The Caenorhabditis elegans gene lin-36 acts to antagonize Ras-mediated vulval induction in a pathway that includes genes with products similar to the mammalian retinoblastoma (Rb) protein and the Rb-binding protein p48. We report that lin-36 encodes a novel protein of 962 amino acids. We demonstrate that lin-36 functions in and is expressed in the vulval precursor cells, establishing that the lin-36 pathway is involved in intercellular signaling. We also report that the lin-36 pathway and/or another pathway that is functionally redundant with the lin-36 pathway antagonize a ligand-independent activity of the receptor tyrosine kinase/Ras vulval induction pathway.

Receptor tyrosine kinase (RTK) and Ras-mediated signal transduction pathways have been found to function normally in the control of the development of numerous organisms (Pawson and Bernstein, 1990; Perrimon, 1994) and have been implicated in oncogenesis (Egan and Weinberg, 1993). The core signal transduction pathway is well conserved and well understood (Egan and Weinberg, 1993). However, little is known about how this pathway is regulated. To approach this problem, we are analyzing the regulation of the RTK- and Ras-mediated signal transduction pathway of vulval induction in Caenorhabditis elegans (for reviews, see Horvitz and Sternberg, 1991; Eisenmann and Kim, 1994; Sundaram and Han, 1996).

In vulval induction a signal from a single cell in the developing gonad, the anchor cell, activates an RTK- and Ras-mediated signal transduction pathway to induce three nearby hypodermal blast cells, P5.p, P6.p and P7.p, to adopt vulval fates. These three cells are members of the vulval equivalence group, a set of six cells collectively referred to as P(3-8).p, all of which have the potential to adopt either of two vulval fates (called 1° and 2°) or a non-vulval fate (called 3°). Mutations that decrease activity in this signaling pathway block vulval induction and result in a vulvaless (Vul) phenotype, whereas mutations that increase activity in this pathway can cause the P3.p, P4.p and P8.p cells to adopt vulval fates and result in a multivulva (Muv) phenotype.

A number of genes have been identified that appear to regulate the activity of the signal transduction pathway of vulval induction. Mutations in the genes lin-2, lin-7 and lin-10 produce a Vul phenotype similar to that produced by reduction-of-function mutations in components of the let-60 ras pathway. lin-2, lin-7 and lin-10 appear to act to localize the LET-23 receptor to the basolateral membrane of the P(3-8).p cells, thereby upregulating the activity of the receptor tyrosine kinase pathway (Ferguson et al., 1987; Hoskins et al., 1996; Simske et al., 1996).

Antagonists of the vulval inductive pathway have also been identified. A group of genes known as the synthetic Multivulva (synMuv) genes inhibit the adoption of vulval fates by the P(3-8).p cells. A synthetic interaction between two different synMuv mutations is required to produce a synMuv phenotype (Horvitz and Sulston, 1980; Ferguson and Horvitz, 1985, 1989). Specifically, the synMuv mutations define two classes of genes, A and B. Animals must carry both a class A and a class B mutation to have a Muv phenotype; animals that carry mutations in either class but not in both classes have a wild-type vulval phenotype. Ferguson and Horvitz (1989) proposed that the synMuv genes encode the components of two functionally redundant pathways that negatively regulate vulval development: if both pathways are blocked, P(3-8).p cells that would normally be prevented from adopting vulval fates instead adopt vulval fates. Four class A genes (lin-8, lin-15A, lin-38 and lin-56) and ten class B genes (lin-9, lin-15B, lin-35, lin-36, lin-37, lin-51, lin-52, lin-53, lin-54, lin-55) have been identified (Horvitz and Sulston, 1980; Ferguson and Horvitz, 1989; J. H. Thomas and H. R. Horvitz, unpublished observations). lin-15 is a complex locus with genetically separable class A and class B activities encoded by two nonoverlapping transcripts (Ferguson and Horvitz, 1989; Clark et al., 1994; Huang et al., 1994). Five synMuv genes have been cloned. Three, lin-15A, lin-15B and lin-9, encode novel proteins (Clark et al., 1994; Huang et al., 1994; Beitel, 1994). lin-35, a class B gene, encodes a protein similar to the retinoblastoma-susceptibility gene product Rb, and lin-53, another class B gene, encodes a protein similar to the RbAp48 protein, which interacts with Rb (Lu and Horvitz, 1998). Genetic mosaic analyses suggest that lin-15A, lin-15B and lin-37 act cell non-autonomously and most likely in the hypodermal syncytium, suggesting that the synMuv genes encode the components of two redundant systems by which the hypodermis inhibits vulval induction (Herman and Hedgecock, 1990; Hedgecock and Herman, 1995).

In this paper, we describe studies of the class B synMuv gene lin-36. We report that lin-36 encodes a novel protein, is likely expressed in the nuclei of the P(3-8).p cells, and acts cell autonomously.

General genetic methods and strains

C. elegans strains were maintained and genetically manipulated as described by Brenner (1974) and were grown at 20°C unless otherwise stated. N2 was the wild-type strain. Mutations used in this study are described by Hodgkin et al. (1988) or as indicated in the text.

Triple and double mutants for gene interaction studies were constructed essentially as described by Ferguson et al. (1987). For all genes tested, either null alleles or the strongest known mutations were used.

Non-complementation screens for lin-36 alleles

L4 males of genotype lin-8(n111) were mutagenized with 4,5´,8-trimethylpsoralen (TMP) and ultraviolet radiation (Yandell et al., 1994) and mated with hermaphrodites of genotype lin-8(n111); sma-3(e491) lin-36(n766) unc-32(e189). A total of 23,666 haploid genomes were screened, and one Muv non-Sma non-Unc F1 animal was isolated. The new mutation was made homozygous and backcrossed three times.

In a second non-complementation screen, L4 males of genotype lin-8(n111) were mutagenized with ethyl methanesulfonate (EMS) (Brenner, 1974) and mated with hermaphrodites of genotype lin-8(n111); lin-36(n766) unc-32(e189); lon-2(e678) xol-1(y70). A total of 21,003 haploid genomes were screened, and five Muv non-Unc non-Lon F1 animals were isolated. The new mutations were made homozygous and backcrossed three or four times.

P(3-8).p cell lineage analysis

P(3-8).p cell lineages were observed using Nomarski optics as described by Sulston and Horvitz (1977).

General DNA manipulation

General DNA manipulations and analyses were performed essentially according to the protocols of Sambrook et al. (1989). The RNA filter used for northern blots was a gift from L. Bloom.

