The Drosophila gene vasa is required for pole plasm assembly and function, and also for completion of oogenesis. To investigate the role of vasa in oocyte development, we generated a new null mutation of vasa, which deletes the entire coding region. Analysis of vasa-null ovaries revealed that the gene is involved in the growth of germline cysts. In vasa-null ovaries, germaria are atrophied, and contain far fewer developing cysts than do wild-type germaria; a phenotype similar to, but less severe than, that of a null nanos allele. The null mutant also revealed roles for vasa in oocyte differentiation, anterior-posterior egg chamber patterning, and dorsal-ventral follicle patterning, in addition to its better-characterized functions in posterior embryonic patterning and pole cell specification. The anterior-posterior and dorsal-ventral patterning phenotypes resemble those observed in gurken mutants. vasa-null oocytes fail to efficiently accumulate many localized RNAs, such as Bicaudal-D, orb, oskar, and nanos, but still accumulate gurken RNA. However, GRK accumulation in the oocyte is severely reduced in the absence of vasa function, suggesting a function for VASA in activating gurken translation in wild-type ovaries.

Segregation of the germline from the soma is a central feature of animal development. In Drosophila, the germline is determined through the activities of maternally expressed RNAs and proteins which colocalize in the pole plasm at the posterior pole of the egg (reviewed by Rongo and Lehmann, 1996). Pole cells, the progenitors of the germline, form very early in embryogenesis, then, beginning at gastrulation, they migrate into the interior of the embryo and ultimately associate with the gonadal mesoderm to form the embryonic gonads (reviewed by Williamson and Lehmann, 1996). Beginning in larval development, germ cells proliferate and differentiate in order to carry out spermatogenesis and oogenesis; among the structures assembled during oogenesis is new pole plasm, which specifies the germline for the subsequent generation of individuals.

Genetic and molecular studies have identified numerous genes which are required for pole plasm assembly and subsequent posterior segment specification and germ cell formation; many of these genes are expressed during oogenesis and produce mRNAs and/or proteins which localize in pole plasm or in polar granules, specialized organelles contained within the pole plasm (reviewed by Rongo and Lehmann, 1996). Analysis of the expression of these genes supports an early hypothesis (Mahowald, 1968) that translational control is a major mechanism regulating Drosophila germline development. The product of the vasa (vas) gene, a DEAD-box-family protein which is localized in polar granules and which shares the enzymatic functions of the translation initiation factor eIF4A (Hay et al., 1988; Lasko and Ashburner, 1988; Liang et al., 1994), is a candidate germline-specific translational regulator. For instance, levels of the short isoform of OSKAR protein (OSK), a molecule central to pole plasm assembly (Ephrussi et al., 1991; Kim-Ha et al., 1991; Ephrussi and Lehmann, 1992), are greatly reduced in vas mutant ovaries (Markussen et al., 1995; Rongo et al., 1995). Another pole plasm mRNA whose translation may be activated by VAS is nanos (nos), as nos RNA carrying an intact translational regulation element in its 3′UTR is completely inactive in embryos derived from vas mutant ovaries (Gavis et al., 1996; Dahanukar and Wharton, 1996).

While the activities of pole plasm components such as VAS have been most thoroughly studied with respect to their function in pole cell formation and specification of the posterior soma, clearly some genes involved in pole plasm assembly also function in other stages of germline development. For instance, females homozygous for either of two strong nos alleles exhibit defects in germ cell proliferation (Lehmann and Nüsslein-Volhard, 1991; Wang et al., 1994). Furthermore, pole cells lacking maternal nos function fail to complete migration and do not associate with the embryonic gonadal mesoderm (Kobayashi et al., 1996), indicating a role for nos in the transition from pole cell to functional germ cell.

Similarly, various vas alleles have defects in oogenesis and lay few or no eggs (Lasko and Ashburner, 1988, 1990; Lehmann and Nüsslein-Volhard, 1991; Schüpbach and Wieschaus, 1991). Females trans-heterozygous for Df(2L)A267 and Df(2L)TE116-GW18, two large deletion mutations which both include vas, were reported to be blocked in early vitellogenic stages of oogenesis (Lasko and Ashburner, 1988). Analysis of whether this phenotype was caused solely by loss of vas function has been confounded by the fact that these trans-heterozygous deficiency lines are haploid for a large number of genes, but that, aside from large deficiencies, a clearly null allele of vas did not exist. Four EMS-induced alleles of vas, vasD1, vasQ6, vasQ7 and vasD5, also lead to greatly reduced fertility, with many egg chambers blocked as for the trans-heterozygous deficiency females (Lehmann and Nüsslein-Volhard, 1991). The few eggs produced by females homozygous for these alleles often lack dorsal appendages and have the micropyle, a specialized vitelline membrane structure normally found only at the anterior of the egg, duplicated at the posterior (Lehmann and Nüsslein-Volhard, 1991). Again, whether these phenotypes represent the results of a complete loss of vas function is unknown. vasQ6 and vasD5 are missense mutations which alter single amino acids of VAS and both alleles produce substantial amounts of mutant protein (Liang et al., 1994), so neither of these mutations is likely to be null. For vasD1 and vasQ7 the molecular nature of the mutation is unknown, but the vas coding region is unaffected in these mutant alleles.

In this paper, we have used a new vas null allele, vasPH165, a small deletion which we generated by imprecise P-element excision, to investigate in detail the role of vas in events of oogenesis prior to pole plasm assembly. We found that abrogation of vas function results in defects in many aspects of oogenesis including control of cystocyte divisions, oocyte differentiation, and specification of posterior and dorsal follicle cell-derived structures. Furthermore, vasPH165 oocytes only weakly concentrate many oocyte-localized RNAs, although some oocyte-specific molecules, including gurken (grk; Schüpbach, 1987; Neuman-Silberberg and Schüpbach, 1993) RNA, remain concentrated in the oocyte in vas mutant ovaries. However, in the case of grk, translation is severely reduced in the absence of vas function. This provides evidence that VAS is involved in translational control mechanisms operating in early stages of oogenesis.

Fly stocks

To create a null allele of vas, excision lines were generated through the introduction of the Δ2-3 transposase source into vasP(ry[+])LYG2 (Rittenhouse and Berg, 1995). To do this, +/Y; vasP(ry[+])LYG2cn; ry506 males were crossed to w; Bic-DPA66Su(Bic-DPA66) cn/CyO; Δ2-3Sb/TM3 Ser virgin females. w/Y; vasP(ry[+])LYG2cn/CyO; Δ2-3Sb/ ry506 males were then crossed to +/+; l(2)05084P(ry[+]l(2)05084/CyO; ry506 virgin females. Individual ry F1 males representing excisions of vasP(ry[+])LYG2were then individually crossed to l(2)05084P(ry[+]l(2)05084/CyO; ry506 virgin females. Balanced ry stocks were generated and females homozygous for an excision chromosome were crossed to Oregon R males to test for fertility. Excision lines were screened for deletions through Southern analysis using a 1.9-kb EcoRI genomic fragment from the vas first intron, and which includes the vasP(ry[+])LYG2 insertion site, as a probe. VAS protein levels from excision lines were determined by western analysis.

vasP(ry[+])LYG2 was provided by Celeste Berg (University of Washington, Seattle); l(2)05084P(ry[+]l(2)05084/CyO was provided by the Berkeley Drosophila Genome Project; nosRC, which disrupts a splice donor site and is null for RNA and protein expression (Wang et al., 1994; Curtis et al., 1997) was received from Takahiro Akiyama (Azabu University, Kanagawa, Japan); and grkHK36, a strong grk allele (Schüpbach, 1987) was obtained from Trudi Schüpbach (Princeton University). The wild-type strain employed was Oregon R.

