Using the Capsuléen (Csul) methyltransferase as bait in the yeast two-hybrid system, we have identified a novel Drosophila protein containing multiple WD repeats and encoded by the valois(vsl) gene, which acts in pole plasm function. Vls is homologous to human MEP50, which forms a complex with the PRMT5 methyltransferase –the human homologue of Csul. We found that Vls localizes to the nuage in the nurse cells and to the pole plasm in the oocyte. Moreover vls is required for the synthesis and/or stability of Oskar and the localization of Tudor (Tud) in both the nuage and at the posterior pole of the oocyte. Furthermore, we show that Vls and a fragment of Tud interact directly in binding assay. As the PMRT5/MEP50 complex is involved in ribonucleoprotein complex assembly, we hypothesize that the Vls complex may play a similar function in assembling the nuage in nurse cells and the polar granules in the oocyte.

In Drosophila, the formation and fate of the germ cells depend on components distinctly located at the posterior pole of the egg – the pole plasm (Geigy, 1931). Functional and structural characteristics of the germ plasm confer specific properties to this region of the egg. First, transplantation of posterior plasm from oocytes or early cleavage embryos induces ectopic formation of pole cells within the recipient embryos(Illmensee and Mahowald, 1974; Okada et al., 1974). Second,the pole plasm is devoid of yolk particles, but exhibits electron-dense,ribonucleoprotein-rich organelles without delimiting membranes, called polar granules (Huettner, 1923; Mahowald, 1962; Illmensee et al., 1976). Third, the respiratory activity of the mitochondria located in the pole plasm is higher than that detected in other regions of preblastoderm embryos(Akiyama and Okada, 1992). Fourth, in addition to its ability to induce formation of ectopic pole cells,the pole plasm also contains an activity required for abdominal development,designated as the posterior determinant(Frohnhöfer et al., 1986; Lehmann and Nüsslein-Volhard,1986).

Although germline formation and abdominal development represent distinct developmental processes, molecular and genetic analyses of the posterior-grandchild-less group of genes indicate that both processes share a common pathway for producing functional determinants localized at the posterior pole of the embryo. So far four genes are known to be directly required in pole plasm function, including osk(Lehmann and Nüsslein-Volhard,1986), vasa (vas)(Schüpbach and Wieschaus,1986), tudor (tud)(Boswell and Mahowald, 1985)and valois (vls)(Schüpbach and Wieschaus,1986). Mutants in all four genes are characterized by the absence or strong reduction of discernible polar granules(Schüpbach and Wieschaus,1986; Boswell and Mahowald,1985).

Formation of pole plasm initially depends upon the localization of osk mRNA (Ephrussi and Lehmann,1992), which becomes concentrated at the posterior end of the oocyte in stage 8 egg chambers (Ephrussi et al., 1991; Kim-Ha et al.,1991). Shortly thereafter, osk mRNA is translated into two isoforms by use of different initiation codons(Markussen et al., 1995; Rongo et al., 1995), and the short form of Osk induces the assembly of polar granules by recruiting directly the Vas protein, a member of the DEAD-box family of putative RNA helicases (Lasko and Ashburner,1988; Hay et al.,1988b) at the posterior pole of the oocyte(Breitwieser et al., 1996). Vas may contribute to the synthesis of specific polar granule components, as suggested by its requirement for promoting translation of nanos(nos) (Gavis et al.,1996) and gurken(Styhler et al., 1998; Tomancak et al., 1998). The other factor involved in pole plasm function is Tud(Golumbeski et al., 1991),which accumulates in polar granules but is also detected in the nuage of nurse cells and at the periphery of embryonic nuclei(Bardsley et al., 1993). Although the molecular nature of vls remained elusive, this gene appears to be involved in pole plasm function(Schüpbach and Wieschaus,1986).

Formation of abdomen and pole cells can be mechanistically uncoupled. The nos gene contributes to abdominal patterning, albeit not to pole cell formation (Lehmann and Nüsslein-Volhard, 1991). The localization of nosmRNA to the posterior pole of the embryo(Wang and Lehmann, 1991)depends upon genes acting in pole plasm and produces a transient gradient of Nos proteins emanating from the posterior pole(Wang et al., 1994). Unlike the posterior determinant, the determinant involved in pole cell formation is likely to consist of multiple elements. The gene germ cell-less(gcl) is one of them (Jongens et al., 1992), and ectopic expression of gcl mRNA at the anterior pole of the embryo produces nuclei with characteristics of pole cell nuclei with which Gcl protein becomes associated on the inner rim of the nuclear pores (Jongens et al.,1994). Furthermore, mitochondrial ribosomal RNA (mtlrRNA) also contributes to germ-line determination(Iida and Kobayashi, 1998). Transport of the mtlrRNA from mitochondria to the surface of polar granules is mediated by Tud protein through a yet unknown mechanism(Amikura et al., 2001).

During our analysis of the capsuléen gene (csul; CG3730), we found one major protein partner by using the yeast two-hybrid system. In this report, we show that the protein interacting with Csul is encoded by the vls gene and describe its function in the process of pole cell formation.

Molecular biology

Plasmid constructs were generated by amplifying selected fragments by PCR(High Fidelity PCR Master, Roche) and cloned into appropriate vectors. vas and tud cDNA plasmids were kindly provided by P. Lasko and R. Boswell, respectively. vls cDNA (clone GM04727, BDGP-EST project) was purchased from Invitrogen and gus cDNA (clone LD34464,BDGP-EST project) was obtained from the Drosophila Genomics Resource Center (DGRC; Bloomington, IN). The plasmid pSM 492 (3×HA) was kindly provided by S. Michaelis.

To rescue the vls phenotype a 2.04 kb genomic DNA fragment containing the vls transcription unit and 535 and 195 nucleotides of the 5′ and 3′ genomic regions, respectively, was cloned into the P-element vector CaSpeR4. The construct was named P-[vls]. Details will be provided upon request. To generate a Vls-HA fusion protein construct, a BglII site was created at the 3′end of the vls-coding sequence of a vls cDNA containing identical 5′ and 3′ untranslated regions as the genomic transgene. A fragment with 3×HA tags was cut with BglII from plasmid pSM 492 and cloned into the vsl-coding sequence. Details will be provided upon request. A clone containing six copies of HA was selected for microinjection.

Sequencing of mutant alleles

The vls-coding region was amplified by PCR (High Fidelity PCR Master; Roche) from genomic DNA obtained from single vls1,vls2 and vls3 flies (primers: sense 5′-gttgctcttccttgctggccgattctc-3′, located at position –58 from the translation initiation codon; and antisense,5′-tgcatttaaactgggctgctgcttcac-3′). The PCR products of 1.46 kb were cloned into pBluescript. Two clones from independent PCR reactions were sequenced for each allele using primers covering the coding region. Nucleotide differences from the wild-type vls sequence were distinguished from PCR errors by their appearance in both independent clones.

Yeast two-hybrid screen

Standard yeast two-hybrid methods were employed, using the GAL4 system(Saccharomyces cerevisiae strain Y190). The Csul-coding region was cloned into the yeast vector pGBT9 to produce a bait construct. This construct was used to screen the Drosophila embryo MATCHMAKER cDNA library(Clontech). About 1.2 millions clones were screened.

GST pull-down assay

Full-length vls and vas cDNAs were subcloned into pGEX6P2(Pharmacia), full-length csul, gus and tud cDNAs into pCITE-4 (Novagen). Recombinant proteins were synthesized in vitro using the TNT T7 Coupled Reticulocyte Lysate System (Promega) in the presence of unlabeled amino acids. GST-fusion proteins expressed in E. coli were purified with glutathione sepharose (Pharmacia) and washed with the binding buffer [20 mM HEPES (pH 7.8), 10% glycerol, 300 mM NaCl, 0.1% sodium deoxycholate, 0.1% NP40 and 0.1% Triton X-100] plus protease inhibitors(Complete EDTA free from Roche; 1:50 dilution). Recombinant proteins (10% of the reaction volume) were added to this mixture (in 1 ml) and incubated for 3 hours at room temperature. The beads were washed six times (10 minutes each)with binding buffer, boiled in loading buffer, and the proteins were separated by electrophoresis on 8% or 12% SDS-polyacrylamide gels. After transfer to a polyvinylidene difluoride (PVDF) membrane, the bound proteins were detected by Western blotting using an S-protein Alkaline Phosphatase conjugate(Novagen).