Subclones were constructed from cosmid E02E3. MluI and NcoI were used to delete cosmid DNA to construct pJHT1 and pJHT2, respectively. The following subclones contained the DNA between the listed restriction sites: pJHT9, MluI and KpnI; pJHT12, SpeI and KpnI; pJHT13, SalI and KpnI; pJHT15, SalI and SacI; pJHT16, SalI and XbaI; pJHT18, ClaI and XbaI.

Germline transformation

Germline transformation by microinjection was done according to Mello and Fire (1995). DNA for injection was purified using Qiagen columns (Qiagen Inc., Chatsworth, CA). Cosmids and plasmids for rescue experiments were injected at concentrations of 20-80 µg/ml. The lin-36::GFP reporter was injected at 50 µg/ml. pRF4, a plasmid containing the rol-6(su1006) dominant allele, was used as a coinjection marker in all experiments at a concentration of 80 µg/ml (Mello and Fire, 1995).

Isolation of cDNAs and sequence determination

We screened a mixed-stage cDNA library (Barstead and Waterston, 1989) with the 32P-labeled 5.0 kb insert in pJHT16 and isolated one cDNA. After limited ExoIII digestion, M13 phage was used to prepare single-stranded DNA (Sambrook et al., 1989), and DNA sequences were determined using the method of dideoxy chain termination and the Sequenase enzyme (US Biochemical, Cleveland, OH). One region in the 3´ untranslated region required the use of Taq polymerase at 95°C for sequence determination (Promega, Madison, WI).

Molecular analysis of lin-36 lesions

Genomic DNA was purified from N2 and strains carrying all known lin-36 mutations. lin-36(n3090) DNA was digested with SalI and XbaI, NcoI, NcoI and XbaI, EcoRV and HinfI, and analyzed by Southern blot analysis (Sambrook et al., 1989) using the 32P-labeled insert of pJHT16 as a probe. The lin-36 coding region and the regions of introns near the splice sites were amplified using the polymerase chain reaction (PCR), and the sequences of these PCR products were determined using an automated ABI 373A cycle sequencer (Applied Biosystems, Foster City, CA).

Genetic mosaic analysis

Genetic mosaic analysis of lin-36 was conducted using the method of free duplication loss described by Herman (1984). Two strategies were adopted to find mosaic animals using a strain of genotype lin-8(n111); dpy-17(e164) ncl-1(e1865) lin-36(n766) unc-36(e251); sDp3[dpy-17(+) ncl-1(+) lin-36(+) unc-36(+)]. First, mosaic L3 and L4 larvae were isolated on the basis of gross morphology and then observed using Nomarski optics to define the cell division at which duplication loss occurred by observing the enlarged nucleoli seen in cells mutant for ncl-1 (Hedgecock and Herman, 1995). Specifically, Unc non-Dpy animals result from duplication loss in AB or ABp; Dpy non-Unc animals result from loss in P1, P2 or C (Yuan and Horvitz, 1990). In these animals, a P2 loss can be distinguished from a C loss by the segregation of all Dpy Ncl Unc progeny. Semi-Unc non-Dpy animals are the result of a loss in either ABpl or ABpr. The second strategy was to examine 1,515 seemingly wild-type L3 and L4 larvae for the Ncl phenotype using Nomarski optics. This approach enabled us to identify ABa losses and losses in the ABp lineages later than ABpl/r. The Ncl phenotype was scored primarily in neurons; it cannot be scored reliably in the Pn.p cells themselves. From the P1 lineage, the pharyngeal neurons I4, I6, M1, M4 and M5, the tail neurons PVR and DVC, and occasionally hyp11 were scored. From ABp, the HSNs and most of the ABp-derived neurons of the head and tail were scored. The subset of ABp-derived neurons with a Ncl phenotype and their lineage relationship (Sulston and Horvitz, 1977; Sulston et al., 1983; White et al., 1986) allowed us to determine the cell division at which duplication loss occurred. From ABa, the BDUs, the ALMs, the pharyngeal neurons I5, MI, the M2s and the M3s, and most of the ABa-derived neurons of the head were scored.

Construction of a lin-36::GFP reporter

A lin-36::GFP reporter was constructed by ligating a 4.4 kb SalI SacI fragment from pJHT16 to a linker fragment and SalI XmaI double-digested pPD95.79, a vector containing a green fluorescent protein (GFP) coding region with a C. elegans unc-54 3´ end and a fluorescence enhancing S65C mutation (Chalfie et al., 1994; Heim et al., 1994, 1995; A. Fire, personal communication). The linker fragment was constructed by using the primers pG5F (5´ AGAGCAATGGAAGACGC) and pG3W (5´ TCCCCCCGGGTTGTTGAGAATCCGATG) to amplify DNA from pJHT16 and digesting with XmaI and SacI. pG3W anneals to the last codons of lin-36, and adds an XmaI site followed by four additional nucleotides.

Five rescued lines were isolated after germline transformation of lin-36(n766); lin-15(n767) animals. To induce chromosomal integration, transformants were exposed to 4,100 rads of γ-irradiation from a 60Co source (Mello and Fire, 1995). Two lines that retained lin-36 rescue ability and GFP staining were isolated.

Laser microsurgery

Laser microsurgery was conducted essentially as described previously (Sulston and White, 1980; Avery and Horvitz, 1987; Thomas et al., 1990). 60 mM sodium azide was used as an anesthetic. Mock-ablated animals were used as an operation control.

Identification of new lin-36 alleles

Seven lin-36 mutations had been identified previously in screens for class B synMuv mutations (Ferguson and Horvitz, 1989; J. H. Thomas and H. R. Horvitz, unpublished observations). To isolate additional lin-36 mutations, we conducted two non-complementation screens using a lin-8(n111) class A background. In one screen, 4,5´,8-trimethylpsoralen was used as a mutagen. One lin-36 mutation, n3090, was isolated and shown by Southern hybridization to be associated with an EcoRV restriction fragment length polymorphism (RFLP). This RFLP was produced by a point mutation rather than a deletion (see below). In the other screen, EMS was used as a mutagen, and five lin-36 mutations (n3093, n3094, n3095, n3096, n3097) were isolated. This frequency of EMS-induced lin-36 alleles, 2.4×10−4 per haploid genome, is close to that expected for the average EMS-induced loss-of-function mutation in C. elegans, 5×10−4 (Brenner, 1974; Meneely and Herman, 1979; Greenwald and Horvitz, 1980).