In situ hybridization and antibody staining

In situ hybridizations with DIG-labeled RNA probes and antibody stainings were carried out on ovaries and embryos as described by Kobayashi et al. (1998), except that DMSO was omitted from the fixation solution used for ovaries. Primary antibodies were used at the following dilutions: α-ORB, 1:20; α-BIC-D, 1:10; α-NOS, 1:150; α-GRK, 1:3000; α-ADD-87, 1:20. α-NOS was pre-adsorbed with embryos from nosBN females, and α-GRK was preadsorbed with grkHK36 ovaries. For bright-field microscopy, antibody stainings were detected with DAB, enhanced using the Vectastain ABC or ABC Elite kits (Vector Laboratories) and biotinylated secondary antibodies. For confocal microscopy, antibody stainings were detected using Texas Red-conjugated secondary antibodies (Molecular Probes). β-gal staining of khc:lacZ ovaries was carried out as described by Clark et al. (1994).

OliGreen/Texas Red-phalloidin staining

Ovaries were dissected in PBS and fixed for 20 minutes in a mixture of 600 μl heptane, 200 μl 4% paraformaldehyde in PBS + 0.2% Tween-20, and 20 μl DMSO. Following fixation, samples were rinsed 3 times with PBT (PBS + 0.1% Tween-20), incubated for 1 hour in RNase A (100 mg/ml), and rinsed again 3 times with PBT. Ovaries were incubated for 1 hour in OliGreen (Molecular Probes, 1:1000 dilution) and Texas Red-phalloidin (Molecular Probes), washed with PBT, and mounted in 70% glycerol in PBS.

vasPH165 is a vas null allele

Identification of additional vas cDNA clones has indicated the presence of a 127-bp exon upstream of the previously reported 5′ end of vas, which extends the 5′ UTR of the gene. This exon is separated by a large intron of 6603 bp from the remainder of the gene (Fig. 1A,B; Berkeley Drosophila Genome Project, unpublished results, submitted to GenBank under accession numbers L81347, L81348, L81449, AC000466, and AC000469). vasLYG2 is a P-element-induced vas allele (Rittenhouse and Berg, 1995), which we mapped to the large first intron of vas (Fig. 1A), and which produces about 2% of the wild-type level of VAS (Fig. 1C), a level essentially undetectable in tissue staining experiments. Despite the very low level of VAS in vasLYG2 ovaries its phenotype is hypomorphic; oogenesis in vasLYG2 females is less severely compromised than in Df(2L)A267/Df(2L)TE116-GW18 flies, and vasLYG2 females lay numerous eggs. This suggested to us either that trace amounts of VAS are sufficient for oogenesis to often proceed to completion, or that the more severe phenotype observed in the double-deficiency lines results from the effects of a reduction in dosage of a second gene which enhances the vas phenotype.

Fig. 1.

(A) Organization of the vas gene. Exons (boxes) are numbered with Roman numerals (I-VIII), the translational start codon is in exon II; E, EcoRI site. The site of the P{ry+, vasLYG2} insertion is indicated, and the insertion site is immediately after position 75581 in the BDGP P1 clone DS00929 (GenBank accession number AC002502). vasPH165 results from an imprecise excision of P{ry+, vasLYG2}, in which 7343 bp of genomic DNA, including the entire vas coding region, are deleted, and replaced with 16 bp from the P element (sequence highlighted). (B) The 5′?UTR of vas. The first 127 nucleotides make up exon I, the 6603-bp first intron follows after nucleotide 127 (solid triangle) and nucleotide 128 corresponds to nucleotide 76 as reported in Lasko and Ashburner (1988). Nucleotides 151-153 are the initiator ATG (underlined). (C; top panel) Western blot probed with α-VAS antiserum to compare the levels of VAS in Oregon R, vasLYG2 and vasPH165 ovarian extracts. Extracts were loaded in various amounts as indicated at the top of the figure. (Bottom panel) The same blot probed with α-eIF4A antiserum (Lavoie, 1995) as a loading control.

Fig. 1.

(A) Organization of the vas gene. Exons (boxes) are numbered with Roman numerals (I-VIII), the translational start codon is in exon II; E, EcoRI site. The site of the P{ry+, vasLYG2} insertion is indicated, and the insertion site is immediately after position 75581 in the BDGP P1 clone DS00929 (GenBank accession number AC002502). vasPH165 results from an imprecise excision of P{ry+, vasLYG2}, in which 7343 bp of genomic DNA, including the entire vas coding region, are deleted, and replaced with 16 bp from the P element (sequence highlighted). (B) The 5′?UTR of vas. The first 127 nucleotides make up exon I, the 6603-bp first intron follows after nucleotide 127 (solid triangle) and nucleotide 128 corresponds to nucleotide 76 as reported in Lasko and Ashburner (1988). Nucleotides 151-153 are the initiator ATG (underlined). (C; top panel) Western blot probed with α-VAS antiserum to compare the levels of VAS in Oregon R, vasLYG2 and vasPH165 ovarian extracts. Extracts were loaded in various amounts as indicated at the top of the figure. (Bottom panel) The same blot probed with α-eIF4A antiserum (Lavoie, 1995) as a loading control.

To distinguish between these possibilities, it was necessary to obtain a molecularly characterizable null vas allele. For this purpose, we generated a series of derivative lines by mobilizing the P element in vasLYG2. Out of 201 excision events, stable lines were generated from 129. Of these, 119 caused reversion to a wild-type phenotype, indicating that the vasLYG2 phenotype results solely from the P-element insertion, and that the vasLYG2 chromosome is free of other female-sterile mutations. Three other derivatives were recessive lethal, and seven, although having excised the ry+ marker on the P element of vasLYG2, remained defective in oogenesis. These were checked by PCR and Southern blots to determine whether they carried a deletion confined to the vas gene, and by western blots using anti-VAS antiserum to determine whether they expressed VAS. From these analyses, one line, vasPH165, was identified, which produced no detectable VAS protein and carried a 7343-bp deletion which removes the entire coding region of vas (Fig. 1A,C). We confirmed the breakpoints of vasPH165 by nucleotide sequencing, comparing the mutant sequence to wild-type sequence provided by the Berkeley Drosophila Genome Project. From the nature of the vasPH165 mutation and the fact that no VAS protein is detectable on western blots even on long overexposures (Fig. 1C), we conclude that vasPH165 is null for VAS. The vasPH165 deletion is mostly limited to vas, as one of its breakpoints lies within vas and the other is 1270 bp downstream of its 3′ end. A nested gene may be located within the 3.5-kb third intron of vas, which is deleted in vasPH165, as a 3-kb transcript present throughout all developmental stages is detected on northern blots using genomic probes including this intron but not with vas cDNA probes (Lasko and Ashburner, 1988). However, any gene other than vas which may be disrupted in vasPH165 is almost certainly irrelevant to the discussion below, as a vas-GFP transgene, constructed from a vas cDNA fused to the vas promoter, and therefore lacking any nested or 3′-flanking genes, rescues vasPH165 homozygotes, and Df(2L)A267/ Df(2L)TE116-GW18 flies to fertility. The phenotypes of vasPH165/vasPH165, vasPH165/ Df(2L)A267, and Df(2L)A267/ Df(2L)TE116-GW18 ovaries are essentially identical.

