oskar


DEVELOPMENTAL BIOLOGY

Oocyte

Oskar mRNA is present in the germarium before 16 cell clusters bud to form individual egg chambers. The Oskar mRNA is confined to the oocyte, the one cell among the 16 within the cluster that is not a nurse cell. By late stage 8, OSK mRNA is still present throughout the oocyte, but has become more concentrated at the anterior margin and the posterior pole. By stage 9, (middle egg chamber), OSK mRNA is no longer concentrated at the anterior margin, and soon becomes completely localized to the posterior pole. Localization takes place prior to the onset of cytoplasmic streaming and prior to bulk flow from the nurse cells into the oocyte. During stage 10A, OSK mRNA begins to accumulate at high levels in the nurse cells.

Different combinations of gap repressors for common stripes in Anopheles and Drosophila embryos

Drosophila segmentation is governed by a well-defined gene regulation network. The evolution of this network was investigated by examining the expression profiles of a complete set of segmentation genes in the early embryos of the mosquito, Anopheles gambiae. There are numerous differences in the expression profiles as compared with Drosophila. The germline determinant Oskar is expressed in both the anterior and posterior poles of Anopheles embryos but is strictly localized within the posterior plasm of Drosophila. The gap genes hunchback and giant display inverted patterns of expression in posterior regions of Anopheles embryos, while tailless exhibits an expanded pattern as compared with Drosophila. These observations suggest that the segmentation network has undergone considerable evolutionary change in the dipterans and that similar patterns of pair-rule gene expression can be obtained with different combinations of gap repressors. The evolution of separate stripe enhancers in the eve loci of different dipterans is discussed (Goltsev, 2004).

Anopheles lacks bicoid and contains a lone Hox3 gene that is more closely related to zen and specifically expressed in the serosa. How is hunchback activated in the presumptive head and thorax in Anopheles? The homeobox gene orthodenticle can substitute for bicoid in Tribolium. However, orthodenticle does not appear to be maternally expressed in Anopheles, but instead, staining is strictly zygotic and restricted to anterior regions, similar to the pattern seen in Drosophila. Sequential patterns of orthodenticle, giant, and hunchback expression are established by differential threshold readouts of the Bicoid gradient in Drosophila. It is possible that an unknown maternal regulatory gradient emanating from the anterior pole is responsible for producing similar patterns of expression in Anopheles. It is proposed that this unknown regulatory factor may be localized to the anterior pole by Oskar. Oskar coordinates the assembly of polar granules and is essential for the localization of Nanos in the posterior plasm. It might also localize one or more unknown determinants in anterior regions of Anopheles embryos (Goltsev, 2004).

Stimulation of endocytosis and actin dynamics by Oskar polarizes the Drosophila oocyte

In Drosophila, localized activity of oskar at the posterior pole of the oocyte induces germline and abdomen formation in the embryo. Oskar has two isoforms, a short isoform encoding the patterning determinant and a long isoform of unknown function. This study shows by immuno-electron microscopy that the two Oskar isoforms have different subcellular localizations in the oocyte: Short Oskar mainly localizes to polar granules, and Long Oskar is specifically associated with endocytic membranes along the posterior cortex. Cell biological and genetic analyses reveal that Oskar stimulates endocytosis, and its two isoforms are required to regulate this process. Furthermore, long F-actin projections at the oocyte posterior pole are described that are induced by and intermingled with Oskar protein. It is proposed that Oskar maintains its localization at the posterior pole through dual functions in regulating endocytosis and F-actin dynamics (Vanzo, 2007).

This investigation of the subcellular localization of Oskar in the Drosophila oocyte has uncovered an unanticipated function of Oskar in endocytosis. Using immuno-electron microscopy, it was found that Short and Long Oskar are mostly concentrated on distinct cellular structures in the oocyte—the polar granules and the endocytic compartment, respectively. Using stereological methods and functional assays, it has been shown that endocytosis levels are asymmetric around the cortex of wild-type oocytes, and that high levels of endocytosis at the posterior pole require Oskar expression. In addition, oskar mutant oocytes exhibit both a reduced endocytosis of yolk proteins and a reduction in endocytic structures at the posterior plasma membrane. Finally, a function of Oskar was identified in the asymmetric organization of the F-actin cytoskeleton in the oocyte. These data strongly suggest that both isoforms are potent regulators of endocytosis and F-actin dynamics (Vanzo, 2007).

The localization of the two Oskar isoforms to distinct subcellular structures during oogenesis may account for the previous unexplained observation of their distinct segregation during early embryogenesis, when the primordial germ cells form. Indeed, during these stages, Short Oskar accumulates in the pole cells, the future germ cells. During oogenesis, Short Oskar mainly concentrates in the polar granules. This localization is consistent with the previously reported function of this isoform in the establishment of the germ cell lineage in the embryo, a process thought to be instructed by polar granules. In contrast, Long Oskar is selectively excluded from pole cells in the early embryo and shows a specific affinity for membranous structures during oogenesis. Thus, the differential localization of the two Oskar isoforms that originates in oogenesis could persist over the lifetime of the two proteins and specify distinct destinies—Short Oskar being incorporated in the germ cells, and Long Oskar not (Kempkens, 2006).

Distinct localizations of the two Oskar isoforms could be due to the amino-terminal extension of Long Oskar (M1M2). However, this extension is unlikely sufficient to explain the specific association of Long Oskar with endocytic membranes, since Short Oskar can also associate with these, at a very low level. This extension might either increase the membrane affinity of Long Oskar and/or provide additional elements for efficient targeting/recruitment. It is noteworthy that Long Oskar is not an integral membrane protein. Neither a signal peptide, required for ER targeting, nor a significant hydrophobic stretch, required for membrane insertion, are apparent in its primary sequence. It is therefore presumably a peripheral protein recruited from the ooplasm to the cytosolic face of the endocytic membrane (Vanzo, 2007).

The absence of Long Oskar from polar granules is surprising for two reasons. First, Long Oskar is required for the maintenance of Short Oskar at the posterior pole of the oocyte. This implies that the maintenance function of Long Oskar must operate in an indirect way. Second, in addition to its specific amino-terminal extension, Long Oskar contains the entire Short Oskar peptide. This suggests that the N-terminal extension inhibits polar granule association of Long Oskar, in addition to promoting its association with membranes. As the polar granule component Vasa is not found on Long Oskar-positive membranes, this extension might also prevent Long Oskar from recruiting/assembling polar granule components, explaining its failure to specify the germline and the abdomen (Vanzo, 2007).

In oskar null oocytes, clathrin-mediated endocytosis is affected, both in its efficiency and its asymmetry. The aberrant prominence of flat, clathrin-coated areas of plasma membrane, instead of coated pits and vesicles, in those oocytes suggests impairment of endocytosis at an early stage, possibly in the initial invagination of the plasma membrane. This reveals that Oskar is a novel regulator of endocytosis (Vanzo, 2007).

The finding that sole expression of Long Oskar in oskar null oocytes triggers the formation of long and dense membrane sheet invaginations whose formation is abrogated upon coexpression of Short Oskar from a second transgene provides plausible explanations for the function of Oskar isoforms. Long Oskar could trigger an early step in the formation of clathrin-mediated membrane invaginations, which, in the absence of Short Oskar, fail to pinch off as coated vesicles and, instead, become protrusions, as observed in the dynamin mutant shibire (shits) at the restrictive temperature. Very reminiscent of abortive coated extensions also described in shibire oocytes, numerous clathrin-coated buds forming at the tip and sides of the large plasma membrane invaginations were identified in oocytes expressing only Long Oskar. In these oocytes, the invaginations may eventually detach and mature into yolk granules. The fact that these invaginations are no longer observed when Short Oskar is present suggests that Short Oskar could have a specific function in the subcellular localization and/or activation of dynamin at the posterior pole. The second, not exclusive, possibility is that Long Oskar activates membrane invagination through a clathrin-independent mechanism of endocytosis that Short Oskar antagonizes/overcomes to promote classical clathrin-mediated endocytosis. Clathrin-independent endocytosis has long been proposed and has recently been shown to involve tubular pleiomorphic structures very similar to those observed in the ooplasm of oocytes expressing Long Oskar only (Vanzo, 2007).

In any event, it is speculated that maintenance of a physiological ratio of the two isoforms (estimated to 1/4, Long/Short Oskar) is crucial to stimulate the clathrin-mediated endocytosis. Consistent with this, overexpressing Long Oskar in wild-type oocytes also induces an alternative, non-clathrin-dependent endocytic pathway, confirming that the balance of expression of the two Oskar isoforms is critical (Vanzo, 2007).

In contrast to Long Oskar, whose sublocalization agrees with a function in endocytosis, it is not obvious how Short Oskar could directly regulate endocytosis and/or Long Oskar function. Only residual membrane association of Short Oskar is detected in transgenic oocytes expressing this isoform alone. However, it is possible that Long Oskar enhances this association in wild-type oocytes. In this hypothesis, Short Oskar might interact functionally with the endocytic membrane to directly regulate either endocytosis or Long Oskar activity. Alternatively, Short Oskar might act indirectly by promoting local concentration of critical regulators of clathrin-mediated endocytosis at the posterior pole, possibly dependent on polar granule assembly. Further investigation of the specific function of Short Oskar in endocytosis will require the development of molecular tools circumventing the requirement for Long Oskar in Short Oskar expression and localization (Vanzo, 2007).

Long Oskar is required for efficient maintenance of Short Oskar at the posterior pole of the oocyte. The finding that Short Oskar largely accumulates in polar granules during oogenesis therefore implies that Long Oskar is essential for polar granule maintenance or integrity. Polar granules are only transiently found in close proximity to endocytic membranes on which Long Oskar localizes, when they form at stage 9. Then, as they mature from stage 9 to stage 10, they become larger, denser, and move away from the area where endocytosis occurs, to accumulate more internally at the edge of the endocytic zone. This observation suggests that polar granules are not maintained at the posterior pole by direct anchoring to the endocytic membranes (Vanzo, 2007).

The cortical F-actin cytoskeleton has been implicated in posterior anchoring of Oskar in oocytes, but the mechanism by which it acts was not addressed. In light of the present results, an attractive hypothesis is that polar granule anchoring involves the F-actin projections that are observed at the posterior pole of wild-type oocytes. These projections would be of sufficient length to span the first micrometer internally from the plasma membrane and to contact the underlying polar granules. In addition, they become detectable from stage 10 onward, when anchoring is required. These projections are largely reduced in oskar null oocytes. As Long Oskar expression per se does not restore their formation, this model suggests the existence of a positive feedback loop mechanism of maintenance, in which polar granules, possibly in concert with Long Oskar, could enhance their own anchoring at the posterior pole (Vanzo, 2007).

In light of the subcellular localization of Long Oskar, its unique competence to anchor at the posterior of the oocyte is quite notable. Interestingly, in addition to its well-documented role in nutrient uptake, endocytosis has emerged as a mechanism restricting the localization of proteins to plasma membrane subdomains in different polarized cells. In yeast, polarized exocytosis coupled with local endocytic recycling localizes membrane proteins at growing shmoo tips. More recently, endocytic trafficking has been shown to localize epithelial polarity proteins and restrict membrane receptor-dependent signaling during cell migration. By analogy, Long Oskar might be maintained and concentrated by continuous endocytic cycles at the posterior pole. By upregulating endocytosis, Oskar might promote its own accumulation, in addition to that of other membrane-associated factors, at the posterior pole. Consistent with this, Oskar has been shown to enhance the posterior accumulation of Rab11, a small GTPase recruited to endocytic membranes and required for endocytic recycling in the oocyte (Vanzo, 2007).

In addition to being actin dependent, Oskar anchoring also relies on an uncharacterized actin-independent mechanism. This mechanism was inferred from the observation that drug-induced F-actin depolymeriation caused only mild Oskar-anchoring defects in wild-type, in vitro-cultured oocytes. In light of the cureent findings, it is proposed that endocytosis might be this alternative mechanism. Consistent with this, the posterior localization of a Long Oskar reporter construct is essentially unaffected by actin depolymerization. Significant Oskar-anchoring defects are provoked by F-actin depolymerization in homer mutant oocytes. Although the molecular function of Homer in anchoring is unknown, it has been proposed that the protein is a key player in the actin-independent anchoring mechanism. Strikingly, using immuno-electron microscopy, it has been observed that a fraction of Homer associates with endocytic membranes and partially colocalizes with Long Oskar-containing endocytic structures. This observation supports a role of Homer and endocytosis in the mechanism of Long Oskar maintenance (Vanzo, 2007).