The phenotypes caused by the strongest lin-36 mutant alleles (n747, n766, n2243, n3093, n3095, n3096) were essentially identical (Table 1), suggesting that they may be complete loss-of-function alleles. The frequency of isolation in our non-complementation screen is consistent with this hypothesis. However, animals heterozygous for a strong lin-36 allele and a deficiency of the lin-36 region displayed a phenotype weaker than that of the corresponding homozygote. For example, the penetrance of the Muv phenotype of lin-8; lin-36(n766)/nDf20 animals was 87% (n=259), whereas that of lin-8; lin-36(n766) homozygotes carrying the same heterozygous marker, unc-32(e189), was 95% (n=214). Similar results were observed for lin-36(n3096). In addition, animals heterozygous for a weak lin-36 allele and a strong lin-36 allele displayed a stronger phenotype than animals heterozygous for a weak allele and a deficiency. Specifically, the penetrance of the Muv phenotype of lin-8; lin-36(n2235)/nDf20 animals was 74% (n=194), whereas that of lin-8; lin-36(n2235)/lin-36(n766) animals was 87% (n=214) and that of lin-8; lin-36(n2235)/lin-36(n747) animals was 91% (n=188); all animals scored were heterozygous for unc-32(e189). These observations suggest that strong alleles of lin-36 reduce lin-36 function more than does a deficiency. Using the polymerase chain reaction (PCR), we showed that all of the lin-36 coding sequence was deleted from the nDf20 chromosome (data not shown). It is possible that a linked, haplo-insufficient suppressor of the synMuv phenotype is deleted by this deficiency. Alternatively, the strong lin-36 alleles may interfere with the activities of other class B synMuv genes in addition to reducing lin-36 activity. Although our data are consistent with the hypothesis that we have isolated complete loss-of-function alleles, it remains possible that the strongest lin-36 alleles are not nulls but rather partially reduce lin-36 activity and also act to antagonize the activity of other members of the lin-35 Rb pathway.

Table 1.

Penetrance of the Muv phenotype in strains carrying lin-36 mutations

Penetrance of the Muv phenotype in strains carrying lin-36 mutations
Penetrance of the Muv phenotype in strains carrying lin-36 mutations

Vulval lineage defects of lin-36

We analyzed the vulval defects of strains carrying lin-36 mutations at the cell lineage level by directly observing the lineages of the P(3-8).p descendants. Animals carrying only a strong lin-36 mutation, n766, displayed wild-type P(3-8).p cell lineages, as did animals carrying only the strong class A mutation, lin-15(n767) (Table 2). The cells of the vulval equivalence group that are not induced in the wild type, P(3,4,8).p, were induced in animals of genotype lin-36(n766); lin-15(n767). In the 10 animals examined, 29/30 of these cells were completely induced and the remaining 1/30 (a P8.p cell that divided to give an OOS lineage; see Sternberg and Horvitz, 1986, 1989, for nomenclature) showed partial induction.

Table 2.

P(3-8).p cell lineages in lin-36 mutants

P(3-8).p cell lineages in lin-36 mutants
P(3-8).p cell lineages in lin-36 mutants

Cloning of lin-36

Previous data placed lin-36 between lin-13 and unc-36 on linkage group III (Ferguson and Horvitz, 1989). We mapped lin-36 under nDf20 and between egl-5 and unc-36 (Fig. 1A), both of which had been cloned and positioned on the physical map (Wilson et al., 1994). The genetic data suggested that lin-36 was located about midway between egl-5 and unc-36, and we injected cosmid pools from this region into animals bearing a class A synMuv mutation and a lin-36 mutation in germline transformation experiments. Cosmid E02E3 rescued the Muv phenotype of both lin-8(n111); lin-36(n747) and lin-36(n766); lin-15(n767) animals and was used to construct subclones. These experiments defined a 5.0 kb minimal rescuing fragment, the XbaI SalI insert of pJHT16 (Fig. 1B).

Fig. 1.

Cloning of lin-36. (A) Genetic map of the lin-36 region of linkage group III. The deficiency nDf20 failed to complement lin-36, and of 17 recombination events scored between egl-5 and unc-36, nine were between egl-5 and lin-36 and eight were between lin-36 and unc-36. Three cosmids from the corresponding region of the physical map (Wilson et al., 1994), C04H11, F44B9 and E02E3, rescued the Muv phenotype of lin-8(n111); lin-36(n747) animals. (B) Subclones constructed from cosmid E02E3. The E02E3 F2 rescue data shown combines the results of injection experiments using both lin-8(n111); lin-36(n747) animals and lin-36(n766); lin-15(n767) animals; the subclones were tested only in the latter strain. The subclone insert is depicted by the line shown adjacent to each subclone plasmid name (see Materials and methods). (C) Gene structure of lin-36. Exons are shown as blocks. Start and stop sequences, poly(glutamine) (poly-Q) and cysteine- and histidine-rich regions and mutational changes are indicated. SL1, trans-spliced leader sequence. NLS, nuclear localization sequence.

Fig. 1.

Cloning of lin-36. (A) Genetic map of the lin-36 region of linkage group III. The deficiency nDf20 failed to complement lin-36, and of 17 recombination events scored between egl-5 and unc-36, nine were between egl-5 and lin-36 and eight were between lin-36 and unc-36. Three cosmids from the corresponding region of the physical map (Wilson et al., 1994), C04H11, F44B9 and E02E3, rescued the Muv phenotype of lin-8(n111); lin-36(n747) animals. (B) Subclones constructed from cosmid E02E3. The E02E3 F2 rescue data shown combines the results of injection experiments using both lin-8(n111); lin-36(n747) animals and lin-36(n766); lin-15(n767) animals; the subclones were tested only in the latter strain. The subclone insert is depicted by the line shown adjacent to each subclone plasmid name (see Materials and methods). (C) Gene structure of lin-36. Exons are shown as blocks. Start and stop sequences, poly(glutamine) (poly-Q) and cysteine- and histidine-rich regions and mutational changes are indicated. SL1, trans-spliced leader sequence. NLS, nuclear localization sequence.