VAS is involved in the maintenance of germline cysts

Upon cursory examination of vasPH165 ovaries under the light microscope, we noticed an obvious atrophy of the germaria as compared with the wild-type, suggesting that fewer germ cells were present. To investigate this more closely, we used an antibody recognizing ADD-87 protein (Lin et al., 1994; Zaccai and Lipshitz, 1996) which stains spectrosomes and fusomes, specific structures present in stem cells, cystoblasts and dividing cystocyte clusters. In vasPH165 ovaries, a reduction in the number of ADD-87-staining structures, and therefore a reduction in the number of developing cysts, is readily apparent as compared with wild-type (Fig. 2A-D). This phenotype has high expressivity, and increases in severity with the age of the female (Fig. 2C-D). In germaria from 7-day-old vasPH165 females, region 1 frequently consists of only a few stem cells and developing cysts. Posterior to these there is often what appears to be an extended interfollicular stalk (compare Fig. 2D with Fig. 2C), likely formed by follicle cells in the absence of cystocyte clusters. This suggests that the cystocyte clusters remaining in the germarium have ceased to develop further and have degenerated, while the follicle cells continue to divide.

Fig. 2.

Confocal micrographs of germaria from (A) wild-type, (B) grkHK36/grkHK36, (C) 4-day-old vasPH165/vasPH165, (D) 7-day-old vasPH165/vasPH165, and (E,F) 1-to 2-day old nosRC/nosRC ovaries stained for the ADD-87 protein, which marks spectrosome and fusome structures and which is diagnostic for stem cells and developing germline cysts. Substantially fewer foci of ADD-87 are observed in all vasPH165 or nosRC germaria as compared with the wild-type or with grkHK36. Similar phenotypes were observed in vasPH165/Df(2L)A267 ovaries.

Fig. 2.

Confocal micrographs of germaria from (A) wild-type, (B) grkHK36/grkHK36, (C) 4-day-old vasPH165/vasPH165, (D) 7-day-old vasPH165/vasPH165, and (E,F) 1-to 2-day old nosRC/nosRC ovaries stained for the ADD-87 protein, which marks spectrosome and fusome structures and which is diagnostic for stem cells and developing germline cysts. Substantially fewer foci of ADD-87 are observed in all vasPH165 or nosRC germaria as compared with the wild-type or with grkHK36. Similar phenotypes were observed in vasPH165/Df(2L)A267 ovaries.

Two other genes that function in the same pathway as vas in posterior patterning, nos and pum, have also been implicated in germ cell proliferation (Lehmann and Nüsslein-Volhard, 1991; Wang et al., 1994; Lin and Spradling, 1997), raising the possibility that they also may be interacting with vas in the germarium. pum mutants produce ovarioles which contain only two or three clusters of undifferentiated germ cells which lack spectrosome/fusome structures or ring canals (Lin and Spradling, 1997), a phenotype we never observed in vas mutants. Upon examination of nosRC mutant ovaries with the α-ADD-87 antibody, we observed a phenotype which appears similar, but somewhat more severe, than that of vasPH165. Many nosRC ovarioles consist of a germarium with one to three cysts (Fig. 2E), followed by an extended stalk and one to three normal-looking egg chambers. In more extreme cases, only remnants of spectrosome/fusome material can be detected in the anterior of the germarium (Fig. 2F), consistent with the conclusion that these germline cells have arrested development.

VAS is involved in oocyte differentiation

At a low frequency (aprox. 1% for each), we observed defects in germline differentiation and oocyte determination in vasPH165 ovaries, including tumorous egg chambers (Fig. 3A), egg chambers with 16 nurse cells and no oocyte, others with two oocytes, and again others with a mislocalized oocyte (Fig. 3B-D). Far more frequently the normal 15 nurse cells and one oocyte are present; however, by at least two criteria the oocytes produced in vasPH165 egg chambers are not fully differentiated. In wild-type development, the nurse cell nuclei endoreplicate during pre-vitellogenic oogenesis and become highly polyploid, whereas the oocyte nucleus remains diploid and condenses into a tight karyosome (Mahowald and Kambysellis, 1980). However, in vasPH165 the oocyte nucleus appears more diffuse than does the wild-type oocyte nucleus (Fig. 3E,F).

Fig. 3.

Phenotypes of vas-null egg chambers affecting oocyte determination. Confocal micrographs of (A-B) vasPH165/vasPH165 ovaries stained with the nuclear dye Oli-Green (Molecular Probes), illustrating the following phenotypes: (A) tumorous germline cyst, (B) 16 nurse cells and no oocyte. The 16 polyploid nuclei are numbered in this panel. (C) vasPH165/ vasPH165 egg chamber with two oocytes, stained for F-actin with Texas Red-phalloidin (Molecular Probes). (D) vasPH165/Df(2L)A267 egg chamber doubly stained with Oli-Green and with Texas Red-phalloidin, illustrating a bipolar egg chamber with its oocyte in the center. We also observed bipolar egg chambers in vasPH165/vasPH165 ovaries. (E,F) Higher-magnification view of oocyte nuclei stained with Oli-Green; (E) wild-type, (F) vasPH165/vasPH165. In each of these panels, an arrowhead points to the oocyte nucleus; note the more diffuse staining in F as compared with E, apparent in almost all vasPH165 egg chambers. Similar phenotypes were observed in vasPH165/Df(2L)A267 ovaries. Scale bars 20 μm (A-C), 100 μm (D), 10 μm (E,F).

Fig. 3.

Phenotypes of vas-null egg chambers affecting oocyte determination. Confocal micrographs of (A-B) vasPH165/vasPH165 ovaries stained with the nuclear dye Oli-Green (Molecular Probes), illustrating the following phenotypes: (A) tumorous germline cyst, (B) 16 nurse cells and no oocyte. The 16 polyploid nuclei are numbered in this panel. (C) vasPH165/ vasPH165 egg chamber with two oocytes, stained for F-actin with Texas Red-phalloidin (Molecular Probes). (D) vasPH165/Df(2L)A267 egg chamber doubly stained with Oli-Green and with Texas Red-phalloidin, illustrating a bipolar egg chamber with its oocyte in the center. We also observed bipolar egg chambers in vasPH165/vasPH165 ovaries. (E,F) Higher-magnification view of oocyte nuclei stained with Oli-Green; (E) wild-type, (F) vasPH165/vasPH165. In each of these panels, an arrowhead points to the oocyte nucleus; note the more diffuse staining in F as compared with E, apparent in almost all vasPH165 egg chambers. Similar phenotypes were observed in vasPH165/Df(2L)A267 ovaries. Scale bars 20 μm (A-C), 100 μm (D), 10 μm (E,F).

This would be consistent either with a failure to form the karyosome structure, perhaps involving a premature meiotic arrest in diplotene rather than in metaphase, or an increase in ploidy in vasPH165 oocytes. A very similar nuclear morphology has been observed in spindle (spn) mutant oocytes, which has been interpreted as resulting from a delay in oocyte determination (González-Reyes et al., 1997).