In conclusion, this work has revealed unexpected cellular functions of Oskar. Beyond its known functions in posterior patterning and germline induction, Oskar regulates asymmetry in clathrin-mediated endocytosis and F-actin organization in the Drosophila oocyte. It is proposed that, by regulating these two cellular processes, in positive feedback loops, Oskar isoforms promote their own maintenance at the posterior pole, thus reinforcing oocyte polarity (Vanzo, 2007).

The endocytic pathway acts downstream of Oskar in Drosophila germ plasm assembly

Cell fate is often determined by the intracellular localization of RNAs and proteins. In Drosophila oocytes, oskar (osk) RNA localization and the subsequent Osk synthesis at the posterior pole direct the assembly of the pole plasm, where factors for the germline and abdomen formation accumulate. osk RNA produces two isoforms, long and short Osk, which have distinct functions in pole plasm assembly. Short Osk recruits downstream components of the pole plasm, whose anchoring to the posterior cortex requires long Osk. The anchoring of pole plasm components also requires actin cytoskeleton, and Osk promotes long F-actin projections in the oocyte posterior cytoplasm. However, the mechanism by which Osk mediates F-actin reorganization remains elusive. Furthermore, although long Osk is known to associate with endosomes under immuno-electron microscopy, it was not known whether this association is functionally significant. This study shows that Rabenosyn-5 (Rbsn-5), a Rab5 effector protein required for the early endocytic pathway, is crucial for pole plasm assembly. rbsn-5- oocytes fail to maintain microtubule polarity, which secondarily disrupts osk RNA localization. Nevertheless, anteriorly misexpressed Osk, particularly long Osk, recruits endosomal proteins, including Rbsn-5, and stimulates endocytosis. In oocytes lacking rbsn-5, the ectopic Osk induces aberrant F-actin aggregates, which diffuse into the cytoplasm along with pole plasm components. It is proposed that Osk stimulates endosomal cycling, which in turn promotes F-actin reorganization to anchor the pole plasm components to the oocyte cortex (Tanaka, 2008).

The polarized targeting and anchoring of specific molecules and organelles to particular subcellular regions are crucial for many cellular processes, including cell-polarity establishment and cell-fate determination. In many animals, germline fate is controlled by maternal factors localized to a specialized cytoplasmic region within the egg, called the germ plasm. Germ plasm contains germ granules, which are electron-dense, and non-membranous structures consisting of maternal RNAs and proteins required for the formation of germ cells. Drosophila germ plasm, also called pole plasm, forms at the posterior pole of the embryo and is inherited by the germline precursors, or pole cells. Because the cytoplasmic transplantation of the pole plasm into recipient embryos causes the ectopic formation of pole cells, the pole plasm contains sufficient factors for germ-cell formation. This observation also highlights the importance of retaining the pole plasm at the posterior cortex of the embryo to ensure the germ cells form at the appropriate location (Tanaka, 2008).

In Drosophila, the pole plasm is assembled during oogenesis, which is divided into 14 morphologically distinct stages of egg chamber development. The egg chamber is composed of a single oocyte and 15 nurse cells, surrounded by a monolayer of somatic follicle cells. During oogenesis, most components of pole plasm are synthesized in the nurse cells and transported into the oocyte via ring canals, which are cytoplasmic bridges interconnecting the oocyte with nurse cells. Within the oocyte, these factors become concentrated at the posterior pole and are assembled into the polar (germ) granules. These factors are transported by a polarized microtubule (MT) array that is initially nucleated at the oocyte posterior and extends into the nurse cells through the ring canals. During stages 6-7, the MT array is reorganized by the transforming growth factor alpha-like Gurken (Grk) signal. In the stage-6 oocyte, posteriorly restricted Grk induces neighboring follicle cells to adopt the posterior fate. These cells send back as-yet unknown signals to the oocyte to trigger the reorganization of the MT cytoskeleton. Consequently, the MT array within the oocyte becomes polarized along the anteroposterior (AP) axis, with the minus ends abundant at the anterior of the oocyte and the plus ends extending toward the posterior. This MT organization promotes the migration of the oocyte nucleus and associated grk RNA to the future anterior-dorsal corner, where Grk signals the follicle cells to define the dorsoventral axis. The polarized MT array also directs the localization of bicoid (bcd) RNA to the anterior and oskar (osk) RNA to the posterior within the oocyte. The anterior accumulation of bcd RNA is required for the proper development of the embryonic head and thoracic structures. The posterior localization of osk RNA is essential for the formation of the germ cells and abdomen (Tanaka, 2008).

osk RNA localization is tightly coupled to translational control: only the posteriorly localized osk message is translated. The localized Osk protein, in turn, recruits downstream components of the pole plasm, such as Vasa (Vas) and Tudor (Tud) proteins, and the nanos, germ cell-less and polar granule component RNAs. Misexpression of Osk at the anterior of the oocyte causes ectopic pole plasm assembly and the formation of germ cells at the new site, indicating that Osk organizes pole plasm assembly (Tanaka, 2008).

Although osk has no known alternatively spliced variants, the osk message produces two protein isoforms, long and short Osk, by translation from in-frame alternative start codons. Short Osk shares its entire sequence with the long isoform. Nevertheless, genetic evidence shows that the two Osk isoforms have distinct functions in the assembly of the pole plasm. Long Osk is required for all the components of the pole plasm, including Osk itself, to be anchored to the posterior cortex, preventing their diffusion into the cytoplasm. However, the mechanism by which long Osk retains pole plasm components at the posterior cortex remains unknown (Tanaka, 2008).

A recent immuno-electron microscopic study revealed that the two Osk isoforms localize to distinct organelles in the oocyte posterior: long Osk associates with endosomes and short Osk is concentrated in the polar granules (Vanzo, 2007). Long Osk also upregulates endocytosis, which occurs preferentially at the oocyte posterior (Vanzo, 2007). Therefore, the endocytic pathway may be involved in pole plasm assembly downstream of long Osk, although data are lacking to show that the association between long Osk and endosomes is functionally significant. Several reports have suggested that vesicular trafficking is involved in pole plasm assembly and germ cell formation. For example, in mutants for Rab11, which encodes a small GTPase involved in the recycling of endosomes, osk RNA fails to be transported to the oocyte posterior, instead forming aggregates close to the posterior. However, the defects in osk RNA localization in Rab11 mutants are thought to be an indirect consequence of the disrupted MT polarization (Tanaka, 2008).

This study shows that Drosophila Rabenosyn-5 (Rbsn-5), a Rab5 effector protein involved in the early endocytic pathway, is required for osk RNA localization and pole plasm assembly. Although the primary defect of the rbsn-5 mutation is, as in the Rab11 mutant, caused by the failure to maintain MT polarity, which secondarily affects osk RNA localization, evidence is provided that the endocytic pathway also functions downstream of Osk to anchor the pole plasm components to the oocyte cortex (Tanaka, 2008).

Vas is a reliable marker for the germline throughout Drosophila development. A GFP-Vas fusion protein enables the direct visualization of the pole plasm and germ cells in the living organism. During oogenesis, GFP-Vas accumulates at the oocyte posterior from stage 9 onward. Using GFP-Vas as a marker, a germline clonal screen was performed targeting chromosome 2L for mutations that disrupted pole plasm assembly. From 5122 lines mutagenized with EMS, 66 mutants were isolated defective in GFP-Vas localization. Twenty-seven of these were alleles of cappuccino, spire or profilin (chickadee), three genes on 2L that are known to be involved in osk RNA localization, which validates the screening strategy (Tanaka, 2008).

Among the other mutants recovered was a recessive lethal mutation, C241, that mapped to 28C2-29E2. Subsequent deficiency mapping and sequencing of the mutant chromosome revealed that the C241 mutation was a single nucleotide substitution in the CG8506 gene, which resulted in a premature stop codon at position 315 of the 505 amino acid open reading frame (ORF). The introduction of a transgene containing a genomic DNA fragment with the CG8506 transcriptional unit rescued the C241 mutant phenotypes (described below). These data show that CG8506 corresponds to the gene that was mutated at the C241 locus. Rabbit and rat polyclonal antisera raised against full-length CG8506 did not detect a truncated form of CG8506 in ovarian extracts from C241 heterozygotes. Furthermore, neither antibody showed immunoreactivity in C241 homozygous clones, suggesting that the truncated protein was not expressed at detectable levels and/or was unstable. Therefore, C241 appeared to be a strong loss-of-function, presumably a protein-null, allele of CG8506 (Tanaka, 2008).

CG8506 (Rabenosyn - FlyBase) encodes a protein homologous to Rabenosyn-5 (Rbsn-5). Rbsn-5 interacts with several Rab proteins, including Rab5, which functions in early endosomal transport. Several Rbsn-5 protein domains are conserved across species, including the FYVE domain, which binds phosphatidylinositol-3-phosphate. However, invertebrate Rbsn-5 homologs lack the C-terminal domain common to the mammalian homologs of this protein. Since the C-terminal domain of mammalian Rbsn-5 is responsible for its interaction with Rab5, whether CG8506 interacted with Rab5 was examined. Pull-down assays showed that GST-Rab5 efficiently pulled down in-vitro-synthesized CG8506 protein in the presence of a GTP analog, GTP-γS, but inefficiently in the presence of GDP. The interaction between CG8506 and Rab5-GTP was specific, because the interactions of CG8506 with Rab11 and Rab7 were at background levels. Consistent with a physical interaction between CG8506 and Rab5 in vitro, in CG8506C241 GLCs, neither auto-fluorescent granules derived from endocytosed yolk proteins nor the incorporation of a fluorescent marker for endocytosis, FM4-64, were observed in the oocytes, suggesting that CG8506 functions cooperatively with Rab5 in the early endocytic pathway. Thus, CG8506 is the Drosophila ortholog of Rbsn-5 and has an evolutionarily conserved function in the endocytic pathway (Tanaka, 2008).

This study shows that that Osk maintains, but does not establish, the posterior accumulation of endosomal proteins and asymmetric endocytosis, and that Osk can recruit endosomal proteins and stimulate endocytosis even at an ectopic site. It is further shown that the anchoring of the pole plasm components to the oocyte cortex requires the Osk-dependent stimulation of endocytic activity. These data reveal an interdependent relationship between Osk anchoring and localized endocytic activity at the oocyte posterior (Tanaka, 2008).

In rbsn-5- oocytes, the anterior misexpression of Osk induces aberrant F-actin aggregates, which diffuse along with pole plasm components into the cytoplasm. Several lines of evidence suggest that the anchoring of pole plasm components requires the proper organization of F-actin. Since endosomal proteins are recruited by long Osk, the idea is favored that the endocytic pathway functions downstream of long Osk to anchor the pole plasm components at the cortex by regulating F-actin dynamics. Supporting this idea, in addition to its roles in early endosomal sorting, Rab5 acts as a signaling molecule that remodels F-actin networks. Rab11, which regulates the recycling of endosomes, is also involved in F-actin organization during cellularization in Drosophila blastoderm embryos. Intriguingly, the recruitment of endosomal proteins by Osk is not sufficient for proper F-actin reorganization to anchor the pole plasm components at the cortex, because their recruitment occurs even in oocytes lacking Rbsn-5, in which cortical anchoring fails. It is therefore proposed that the continuous cycling of endosomes is required for pole plasm components to be anchored to the oocyte cortex. This scenario is compatible with a model in yeasts, which use endocytic cycling coupled with localized exocytosis to maintain their polarity, although it is unclear if F-actin reorganization is involved in this process (Tanaka, 2008).

Rbsn-5 is primarily required for the maintenance of MT polarity that directs posterior localization of osk RNA. Rab11 is also required for MT polarization in the oocyte. However, the accumulation of endosomal proteins and upregulation of endocytosis at the oocyte posterior require the oocyte polarization, which promotes the reorganization of the MT array. Thus, MT polarization and asymmetric activation of the endocytic pathway are probably interdependent as well. Furthermore, maintenance of polarized endocytic activity depends on Osk. Intriguingly, Osk is also thought to maintain MT polarity, as posterior accumulation of Kin-βgal is partially defective in the absence of Osk (Zimyanin, 2007). It is therefore likely that the endocytic pathway and Osk form a positive-feedback loop that maintains oocyte polarity: Osk may maintain MT polarity through recruiting endosomal proteins. Based on these results, a model is proposed in which the endocytic pathway is involved in several distinct steps in pole plasm assembly (Tanaka, 2008).