This 5.0 kb fragment was used to probe a northern blot containing poly(A)+ RNA from mixed-stage animals and recognized a single band of approximately 3.5 kb. The same fragment was used as a probe to screen a cDNA library and isolate a single clone. The 3,400 nucleotide cDNA included a poly(A) tail and nine nucleotides of the SL1 trans-spliced leader (Krause and Hirsh, 1987; Huang and Hirsh, 1989), indicating that it was full-length. The sequence of the cDNA was compared to the genomic sequence from the region generated by the C. elegans genome sequencing project (Wilson et al., 1994) to define the exon and intron structure of the gene (Fig. 1C).

lin-36 encodes a novel protein

The coding sequence of lin-36 was determined from the cDNA clone (Fig. 2). Conceptual translation yielded a novel gene product of 962 amino acids with a predicted molecular mass of 108 kDa. The predicted protein product is highly hydrophilic and has no hydrophobic stretches of significant length. There are few obvious protein motifs. There is a potential nuclear localization sequence (AKKRK) at amino acids 528-532 (for review, see Goldfarb, 1989) and a poly(glutamine) stretch at the C terminus. The region between residues 101 and 376 is cysteine- and histidine-rich; it contains 15 histidines and 14 cysteines. One region containing two cysteines (residues 268 and 271), two histidines (residues 284 and 288) and a leucine (residue 281) fits zinc-finger spacing requirements; however, other amino acids are not conserved (Rosenberg et al., 1986; Tautz et al., 1987), and we suspect that this region does not form a zinc finger.

Fig. 2.

lin-36 cDNA and LIN-36 predicted amino acid sequences. Nucleotides (upper lines) are numbered beginning at the first nucleotide after the EcoRI linker. The first nine nucleotides correspond to the 3´ end of the SL1 trans-spliced leader. Amino acids (lower lines) are numbered beginning at the presumptive ATG start codon and ending at the residue just before the presumptive TAA stop codon. Arrowheads indicate the locations of intervening sequences. The nine introns are 56, 75, 101, 75, 101, 123, 66, 48 and 215 nucleotides long, respectively. The GenBank accession number for the lin-36 cDNA is AF104917

Fig. 2.

lin-36 cDNA and LIN-36 predicted amino acid sequences. Nucleotides (upper lines) are numbered beginning at the first nucleotide after the EcoRI linker. The first nine nucleotides correspond to the 3´ end of the SL1 trans-spliced leader. Amino acids (lower lines) are numbered beginning at the presumptive ATG start codon and ending at the residue just before the presumptive TAA stop codon. Arrowheads indicate the locations of intervening sequences. The nine introns are 56, 75, 101, 75, 101, 123, 66, 48 and 215 nucleotides long, respectively. The GenBank accession number for the lin-36 cDNA is AF104917

Molecular determination of lin-36 lesions

We used PCR to amplify lin-36 DNA from mutants to determine the sequence changes in these strains. Six mutations are missense mutations, five are nonsense mutations, one is a splice-site mutation, and one contains a missense mutation as well as a silent mutation (Table 3; Fig. 1C). Three of the ochre mutations, lin-36(n766), lin-36(n3095), lin-36(n3096) and the amber mutation lin-36(n3093) eliminate 167, 394, 521 and 384 amino acids of the 962 amino acid coding sequence, respectively. All produced a strong phenotype (Table 1). The fourth ochre mutation, lin-36(n2235), eliminates only 65 amino acids of the coding sequence and produced a weaker phenotype.

Table 3.

Sequences of lin-36 mutations

Sequences of lin-36 mutations
Sequences of lin-36 mutations

lin-36 acts cell autonomously

To determine the site of action of lin-36, we conducted a genetic mosaic analysis. A strain of genotype lin-8(n111); dpy-17(e164) ncl-1(e1865) lin-36(n766) unc-36(e251); sDp3(f) was used to generate genetic mosaics as described in Materials and methods. Mosaics were identified and characterized as L3 and L4 larvae and subsequently scored for the Muv phenotype upon reaching adulthood. Strains mutant for lin-36 exhibit maternal rescue of the Muv phenotype (Ferguson and Horvitz, 1989; J. H. Thomas and H. R. Horvitz, unpublished observations). Presumably, maternally deposited wild-type lin-36 RNA or protein persists through several cell divisions and is present in sufficient quantities in the relevant cells to affect the penetrance of the Muv phenotype. This maternal effect affects the mosaic analysis, so we determined that the penetrance of the Muv phenotype of Dpy Ncl Unc animals homozygous for lin-36 descended from duplication-bearing mothers was 69% (n=240). This number sets an upper limit on the fraction of Muv animals we expected from a mitotic loss of the duplication in the lineage from which the cellular focus of lin-36 is derived.

Duplication loss in AB, ABp, ABpl or ABpr often produced Muv animals (Fig. 3). Loss in ABa or P1, P2 or C failed to result in Muv animals. These data are consistent with an anatomical focus in both the ABpl and ABpr lineages. The P(3-8).p cells derive from both of these lineages. The somatic gonad is derived from EMS, a daughter of P1. By contrast, the major hypodermal syncytium, hyp7, derives from ABa, ABp and C; C, a descendant of P1, makes the largest hyp7 contribution during embryogenesis. ABa makes a larger contribution to hyp7 than either ABpl or ABpr, yet losses in ABpl or ABpr often resulted in a Muv phenotype, whereas losses in ABa or P1 did not (Fig. 3). Since vulval development is thought to involve the P(3-8).p cells, hyp7 and the gonadal anchor cell, these observations indicate that the P(3-8).p cells, which are derived from ABpl and ABpr, are the best candidates for the focus of lin-36 activity. Thus, lin-36 probably acts cell autonomously. As discussed below, synMuv genes with a postulated focus in hyp7 show very different clonal behavior. We observed three double losses (Fig. 3 legend). One occurred in AB and MS/EMS and resulted in a Muv phenotype, presumably because of the AB loss. Another occurred in ABpl and C and resulted in a Muv phenotype, presumably because of the ABpl loss. A third occurred in ABprapp and ABa and resulted in a Muv phenotype, presumably because of the ABprapp loss. Thus, we can account for the phenotypes of mosaic animals with double losses on the basis of one of the two losses.

Fig. 3.

