT is clearly not the only regulatory influence on profilin expression as profilin is an essential gene expressed in many cell types, including those in which STAT is not active or detected. It is likely that transcription factors other than STAT turn on profilin expression in many cell types and that STAT acts along with other regulators to reinforce profilin expression in GSCs and CySCs. Conversely, overexpression of profilin was not sufficient to re-establish attachment of stat-depleted GSCs, suggesting that STAT probably regulates a number of genes to ensure that GSCs remain within the stem cell niche (Shields, 2014).
Loss of chic function in somatic cyst cells impaired the ability of cyst cells to build and/or maintain the cytoplasmic extensions through which they embrace and enclose spermatogonial cysts. Two somatic cyst cells normally surround each gonialblast and enclose its mitotic and meiotic progeny throughout Drosophila spermatogenesis. The cyst cells co-differentiate with the germ cells they enclose. Several lines of evidence support the model that either the ability of somatic cyst cells to enclose germ cells or their ability to send signals to adjacent germ cells is important to restrict proliferation and promote differentiation of germ cells. In either case, activation of EGFR in cyst cells is required for cyst cells to enclose germ cells and/or send the signals for germ cells to differentiate. The similarities in phenotype between loss of chic function and loss of EGFR activation in somatic cyst cells raise the possibility that chic/profilin may act downstream of activated EGFR to modulate the actin cytoskeleton for the remodeling of cyst cells to form or maintain the cytoplasmic extensions that enclose germ cells. Indeed, activated EGFR is known in other systems to tyrosine phosphorylate phospholipase C-γ1 (PLC-γ1), a soluble enzyme in quiescent cells like daughter cyst cells, activating it to catalyze hydrolysis of the membrane lipid phosphatidylinositol 4,5-bisphosphate (PIP2), which binds profilin protein with high affinity, which inhibits the interaction between profilin and actin. The hydrolysis of PIP2 by activated PLC-γ1 results in localized release of profilin and other actin-binding proteins, enabling them to interact with actin and participate in cytoskeletal rearrangement and membrane protrusion. Thus, based on biochemical analysis in other systems, a link between EGFR activation and profilin leading to local remodeling of the actin cytoskeleton is plausible in somatic cyst cells, although it remains to be directly tested (Shields, 2014).
In many tissues, the stem cell niche must coordinate behavior across multiple stem cell lineages. How this is achieved is largely unknown. This study has identified delayed completion of cytokinesis in germline stem cells (GSCs) as a mechanism that regulates the production of stem cell daughters in the Drosophila testis. Through live imaging, a secondary F-actin ring was shown to form through regulation of Cofilin activity to block cytokinesis progress after contractile ring disassembly. The duration of this block is controlled by Aurora B kinase. Additionally, a requirement was identified for somatic cell encystment of the germline in promoting GSC abscission. It is suggested that this non-autonomous role promotes coordination between stem cell lineages. These findings reveal the mechanisms by which cytokinesis is inhibited and reinitiated in GSCs and why such complex regulation exists within the stem cell niche (Lenhart, 2015).
This first real-time analysis of GSCs through abscission has revealed surprising complexities layered in cytokinesis. First, cytokinesis is blocked after central spindle and contractile ring disassembly and before entry to the abscission phase. This block is imposed by a secondary F-actin-ring. Second, AurB regulates the transition between phase one and phase two. That transition marks a vital step in the reinitiation of cytokinesis, permitting cytoplasmic isolation and recruitment of abscission machinery. Finally, somatic cell encystment is essential to abscission. Thus, three discrete nodes of regulation are layered on top of the canonical cytokinesis program to achieve tight temporal control over daughter cell production, and thus tissue maintenance by the resident stem cells (Lenhart, 2015).
Incomplete cytokinesis is a deeply conserved feature of germ cells that establishes the syncytium necessary for robust germline development. Differentiating germ cells appear to arrest cytokinesis immediately following contractile ring ingression because the known components of stable ring canals are identical to those of the contractile ring. It was thought that delayed cytokinesis in GSCs was simply a remnant of this conserved program. In contrast, this study found that the delay is mechanistically distinct from that occurring in differentiating germ cells. GSCs complete ingression, disassemble their contractile ring F-actin, and dissolve central spindle microtubules before engaging a ROK-LimK-Cofilin pathway to regulate a secondary F-actin ring that blocks cytokinesis progression until its disassembly at the entry to phase two (Lenhart, 2015).
Interestingly, the F-actin rings of gonial cells were not disrupted by manipulation of Cofilin activity, in contrast to their precocious disassembly in GSC-Gb pairs. This functional distinction is likely tied to the different biological goal of the stem cell versus the differentiating germ cell. One must release a differentiating daughter cell while the other must communicate syncitially for differentiation to progress normally. Ultimately, because the stem cell niche confers this functional distinction, future work will investigate whether it directly controls F-actin dynamics in the stem cell by possibly modulating Cofilin, or acts indirectly through other stem cell factors to do so (Lenhart, 2015).
These data strongly indicate that the secondary F-actin ring must be disassembled for abscission to be reinitiated. This suggests that F-actin at the IC bridge inhibits abscission, and work in other cells supports this. Inhibition of the Cofilin phosphatase, activation of AurB, depletion of phosphoinositide 5-phosphatase, or of Rab35 all lead to retention of F-actin at the IC bridge and inhibit abscission. Importantly, abscission could be restored after Rab35 depletion by forcing F-actin disassembly (Lenhart, 2015).
GSC-Gb pairs depleted for aurB fail to complete abscission prior to mitotic entry and form interconnected germ cells attached to the hub. This could suggest that AurB is normally required to promote abscission. However, expressing an activated form of Svn did not induce precocious abscission as would be expected in this model. Rather, SvnS125E expression advanced the transition from phase one to two, while aurB depletion delayed it. These reciprocal effects suggest instead that AurB times the phase one-phase two transition. In this model, the lack of abscission in aurB mutants is an indirect consequence of spending a shorter fraction of the total cycle in phase two. For example, this study has shown that ESCRTIII is localized during phase two and in the apparent absence of central spindle microtubules. In aurB-depleted cells, there simply may not be enough time during the shortened phase two for the already compromised recruitment of ESCRTIII machinery to promote abscission prior to mitotic entry. It is also noted that the lack of a central spindle raises the issue of how ESCRTIII components are delivered to the IC bridge. Perhaps the midbody performs this role, as has been suggested for the C. elegans first cell division (Lenhart, 2015).
Recent studies have found that shrub is negatively regulated by AurB in female GSCs (Matias, 2015). Although the current results suggest that AurB activity should promote ESCRTIII function in the testis, it is compelling to speculate that AurB might control the phase one-phase two transition through shrub. Alternatively, AurB could directly control this transition by regulating disassembly of the secondary F-actin ring, as there is precedent for AurB controlling actin dynamics. For example, in the 'No Cut' pathway, maintenance of AurB activity late in cytokinesis is associated with persistence of F-actin at the IC bridge. Intriguingly, AurB can phosphorylate formin proteins and thereby regulate actin stress fiber formation. Although in this context AurB activity positively regulates actin polymerization, the interaction between AurB and formin suggests a direct link between CPC activity and actin dynamics. This connection is particularly compelling given that formins can also promote severing of actin filaments. Thus, it is intriguing to speculate that AurB phosphorylation of formins at the IC bridge in GSC-Gb pairs may promote severing of actin filaments in the secondary ring and thereby promote transition from phase one to phase two of delay (Lenhart, 2015).
Perhaps most excitingly, this study has identified non-autonomous control over GSC-Gb abscission by somatic cell encystment. This sheds light on the functional relevance of abscission delay. Encystment of spermatogonia by two somatic cells is required for proper germ cell differentiation. However, GSCs and their flanking CySCs do not coordinate daughter cell production by synchronizing their cell cycles. Linking abscission to encystment is an elegant alternative for promoting coordinated release of stem cell daughters from the niche (Lenhart, 2015).
Several questions are raised by the current observations, such as precisely when abscission is triggered relative to cyst cell engulfment of the Gb. It would be necessary to carry out live imaging simultaneously on germline and adjacent somatic cells to address this. However, imaging CySCs and cyst cells is fraught with difficulty due to their irregular morphology and small size. Thus, it has not yet been possible to image somatic cells with anywhere near the resolution achieved for GSC-Gb pairs (Lenhart, 2015).
Encystment could promote abscission through contact-dependent signaling, where CySCs or cyst cells produce the ligand. Alternatively, the abscission trigger might be mechanical, because tension has been suggested to regulate abscission in cultured cells. Here, as daughter cells migrated apart in culture following mitosis, tension along the bridge connecting them increased and this lengthened the time to abscission. Experimentally decreasing bridge tension triggered earlier abscission. In the current system, most Gbs are displaced some distance from the hub during phase two, with a consequent elongation of the IC bridge connecting those cells to the GSC. Perhaps movement of the Gb away from its mother GSC generates increased tension along the bridge. Symmetric encystment might relieve that tension by providing equalizing forces on both sides of the IC bridge, inducing abscission while ensuring that the Gb is properly associated with two somatic cells. In culture, increased tension delayed abscission by disrupting assembly of functional ESCRTIII complexes at the IC bridge. Therefore, it will be interesting to address whether ESCRTIII complexes in GSCs are temporally regulated by encystment. Whatever the mechanism, the cyst cells are clearly poised for intimate contact at the appropriate time, because the midbody remnant is sometimes taken up byÊencysting somatic cells after abscission (Lenhart, 2015).
This work has clarified the mechanism by which cytokinesis is delayed in GSCs, identifying three distinct regulatory events layered on top of the traditional program of cytokinesis. These events impose an appropriate delay, a timed reinitiation, and a regulated abscission in the GSCs. This stem cell-specific program assists in the coordinate release of differentiating daughter cells from the resident stem cell populations in this niche. Because similar requirements for synchronized daughter cell production between multiple stem cell populations exist in other tissues, it is enticing to speculate that regulated abscission might be used to promote coordination in other niches. Membrane scission is difficult to demonstrate in vivo in many systems, so it is not yet known if stem cells other than the germline exhibit abscission delay. As higher resolution methods are developed to visualize stem cell dynamics within endogenous niches, it will be interesting to see if abscission delay emerges as a conserved mechanism of niche-dependent control over stem cell proliferation (Lenhart, 2015).
The tight interaction between somatic and germline cells is conserved in animal spermatogenesis. The testes of Drosophila melanogaster are the model of choice to identify processes responsible for mature gamete production. However, processes of differentiation and soma-germline interactions occurring in somatic cyst cells are currently understudied. This study focused on the comparison of transcriptome expression patterns of early and mature somatic cyst cells to find out the developmental changes taking place in them. A FACS-based approach was employed for the isolation of early and mature somatic cyst cells from fly testes, subsequent preparation of RNA-Seq libraries, and analysis of gene differential expression in the sorted cells. Increased expression was found of genes involved in cell cycle-related processes in early cyst cells, which is necessary for the proliferation and self-renewal of a crucial population of early cyst cells, cyst stem cells. Genes proposedly required for lamellipodium-like projection organization for proper cyst formation were also detected among the upregulated ones in early cyst cells. Gene Ontology and interactome analyses of upregulated genes in mature cyst cells revealed a striking over-representation of gene categories responsible for metabolic and catabolic cellular processes, as well as genes supporting the energetic state of the cells provided by oxidative phosphorylation that is carried out in mitochondria. This comparative analyses of differentially expressed genes revealed major peculiarities in early and mature cyst cells and provide novel insight into their regulation, which is important for male fertility (Adashev, 2022).
Phagoptosis is a frequently occurring nonautonomous cell death pathway in which phagocytes eliminate viable cells. While it is thought that phosphatidylserine (PS) 'eat-me' signals on target cells initiate the process, the precise sequence of events is largely unknown. This study shows that in Drosophila testes, progenitor germ cells are spontaneously removed by neighboring cyst cells through phagoptosis. Using live imaging with multiple markers, it was demonstrated that cyst cell-derived early/late endosomes and lysosomes fused around live progenitors to acidify them, before DNA fragmentation and substantial PS exposure on the germ cell surface. Furthermore, the phagocytic receptor Draper is expressed on cyst cell membranes and is necessary for phagoptosis. Significantly, germ cell death is blocked by knockdown of either the endosomal component Rab5 or the lysosomal associated protein Lamp1, within the cyst cells. These data ascribe an active role for phagocytic cyst cells in removal of live germ cell progenitors (Zohar-Fux, 2022).
Niches maintain a finite pool of stem cells via restricted space and short-range signals. Stem cells compete for limited niche resources, but the mechanisms regulating competition are poorly understood. Using the Drosophila testis model, this study showed that germline stem cells (GSCs) lacking the transcription factor Chinmo gain a competitive advantage for niche access. Surprisingly, chinmo(-/-) GSCs rely on a new mechanism of competition in which they secrete the extracellular matrix protein Perlecan to selectively evict non-mutant GSCs and then upregulate Perlecan-binding proteins to remain in the altered niche. Over time, the GSC pool can be entirely replaced with chinmo-/- cells. As a consequence, the mutant chinmo allele acts as a gene drive element; the majority of offspring inherit the allele despite the heterozygous genotype of the parent. These results suggest that the influence of GSC competition may extend beyond individual stem cell niche dynamics to population-level allelic drift and evolution (Tseng, 2022).
This work reveals an unexpected model of GSC competition that results in biased inheritance. These results provide a mechanistic demonstration of the postulated 'mitotic drive' by which GSCs with a competition advantage transmit competitive alleles at a greater than the expected 50% Mendelian ratio. Studies in plants, yeast, flies, and mice have shown that selfish genetic elements can cause gene drive through various molecular mechanisms. These include 'meiotic drivers' that co-opt meiotic divisions or that kill viable gametes that do not inherit the selfish element. 'Mitotic drive' occurs earlier in the germline lineage, at the stem cell level, and is an understudied area. Prior work in the Drosophila ovary has shown that dedifferentiation-defective female GSCs 'win' by upregulating E-Cad at the GSC-niche interface and gradually pushing WT GSCs out of the niche. Since differentiation-defective GSCs do not differentiate into gametes, this study could not examine biased allele inheritance. Female GSCs with 4x gene dose of Myc were reported to be 'winners,' but biased inheritance was not assessed (Tseng, 2022).