Secondly, vasPH165 oocytes do not efficiently accumulate at least four oocyte-localized RNAs. In wild-type ovaries, the Bicaudal-D, orb, osk and nos RNAs all accumulate efficiently in the oocyte within the germarium and remain concentrated therein throughout the early stages of oogenesis (Ephrussi et al., 1991; Kim-Ha et al., 1991; Suter and Steward, 1991; Lantz et al., 1992; Wang et al., 1994; Fig. 4A,B,D,E,I,J,M). Oocyte accumulation of these RNAs is much less pronounced in vasPH165 egg chambers in which Bic-D RNA localization is essentially undetectable (Fig. 4C), and the concentration of orb RNA in the oocyte is only slightly above the levels in the nurse cells (Fig. 4F). In vasPH165 egg chambers, osk RNA is also not observed to accumulate in the oocyte in the germarium or first vitellarial stages of development. Rather, osk RNA tends to concentrate in a somewhat diffuse manner in the center of the egg chamber (Fig. 4K,L). Finally, loss of vas function has the least pronounced effect on the accumulation of nos RNA into the oocyte, but even in this case oocyte localization is incomplete and poorly maintained (Fig. 4N,O).

Fig. 4.

Localization of various RNAs and proteins in wild-type and vasPH165 ovaries. Micrographs of wild-type ovaries are arranged on the left of the vertical bar, and micrographs of mutant ovaries are arranged on the right. (A-C) In situ hybridizations illustrating the distribution of Bic-D RNA in (A,B) wild-type and (C) vasPH165/vasPH165 ovaries. Localized signal is virtually absent in the vas mutant. (D-F) In situ hybridizations illustrating the distribution of orb RNA in (D,E) wild-type and (F) vasPH165/vasPH165 ovaries. Localized signal is much fainter in the vas mutant. (G,H) Ovaries from (G) wild-type and (H) vasPH165/vasPH165 females stained with an antiserum recognizing ORB protein. Despite the weak localization of orb RNA to the oocyte in the vas mutant, ORB protein is abundant and oocyte-specific from about stage 5. In earlier stages of oogenesis, ORB staining is much more diffuse in the vas mutant than in the wild-type. (I-L) In situ hybridizations illustrating the distribution of osk RNA in (I,J) wild-type and (K,L) vasPH165/vasPH165 ovaries. Accumulation of osk RNA in the oocyte is greatly reduced in the vas mutant, and diffuse signal is often apparent in the center of egg chambers (K). Weak localization to the oocyte is also frequently observed (L). (M-O) In situ hybridizations illustrating the distribution of nos RNA in (M) wild-type and (N,O) vasPH165/vasPH165 ovaries. The distribution of nos RNA in vas mutant egg chambers is far more diffuse than in the wild-type, but a higher concentration of nos RNA in the oocyte than in the nurse cells is usually apparent. (P,Q) Ovaries from (P) wild-type and (Q) vasPH165/vasPH165 females stained with an antiserum recognizing NOS protein. The peak of NOS expression in developing cysts at the posterior of germarium region 1 (Wang et al., 1994) is apparent in both the wild-type and mutant ovaries (arrowheads). For each probe and antiserum presented, the control and vasPH165/vasPH165 stainings were processed in parallel, and color reactions proceeded for identical lengths of time.

Fig. 4.

Localization of various RNAs and proteins in wild-type and vasPH165 ovaries. Micrographs of wild-type ovaries are arranged on the left of the vertical bar, and micrographs of mutant ovaries are arranged on the right. (A-C) In situ hybridizations illustrating the distribution of Bic-D RNA in (A,B) wild-type and (C) vasPH165/vasPH165 ovaries. Localized signal is virtually absent in the vas mutant. (D-F) In situ hybridizations illustrating the distribution of orb RNA in (D,E) wild-type and (F) vasPH165/vasPH165 ovaries. Localized signal is much fainter in the vas mutant. (G,H) Ovaries from (G) wild-type and (H) vasPH165/vasPH165 females stained with an antiserum recognizing ORB protein. Despite the weak localization of orb RNA to the oocyte in the vas mutant, ORB protein is abundant and oocyte-specific from about stage 5. In earlier stages of oogenesis, ORB staining is much more diffuse in the vas mutant than in the wild-type. (I-L) In situ hybridizations illustrating the distribution of osk RNA in (I,J) wild-type and (K,L) vasPH165/vasPH165 ovaries. Accumulation of osk RNA in the oocyte is greatly reduced in the vas mutant, and diffuse signal is often apparent in the center of egg chambers (K). Weak localization to the oocyte is also frequently observed (L). (M-O) In situ hybridizations illustrating the distribution of nos RNA in (M) wild-type and (N,O) vasPH165/vasPH165 ovaries. The distribution of nos RNA in vas mutant egg chambers is far more diffuse than in the wild-type, but a higher concentration of nos RNA in the oocyte than in the nurse cells is usually apparent. (P,Q) Ovaries from (P) wild-type and (Q) vasPH165/vasPH165 females stained with an antiserum recognizing NOS protein. The peak of NOS expression in developing cysts at the posterior of germarium region 1 (Wang et al., 1994) is apparent in both the wild-type and mutant ovaries (arrowheads). For each probe and antiserum presented, the control and vasPH165/vasPH165 stainings were processed in parallel, and color reactions proceeded for identical lengths of time.

Many oocyte-specific RNAs are still translated, and accumulate in vas-null oocytes despite defects in RNA localization

Despite the poor localization of orb RNA, ORB protein accumulates efficiently in vasPH165 oocytes, although somewhat later in oogenesis than in the wild-type (compare Fig. 4H with Fig. 4G). NOS protein expression is also broadly similar to wild-type in vasPH165 ovaries. The reduced number of developing cysts in vasPH165 germaria complicates the analysis, but the peak of NOS protein expression in 4-to 8-cell cysts at the posterior of region 1 (Wang et al., 1994) remains apparent in vasPH165 (Fig. 4P,Q). OSK protein is not detectably translated in early oogenesis in wild-type nor in vasPH165 egg chambers (Kim-Ha et al., 1995; Rongo et al., 1995; data not shown).

In wild-type oocytes, BIC-D and EGALITARIAN (EGL) proteins colocalize in the oocyte cytoplasm, in a posterior crescent from stages 2-7 (Suter and Steward, 1991; Mach and Lehmann, 1997; Fig. 5A,C,D). In vasPH165, BIC-D and EGL are efficiently translated and are localized to the oocyte. However, the two proteins are tightly concentrated in a focus which, in higher magnification in both bright-field and confocal microscopy, appears to be adjacent to the oocyte nucleus (Fig. 5B,E, and data not shown).

Fig. 5.

(A,B) Ovaries from (A) wild-type and (B) vasPH165/vasPH165 females stained with a monoclonal antibody recognizing BIC-D protein. In the vas mutant, BIC-D protein is localized in a tight focus adjacent to the oocyte nucleus throughout previtellogenic stages of oogenesis. (C-E) Ovaries from (C,D) wild-type and (E) vasPH165/vasPH165 females stained with an antiserum recognizing EGL protein. The distribution of EGL is identical to that of BIC-D in both the wild-type and the vas mutant.

Fig. 5.