Immunoprecipitation

Fly ovaries were homogenized using a plastic pestle in ice-cold IP buffer[145 mM NaCl, 10% glycerol, 1 mM MgCl2, 1.5 mM NaH2PO4, 8 mM Na2HPO4 (pH 7.4) and 0.5% NP40] containing protease inhibitors. Lysates were clarified by centrifugation at 14,000 g for 10 minutes at 4°C. The supernatants were removed and mixed with 400 units of RNasin (Promega), 30μl of anti-HA antibody-coupled protein A beads for 5 hours at 4°C on a rotator. The IP mix was centrifuged at low speed to remove the supernatant. Protein A-beads were then washed four times with IP buffer for 10 minutes each and TAP-Csul was detected on a western blot by using alkaline phosphatase-conjugated IgG (Sigma). TAP-Csul purification was essentially performed as described by Rigaut et al.(Rigaut et al., 1999).

Immunocytochemistry

Immunostaining was performed with primary antibodies, including anti-Mael from rabbit (Findley et al.,2003), anti-Osk from rabbit (gift of A. Ephrussi), anti-Vas from rabbit and rat (gift of P. Lasko and A. Nakimura, respectively), anti-Tud from rabbit (TUD56; gift of S. Kobayashi), and monoclonal anti-HA (clone 16B12;BAbCO). To reduce background, we incubated first the anti-HA antibodies(1:400) for 3 hours at room temperature with wild-type ovaries and then overnight with experimental ovaries. For immunostaining, the ovaries were fixed for 10 minutes in 4% paraformaldehyde. To visualize components of the nuage, we used the protocol of Findley et al.(Findley et al., 2003). Chromatin was visualized by staining with Oli-Green (Molecular Probes). Cy3-,Cy5- and FITC-conjugated secondary antibodies (Jackson ImmunoResearch Laboratories) were used at 1:200.

Isolation of a major Capsuléen partner

The capsuléen gene (csul; CG3730) is required for pole cell formation in the pathway controlled by osk(J.A., R. Ollo, A. Ephrussi and B.M.M., unpublished). To identify Csul-interacting proteins active in pole plasm, we performed a yeast two-hybrid screen. The bait construct consisted of the sequence of the Csul protein (610 amino acid residues) fused to the DNA binding domain (BD) of Gal4. By screening a Drosophila cDNA library, we obtained six independent interacting clones of which four of them corresponded to the CG10728 coding sequence predicted by the Drosophila genome project. Conceptual translation of CG10728 cDNA revealed a protein of 367 amino acids. Two cDNAs encoded a nearly full-size protein lacking the first 10 amino acid residues, whereas the other two isolates encoded a protein in which 37 residues were missing from the N terminus. Confirmation of the interaction was obtained by performing the reciprocal two-hybrid assay,showing that BD-CG10728 could strongly bind the activation domain of Gal4 fused to Csul (data not shown). Sequence analysis on the Protein Sequence Analysis Server(bmerc-www.bu.edu/psa/)indicated that CG10728 contains four WD repeats(Fig. 1A).

Fig. 1.

Molecular analysis of vls. (A) Location of three EMS mutations in vls. All mutations lead to a premature stop codon. WD repeats are shown in grey. (B) The vls transcript and restriction map of the genomic DNA used for generating a vls rescue transgene. Exons are drawn as boxes and the putative open reading frame is indicated in black. B, BamHI; Bs, BstEII; P, PstI; R, EcoRI. (C)Expression of the P[HA-Vls] transgene in ovaries revealed by western blotting using anti-HA antibodies. A band of ∼50 kDa is detected in transgenic but not in wild-type ovaries. (D) Alignment of the amino acid sequences of Vls and homologous proteins: CP8824, an incomplete Anopheles gambiae homolog; human MEP50; and MGC65780, a zebrafish protein. These proteins display strong conservation of the WD repeats (grey boxes). MEP50 and the zebrafish homolog display six potential WD repeats, whereas Drosophila Vls contains only four such repeats, as predicted by the Protein Sequence Analysis server of the BioMolecular Engineering Research Center of Boston University(http://bmerc-www.bu.edu/psa/). Multiple sequence alignment of Vls and related proteins was performed using the Pileup program of the Wisconsin Package (Genetics Computer Group). Gaps in the amino acid sequence, indicated by dots, were introduced for optimal alignment. The GenBank accession numbers of these sequences are the following:CP8824_AG, EAA14269; MEP50_HS, AAL79917; MGC65780_DR, AAH56278.

Fig. 1.

Molecular analysis of vls. (A) Location of three EMS mutations in vls. All mutations lead to a premature stop codon. WD repeats are shown in grey. (B) The vls transcript and restriction map of the genomic DNA used for generating a vls rescue transgene. Exons are drawn as boxes and the putative open reading frame is indicated in black. B, BamHI; Bs, BstEII; P, PstI; R, EcoRI. (C)Expression of the P[HA-Vls] transgene in ovaries revealed by western blotting using anti-HA antibodies. A band of ∼50 kDa is detected in transgenic but not in wild-type ovaries. (D) Alignment of the amino acid sequences of Vls and homologous proteins: CP8824, an incomplete Anopheles gambiae homolog; human MEP50; and MGC65780, a zebrafish protein. These proteins display strong conservation of the WD repeats (grey boxes). MEP50 and the zebrafish homolog display six potential WD repeats, whereas Drosophila Vls contains only four such repeats, as predicted by the Protein Sequence Analysis server of the BioMolecular Engineering Research Center of Boston University(http://bmerc-www.bu.edu/psa/). Multiple sequence alignment of Vls and related proteins was performed using the Pileup program of the Wisconsin Package (Genetics Computer Group). Gaps in the amino acid sequence, indicated by dots, were introduced for optimal alignment. The GenBank accession numbers of these sequences are the following:CP8824_AG, EAA14269; MEP50_HS, AAL79917; MGC65780_DR, AAH56278.

CG10728 codes for valois

The CG10728 gene is located at chromosomal band 38B2. As vls was genetically assigned to this region(Schüpbach and Wieshaus,1986), we investigated whether the CG10728 sequence was altered in vls mutants. To determine the lesions in vls1, vls2 and vls3, we amplified by PCR genomic DNA fragments encompassing the CG10728-coding sequence, then cloned and sequenced the amplified fragments from each mutant. For each vls mutant we found single nucleotide substitutions that result in premature termination of the coding sequence (Fig. 1A). In vls1 and vls2 the nucleotide substitutions changed a TGG Trp codon into TAG and TGA stop codons, producing truncated proteins of 227 and 52 amino acids, respectively. In vls3, the nucleotide substitution transformed a CGA Arg codon into a TGA stop codon producing a truncated protein of 69 amino acids. To confirm that we cloned vls, we constructed two transgenes containing the presumed vls sequence, including its 5′regulatory sequences. The first transgene contained a genomic DNA fragment comprising the CG10728 transcription unit(Fig. 1B). The second transgene harboured the same sequence to which we fused six copies of the hemaglutinin tag (HA), in frame with the vls-coding sequence. Western blots of proteins extracted from ovaries producing HA-Vls probed with anti-HA antibodies displayed a single polypeptide with the expected mass of 50 kDa(Fig. 1C). Both transgenes restored full viability and fertility to embryos produced by vls1 and vls3 homozygous females (data not shown). Taken together, our results demonstrate that CG10728corresponds to vls.