Given the competitive advantage of GSCs lacking chinmo, why has evolution not selected for male GSCs with no chinmo expression? Since chinmo is an essential gene required for development (Zhu et al., 2006), loss of chinmo in GSCs might cause reduced chinmo expression in other tissues, likely reducing organismal fitness. Furthermore, chinmo- dependent competition is a progressive phenotype requiring at least 2 weeks of adulthood. If males with mosaic chinmo- clones in the germline were mated as young adults, both the chinmo+ and chinmo- allele would be passed on to offspring, and this would be sufficient to maintain the chinmo+ allele in the population. Although the GSC pool in testes with chinmo- cells is often monoclonal, why is the chinmo- allele is not passed on to 100% of offspring? Males are maintained as virgins until their testes are dissected or until mating. This means that both chinmo- GFP-negative and ubi-GFP-positive spermatids are stored in the seminal vesicle throughout the male's lifetime, and the single round of mating in these experiments allows for the transmission of both types of sperm (Tseng, 2022).
This work identified three phenotypes-competition, aging, and transdifferentiation-that are dependent on ectopic Pcan expression in chinmo-deficient stem cells. This remarkable finding raises the important question of which factors regulate Chinmo expression in GSCs during adulthood and which genes are direct targets of Chinmo. This study identified chinmo as a JAK/STAT target gene, but STAT-deficient GSCs still express Chinmo protein, indicating that as-yet unidentified factors regulate Chinmo in GSCs and perhaps in other stem cells. Additionally, future molecular work will be needed to determine whether Pcan, Dg, and βPS are direct Chinmo target genes in germline and somatic stem cells (Tseng, 2022).
This work raises the possibility that other mutant stem cells can 'cheat' by resculpting their microenvironment and then ensuring their own retention in this remodeled milieu. Paternal age effect disorders (PAEs) encompass a broad spectrum of spontaneous congenital disorders and are thought to arise from rare selfish GSCs that are positively selected and clonally expanded. While the current model of PAE postulates increased proliferation of mutant GSCs as the competitive mechanism, other selfish cellular behaviors, such as the ones this study has discovered, could also be functioning in the mammalian testis. Cancer stem cells could utilize the mechanisms described in this study to colonize a tissue. Leukemic stem cells induce progressive remodeling of the bone marrow niche, and this altered niche favors the mutant stem cells while impairing normal hematopoietic stem cell residence and contributes to bone marrow fibrosis. In sum, this work raises the possibility that selfish stem cells across species cheat using the mechanisms this study has discovered in competitive GSCs in the Drosophila testis (Tseng, 2022).
Although this study found significant increases in Pcan, Dg, and βPS transcripts in GSCs deficient for Chinmo, HCR-FISH analyses of relative intensity are not strictly quantitative. As such, it cannot be ruled out that Chinmo affects other aspects of mRNA regulation, such as splicing. Because the transcriptome of chinmo-/- GSCs have not been attained at sufficient resolution and this study has not been able to perform Chinmo ChIP-seq in GSCs, it is not possible to conclude that Chinmo directly represses these genes. It will be important in the future to determine Chinmo occupancy on chromatin in GSCs and in other stem cells. Surprisingly, the niche spaces vacated by non-mutant neighbor GSCs are occupied by cyst stem cells (CySCs) and not by chinmo-/- GSCs. It is not understood why the CySCs predominate, and future studies employing ex vivo live-imaging will be important to gain insights into this process (Tseng, 2022).
From insects to mammals, oocytes and sperm develop within germline cysts comprising cells connected by intercellular bridges (ICBs). In numerous insects, formation of the cyst is accompanied by growth of the fusome-a membranous organelle that permeates the cyst. Fusome composition and function are best understood in Drosophila melanogaster: during oogenesis, the fusome dictates cyst topology and size and facilitates oocyte selection, while during spermatogenesis, the fusome synchronizes the cyst's response to DNA damage. Despite its distinct and sex-specific roles during insect gametogenesis, elucidating fusome growth and inheritance in females and its structure and connectivity in males has remained challenging. This study took advantage of advances in three-dimensional (3D) confocal microscopy and computational image processing tools to reconstruct the topology, growth, and distribution of the fusome in both sexes. In females, the experimental findings inform a theoretical model for fusome assembly and inheritance and suggest that oocyte selection proceeds through an 'equivalency with a bias' mechanism. In males, it was found that cell divisions can deviate from the maximally branched pattern observed in females, leading to greater topological variability. This work consolidates existing disjointed experimental observations and contributes a readily generalizable computational approach for quantitative studies of gametogenesis within and across species (Diegmiller, 2023).
Molano-Fernandez, M., Hickson, I. D. and Herranz, H. (2022). Cyclin E overexpression in the Drosophila accessory gland induces tissue dysplasia. Front Cell Dev Biol 10: 992253. PubMed ID: 36704199
The regulation of the cell division cycle is governed by a complex network of factors that together ensure that growing or proliferating cells maintain a stable genome. Defects in this system can lead to genomic instability that can affect tissue homeostasis and thus compromise human health. Variations in ploidy and cell heterogeneity are observed frequently in human cancers. This study examined the consequences of upregulating the cell cycle regulator Cyclin E in the Drosophila melanogaster male accessory gland. The accessory gland is the functional analog of the human prostate. This organ is composed of a postmitotic epithelium that is emerging as a powerful in vivo system for modelling different aspects of tumor initiation and progression. Cyclin E upregulation in this model was shown to be sufficient to drive tissue dysplasia. Cyclin E overexpression drives endoreplication and affects DNA integrity, which results in heterogeneous nuclear and cellular composition and variable degrees of DNA damage. Evidence is presented showing that, despite the presence of genotoxic stress, those cells are resistant to apoptosis and thus defective cells are not eliminated from the tissue. It was also shown that Cyclin E-expressing cells in the accessory gland display mitochondrial DNA aggregates that colocalize with Cyclin E protein. Together, these findings show that Cyclin E upregulation in postmitotic cells of the accessory gland organ causes cellular defects such as genomic instability and mitochondrial defects, eventually leading to tissue dysplasia. This study highlights novel mechanisms by which Cyclin E might contribute to disease initiation and progression (Molano-Fernandez, 2022).
Asymmetric cell division is utilized by a broad range of cell types to generate two daughter cells with distinct cell fates. In stem cell populations asymmetric cell division is believed to be crucial for maintaining tissue homeostasis, failure of which can lead to tissue degeneration or hyperplasia/tumorigenesis. Asymmetric cell divisions also underlie cell fate diversification during development. Accordingly, the mechanisms by which asymmetric cell division is achieved have been extensively studied, although the check points that are in place to protect against potential perturbation of the process are poorly understood. Drosophila melanogaster male germline stem cells (GSCs) possess a checkpoint, termed the centrosome orientation checkpoint (COC), that monitors correct centrosome orientation with respect to the component cells of the niche to ensure asymmetric stem cell division. The COC is the only checkpoint mechanism identified to date that specializes in monitoring the orientation of cell division in multicellular organisms. By establishing colcemid-induced microtubule depolymerization as a sensitive assay, this study examined the characteristics of COC activity and found that it functions uniquely in GSCs but not in their differentiating progeny. The COC operates in the G2 phase of the cell cycle, independently of the spindle assembly checkpoint. This study may provide a framework for identifying and understanding similar mechanisms that might be in place in other asymmetrically dividing cell types (Venkei, 2015).
Stereotypical orientation of the mitotic spindle is a widely utilized mechanism to achieve asymmetric cell division. Although considerable knowledge has accumulated regarding how spindle orientation is established, little is known about whether cells possess a mechanism that monitors successful spindle orientation. In the present study, using colcemid treatment as a sensitive assay, the nature of the COC, which is the only known orientation/polarity checkpoint in multicellular organisms, was was investigated. It was established that: (1) the COC specifically operates in GSCs, but not differentiating germ cells (GBs and SGs); (2) the COC operates in G2 phase of the cell cycle to prevent precocious entry into mitosis upon centrosome misorientation; and (3) as a checkpoint mechanism, the COC is distinct from the SAC (Venkei, 2015).
These results show that the COC is a GSC-specific checkpoint that monitors centrosome orientation and arrests cells in G2 phase when centrosomes are not correctly oriented. It remains unclear whether the COC-mediated G2 arrest might eventually undergo adaptation to allow mitotic entry, as is the case with mitotic slippage in the SAC. It is worth noting in this context that the GSC mitotic index never increased during 6 h of colcemid treatment, whereas SGs seem to undergo mitotic slippage by 6 h of colcemid treatment. This suggests that the COC-mediated G2 arrest is relatively strong. The findings that the mad2 mutation has no effect on G2 arrest of GSCs and that par-1 and cnn mutants have no effect on mitotic arrest in GBs/SGs upon colcemid treatment strongly support the notion that the COC and SAC constitute distinct checkpoint mechanisms. Although CySCs and SGs also orient their mitotic spindles, the present study shows that spindle orientation in these cells is not under the control of the COC. However, the lack of a COC in these cell types does not exclude the possibility that distinct polarity checkpoint mechanisms are in place to ensure correct spindle orientation. If the arrest points of such checkpoints are not prior to the arrest point of the SAC (i.e. metaphase), the assay using colcemid would not reveal their presence (Venkei, 2015).
Mutant analysis of multiple centrosomal components in this study revealed a selective requirement for centrosomal components in the function of the COC. Sas-4 and Cnn are crucial for COC function, whereas Spd-2 and Apc1 are not. This indicates that not all of the centrosomal proteins are involved in COC function. Conversely, not all COC components are localized to the centrosome. As shown in a previous study, an essential component of the COC, Par-1, is localized to the spectrosome, where it regulates the localization of Cyclin A to regulate mitotic entry. How the information on centrosomal orientation is communicated to the spectrosome, where Par-1 and Cyclin A localize, remains to be determined. A major outstanding question in understanding the COC is how it senses the location of the centrosome with respect to the hub cells. Previous studies have shown that the mother centrosome is anchored to the adherens junctions formed at the hub-GSC interface via MTs. Therefore, it is plausible that the COC senses aspect(s) of these interactions. It awaits future investigation to understand how the association of the centrosome with the hub-GSC interface is mechanistically sensed, and how such information is integrated with the activity of COC component(s) on the spectrosome (Venkei, 2015).
In summary, this present work establishes that the COC is a checkpoint mechanism that is distinct from the SAC and monitors correct centrosome orientation specifically in GSCs. It is speculated that a similar mechanism might be in place in other systems that rely on asymmetric cell division (Venkei, 2015).
Establishment of germline sexual identity is critical for production of male and female germline stem cells, as well as sperm versus eggs. This study identified PHD Finger Protein 7 (PHF7) as an important factor for male germline sexual identity in Drosophila. PHF7 exhibits male-specific expression in early germ cells, germline stem cells, and spermatogonia. It is important for germline stem cell maintenance and gametogenesis in males, whereas ectopic expression in female germ cells ablates the germline. Strikingly, expression of PHF7 promotes spermatogenesis in XX germ cells when they are present in a male soma. PHF7 homologs are also specifically expressed in the mammalian testis, and human PHF7 rescues Drosophila Phf7 mutants. PHF7 associates with chromatin, and both the human and fly proteins bind histone H3 N-terminal tails with a preference for dimethyl lysine 4 (H3K4me2). It is proposed that PHF7 acts as a conserved epigenetic 'reader' that activates the male germline sexual program (Yang, 2012).
Sex determination is key to sexual reproduction, and both somatic cells and germ cells need to establish sex-specific developmental fates. Germline sexual development is essential for the production of two distinct gametes, and underlies important differences in the regulation of male versus female fertility. In some species, germline stem cells are present in both males and females to sustain constant gamete production, but are regulated differently throughout development. In other species such as humans, sex-specific germ cell development produces a germline stem cell population only in males, whereas females have a much more limited capacity in making eggs. Defects in germline sexual development lead to a failure in gametogenesis, thus the study of germline sex determination is essential for understanding normal reproductive potential and treating infertility (Yang, 2012).
In some animals, such as mammals and Drosophila, the sex chromosome compositions of the soma and germline are interpreted independently, and the 'sex' of the germline must match that of the soma for proper germ cell development to occur. For example, patients with Klinefelter's Syndrome have an XXY sex chromosome constitution and are almost always infertile. These individuals develop somatically as males due to the presence of a Y chromosome but the germline suffers from severe atrophy, including the loss of premeiotic germline and germline stem cells. This is due to the presence of two X chromosomes in the germ cells, as the limited spermatogenesis in these patients is from germ cells that have lost one of the X chromosomes. In Drosophila, XX females that are somatically transformed into males exhibit a similar germline loss due to a conflict in sexual identity between the masculinized soma and XX germline. Thus, fruit flies are a valuable model organism for studying how germ cells establish a proper sexual identity by coordinating intrinsic signals and those coming from the soma (Yang, 2012).