The localization of bcd RNA to the anterior pole of the oocyte requires the ESCRT-II (endosomal sorting complex required for transport II) complex, which sorts mono-ubiquitinated endosomal transmembrane proteins into multivesicular bodies. Furthermore, Vps36p, a component of the ESCRT-II complex, binds bcd 3' UTR in vitro and co-localizes with bcd RNA at the oocyte anterior, suggesting the direct involvement of ESCRT-II in bcd RNA localization. osk RNA, however, appears to use another mechanism for its posterior localization, since its localization is unaffected in the absence of ESCRT-II function. Several lines of evidence suggest that ER organization and RNA localization are linked. However, it is considered unlikely that the ER directs the posterior localization of osk RNA, because ER components and osk RNP distributed differentially in developing oocytes. Interestingly, the osk RNP and the endosomal proteins are in close proximity during their transport to the oocyte posterior. Although their close association may simply be owing to the dynamic rearrangements of the MT array during stages 7-8, these findings suggest that the endocytic pathway may also play a role in the targeting of osk RNP to the posterior pole of the oocyte. Retroviral genomic RNAs are known to hitchhike on endosomal vesicles to reach the plasma membrane. Therefore, it will be interesting to learn if osk RNA is also transported to the posterior pole of the oocyte along with the endosomes (Tanaka, 2008).

Effects of Mutation or Deletion

Embryos derived from oskar females lack the specialized pole plasm including polar granules. They also lack pole cells, the zygotic stem cells for the gonads. In addition, the abdominal region remains unsegmented and eventually dies. Transplantation of cytoplasm from normal embryos into mutant embryos reveals that osk-dependent activity is strictly localized at the posterior pole and has two distinct functions. In mutant embryos the activity will normalize pole cell formation when transplanted into the posterior pole. Furthermore, osk activity can provoke the formation of a second "posterior center" at the anterior (Lehmann, 1986).

osk mutants fall into two classes: one in which OSK mRNA localization is normal, and one in which initial localization to the posterior pole is normal, but then becomes delocalized.

Oskar is involved in Drosophila long-term memory

Memory formation after olfactory learning in Drosophila displays behavioral and molecular properties similar to those of other species. Particularly, long-term memory requires CREB-dependent transcription, suggesting the regulation of 'downstream' genes. At the cellular level, long-lasting synaptic plasticity in many species also appears to depend on CREB-mediated gene transcription and subsequent structural and functional modification of relevant synapses. To date, little is known about the molecular-genetic mechanisms that contribute to this process during memory formation. Two complementary strategies were used to identify these genes. From DNA microarrays, 42 candidate memory genes were identified that appear to be transcriptionally regulated in normal flies during memory formation. Via mutagenesis, 60 mutants with defective long-term memory have been independently identified and molecular lesions have been identified for 58 of these. The pumilio translational repressor was found from both approaches, along with six additional genes with established roles in local control of mRNA translation. In vivo disruptions of four genes, staufen, pumilio, oskar, and eIF-5C, yield defective memory. It is concluded that convergent findings from the behavioral screen for memory mutants and DNA microarray analysis of transcriptional responses during memory formation in normal animals suggest the involvement of the pumilio/staufen pathway in memory. Behavioral experiments confirm a role for this pathway and suggest a molecular mechanism for synapse-specific modification (Daubnau, 2003).

The 60 memory mutants were generated with enhancer-trap transposons, which often drive reporter genes (lacZ or GFP) in patterns of expression similar to those of the endogenous genes they disrupt. Thus, reporter gene expression patterns were examined for milord-1 and -2 (pum), norka (oskar), and krasavietz (eIF-5C) in the adult CNS. Each of these enhancer-trap memory mutants shows common reporter gene expression in the mushroom body, an anatomical focus with a demonstrated role for olfactory memory. The norka and krasavietz strains carry a PGAL4 transposon that can drive expression of GFP in neuronal somata and processes. These data clearly reveal a common site of expression in a subset of intrinsic mushroom body neurons (Kenyon cells) that comprise the α and β lobes. The milord-1 and milord-2 strains, in contrast, carry a PlacZ transposon that expresses β galactosidase only in somata and, thus, only around the calyx region of mushroom bodies (Daubnau, 2003).

An existing mouse polyclonal antibody was to determine the expression pattern of Pum protein in the adult CNS. Consistent with the pattern of enhancer-trap expression for milord-1 and -2, Pum immunoreactivity is detected broadly in the CNS but appears to be expressed at high levels in mushroom body neurons. Strong immunoreactivity appears to be perinuclear in Kenyon cells, whereas weaker, punctate expression is detected in mushroom body neuropil (calyx). This antibody shows appreciable specificity on Western blots of embryonic extracts. It was not possible to use pumilio null mutants to establish antibody specificity for adult brain tissue, however, because the null mutants are not viable as adults. Coexpression in mushroom bodies of the reporter genes for oskar, pum, and eIF-5C and anti-Pum immunostaining are consistent with the notion that these genes function together in the CNS during long-term memory formation (Daubnau, 2003).

STAU already has been implicated in mRNA localization in mammalian CNS. In cultured hippocampal neurons, for instance, STAU has a punctate, somato-dendritic distribution and is a component of large RNP-containing neural granules, which themselves are associated with microtubules. These neural granules seem to play an analogous role in targeting mRNA translation to subcellular (synaptic) compartments in neurons, as do STAU-containing RNP particles (polar granules) in Drosophila oocytes. In cultured hippocampal neurons, neural granules are located near dendritic spines, appear to dissociate in response to local synaptic activity, and thereby release translationally repressed mRNAs. This process has been proposed as a mechanism for synapse-specific modification via local protein synthesis in response to neural activity in vitro. The data indicate that this staufen-dependent pathway underlies memory formation per se. Moreover, the further identification of oskar as a memory mutant and of moesin and orb as confirmed candidate memory genes suggests that additional genetic components of the machinery used for mRNA translocation and translation in oocytes also may function in neurons (Daubnau, 2003).

Combined with these observations from the literature, the data suggest a molecular mechanism for synapse-specific delivery of gene products during long-term memory formation. (1) Behavioral training results in the activation of CREB-mediated transcription, and nascent mRNAs are packaged into an RNP complex, a neural granule. In addition to staufen, oskar, and moesin, these granules may well include other components of polar granules such as mago and faf. (2) These neural granules then are transported into dendritic shafts along an organized microtubule network, as proposed above for vertebrate neurons. These activity-induced transcripts may be delivered nonselectively throughout the neuron or selectively to sites of recent synaptic activity. In either case, packaged mRNAs probably are translationally quiescent while in transport, thereby preventing ubiquitous expression of protein products. The data implicate pumilio as part of this translational repression complex. Finally, synapse-specific modification may result from the depolarization-dependent release of neural granule-associated mRNAs and a concomittant translational derepression (Daubnau, 2003 and references therein).

Local derepression of translation, in part, may involve phosphorylation of CPEB (orb) by aurora kinase, resulting in cytoplasmic polyadenylation and the dissociation of MASKIN from eIF-4E, which then allows interaction between eIF-4E and eIF-4G. Release of eIF-4E via phosphorylation of other 4E binding proteins also may promote assembly of the rest of the translation initiation complex. The presence during synaptic or behavioral plasticity of several other persistently active kinases also might contribute to such phosphorylation. CPEB-mediated translational activation in Xenopus oocytes, for instance, is associated with phosphorylation of ORB by CDC2 kinase (which is a dimer of CycB and CDC2) and ubiquitin-mediated degradation of Orb, perhaps modulated by faf or another ubiquitin hydrolase. Here again, DNA chip and memory mutant experiments have identified several of these additional components (Daubnau, 2003 and references therein).

Mis-specified cells die by an active gene-directed process, and inhibition of this death results in cell fate transformation in Drosophila

Incorrectly specified or mis-specified cells often undergo cell death or are transformed to adopt a different cell fate during development. The underlying cause for this distinction is largely unknown. In many developmental mutants in Drosophila, large numbers of mis-specified cells die synchronously, providing a convenient model for analysis of this phenomenon. The maternal mutant bicoid is a particularly useful model with which to address this issue because its mutant phenotype is a combination of both transformation of tissue (acron to telson) and cell death in the presumptive head and thorax regions. A subset of these mis-specified cells die through an active gene-directed process involving transcriptional upregulation of the cell death inducer hid. Upregulation of hid also occurs in oskar mutants and other segmentation mutants. In hid bicoid double mutants, mis-specified cells in the presumptive head and thorax survive and continue to develop, but they are transformed to adopt a different cell fate. Evidence is provided that the terminal torso signaling pathway protects the mis-specified telson tissue in bicoid mutants from hid-induced cell death, whereas mis-specified cells in the head and thorax die, presumably because equivalent survival signals are lacking. These data support a model whereby mis-specification can be tolerated if a survival pathway is provided, resulting in cellular transformation (Werz, 2005).

Although this study largely focus on the maternal effect mutants bicoid and oskar, it is likely that the principles uncovered are of broader significance. Segmentation mutants acting downstream of bicoid and oskar, including mutants of gap genes (Krüppel, knirps), pair-rule genes (odd, fushi-tarazu) and segment polarity genes (wg, hedgehog, engrailed) induce expression of hid. These mutants are characterized by loss of larval tissue. As in the case of bicoid and oskar, hid expression is upregulated during stage 9 of embryogenesis in the regions of the mutant embryos that are later deleted in the larvae. In addition, hid mutants rescue the cuticle phenotype of armadillo mutants. Finally, hid expression accompanied by TUNEL-positive cell death was found in dorsal and Toll10b mutants, which cause dorsalizing and ventralizing phenotypes, respectively, along the dorsoventral axis of Drosophila embryos. Thus, these data support the notion that upregulation of hid appears to be a common trigger for a caspase-dependent cell death program in mis-specified cells of patterning mutants (Werz, 2005).

Furthermore, mutations affecting imaginal disc development result in loss of the adult appendage due to inappropriate cell death. It is currently being determined whether these mutants also require hid expression to develop the final phenotypes. Moreover, many gene disruptions in mice result in inappropriate cell death in the tissue that requires the function of the disrupted gene, suggesting that similar mechanisms might exist in mammalian development. Finally, cell death may be an important contributing factor to human congenital birth defects. Thus, an understanding of the underlying mechanisms is of general interest (Werz, 2005).

Interestingly, not all segment polarity mutants analyzed induce hid expression and cell death. Embryos mutant for patched, which encodes the hedgehog receptor, were not found to express hid and do not exhibit increased cell death, although hedgehog mutants both upregulate hid and contain increased amounts of cell death. The reasons for these differences are not known, but partial redundancy might account for lack of hid expression in patched mutants. The Drosophila genome encodes another patched homolog, patched-related, which might provide the survival requirement for mis-specified cells in patched mutants (Werz, 2005).

Mis-specified cells in bicoid and oskar mutants induce expression of hid. No increased reaper or grim expression was observed in these mutants. However, expression of reaper has been reported in crumbs mutants, which affect epithelial integrity. X-ray-treated embryos also preferentially respond by upregulation of reaper, rather than hid. Although crumbs mutants were not analyzed for hid expression, it appears that cells contain several developmental checkpoints, which activate different cell death-inducing regulators depending on the type of abnormal cellular development (Werz, 2005 and references therein).

Mis-specified cells can survive if an alternative survival pathway is provided. The example presented here is the acron into telson transformation in bicoid mutants, which is mediated by the torso signaling pathway. Although the cells giving rise to telson structures at the anterior tip are mis-specified based on Abd-B-labeling experiments, they survive because they receive a survival signal from the torso signaling system. In this case, transformation rather than cell death is favored. It has been shown that activation of the Ras/Mapk pathway protects cells from hid-induced apoptosis, both by transcriptional repression of hid and by phosphorylation of Hid protein by Mapk. Because Torso, which encodes a receptor tyrosine kinase (RTK), is known to activate Ras and Mapk, tests were performed to see whether manipulation of active Mapk levels using a gain-of-function allele, MapkSem, can suppress hid expression and cell death in bicoid mutants. However, this was found not to be the case. Thus, torso appears to protect mis-specified cells independently of Mapk activation (Werz, 2005).