Csul and Vls interact in vivo

Among the nearest homologues of Vls we identified in protein databases the methylosome protein 50 (MEP50) (Fig. 1D). MEP50, the human homologue of Vls, contains six WD repeats and interacts with PRMT5 (Friesen et al.,2002), the human homologue of Csul. The finding of an interaction between Csul and Vls in yeast prompted us to analyze whether such binding occurs in Drosophila ovaries. To this end, we constructed transgenic flies bearing both the HA-vls transgene and a Tandem Affinity Purification (TAP) (Rigaut et al.,1999) tagged csul transgene (TAP-csul) shown to restore csul development (data not shown). Protein extracts of P{TAP-csul}; vls3; P{HA-vls} ovaries were immunoprecipitated using rabbit IgG-sepharose beads and the bound Csul-complexes were released by treatment with recombinant TEV protease. Blots of released proteins probed with anti-HA antibodies showed an immunoreactive band of 50 kDa exhibiting the predicted mass of HA-Vls fusion protein(Fig. 2A). Reciprocally, blots of ovarian proteins immunoprecipitated with anti-HA antibodies and probed with IgG antibodies displayed an immunoreactive band of the mass predicted for TAP-Csul (90 kDa; Fig. 2B). These results indicate that Csul and Vls can associate in a protein complex in Drosophila ovaries.

Fig. 2.

Vls interacts physically with Csul. (A) Vls interacts with Csul in vivo. Ovarian protein extracts of homozygous vls3 females producing both TAP-Csul and HA-Vls (lanes 1 and 2), or only HA-Vls (lane 3),were directly loaded on a gel (one tenth of total input, lane1) or immunoprecipitated with Sepharose IgG (lanes 2 and 3). Following separation by SDS-PAGE electrophoresis, bound-HA-Vls was detected by immunoblotting using anti-HA antibodies. (B) Reciprocally, Vls was immunoprecipitated using anti-HA antibodies and bound Csul was detected by immunoblotting using alkaline phosphatase-conjugated anti-rabbit antibodies. (C) Binding of the C terminus of Vls to Csul. Full-length GST-Vls or derivatives were purified from bacterial extracts and incubated with S•Tag-Csul. Bound S•Tag-Csul(75 kDa) was separated by SDS-PAGE electrophoresis and detected by immunoblotting using alkaline phosphatase-conjugated S proteins. Input:one-tenth of protein extract was loaded on the gel. The amount of used GST-fusion proteins was evaluated by SDS-PAGE followed by Coomassie staining(lower panel). Representation of the GST-Vls constructs used for the mapping and summary of results are indicated on the right. The WD-repeats and the putative Csul-binding domain of Vls are indicated by grey and green boxes,respectively. (D) The C-terminal region of Csul interacts with Vls. Full-size S•Tag-Csul or derivatives were synthesized in vitro and incubated with full-size GST-Vls. Following separation by SDS-PAGE electrophoresis the bound S•Tag-Csul proteins were detected by immunoblotting using alkaline phosphatase-conjugated S proteins. Left panel: input S•Tag-Csul proteins. The arrow indicates an endogenous protein synthesized in the reticulocyte system and reactive to alkaline phosphatase-conjugated S proteins. White squares indicate the position of the different S•Tag-Csul proteins. Middle panel: S•Tag-Csul proteins bound to GST-Vls are indicated by white squares. Representation of S•Tag-Csul constructs used for the mapping and summary of the results are indicated in the right panel. The putative Vls-binding domain of Csul is depicted in blue.

Fig. 2.

Vls interacts physically with Csul. (A) Vls interacts with Csul in vivo. Ovarian protein extracts of homozygous vls3 females producing both TAP-Csul and HA-Vls (lanes 1 and 2), or only HA-Vls (lane 3),were directly loaded on a gel (one tenth of total input, lane1) or immunoprecipitated with Sepharose IgG (lanes 2 and 3). Following separation by SDS-PAGE electrophoresis, bound-HA-Vls was detected by immunoblotting using anti-HA antibodies. (B) Reciprocally, Vls was immunoprecipitated using anti-HA antibodies and bound Csul was detected by immunoblotting using alkaline phosphatase-conjugated anti-rabbit antibodies. (C) Binding of the C terminus of Vls to Csul. Full-length GST-Vls or derivatives were purified from bacterial extracts and incubated with S•Tag-Csul. Bound S•Tag-Csul(75 kDa) was separated by SDS-PAGE electrophoresis and detected by immunoblotting using alkaline phosphatase-conjugated S proteins. Input:one-tenth of protein extract was loaded on the gel. The amount of used GST-fusion proteins was evaluated by SDS-PAGE followed by Coomassie staining(lower panel). Representation of the GST-Vls constructs used for the mapping and summary of results are indicated on the right. The WD-repeats and the putative Csul-binding domain of Vls are indicated by grey and green boxes,respectively. (D) The C-terminal region of Csul interacts with Vls. Full-size S•Tag-Csul or derivatives were synthesized in vitro and incubated with full-size GST-Vls. Following separation by SDS-PAGE electrophoresis the bound S•Tag-Csul proteins were detected by immunoblotting using alkaline phosphatase-conjugated S proteins. Left panel: input S•Tag-Csul proteins. The arrow indicates an endogenous protein synthesized in the reticulocyte system and reactive to alkaline phosphatase-conjugated S proteins. White squares indicate the position of the different S•Tag-Csul proteins. Middle panel: S•Tag-Csul proteins bound to GST-Vls are indicated by white squares. Representation of S•Tag-Csul constructs used for the mapping and summary of the results are indicated in the right panel. The putative Vls-binding domain of Csul is depicted in blue.

Interaction domains in Vls and Csul

To further characterize the interaction between Csul and Vls, we mapped the binding domains in each protein by using a GST-pull down assay. For this purpose we fused Vls or fragments of Vls to GST, and Csul or fragments of Csul to S•Tag (Kim and Raines,1993). In vitro translated S•Tag-Csul polypeptides were incubated with immobilized GST-Vls proteins and, after washing, the bound S•Tag-Csul proteins were detected by using S-protein coupled to alkaline phosphatase. As shown in Fig. 2C, Csul bound specifically to immobilized full-length Vls, or to the 30 amino acid C terminal region of Vls, but not to the rest of the protein. GST alone displayed no binding to S•Tag-Csul (data not shown). Conversely, we mapped the region of Csul mediating the interaction with Vls to the C terminal region of Csul (Fig. 2D). Neither the N-terminal region nor the central region of Csul showed strong binding to Vls. Thus, our data indicate that Csul and Vls interact through their C-terminal regions.

Vls is present in nuage and pole plasm

To determine the distribution of Vls during oogenesis, we analyzed the localization of HA-Vls protein using anti-HA antibodies. Immunostaining of vls3; P{HA-vls} egg chambers revealed that Vls accumulates at the posterior pole of the oocyte in stage 10 egg chambers(Fig. 3A,C) and in the pole plasm of early syncytial embryos (Fig. 3E,F). We also found Vls in pole cells(Fig. 3H,I) and in migrating primordial germ cells (data not shown). During oogenesis, Vls decorates a region surrounding the nurse cell nuclei(Fig. 3K), which bears strong similarity to the nuage (Findley et al.,2003). By contrast, immunostaining of wild type ovaries using anti-HA antibodies displayed only an overall residual staining(Fig. 3B,D,G,J,L). These data show that Vls is a component of both the nuage and the pole plasm.

Fig. 3.

Vls is a component of nuage and pole plasm. Fixed, whole-mount egg chambers stained with Oli-Green for DNA and anti-HA antibodies in homozygous vls3 flies expressing P{HA-Vls} (A,C,E,F,H,I,K)and in wild-type flies as controls (B,D,G,J,L). HA-Vls is detected at the posterior pole of the oocyte in transgenic vls3; P{HA-Vls}stage 10 egg chambers (A,C). (C,D) Enlarged views of similar stage 10 oocytes,as shown in A and B, respectively. HA-Vls is observed at the posterior pole of early vls3; P{HA-Vls} embryos and in pole cells (E,F,H,I).(F,G,I,J) Enlarged views of similar syncytial and blastoderm embryos, as shown in E and H, respectively. HA-Vls is detected in the nuage surrounding the nurse cell nuclei (K).

Fig. 3.