Mosaic analysis of lin-36. A partial embryonic cell lineage (Sulston et al., 1983) showing the origins of cells involved in vulva development. The gonadal anchor cell (ac) is derived from EMS; P(3-8).p are derived from ABplapp and ABprapp; hyp7, the main body hypodermal syncytium, is derived from ABa, ABp and C. The number of hyp7 nuclei that each of these lineages contributes at each developmental stage (E, embryonic; L1, L2, L3, first, second and third larval stages, respectively) is noted under the lineage (Sulston and Horvitz, 1977; Sulston et al., 1983). Contributions from the P3.p, P4.p and P8.p 3° cell lineages are noted in the L3 totals even though it is not likely that these hyp7 nuclei play a role in vulval cell determination; these totals include two nuclei each from ABpl and ABpr and two additional nuclei from P8.p, which descends either from ABpl or ABpr, accounting for the variable contribution to hyp7 of the L3 ABpl and ABpr lineages. In Muv animals, P3.p, P4.p and P8.p descendants often adopt vulval cell fates rather than hyp7 cell fates. Brackets below the lineage diagram show the lineages in which duplication loss occurred and the fraction of mosaic animals in this category with a Muv phenotype. The summary of losses and phenotypes in the figure is derived from the following data where the cell in which the loss of sDp3 occurred is reported followed by the fraction of such mosaic animals that were Muv: AB, 11/13; ABp, 9/22; AB or ABp, 8/9; ABpl, 1/6; ABpla, 0/1; ABpr, 1/2; ABpra, 0/1; ABa, 0/6; ABar, 0/1; P1, 0/1; P2, 0/3; C, 0/4; P1 or P2 or C, 0/5; P1 or P2, 0/1; P2 or C, 0/4; AB and EMS or MS, 1/1; ABpl and C, 1/1; ABprapp and ABa, 1/1. Animals in which double losses occurred are excluded.

Fig. 3.

Mosaic analysis of lin-36. A partial embryonic cell lineage (Sulston et al., 1983) showing the origins of cells involved in vulva development. The gonadal anchor cell (ac) is derived from EMS; P(3-8).p are derived from ABplapp and ABprapp; hyp7, the main body hypodermal syncytium, is derived from ABa, ABp and C. The number of hyp7 nuclei that each of these lineages contributes at each developmental stage (E, embryonic; L1, L2, L3, first, second and third larval stages, respectively) is noted under the lineage (Sulston and Horvitz, 1977; Sulston et al., 1983). Contributions from the P3.p, P4.p and P8.p 3° cell lineages are noted in the L3 totals even though it is not likely that these hyp7 nuclei play a role in vulval cell determination; these totals include two nuclei each from ABpl and ABpr and two additional nuclei from P8.p, which descends either from ABpl or ABpr, accounting for the variable contribution to hyp7 of the L3 ABpl and ABpr lineages. In Muv animals, P3.p, P4.p and P8.p descendants often adopt vulval cell fates rather than hyp7 cell fates. Brackets below the lineage diagram show the lineages in which duplication loss occurred and the fraction of mosaic animals in this category with a Muv phenotype. The summary of losses and phenotypes in the figure is derived from the following data where the cell in which the loss of sDp3 occurred is reported followed by the fraction of such mosaic animals that were Muv: AB, 11/13; ABp, 9/22; AB or ABp, 8/9; ABpl, 1/6; ABpla, 0/1; ABpr, 1/2; ABpra, 0/1; ABa, 0/6; ABar, 0/1; P1, 0/1; P2, 0/3; C, 0/4; P1 or P2 or C, 0/5; P1 or P2, 0/1; P2 or C, 0/4; AB and EMS or MS, 1/1; ABpl and C, 1/1; ABprapp and ABa, 1/1. Animals in which double losses occurred are excluded.

The simplest interpretation of these results is that lin-36 acts cell autonomously in the P(3-8).p cells, although it is conceivable that lin-36 is instead required in close relatives of the P(3-8).p cells, possibly hyp7 cells derived from V3 or V5 in both the ABpl and ABpr lineages. It is also conceivable that lin-36 acts in both the P(3-8).p cells and critical role in the P(3-8).p cells. In such a model, loss of lin-36 activity in both the P(3-8).p cells and some hyp7 nuclei results in a Muv phenotype, although loss in only hyp7 does not. Although the penetrance of the Muv phenotype in animals that have lost lin-36 activity in AB or ABp (64%, n=44) is essentially the same as that in animals that have lost lin-36 activity meiotically (69%, n=240), the penetrance of the Muv phenotype in animals that have lost lin-36 activity in ABpl or ABpr is substantially lower (2/10, excluding double losses). These data are consistent with a lin-36 contribution from hyp7 in addition to that from P(3-8).p cells. However, these data are also consistent with the cell autonomous model. Specifically, the control animals that lost lin-36 activity meiotically displayed not only incomplete penetrance but also variable expressivity of the Muv phenotype, in that not all P(3-8).p cells adopted vulval fates. Since ABpl and ABpr mosaic animals have only one or two P(3-8).p cells that can adopt ectopic cell fates, the cell autonomous model predicts a reduced frequency of Muv animals from these mosaic classes compared to the AB and ABp mosaic classes.

A lin-36::GFP reporter is expressed in the nuclei of Pn.p cells

We determined the expression pattern of a lin-36::GFP reporter construct containing 1 kb of sequence 5´ to the ATG start site, the entire lin-36 open reading frame, all lin-36 introns and the GFP gene fused in-frame to the last codon of lin-36. This lin-36::GFP reporter transgene rescued the Muv phenotype of a lin-36; lin-15(n767) strain, indicating that it was expressed in cells needed for lin-36 function. lin-36::GFP was expressed in the nuclei of the P(3-8).p cells and their descendants during vulval cell determination, division and invagination (Fig. 4). Similar observations have been made with an independently constructed lin-36::GFP reporter (A. Hajnal and S. Kim, personal communication). The expression pattern of lin-36 in the P(3-8).p cells and their descendants is consistent with our conclusion from the genetic mosaic analysis that lin-36 acts cell autonomously. The P(3-8).p cell expression of lin-36::GFP was not altered by mutations in other class B genes, specifically lin-9, lin-15, lin-35, lin-51, lin-52, lin-53 and lin-55 (Ferguson and Horvitz, 1989; J. H. Thomas and H. R. Horvitz, unpublished observations) (data not shown).

Fig. 4.

lin-36::GFP reporter expression. Anterior is to the left; dorsal is up. (A,C) Photomicrographs taken using Nomarski optics. (B,D) Fluorescent images of the animals seen in A and C observed using an LP-FITC filter. Animals are from the rescued Roller line 3L26a, which carries an integrated reporter transgene and a co-integrated marker transgene. (A,B) L3 hermaphrodite with reporter expression in the nuclei of the P(3-8).p cells. The nuclei of ventral cord neurons adjacent to the P(3-8).p cells also expressed the reporter construct. (C,D) L3 hermaphrodites showing the nuclear expression of the reporter in the descendants of the P(3-8).p cells during vulval cell divisions. The lineage relationships of these cells are shown by lineage diagrams superimposed on the images. Bar, 20 µm.