(A,B) Ovaries from (A) wild-type and (B) vasPH165/vasPH165 females stained with a monoclonal antibody recognizing BIC-D protein. In the vas mutant, BIC-D protein is localized in a tight focus adjacent to the oocyte nucleus throughout previtellogenic stages of oogenesis. (C-E) Ovaries from (C,D) wild-type and (E) vasPH165/vasPH165 females stained with an antiserum recognizing EGL protein. The distribution of EGL is identical to that of BIC-D in both the wild-type and the vas mutant.

vasPH165 egg chambers often degenerate, usually at stage 6 but occasionally later

Most vasPH165 egg chambers develop into germline cysts containing 15 nurse cells and a posteriorly localized oocyte. These appear normal until about stage 6, whereupon all the nuclei undergo pycnosis and further development ceases (Fig. 6A). Approximately 70% of all vasPH165 egg chambers examined terminate development in this manner. A minority (approx. 20%) of vasPH165 oocytes continue developing beyond stage 6; most of these are blocked at stage 9-10, after which the oocyte often loses its integrity at the anterior end, and nurse cell nuclei invade the region of the egg chamber normally occupied only by the oocyte (Fig. 6B). Finally, we found that a small number of oocytes produced by vasPH165 females complete oogenesis, but that these frequently have duplicated micropyles at both the anterior and posterior ends (Fig. 6C). These eggs also often have dorsal appendage defects consistent with ventralization of the chorion: of 73 eggs analyzed, 14 (19%) lacked dorsal appendages; 48 (66%) had a single fused dorsal appendage; and only 11 (15%) had two dorsal appendages.

Fig. 6.

Phenotypes of vas-null egg chambers affecting later stages of oogenesis. (A) Confocal micrograph of a vasPH165/vasPH165 ovariole, stained with the nuclear dye Oli-Green, showing pycnosis and degeneration after stage 6. (B) Confocal micrograph of a vasPH165/Df(2L)A267 egg chamber, stained with Oli-Green and Texas Red-phalloidin, illustrating a nurse cell nucleus invading the anterior of a stage-10 oocyte. (C) Nomarski optics image of a vasPH165/Df(2L)A267 egg chamber with duplicated micropyles (arrows). The phenotypes depicted in B and C are also observed in vasPH165/vasPH165 ovaries, at a similar frequency to vasPH165/Df(2L)A267.

Fig. 6.

Phenotypes of vas-null egg chambers affecting later stages of oogenesis. (A) Confocal micrograph of a vasPH165/vasPH165 ovariole, stained with the nuclear dye Oli-Green, showing pycnosis and degeneration after stage 6. (B) Confocal micrograph of a vasPH165/Df(2L)A267 egg chamber, stained with Oli-Green and Texas Red-phalloidin, illustrating a nurse cell nucleus invading the anterior of a stage-10 oocyte. (C) Nomarski optics image of a vasPH165/Df(2L)A267 egg chamber with duplicated micropyles (arrows). The phenotypes depicted in B and C are also observed in vasPH165/vasPH165 ovaries, at a similar frequency to vasPH165/Df(2L)A267.

vas activity is required for efficient accumulation of GRK protein

The phenotypes of late-stage vasPH165 oocytes suggest that the GRK signalling pathway (González-Reyes et al., 1995; Roth et al., 1995) may be inactive in these ovaries, and led us to investigate the expression and distribution of grk RNA and protein in vasPH165. Unlike Bic-D, orb, osk and nos RNA, localization of grk RNA to the oocyte remains highly efficient in vasPH165 egg chambers (Fig. 7). However, while in wild-type oocytes from about stage 2-3 the distribution of grk RNA forms an obvious posterior crescent (Fig. 7A), grk RNA remains tightly concentrated in an extremely small area in vasPH165 oocytes (Fig. 7B), similar to the BIC-D/EGL localization pattern in these oocytes (Fig. 5). Later in oogenesis, grk RNA becomes anteriorly localized in both wild-type and vasPH165 oocytes, although in the mutant its distribution may extend further ventrally (Fig. 7C and D).

Fig. 7.

(A-D) In situ hybridizations illustrating the distribution of grk RNA in (A,C) wild-type and (B,D) vasPH165/vasPH165 ovaries. In the vas mutant, grk RNA is localized in a tight focus within the oocyte throughout pre-vitellogenic stages of oogenesis. This distribution resolves into an anterior ring at stage 8, with a variably higher concentration at the anterodorsal corner, but not so sharp an asymmetry as is evident in the wild-type. (E-F) Ovaries from (E) wild-type and (F) vasPH165/vasPH165 females stained with an antiserum recognizing GRK protein (Neuman-Silberberg and Schüpbach, 1996). Essentially no localized signal is observed in the vas mutant. For both the in situ hybridization and antibody staining reactions, the control and vasPH165/vasPH165 stainings were processed in parallel, and color reactions proceeded for identical lengths of time. (G) Western blot prepared from extracts of ovaries dissected from OreR, wild-type; grk, grkHK36/grkHK36; vas, vasPH165/vasPH165 females, and probed with the antiserum directed against GRK (top panel) or with an antiserum directed against eIF4A (Lavoie, 1995) as a loading control (bottom panel). GRK protein level is greatly reduced in the grkHK36 and vasPH165 extracts. (H-M) Distribution of β-galactosidase activity in ovaries bearing the khc:lacZ transgene (Clark et al., 1994). H-J) Stage-8, -9, and -10 egg chambers from CyO/vasPH165; khc:lacZ individuals. (K-M) Stage-8, -9, and -10 egg chambers from vasPH165/vasPH165; khc:lacZ individuals. The control and vasPH165/vasPH165 stainings were processed in parallel for identical lengths of time.

Fig. 7.

(A-D) In situ hybridizations illustrating the distribution of grk RNA in (A,C) wild-type and (B,D) vasPH165/vasPH165 ovaries. In the vas mutant, grk RNA is localized in a tight focus within the oocyte throughout pre-vitellogenic stages of oogenesis. This distribution resolves into an anterior ring at stage 8, with a variably higher concentration at the anterodorsal corner, but not so sharp an asymmetry as is evident in the wild-type. (E-F) Ovaries from (E) wild-type and (F) vasPH165/vasPH165 females stained with an antiserum recognizing GRK protein (Neuman-Silberberg and Schüpbach, 1996). Essentially no localized signal is observed in the vas mutant. For both the in situ hybridization and antibody staining reactions, the control and vasPH165/vasPH165 stainings were processed in parallel, and color reactions proceeded for identical lengths of time. (G) Western blot prepared from extracts of ovaries dissected from OreR, wild-type; grk, grkHK36/grkHK36; vas, vasPH165/vasPH165 females, and probed with the antiserum directed against GRK (top panel) or with an antiserum directed against eIF4A (Lavoie, 1995) as a loading control (bottom panel). GRK protein level is greatly reduced in the grkHK36 and vasPH165 extracts. (H-M) Distribution of β-galactosidase activity in ovaries bearing the khc:lacZ transgene (Clark et al., 1994). H-J) Stage-8, -9, and -10 egg chambers from CyO/vasPH165; khc:lacZ individuals. (K-M) Stage-8, -9, and -10 egg chambers from vasPH165/vasPH165; khc:lacZ individuals. The control and vasPH165/vasPH165 stainings were processed in parallel for identical lengths of time.

Despite the relatively normal accumulation of grk RNA in vasPH165 oocytes, and unlike the other proteins discussed above, the effects of a loss of vas activity are striking with regard to GRK protein accumulation; essentially no localized GRK is observed in vasPH165 oocytes (Fig. 7E and F). Furthermore, as measured on western blots, the level of GRK protein is greatly reduced in vasPH165 ovaries as compared with wild-type (Fig. 7G). Upon overexposure of such western blots a small amount of GRK can be detected in both grkHK36 and vasPH165 extracts (Neuman-Silberberg and Schüpbach, 1996; data not shown). As grk is required for specification of dorsal and posterior follicle structures, the duplicated micropyles and dorsal appendage defects found in vasPH165 eggs (and in eggs produced by other vas alleles; Lehmann and Nüsslein-Volhard, 1991; Schüpbach and Wieschaus, 1991) are likely to be caused by the reduced level of GRK in vas mutants. These results also suggest that VAS may activate GRK translation in wild-type oocytes.