Vls is a component of nuage and pole plasm. Fixed, whole-mount egg chambers stained with Oli-Green for DNA and anti-HA antibodies in homozygous vls3 flies expressing P{HA-Vls} (A,C,E,F,H,I,K)and in wild-type flies as controls (B,D,G,J,L). HA-Vls is detected at the posterior pole of the oocyte in transgenic vls3; P{HA-Vls}stage 10 egg chambers (A,C). (C,D) Enlarged views of similar stage 10 oocytes,as shown in A and B, respectively. HA-Vls is observed at the posterior pole of early vls3; P{HA-Vls} embryos and in pole cells (E,F,H,I).(F,G,I,J) Enlarged views of similar syncytial and blastoderm embryos, as shown in E and H, respectively. HA-Vls is detected in the nuage surrounding the nurse cell nuclei (K).

Distribution of pole plasm and nuage components in vls egg chambers

Embryos produced by vls females fail to form pole cells at the posterior pole (Schüpbach and Wieschaus, 1986) (Fig. 4B). To check whether pole cell formation strictly depends on vls activity, we introduced an osk-bcd 3′UTR transgene into vls1 and vls3 backgrounds, and found no evidence for ectopic formation of pole cells at the anterior pole(Fig. 4D), although we could detect residual Vas staining in this region of the egg (data not shown). This result concurs with the suppressor effect of vls1 on ectopic pole cell formation observed after overexpression of osk(Smith et al., 1992), but contrasts the data using an osk-bcd 3′UTR transgene expressed in vls2/Df(2R)TW2 flies by Ephrussi and Lehmann(Ephrussi and Lehmann, 1992). From our data, we conclude that vls function is essential for pole cell formation.

Fig. 4.

Defective pole plasm assembly in vls. (A,B) Absence of pole cell formation in vls3. Staining of Vas with rat anti-Vas antibodies (red) and DNA with Oli-Green (green) in wild-type (A) and vls3 (B) blastoderm embryos. (C,D) vls suppresses ectopic anterior pole cell formation induced by osk-bcd 3UTR. Staining of Vas protein (red) and DNA (green) in osk-bcd 3UTR (C) and vls3; osk-bcd 3UTR (D) blastoderm embryos. No pole cells form at the anterior pole of vls3; osk-bcd 3UTR embryos. (E,F) Osk is localized at the posterior pole of the oocyte in vls3 stage 10 egg chambers, albeit in reduced amount. (E) Wild-type and (F) vls3 egg chambers. (G) The level of Osk protein is reduced in vls. Ovarian protein extracts of wild-type (WT), vls1,aubN11/aubHN2 and osk84/Df(3R)pXT103 females were separated by SDS-PAGE electrophoresis, blotted and detected using anti-Osk antibodies. The abundance of short Osk is reduced in vls mutant ovaries, whereas the long form is maintained at the wild-type level. Both Osk forms are strongly reduced in aub ovaries, with a more substantial effect on short Osk. In contrast to aub, the phosphorylation of short Osk is not affected in vls. The same blot was successively probed for ribosomal P40, as loading control (Török et al.,1999). (H,I) Vas is localized at the posterior pole of the oocyte in wild type (H) and vls3 (I) stage 10 egg chambers. There is less localized Vas in vls3 compared with wild type. Ovaries were stained with rabbit anti-Vas antibodies (red) and Oli-Green for DNA (green). (J,K) Tud is detected in nuage and at the oocyte posterior pole in wild-type stage 10 egg chambers (J) but is not detected at these locations in vls3 stage 10 egg chambers (K). Ovaries were stained with anti-Tud antibodies (red) and Oli-Green for DNA (green).

Fig. 4.

Defective pole plasm assembly in vls. (A,B) Absence of pole cell formation in vls3. Staining of Vas with rat anti-Vas antibodies (red) and DNA with Oli-Green (green) in wild-type (A) and vls3 (B) blastoderm embryos. (C,D) vls suppresses ectopic anterior pole cell formation induced by osk-bcd 3UTR. Staining of Vas protein (red) and DNA (green) in osk-bcd 3UTR (C) and vls3; osk-bcd 3UTR (D) blastoderm embryos. No pole cells form at the anterior pole of vls3; osk-bcd 3UTR embryos. (E,F) Osk is localized at the posterior pole of the oocyte in vls3 stage 10 egg chambers, albeit in reduced amount. (E) Wild-type and (F) vls3 egg chambers. (G) The level of Osk protein is reduced in vls. Ovarian protein extracts of wild-type (WT), vls1,aubN11/aubHN2 and osk84/Df(3R)pXT103 females were separated by SDS-PAGE electrophoresis, blotted and detected using anti-Osk antibodies. The abundance of short Osk is reduced in vls mutant ovaries, whereas the long form is maintained at the wild-type level. Both Osk forms are strongly reduced in aub ovaries, with a more substantial effect on short Osk. In contrast to aub, the phosphorylation of short Osk is not affected in vls. The same blot was successively probed for ribosomal P40, as loading control (Török et al.,1999). (H,I) Vas is localized at the posterior pole of the oocyte in wild type (H) and vls3 (I) stage 10 egg chambers. There is less localized Vas in vls3 compared with wild type. Ovaries were stained with rabbit anti-Vas antibodies (red) and Oli-Green for DNA (green). (J,K) Tud is detected in nuage and at the oocyte posterior pole in wild-type stage 10 egg chambers (J) but is not detected at these locations in vls3 stage 10 egg chambers (K). Ovaries were stained with anti-Tud antibodies (red) and Oli-Green for DNA (green).

To determine vls hierarchical position among known components of the pole plasm, we investigated the distribution of Osk, Vas and Tud in vls1 and vls3 egg chambers. Use of specific antibodies against Osk revealed that Osk was normally localized at the posterior pole from stage 9 onwards egg chambers in vls(Fig. 4F). However, we often noticed that the amount of Osk protein was reduced in vls in comparison with wild type. To determine whether vls regulates Osk protein accumulation, we examined by immunoblotting the amount of Osk protein produced in wild-type and vls ovaries. We found that the level of Osk short form was markedly reduced in vls in comparison with that in wild type (Fig. 4G)(Rongo et al., 1995) and even more reduced in aubergine (aub)(Wilson et al., 1996). However, the phosphorylation of this isoform seems unaffected in vls. In addition, we noticed that the slower migrating form of the short Osk protein was particularly reduced in abundance in aub protein extract,suggesting a defective phosphorylation in aub ovaries. This result extends the previous observation of Wilson et al.(Wilson et al., 1996), showing a strongly reduced level of short Osk in aub ovaries and may explain the aub requirement for osk activity induced anteriorly by the OB1 osk-bcd 3′UTR transgene. Examination of Vas distribution showed a normal localization at the posterior pole in stage 10 vls egg chambers (Fig. 4I), albeit at a reduced level compared with wild-type egg chambers (Fig. 4H).

Checking the distribution of Tud during oogenesis revealed that Tud was absent from the posterior pole of vls mutant oocytes in stage 10 egg chambers (Fig. 4K), indicating that Vls may play a role in pole plasm accumulation of Tud. As Tud is also a component of the nuage encircling nurse cell nuclei(Bardsley et al., 1993), we looked at this structure by examining the distribution of three components of the nuage including Tud, Vas (Hay et al.,1988a) and Maelstrom (Mael)(Findley et al., 2003). The use of a transgene expressing a GFP-Vas fusion protein(Sano et al., 2002) associated with the vls chromosome revealed that GFP-Vas was normally localized in the nuage of vls nurse cells, albeit at a lesser degree compared with wild type (Fig. 5B)(Hay et al., 1990; Lasko and Ashburner, 1990). Examination of Mael showed that this protein is also present in the nuage,although its distribution diverges from the wild-type pattern. We found that Mael is concentrated in brighter spots in vls than in wild type at the periphery of the nurse cell nuclei(Fig. 5D). By contrast, we were unable to detect Tud in the nuage of vls nurse cells(Fig. 5F), although Tud accumulates in vls oocytes (Fig. 5F) and transiently localizes at their anterior margin (data not shown). These findings indicate that: (1) vls acts downstream of Vas for the recruitment of Tud in the nuage; and (2) events occurring in the nuage are dispensable for the transport of Tud from the nurse cells to the oocyte but are required for Tud recruitment in the pole plasm.