In Drosophila, the presence of two X chromosomes promotes female somatic identity by activating an alternative splicing cascade that acts through Sex lethal (SXL) and Transformer (TRA), and ultimately leads to production of either the male or female forms of the transcription factors Doublesex (DSX) and Fruitless (FRU). DSX and FRU are responsible for virtually all sexually dimorphic somatic traits in Drosophila, with DSX being the key factor in the somatic gonad. In contrast, the germline does not determine its sex with this cascade and factors like TRA and DSX are not required in germ cells. Although SXL is required to promote female germ cell identity, its targets and mechanism of action in the germline are not known. The transcription factor OVO and the ubiquitin protease Ovarian Tumor (OTU) are also required in the female germline and thought to function upstream of SXL. Even less is known about how sexual identity is specified in male germ cells. Male germ cells receive a signal through the JAK/STAT pathway that promotes their sexual identity, but the downstream factors that are subsequently activated are not known. Similarly, how male germ cells respond to their own sex chromosome constitution is also not known (Yang, 2012).
This study reports a histone code reader, Plant Homeodomain (PHD) Finger 7 (PHF7), that acts in the Drosophila germline to promote male sexual identity. PHF7 is specifically expressed in male germ cells from early stages of development and is restricted to male germline stem cells (GSCs) and spermatogonia. Phf7 is required for GSC maintenance and proper entry into spermatogenesis. Interestingly, expression of Phf7 in female germ cells causes ablation of the female germline. Moreover, Phf7 affects sexual compatibility between germline and soma. Loss of Phf7 in XY germ cells alleviates the germline loss typically observed when XY germ cells are surrounded by a female soma, and expression of Phf7 can induce spermatogenesis in XX germ cells nurtured by male soma. These findings indicate that Phf7 is an essential factor in determining sexual development in the Drosophila germline, and suggest that activation of the male identity occurs through interaction with the germline epigenome (Yang, 2012).
The data indicate that Phf7 acts to promote a male identity in the germline. Loss of Phf7 function affected male GSC maintenance and spermatogenesis, but had no effect in females. Phf7-mutant GSCs exhibited a more female-like pattern of spectrosome localization, and male (XY) germ cells mutant for Phf7 were more compatible with a female soma than were wild-type male germ cells. Further, expression of PHF7 was able to masculinize the female germline: PHF7 expression induced apoptosis in developing XX germ cells and interacted with mutations in otu in a manner that indicates XX germ cells that express PHF7 are more male-like. Strikingly, PHF7 expression was able to induce spermatogenesis in XX germ cells when they are present in a male soma, something that XX germ cells are normally not able to do. Taken together, these results indicate that Phf7 promotes and is sufficient to induce male identity in the germline (Yang, 2012).
Sex determination is thought to be initiated early during development, and sex-specific differences in the male and female germline are first observed during embryogenesis. The data indicate that Phf7 plays a role in early germline sexual development, rather than a late role to regulate germ cell differentiation and gametogenesis. First, PHF7 expression is observed in the embryonic gonad and, in the adult, PHF7 is found in the GSCs and early gonia and disappears dramatically as gonia transition to spermatocytes. Further, forced PHF7 expression disrupts early female germ cell development, around the time when they are first forming GSCs. Expression of PHF7 after the early cystoblast stage (Bam-Gal4, UAS-Gal4) had no effect on the female germline, indicating that it can only affect early stages of female germ cell development. Phf7 mutants show defects in male GSC behavior and maintenance, and in the initial progression to form spermatocytes, but it is possible that these defects are due to even earlier problems in male sexual identity (Yang, 2012).
Germline sexual identity is determined by both the germ cell sex chromosome constitution and signals from the surrounding soma. Phf7 expression is activated in XX germ cells when in contact with a male soma and repressed in XY germ cells when contacting a female soma. However, in a female somatic environment, XY germ cells are somewhat more likely than XX germ cells to express Phf7, indicating that Phf7 may also respond to the sex chromosome constitution of the germ cells in addition to being regulated by the soma. Further, exogenous expression of Phf7 is required to promote spermatogenesis in XX germ cells when in a male soma. Thus, the Phf7 expression that is normally initiated in such germ cells by the male soma must either not be maintained, or may be insufficient to overcome the influence of the XX sex chromosome genotype (Yang, 2012).
It is likely that Phf7 is not acting alone to control male sexual identity. Phf7 mutant males are still able to undergo spermatogenesis, but at a much reduced capacity. This appears to be the null phenotype for Phf7 as ther mutants have lost significant portions of the coding sequence. Further, when PHF7 is expressed in XX germ cells present in a male soma, these germ cells can undergo spermatogenesis, but the penetrance of this phenotype is low. Interestingly, the rescue of spermatogenesis in these XX germ cells follows an 'all or nothing' pattern; either the rescue is largely complete to give full testes and sperm production, or little rescue is observed. Therefore, there appears to be a threshold that must be crossed to promote male germline sexual identity, and that once this threshold is met, those germ cells either take over the testis, or induce other germ cells to also follow the male pathway. The simplest explanation for both the incomplete block to spermatogenesis in Phf7 mutants and the incomplete rescue of spermatogenesis by Phf7 in XX males is that an additional factor (or factors) exists that promotes male identity in addition to Phf7. Such a factor could function parallel to Phf7 in a single pathway, or represent independent input regarding germline sex determination (e.g., independent signals from the soma that influence germline sex) (Yang, 2012).
PHD fingers, such as those found in PHF7, are best known for their ability to specifically bind histones that have been modified on their N-terminal tails, in particular methylated H3K4. This study shows that both Drosophila and human PHF7 can directly associate with dimethylated H3K4, indicating that PHF7 is indeed a histone code reader. It is uncommon for PHD domains to associate preferentially with H3K4me2 over H3K4me3, but this specificity has been observed previously, and is likely important for how PHF7 modulates expression of its targets. Both di- and trimethylated H3K4 are found at actively transcribed genes, but H3K4me2 is normally localized at the 5′ end of coding sequences, downstream of H3K4me3, which is near promoters. The two marks are also regulated by different demethylases. A few recent studies have started to dissect effects of H3K4me2 on gene transcription, but the exact mechanisms are not well understood. Some PHD finger proteins also contain other domains, such as those that modify histones enzymatically. This does not appear to be the case for PHF7, and the region of homology between PHF7 homologs of different species is restricted to the PHD domains. However, individual PHD fingers can bind modified histone tails independently, and it is yet unclear which PHD finger in PHF7 contacts H3K4me2 and what activities the others might have. The logic of how PHF7 is recruited to specific loci and affects chromatin structure and gene activity are interesting questions for future work (Yang, 2012).
Another point of interest is how a reader of such a common epigenetic mark would have a sex-specific role in regulating male germline identity. It has been observed that mutation of an H3K4me2 demethylase in Caenorhabditis elegans, which leads to increased dimethylation at H3K4, results in ectopic activation of male-specific germline genes. A similar mutation in Drosophila causes female germline developmental defects, which may be related to the germline atrophy observed when PHF7 expression was upregulated in female germ cells. These data are consistent with the hypothesis that H3K4me2 has a role in regulating the male germline genome. Interestingly, another germline chromatin factor, No child left behind (NCLB), has been identifed that is expressed in germ cells of both sexes but required for GSC function only in males. Thus, NCLB may cooperate with PHF7 in regulating the male GSC transcriptional program (Yang, 2012).
Based on sequence homology, orthologs of Phf7 are present in a wide range of mammalian species. Human and mouse PHF7 share extensive homology to Drosophila PHF7 throughout the N-terminus where the PHD fingers are present, and the results confirm that human PHF7 recognizes H3K4me2, similar to the fly protein. Interestingly, EST profiling indicates strong testis biases for Phf7 expression in many species, including humans, mice, rats, and dogs. Moreover, several studies that performed genome-wide RNA profiling from purified mouse germline populations indicate that mouse Phf7 expression is present in spermatogonia and is further induced in spermatocytes. Remarkably, human PHF7 was able to rescue fecundity defects in male flies mutant for Phf7. Thus, the sequence conservation observed between mammalian and Drosophila Phf7 represents true functional orthology (Yang, 2012).
As in Drosophila, germline sex determination in mouse is regulated at an early stage and is controlled by important signals from the soma, which for the mouse include retinoic acid and FGF9. Such signals regulate the timing of meiotic entry, which is different between the sexes, and also influence sex-specific programs of germline gene expression, such as expression of the key male-specific factor nanos2. Significant changes in germ cell chromatin occur during this critical time in germ cell development, including changes in the H3K4 methylation state. Thus, Phf7 represents a prime candidate for interpreting these chromatin changes in a sex-specific manner to regulate male-specific gene expression. It will be of great interest to determine whether Phf7 plays a critical role in mouse and human male germ cell development, as is proposed for Drosophila (Yang, 2012).
Doublesex (Dsx) and Fruitless (Fru) are the two downstream transcription factors that actuate Drosophila sex determination. While Dsx assists Fru to regulate sex-specific behavior, whether Fru collaborates with Dsx in regulating other aspects of sexual dimorphism remains unknown. One important aspect of sexual dimorphism is found in the gonad stem cell (GSC) niches, where male and female GSCs are regulated to create large numbers of sperm and eggs. This study reports that Fru is expressed male-specifically in the GSC niche and plays important roles in the development and maintenance of these cells. Unlike previously-studied aspects of sex-specific Fru expression, which is regulated by Transformer (Tra)-mediated alternative splicing, this study shows that male-specific expression of fru in the gonad is regulated downstream of dsx, and is independent of tra. fru genetically interacts with dsx to support maintenance of the niche throughout development. Ectopic expression of fru inhibited female niche formation and partially masculinized the ovary. fru is also required autonomously for cyst stem cell maintenance and cyst cell survival. Finally, this study identified a conserved Dsx binding site upstream of fru promoter P4 that regulates fru expression in the niche, indicating that fru is likely a direct target for transcriptional regulation by Dsx. These findings demonstrate that fru acts outside the nervous system to influence sexual dimorphism and reveal a new mechanism for regulating sex-specific expression of fru that is regulated at the transcriptional level by Dsx, rather than by alternative splicing by Tra (Zhou, 2021).
In sexually reproducing animals, the proper production of gametes and successful copulation are equally critical for reproductive success. It is therefore important that both the gonad and the brain know their sexual identity. The Doublesex/Mab-3 Related Transcription Factors act downstream of sex determination and play an evolutionarily conserved role to establish and maintain sexual dimorphism in the gonad. Meanwhile, sexual dimorphism in other tissues such as the brain is controlled, to varying degrees in different animals, through autonomous control by the sex determination and non-autonomous signaling from the gonads. In many invertebrate species, another sex-determination gene fruitless (fru), which encodes multiple BTB-Zinc finger transcription factors, plays a central role in controlling mate choice, courtship behavior and aggression. How sex determination in the gonad and the nervous system are related and coordinated in these species remains unclear (Zhou, 2021).
The founding member of the DMRT family is Drosophila doublesex (dsx). dsx and fru undergo sex-specific alternative mRNA splicing by the sex determination factor Transformer (Tra), together with its co-factor Transformer-2 (Tra-2), to produce transcripts encoding sex-specific protein isoforms. It was once thought that dsx controls sexual dimorphism outside the nervous system while fru regulates sex-specific nervous system development and behavior. But more recent evidence shows that dsx cooperates with fru to specify sex-specific neural circuitry and regulate courtship behaviors. However, whether fru acts along with dsx to control sexual dimorphism outside the nervous system remains unknown (Zhou, 2021).
The fru gene locus contains a complex transcription unit with multiple promoters and alternative splice forms (see Fruitless is expressed male-specifically in the germline stem cell niche and is independent of FruM). Sex-specific regulation of fru was only known to occur through alternative splicing of transcripts produced from the P1 promoter, which produces the FruM isoforms. The downstream promoters (P2-P4) produce Fru isoforms (collectively named FruCom) encoded by transcripts that are common to both sexes and are required for viability in both males and females. fru P1 transcripts have only been detected in the nervous system, suggesting that sex-specific functions of fru are limited to neural tissue. However, FruCom is expressed in several non-neural tissues, including sex-specific cell types of the reproductive system. Further, from a recent genome-wide search for putative Dsx targets, fru was identified as a candidate for transcriptional regulation by Dsx. These data raise the possibility that fru functions cooperatively with dsx to regulate gonad development (Zhou, 2021).
Over the past decades, much effort has been focused on understanding the functions of fru in regulating sex-specific behaviors, yet it remained unclear whether fru plays a role in regulating sexual dimorphism outside the nervous system. The work presented in this study demonstrates that Fru is expressed male-specifically in the gonad stem cell niche, and is required for CySC maintenance, cyst cell survival, and for the maintenance of the hub during larval development. Further, male-specific expression of Fru in the gonad is independent of the previously described mechanism of sex-specific alternative splicing by Tra, and is instead dependent on dsx. fru appears to be a direct target for transcriptional regulation by Dsx. This work provides evidence that fru regulates sex-specific development outside the nervous system and alters traditional thinking about the structure of the Drosophila sex determination pathway (Zhou, 2021).
While it was previously reported that fru is expressed in tissues other than the nervous system, including in the gonad, a function for fru outside the nervous system was previously unknown. This study found that Fru is expressed in the developing and adult testis in the hub, the CySC, and the early developing cyst cells. Importantly, it was found that fru is important for the proper function of these cells (Zhou, 2021).
Fru is not expressed at the time of hub formation during embryogenesis, but expression is initiated during the L2/L3 larval stage. This correlates with a time period when the hub must be maintained and resist transforming into female niche structures: in dsx mutants, all gonads in XX and XY animals develop hubs, but in half of each, hubs transform into terminal filament cells and cap cells. fru is not required for initial hub formation, consistent with it not being expressed at that time. fru is also not, by itself, required for hub maintenance under the conditions that were possible to assay (prior to the pupal lethality of fru null mutant animals). However, under conditions where hub maintenance is compromised by loss of dsx function, fru clearly plays a role in influencing whether a gonad will retain a hub, or transform into TF. Fru expression in dsx mutant gonads correlates with whether they formed male or female niche structures, and removing even a single allele of fru is sufficient to induce more hubs to transform into TFs. Finally, ectopic expression of Fru in females is sufficient to inhibit TF formation and partially masculinize the gonad, but does not induce hub formation. Thus, it is proposed that fru is one factor acting downstream of dsx in the maintenance of the male gonad stem cell niche, but that it acts in combination with other factors that also regulate this process (Zhou, 2021).