We examined the organization of the microtubule cytoskeleton in vasPH165 oocytes using a khc:lacZ reporter gene construct (Clark et al., 1994). In the minority of vasPH165 egg chambers that develop normally through stage 10 and can thus be assessed in this way, the distribution of KHC-β-gal in oocytes is similar to that in the wild-type (Fig. 7H-M), with the exception that the intensity of signal is much lower in mutant egg chambers. The reduced signal may indicate that only a subset of microtubules are correctly organized in vasPH165 oocytes. An alternative explanation, that VAS may be required for efficient translation of KHC-β-gal protein in oocytes, is formally possible.

Partially redundant roles for VAS in early oogenesis

We have isolated a null allele of vas, vasPH165, and its analysis reveals that VAS is involved in many stages of oogenesis, including cystocyte differentiation, oocyte differentiation, and specification of anterior-posterior polarity in the developing cysts. However, the phenotype of vasPH165 is inconsistent with a total failure of any of these processes, indicating that VAS has a partially redundant role in bringing them about. Some vas-null egg chambers, namely those with two oocytes or with mislocalized oocytes (Fig. 3C,D) are similar to those produced in spn-A, spn-B, spn-C, spn-D and homeless (spn-E) mutants (González-Reyes and St Johnston, 1994; Gillespie and Berg, 1995; González-Reyes et al., 1997). Furthermore, the vas-null oocyte nucleus appears to be very similar to those observed in spn mutants (Fig. 3F, González-Reyes et al., 1997), and defects in anterodorsal accumulation of grk mRNA in stage 9 have also been reported for spn mutants (Gillespie and Berg, 1995; González-Reyes et al., 1997). It is likely that vas and the spn genes have overlapping functions in oocyte determination, and it is noteworthy in this context that homeless, which, like vas, is required in the germline, encodes a DE-H-box protein somewhat related to VAS. A role for the spn gene products in translational activation of grk has been suggested (González-Reyes et al., 1997). Construction of multiple mutants with vasPH165 and spn mutations will be important in understanding the degree to which their functions overlap.

vas and nos are both required for germline cyst development

We have observed a decrease in the number of developing germline cysts in vasPH165 ovaries, and a similar, but more severe, phenotype occurs in ovaries from females null for nos function (Fig. 2; Wang et al., 1994; Curtis et al., 1997). Because of the similarities between this aspect of the vas and nos mutant phenotypes, and because VAS has been implicated in translational activation of nos in the pole plasm (Gavis et al., 1996; Dahanukar and Wharton, 1996), it is tempting to speculate that VAS and NOS function in the same pathway to promote germline cyst development, and that VAS may activate (but not be absolutely essential for) translation of NOS in this stage of germline development as well. Our immunostaining for NOS neither supports nor refutes this possibility. While we see no obvious lack of NOS in vasPH165 germaria (Fig. 4Q), it is presently impossible to detect differences in NOS levels in stem cells or cystoblasts, as even in wild-type ovaries these cells are not detectably stained with NOS antibody (Wang et al., 1994; Fig. 4Q). The higher level of NOS in 4-8 cell cysts, which appears to be unaffected by vasPH165, would not be responsible for cyst development, as it occurs at a later developmental stage than that at which the cysts are blocked in nosRC ovaries.

The reduction in the number of developing germline cysts we observe in vas and nos null mutants could result from a failure of the pole cells to migrate to the gonad during embryogenesis, a failure of the germ cells to regulate the cell cycle or to respond to cell division signals, or from inappropriate differentiation of stem cells into cystoblasts, depleting the germarium of proliferative stem cells. Of these three possibilities, we do not think the defect involves pole cell migration, as we never observe completely agametic ovaries in vas or nos mutants. Agametic ovaries are observed even in cases where pole cell migration is incompletely blocked, such as in the progeny of AS-Pgc females (Nakamura et al., 1996). Direct assessment of the role of zygotic VAS in pole cell migration is confounded by the fact that maternal VAS perdures until embryonic stage 15-16, long after pole cell migration is complete.

Different threshold levels of VAS are required for its various activities

In wild-type flies, vas is abundantly expressed in the female germline, and, while VAS is concentrated at the posterior pole of the oocyte from stage 10 of oogenesis, easily detectable levels of VAS are present uniformly throughout the early embryo (Lasko and Ashburner, 1990). These uniform levels are sufficient for ectopic posterior somatic structures to develop in circumstances where OSK is present outside of the pole plasm, for instance in females overexpressing osk or carrying the P[oskBRE] transgene (Smith et al., 1992; Kim-Ha et al., 1995). For embryos produced by females carrying a Bic-D dominant allele, ectopic posterior segments form dependent on vas activity (Mohler and Wieschaus, 1986), but ectopic pole cells do not form, suggesting that the threshold level of VAS necessary for determination of the posterior soma is less than that necessary to induce formation of pole cells (Wharton and Struhl, 1989). A much lower threshold level of VAS is required in early oogenesis than for either later function, since vasLYG2 homozygotes, which produce very low levels of unaltered VAS (Fig. 1C), lay a far larger number of eggs than do vasPH165 homozygotes, and show defects in oogenesis at lower penetrance and of less severity than for the null (S. S. and P. L., unpublished observations).

RNA localization is dispensable for localizing proteins to the early oocyte

A comparison of the distribution of RNAs and proteins in vas-null egg chambers suggests that localization of RNAs, such as Bic-D and orb, is largely redundant for targeting their proteins to the oocyte. Furthermore, as most vas-null egg chambers proceed through development at least to stage 6, despite the very weak localization of such RNAs, it can also be argued that RNA localization is relatively unimportant for the functions of these genes. Therefore translational control, or protein targeting by means independent of RNA localization, must have a critical role in establishing the correct distribution of these molecules. This is perhaps similar to the interplay between translational control and RNA localization apparent in later oogenesis and in embryos; targeting of OSK and NOS to the posterior pole is accomplished through a combination of RNA localization and translational repression of the unlocalized RNA (Kim-Ha et al. 1995; Markussen et al., 1995; Rongo et al., 1995; Gavis et al., 1996; Smibert et al., 1996), and, at least for nos, it has been argued that RNA localization comes about as a consequence of translational regulation (Dahanukar and Wharton, 1996).

Two separable stages of oocyte localization for grk RNA and BIC-D/EGL?