Fig. 5.

Displacement of nuage components during early vls oogenesis. (A,B)GFP-Vas and (C,D) Mael accumulate in nuage in stage 7 egg chambers in wild type (A,C) and vls3 (B,D). GFP-Vas (green) was detected in fixed egg chambers stained with Hoechst 33258 for DNA (red). (E,F) Tud accumulates in nuage in wild type stage 7 nurse cells (E), whereas it is homogenously dispersed in the nurse cell cytoplasm in vls3stage 7 egg chambers (F). Ovaries were stained using anti-Tud antibodies (red)and Oli-Green for DNA (green). Tud is found at high levels in wild type and vls3 oocytes. (G,H) Staining of DNA with Oli-Green reveals that the karyosome, which appears as a single compact dot in wild-type oocytes(G), is frequently fragmented in vls3 oocytes (H). The oocyte nuclear membrane is indicated by a broken line.

Fig. 5.

Displacement of nuage components during early vls oogenesis. (A,B)GFP-Vas and (C,D) Mael accumulate in nuage in stage 7 egg chambers in wild type (A,C) and vls3 (B,D). GFP-Vas (green) was detected in fixed egg chambers stained with Hoechst 33258 for DNA (red). (E,F) Tud accumulates in nuage in wild type stage 7 nurse cells (E), whereas it is homogenously dispersed in the nurse cell cytoplasm in vls3stage 7 egg chambers (F). Ovaries were stained using anti-Tud antibodies (red)and Oli-Green for DNA (green). Tud is found at high levels in wild type and vls3 oocytes. (G,H) Staining of DNA with Oli-Green reveals that the karyosome, which appears as a single compact dot in wild-type oocytes(G), is frequently fragmented in vls3 oocytes (H). The oocyte nuclear membrane is indicated by a broken line.

Examination of vls egg chambers revealed also a modification in the structure of the oocyte nucleus in which the karyosome is fragmented into two fuzzy spots (Fig. 5H)instead of forming a single compact dot in wild-type nuclei(Fig. 5G). A similar phenotype was described for spindle(González-Reyes et al.,1997) and vas mutants(Tomancak et al., 1998). As vls oocytes undergo normal meiosis and support normal embryonic development, this finding suggests that Vls may exert a dispensable function on the karyosome organization.

Vls physically interacts with Tud in vitro

The presence in Vls of four WD repeats able to interact with other proteins(reviewed by Smith et al.,1999) prompted us to examine whether Vls could physically interact with proteins localized in the nuage. For this purpose, we performed pull-down assays using tagged Tud, Vas and Gustavus (Gus)(Styhler et al., 2002)polypeptides.

Five segments of Tud fused to the S•Tag, including JOZ (amino acid residues 3-273), 9A1 (residues 198-1199), 3ZS+L-N (residues 1198-1981) and 3ZS+L-C (residues 1941-2515) (Golumbeski et al., 1991), together comprising the complete Tud protein(Fig. 6A), were in vitro translated and the synthesized polypeptides were incubated with immobilized full-length and truncated GST-Vls proteins. Of the five Tud fragments, we found that only the 9A1 fragment could interact with Vls, whereas JOZ and the two subfragments of 3ZS+L showed only weak binding(Fig. 6C). The 9A1 fragment was further divided into two fragments. The 9A1-N and 9A1-C fragments (residues 198-770 and 751-1199, respectively) displayed a strong binding to Vls(Fig. 6C).

Fig. 6.

Vls physically interacts with Tud. (A) The eight Tudor (1-8) and two Tudor-like domains (1′-2′) are depicted as purple boxes. Fragments of Tud (Golumbeski et al.,1991) used in pull-down assays are indicated below the map. N-and C-terminal amino acid residues of each fragment are indicated. (B) Western blot analysis of in vitro produced S•Tag-Tud fragments detected by using alkaline phosphatase-conjugated S proteins. (C) After incubation with GST or GST-Vsl, the bound Tud fragments were detected by immunoblotting using alkaline phosphatase-conjugated S proteins: the Tud-9A1-N and Tud-9A1-C fragments display a strong binding to GST-Vls, whereas Tud-JOZ and Tud-3ZS+L bind only weakly to it. GST alone (negative control) exhibits no binding to Tud-JOZ and Tud-9A1. Lower panels: input GST-Vls and GST proteins separated on a SDS-PAGE gel and stained with Coomassie.

Fig. 6.

Vls physically interacts with Tud. (A) The eight Tudor (1-8) and two Tudor-like domains (1′-2′) are depicted as purple boxes. Fragments of Tud (Golumbeski et al.,1991) used in pull-down assays are indicated below the map. N-and C-terminal amino acid residues of each fragment are indicated. (B) Western blot analysis of in vitro produced S•Tag-Tud fragments detected by using alkaline phosphatase-conjugated S proteins. (C) After incubation with GST or GST-Vsl, the bound Tud fragments were detected by immunoblotting using alkaline phosphatase-conjugated S proteins: the Tud-9A1-N and Tud-9A1-C fragments display a strong binding to GST-Vls, whereas Tud-JOZ and Tud-3ZS+L bind only weakly to it. GST alone (negative control) exhibits no binding to Tud-JOZ and Tud-9A1. Lower panels: input GST-Vls and GST proteins separated on a SDS-PAGE gel and stained with Coomassie.

In the reciprocal experiment, we found that the Tud 9A1-N fragment can bind to the region encompassing residues 90-190 of Vls. When this segment was further divided into two A and B fragments, containing residues 90-139 and 139-192, respectively, we discovered that only the fragment A was able to bind to Tud 9A1-N. These results indicate that the first WD repeat of Vls can directly interact with Tud (Fig. 7).

Fig. 7.

Tud-binding domain in Vls. As performed in Fig. 2C, full-length GST-Vls or derivatives were incubated with S•Tag-Tud9A1-N. Bound S•Tag-Tud9A1-N was detected as described in Fig. 6B. Input: one-tenth of the S•Tag-Tud9A1-N extract. The smallest fragment of Vls showing binding to Tud9A1-N encompasses the first WD domain shown in green. The amount of GST-Vls proteins was visualized by Coomassie staining (lower panel). Representation of the GST-Vls fragments used for mapping and summary of the results are indicated on the right panel.

Fig. 7.

Tud-binding domain in Vls. As performed in Fig. 2C, full-length GST-Vls or derivatives were incubated with S•Tag-Tud9A1-N. Bound S•Tag-Tud9A1-N was detected as described in Fig. 6B. Input: one-tenth of the S•Tag-Tud9A1-N extract. The smallest fragment of Vls showing binding to Tud9A1-N encompasses the first WD domain shown in green. The amount of GST-Vls proteins was visualized by Coomassie staining (lower panel). Representation of the GST-Vls fragments used for mapping and summary of the results are indicated on the right panel.

No specific Vls binding with Vas and Gus polypeptides was detected in GST pull-down assays using GST, GST-Vls and GST-Vls (residues 90-190). Similarly,no interaction could be uncovered between Vls and Vas in the yeast two-hybrid assay (data not shown). These data indicate that the binding between Vls and Tud represents a specific interaction.

By analyzing potential interactors of the putative Csul methyltransferase,we isolated the vls gene. The confirmation that we cloned vls was obtained by determining the molecular lesions in three EMS-induced vls mutations and by showing that development and fertility of vls mutant flies is restored by transgenes containing the vls genomic sequence.

Attempts to isolate additional partners of Vls by using the yeast-two hybrid system were unsuccessful because of the occurrence of a too large number of clones growing on selective medium (data not shown), presumably reflecting the occurrence of WD repeats that are known to mediate interactions with numerous proteins in eukaryotic cells(Smith et al., 1999). By using direct binding assays with proteins involved in pole plasm function, we found that Vls interacts with Tud.