The hid bicoid double mutant analysis reveals that the transformation of anterior into posterior identity expands beyond the telson, and that this expansion undergoes hid-induced cell death in bicoid single mutants. The rescued cells secrete larval cuticle elements, suggesting that mis-specified cells have the developmental capacity to terminally differentiate. However, in hid+ background, they instead die, presumably because equivalent survival signals are lacking. It is proposed that mis-specified cells undergo cell death if no alternative survival pathway is provided to protect them (Werz, 2005).

An alternative survival mechanism might also operate in other developmental mutants where transformation rather than cell death occurs. Mutations in the sev RTK and its ligand boss result in transformation of the R7 photoreceptor cell into a non-neuronal cone cell. Survival of this cell could be mediated by the Drosophila Egf receptor (Egfr), another RTK, which is required to maintain cell survival in the developing eye disc. Accordingly, activation of the Ras/Mapk pathway by Egfr would inhibit hid expression and support survival of the presumptive R7 photoreceptor cell. This interpretation is also consistent with observations that egfr- clones are small and undergo cell death, and that this death can be suppressed in hid mutants. Thus, transformation of the R7 photoreceptor to a cone cell rather than R7 cell death in sev and boss mutants could occur because of survival signaling by the Egfr (Werz, 2005).

The hid bicoid double mutant analysis suggests that mis-specified cells can continue to develop and differentiate. Yet, they die. Presumably, this cell death protects the organism from potentially dangerous cells. For example, it is conceivable that in mammals, surviving mis-specified cells might lie dormant in the host organism for years. During this time, they might acquire additional genetic alterations that could drive the progressive transformation of these cells into malignant cancer. In wild-type embryos, mis-specification probably occurs in cells in isolation, and elimination of these cells does not interfere with development and survival of the organism. Only in extreme situations, such as the patterning mutants analyzed in this study, is the mis-specification caused by aberrant development so severe that the affected organism dies (Werz, 2005).

The cause of mis-specification in each segmentation mutant is different. Usually, the expression of other segmentation genes is shifted and expanded, resulting in flattened gradients. Yet, irrespective of the cause of mis-specification, most of these mutants have in common that they induce hid expression. It is currently unknown how the mis-specified fate of cells is recognized, and how hid expression is induced. One possibility might be that the protein gradients established by bicoid+ and oskar+, as well as other segmentation genes are used as readout for proper cellular specification. The steepness of protein gradients as a means to determine life or death decisions has recently been proposed. Such a model would imply that cells are able to determine their position in a graded field and compare this readout with their neighbors. Because in bicoid and oskar mutants these gradients do not form, the concentration difference between neighboring cells would be zero. If the concentration difference between two neighboring cells is below a crucial threshold, they induce the expression of hid and undergo cell death. This model could also explain embryonic pattern repair, which was described in embryos that express six copies of the bicoid gene. In these embryos, the head and thorax primordia are expanded because of the presence of six copies of bicoid. However, this expansion is corrected for by induction of cell death, and relatively normal larvae develop. In this case, the Bicoid protein gradient does form, but would be flatter compared with wild type. Thus, the concentration difference between neighboring cells would be below a critical threshold, sufficient to induce hid-dependent cell death. However, it is largely unknown how cells compare their position in a graded field with those of their neighbors. It has been proposed that short-range cell interactions mediated via the cell-surface proteins Capricious and Tartan provide cues that support cell survival during wing development. Cells unable to participate in these interactions are eliminated by cell death. It is unclear, however, whether short-range interactions are sufficient to explain the cell death phenotype in bicoid and oskar mutants (Werz, 2005).

Irrespective of the underlying mechanism for sensing mis-specification, the current results highlight the role of an active gene-directed process that removes mis-specified cells during development. However, if a survival mechanism is provided, mis-specified cells can survive and adopt a different fate. In wild-type embryos, mis-specification probably occurs in cells in isolation, and hence is difficult to study. However, in bicoid and oskar mutants, large regions of neighboring cells are mis-specified and undergo cell death simultaneously, providing a unique opportunity to clarify the signals that initiate cell death in situations where cells are developmentally mis-specified (Werz, 2005).

A translation-independent role of oskar RNA in early Drosophila oogenesis

The Drosophila maternal effect gene oskar encodes the posterior determinant responsible for the formation of the posterior pole plasm in the egg, and thus of the abdomen and germline of the future fly. Previously identified oskar mutants give rise to offspring that lack both abdominal segments and a germline, thus defining the 'posterior group phenotype'. Common to these classical oskar alleles is that they all produce significant amounts of oskar mRNA. By contrast, two new oskar mutants in which oskar RNA levels are strongly reduced or undetectable are sterile, because of an early arrest of oogenesis. This egg-less phenotype is complemented by oskar nonsense mutant alleles, as well as by oskar transgenes, the protein-coding capacities of which have been annulled. Moreover, expression of the oskar 3' untranslated region (3'UTR) is sufficient to rescue the egg-less defect of the RNA null mutant. This analysis thus reveals an unexpected role for oskar RNA during early oogenesis, independent of Oskar protein. These findings indicate that oskar RNA acts as a scaffold or regulatory RNA essential for development of the oocyte (Jenny, 2006).

The new oskar alleles oskA87 and osk187 reveal that oskar plays a crucial role during early oogenesis. This early function of oskar is unusual, since it is mediated by oskar RNA, and not by Oskar protein. The low but detectable amount of oskar RNA observed in osk187 mutants indicates that a threshold exists, below which egg chambers fail to develop. These findings raise the possibility that many protein encoding genes may fulfill additional non-coding functions and thus increase the spectrum of gene function (Jenny, 2006).

What might the function of oskar RNA be during early oogenesis? Oskar protein serves as a scaffold for the assembly of cytoplasmic structures essential for germline development, the polar granules, at the posterior pole of the oocyte and embryo. During early oogenesis, oskar RNA might provide a similar scaffold function for assembly of cytoplasmic complexes essential for the progression of oocyte development. RNAs are associated with proteins in ribonucleoprotein complexes during most of their existence, from their emergence as nascent transcripts in the nucleus, during nucleo-cytoplamic transport, to their final cytoplasmic localization, translation or degradation. Staufen protein requires oskar RNA for its transport from the nurse cells into the oocyte. In addition, it is the oskar 3'UTR that mediates accumulation of Staufen protein in the oocyte. This reveals that the well-known mutual interdependence of Staufen and oskar mRNA in their localization during oogenesis is mediated by interaction of Staufen with the oskar 3'UTR. Of the candidate proteins examined, only Staufen showed a clearly altered distribution in the oskar RNA null mutant, yet Staufen itself does not appear to play a role during early oogenesis. It is therefore reasonable to assume that oskar RNA acts as a structural partner for the transport into the oocyte of additional, so far unidentified proteins or RNAs essential for its development. In this regard it is interesting to note that both VgT RNA and the non-coding Xlsirts RNA have been shown to mediate anchoring of several RNAs at the vegetal pole of the Xenopus oocyte (Jenny, 2006).

Alternative functions of the oskar 3'UTR are also plausible. In particular, the oskar 3'UTR might bind and sequester a negative regulator that, in its free form (i.e. in an oskar RNA null background), inhibits early oogenesis. One candidate that has been shown to bind to the oskar 3'UTR is the translational regulator Bruno. However, overexpression of Bruno - at least in the presence of wild-type levels of oskar mRNA - does not cause a phenotype similar to that of the oskar RNA null mutant. Thus, to fully understand the mechanism of oogenesis arrest resulting from absence of oskar mRNA, it will be important to identify other proteins and RNAs binding to the oskar 3'UTR that are required for egg chamber development (Jenny, 2006).

Oskar controls morphology of polar granules and nuclear bodies in Drosophila

Germ cell formation in Drosophila relies on polar granules, which are large ribonucleoprotein complexes found at the posterior end of the embryo. The granules undergo characteristic changes in morphology during development, including the assembly of multiple spherical bodies from smaller precursors. Several polar granule components, both protein and RNA, have been identified. One of these, the protein Oskar, acts to initiate granule formation during oogenesis and to recruit other granule components. To investigate whether Oskar has a continuing role in organization of the granules and control of their morphology, advantage was taken of species-specific differences in polar granule structure. The polar granules of D. immigrans fuse into a single large oblong aggregate, as opposed to the multiple, distinct, spherical granules of D. melanogaster embryos. D. immigrans oskar rescues the body patterning and pole cell defects of embryos from D. melanogaster oskar- mothers, and converts the morphology of the polar granules to that of D. immigrans. The nuclear bodies, which are structures that appear to be closely related to polar granules, are also converted to the D. immigrans type morphology. It is concluded that oskar plays a persistent and central role in the polar granules, not only initiating their formation but also controlling their organization and morphology (Jones, 2007).

A functional link between localized Oskar, dynamic microtubules, and endocytosis

Many cell types including developing oocytes, fibroblasts, epithelia and neurons use mRNA localization as a means to establish polarity. The Drosophila oocyte has served as a useful model in dissecting the mechanism of mRNA localization. The polarity of the oocyte is established by the specific localization of three critical mRNAs - oskar, bicoid and gurken. The localization of these mRNAs requires microtubule integrity, and the activity of microtubule motors. However, the precise organization of the oocyte microtubule cytoskeleton remains an open question. In order to examine the polarity of oocyte microtubules, the localization of canonical microtubule plus end binding proteins, EB1 and CLIP-190, was visualized. Both proteins were enriched at the posterior of the oocyte, with additional foci detected within the oocyte cytoplasm and along the cortex. Surprisingly, however, it was found that this asymmetric distribution of EB1 and CLIP-190 was not essential for oskar mRNA localization. However, Oskar protein was required for recruiting the plus end binding proteins to the oocyte posterior. Lastly, these results suggest that the enrichment of growing microtubules at the posterior pole functions to promote high levels of endocytosis in this region of the cell. Thus, multiple polarity-determining pathways are functionally linked in the Drosophila oocytes (Sanghavi, 2012).

Although posterior EB1 and CLIP-190 are not essential for oskar mRNA localization, Oskar protein is required for recruiting EB1 and CLIP-190 to the posterior pole. The mechanism by which Oskar performs this function is less clear. Oskar protein has been shown to induce the formation of long F-actin fibers at the posterior of the oocyte. It was therefore hypothesized that these actin fibers might recruit growing microtubule plus ends. Consistent with this hypothesis, treatment of ovaries with Lat-A resulted in loss of EB1 and CLIP-190 foci from the posterior pole. Under these treatment conditions however, Oskar protein was still detected at the posterior. This suggests that EB1 and CLIP-190 recruitment to the posterior pole occurs downstream of Oskar protein localization and requires integrity of the actin cytoskeleton. How might actin function in recruiting EB1 and CLIP-190 to the posterior pole? One possibility is via the activity of cross-linking proteins that have been shown to bind microtubule plus ends as well actin. Another possibility is that the actin network might be required for polarizing oocyte microtubules. Consistent with this notion, mutations in the actin nucleators, capu and spire, result in oocytes that lack microtubule polarity (Sanghavi, 2012).

The results also suggest that Oskar is not sufficient for recruiting EB1 and CLIP-190. Ectopic expression of Oskar protein along the entire oocyte cortex or within a central focus failed to efficiently recruit EB1 and CLIP-190. It is therefore likely that multiple factors function to recruit these plus end binding proteins to the posterior pole. Curiously, however, mis-expression of Oskar reduced the posterior accumulation of EB1. Posterior localization of CLIP-190 was also affected under these conditions, but to a lesser degree. The reason for this phenotype is unclear, but it nonetheless suggests that mis-expression of Oskar affects the dynamics of microtubules within the oocyte (Sanghavi, 2012).

In contrast to EB1 and CLIP-190, the ectopic Oskar focus contained high levels of Khc. A similar result was also observed by Z using Kin:βgal as a reporter. At present it is unclear whether Khc and Kin:βgal localization are truly indicative of microtubule plus ends. In addition to the ectopic Oskar focus, Khc was found in the center of par-1 mutant oocytes, but EB1 and CLIP-190 were not. Similarly, Khc, but not EB1 and CLIP-190, was localized to the posterior in oskar protein-null mutants. As mentioned previously, Khc and Kin:βgal might label a population of microtubules that are polarized, yet relatively static in vivo. Further work is needed to explore this possibility (Sanghavi, 2012).