This study also demonstrated that fru is required for CySC maintenance and for the survival of differentiating cyst cells. Loss of fru from the CySC lineage led to rapid loss of these CySCs from the testis niche. Since precocious differentiation of CySCs or an increase in their apoptosis was not observed, these mechanisms do not appear to contribute to CySC loss. One possibility is that fru is needed for CySCs to have normal expression of adhesion proteins and compete with other stem cells for niche occupancy. It has been shown that fru regulates the Slit-robo pathway and that robo1 is a direct target of fru in the CNS. Interestingly, the Slit-Robo pathway also functions in the CySCs to modulate E-cadherin levels and control the ability of CySCs to compete for occupancy in the niche. Therefore, fru may use similar mechanisms to maintain CySC attachment to the hub. fru also influences survival in the differentiating cyst cells, as an increase in cell death was observed in these cells in fru mutants. Several reports have demonstrated that fru represses programmed cell death in the nervous system. It was further indicated that the cell death gene reaper is a putative target of Fru. Thus, fru may play a role in repressing the apoptosis of cyst cells (Zhou, 2021).
In summary, fru function is clearly important for male niche maintenance and the function of the CySCs and their differentiating progeny. This provides clear evidence that fru regulates sex-specific development in tissues other than the nervous system. Whether additional tissues are also regulated by fru remains to be determined (Zhou, 2021).
Previously, it was thought that the only mechanism by which sex-specific functions of fru were regulated was through Tra-dependent alternative splicing of the P1 transcripts. fru null alleles are lethal in both sexes and Fru proteins derived from non-P1 promoters were thought to be sex-nonspecific and not to contribute to sex determination. Thus, fru and dsx were considered as parallel branches of the sex determination pathway, each independently regulated by Tra. This study demonstrates that fru can also be regulated in a manner independent of tra and dependent on dsx, and provides evidence that fru is a direct target for transcriptional regulation by Dsx (see Proposed model of the Drosophila sex determination pathway. First, Fru expression in the testis is independent of the P1 transcript that is regulated by Tra. A P1 Gal4 reporter is not expressed in the testis and a mutation that prevents FruM expression from P1 does not affect Fru immunoreactivity in the testis. Second, in animals that simultaneously express the female form of tra (Tra on) and the male form of Dsx [XX; dsxD/Df(3R)dsx3], Fru is expressed in the male mode in the testis, demonstrating that it is regulated by dsx and not tra. Finally, an evolutionarily conserved Dsx consensus binding site upstream of the P4 promoter is required for proper expression levels of a fru P4 reporter in the testis. Together, these data demonstrate a novel mode for fru regulation by the sex determination pathway, where sex-specific expression of fru is regulated by dsx. It also means that the large number of fru transcripts that do not arise from the P1 promoter can be expressed in a sex-specific manner to contribute to sexual dimorphism (Zhou, 2021).
The male and female forms of Dsx contain the same DNA binding domain and can regulate the same target genes, but often have opposite effects on gene expression. Prior to this study, the documented Dsx targets (Yolk proteins 1 and 2, bric-a-brac and desatF), along with other proposed targets, were all expressed at higher levels in females than males. Thus, for these targets, DsxF acts as an activator and DsxM acts as a repressor (or DsxM has no role). Interestingly, fru is the first identified Dsx target that is expressed in a male-biased manner. Thus, for direct regulation of fru, DsxM would activate expression while DsxF represses. Mechanistically for Dsx, this implies that the male and female isoforms are not dedicated repressors and activators, respectively, but may be able to switch their mode of regulation in a tissue-specific or target-specific manner. Mouse DMRT1 has also been shown to regulate gene expression both as transcriptional activator and repressor. Thus, it is quite possible that bifunctional transcriptional regulation is a conserved characteristic of DMRTs (Zhou, 2021).
It is possible that dsx regulation of fru occurs in the nervous system as well, where it co-exists with direct regulation of fru alternative splicing by Tra. It was originally thought that alternative splicing of the fru P1 transcript by tra was essential for male courtship behavior. However, more recently it was found that these animals could exhibit male courtship behavior if they were simply stimulated by other flies prior to testing. Interestingly, the courtship behavior exhibited by these males was dependent on dsx. It is proposed that fru might still be essential for male courtship in these fru P1-mutants, but that sex-specific fru expression is dependent on transcriptional regulation of other fru promoters by Dsx (Zhou, 2021).
If sex-specific fru function can be regulated both through alternative splicing by Tra and through transcriptional regulation by Dsx, it raises the question of what is the relationship between these two modes of regulation? It is proposed that regulation of fru by Dsx is the more ancient version of the sex determination pathway and that additional regulation of fru by Tra evolved subsequently, through the acquisition of regulatory RNA elements in the fru P1 transcript. This model is supported by studies of fru gene structures in distantly related Dipteran species, and species of other insect orders, that illustrate the considerable variability in the organization of sequences controlling fru splicing. Further, in some insects, no evidence for alternative splicing of fru has been found, yet fru still plays an important role in males to control courtship behaviors. Finally, in the Hawaiian picture-winged group of subgenus Drosophila, the fru orthologues lack the P1 promoter, and non-P1 fru transcripts exhibit male-specific expression, similar to what is proposed for non-P1 fru transcripts in D. melanogaster. Thus, it appears that regulation of fru by dsx may be the evolutionarily more ancient mechanism for sex-specific control of fru, while Tra-dependent splicing of P1 transcripts is a more recent adaptation. More broadly, tra is not conserved in the sex determination pathway in the majority of animal groups, while homologs of Dsx, the DMRTs, are virtually universal in animal sex determination. Thus, if Fru orthologs are involved in the creation of sexual dimorphism in the body or the brain in other animals, they cannot be regulated by Tra but may be regulated by DMRTs (Zhou, 2021).
Germ cells in Drosophila melanogaster need intrinsic factors along with somatic signals to activate proper sexual programs. A key factor for male germline sex determination is PHD finger protein 7 (Phf7), a histone reader expressed in the male germline that can trigger sex reversal in female germ cells and is also important for efficient spermatogenesis. This study found that the evolutionarily novel C-terminus in Phf7 is necessary to turn on the complete male program in the early germline of D. melanogaster, suggesting that this domain may have been uniquely acquired to regulate sexual differentiation. Genes were sought that regulated by Phf7 related to sex determination in the embryonic germline by transcriptome profiling of FACS-purified embryonic gonads. One of the genes positively-regulated by Phf7 in the embryonic germline was an HP1 family member, Heterochromatin Protein 1D3 chromoshadow domain (HP1D3csd). This gene is needed for Phf7 to induce male-like development in the female germline, indicating that HP1D3csd is an important factor acting downstream of Phf7 to regulate germline masculinization (Yang, 2021).
This study investigated how Phf7 regulates sex determination in the embryonic germline, and one of the interesting findings is that the unusual C-terminus of Phf7 is necessary for its effects in germline masculinization. The N-terminus of Phf7 is a conserved module comprised of three zinc fingers, of which at least one is functionally essential, and this part of the Phf7 protein evolved from G2E3 (G2/M E3 Ubiquitin Ligase), a protein also made up of three zinc fingers. In contrast, the C-terminus of Phf7 is evolutionarily novel and is not similar to any known domains, suggesting that this domain is undergoing very rapid evolution, a feature not uncommon for factors involved in sex determination (Yang, 2021).
Previously a phylogenetic analysis of Phf7 proteins was conducted across the species tree, and surprisingly it was found that Phf7 in insects and amniotes do not share a common ancestor. Those findings with the latest results indicate that Phf7 in these two animal branches are not orthologous to each other, and that the emergence of the novel C-terminus is likely a unique event that occurred in Drosophila to regulate sexual differentiation in the germline. Recently, mouse Phf7 was demonstrated to be expressed in spermatocytes and can ubiquitinate histones to facilitate histone to protamine exchange. This shows that the expression patterns and functions of D. melanogaster and mouse Phf7 are different, albeit both acting on the male germline. These observations further suggest that the C-terminus of D. melanogaster PHF7 evolved onto an existing module of three zinc fingers, thereby creating new ways to regulating germline sexual development. This is a very interesting example that adds to the collection of diverse mechanisms in sex determination (Yang, 2021).
What does this uncommon C-terminus of PHF7 do? The two most intuitive ideas are that it acts as a transactivation domain like those found in transcription factors, or that it can recruit other effector molecules through protein-protein interactions. The former idea did not hold up when tested in S2 cells. The possibility that the Phf7 C-terminus acts as a bridge between its histone-associating N-terminus and other transcription factors or chromatin factors to alter target gene expression is an appealing one but there is currently no direct data that support this idea (Yang, 2021).
Downstream effectors of Phf7 were sought in the embryonic germline, and it was revealed that HP1D3csd is activated by Phf7 to regulate germline masculinization. Two different genetic tests were perfomed, and while both indicated that Phf7 and HP1D3csd genetically interact, there were some differences in the results. In the Phf7-induced female germline loss assay, both reduction and gain of HP1D3csd expression were found to be exacerbated the Phf7-induced phenotype. In comparison, in the spermatogenesis rescue assay, loss of one HP1D3csd copy hampered rescue whereas HP1D3csd overexpression enhanced spermatogenesis in XX germ cells. The latter experiment is a more direct assay of germline masculinization whereas germline loss can potentially be caused by secondary effects unrelated to sexual development. Therefore, it is thought the results of the spermatogenesis rescue experiments more accurately reflect the relationship between Phf7 and HP1D3csd. In addition to the transcriptome results, HP1D3csd has been identified along with Phf7 to be a part of sex-biased mechanisms in other contexts not limited to the germline. These also support the model that Phf7 and HP1D3csd function synergistically (Yang, 2021).
Phf7 regulates male germline development, and it can associate with the active histone mark methylated H3K47, but it is unclear what Phf7 then does to regulate expression of target genetic loci. H3K9 methylation has also been reported to be important for maintaining sexual differentiation programs in the Drosophila germline. The identification of HP1D3csd as an important downstream factor provides interesting new ideas regarding how the male germline program is initiated and regulated. CSDs have been shown to interact with various chromatin remodelers, thus one appealing model would be that Phf7 can activate or even recruit HP1D3csd to loci important for germline masculinization. This would in turn bring chromatin remodelers to such genes for expression activation and regulation and initiate male-development of the germline. Given the other finding in this study that the C-terminus of Phf7 is an essential part of this process, it would be very interesting to now study which of these factors interact and cooperate with one another (Yang, 2021).
Stem cells reside within specialized microenvironments, or niches, that control many aspects of stem cell behavior. Somatic hub cells in the Drosophila testis regulate the behavior of cyst stem cells (CySCs) and germline stem cells (GSCs) and are a primary component of the testis stem cell niche. The shutoff (shof) mutation, characterized by premature loss of GSCs and CySCs, was mapped to a locus encoding the evolutionarily conserved transcription factor Escargot (Esg). Hub cells depleted of Esg acquire CySC characteristics and differentiate as cyst cells, resulting in complete loss of hub cells and eventually CySCs and GSCs, similar to the shof mutant phenotype. Esg-interacting proteins were identified, and an interaction was demonstrated between Esg and the corepressor C-terminal binding protein (CtBP), which is also required for maintenance of hub cell fate. These results indicate that niche cells can acquire stem cell properties upon removal of a single transcription factor in vivo (Voog, 2014)
The homeostasis of self-renewal and differentiation in stem cells is strictly controlled by intrinsic signals and their niche. A large-scale RNA interference (RNAi) screen was conducted in Drosophila testes and 221 genes required for germline stem cell (GSC) maintenance or differentiation were identified. Knockdown of these genes in transit-amplifying spermatogonia and cyst cells further revealed various phenotypes. Complex analysis uncovered that many of the identified genes are involved in key steps of protein synthesis and degradation. A group of genes that are required for mRNA splicing and protein translation contributes to both GSC self-renewal and early germ cell differentiation. Loss of genes in protein degradation pathway in cyst cells leads to testis tumor with overproliferated germ cells. Importantly, in the Cullin 4 - Ring E3 ubiquitin ligase (CRL4) complex, multiple proteins were identified that are critical to GSC self-renewal. pic/DDB1, the linker protein of CRL4, is not only required for GSC self-renewal in flies but also for maintenance of spermatogonial stem cells (SSCs) in mice (Yu, 2016).
Stem cells are regulated both intrinsically and externally, including by signals from the local environment and distant organs. To identify genes and pathways that regulate stem-cell fates in the whole organism, a genome-wide transgenic RNAi screen was performed through ubiquitous gene knockdowns, focusing on regulators of adult Drosophila testis germline stem cells (GSCs). This study identified 530 genes that regulate GSC maintenance and differentiation. Of these, 113 selected genes were further knocked down using cell-type-specific Gal4s, and more than half were found to be external regulators, that is, from the local microenvironment or more distal sources. Some genes, for example, versatile (vers), encoding a Myb/SANT-like DNA-binding domain-containing heterochromatin protein, regulates GSC fates differentially in different cell types and through multiple pathways. It was also found that mitosis/cytokinesis proteins are especially important for male GSC maintenance. These findings provide valuable insights and resources for studying stem cell regulation at the organismal level (Liu, 2016).
How tissues adapt to varying nutrient conditions is of fundamental
importance for robust tissue homeostasis throughout the life of an
organism, but the underlying mechanisms are poorly understood. This study
shows that Drosophila testis
responds to protein starvation by eliminating transit-amplifying
spermatogonia (SG) while maintaining a reduced pool of actively
proliferating germline stem cells (GSCs). During protein starvation, SG
died in a manner that was mediated by the apoptosis of somatic cyst cells
(CCs) that encapsulated SG and regulated their development. Strikingly,
GSCs could not be maintained during protein starvation when CC-mediated SG
death was inhibited, leading to an irreversible collapse of tissue
homeostasis. The study proposes that the regulated elimination of
transit-amplifying cells is essential to preserve stem cell function and
tissue homeostasis during protein starvation (Yang, 2015).