Vls is a member of the WD-repeat protein superfamily

The feature characterizing the superfamily of WD-repeat proteins is the WD motif, a ∼40 amino acid stretch typically containing a GH dipeptide 11-24 residues from its N-terminus and a WD dipeptide at its C terminus, albeit exhibiting only limited amino acid sequence conservation(Yu et al., 2000). When present in a protein, the WD motif typically occurs in multiple tandem repeated units. Based on structural analysis, the conformation of the WD repeat is defined as a series of four-stranded anti-parallel β sheets,which fold into a higher-order structure termed a β-propeller. At least four repeats are required to constitute a β-propeller(Chothia et al., 1997). We found that Vls contains four conserved WD repeats and can potentially form aβ-propeller structure. Interestingly, the nearest mammalian homologue of Vls, the human MEP50 protein, displays six WD repeats. Although the sequence of the two additional repeats is conserved overall in Drosophila Vls,no characteristic GH and WD di-peptides could be found in these repeats(Fig. 1D).

WD-repeat proteins act as scaffolding/anchoring proteins for a number of binding partners (Smith et al.,1999). WD-repeat motifs within one protein can simultaneously bind several proteins and foster transient interactions with other proteins. Moreover, WD-repeat proteins occur in relatively stable protein complexes in which they play a structural role. A similar function can be assigned to Vls in promoting either permanent or transient interaction with other proteins.

Evolutionary conserved interaction between Vls and Csul

The human homologue of Vls, the MEP50 protein, was identified through its association with PRMT5 (Friesen et al.,2002), the human homologue of Drosophila Csul methyltransferase. We show that the domain of Vls required for binding to Csul is small and confined to its C-terminal end. By contrast, the domain of MEP50 specifically involved in the interaction with PRMT5 remains unknown.

MEP50 acts as an adaptor binding to a subset of the Sm proteins and contributing to their methylation by the PRMT5 methyltransferase(Friesen et al., 2002). Biochemical assays indicate that MEP50 is necessary for the methyltransferase activity of the methylosome, as anti-MEP50 antibodies significantly reduce the methylation of Sm proteins (Friesen et al., 2002). However, the precise role of MEP50 remains elusive. Its possible functions include a regulation of the enzymatic activity of PMRT5 and the control of the positioning of the substrate for methylation.

The human methylosome complex is involved in the assembly of spliceosomal U-rich small nuclear ribonucleoproteins (snRNPs) mediated by the survival motoneuron (SMN) protein, a gene product that is affected in spinal muscular atrophy (Lefebvre et al.,1995; Meister et al.,2002). SMN is produced ubiquitously and contains a single Tudor domain that associates with SmB/B′, SmD1-D3 and SmE proteins of snRNPs(Liu et al., 1997; Buhler et al., 1999; Selenko et al., 2001). The assembly of snRNPs mediated by SMN occurs in the cytoplasm and is stimulated by the PMRT5-methylosome complex that converts specific arginine residues in the Sm proteins into dimethylarginines(Meister at al., 2001; Friesen et al., 2001a; Branscombe et al., 2001),facilitating the binding of the Sm proteins to SMN and their association with snRNA molecules (Friesen et al.,2001b; Brahms et al.,2001). Ultimately, the assembled snRNPs are released and targeted to the nucleus, whereas the SMN-PRMT5 complex may dissociate before its components associate again for a new round of assembly.

In Drosophila, the Tud protein differs from SMN by containing eight Tudor domains and two Tudor-like domains(Callebaut and Mornon, 1997),whereas SMN contains a single Tudor domain(Pontig, 1997). It is thus possible to envisage that Drosophila Tud may bind different categories of cytoplasmic RNPs through its multiple Tudor domains. However, in contrast to the PMRT5/MEP50 complex that apparently binds SMN through other protein(s) present in the complex, we found that Vls can directly bind to Tud through its first WD repeat. As the C-terminal tail of Vls binds to Csul, it is possible that the Drosophila Csul/Vls methylosome associates with Tud through Vls. Thus, we propose that the association between the methylosome and Tud promotes binding and assembly of specific RNPs on Tud. Further experiments are needed to unravel the relationship between these proteins, the targets of Csul methyltransferase activity and the nature of the RNPs associated to Tud.

vls functions in nuage and pole plasm

vls mutations causing a grandchild-less phenotype are characterized by agametic larvae exhibiting defects in abdominal patterning(Schüpbach and Wieschaus,1986). Eggs laid by homozygous vls females are devoid of polar granules and consequently the embryos produce no pole cells(Schüpbach and Wieschaus,1986). In these embryos, Tud is absent from the posterior pole(Bardsley et al., 1993), and Vas rapidly disappears from this location during the period of nuclear cleavage (Hay et al., 1990; Lasko and Ashburner, 1990). Our analysis reveals that localization of Tud is already absent from the nuage surrounding the vls nurse cell nuclei. However, the occurrence of Vas and Mael in the nuage of vls nurse cells indicates that aspects of this structure can be made independently of Tud.

A pivotal role in the organization of the nuage was assigned to Vas on the basis of its involvement in localizing specific components such as Aub(Wilson et al., 1996) and Mael(Clegg et al., 1997) in this structure (Harris and Macdonald,2001; Findley et al.,2003). To confirm a Vas-dependence of Tud localization in the nuage, we examined Tud distribution in vas egg chambers and found that the localization of Tud around nurse cell nuclei is fully abolished (data not shown). Similarly the absence of Tud in the nuage of vls nurse cells shows that Tud localization in the nuage is vls dependent. However, we detected significant amounts of Tud in vls oocytes and at their anterior border, indicating that Tud accumulation in the nuage is not required for Tud transport into the oocyte and its anterior margin. Furthermore, as Vls-HA does not accumulate in early oocytes nor localize at their anterior margin (data not shown), we conclude that the interaction between Vls and Tud should be spatiotemporally regulated.

We show that Vls is a component of the nuage and pole plasm. Only a limited number of proteins display a similar pattern of distribution, including Vas(Hay et al., 1988a), Aub(Harris and Macdonald, 2001)and Tud (Bardsley et al.,1993). The dual location of these proteins indicates that they either perform distinct functions at each location or exert a function in the nuage required for their accumulation in the pole plasm. The finding that Vas absence in the pole plasm correlates with its absence in the nuage supports the latter possibility (Liang et al.,1994).

Inactivation of vls function exerts a further effect on the production of the short form of Osk protein. As osk mRNA localization and amount seems normal in vls embryos(Ephrussi et al., 1991), we assume that the lower amount of this form detected by immunoblotting in vls ovaries corresponds to either a defect in Osk synthesis or stability. We notice, however, that the level of Osk abundance varies considerably between individual vls oocytes with an apparently normal level in a small number of oocytes and a markedly reduced level in the majority of oocytes. The lower amount of Vas detected at the posterior pole of stage 10 vls or vls2/Df(2L)TW2 oocytes(Hay et al., 1990), can be interpreted as a consequence of the reduced amount of Osk protein, as Vas is absent from the posterior pole of osk oocytes(Hay et al., 1990; Lasko and Ashburner,1990).

The mechanism by which Vls regulates Osk synthesis and/or stability remains unknown. However, on the basis of Vls localization during oogenesis, we envisage that vls could regulate the production of the short form of Osk by two distinct mechanisms. One the one hand, vls can regulate Osk synthesis by recruiting specific enhancing factors in the pole plasm. On the other hand, Osk synthesis may be dependent on events mediated by vls occurring in the nuage. Similarly, efficient osk mRNA translation in the pole plasm could also be mediated by Aub in the nuage(Harris and Macdonald, 2001). Furthermore, recent data point out that the nuage may function in assembling or reorganizing ribonucleoprotein complexes, particularly those involving localized or translationally regulated mRNAs(Snee and Macdonald,2003).

The formation of polar granules fully depends upon vls activity(Schüpbach and Wieshaus,1986), but only partially upon tud function as polar granules in reduced number and altered morphology are observed in amorphic tud pre-blastoderm embryos(Thomson and Lasko, 2004). This raises the question of what the targets of vls function are in addition to tud and osk. Further experiments will reveal the components required for vls-dependent formation of polar granules.