Fig. 4.

lin-36::GFP reporter expression. Anterior is to the left; dorsal is up. (A,C) Photomicrographs taken using Nomarski optics. (B,D) Fluorescent images of the animals seen in A and C observed using an LP-FITC filter. Animals are from the rescued Roller line 3L26a, which carries an integrated reporter transgene and a co-integrated marker transgene. (A,B) L3 hermaphrodite with reporter expression in the nuclei of the P(3-8).p cells. The nuclei of ventral cord neurons adjacent to the P(3-8).p cells also expressed the reporter construct. (C,D) L3 hermaphrodites showing the nuclear expression of the reporter in the descendants of the P(3-8).p cells during vulval cell divisions. The lineage relationships of these cells are shown by lineage diagrams superimposed on the images. Bar, 20 µm.

Besides the P(3-8).p cells, many other cells expressed GFP in strains bearing our reporter construct. Most notably, neurons of the head, tail and ventral cord expressed lin-36::GFP throughout development, and sporadic fluorescence was infrequently observed in the germline (data not shown). Very weak staining was observed in hypodermal and intestinal nuclei.

lin-36::GFP expression in all cases was localized to nuclei. The GFP reporter gene used in the construction of this reporter construct did not contain a nuclear localization sequence, but as mentioned above, the lin-36 coding region contains a potential nuclear localization sequence. These data suggest that lin-36 may act in the nuclei of the P(3-8).p cells to mediate its effect on vulval development.

The synMuv genes inhibit a ligand-independent activity of the let-23 RTK

The Vul phenotype of mutations in the let-23 receptor tyrosine kinase gene is epistatic to the Muv phenotype of the lin-8; lin-9 and lin-15AB synMuv mutations (Ferguson et al., 1987; Huang et al., 1994). To determine if lin-36 acts similarly in the genetic pathway controlling vulval induction, we constructed triple mutants using let-23 mutations and a lin-36 mutation (Table 4). The Vul phenotype of let-23 mutations is epistatic to the lin-36; lin-15A synMuv phenotype. Thus, let-23 activity is required for the adoption of vulval fates by P(3-8).p cells in synMuv mutants. The let-23 receptor tyrosine kinase acts genetically downstream of, or in parallel to, lin-36 as well as the other synMuv genes.

Table 4.

The synMuv phenotype is coexpressed with the lin-2,7,10 Vul phenotype

The synMuv phenotype is coexpressed with the lin-2,7,10 Vul phenotype
The synMuv phenotype is coexpressed with the lin-2,7,10 Vul phenotype

lin-2, lin-7 and lin-10 positively regulate the let-23-mediated signal transduction cascade, apparently by localizing the LET-23 receptor to the basolateral membrane of the P(3-8).p cells (Simske et al., 1995; Hoskins et al., 1996). A mixture of Muv and Vul phenotypes, i.e. animals both with ectopic ventral protrusions and bloated with eggs (or bags of worms), has been reported for animals triply mutant for lin-8; lin-9 and either lin-2, lin-7 or lin-10; however, double mutants between the synMuv mutation lin-15(n309) and a lin-2, lin-7 and lin-10 mutation have been reported to be Muv (Ferguson et al., 1987). Since the lin-9 mutation used in these experiments was not a null allele (Ferguson and Horvitz, 1989), we repeated these experiments using the strong lin-36(n766) allele. Some animals were Muv, some were Vul, some had a wild-type phenotype, and some expressed both a Vul and a Muv phenotype (Table 4). Thus, in these triple mutants the Muv and Vul phenotypes were coexpressed. Since this result is similar to that observed by Ferguson et al. (1987) with lin-8; lin-9 and lin-2, lin-7 or lin-10 triple mutants, we repeated the lin-15AB experiment, but used another allele, lin-15(e1763), which is a slightly larger deletion of lin-15 than is lin-15(n309). Both alleles eliminate both the lin-15A and lin-15B gene products and their activities (Clark et al., 1994). The lin-15AB double mutants with lin-2, lin-7 and lin-10 mutations showed coexpression of the Muv and Vul phenotypes (Table 4). These results are consistent with our observations of the lin-36; lin-15A mutants; however, unlike the lin-36 triple mutants, most animals of these genotypes were either Muv or Muv Vul; few were wild type or Vul. Since almost all animals expressed the Muv phenotype and some expressed the Vul phenotype in addition to the Muv phenotype, it is likely that in the previous studies by Ferguson et al. (1987) the Vul phenotype was not noticed. We conclude that in mutants carrying class A and class B synMuv mutations and a lin-2, lin-7 or lin-10 mutation there is a coexpression of the Muv and Vul phenotypes.

The anchor cell is necessary for the adoption of vulval cell fates by the P(3-8).p cell in wild-type animals, as demonstrated by experiments in which the anchor cell was ablated by laser microsurgery (Kimble, 1981). lin-8; lin-9 and lin-15AB synMuv mutants in which the anchor cell was ablated using laser microsurgery are Muv, suggesting that in the absence of negative regulation by the synMuv genes, P(3-8).p cells do not require the anchor cell signal to adopt vulval cell fates (Ferguson et al., 1987). We tested whether anchor cell-deficient synMuv mutants carrying lin-36 mutations were similarly Muv. We eliminated the entire somatic gonad, which includes the anchor cell progenitor, using laser microsurgery to ablate Z1 and Z4, the precursors of the somatic gonad (Kimble and Hirsh, 1979). These operations were performed at the early L1 larval stage. The ablation of Z1 and Z4 prevents germline proliferation and produces a sterile animal with no gonad (Kimble and White, 1981). This phenotype served as an internal control to confirm the ablation of the somatic gonad. We ablated Z1 and Z4 in lin-36; lin-15A animals and found that 11/11 of these animals produced ectopic vulval tissue.