It has been demonstrated that endocytosis within the oocyte is asymmetric, occurring at a higher level at the posterior). Oskar protein is required for maintaining this polarized distribution of endocytosis. The mechanism by which Oskar performs this function, however, is largely unknown (Sanghavi, 2012).

The current findings indicate that growing microtubule plus ends are recruited to the oocyte posterior by Oskar. In this regard, it is interesting to note that CLIP-170, the vertebrate homolog of CLIP-190, was initially identified as a protein that directly bound to endocytic vesicles as well as microtubule plus ends. More recently, CLIP-170 was also shown to function at microtubule plus ends to stimulate phagocytosis in macrophage cells and to promote the capture and minus end transport of melanophores in Xenopus). These findings highlight the importance of microtubule plus ends and CLIP-170 in vesicle transport (Sanghavi, 2012).

Based on these observations, it is hypothesized that Oskar promotes endocytosis by recruiting microtubule plus end binding proteins to the posterior pole. Consistent with this hypothesis, EB1 and CLIP-190 localization, as well as endocytosis was drastically reduced at the oocyte posterior in oskar protein-null and staufen mutant. Treatment of egg chambers with a microtubule-stabilizing drug also had a similar effect on microtubule plus end localization and endocytosis. These results suggest an intimate relationship between dynamic microtubules and efficient endocytosis at the posterior pole. One possibility is that the dynamic microtubules themselves are important for posterior endocytosis. Alternatively, EB1 and CLIP-190 might deliver factors to the posterior that facilitate high endocytic activity (Sanghavi, 2012).

These results indicate that oocyte microtubules are polarized, with an enrichment of plus ends at the posterior pole. The results also suggest that the primarily function of the polarized microtubule network is to promote high endocytic activity at the oocyte posterior (Sanghavi, 2012).

oskarRNA plays multiple noncoding roles to support oogenesis and maintain integrity of the germline/soma distinction

The Drosophila oskar (osk) mRNA is unusual in that it has both coding and noncoding functions. As an mRNA, osk encodes a protein required for embryonic patterning and germ cell formation. Independent of that function, the absence of osk mRNA disrupts formation of the karyosome and blocks progression through oogenesis. This study shows that loss of osk mRNA also affects the distribution of regulatory proteins, relaxing their association with large RNPs within the germline, and allowing them to accumulate in the somatic follicle cells. This and other noncoding functions of the osk mRNA are mediated by multiple sequence elements with distinct roles. One role, provided by numerous binding sites in two distinct regions of the osk 3' UTR, was to sequester the translational regulator Bruno (Bru), which itself controlled translation of osk mRNA. This defined a novel regulatory circuit, with Bru restricting the activity of osk, and osk in turn restricting the activity of Bru. Other functional elements, which did not bind Bru and were positioned close to the 3' end of the RNA, acted in the oocyte and were essential. Despite the different roles played by the different types of elements contributing to RNA function, mutation of any led to accumulation of the germline regulatory factors in the follicle cells (Kanke, 2015).

Translational activation of Oskar mRNA: Reevaluation of the role and importance of a 5' regulatory element

Local translation of oskar (osk) mRNA at the posterior pole of the Drosophila oocyte is essential for axial patterning of the embryo, and is achieved by a program of translational repression, mRNA localization, and translational activation. Multiple forms of repression are used to prevent Oskar protein from accumulating at sites other than the oocyte posterior. Activation is mediated by several types of cis-acting elements, which presumably control different forms of activation. This study characterized a 5' element, positioned in the coding region for the Long Osk isoform and in the extended 5' UTR for translation of the Short Osk isoform. This element was previously thought to be essential for osk mRNA translation, with a role in posterior-specific release from repression. From this work, which includes assays which separate the effects of mutations on RNA regulatory elements and protein coding capacity, it was found that the element is not essential, and the study concludes that there is no evidence supporting a role for the element only at the posterior of the oocyte. The 5' element has a redundant role, and is only required when Long Osk is not translated from the same mRNA. Mutations in the element do disrupt the anchoring function of Long Osk protein through their effects on the amino acid sequence, a confounding influence on interpretation of previous experiments (Kanke, 2015).


REFERENCES

Abdu, U., Bar D. and Schüpbach T. (2006). spn-F encodes a novel protein that affects oocyte patterning and bristle morphology in Drosophila. Development 133: 1477-1484. PubMed Citation: 16540510

Ahuja, A. and Extavour, C. G. (2014). Patterns of molecular evolution of the germ line specification gene oskar suggest that a novel domain may contribute to functional divergence in Drosophila. Dev Genes Evol. [Epub ahead of print] PubMed ID: 24407548

Alexandrov, A., Colognori, D., Shu, M. D. and Steitz, J. A. (2012). Human spliceosomal protein CWC22 plays a role in coupling splicing to exon junction complex deposition and nonsense-mediated decay. Proc Natl Acad Sci U S A 109(52): 21313-21318. PubMed ID: 23236153

Babu, K., Cai, Y., Bahri, S. Yang, X. and Chia, W. (2004). Roles of Bifocal, Homer, and F-actin in anchoring Oskar to the posterior cortex of Drosophila oocytes. Genes Dev. 18: 138-143. 14752008

Barbosa, I., Haque, N., Fiorini, F., Barrandon, C., Tomasetto, C., Blanchette, M. and Le Hir, H. (2012). Human CWC22 escorts the helicase eIF4AIII to spliceosomes and promotes exon junction complex assembly. Nat Struct Mol Biol 19(10): 983-990. PubMed ID: 22961380

Baum, B., Li, W. and Perrimon, N. (2000). A cyclase-associated protein regulates actin and cell polarity during Drosophila oogenesis and in yeast. Curr. Biol. 10: 964-973. 10985383

Benoit, B., et al. (1999). The Drosophila poly(A)-binding protein II is ubiquitous throughout Drosophila development and has the same function in mRNA polyadenylation as its bovine homolog in vitro. Nucleic Acids Res. 27(19): 3771-8. Medline abstract: 10481015

Benoit, B., et al. (2005). An essential cytoplasmic function for the nuclear poly(A) binding protein, PABP2, in poly(A) tail length control and early development in Drosophila. Dev. Cell. 9(4): 511-22. Medline abstract: 16198293

Besse, F., et al. (2009). Drosophila PTB promotes formation of high-order RNP particles and represses oskar translation. Genes Dev. 23(2): 195-207. PubMed Citation: 19131435

Braat, A. K., et al. (2004). Localization-dependent Oskar protein accumulation control after the initiation of translation. Dev. Cell 7: 125-131. 15239960

Breitwieser, W., et al. (1996). Oskar protein interaction with Vasa represents an essential step in polar granule assembly. Genes. Dev. 10: 2179-88. PubMed Citation: 8804312

Brendza, R. P., et al. (2000). A function for kinesin I in the posterior transport of Oskar mRNA and Staufen protein. Science 289(5487): 2120-2. 11000113

Broyer, R., Monfort, E. and Wilhelm, J. E. (2016). Cup regulates oskar mRNA stability during oogenesis. Dev Biol [Epub ahead of print]. PubMed ID: 27554167

Bullock, S. L. and Ish-Horowicz, D. (2001). Conserved signals and machinery for RNA transport in Drosophila oogenesis and embryogenesis. Nature 414(6864): 611-6. 11740552

Bullock, S. L., Ringel, I., Ish-Horowicz, D. and Lukavsky, P. J. (2010). A'-form RNA helices are required for cytoplasmic mRNA transport in Drosophila. Nat Struct Mol Biol 17: 703-709. PubMed ID: 20473315

Castagnetti, S., et al. (2000). Control of oskar mRNA translation by Bruno in a novel cell-free system from Drosophila ovaries. Development 127: 1063-1068. PubMed Citation: 10662645

Castagnetti, S. and Ephrussi, A. (2003). Orb and a long poly(A) tail are required for efficient oskar translation at the posterior pole of the Drosophila oocyte. Development 130: 835-843. 12538512

Cavey, M., Hijal, S., Zhang, X. and Suter, B. (2005). Drosophila valois encodes a divergent WD protein that is required for Vasa localization and Oskar protein accumulation. Development 132(3):459-68. 15634703

Chang, J. S., Tan, L. and Schedl P. (1999). The Drosophila CPEB homolog, Orb, is required for Oskar protein expression in oocytes. Dev. Biol. 215(1): 91-106. PubMed Citation: 10525352

Chekulaeva, M., Hentze, M. W. and Ephrussi, A. (2006). Bruno acts as a dual repressor of oskar translation, promoting mRNA oligomerization and formation of silencing particles. Cell 124(3): 521-33. 16469699

Christerson, L. B. and McKearin, D. M. (1994). orb is required for anteroposterior and dorsoventral patterning during Drosophila oogenesis. Genes Dev. 8: 614-628. PubMed Citation: 7926753

Clegg, N. J., et al. (1997). maelstrom is required for an early step in the establishment of Drosophila oocyte polarity: posterior localization of grk mRNA. Development 124(22): 4661-4671. PubMed Citation: 9409682

Coller, J. and Parker, R. (2005). General translational repression by activators of mRNA decapping. Cell 122: 875-886. 16179257

Cook, H. A., Koppetsch, B. S., Wu, J. and Theurkauf, W. E. (2004). The Drosophila SDE3 homolog armitage is required for oskar mRNA silencing and embryonic axis specification. Cell 116(6): 817-29. 15035984

Cornelis, S. (2005). UNR translation can be driven by an IRES element that is negatively regulated by polypyrimidine tract binding protein. Nucleic Acids Res. 33: 3095-3108. PubMed Citation: 15928332

Coutelis, J. B. and Ephrussi, A. (2007). Rab6 mediates membrane organization and determinant localization during Drosophila oogenesis. Development 134(7): 1419-30. Medline abstract: 17329360

Cox, R. T. and Spradling, A. C. (2003). A Balbiani body and the fusome mediate mitochondrial inheritance during Drosophila oogenesis. Development 130: 1579-1590. 12620983

Dahanukar, A., Walker, J. A. and Wharton, R. P. (1999). Smaug, a novel RNA-binding protein that operates a translational switch in Drosophila. Mol. Cell 4: 209-218

Ding, D., et al. (1993). Dynamic Hsp83 RNA localization during Drosophila oogenesis and embryogenesis. Mol. Cell Biol. 13: 3773-81

Ding, D., Whittaker, K. L. and Lipshitz, H. D. (1994). Mitochondrially encoded 16S large ribosomal RNA is concentrated in the posterior polar plasm of early Drosophila embryos but is not required for pole cell formation. Dev. Biol. 163: 503-15

Deng, W. and Lin, H. (1997). Spectrosomes and fusomes anchor mitotic spindles during asymmetric germ cell divisions and facilitate the formation of a polarized microtubule array for oocyte specification in Drosophila. Dev. Biol. 189(1): 79-94

Doerflinger, H., Benton, R., Torres, I. L., Zwart, M. F. and St Johnston, D. (2006). Drosophila anterior-posterior polarity requires actin-dependent PAR-1 recruitment to the oocyte posterior. Curr. Biol. 16(11): 1090-5. 16753562

Dold, A., Han, H., Liu, N., Hildebrandt, A., Bruggemann, M., Ruckle, C., Hanel, H., Busch, A., Beli, P., Zarnack, K., Konig, J., Roignant, J. Y. and Lasko, P. (2020). Makorin 1 controls embryonic patterning by alleviating Bruno1-mediated repression of oskar translation. PLoS Genet 16(1): e1008581. PubMed ID: 31978041

Dollar, G., Struckhoff, E., Michaud, J. and Cohen, R. S. (2002). Rab11 polarization of the Drosophila oocyte: a novel link between membrane trafficking, microtubule organization, and oskar mRNA localization and translation. Development 129: 517-526. 11807042

Doyle, M. and Kiebler, M. A. (2012). A numbers game underpins cytoplasmic mRNA transport. Nat Cell Biol 14: 333-335. PubMed ID: 22469827

Dubin-Bar D., et al. (2008). The Drosophila IKK-related kinase (Ik2) and Spindle-F proteins are part of a complex that regulates cytoskeleton organization during oogenesis. BMC Cell Biol. 9: 51. PubMed Citation: 18796167