As this work was being completed, another group reported the cloning of vls by positional mapping (Cavey et al., 2005). Although Cavey and colleagues identified only two WD repeats in Vls, they clearly recognized that its nearest homologue is MEP50. As the human methylosome is formed by MEP50 and PMRT5 homologous to Drosophila Vls and Csul, respectively, our finding that Vls can specifically bind to Csul indicates that it is the orthologue of MEP50 and not a `divergent WD protein'. In contrast to Cavey and colleagues, we found a restricted and dynamic Vls distribution during oogenesis, first in the nuage and then at the posterior pole of the growing oocyte. Finally, we observed that Vls is preferentially incorporated in the forming pole cells. These findings show that vls may crucially act in the nuage, germ plasm and pole cells, and are consistent with the vls mutant phenotype. Similar to Cavey and colleagues, we detected a lower amount of Osk at the posterior pole of growing vls oocytes, but, in a discrepancy that we cannot explain, we found that Osk levels were already lower in stage 9 egg chambers,whereas Cavey et al. only observed a decrement of Osk later during oogenesis in stage 11 egg chambers.

In conclusion, we demonstrate that Vls can interact physically with at least two proteins, Csul and Tud, which are specifically involved in germ-line determination. Vls, in association with Csul, constitutes the first example of a partner of a dimethylarginine protein methyltransferase whose function has been characterized in vivo. These findings reinforce their cardinal function in a pathway first elucidated through genetic investigations. Our work sets the basis for further investigations on the role of Vls, its dependence upon Csul and its involvement in specific localization of cytoplasmic proteins during the formation of a functional pole plasm.

We are grateful to R. Boswell, A. Ephrussi, S. Findley, S. Kobayashi, P. Lasko, R. Lehmann and A. Nakimura, for providing reagents. We are particularly thankful to Anne Ephrussi for her numerous inputs given to this work. We also thank the members of the DKFZ group, present and past, especially Gunter Merdes for his advice, and Joachim Marhold, Rolf Schmitt, the students of the Fachhochschule für Technik und Gestaltung (Mannheim), Tanja Hönig,Sylvia Lux and Jens Hohmann for their help.