We tested whether the anchor cell signal is responsible for the adoption of vulval fates by P(3-8).p cells in which receptor localization is disrupted and the negative regulatory pathways are eliminated by triple mutants carrying synMuv mutations and mutations in lin-2, lin-7 or lin-10. We ablated Z1 and Z4 in lin-7; lin-36; lin-15A animals and observed that 14/14 operated animals had ectopic vulval tissue. Thus, in animals in which the gonadal signal, the receptor localization system and the synMuv-mediated inhibitory pathways have been disrupted, the P(3-8).p cells can adopt vulval fates. These results, together with the triple mutant analyses, are consistent with the hypothesis that a ligand-independent and receptor localization-independent activity of the let-23 signal transduction pathway can lead to the adoption of vulval cell fates. Presumably, it is this ligand-independent and receptor localization-independent activity that the synMuv genes normally inhibit.

lin-36 encodes a novel protein and acts cell autonomously

Our results suggest that lin-36 encodes a novel protein that is required within the P(3-8).p cells to negatively regulate vulval development. Our lin-36 genetic mosaic data are in marked contrast to the lin-15 and lin-37 genetic mosaic data that led to the conclusion that these synMuv loci act cell non-autonomously (Herman and Hedgecock, 1990; Hedgecock and Herman, 1995). Specifically, Herman and Hedgecock (1990) found that loss of lin-15 function in P1, AB or ABp, ABpl or ABpr can result in a Muv phenotype. hyp7, which derives from these lineages, was postulated to be the anatomical focus of both lin-15A and lin-15B activity, since both activities must be lost to produce a Muv phenotype. Hedgecock and Herman (1995) found that loss of lin-37 function in P1, ABp, ABpl or ABpr did not cause a Muv phenotype and that loss in AB infrequently resulted in a Muv phenotype. Loss in AB results in many hyp7 nuclei lacking lin-37 function. Hedgecock and Herman (1995) postulated that lin-37 is also required in hyp7, but that a relatively low dosage is sufficient for wild-type function. Our results, which show a requirement for lin-36 function in only those lineages that generate the P(3-8).p cells (e.g. ABp but not P1), suggest that lin-36 acts in the P(3-8).p cells and not in hyp7. This hypothesis is consistent with our lin-36 expression data.

Taken together, our findings and those of Herman and Hedgecock (1990) and Hedgecock and Herman (1995) suggest that some class B synMuv genes function in the P(3-8).p cells and others function in hyp7. Thus, the class B synMuv genes appear to encode components of an intercellular signaling system between hyp7 and P(3-8).p. Since lin-36 is expressed in the nucleus, it is likely that the inhibitory signal from the class B pathway is interpreted within the nucleus of the P(3-8).p cell.

Antagonism of the inductive signaling pathway by the Rb-mediated class B synMuv pathway

The class B synMuv genes lin-35 and lin-53 encode proteins similar to the products of the mammalian retinoblastoma (Rb) tumor suppressor gene and the Rb-binding protein RbAp48, respectively (Lu and Horvitz, 1998). Both Rb and RbAp48 are believed to be involved in transcriptional regulation (Qian et al., 1993; Taya, 1997). lin-35 Rb appears to be expressed in the nuclei of the P(3-8).p cells (Lu and Horvitz, 1998). Since lin-36 is expressed in the nuclei of the P(3-8).p cells and seems to act in these cells, it is possible that lin-36 acts as a regulator or effector of lin-35 Rb activity.

How do lin-36 and the other class B synMuv genes antagonize the let-23 signal transduction cascade? Our observations suggest that the class B synMuv genes may inhibit the adoption of vulval cell fates by acting in the nucleus, possibly at the level of transcription. If so, the class B genes would provide an input into the cell-fate decision independently of the let-23 signal transduction pathway. The nuclear role of Rb-like proteins is consistent with this hypothesis.

SynMuv mutants are Vul in a strain in which let-23 receptor gene activity is reduced but are Muv in gonad-ablated animals. These observations indicate that whereas activation of the inductive signal transduction cascade by the gonadal ligand is not necessary for the adoption of vulval fates in the absence of negative regulation by lin-36 and the other synMuv genes, the presence of the intracellular signal transduction cascade nonetheless is necessary. These results are consistent with a model in which the two classes of synMuv genes antagonize either the gonadal ligand-independent (basal) activity of the let-23-mediated inductive pathway or a weaker activity stimulated by a second, nongonadal signal of the let-23-mediated pathway. In either case, a gonadal ligand-independent activity is necessary for the adoption of vulval fates by the P(3-8).p cells in a synMuv mutant. The inhibition of this activity by the synMuv genes can be overcome by the activation of the inductive pathway by the binding of the lin-3 ligand to receptor in the subset of P(3-8).p cells nearest the anchor cell.

lin-2, lin-7 and lin-10 positively regulate the inductive signal transduction cascade by localizing the LET-23 receptor to the P(3-8).p cell junctions (Hoskins et al., 1996; Simske et al., 1996). Our gene interaction studies showed that the synMuv phenotype is coexpressed with the Vul phenotype of lin-2, lin-7 and lin-10 in mutants with lin-36 and lin-15A or with lin-15AB. This coexpression suggests that the synMuv genes act to antagonize the let-23 pathway independently of the lin-2, lin-7, lin-10 regulatory system, since neither set of genes requires the presence of the other to exert its effect on vulval development. In the absence of the activity of the synMuv genes, the proper subcellular localization of LET-23 receptor is sufficient to allow all P(3-8).p cells to adopt vulval fates. However, in the absence of both synMuv gene-mediated negative regulation and proper subcellular localization of LET-23 receptor, the P(3-8).p cells variably adopt either vulval or nonvulval fates, despite the presence of the inductive signal.

Our laser ablation experiments showed that the P(3-8).p cell fates seen in animals lacking the stimulatory anchor cell signal, positive regulation by lin-7 and negative regulation by lin-36 and lin-15A, are similar to the fates seen in animals lacking just the positive regulatory pathway mediated by lin-2, lin-7, lin-10 and the negative regulatory pathway mediated by the synMuv genes. This observation suggests that the P(3-8).p cells that adopt vulval cell fates in animals lacking both synMuv activity and receptor localization do not adopt these fates in response to the anchor cell signal. We hypothesize that LET-23 receptor clustering can amplify the signal independently of ligand binding. Not all P(3-8).p cells adopt vulval cell fates, however. Unstimulated and unlocalized let-23 activity is sufficient only for some cells to adopt vulval fates in the absence of synMuv gene-mediated inhibition.

The data described above suggest the following model. The six Pn.p cells are predisposed to adopt vulval fates by an unstimulated basal activity of the let-23 receptor tyrosine kinase (Fig. 5A). In the absence of this activity, the P(3-8).p cells cannot adopt vulval fates. However, this predisposition to adopt vulval fates is inhibited by the action of the two functionally redundant signaling pathways encoded by the synMuv genes, which antagonize the basal activity of the let-23 signal transduction pathway. The inductive signal from the anchor cell overcomes this inhibition and causes nearby P(3-8).p cells to adopt vulval cell fates, either by overriding or by inhibiting the synMuv genes (Fig. 5B).

Fig. 5.