Two broadly known characteristics of germ cells in many organisms are their development as a 'cyst' of interconnected cells and their high sensitivity to DNA damage. This study provides evidence that in the Drosophila testis, connectivity serves as a mechanism that confers to spermatogonia a high sensitivity to DNA damage. All spermatogonia within a cyst die synchronously even when only a subset of them exhibit detectable DNA damage. Mutants of the fusome, an organelle that is known to facilitate intracyst communication, compromise synchronous spermatogonial death and reduces overall germ cell death. These data indicate that a death-promoting signal is shared within the cyst, leading to death of the entire cyst. Taken together, it is proposed that intercellular connectivity supported by the fusome uniquely increases the sensitivity of the germline to DNA damage, thereby protecting the integrity of gamete genomes that are passed on to the next generation (Li, 2017).
Fatty acids are precursors of potent lipid signaling molecules. They are stored in membrane phospholipids and released by phospholipase A2 (PLA2). Lysophospholipid acyltransferases (ATs) oppose PLA2 by re-esterifying fatty acids into phospholipids, in a biochemical pathway known as the Lands Cycle. Drosophila Lands Cycle ATs oys and nes, as well as 7 predicted PLA2 genes, are expressed in the male reproductive tract. Oys and Nes are required for spermatid individualization. Individualization, which occurs after terminal differentiation, invests each spermatid in its own plasma membrane and removes the bulk of the cytoplasmic contents. This study developed a quantitative assay to measure individualization defects. Individualization is demonstrated to be sensitive to temperature and age but not to diet. Mutation of the cyclooxygenase Pxt, which metabolizes fatty acids to prostaglandins, also leads to individualization defects. In contrast, modulating phospholipid levels by mutation of the phosphatidylcholine lipase Swiss cheese (Sws) or the ethanolamine kinase Easily shocked (Eas) does not perturb individualization, nor does Sws overexpression. These results suggest that fatty acid derived signals such as prostaglandins, whose abundance is regulated by the Lands Cycle, are important regulators of spermatogenesis (Ben-David, 2015).
How cells avoid excessive caspase activity and unwanted cell death during apoptotic caspase-mediated removal of large cellular structures is poorly understood. This study investigated caspase-mediated extrusion of spermatid cytoplasmic contents in Drosophila during spermatid individualization. It was shown that a Krebs cycle component, the ATP-specific form of the succinyl-CoA synthetase β subunit (A-Sβ), binds to and activates the Cullin-3-based ubiquitin ligase (CRL3) complex required for caspase activation in spermatids. In vitro and in vivo evidence suggests that this interaction occurs on the mitochondrial surface, thereby limiting the source of CRL3 complex activation to the vicinity of this organelle and reducing the potential rate of caspase activation by at least 60%. Domain swapping between A-Sβ and the GTP-specific SCSβ (G-Sβ), which functions redundantly in the Krebs cycle, show that the metabolic and structural roles of A-Sβ in spermatids can be uncoupled, highlighting a moonlighting function of this Krebs cycle component in CRL activation (Aram, 2016). In Drosophila early post-meiotic spermatids, mitochondria undergo dramatic shaping into the Nebenkern, a spherical body with complex internal structure that contains two interwrapped giant mitochondrial derivatives. The purpose of this study was to elucidate genetic and molecular mechanisms underlying the shaping of this structure. The knotted onions (knon) gene encodes an unconventionally large testis-specific paralog of ATP synthase subunit d and is required for internal structure of the Nebenkern as well as its subsequent disassembly and elongation. Knon localizes to spermatid mitochondria and, when exogenously expressed in flight muscle, alters the ratio of ATP synthase complex dimers to monomers. By RNAi knockdown mitochondrial shaping roles were uncovered for other testis-expressed ATP synthase subunits. This study demonstrates the first known instance of a tissue-specific ATP synthase subunit affecting tissue-specific mitochondrial morphogenesis. Since ATP synthase dimerization is known to affect the degree of inner mitochondrial membrane curvature in other systems, the effect of Knon and other testis-specific paralogs of ATP synthase subunits may be to mediate differential membrane curvature within the Nebenkern (Sawyer, 2017).
Cell type-specific transcriptional programs that drive differentiation of specialized cell types are key players in development and tissue regeneration. One of the most dramatic changes in the transcription program in Drosophila occurs with the transition from proliferating spermatogonia to differentiating spermatocytes, with >3000 genes either newly expressed or expressed from new alternative promoters in spermatocytes. This study shows that opening of these promoters from their closed state in precursor cells requires function of the spermatocyte-specific tMAC complex, localized at the promoters. The spermatocyte-specific promoters lack the previously identified canonical core promoter elements except for the Inr. Instead, these promoters are enriched for the binding site for the TALE-class homeodomain transcription factors Achi/Vis and for a motif originally identified under tMAC ChIP-seq peaks. The tMAC motif resembles part of the previously identified 14-bp beta2UE1 element critical for spermatocyte-specific expression. Analysis of downstream sequences relative to transcription start site usage suggested that ACA and CNAAATT motifs at specific positions can help promote efficient transcription initiation. These results reveal how promoter-proximal sequence elements that recruit and are acted upon by cell type-specific chromatin binding complexes help establish a robust, cell type-specific transcription program for terminal differentiation (Lu, 2020).
Transcriptional regulation plays a central role in producing different cell types from the same genomic content. Throughout embryonic development, cells make and respond to cell fate decisions by turning on new transcription programs required to generate progressively more specialized cell types. Similar events drive differentiation of specialized cells from proliferating precursors in the adult stem cell lineages that maintain and repair many tissues throughout the life span. Understanding how cell type-specific transcription is achieved forms the very basis of understanding differentiation and development in multicellular organisms (Lu, 2020).
Tissue and stage-specific transcription programs are established by intricate interplay among promoter-proximal and distal DNA elements, and protein complexes that interact with them. Much recent work has focused on the role of stage or tissue-specific transcriptional activators and repressors acting upon distal enhancer elements to control the time and place of expression of developmental genes. However, evidence has emerged that variant forms of core promoter motifs and their recognition factors can play roles in cell type-specific transcription programs in certain tissues (Lu, 2020).
Several canonical core promoter motifs and promoter types have been identified and extensively studied. In Drosophila, TATA-box and/or downstream promoter element (DPE)-containing promoters tend to initiate transcription from a narrow region. The TATA box is bound by the TATA-binding protein (TBP), while the DPE is bound by certain TBP-associated factors (TAFs) in the general transcription factor TFIID to help precisely position RNA polymerase II for transcript initiation. On the other hand, promoters containing the DNA replication-related element (DRE) and/or other Ohler motifs (Ohler, 2002) tend to initiate transcription from a broad region and are thought to be associated with housekeeping genes. Recent work in Drosophila has shown that thousands of enhancers exhibit a distinct preference for one or the other of these promoter types (Zabidi, 2015; Rennie, 2018), suggesting that sequences near the transcription start site can play key roles in geneselective transcriptional regulation. However, the extent to which these canonical versus other core promoter motifs contribute to cell type-specific gene regulatory programs in differentiating cells and the molecular mechanisms by which they do so are not understood (Lu, 2020).
Male germ cell differentiation in Drosophila provides an excellent opportunity to study cell type-specific transcriptional regulation, as more than a thousand genes turn on for the first time in development when male germ cells become spermatocytes. In Drosophila, one germline stem cell normally produces a new stem cell and a gonialblast, which founds a clone of proliferating spermatogonia through four rounds of mitosis. The resulting 16 interconnected germ cells undergo a last round of DNA synthesis; then, as spermatocytes, they enter meiotic prophase. During this ~3-day period the spermatocytes express the vast majority of genes needed for later stages of male germ cell development. A recently developed heat-shock-Bam time course system (Kim, 2017) provided a way to obtain large quantities of germ cells at similar stages and greatly empowered study of the temporal events and molecular mechanisms that underlie the dramatic, cell type-specific change in transcriptional regulation that accompanies the transition from spermatogonia to spermatocyte (Lu, 2020).
Genetic and biochemical studies have identified two sets of cell type-specific proteins that regulate the spermatocyte transcription program, the tMAC complex (Beall, 2007) required for turning on most of the spermatocyte-specific gene expression program (Perezgasga, 2004; Doggett, 2011) and the testis-specific TAFs (tTAFs) required for full levels of expression of many genes in that program (Doggett, 2011; Lu, 2015). tMAC interacts physically with Achi/Vis, two highly similar TGIF-related TALE-class homeodomain proteins encoded by tandemly duplicated genes that are required for transcription in spermatocytes of tMAC-dependent genes (Ayyar, 2003; Wang, 2003). In flies null mutant for tMAC components, Achi/Vis, or tTAFs, germ cells arrest as mature spermatocytes. tMAC is a spermatocyte-specific version of the widely expressed and evolutionarily conserved MuvB core complex, which binds different protein partners, including members of the E2F, DP, Rb, and Myb protein families, to repress or activate key cell cycle and developmental genes (Sadasivam, 2013; Fischer, 2017). The tMAC complex expressed in Drosophila spermatocytes contains two proteins shared with the MuvB core: p55 Caf1 (RBBP4 in humans) and Mip40 (LIN37). tMAC also contains testis-specific paralogs of three of the other MuvB core components: Aly (paralog of Mip130 [LIN9]), Tomb (paralog of Mip120 [LIN54]), and Wuc (paralog of Lin52 [LIN52]). In addition, tMAC includes the testis-specific proteins Topi and Comr. Among the known tMAC subunits, Tomb, Topi, and Comr have predicted DNA-binding domains. Despite the importance of tMAC for turning on expression of most of the genes newly expressed in spermatocytes, the mechanism by which tMAC carries out this function is not known (Lu, 2020).
To investigate how the cell type-specific gene expression program for spermatocyte differentiation turns on, this study used RNA-seq to map transcript levels, CAGE to quantitatively map transcription start site (TSS) usage, and ATAC-seq to map chromatin accessibility as proliferating spermatogonia transition to differentiating spermatocytes. Combining these data, the promoters that turn on when germ cells become spermatocytes were shown to lack most of the canonical core promoter motifs. Instead, these promoters are enriched for the tMAC-ChIP motif and putative Achi/Vis-binding motif, and require tMAC function to become accessible and initiate active transcription as germ cells transition to the spermatocyte state. The findings from genome-wide analyses and selected reporter constructs are consistent with published results that regulatory elements sufficient for expression in spermatocytes lie close to the promoter regions in several genes expressed specifically in spermatocytes (Lu, 2020).
One of the striking findings from this study is that many genes expressed both in bam-/- mutant testes, in which spermatogonia continue to proliferate, and in 72 hr post-heat shock testes, in which many germ cells have progressed to the spermatocyte state, use an alternate promoter that turns on only in the spermatocyte-containing sample. It is possible that the conditions for productive transcription initiation are so different in spermatocytes compared with spermatogonia that many genes evolved alternative promoters to allow transcription in both cell types. Use of promoters with distinct proximal motifs bound by cell type-specific promoter-interacting factors such as are describe in this study may contribute to down-regulation of the old program as well as to turning on a new cell type-specific differentiation program. Indeed, the DRE motif, enriched at broad promoters down-regulated when spermatogonia transition to spermatocytes, is bound by the DRE-binding factor protein DREF, which recruits the TBP paralog TRF2 to facilitate transcript initiation. Because DREF protein is expressed in spermatogonia but down-regulated as spermatocytes mature, promoters that depend on the DRE motif and DREF-TRF2 may no longer express efficiently in late spermatocyte stages (Lu, 2020).
Function of the testis-specific tMAC complex is required for the vast majority of both the off-to-on gene and alternative new spermatocyte-specific promoters to become open and accessible as spermatogonia transition to the spermatocyte state. Where a new alternative spermatocyte-specific promoter is located within the stretch of DNA that wraps around the -1 nucleosome of the old spermatogonial promoter, displacement of this -1 nucleosome is likely a prerequisite for the expression of the new promoter. Together with the requirement for tMAC function for opening and expression of the new alternative promoter, this suggests that tMAC may play a role in remodeling nucleosomes. It is not known whether tMAC opens local chromatin by binding DNA that is transiently detached from nucleosomes due to loosening or breathing of nucleosomal arrays or other chromosomal events, and then holding it in an open position, or whether tMAC can bind to DNA wrapped around nucleosomes, similar to a pioneer transcription factor. The tMAC subunit Comr has a winged helix domain, and certain winged helix domain proteins are able to bind DNA on one side and allow simultaneous histone binding. Although none of the core components of tMAC have as yet been shown to have nucleosome remodeling activity, genetic analysis in C. elegans suggests that the generally expressed MMB/dREAM components (SynMuvB genes in worms) can interact with and may recruit components of the nucleosome remodeling and histone deacetylase (NuRD) complex (Lu, 2020).
It is possible that many of the alternative new promoters arose as a by-product of an ability of tMAC to promote chromatin opening and/or transcription initiation at many sites in the genome. In fact, mechanisms have evolved to keep this propensity under restraint: Recent work has showen that action of the multiple zinc finger protein Kmg expressed in spermatocytes is required to limit activity of tMAC to its normal target genes. With loss of Kmg function, tMAC binds to many additional sites in the genome that are not previously annotated promoters, with some of these being activated for transcription initiation (Kim, 2007; Lu, 2020 and references therein).