Akiyama, T. and Okada, M. (
1992
). Spatial and developmental changes in the respiratory activity of mitochondria in early Drosophila embryos.
Development
115
,
1175
-1182.
Amikura, R., Hanyu, K., Kashikawa, M. and Kobayashi, S.(
2001
). Tudor protein is essential for the localization of mitochondrial RNAs in polar granules of Drosophila embryos.
Mech. Dev.
107
,
97
-104.
Bardsley, A., McDonald, K. and Boswell, R. E.(
1993
). Distribution of Tudor protein in the Drosophilaembryo suggests separation of functions based on site of localization.
Development
119
,
207
-219.
Boswell, R. E. and Mahowald, A. P. (
1985
). tudor, a gene required for assembly of the germ plasm in Drosophila melanogaster.
Cell
43
,
97
-104.
Breitwieser, W., Markussen, F.-H., Horstmann, H. and Ephrussi,A. (
1996
). Oskar protein interaction with Vasa represents an essential step in polar granule assembly.
Genes Dev.
10
,
2179
-2188.
Brahms, H., Meheus, L., de Brabandere, V., Fischer, U. and Lührmann, R. (
2001
). Symmetrical dimethylation of arginine residues in spliceosomal Sm protein B/B′ and the Sm-like protein LSm4, and their interaction with the SMN protein.
RNA
7
,
1531
-1542.
Branscombe, T. L., Frankel, A., Lee, J.-H., Cook, J. R., Yang,Z.-h., Pestka, S. and Clarke, S. (
2001
). PRMT5 (Janus kinase-binding protein 1) catalyzes the formation of symmetric dimethylarginine residues in proteins.
J. Biol. Chem.
276
,
32971
-32976.
Bühler, D., Raker, V., Lührmann, R. and Fischer,U. (
1999
). Essential role fort he tudor domain of SMN in spliceosomal U snRNP assembly: implications for spinal muscular atrophy.
Hum. Mol. Genet.
8
,
2351
-2357.
Callebaut, I. and Mornon, J.-P. (
1997
). The human EBNA-2 coactivator p100: multidomain organization and relationship to the staphylococcal nuclease fold and to the tudor protein involved in Drosophila melanogaster development.
Biochem. J.
321
,
125
-132.
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
,
459
-468.
Chothia, C., Hubbard, T., Brenner, S., Barns, H. and Murzin,A. (
1997
). Protein folds in the all-beta and all-alpha classes.
Annu. Rev. Biophys. Biomol. Struct.
26
,
597
-627.
Clegg, N. J., Frost, D. M., Larkin, M. K., Subrahmanyan, L.,Bryant, Z. and Ruohola-Baker, H. (
1997
). maelstromis required for an early step in the establishment of Drosophilaoocyte polarity: posterior localization of grk mRNA.
Development
124
,
4661
-4671.
Ephrussi, A. and Lehmann, R. (
1992
). Induction of germ cell formation by oskar.
Nature
358
,
387
-392.
Ephrussi, A., Dickenson, L. K. and Lehmann, R.(
1991
). oskar organizes the germ plasm and directs localization of the posterior determinant nanos.
Cell
66
,
37
-50.
Findley, S. D., Tamanaha, M., Clegg, N. J. and Ruohola-Baker,H. (
2003
). Maelstrom, a Drosophila spindle-class gene, encodes a protein that colocalizes with Vasa and RDE1/AGO1 homolog, Aubergine, in nuage.
Development
130
,
859
-871.
Friesen, W. J., Massenet, S., Paushkin, S., Wyce, A. and Dreyfuss, G. (
2001a
). SMN, the product of the spinal muscular atrophy gene, binds preferentially to dimethylarginine-containing protein targets.
Mol. Cell
7
,
1111
-1117.
Friesen, W. J., Paushkin, S., Wyce, A., Massenet, S., Pesiridis,G. S., van Duyne, G., Rappsilber, J., Mann, M. and Dreyfuss, G.(
2001b
). The methylosome, a 20S complex containing JBP1 and pICln, produces dimethylarginine-modified Sm proteins.
Mol. Cell. Biol.
21
,
8289
-8300.
Friesen, W. J., Wyce, A., Paushkin, S., Abel, L., Rappsilber,J., Mann, M. and Dreyfuss, G. (
2002
). A novel WD repeat protein component of the methylosome binds Sm proteins.
J. Biol. Chem.
277
,
8243
-8247.
Frohnhöfer, H. G., Lehmann, R. and Nüsslein-Volhard,C. (
1986
). Manipulating the anteroposterior pattern of the Drosophila embryo.
J. Embryol. Exp. Morph. Suppl
97
,
169
-179.
Gavis, E. R., Lunsford, L., Bergsten, S. E. and Lehmann, R.(
1996
). A conserved 90 nucleotide element mediates translational repression of nanos RNA.
Development
122
,
2791
-2800.
Geigy, R. (
1931
). Action de l'ultraviolet sur le pôle germinal dans l'oeuf de Drosophila melanogaster:castration et mutabilité.
Rev. Suisse Zool.
38
,
187
-288.
Golumbeski, G. S., Bardsley, A., Tax, F. and Boswell, R. E.(
1991
). tudor, a posterior-group gene of Drosophila melanogaster, encodes a novel protein and an mRNA localized during mid-oogenesis.
Genes Dev.
5
,
2060
-2070.
González-Reyes, A., Elliott, H. and St Johnston, D.(
1997
). Oocyte determination and the origin of polarity in Drosophila: the role of the spindle genes.
Development
124
,
4927
-4937.
Harris, A. N. and Macdonald, P. M. (
2001
). Aubergine encodes a Drosophila polar granule component required for pole cell formation and related to eIF2C.
Development
128
,
2823
-2832.
Hay, B., Ackerman, L., Barbel, S., Jan, L. Y. and Jan, Y. N.(
1988a
). Identification of a component of Drosophilapolar granules.
Development
103
,
625
-640.
Hay, B., Jan, L. Y. and Jan, Y. N. (
1988b
). A protein component of the Drosophila polar granules is encoded by vasa and has extensive sequence similarity to ATP-dependent helicases.
Cell
55
,
577
-587.
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
-433.
Huettner, A. F. (
1923
). The origin of the germ cells in Drosophila melanogaster.
J. Morphol.
37
,
385
-423.
Iida, T. and Kobayashi, S. (
1998
). Essential role of mitochondrially encoded large rRNA for germ-line formation in Drosophila embryos.
Proc. Natl. Acad. Sci. USA
95
,
11274
-11278.
Illmensee, K. and Mahowald, A. P. (
1974
). Transplantation of posterior polar plasm in Drosophila. Induction of germ cells at the anterior pole of the egg.
Proc. Natl. Acad. Sci. USA
71
,
1016
-1020.
Illmensee, K., Mahowald, A. P. and Loomis, M. R.(
1976
). The ontogeny of germ plasm during oogenesis in Drosophila.
Dev. Biol.
49
,
40
-65.
Jongens, T. A., Hay, B., Jan, L. Y. and Jan, Y. N.(
1992
). The germ cell-less gene product: a posteriorly localized component necessary for germ cell development in Drosophila.
Cell
70
,
569
-584.
Jongens, T. A., Ackerman, L. D., Swedlow, J. R., Jan, L. Y. and Jan, Y. N. (
1994
). Germ cell-less encodes a cell type-specific nuclear pore-associated protein and functions early in the germ-cell specification pathway of Drosophila.
Genes Dev.
8
,
2123
-2136.
Kim, J.-S. and Raines, R. T. (
1993
). Ribonuclease S-peptide as a carrier in fusion proteins.
Protein Sci.
2
,
348
-356.
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.
Lasko, P. and Ashburner, M. (
1988
). The product of the Drosophila gene vasa is very similar to eukaryotic initiation factor-4A.
Nature
335
,
611
-617.
Lasko, P. and Ashburner, M. (
1990
). Posterior localization of vasa protein correlates with, but is not sufficient for, pole cell development.
Genes Dev.
4
,
905
-921.
Lefebvre, S., Bürglen, L., Reboullet, S., Clermont, O.,Burlet, P., Viollet, L., Benichou, B., Cruaud, C., Millasseau, P., Zeviani, M. et al. (
1995
). Identification and characterization of a spinal muscular atrophy-determining gene.
Cell
80
,
155
-165.
Lehmann, R. and Nüsslein-Volhard, C.(
1986
). Abdominal segmentation, pole cell formation, and embryonic polarity require the localized activity of oskar, a maternal gene of Drosophila.
Cell
47
,
141
-152.
Lehmann, R. and Nüsslein-Volhard, C.(
1991
). The maternal gene nanos has a central role in posterior pattern formation in the Drosophila embryo.
Development
112
,
679
-691.
Liang, L., Diehl-Jones, W. and Lasko, P.(
1994
). Localization of vasa protein to the Drosophilapole plasm is independent of its RNA-binding and helicase activities.
Development
120
,
1201
-1211.
Liu, Q., Fischer, U., Wang, F. and Dreyfuss, G.(
1997
). The spinal muscular atrophy disease gene product, SMN,and ist associated protein SIP1 are in a complex with spliceosomal snRNP proteins.
Cell
90
,
1013
-1021.
Mahowald, A. P. (
1962
). Fine structure of pole cells and polar granules in Drosophila melanogaster.
J. Exp. Zool.
151
,
201
-215.
Markussen, F.-H., Michon, A.-M., Breitwieser, W. and Ephrussi,A. (
1995
). Translational control of oskar generates Short OSK, the isoform that induces pole plasm assembly.
Development
121
,
3723
-3732.
Meister, G., Eggert, C., Bühler, D., Brahms, H., Kambach,C. and Fischer, U. (
2001
). Methylation of Sm proteins by a complex containing PRMT5 and the putative U snRNP assembly factor pICln.
Curr. Biol.
11
,
1990
-1994.
Meister, G., Eggert, C. and Fischer, U. (
2002
). SMN-mediated assembly of RNPs: a complex story.
Trends Cell Biol.
12
,
472
-478.
Okada, M., Kleinman, I. A. and Schneiderman, H. A.(
1974
). Restoration of fertility in sterilized Drosophila eggs by transplantation of polar cytoplasm.
Dev. Biol.
37
,
43
-54.
Pontig, C. P. (
1997
). Tudor domains in poteins that interact with RNA.
Trends Biochem. Sci.
22
,
51
-52.
Rigaut, G, Shevchenko, A., Rutz, B., Wilm, M., Mann, M. and Séraphin, B. (
1999
). A generic protein purification method for protein complex characterization and proteome exploration.
Nat. Biotechnol.
17
,
1030
-1032.
Rongo, C., Gavis, E. R. and Lehmann, R. (
1995
). Localization of oskar RNA Regulates oskar translation and requires Oskar protein.
Development
121
,
2737
-2746.
Sano, H., Nakamura, A. and Kobayashi, S.(
2002
). Identification of a transcriptional regulatory region for germline-specific expression of vasa gene in Drosophila melanogaster.
Mech. Dev.
112
,
129
-139.
Schüpbach, T. and Wieschaus, E. (
1986
). Maternal-effect mutations altering the anterior-posterior pattern of the Drosophila embryo.
Roux's Arch. Dev. Biol.
195
,
302
-317.
Selenko, P., Sprangers, R., Stier, G., Bühler, D., Fischer,U. and Sattler, M. (
2001
). SMN Tudor domain structure and its interaction with the Sm proteins.
Nat. Struct. Biol.
8
,
27
-31.
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
-859.
Smith, T. F., Gaitatzes, C., Saxena, K. and Neer, E. J.(
1999
). The WD repeat: a common architecture for diverse functions.
Trends Biochem. Sci.
24
,
181
-185.
Snee, M. J. and Macdonald, P. M. (
2003
). Live imaging of nuage and polar granules: evidence against a precursor-product relationship and a novel role of Oskar in stabilization of polar granule components.
J. Cell Sci.
117
,
2109
-2120.
Styhler, S., Nakamura, A., Swan, A., Suter, B. and Lasko, P.(
1998
). vasa is required for Gurken accumulation in the oocyte, and is involved in oocyte differentiation and germline cyst development.
Development
125
,
1569
-1578.
Styhler, S., Nakamura, A. and Lasko, P. (
2002
). VASA localization requires the SPRY-domain and SOCS-box containing protein,GUSTAVUS.
Dev. Cell
3
,
865
-876.
Thomson, T. and Lasko, P. (
2004
). Drosophila tudor is essential for polar granule assembly and pole cell specification, but not for posterior patterning.
Genesis
40
,
164
-170.
Tomancak, P., Guichet, A., Zavorszky, P. and Ephrussi, A.(
1998
). Oocyte polarity depends on regulation of gurkenby Vasa.
Development
125
,
1723
-1732.
Török, I., Herrmann-Horle, D., Kiss, I., Tick, G.,Speer, G., Schmitt, R. and Mechler, B. M. (
1999
). Down-regulation of RpS21, a putative translation initiation factor interacting with P40, produces viable minute imagos and larval lethality with overgrown hematopoietic organs and imaginal discs.
Mol. Cell. Biol.
19
,
2308
-2321.
Wang, C. and Lehmann, R. (
1991
). nanosis the localized posterior determinant in Drosophila.
Cell
66
,
637
-647.
Wang, C., Dickinson, L. K. and Lehmann, R.(
1994
). Genetics of nanos localization in Drosophila.
Dev. Dyn.
199
,
103
-115.
Wilson, J. E., Connell, J. E. and Macdonald, P. M.(
1996
). aubergine enhances oskar translation in the Drosophila ovary.
Development
122
,
1631
-1639.
Yu, L., Gaitatzes, C., Neer, E. and Smith, T. F.(
2000
). Thirty-plus functional families from a single motif.
Protein Sci.
9
,
2470
-2476.