In vasPH165 ovaries, grk RNA and the BIC-D/EGL complex still localize efficiently to the oocyte, but these molecules are concentrated in a tight focus rather than being distributed more widely in the oocyte cytoplasm. This suggests that in wild-type egg chambers, localization of these molecules may occur in two steps: a vas-independent step, which concentrates them in the oocyte, and a second step which distributes them to a specific region within the oocyte, depending directly or indirectly on vas activity. Alternatively, loss of vas function results in the formation of a novel structure or structures in the oocyte cytoplasm which trap grk RNA and the BIC-D/EGL complex. The distribution of BIC-D and EGL is unaffected in grkHK36 ovaries, indicating that the abnormal localization of these two proteins in vasPH165 oocytes is not a consequence of the failure to accumulate GRK in vasPH165 (data not shown). We have found that vas-null mutations affect some of the same processes as do Bic-D and egl mutations; namely, oocyte determination (in rare 16-nurse-cell and two-oocyte egg chambers) and establishment of dorsal-ventral polarity (Swan and Suter, 1996; Mach and Lehmann, 1997; this paper). In this context, it is possible that these vas phenotypes are a result of the altered distribution of BIC-D/EGL we observed. The altered distribution of grk RNA within the early oocyte is similar to that observed in maelstrom (mael) mutants which affect the organization of oocyte microtubules (Clegg et al., 1997). However, unlike in vas-null egg chambers in which many RNAs, such as osk, are poorly transported from the nurse cells to the oocyte (Fig. 4), mael only affects RNA distribution within the oocyte (Clegg et al., 1997). Furthermore, posterior localization of KHC-βGAL in vas-null oocytes (Fig. 7) suggests that the organization of the microtubule cytoskeleton is at least partially maintained.

Is VAS a direct activator of GRK translation?

In the yeast S. cerevisiae, a protein highly similar to VAS, DED1p, has recently been implicated in translation initiation (Chuang et al., 1997; de la Cruz et al., 1997). While the loss-of-function experiments presented here cannot distinguish whether VAS is required directly or indirectly for GRK accumulation, a simple model which takes into account the molecular nature of VAS and explains the reduction of GRK protein expression in vas-null ovaries is that VAS interacts directly with grk RNA and activates its translation. This is consistent, as well, with the striking reduction we observed in GRK accumulation in vas-null oocytes despite the high level of grk RNA in these oocytes. Importantly, other RNAs, such as Bic-D and orb, whose localization to the oocyte is more severely perturbed in vas-null ovaries, are still translated at levels comparable to wild-type in vas-null oocytes. Nevertheless, demonstration of a direct effect of VAS on grk translation must await more detailed experimentation. Bacterially expressed VAS has been demonstrated to bind RNA in vitro (Liang et al., 1994). However, no specificity for RNA binding by VAS in vitro has thus far been demonstrated, though grk was not among the RNAs tested in these in vitro binding experiments. There appears to be elaborate posttranscriptional regulation of GRK in the developing oocyte, involving the activities of numerous genes. In addition to vas, the genes aubergine and encore have been implicated as essential for efficient GRK accumulation (Wilson et al., 1996; Hawkins et al., 1997).

Reduction of GRK activity explains only a subset of vas phenotypes

Duplication of anterior vitelline membrane structures occurs in strong grk mutant alleles, and GRK signalling has been implicated in the specification of posterior polar follicle cells (González-Reyes et al., 1995; Roth et al., 1995). We demonstrate in this paper that many of the oocytes that complete oogenesis in a loss-of-function vas mutant have duplicated micropyles and/or fused or reduced dorsal appendages, and four other vas alleles (PW72, QS17, AQB3, RG53) have previously been shown to produce eggs with fused and reduced dorsal appendages when hemizygous over Df(2L)osp29 (Schüpbach and Wieschaus, 1991). All of these phenotypes resemble those observed in grk mutants, making it likely that these aspects of the vas mutant phenotype are brought about by a reduction of GRK activity, particularly since we show that GRK accumulation is negligible in vas-null oocytes. However, aspects of the vas-null phenotype which manifest themselves earlier in oogenesis, such as the reduced number of developing cysts, the defects in RNA localization to the oocyte, and the aberrant localization of BIC-D and EGL within the oocyte, are not found in grk mutant ovaries (Fig. 2; S.S. and P.L., unpublished results). This suggests that VAS may regulate the activity of numerous genes. Evidence similar to that presented here implicates VAS in translational activation of Short OSK in the pole plasm (Markussen et al., 1995). Furthermore, it is possible that VAS regulates translation of other, presently unknown, RNAs whose products are involved in establishing cytoskeletal polarity in the early cyst and/or in localizing RNAs to the oocyte.

We are grateful to Takahiro Akiyama, Celeste Berg, Ruth Lehmann, John Roote, and Trudi Schüpbach for fly stocks, to Howard Lipshitz for α-ADD-87 antibodies, to Paul Schedl for α-ORB antibodies, to Shira Neuman-Silberberg and Trudi Schüpbach for α-GRK antibodies, and to Robin Wharton for α-NOS antibodies. Thanks also to Antoine Guichet, Pavel Tomancak, Anne Ephrussi, Jean-Louis Couderc, Don Rio, Lin Yue, Frank Laski, and the Berkeley Drosophila Genome Project for sharing unpublished data and reagents, to an anonymous reviewer for comments on the manuscript, and to Cate Lévesque for fly food. This research was supported by operating grants to P. L. from the National Cancer Institute of Canada (NCIC) with funds from the Canadian Cancer Society, the Natural Science and Engineering Research Council of Canada (NSERC), and by a team grant from the Fonds pour la formation de chercheurs et l’aide à la recherche du Québec (FCAR). A. N. was supported by a Japan Society for the Promotion of Science postdoctoral fellowship, and a Medical Research Council of Canada postdoctoral fellowship. A. S. and S. S. acknowledge graduate bursaries from FCAR. P. L. and B. S. are Research Scientists of the National Cancer Institute of Canada.