These motif and TSS analyses suggest that the off-to-on genes that express from a narrow region of efficient transcription initiation differ from the canonical, previously described narrow promoters. Among the down-regulated genes, most of the promoters with a narrow 'region of efficient transcription initiation' also had narrow total span of cap analysis gene expression (CAGE) signals, likely a result of TFIID precisely positioning Pol II for transcription initiation. In contrast, most of the off-to-on genes with narrow regions of efficient transcript initiation had a wide total span of CAGE signals. In other words, although these off-to-on promoters had many usable and permissive TSS positions, they were dominantly expressed from just a few TSSs within a narrow region, likely facilitated by the upstream and downstream promoter motifs and Inr at optimal positions (Lu, 2020).
Traditionally, there has been much focus on the importance of distal enhancer elements rather than promoters in specifying cell type-specific transcription programs. This view, however, may be somewhat biased by the intense analysis of developmental regulatory genes like even-skipped, the cell cycle regulatory phosphatase string/cdc25, or the close-range developmental signaling molecule BMP5, which are expressed and function in several disparate places in the body. Because the regionally expressed transcriptional activators and repressors that combine to establish positional identity differ in different regions of the body, it stands to reason that a gene that is expressed in different specific places, such as Pitx1 in jaw, pituitary, and pelvis, would need to use different enhancer elements to specify activation in different regions, with the regulatory input from the different enhancers perhaps feeding in to a common generic promoter. The situation and constraints may be quite different for terminal differentiation genes that are only expressed in a single tissue. In this case, as is shown in this study for differentiation of male germ cells, it may be more possible for the key regulatory sequences that specify cell type-specific transcriptional activation to be built into the core promoter and promoter-proximal regulatory sequences (Lu, 2020).
Drosophila spermatogenesis constitutes a paradigmatic system to study maintenance, proliferation, and differentiation of adult stem cell lineages. Each Drosophila testis contains 6-12 germ stem cells (GSCs) that divide asymmetrically to produce gonialblast cells that undergo four transit-amplifying (TA) spermatogonial divisions before entering spermatocyte differentiation. Mechanisms governing these crucial transitions are not fully understood. This study reports the essential role of the germline linker histone dBigH1 during early spermatogenesis. These results suggest that dBigH1 is a general silencing factor that represses Bam, a key regulator of spermatogonia proliferation that is silenced in spermatocytes. Reciprocally, Bam represses dBigH1 during TA divisions. This double-repressor mechanism switches dBigH1/Bam expression from off/on in spermatogonia to on/off in spermatocytes, regulating progression into spermatocyte differentiation. dBigH1 is also required for GSC maintenance and differentiation. These results show the critical importance of germline H1s for male GSC lineage differentiation, unveiling a regulatory interaction that couples transcriptional and translational repression (Carbonell, 2017).
Studies in Drosophila have provided important insights into the cellular pathways governing maintenance, proliferation, and differentiation of adult stem cell lineages, which is central to understanding normal tissue homeostasis and its alteration in disease. In particular, Drosophila spermatogenesis has become an ideal model system to study these questions. In the Drosophila testis, germ stem cells (GSCs) localize anterior, anchored to a niche of somatic cells (hub), and divide asymmetrically for self-renewal and to produce daughter progenitor gonialblast cells (GBs), which start the complex differentiation program that leads to the production of functional gametes. GBs are surrounded by 2 somatic cyst cells (Cs) and undergo four successive rounds of transit-amplifying (TA) mitoses with incomplete cytokinesis to produce a cyst of 16 sister spermatogonial cells that remain interconnected. Then, cysts differentiate to spermatocytes and undergo two meiotic divisions to produce 64 spermatids that develop to mature sperm cells (Carbonell, 2017).
bag-of-marbles (bam) is an important regulator of the first stages of spermatogenesis. Upon asymmetric division, daughter cells move away from the niche and escape Dpp/BMP-mediated repression. Bam expression increases during the first TA divisions, reaching a maximum at the 8-cell stage. Then, at the 16-cell stage, Bam expression decreases rapidly, TA proliferation stops, and differentiation into spermatocytes proceeds. How these crucial developmental transitions occurring during early male GSC lineage differentiation are regulated is not fully understood. Bam is a translational repressor that interacts with Bgcn (Benign gonial cell neoplasm) and Tut (tumorous testis) to repress Mei-P26 expression, establishing a regulatory feedback loop that governs spermatogonia proliferation. Bam also plays an important function in female oogenesis, in which it is repressed in GSCs by Dpp/BMP signaling and interacts with Bgcn to prevent translation of GSC maintenance factors (Carbonell, 2017).
This study report on the essential contribution of the Drosophila germline-specific linker histone H1 (dBigH1) to male GSC lineage development and differentiation. Linker H1s are intrinsic components of chromatin that interact with the nucleosome and regulate chromatin higher-order organization. In comparison to core histones, H1s are less well conserved, with most species containing several variants that play partially redundant functions. A conserved feature in metazoans is the presence of germline-specific variants that replace somatic H1s in germ cells (GCs). Vertebrates generally contain several male-specific variants (i.e., H1t, HILS1, and H1T2 in mice and humans) and one female-specific H1 (i.e., B4 in Xenopus and H1oo in mice and humans). In contrast, a single germline-specific linker histone dBigH1 exists in Drosophila, which is present in both the female and the male germline. Female-specific H1s are generally retained during early embryogenesis until zygotic genome activation (ZGA) . In this regard, in Drosophila, dBigH1 has been shown to maintain the zygotic genome silenced until ZGA is completed at cellularization, when dBigH1 is replaced by somatic dH1 (Carbonell, 2017 and references therein).
Little is known about the functions that germline-specific H1s play in GSC lineage development and differentiation. In mammals, h1t2 mutant mice show several abnormalities during spermatogenesis and have reduced fertility. Similarly, hils1 expression is reduced in men suffering from reduced sperm motility. However, h1t mutants do not show detectable abnormality or fertility defects. In females, H1oo is required for maturation of germinal-vesicle stage oocytes. Finally, in Caenorhabditis elegans, depletion of H1.1/HIS-24, which is abundant in the germline, affects GCs proliferation and differentiation and reduces fertility. This study shows that in Drosophila, dBigH1 is essential for male GSC lineage differentiation. dBigH1 and Bam form a double-repressor loop that regulates progression into spermatocyte differentiation. It study also shows that dBigH1 acts as a general repressor in spermatocytes and that dBigH1 is required for male GSC maintenance. Altogether, these results unveil the essential contribution of germline-specific linker histone H1 variants to GSC lineage development and differentiation (Carbonell, 2017).
This study shows that dBigH1 is required to silence bam, which is a master regulator of spermatogonia proliferation and differentiation. During the first three TA divisions, Mei-P26 facilitates accumulation of Bam, which reaches a maximum at the 8-cell stage. Then, Bam levels decrease and spermatogonia stop proliferation and differentiate to spermatocytes (see dBigH1 and Bam Form a Double-Repressor Loop that Regulates Entrance into Spermatocyte Differentiation). Several mechanisms are known to contribute to Bam downregulation after the 8-cell stage. High Bam levels downregulate Mei-P26 translation, establishing a regulatory feedback loop. In addition, several microRNAs have been shown to downregulate Bam translation. The current results suggest a model by which, in addition to translational regulation, dBigH1-mediated transcriptional repression is required to silence bam during spermatocyte differentiation. In the absence of dBigH1, bam is not silenced; thus, entrance to the spermatocyte differentiation program is blocked and spermatogonial cells accumulate. This accumulation is not accompanied by increased spermatogonia proliferation; since dBigH1 is absent during the TA divisions and, therefore, its depletion is not affecting Bam accumulation to reach the threshold that dictates proliferation stop. The current results indicate that in addition to bam, dBigH1 represses expression of multiple other genes in spermatocytes, suggesting that like in early embryogenesis, dBigH1 acts as a general silencing factor in spermatocytes. Altogether, these observations support a model by which dBigH1 acts after spermatogonia cease proliferation to set up the specific gene expression program that governs spermatocyte differentiation (Carbonell, 2017).
The results also show that Bam, which is an important translational repressor, downregulates dBigH1 expression during TA divisions. dBigH1 expression in TA cells decreases parallel to the progressive accumulation of Bam, being detectable in all 2-cell cysts and in some 4-cell cysts. It is not known whether Bam directly interacts with dBigH1 mRNAs. However, dBigH1 mRNAs are likely present during TA divisions, as they are detected before spermatocyte differentiation and bam-GAL4-induced dBigH1 depletion in TA cells reduces dBigH1 content, blocking spermatocyte differentiation. Moreover, Bam represses dBigH1 expression specifically in TA spermatogonial cells when it is driven by the ubiquitously active vasa promoter. Altogether, these observations suggest that Bam expression during the TA divisions inhibits dBigH1 mRNA translation. Later, when Bam levels decrease, dBigH1 translation resumes, reinforcing Bam downregulation through transcriptional silencing. The results show that this dBigH1/Bam double-repressor loop is crucial to license spermatogonia into spermatocyte differentiation. The important contribution of mechanisms that regulate mRNA translation during spermatogenesis has been extensively studied. However, the actual role of transcription regulation in these processes is not well understood. From this point of view, this work unveils a functional interaction during the early stages of spermatogenesis that integrates both translational and transcriptional regulation (Carbonell, 2017).
These results also suggest that dBigH1 is required for GSC maintenance, as shown by the strong developmental defects observed in nos > bigH1RNAi testes in which dBigH1 depletion was induced in GSCs. These defects include the lack of testes in ∼10% of cases and the drastic loss of GCs in the rest of the affected testes. bam overexpression results in GSC loss. However, the contribution of dBigH1 to GSC maintenance is not likely reflecting a role in bam repression since, in knockdown nos > bigH1RNAi testes, no derepression of a bamP-GFP reporter was observed in vasa-positive hub-attached cells that showed no detectable dBigH1 expression. In this regard, it is known that bam is actively repressed in GSCs by the DNA binding proteins PMad/Medea that are downstream effectors of Dpp/BMP signals emanating from the somatic cells of the niche. Repression imposed by specific DNA binding proteins likely prevails over transcriptional silencing induced by general repressors such as dBigH1. dBigH1 expression is constrained to the primordial GSCs early in embryogenesis, being present in somatic cells as long as their transcriptional program is not turned on. In this scenario, it is tempting to speculate that dBigH1 is required in GSCs to repress the somatic gene expression program throughout development. In this regard, its replacement by somatic dH1 during TA divisions is particularly intriguing. How this replacement takes place and what the consequences of its misregulation are remain to be determined (Carbonell, 2017).
The presence of germline-specific histone H1 is conserved in metazoans. However, to date, detailed functional analysis of their contribution to germline development and differentiation was largely missing. From this point of view, this study unveils the fundamental functions that germline-specific linker histone H1 variants play in male GSC lineage differentiation, providing further understanding of the factors and mechanisms that regulate the dramatic developmental transitions associated with spermatogenesis (Carbonell, 2017).
The generation of reactive oxygen species (ROS) widely occurs in metabolic reactions and affects stem cell activity by participating in stem cell self-renewal. However, the mechanisms of transit-amplifying (TA) spermatogonial divisions mediated by oxidative stress are not fully understood. Through genetic manipulation of Drosophila testes, this study demonstrated that CG8005 regulated TA spermatogonial divisions and redox homeostasis. Using in vitro approaches, it was shown that the knockdown of CG8005 increased ROS levels in S2 cells; the induced ROS generation was inhibited by NAC and exacerbated by H(2)O(2) pretreatments. Furthermore, the silencing of CG8005 increased the mRNA expression of oxidation-promoting factors Keap1, GstD1, and Mal-A6 and decreased the mRNA expression of antioxidant factors cnc, Gclm, maf-S, ND-42, and ND-75. The functions of the antioxidant factor cnc, a key factor in the Keap1-cnc signaling pathway was further investigated; cnc mimicked the phenotype of CG8005 in both Drosophila testes and S2 cells. These results indicated that CG8005, together with cnc, controlled TA spermatogonial divisions by regulating oxidative stress in Drosophila (Chen, 2020).
Gamete development ultimately influences animal fertility. Identifying mechanisms that direct gametogenesis, and how they deteriorate with age, may inform ways to combat infertility. Recentl work has shown that lysosomes acidify during oocyte maturation in Caenorhabditis elegans, suggesting that a meiotic switch in lysosome activity promotes female germ-cell health. Using Drosophila melanogaster, this study reports that lysosomes likewise acidify in male germ cells during meiosis. Inhibiting lysosomes in young-male testes causes E-cadherin accumulation and loss of germ-cell partitioning membranes. Notably, analogous changes occur naturally during aging; in older testes, a reduction in lysosome acidity precedes E-cadherin accumulation and membrane dissolution, suggesting one potential cause of age-related spermatocyte abnormalities. Consistent with lysosomes governing the production of mature sperm, germ cells with homozygous-null mutations in lysosome-acidifying machinery fail to survive through meiosis. Thus, lysosome activation is entrained to meiotic progression in developing sperm, as in oocytes, and lysosomal dysfunction may instigate male reproductive aging (Butsch, 2022).
In polyandrous internally fertilizing species, a multiply-mated female can use stored sperm from different males in a biased manner to fertilize her eggs. The female's ability to assess sperm quality and compatibility is essential for her reproductive success, and represents an important aspect of postcopulatory sexual selection. In Drosophila melanogaster, previous studies demonstrated that the female nervous system plays an active role in influencing progeny paternity proportion, and suggested a role for octopaminergic/tyraminergic Tdc2 neurons in this process. This study reports that inhibiting Tdc2 neuronal activity causes females to produce a higher-than-normal proportion of first-male progeny. This difference is not due to differences in sperm storage or release, but instead is attributable to the suppression of second-male sperm usage bias that normally occurs in control females. It was further shown that a subset of Tdc2 neurons innervating the female reproductive tract is largely responsible for the progeny proportion phenotype that is observed when Tdc2 neurons are inhibited globally. On the contrary, overactivation of Tdc2 neurons does not further affect sperm storage and release or progeny proportion. These results suggest that octopaminergic/tyraminergic signaling allows a multiply-mated female to bias sperm usage, and identify a new role for the female nervous system in postcopulatory sexual selection (Chen, 2022).