Dubin-Bar, D., Bitan, A., Bakhrat, A., Amsalem, S. and Abdu, U. (2011). Drosophila javelin-like encodes a novel microtubule-associated protein and is required for mRNA localization during oogenesis. Development 138(21): 4661-71. PubMed Citation: 21989913

Dubnau, J., et al. (2003). The staufen/pumilio pathway is involved in Drosophila long-term memory. Curr. Biol. 13: 286-296. 12593794

Ephrussi, A. Dickinson, L. K. and Lehmann, R. (1991). Oskar organizes the germ plasm and directs localization of the posterior determinant nanos. Cell 66: 37-50

Ephrussi, A. and Lehmann, R. (1992). Induction of germ cell formation by oskar. Nature 358: 387-92

Erdelyi, M., et al. (1995). Requirement for Drosophila cytoplasmic tropomyosin in oskar mRNA localization. Nature 377: 524-527

Fan, S. J., Marchand, V. and Ephrussi, A. (2011). Drosophila Ge-1 promotes P body formation and oskar mRNA localization. PLoS One. 6(5): e20612. PubMed Citation: 21655181

Ferrandon, D., et al. (1994). Staufen protein associates with the 3'UTR of Bicoid mRNA to form particles that move in a microtubule-dependent manner. Cell 79: 1221-1232

Fischer-Vize, J. A., Rubin, J. M. and Lehmann, R. (1992). The fat facets gene is required for Drosophila eye and embryo development. Development 116: 985-1000

Frydman, H. M. and Spradling, A. C. (2001). The receptor-like tyrosine phosphatase Lar is required for epithelial planar polarity and for axis determination within Drosophila ovarian follicles. Development 128: 3209-3220. 11688569

Fusco, D., et al. (2003). Single mRNA molecules demonstrate probabilistic movement in living mammalian cells. Curr. Biol. 13: 161-167. PubMed Citation: 12546792

Gajewski, K. M. and Schulz, R. A. (1995). Requirement of the ETS domain transcription factor D-ELG for egg chamber patterning and development during Drosophila oogenesis. Oncogene 11: 1033-1040

Galban, S. (2008). RNA-binding proteins HuR and PTB promote the translation of hypoxia-inducible factor 1α. Mol. Cell. Biol. 28: 93-107. PubMed Citation: 17967866

Gaspar, I., Yu, Y. V., Cotton, S. L., Kim, D. H., Ephrussi, A. and Welte, M. A. (2014). Klar ensures thermal robustness of oskar localization by restraining RNP motility. J Cell Biol 206: 199-215. PubMed ID: 25049271

Geng, C. and MacDonald, P. M. (2006). Imp associates with Squid and Hrp48 and contributes to localized expression of gurken in the oocyte. Mol. Cell. Biol. 26: 9508-9516. PubMed citation: 17030623

Ghosh, S., Marchand, V., Gaspar, I. and Ephrussi, A. (2012). Control of RNP motility and localization by a splicing-dependent structure in oskar mRNA. Nat Struct Mol Biol 19: 441-449. PubMed ID: 22426546

Gillespie, D. E., and Berg, C. A. (1995). Homeless is required for RNA localization in Drosophila oogenesis and encodes a new member of the DE-H family of RNA-dependent ATPases. Genes Dev 9: 2495-2508

Goltsev, Y., et al. (2004). Different combinations of gap repressors for common stripes in Anopheles and Drosophila embryos. Dev. Biol. 275: 435-446. 15501229

Gonsalvez, G. B., et al. (2010). Sm proteins specify germ cell fate by facilitating oskar mRNA localization. Development 137(14): 2341-51. PubMed Citation: 20570937

Guichet, A., Peri, F. and Roth, S. (2001). Stable anterior anchoring of the oocyte nucleus is required to establish dorsoventral polarity of the Drosophila egg. Dev. Bio. 237: 93-106. 11518508

Gunkel, N., Yano, T., Markussen, F.-H., Olsen, L. C. and Ephrussi, A. (1998). Localization-dependent translation requires a functional interaction between the 5' and 3' ends of oskar mRNA. Genes Dev. 12: 1652-1664

Hachet, O. and Ephrussi, A. (2001). Drosophila Y14 shuttles to the posterior of the oocyte and is required for oskar mRNA transport. Cur. Bio. 11: 1666-1674. 11696323

Hachet, O. and Ephrussi, A. (2004). Splicing of oskar RNA in the nucleus is coupled to its cytoplasmic localization. Nature 428: 959-963. PubMed Citation: 15118729

Hay, B., Jan, L. Y. and Jan, Y. N. (1990). Localization of vasa, a component of Drosophila polar granules, in maternal-effect mutants that alter embryonic anteroposterior polarity. Development 109: 425-33

Huang, F., Paulson, A., Dutta, A., Venkatesh, S., Smolle, M., Abmayr, S. M. and Workman, J. L. (2014). Histone acetyltransferase Enok regulates oocyte polarization by promoting expression of the actin nucleation factor spire. Genes Dev 28: 2750-2763. PubMed ID: 25512562

Hurd, T.R., Herrmann, B., Sauerwald, J., Sanny, J., Grosch, M. and Lehmann, R. (2016). Long Oskar controls mitochondrial inheritance in Drosophila melanogaster. Dev Cell 39: 560-571. PubMed ID: 27923120

Huynh, J.-R., et al. (2001). PAR-1 is required for the maintenance of oocyte fate in Drosophila. Development 128: 1201-1209. 11245586

Huynh, J.-R., et al. (2004). The Drosophila hnRNPA/B homolog, Hrp48, is specifically required for a distinct step in osk mRNA localization. Dev. Cell 6: 625-635. 15130488

Irion, U., et al. (2006). Miranda couples oskar mRNA/Staufen complexes to the bicoid mRNA localization pathway. Dev. Biol. 297: 522-533. Medline abstract: 16905128

Jankovics, F., Sinka, R. and Erdelyi, M. (2001). An interaction type of genetic screen reveals a role of the Rab11 gene in oskar mRNA localization in the developing Drosophila melanogaster oocyte. Genetics 158: 1177-1188. 11454766

Jankovics, F., et al. (2002). MOESIN crosslinks actin and cell membrane in Drosophila oocytes and is required for OSKAR anchoring. Curr. Biol. 12: 2060-2065. 12477397

Januschke, J., Nicolas, E., Compagnon, J., Formstecher, E., Goud, B. and Guichet, A. (2007). Rab6 and the secretory pathway affect oocyte polarity in Drosophila. Development 134(19): 3419-25. Medline abstract: 17827179

Jang, S. K. (2006). Internal initiation: IRES elements of picornaviruses and hepatitis c virus. Virus Res. 119: 2-15. PubMed Citation: 16377015

Jenny, A., Hachet, O., Zavorszky, P., Cyrklaff, A., Weston, M. D., Johnston, D. S., Erdelyi, M. and Ephrussi, A. (2006). A translation-independent role of oskar RNA in early Drosophila oogenesis. Development 133: 2827-2833. PubMed ID: 16835436

Jeong, E. B., Jeong, S. S., Cho, E. and Kim, E. Y. (2019). Makorin 1 is required for Drosophila oogenesis by regulating insulin/Tor signaling. PLoS One 14(4): e0215688. PubMed ID: 31009498

Jeske, M., Meyer, S., Temme, C., Freudenreich, D. and Wahle, E. (2006). Rapid ATP-dependent deadenylation of nanos mRNA in a cell-free system from Drosophila embryos. J. Biol. Chem. 281: 25124-25133. Medline abstract: 16793774

Jeske, M., Bordi, M., Glatt, S., Muller, S., Rybin, V., Muller, C. W. and Ephrussi, A. (2015). The crystal structure of the Drosophila germline inducer Oskar identifies two domains with distinct Vasa helicase- and RNA-binding activities. Cell Rep 12: 587-598. PubMed ID: 26190108

Jeske, M., Muller, C. W., Ephrussi, A. (2017). The LOTUS domain is a conserved DEAD-box RNA helicase regulator essential for the recruitment of Vasa to the germ plasm and nuage. Genes Dev. 31(9):939-952. PubMed ID: 28536148

Jones, J. R. and Macdonald, P. M. (2007). Oskar controls morphology of polar granules and nuclear bodies in Drosophila. Development 134(2): 233-6. Medline abstract: 17151014

Juge, F., Zaessinger, S., Temme, C., Wahle, E. and Simonelig, M. (2002). Control of poly(A) polymerase level is essential to cytoplasmic polyadenylation and early development in Drosophila. EMBO J. 21(23): 6603-13. 12456666

Kalifa, Y., Huang, T., Rosen, L. N., Chatterjee, S. and Gavis, E. R. (2006). Glorund, a Drosophila hnRNP F/H Homolog, is an ovarian repressor of nanos translation. Dev. Cell 10: 291-301. Medline abstract: 16516833

Kanamori T., et al. (2008). β-Catenin asymmetry is regulated by PLA1 and retrograde traffic in C. elegans stem cell divisions. EMBO J. 27: 1647-1657. PubMed Citation: 18497747

Kanke, M., Jambor, H., Reich, J., Marches, B., Gstir, R., Ryu, Y. H., Ephrussi, A. and Macdonald, P. M. (2015). oskar RNA plays multiple noncoding roles to support oogenesis and maintain integrity of the germline/soma distinction. RNA 21: 1096-1109. PubMed ID: 25862242

Kanke, M. and Macdonald, P. M. (2015). Translational activation of Oskar mRNA: Reevaluation of the role and importance of a 5' regulatory element. PLoS One 10: e0125849. PubMed ID: 25938537

Kim, J., Lee, J., Lee, S., Lee, B. and Kim-Ha, J. (2014). Phylogenetic comparison of oskar mRNA localization signals. Biochem Biophys Res Commun. 444(1): 98-103. PubMed ID: 24440702

Kim-Ha, J., Smith, J.L. and Macdonald, P.M. (1991). oskar mRNA is localized to the posterior pole of the Drosophila oocyte. Cell 66: 23-35

Kim-Ha, J., et al. (1993). Multiple RNA regulatory elements mediate distinct steps in localization of oskar mRNA. Development 119: 169-78

Kim-Ha, J., Kerr, K. and Macdonald, P. M. (1995). Translational regulation of oskar mRNA by bruno, an ovarian RNA-binding protein, is essential. Cell 81: 403-412

Kobayashi, S., et al. (1995). Mislocalization of oskar product in the anterior pole results in ectopic localization of mitochondrial large ribosomal RNA in Drosophila embryos. Dev. Biol. 169: 384-6

Kress, T. L., Yoon, Y. J. and Mowry, K. L. (2004). Nuclear RNP complex assembly initiates cytoplasmic RNA localization. J. Cell Biol. 165: 203-211. PubMed Citation: 15096527

Lane, M. E. and Kalderon, D. (1994). RNA localization along the anteroposterior axis of the Drosophila oocyte requires PKA-mediated signal transduction to direct normal microtubule organization. Genes Dev 8: 2986-2995

Lantz, V. A., Clemens, S. E. and Miller, K. G. (1999). The actin cytoskeleton is required for maintenance of posterior pole plasm components in the Drosophila embryo. Mech. Dev. 85: 111-122

Lasko, P. F. and Ashburner, M. (1990). Posterior localization of vasa protein correlates with, but is not sufficient for, pole cell development. Genes Dev. 4(6): 905-21. PubMed citation: 2384213

Lewis, R. A., Gagnon, J. A., and Mowry, K. L. (2008). PTB/hnRNP I is required for RNP remodeling during RNA localization in Xenopus oocytes. Mol. Cell. Biol. 28: 678-686. PubMed Citation: 18039852

Littleton, J. T., Stern, M., Schulze, K., Perin, M. and Bellen, H. J. (1993). Mutational analysis of Drosophila synaptotagmin dem-onstrates its essential role in Ca(21)-activated neurotransmitter release. Cell 74: 1125-1134. PubMed Citation: 8104705

Lehmann, R. and Nusslein-Volhard, C. (1986). Abdominal segmentation, pole cell formation, and embryonic polarity require the localized activity of oskar, a maternal gene in Drosophila. Cell 47: 141-52. PubMed Citation: 3093084

Lehmann, R. and Nusslein-Volhard, C. (1991). The pole plasm is required for germ cell formation and contains the determinant of posterior polarity, encoded by nanos. Development 112: 679-91. PubMed Citation: 1935684