Chuang
,
R. Y.
,
Weaver
,
P. L.
,
Liu
,
Z.
and
Chang
,
T. H.
(
1997
).
Requirement of the DEAD-box protein ded1p for messenger RNA translation
.
Science
275
,
1468
1471
.
Clark
,
I.
,
Giniger
,
E.
,
Ruohola-Baker
,
H.
,
Jan
,
L. Y.
and
Jan
,
Y. N.
(
1994
).
Transient posterior localization of a kinesin fusion protein re?ects anteroposterior polarity of the Drosophila oocyte
.
Curr. Biol
.
4
,
289
300
.
Clegg
,
N. J.
,
Frost
,
D. M.
,
Larkin
,
M. K.
,
Subrahmanyan
,
L.
,
Bryant
,
Z.
and
Ruohola-Baker
,
H.
(
1997
).
maelstrom is required for an early step in the establishment of Drosophila oocyte polarity: posterior localization of grk mRNA
.
Development
124
,
4661
4671
.
Curtis
,
D.
,
Treiber
,
D. K.
,
Tao
,
F.
,
Zamore
,
P. D.
,
Williamson
,
J. R.
and
Lehmann
,
R.
(
1997
).
A CCHC metal-binding domain in Nanos is essential for translational regulation
.
EMBO J
.
16
,
834
843
.
Dahanukar
,
A.
and
Wharton
,
R. P.
(
1996
).
The Nanos gradient in Drosophila embryos is generated by translational regulation
.
Genes Dev
.
10
,
2610
2620
.
de la Cruz,
J.
,
Iost
,
I.
,
Kressler
,
D.
and
Linder
,
P.
(
1997
).
The p20 and Ded1 proteins have antagonistic roles in eIF4E-dependent translation in Saccharomyces cerevisiae
.
Proc. Natl. Acad. Sci. USA
94
,
5201
5206
.
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
392
.
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
.
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
.
González-Reyes
,
A.
and
St Johnston
,
D.
(
1994
).
Role of oocyte position in establishment of anterior-posterior polarity in Drosophila
.
Science
266
,
639
642
.
González-Reyes
,
A.
,
Elliott
,
H.
and
St Johnston
,
D.
(
1995
).
Polarization of both major body axes in Drosophila by gurken-torpedo signalling
.
Nature
375
,
654
658
.
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
.
Hawkins
,
N. C.
,
Van Buskirk
,
C.
,
Grossniklaus
,
U.
and
Schüpbach
,
T.
(
1997
).
Post-transcriptional regulation of gurken by encore is required for axis determination in Drosophila
.
Development
124
,
4801
4810
.
Hay
,
B.
,
Ackerman
,
L.
,
Barbel
,
S.
,
Jan
,
L. Y.
and
Jan
,
Y. N.
(
1988
).
Identi?cation of a component of Drosophila polar granules
.
Development
103
,
625
640
.
Kim-Ha
,
J
,
Smith
,
J. L
and
Macdonald
,
P. M.
(
1991
).
oskar mRNA is localized to the posterior pole of the Drosophila embryo
.
Cell
66
,
23
34
.
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.
,
Yamada
,
M.
,
Asaoka
,
M.
and
Kitamura
,
T.
(
1996
).
Essential role of the posterior morphogen nanos for germline development in Drosophila
.
Nature
380
,
708
711
.
Kobayashi
,
S.
,
Amikura
,
R.
,
Nakamura
,
A.
and
Lasko
,
P. F.
(
1998
). Techniques for analyzing protein and RNA distribution in Drosophila ovaries and embryos at structural and ultrastructural resolution. In
Advances in molecular biology: a comparative methods approach to the study of oocytes and embryos
. (ed.
J.
Richter
).
Oxford
:
Oxford University Press, in press
.
Lantz
,
V.
,
Ambrosio
,
L.
and
Schedl
,
P.
(
1992
).
The Drosophila orb gene is predicted to encode sex-speci?c germline RNA-binding proteins and has localized transcripts in ovaries and early embryos
.
Development
115
,
75
88
.
Lasko
,
P. F.
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. F.
and
Ashburner
,
M.
(
1990
).
Posterior localization of vasa protein correlates with, but is not sufficient for, pole cell formation
.
Genes Dev
.
4
,
905
921
.
Lavoie
,
C. A.
(
1995
).
The molecular and biochemical characterization of proteins involved in translation initiation in Drosophila melanogaster. Ph. D. thesis
,
McGill University
.
Lehmann
,
R.
and
Nüsslein-Volhard
,
C.
(
1991
).
The maternal gene nanos has a central role in posterior pattern formation of the Drosophila embryo
.
Development
112
,
679
691
.
Liang
,
L.
,
Diehl-Jones
,
W.
and
Lasko
,
P. F.
(
1994
).
Localization of vasa protein to the Drosophila pole plasm is independent of its RNA-binding and helicase activities
.
Development
120
,
1201
1211
.
Lin
,
H.
and
Spradling
,
A. C.
(
1997
).
A novel group of pumilio mutations affects the asymmetric division of germline stem cells in the Drosophila ovary
.
Development
124
,
2463
2476
.
Lin
,
H.
,
Yue
,
L.
and
Spradling
,
A. C.
(
1994
).
The Drosophila fusome, a germline-speci?c organelle, contains membrane skeletal proteins and functions in cyst formation
.
Development
120
,
947
955
.
Mach
,
J. M.
and
Lehmann
,
R.
(
1997
).
An Egalitarian-BicaudalD complex is essential for oocyte speci?cation and axis determination in Drosophila
.
Genes Dev
.
11
,
423
435
.
Mahowald
,
A. P.
(
1968
).
Polar granules of Drosophila. II. Ultrastructural changes during early embryogenesis
.
J. Exp. Zool
.
167
,
237
262
.
Mahowald
,
A. P.
and
Kambysellis
,
M. P.
(
1980
). Oogenesis. In
The Genetics and Biology of Drosophila, vol. 2d
(ed.
Ashburner
,
M.
and
Wright
,
T. R. W.
), pp.
141
224
,
London
:
Academic Press
.
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
.
Mohler
,
J.
and
Wieschaus
,
E. F.
(
1986
).
Dominant maternal-effect mutations of Drosophila melanogaster causing the production of double-abdomen embryos
.
Genetics
112
,
803
822
.
Nakamura
,
A.
,
Amikura
,
R.
,
Mukai
,
M.
,
Kobayashi
,
S.
and
Lasko
,
P. F.
(
1996
).
Requirement for a non-coding RNA component of Drosophila polar granules for germ cell establishment
.
Science
274
,
2075
2079
.
Neuman-Silberberg
,
F. S.
and
Schüpbach
,
T.
(
1993
).
The Drosophila dorsoventral patterning gene gurken produces a dorsally localized RNA and encodes a TGFα -like protein
.
Cell
75
,
165
174
.
Neuman-Silberberg
,
F. S.
and
Schüpbach
,
T.
(
1996
).
The Drosophila TGFα -like protein Gurken: expression and cellular localization during Drosophila oogenesis
.
Mech. Dev
.
59
,
105
113
.
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
.
Rongo
,
C.
and
Lehmann
,
R.
(
1996
).
Regulated synthesis, transport and assembly of the Drosophila germ plasm
.
Trends Genet
.
12
,
102
109
.
Roth
,
S.
,
Neuman-Silberberg
,
F. S.
,
Barcelo
,
G.
and
Schüpbach
,
T.
(
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
.
Schüpbach
,
T.
(
1987
).
Germ line and soma cooperate during oogenesis to establish the dorsoventral pattern of egg shell and embryo in Drosophila melanogaster
.
Cell
49
,
699
707
.
Schüpbach
,
T.
and
Wieschaus
,
E.
(
1991
).
Female sterile mutations on the second chromosome of Drosophila melanogaster. II. Mutations blocking oogenesis or altering egg morphology
.
Genetics
129
,
1119
1136
.
Smibert
,
C. A.
,
Wilson
,
J. E.
,
Kerr
,
K.
and
Macdonald
,
P. M.
(
1996
).
smaug protein represses translation of unlocalized nanos mRNA in the Drosophila embryo
.
Genes Dev
.
10
,
2600
2609
.
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
.
Suter
,
B.
and
Steward
,
R.
(
1991
).
Requirement for phosphorylation and localization of the Bicaudal-D protein in Drosophila oocyte differentiation
.
Cell
67
,
917
926
.
Swan
,
A.
and
Suter
,
B.
(
1996
).
Role of Bicaudal-D in patterning the Drosophila egg chamber in mid-oogenesis
.
Development
122
,
3577
3586
.
Wang
,
C.
,
Dickinson
,
L. K.
and
Lehmann
,
R.
(
1994
).
Genetics of nanos localization in Drosophila
.
Dev. Dynam
.
199
,
103
115
.
Wharton
,
R. P.
and
Struhl
,
G.
(
1989
).
Structure of the Drosophila BicaudalD protein and its role in localizing the posterior determinant nanos
.
Cell
59
,
881
892
.
Williamson
,
A.
and
Lehmann
,
R.
(
1996
).
Germ cell development in Drosophila
.
Ann. Rev. Cell Dev. Biol
.
12
,
365
391
.
Wilson
,
J. E.
,
Connell
,
J. E.
and
Macdonald
,
P. M.
(
1996
).
aubergine enhances oskar translation in the Drosophila ovary
.
Development
122
,
1631
1639
.
Zaccai
,
M.
and
Lipshitz
,
H. D.
(
1996
).
Differential distributions of two adducin-like protein isoforms in the Drosophila ovary and early embryo
.
Zygote
4
,
159
166
.