Membrane curvature recruits Bin-Amphiphysin-Rvs (BAR)-domain proteins and induces local F-actin assembly, which further modifies the membrane curvature and dynamics. The downstream molecular pathway in vivo is still unclear. This study shows that a tubular endomembrane scaffold supported by contractile actomyosin stabilizes the somatic cyst cell membrane folded around rigid spermatid heads during the final stages of sperm maturation in Drosophila testis. The structure resembles an actin "basket" covering the bundle of spermatid heads. Genetic analyses suggest that the actomyosin organization is nucleated exclusively by the formins - Diaphanous and Dishevelled Associated Activator of Morphogenesis (DAAM) - downstream of Rho1, which is recruited by the BAR-domain protein Amphiphysin. Actomyosin activity at the actin basket gathers the spermatid heads into a compact bundle and resists the somatic cell invasion by intruding spermatids. These observations reveal a distinct response mechanism of actin-membrane interactions, which generates a cell-adhesion-like strategy through active clamping (Kapoor, 2021).
Successful reproduction is dependent on the transfer of male seminal proteins to females upon mating. These proteins arise from secretory tissues in the male reproductive tract, including the prostate and seminal vesicles in mammals and the accessory gland in insects. Although detailed functional studies have provided important insights into the mechanisms by which accessory gland proteins support reproduction, much less is known about the molecular mechanisms that regulate their expression within this tissue. This study shows that the Drosophila HR39 nuclear receptor is required for the proper expression of most genes that encode male accessory gland proteins. Consistent with this role, HR39 mutant males are infertile. In addition, tissue-specific RNAi and genetic rescue experiments indicate that HR39 acts within the accessory glands to regulate gene expression and male fertility. These results provide new directions for characterizing the mammalian orthologs of HR39, the SF-1 and LRH-1 nuclear receptors, both of which are required for glandular secretions and reproduction. In addition, these studies provide a molecular mechanism to explain how the accessory glands can maintain the abundant levels of seminal fluid production required to support fertility (Praggastis, 2021).
Conservation of genetic toolkits in disparate phyla may help reveal commonalities in organ designs transcending their extreme anatomical disparities. A male accessory sexual organ in mammals, the prostate, for instance, is anatomically disparate from its analogous, phylogenetically distant counterpart - the male accessory gland (MAG) - in insects like Drosophila. It has not been ascertained if the anatomically disparate Drosophila MAG shares developmental parallels with those of the mammalian prostate. This study shows that the development of Drosophila mesoderm-derived MAG entails recruitment of similar genetic toolkits of tubular organs like that seen in endoderm-derived mammalian prostate. For instance, like mammalian prostate, Drosophila MAG morphogenesis is marked by recruitment of fibroblast growth factor receptor (FGFR) - a signalling pathway often seen recruited for tubulogenesis - starting early during its adepithelial genesis. A specialisation of the individual domains of the developing MAG tube, on the other hand, is marked by the expression of a posterior Hox gene transcription factor, Abd-B, while Hh-Dpp signalling marks its growth. Drosophila MAG, therefore, reveals the developmental design of a unitary bud-derived tube that appears to have been co-opted for the development of male accessory sexual organs across distant phylogeny and embryonic lineages (Kumari, 2021).
While the striking effects of seminal fluid proteins (SFPs) on females are fairly conserved among Diptera, most SFPs lack detectable homologues among the SFP repertoires of phylogenetically distant species. How such a rapidly changing proteome conserves functions across taxa is a fascinating question. However, this and other pivotal aspects of SFPs' evolution remain elusive because discoveries on these proteins have been mainly restricted to the model Drosophila melanogaster. This study provides an overview of the current knowledge on the inter-specific divergence of the SFP repertoire in Drosophila and compile the increasing amount of relevant genomic information from multiple species. Capitalizing on the accumulated knowledge in D. melanogaster, novel sets of high-confidence SFP candidates and transcription factors are presented, presumptively involved in regulating the expression of SFPs. This study also addresses open questions by performing comparative genomic analyses that failed to support the existence of many conserved SFPs shared by most dipterans and indicated that gene co-option is the most frequent mechanism accounting for the origin of Drosophila SFP-coding genes. It is hoped this update establishes a starting point to integrate further data and thus widen the understanding of the intricate evolution of these proteins (Hurtado, 2021).
Early work on de novo gene discovery in Drosophila was consistent with the idea that many such genes have male-biased patterns of expression, including a large number expressed in the testis. However, there has been little formal analysis of variation in the abundance and properties of de novo genes expressed in different tissues. This study investigated the population biology of recently evolved de novo genes expressed in the D. melanogaster accessory gland, a somatic male tissue that plays an important role in male and female fertility and the post mating response of females, using the same collection of inbred lines used previously to identify testis-expressed de novo genes, thus allowing for direct cross tissue comparisons of these genes in two tissues of male reproduction. Using RNA-seq data this study identified candidate de novo genes located in annotated intergenic and intronic sequence and determine the properties of these genes including chromosomal location, expression, abundance, and coding capacity. Generally, major differences were found between the tissues in terms of gene abundance and expression, though other properties such as transcript length and chromosomal distribution are more similar. Differences between regulatory mechanisms of de novo genes in the two tissues were also explored and how such differences may interact with selection to produce differences in D. melanogaster de novo genes expressed in the two tissues (Cridland, 2021).
How and when potential becomes restricted in differentiating stem cell daughters is poorly understood. While it is thought that signals from the niche are actively required to prevent differentiation, another model proposes that stem cells can reversibly transit between multiple states, some of which are primed, but not committed, to differentiate. In the Drosophila testis, somatic cyst stem cells (CySCs) generate cyst cells, which encapsulate the germline to support its development. CySCs were found to be maintained independently of niche self-renewal signals if activity of the PI3K/Tor pathway is inhibited. Conversely, PI3K/Tor is not sufficient alone to drive differentiation, suggesting that it acts to license cells for differentiation. Indeed, it was found that the germline is required for differentiation of CySCs in response to PI3K/Tor elevation, indicating that final commitment to differentiation involves several steps and intercellular communication. It is proposed that CySC daughter cells are plastic, that their fate depends on the availability of neighbouring germ cells, and that PI3K/Tor acts to induce a primed state for CySC daughters to enable coordinated differentiation with the germline (Yuen, 2021).
The genitalia present some of the most rapidly evolving anatomical structures in the animal kingdom, possessing a variety of parts that can distinguish recently diverged species. In the Drosophila melanogaster group, the phallus is adorned with several processes, pointed outgrowths, that are similar in size and shape between species. However, the complex three-dimensional nature of the phallus can obscure the exact connection points of each process. Previous descriptions based upon adult morphology have primarily assigned phallic processes by their approximate positions in the phallus and have remained largely agnostic regarding their homology relationships. In the absence of clearly identified homology, it can be challenging to model when each structure first evolved. This study employed a comparative developmental analysis of these processes in eight members of the melanogaster species group to precisely identify the tissue from which each process forms. The results indicate that adult phallic processes arise from three pupal primordia in all species. In some cases the same primordia generate homologous structures whereas in other cases, different primordia produce phenotypically similar but remarkably non-homologous structures. This suggests that the same gene regulatory network may have been redeployed to different primordia to induce phenotypically similar traits. These results highlight how traits diversify and can be redeployed, even at short evolutionary scales (Rice, 2021).
The rapid evolution of seminal fluid proteins (SFPs) has been suggested to be driven by adaptations to postcopulatory sexual selection (e.g. sperm competition). However, it has been recently shown that most SFPs evolve rapidly under relaxed selective pressures. Given the role of SFPs in competition for fertilization phenotypes, like the ability to transfer and store sperm and the modulation of female receptivity and ovulation, the prevalence of selectively relaxed SFPs appears as a conundrum. One possible explanation is that selection on SFPs might be relaxed in terms of protein amino acid content, but adjustments of expression are essential for post-mating function. Interestingly, there is a general lack of systematic implementation of gene expression perturbation assays to monitor their effect on phenotypes related to sperm competition. This study successfully manipulated the expression of 16 SFP encoding genes using tissue-specific knockdowns (KDs) and determined the effect of these genes' perturbation on three important post-mating phenotypes: female refractoriness to remating, defensive (P1), and offensive (P2) sperm competitive abilities in Drosophila melanogaster. These analyses show that KDs of tested SFP genes do not affect female refractoriness to remating and P2, however, most gene KDs significantly decreased P1. Moreover, KDs of SFP genes that are selectively constrained in terms of protein-coding sequence evolution have lower P1 than KDs of genes evolving under relaxed selection. These results suggest a more predominant role, than previously acknowledged, of variation in gene expression than coding sequence changes on sperm competitive ability in D. melanogaster (Patlar, 2022).
The most studied pheromone in Drosophila melanogaster, cis-vaccenyl acetate (cVA), is synthesized in the male ejaculatory bulb and transferred to the female during copulation. Combined with other chemicals, cVA can modulate fly aggregation, courtship, mating and fighting. This study explored the mechanisms underlying both cVA biosynthesis and emission in males of two wild types and a pheromonal mutant line. The effects of ageing, adult social interaction, and maternally transmitted cVA and microbes - both associated with the egg chorion - on cVA biosynthesis and emission were measured. While ageing and genotype changed both biosynthesis and emission in similar ways, early developmental exposure to maternally transmitted cVA and microbes strongly decreased cVA emission but not the biosynthesis of this molecule. This indicates that the release - but not the biosynthesis - of this sex pheromone strongly depends on early developmental context. The mechanism by which the preimaginal effects occur is unknown, but reinforces the significance of development in determining adult physiology and behaviour (Cortot, 2022).
Secretory cells in glands and the nervous system frequently package and store proteins destined for regulated secretion in dense-core granules (DCGs), which disperse when released from the cell surface. Despite the relevance of this dynamic process to diseases such as diabetes and human neurodegenerative disorders, mechanistic understanding is relatively limited, because of the lack of good cell models to follow the nanoscale events involved. This study employed the prostate-like secondary cells (SCs) of the Drosophila male accessory gland to dissect the cell biology and genetics of DCG biogenesis. These cells contain unusually enlarged DCGs, which are assembled in compartments that also form secreted nanovesicles called exosomes. Known conserved regulators of DCG biogenesis, including the small G-protein Arf1 and the coatomer complex AP-1, play key roles in making SC DCGs. Using real-time imaging, this study found that the aggregation events driving DCG biogenesis are accompanied by a change in the membrane-associated small Rab GTPases which are major regulators of membrane and protein trafficking in the secretory and endosomal systems. Indeed, a transition from trans-Golgi Rab6 to recycling endosomal protein Rab11, which requires conserved DCG regulators like AP-1, is essential for DCG and exosome biogenesis. These data allow development of a model for DCG biogenesis that brings together several previously disparate observations concerning this process and highlights the importance of communication between the secretory and endosomal systems in controlling regulated secretion (Wells, 2023).
Although most cells are mononuclear, the nucleus can exist in the form of binucleate or even multinucleate to respond to different physiological processes. The male accessory gland of Drosophila is the organ that produces semen, and its main cells are binucleate. This study observe that CTP synthase (CTPS) forms filamentous cytoophidia in binuclear main cells, primarily located at the cell boundary. In CTPSH355A, a point mutation that destroys the formation of cytoophidia, it was found that the nucleation mode of the main cells changes, including mononucleates and vertical distribution of binucleates. Although the overexpression of CTPSH355A can restore the level of CTPS protein, it will neither form cytoophidia nor eliminate the abnormal nucleation pattern. Therefore, these data indicate that there is an unexpected functional link between the formation of cytoophidia and the maintenance of binucleation in Drosophila main cells (You, 2023).
Seminal fluid plays an essential role in promoting male reproductive success and modulating female physiology and behavior. In the fruit fly, Drosophila melanogaster, Sex Peptide (SP) is the best-characterized protein mediator of these effects. It is secreted from the paired male accessory glands (AGs), which, like the mammalian prostate and seminal vesicles, generate most of the seminal fluid contents. After mating, SP binds to spermatozoa and is retained in the female sperm storage organs. It is gradually released by proteolytic cleavage and induces several long-term postmating responses, including increased ovulation, elevated feeding, and reduced receptivity to remating, primarily signaling through the SP receptor (SPR). This study demonstrates a previously unsuspected SPR-independent function for SP. In the AG lumen, SP and secreted proteins with membrane-binding anchors are carried on abundant, large neutral lipid-containing microcarriers. These microcarriers are transferred to females during mating where they rapidly disassemble. Remarkably, SP is a key microcarrier assembly and disassembly factor. Males expressing nonfunctional SP mutant proteins that affect SP's binding to and release from sperm in females also do not produce normal microcarriers. These data therefore reveal a role for SP in formation of seminal macromolecular assemblies, which may explain the presence of SP in Drosophila species that lack the signaling functions seen in D melanogaster (Wainwright, 2021).
Seminal fluid plays an essential role in male reproductive success. In D. melanogaster, SP, produced from the male AG, has been highlighted as a central player in this process, acting via receptors in the female to stimulate changes that increase fecundity and prevent remating. This study demonstrates that SP has an additional, unsuspected role in males in the assembly of neutral lipid-containing microcarriers in the AG lumen. These microcarriers store SP and can carry other proteins with lipid anchors. Furthermore, proteomics analysis reveals that the normal delivery of subgroups of SFPs to females during mating requires SP, potentially because these subgroups interact differently with microcarriers. Microcarrier interactions are likely to also affect dispersal of these proteins in the female reproductive tract. This analysis of microcarriers in other Drosophila species reveals that SP's microcarrier assembly function may exist in species in which SP has more limited roles in modulating the PMR, suggesting that the former function might have been critical in the evolution of this molecule (Wainwright, 2021).