Lehmann, R. and Ephrussi, A. (1994). Germ plasm formation and germ cell determination in Drosophila. Ciba Found Symp 182: 282-296; discussion 296-300. PubMed Citation: 7530619

Lie, Y. S. and Macdonald, P. M. (1999a). Apontic binds the translational repressor Bruno and is implicated in regulation of oskar mRNA translation. Development 126: 1129-1138. PubMed Citation: 10021333

Lie, Y. S. and Macdonald, P. M. (1999b). Translational regulation of oskar mRNA occurs independent of the cap and poly(A) tail in Drosophila ovarian extracts. Development 126: 4989-4996. PubMed Citation: 10529417

Little, S. C., Sinsimer, K. S., Lee, J. J., Wieschaus, E. F. and Gavis, E. R. (2015). Independent and coordinate trafficking of single Drosophila germ plasm mRNAs. Nat Cell Biol 17(5):558-68. PubMed ID: 25848747

Loiseau, P., et al. (2010). Drosophila PAT1 is required for Kinesin-1 to transport cargo and to maximize its motility. Development 137(16): 2763-72. PubMed Citation: 20630947

Lu, W., Lakonishok, M., Liu, R., Billington, N., Rich, A., Glotzer, M., Sellers, J. R. and Gelfand, V. I. (2020). Competition between kinesin-1 and myosin-V defines Drosophila posterior determination. Elife 9. PubMed ID: 32057294

Lynch, J. A., et al. (2011). The phylogenetic origin of oskar coincided with the origin of maternally provisioned germ plasm and pole cells at the base of the Holometabola. PLoS Genet. 7(4): e1002029. PubMed Citation: 21552321

Macdonald, P. M., Kanke, M. and Kenny, A. (2016). Community effects in regulation of translation. Elife 5 [Epub ahead of print]. PubMed ID: 27104756

Mach, J. M. and Lehmann, R. (1997). An Egalitarian-BicaudalD complex is essential for oocyte specification and axis determination in Drosophila. Genes Dev. 11: 423-435. PubMed Citation: 9042857

Manseau, L., Calley, J. and Phan, H. (1996). Profilin is required for posterior patterning of the Drosophila oocyte. Development 122: 2109-16

Mansfield, J. H., Wilhelm, J. E. and Hazelrigg, T. (2002). Ypsilon Schachtel, a Drosophila Y-box protein, acts antagonistically to Orb in the oskar mRNA localization and translation pathway. Development 129: 197-209. 11782413

Markesich, D. C., Gajewski, K. M., Nazimiec, M. E. and Beckingham, K. (2000). bicaudal encodes the Drosophila beta NAC homolog, a component of the ribosomal translational machinery. Development 127: 559-572. 10631177

Markussen, F. H., et al. (1995). Translational control of oskar generates short OSK, the isoform that induces pole plasma assembly. Development 121: 3723-3732

Megosh, H. B., Cox, D. N., Campbell, C. and Lin, H. (2006). The role of PIWI and the miRNA machinery in Drosophila germline determination. Curr. Biol. 16(19): 1884-94. Medline abstract: 16949822

Meignin, C. and Davis, I. (2008). UAP56 RNA helicase is required for axis specification and cytoplasmic mRNA localization in Drosophila. Dev. Biol. 315(1): 89-98. PubMed Citation: 18237727

Micklem, D. R., et al. (2000). Distinct roles of two conserved Staufen domains in oskar mRNA localization and translation. EMBO J. 19: 1366-1377.

Mohr, S. E., Dillon, S. T. and Boswell, R. E. (2001). The RNA-binding protein Tsunagi interacts with Mago Nashi to establish polarity and localize oskar mRNA during Drosophila oogenesis. Genes Dev. 15: 2886-2899. 11691839

Munro, T. P., Kwon, S., Schnapp, B. J. and St Johnston, D. (2006). A repeated IMP-binding motif controls oskar mRNA translation and anchoring independently of Drosophila melanogaster IMP. J Cell Biol 172(4): 577-588. PubMed ID: 16476777

Morris, J. Z., Hong, A., Lilly, M. A. and Lehmann, R. (2005). twin, a CCR4 homolog, regulates cyclin poly(A) tail length to permit Drosophila oogenesis. Development 132(6): 1165-74. Medline abstract: 15703281

Munro, T. P., Kwon, S., Schnapp, B. J. and St Johnston, D. (2006). A repeated IMP-binding motif controls oskar mRNA translation and anchoring independently of Drosophila melanogaster IMP. J. Cell Biol. 172(4): 577-88. 16476777

Murata, Y. and Wharton, R. P. (1995). Binding of pumilio to maternal hunchback mRNA is required for posterior patterning in Drosophila embryos. Cell 80: 747-756

Nakamura, A., Amikura, R., Hanyu, K. and Kobayashi, S. (2001). Me31B silences translation of oocyte-localizing RNAs through the formation of cytoplasmic RNP complex during Drosophila oogenesis. Development 128: 3233-3242. 11546740

Nakamura, A., Sato, K. and Hanyu-Nakamura, K. (2004). Drosophila Cup is an eIF4E binding protein that associates with Bruno and regulates oskar mRNA translation in oogenesis. Dev. Cell 6: 69-78. 14723848

Newmark, P. A. and Boswell, R. E. (1994). The mago nashi locus encodes an essential product required for germ plasm assembly in Drosophila. Development 120: 1303-1313. PubMed Citation: 8026338

Norvell, A., Debec, A., Finch, D., Gibson, L. and Thoma, B. (2005). Squid is required for efficient posterior localization of oskar mRNA during Drosophila oogenesis. Dev. Genes Evol. 215(7): 340-9. 15791421

Oberstrass, F. C., et al. (2005). Structure of PTB bound to RNA: Specific binding and implications for splicing regulation. Science 309: 2054-2057. PubMed Citation: 16179478

Obrdlik, A., Lin, G., Haberman, N., Ule, J. and Ephrussi, A. (2019). The transcriptome-wide landscape and modalities of EJC binding in adult Drosophila. Cell Rep 28(5): 1219-1236. PubMed ID: 31365866

Ottone, C., et al (2012). The translational repressor Cup is required for germ cell development in Drosophila. J. Cell Sci. 125(Pt 13): 3114-23. PubMed Citation: 22454519

Palacios, I. M. and St Johnston, D. (2002). Kinesin light chain-independent function of the Kinesin heavy chain in cytoplasmic streaming and posterior localisation in the Drosophila oocyte. Development 129: 5473-5485. 12403717

Pane, A., Wehr, K., Schüpbach, T. (2007). zucchini and squash encode two putative nucleases required for rasiRNA production in the Drosophila germline. Dev. Cell 12(6): 851-62. PubMed citation: 17543859

Pickering, B. M., et al. (2004). Bag-1 internal ribosome entry segment activity is promoted by structural changes mediated by poly(rC) binding protein 1 and recruitment of polypyrimidine tract binding protein 1. Mol. Cell. Biol. 24: 5595-5605. PubMed Citation: 15169918

Polesello, C., Delon, I., Valenti, P., Ferrer, P. and Payre F. (2002). Dmoesin controls actin-based cell shape and polarity during Drosophila melanogaster oogenesis. Nat. Cell Biol. 4(10): 782-9. 12360288

Pokrywka, N. J. and Stephenson, E. C. (1995). Microtubules are a general component of mRNA localization systems in Drosophila oocytes. Dev. Biol. 167: 363-70

Pokrywka, N. J., Fishbein, L. and Frederick, J. (2000). New phenotypes associated with the swallow gene of Drosophila: evidence for a general role in oocyte cytoskeletal organization. Dev. Genes Evol. 210: 426-435

Pyronnet, S. (2001). Phosphorylation of the cap-binding protein eIF4E by the MAPK-activated protein kinase Mnk1. Biochem. Pharmacol. 60(8): 1237-43. PubMed Citation: 11007962

Qualmann, B. and Kessels, M. M. (2009). New players in actin polymerization - WH2-domain-containing actin nucleators. Trends Cell Biol. 19: 276-285. PubMed Citation: 19406642

Rabouille, C., Kondo, H., Newman, R., Hui, N., Freemont, P. and Warren, G. (1998). Syntaxin 5 is a common component of the NSF- and P97-mediated reassembly pathways of Golgi cisternae from mitotic Golgi fragments in vitro. Cell 92: 603-610

Ramos, A., et al. (2000). RNA recognition by a Staufen double-stranded RNA-binding domain. EMBO J. 19: 997-1009. PubMed Citation: 10698941

Reveal, B., et al. (2010). BREs mediate both repression and activation of oskar mRNA translation and act in trans. Dev. Cell 18(3): 496-502. PubMed Citation: 20230756

Reyes, R. and Izquierdo, J. M. (2007). The RNA-binding protein PTB exerts translational control on 3'-untranslated region of the mRNA for the ATP synthase β-subunit. Biochem. Biophys. Res. Commun. 357: 1107-1112. PubMed Citation: 17466948

Riechmann, V., Gutierrez, G. J., Filardo, P., Nebreda, A. R. and Ephrussi, A. (2002). Par-1 regulates stability of the posterior determinant Oskar by phosphorylation. Nat. Cell Biol. 4: 337-342. 11951092

Riley, G. R., et al. (1991). Positive and negative control of the Antennapedia promoter P2. Dev. Suppl. 1: 177-85

Rittenhouse, K. R. and Berg, C. A. (1995). Mutations in the Drosophila gene bullwinkle cause the formation of abnormal eggshell structures and bicaudal embryos. Development 121: 3023-3033

Rongo, C., Gavis, E. R. and Lehmann, R. (1995). Localization of oskar RNA regulates oskar translation and requires Oskar protein. Development 121: 2737-2746

Roth, S., et al. (1995). cornichon and the EGF receptor signaling process are necessary for both anterior-posterior and dorsal-ventral pattern formation in Drosophila. Cell 81: 967-978

Ruden, D. M., et al. (2000). Membrane fusion proteins are required for oskar mRNA localization in the Drosophila egg chamber. Dev. Biol. 218: 314-325.

Ruohola, H., et al. (1994). Role of neurogenic genes in establishment of follicle cell fate and oocyte polarity during oogenesis in Drosophila. Cell 66: 433-449

Ryu, Y. H. and Macdonald, P. M. (2015). RNA sequences required for the noncoding function of oskar RNA also mediate regulation of Oskar protein expression by Bicoid Stability Factor. Dev Biol 407(2):211-23. PubMed ID: 26433064

Saffman, E. E., Styhler, S., Rother, K., Li, W., Richard, S. and Lasko, P. (1998). Premature translation of oskar in oocytes lacking the RNA-binding protein Bicaudal-C. Mol. Cell. Biol. 18: 4855-4862. 9671494

Sanghavi, P., Lu, S. and Gonsalvez, G. B. (2012). A functional link between localized Oskar, dynamic microtubules, and endocytosis. Dev Biol 367: 66-77. PubMed ID: 22561189

Sanghavi, P., Laxani, S., Li, X., Bullock, S. L. and Gonsalvez, G. B. (2013). Dynein associates with oskar mRNPs and is required for their efficient net plus-end localization in Drosophila oocytes. PLoS One 8: e80605. PubMed ID: 24244700

Satoh, D., Sato, D., Tsuyama, T., Saito, M., Ohkura, H., Rolls, M. M., Ishikawa, F. and Uemura, T. (2008). Spatial control of branching within dendritic arbors by dynein-dependent transport of Rab5-endosomes. Nat Cell Biol 10: 1164-1171. PubMed ID: 18758452

Sauliere, J., Haque, N., Harms, S., Barbosa, I., Blanchette, M. and Le Hir, H. (2010). The exon junction complex differentially marks spliced junctions. Nat Struct Mol Biol 17: 1269-1271. PubMed ID: 20818392

Semler, B. L. and Waterman, M. L. (2008). IRES-mediated pathways to polysomes: Nuclear versus cytoplasmic routes. Trends Microbiol. 16: 1-5. PubMed Citation: 18083033

Serano, T. L. and Cohen, R. S. (1995). A small predicted stem-loop structure mediates oocyte localization of Drosophila K10 mRNA. Development 121: 3809-3818

Schulze, K. L., Littleton, J. T., Salzberg, A., Halachmi, N., Stern, M., Lev, Z. and Bellen, H. J. (1994). Rop, a Drosophila homolog of yeast Sec1 and vertebrate n-Sec1/Munc-18 proteins, is a negative regulator of neurotransmitter release in vivo. Neuron 13: 1099-1108

Semotok, J. L., et al. (2005). Smaug recruits the CCR4/POP2/NOT deadenylase complex to trigger maternal transcript localization in the early Drosophila embryo. Curr. Biol. 15(4): 284-94. Medline abstract: 15723788

Serbus, L. R., Cha, B. J., Theurkauf, W. E. and Saxton, W. M. (2005). Dynein and the actin cytoskeleton control kinesin-driven cytoplasmic streaming in Drosophila oocytes. Development 132(16): 3743-52. 16077093

Shapiro, R. S. and Anderson, K. V. (2006). Drosophila Ik2, a member of the I kappa B kinase family, is required for mRNA localization during oogenesis. Development 133(8): 1467-75. Medline abstract: 16540511

Shulman, J. M., Benton, R. and St Johnston, D. (2000). The Drosophila homolog of C. elegans PAR-1 organizes the oocyte cytoskeleton and directs oskar mRNA localization to the posterior pole. Cell 101: 377-388.

Siegel, V., et al. (1993). pipsqueak, an early acting member of the posterior group of genes, affects vasa level and germ cell-somatic cell interaction in the developing egg chamber. Development 119: 1187-202. PubMed Citation: 8306882

Simon, B., Masiewicz, P., Ephrussi, A. and Carlomagno, T. (2015). The structure of the SOLE element of oskar mRNA. RNA [Epub ahead of print]. PubMed ID: 26089324

Sinka, R., et al. (2002). poirot, a new regulatory gene of Drosophila oskar acts at the level of the short Oskar protein isoform. Development 129: 3469-3478. 12091316

Sinsimer, K. S., Jain, R. A., Chatterjee, S. and Gavis, E. R. (2011). A late phase of germ plasm accumulation during Drosophila oogenesis requires Lost and Rumpelstiltskin. Development 138(16): 3431-40. PubMed Citation: 21752933

Sinsimer, K. S., Lee, J. J., Thiberge, S. Y. and Gavis, E. R. (2013). Germ plasm anchoring is a dynamic state that requires persistent trafficking. Cell Rep 5(5): 1169-77. PubMed ID: 24290763

Smibert, C. A., et al. (1999). Smaug, a novel and conserved protein, contributes to repression of nanos mRNA translation in vitro. RNA 5(12): 1535-47. Medline abstract: 10606265

Smith, J. L., Wilson, J. E. and Macdonald, P. M. (1992). Overexpression of oskar directs ectopic activation of nanos and presumptive pole cell formation in Drosophila embryos. Cell 70: 849-59. PubMed Citation: 1516136

Snee, M. J. and Macdonald. P. M. (2004). Live imaging of nuage and polar granules: evidence against a precursor-product relationship and a novel role for Oskar in stabilization of polar granule components. J. Cell Sci. 117(Pt 10): 2109-20. 15090597

Snee, M. J. and Macdonald, P. M. (2009). Dynamic organization and plasticity of sponge bodies. Dev. Dyn. 238: 918-930. PubMed Citation: 1930139

Song, Y., et al. (2005). Evidence for an RNA chaperone function of polypyrimidine tract-binding protein in picornavirus translation. RNA 11: 1809-1824. PubMed Citation: 16314455

St Johnston, D., Beuchle, D. and Nusslein-Volhard, C. (1991). Staufen, a gene required to localize maternal RNAs in the Drosophila egg. Cell 66: 51-63. PubMed Citation: 1712672

Stark, A., Brennecke, J., Russell, R. B., and Cohen, R. S. (2003). Identification of Drosophila MicroRNA targets. PLoS Biology 1, e61. 14691536

Steckelberg, A. L., Altmueller, J., Dieterich, C. and Gehring, N. H. (2015). CWC22-dependent pre-mRNA splicing and eIF4A3 binding enables global deposition of exon junction complexes. Nucleic Acids Res 43(9): 4687-4700. PubMed ID: 25870412

Stoneley, M. and Willis, A. E. (2004). Cellular internal ribosome entry segments: Structures, trans-acting factors and regulation of gene expression. Oncogene 23: 3200-3207. PubMed Citation: 15094769

Suyama, R., et al. (2009). The actin-binding protein Lasp promotes Oskar accumulation at the posterior pole of the Drosophila embryo. Development 136(1): 95-105. PubMed Citation: 19036801

Suzanne, M., et al. (1999). The Drosophila p38 MAPK pathway is required during oogenesis for egg asymmetric development. Genes Dev. 13: 1464-1474

Tanaka, T. and Nakamura, A. (2008). The endocytic pathway acts downstream of Oskar in Drosophila germ plasm assembly. Development 135: 1107-1117. PubMed Citation: 18272590

Tanaka, T., et al. (2011). Drosophila Mon2 couples Oskar-induced endocytosis with actin remodeling for cortical anchorage of the germ plasm. Development 138(12): 2523-32. PubMed Citation: 21610029

Tanaka, T. and Nakamura, A. (2011b). Oskar-induced endocytic activation and actin remodeling for anchorage of the Drosophila germ plasm. Bioarchitecture 1(3): 122-126. PubMed ID: 21922042

Temme, C., Zaessinger, S., Meyer, S., Simonelig, M. and Wahle, E. (2004). A complex containing the CCR4 and CAF1 proteins is involved in mRNA deadenylation in Drosophila. EMBO J. 23(14): 2862-71. Medline abstract: 15215893

Tetzlaff, M. T., Jäckle, H. and Pankratz, M. J. (1996). Lack of Drosophila cytoskeletal tropomyosin affects head morphogenesis and the accumulation of oskar mRNA required for germ cell formation. EMBO J. 15: 1247-1254

Tirronen, M., et al. (1995). Two otu transcripts are selectively localised in Drosophila oogenesis by a mechanism that requires a function of the otu protein. Mech. Dev. 52: 65-75

Theurkauf, W. E., et al. (1993). A central role for microtubules in the differentiation of Drosophila oocytes. Development 118: 1169-80

Tomari, Y., Du, T., Haley, B., Schwarz, D. S., Bennett, R., Cook, H. A., Koppetsch, B. S., Theurkauf, W. E. and Zamore, P. D. (2004). RISC assembly defects in the Drosophila RNAi mutant armitage. Cell 116(6): 831-41. 15035985

Trucco, A., Gaspar, I. and Ephrussi, A. (2009). Assembly of endogenous oskar mRNA particles for motor-dependent transport in the Drosophila oocyte. Cell 139(5): 983-98. PubMed Citation: 19945381

van Eeden, F. J. M., et al. (2001). Barentsz is essential for the posterior localization of Oskar mRNA and colocalizes with it to the posterior pole. J. Cell Biol. 154: 511-524. 11481346

Vanzo, N. F. and Ephrussi, A. (2002). Oskar anchoring restricts pole plasm formation to the posterior of the Drosophila oocyte. Development 129: 3705-3714. 12117819

Vanzo, N., Oprins, A., Xanthakis, D., Ephrussi, A. and Rabouille, C. (2007). Stimulation of endocytosis and actin dynamics by Oskar polarizes the Drosophila oocyte. Dev. Cell 12: 543-555. PubMed Citation: 17419993

Veeranan-Karmegam, R., Boggupalli, D. P., Liu, G. and Gonsalvez, G. B. (2016). A new isoform of Drosophila non-muscle Tropomyosin 1 interacts with Kinesin-1 and functions in oskar mRNA localization. J Cell Sci 129: 4252-4264. PubMed ID: 27802167

Wang, L., Eckmann, C. R., Kadyk, L. C., Wickens, M. and Kimble, J. (2002). A regulatory cytoplasmic poly(A) polymerase in Caenorhabditis elegans. Nature 419: 312-316. 12239571

Webster, P. J., Suen, J. and Macdonald, P. M. (1994). Drosophila virilis oskar transgenes direct body patterning but not pole cell formation or maintenance of mRNA localization in D. melanogaster. Development 120: 2027-2037

Webster, P. J., et al. (1997). Translational repressor Bruno plays multiple roles in development and is widely conserved. Genes Dev. 11(19):2510-2521

Werz, C., et al. (2005). Mis-specified cells die by an active gene-directed process, and inhibition of this death results in cell fate transformation in Drosophila. Development 132(24): 5343-52. 16280349

Wilhelm, J. E., et al. (2000). Isolation of a ribonucleoprotein complex involved in mRNA localization in Drosophila oocytes. J. Cell Biol. 148: 427-440. PubMed Citation: 10662770

Wilhelm, J. E., Hilton, M., Amos, Q. and Henzel, W. J. (2003). Cup is an eIF4E binding protein required for both the translational repression of oskar and the recruitment of Barentsz. J. Cell Biol. 163: 1197-1204. 14691132

Williams, L. S., Ganguly, S., Loiseau, P., Ng, B. F. and Palacios, I. M. (2014). The auto-inhibitory domain and ATP-independent microtubule-binding region of Kinesin heavy chain are major functional domains for transport in the Drosophila germline. Development 141: 176-186. PubMed ID: 24257625

Wilson, J. E., Connell, J. E. and Macdonald, P. M. (1996). aubergine enhances oskar translation in the Drosophila ovary. Development 122: 1631-39. PubMed ID: 8625849

Wolfgang, W. J. and Forte, M. (1995). Posterior localization of the Drosophila Gi alpha protein during early embryogenesis requires a subset of the posterior group genes. Int. J. Dev. Biol. 39: 581-586

Xu, X., Brechbiel, J. L. and Gavis, E. R. (2013). Dynein-Dependent Transport of nanos RNA in Drosophila Sensory Neurons Requires Rumpelstiltskin and the Germ Plasm Organizer Oskar. J Neurosci 33: 14791-14800. PubMed ID: 24027279

Yang, N., Yu, Z., Hu, M., Wang, M., Lehmann, R. and Xu, R.M. (2015). Structure of Drosophila Oskar reveals a novel RNA binding protein. Proc Natl Acad Sci U S A [Epub ahead of print]. PubMed ID: 26324911

Yano, T., et al. (2004). Hrp48, a Drosophila hnRNPA/B homolog, binds and regulates translation of oskar mRNA. Genes Dev. 6: 637-648. 15130489

Yoshida, S., Muller, H. A., Wodarz, A. and Ephrussi, A. (2004). PKA-R1 spatially restricts Oskar expression for Drosophila embryonic patterning. Development 131(6): 1401-10. 14993189

Zaessinger, S., Busseau, I. and Simonelig, M. (2006). Oskar allows nanos mRNA translation in Drosophila embryos by preventing its deadenylation by Smaug/CCR4. Development 133(22): 4573-83. Medline abstract: 17050620

Zheng P., Eastman J., Vande Pol S. and Pimplikar S. W. (1998). PAT1, a microtubule-interacting protein, recognizes the basolateral sorting signal of amyloid precursor protein. Proc. Natl. Acad. Sci. 95: 14745-14750. PubMed Citation: 9843960

Zheng, Y., Wildonger, J., Ye, B., Zhang, Y., Kita, A., Younger, S. H., Zimmerman, S., Jan, L. Y. and Jan, Y. N. (2008). Dynein is required for polarized dendritic transport and uniform microtubule orientation in axons. Nat Cell Biol 10: 1172-1180. PubMed ID: 18758451

Zimyanin, V., Lowe, N. and St Johnston, D. (2007). An Oskar-dependent positive feedback loop maintains the polarity of the Drosophila oocyte. Curr. Biol. 17(4): 353-9. Medline abstract: 17275299

Zimyanin, V. L., Belaya, K., Pecreaux, J., Gilchrist, M. J., Clark, A., Davis, I. and St Johnston, D. (2008). In vivo imaging of oskar mRNA transport reveals the mechanism of posterior localization. Cell 134: 843-853. PubMed ID: 18775316


oskar: Biological Overview | Evolutionary Homologs | Regulation | Protein Interactions | Factors affecting Oskar translation | Factors affecting Oskar localization | Developmental Biology | Effects of Mutation

date revised: 26 August 2020

Home page: The Interactive Fly © 1997 Thomas B. Brody, Ph.D.

The Interactive Fly resides on the
Society for Developmental Biology's Web server.