Seminal proteins are produced throughout adult life, but these proteins are only transferred to females sporadically. Some of these proteins are then rapidly activated via mechanisms that are thought to include proteolytic cleavage and pH changes in the female reproductive tract (discussed in Wilson, 2017). The data suggest that microcarriers could contribute to this activation process. They are repositories for main cell-derived seminal proteins, which presumably partition from the aqueous phase of the AG's secretions, either because of their lipophilicity or because they have binding partners on the microcarrier surface. In the male, molecules like SP bind specifically to microcarriers and not to AG epithelial cells, strongly suggesting that these surfaces are structurally distinct. Subsequent microcarrier dissipation in the female reproductive tract provides a mechanism for dispersing proteins like SP so they can associate with receptors and cell membranes following mating (Wainwright, 2021).
Although both staining of normal microcarriers with lipophilic dyes and the homogeneous internal structure of large defective SP1/Df(SP) null 'microcarriers' observed with DIC strongly suggest that neutral lipids are a major component of these structures, their precise composition remains unclear. In addition, their nonspherical shape in wild-type males suggests that architectural proteins are highly likely to be involved in establishing their final structure, a proposal supported by the SP mutant phenotype. It will now be important to identify these other structural constituents and to establish whether any of these, unlike SP, play evolutionarily conserved roles in seminal fluid production outside the Drosophila genus (Wainwright, 2021).
Analysis of transcriptomics data from adult Drosophila organs reveals high level expression in the AG of multiple lipases that are predicted to be secreted (e.g., CG5162, CG11598, CG11600, CG11608, CG13034, CG18258, CG18284, CG31872, and CG34447), with all having been detected in proteomics analyses of seminal fluid. These include proteins sharing homology with triacylglycerol lipases (e.g., CG5162, CG13034, CG18258, and CG34447). These lipases provide a potential mechanism to break down neutral lipid transferred in microcarriers to females so the products can be used as fuel. Mammalian seminal fluid also contains lipases and triacylglycerides, suggesting that the latter may be required, perhaps as a male-derived nutrient source, in the reproductive system of all higher organisms (Wainwright, 2021).
Identification of extracellular neutral lipid microcarriers as accessible stores of specific seminal proteins is reminiscent of the role of intracellular lipid droplets in storage of cytoplasmic and nuclear proteins. Lipid droplets are able to dock with specific intracellular organelles to mediate their functions and deliver their cargos. It will be interesting to investigate whether the remnants of microcarriers, such as the microdomains observed with SP-GFP, are in any way targeted to specific cells or structures after transfer to females as these storage vehicles break down (Wainwright, 2021).
It has previously been reported in Drosophila that males can adaptively modulate the relative balance of seminal proteins, including SP, in the ejaculate, depending on female mating status and the presence of rival males. Loading of selected proteins onto microcarriers might provide a simple mechanism to control such rapid changes if the transfer of these large structures can be differentially regulated compared to soluble proteins: for example, by controlling the opening of the sphincters through which seminal fluid passes from the AGs to the ejaculatory duct (Wainwright, 2021).
This study reveals a previously unsuspected male-specific, SPR-independent role for SP in regulating microcarrier shape and size. SP mutants in D. melanogaster still have neutral lipid-containing structures, but they appear to aggregate and fuse, particularly after mating, to generate large lipid droplet-like structures that no longer retain molecules like SP at their surface. To date, it has not been possible to separate the different activities of SP in males and females through expression of different mutants or altered SP levels, making it difficult to fully gauge the importance of the male-specific microcarrier function. However, the observation that SP mutants, which were known to affect binding of SP to the surface of sperm or its subsequent release, also fail to rescue the microcarrier defect in SP null males suggests that the interpretation of the phenotypes associated with these mutants requires some reevaluation (Wainwright, 2021).
The data suggest that both the C-terminal and N-terminal domains of SP can interact with microcarriers even though they share no structural similarity. This may involve direct binding to the outer surface of the microcarrier or, because both domains contain charged residues, more indirect associations via other molecules attached to microcarriers. The multidomain interaction contrasts with sperm binding and may underlie why SP can transfer to sperm in the female reproductive tract (Wainwright, 2021).
It has been suggested that SP is likely to have roles in addition to its effects mediated via SPR signaling in the female reproductive tract, which include induction of a female sexual refractory period. This is because some SP-expressing species, like D. pseudoobscura and D. persimilis, do not appear to express SPR in this location and additionally show much less female postmating refractoriness relative to other SP-producing species. The current data suggest that microcarrier assembly may be this additional function, with the shape of microcarriers rapidly coevolving with SP. An absence of microcarriers in species with a highly divergent (D. virilis) or no (D. mojavensis) SP homolog, as evidenced by two different staining methods, supports the hypothesis. Not unexpectedly, DIC microscopy suggested that the luminal content of these latter two species is not homogenous, but it is clearly different from the other Drosophila species that ware studied. Interestingly, D. virilis expresses SPR in the female reproductive tract so, unlike in other species, its SP protein may be specifically involved in activating the female PMR, rather than microcarrier formation (Wainwright, 2021).
In light of these findings, it will now be important to investigate whether other proteins with fundamental roles in packaging and storing seminal fluid components have also evolved signaling roles in animals (Wainwright, 2021).
An important conclusion from this study is that the normal transfer of different subgroups of SFPs is dependent on SP. One simple explanation is that this reflects differences in their interactions with microcarriers. Having shown that main cell-expressed GFP-GPI binds to microcarriers, it was interesting to identify the GPI-anchored junctional protein Contactin as one of the proteins, which appears to be retained more in the AGs of SP null males. Furthermore, preferential retention of Dup99B in SP null males is consistent with the idea that this SP-like protein might bind to microcarriers even though it is primarily expressed in the adjacent ejaculatory duct epithelium (Wainwright, 2021).
Elegant studies by Wolfner and coworkers have identified several long-term response (LTR) network genes expressed in the AG that are interdependent and required either in the male or female for SP to be retained in the sperm storage organs. It was noticed that several of these proteins appear to be expressed at higher levels in SP nulls and also that a greater proportion is transferred from mutant males to females upon mating. A previous study has suggested that two of these proteins, CG1652 and CG1656, are present at similar levels in the female reproductive tract 1 h ASM to SP null and SP rescue males (40). This difference cannot be explained, but it is important to emphasize that this study measures the relative quantity of these proteins that leaves the male AG, not what remains in the female reproductive tract some time later. Overall, the proteomics analysis clearly shows that SP modulates the transfer of specific subclasses of SFPs to females, and, particularly in the case of proteins that are retained in SP nulls, this could result from the disruption of microcarriers (Wainwright, 2021).
It will now be important to investigate whether any of the network genes is involved in loading or unloading SP from microcarriers or, indeed, whether they play a role in microcarrier assembly, particularly since they appear to be present in species where SP does not seem to be involved in signaling. The role of secondary cells in microcarrier morphology also needs to be examined in more detail. Furthermore, confirming that other SFPs identified in the proteomics analysis or main cell-expressed GPI-anchored proteins are microcarrier cargos should allow the functions of these structures to be assessed more extensively and may suggest molecular tools that could be used to screen for similar processes in higher organisms (Wainwright, 2021).
Sphingolipids are ubiquitous structural components of eukaryotic cell membranes which are vital for maintaining the integrity of cells. Alkaline ceramidase is a key enzyme in sphingolipid biosynthesis pathway; however, little is known about the role of the enzyme in the male reproductive system of Drosophila melanogaster. To investigate the impact of alkaline ceramidase (Dacer) on male Drosophila, Dacer deficiency mutants (MUs) were obtained and were found to displayed apparent defects in the testis's phenotype. To profile the molecular changes associated with this abnormal phenotype, de novo transcriptome analyses of the MU and wildtype (WT) testes was performed; 1239 upregulated genes and 1102 downregulated genes were obtained. Then, six upregulated genes (papilin [Ppn], croquemort [Crq], terribly reduced optic lobes [Trol], Laminin, Wunen-2, collagen type IV alpha 1 [Cg25C]) and three downregulated genes (mucin related 18B [Mur18B], rhomboid-7 [Rho-7], CG3168) were confirmed through quantitative real-time polymerase chain reaction in WT and MU samples. The differentially expressed genes were mainly associated with catalytic activity, oxidoreductase activity and transmembrane transporter activity, which significantly contributed to extracellular matrix-receptor interaction, fatty acids biosynthesis as well as glycine, serine, and threonine metabolism. The results highlight the importance of Dacer in the reproductive system of D. melanogaster and provide valuable resources to dig out the specific biological functions of Dacer in insect reproduction (Zhang, 2021).
Many ribosomal proteins (RPs) not only play essential roles in ribosome biogenesis, but also have "extraribosomal" functions in various cellular processes. RpL36 encodes ribosomal protein L36, a component of the 60S subunit of ribosomes in Drosophila melanogaster. This paper reports RpL36 is required for spermatogenesis in D. melanogaster. After showing the evolutionary conservation of RpL36 sequences in animals, it was shown that RpL36 expression level in fly testes was significantly higher than in ovaries. Knockdown RpL36 in fly testes resulted in a significantly decreased egg hatch rate when these males mated with wild-type females. Furthermore, 76.67% of the RpL36 knockdown fly testes were much smaller in comparison to controls. Immunofluorescence staining showed that in the RpL36 knockdown testis hub cell cluster was enlarged, while the number of germ cells, including germ stem cells, was reduced. Knockdown of RpL36 in fly testis caused much fewer or no mature sperms in seminal vesicles. The terminal deoxynucleotidyl transferase dUTP nick end labeling (TUNEL) signal was stronger in RpL36 knockdown fly testes than in the control testes, but the TUNEL-positive cells could not be stained by Vasa antibody, indicating that apoptotic cells are not germ cells. The percentage of pH3-positive cells among the Vasa-positive cells was significantly reduced. The expression of genes involved in cell death, cell cycle progression, and JAK/STAT signaling pathway was significantly changed by RpL36 knockdown in fly testes. These results suggest that RpL36 plays an important role in spermatogenesis, likely through JAK/STAT pathway, thus resulting in defects in cell-cycle progression and cell death in D. melanogaster testes (Fang, 2021).
The human orthologue of the tumor suppressor protein FBW7 is encoded by the Drosophila archipelago (ago) gene. Ago is an F-box protein that gives substrate specificity to its SCF ubiquitin ligase complex. It has a central role in multiple biological processes in a tissue-specific manner such as cell proliferation, cellular differentiation, hypoxia-induced gene expression. This study presents a previously unknown tissue-specific role of Ago in spermatid differentiation. A classical mutant of ago was identified that is semi-lethal and male-sterile. During the characterization of ago function in testis, ago was found to play role in spermatid development, following meiosis. Spermatogenesis defects was confirmed by silencing ago by RNAi in testes. The v mutants show multiple abnormalities in elongating and elongated spermatids, including aberration of the cyst morphology, malformed mitochondrial structures, and individualization defects. Additionally, the subcellular localization of Ago protein was determined with mCherry-Ago transgene in spermatids. These findings highlight the potential roles of Ago in different cellular processes of spermatogenesis, like spermatid individualization, and regulation of mitochondrial morphology (Vedelek, 2021).
During spermatogenesis, the process in which sperm for fertilization are produced from germline cells, gene expression is spatiotemporally highly regulated. In Drosophila, successful expression of extremely large male fertility factor genes on Y-chromosome spanning some megabases due to their gigantic intron sizes is crucial for spermatogenesis. Expression of such extremely large genes must be challenging, but the molecular mechanism that allows it remains unknown. This study reports that a novel RNA-binding protein Maca, which contains two RNA-recognition motifs, is crucial for this process. maca null mutant male flies exhibited a failure in the spermatid individualization process during spermatogenesis, lacked mature sperm, and were completely sterile, while maca mutant female flies were fully fertile. Proteomics and transcriptome analyses revealed that both protein and mRNA abundance of the gigantic male fertility factor genes kl-2, kl-3, and kl-5 (kl genes) are significantly decreased, where the decreases of kl-2 are particularly dramatic, in maca mutant testes. Splicing of the kl-3 transcripts was also dysregulated in maca mutant testes. All these physiological and molecular phenotypes were rescued by a maca transgene in the maca mutant background. Furthermore, it was found that in the control genetic background, Maca is exclusively expressed in spermatocytes in testes and enriched at Y-loop A/C in the nucleus, where the kl-5 primary transcripts are localized. These data suggest that Maca increases transcription processivity, promotes successful splicing of gigantic introns, and/or protects transcripts from premature degradation, of the kl genes. This study identified a novel RNA-binding protein Maca that is crucial for successful expression of the gigantic male fertility factor genes, spermatogenesis, and male fertility (Zhu, 2021).
CPEB proteins are conserved translation regulators involved in multiple
biological processes. One of these proteins in Drosophila, Orb2, is a principal player in spermatogenesis. It is required for meiosis and spermatid differentiation. During the later process orb2 mRNAs and proteins are localized within the developing spermatid. To evaluate the role of orb2 mRNA 3'UTR in spermatogenesis, the CRISPR/Cas9 system was used to generate a deletion of the orb2
3'UTR, orb2R. This deletion disrupts the process of spermatid
differentiation but has no apparent effect on meiosis. Differentiation
abnormalities include defects in the initial polarization of the 64-cell
spermatid cysts, mislocalization of mRNAs and proteins in the
elongating spermatid tails, altered morphology of the elongating
spermatid tails, and defects in the assembly of the individualization
complex. These disruptions in differentiation appear to arise because orb2 mRNAs and proteins are not properly localized within the 64-cell spermatid cyst (Gilmutdinov, 2021).