Model for the control of the P(3-8).p cell-fate decision in wild-type animals. See text for details. (A) Cells far from the anchor cell, P(3,4,8).p, adopt nonvulval cell fates. SynMuv gene activity is sufficient to inhibit the ligand-independent (basal) activity of the let-23 signal transduction pathway. (B) Cells closer to the anchor cell, P(5,6,7).p, adopt vulval cell fates. The lin-3 ligand stimulates the let-23 receptor. The synMuv gene-mediated antagonism of the let-23 signal transduction activity is insufficient to block the signal from the activated induction pathway.

Fig. 5.

Model for the control of the P(3-8).p cell-fate decision in wild-type animals. See text for details. (A) Cells far from the anchor cell, P(3,4,8).p, adopt nonvulval cell fates. SynMuv gene activity is sufficient to inhibit the ligand-independent (basal) activity of the let-23 signal transduction pathway. (B) Cells closer to the anchor cell, P(5,6,7).p, adopt vulval cell fates. The lin-3 ligand stimulates the let-23 receptor. The synMuv gene-mediated antagonism of the let-23 signal transduction activity is insufficient to block the signal from the activated induction pathway.

We thank Greg Beitel, Kerry Kornfeld and Shai Shaham for helpful discussions during the course of this work. We thank Xiaowei Lu, Craig Ceol and Rachel Hoang for comments concerning the manuscript, Beth James for technical assistance and Alex Hajnal and Stuart Kim for sharing unpublished data. We thank Stuart Kim and the Caenorhabditis Genetics Center (supported by the National Institutes of Health National Center for Research Resources) for providing some of the strains used in this work. J. H. T. was a Predoctoral Fellow of the Howard Hughes Medical Institute. H. R. H. is an Investigator of the Howard Hughes Medical Institute.

Anderson
,
P.
(
1995
).
Mutagenesis
.
In Caenorhabditis elegans: Modern Biological Analysis of an Organism. Methods Cell Biol
.
48
(ed.
H. F.
Epstein
and
D. C.
Shakes
), pp.
31
58
.
Academic Press
,
New York
.
Arioan
,
R. V.
and
Sternberg
,
P. W.
(
1991
).
Multiple functions of let-23, a Caenorhabditis elegans receptor tyrosine kinase gene required for vulval induction
.
Genetics
128
,
251
267
.
Avery
,
L.
and
Horvitz
,
H. R.
(
1987
).
A cell that dies during wild-type C. elegans development can function as a neuron in a ced-3 mutant
.
Cell
51
,
1071
1078
.
Barstead
,
R. J.
and
Waterston
,
R. H.
(
1989
).
The basal component of the nematode dense-body is vinculin
.
J. Biol. Chem
.
264
,
10177
10185
.
Beitel
,
G. J.
(
1994
).
Genetic and molecular analyses of let-60 ras, lin-1 and lin-9: genes that function in Caenorhabditis elegans vulval induction
.
PhD Thesis, Massachusetts Institute of Technology
.
Brenner
,
S.
(
1974
).
The genetics of Caenorhabditis elegans
.
Genetics
77
,
71
94
.
Chalfie
,
M.
,
Tu
,
Y.
,
Euskirchen
,
G.
,
Ward
,
W. W.
and
Prasher
,
D. C.
(
1994
).
Green fluorescent protein as a marker for gene expression
.
Science
263
,
802
805
.
Clark
,
S. G.
,
Lu
,
X.
and
Horvitz
,
H. R.
(
1994
).
The Caenorhabditis elegans locus lin-15, a negative regulator of a tyrosine kinase signaling pathway, encodes two different proteins
.
Genetics
137
,
987
997
.
Egan
,
S. E.
and
Weinberg
,
R. A.
(
1993
).
The pathway to signal achievement
.
Nature
365
,
781
783
.
Eisenmann
,
D. M.
and
Kim
,
S. K.
(
1994
).
Signal transduction and cell fate specification during Caenorhabditis elegans vulval development
.
Curr. Opin. Genet. Dev
.
4
,
508
516
.
Ferguson
,
E. L.
and
Horvitz
,
H. R.
(
1985
).
Identification and genetic characterization of 22 genes that affect the vulval cell lineages of the nematode Caenorhabditis elegans
.
Genetics
110
,
17
72
.
Ferguson
,
E. L.
and
Horvitz
,
H. R.
(
1989
).
The multivulva phenotype of certain Caenorhabditis elegans mutants results from defects in two functionally redundant pathways
.
Genetics
123
,
109
121
.
Ferguson
,
E. L.
,
Sternberg
,
P. W.
, and
Horvitz
,
H. R.
(
1987
).
A genetic pathway for the specification of the vulval cell lineages of Caenorhabditis elegans
.
Nature
326
,
259
267
.
Fields
,
C.
(
1990
).
Information content of Caenorhabditis elegans splice site sequences varies with intron length
.
Nucl. Acids Res
.
18
,
1509
1512
.
Goldfarb
,
D. S.
(
1989
).
Nuclear transport
.
Curr. Opin. Cell Biol
.
1
,
441
446
.
Greenwald
,
I. S.
and
Horvitz
,
H. R.
(
1980
).
unc-93(e1500): a behavioral mutant of Caenorhabditis elegans that defines a gene with a wild-type null phenotype
.
Genetics
96
,
147
164
.
Hedgecock
,
E. M.
and
Herman
,
R. K.
(
1995
).
The ncl-1 gene and genetic mosaics of Caenorhabditis elegans
.
Genetics
141
,
989
1006
.
Heim
,
R.
,
Prasher
,
D. C.
and
Tsien
,
R. Y.
(
1994
).
Wavelength mutations and posttranslational autoxidation of green fluorescent protein
.
Proc. Natl. Acad. Sci. USA
91
,
12501
12504
.
Heim
,
R.
,
Cubitt
,
A. B.
and
Tsien
,
R. Y.
(
1995
).
Improved green flourescence
.
Nature
373
,
663
664
.
Herman
,
R. K.
(
1984
).
Analysis of genetic mosaics of the nematode C. elegans
.
Genetics
108
,
165
180
.
Herman
,
R. K.
and
Hedgecock
,
E. M.
(
1990
).
Limitation of the size of the vulval primordium of Caenorhabditis elegans by lin-15 expression in the surrounding hypodermis
.
Nature
348
,
169
171
.
Hodgkin
,
J.
,
Edgley
,
M.
,
Riddle
,
D. L.
and
Albertson
,
D. G.
(
1988
).
Appendix 4: Genetics
.
In The Nematode Caenorhabditis elegans
(ed.
W. B.
Wood
and
the Community Of C. Elegans Researchers
), pp.
491
584
.
Cold Spring Harbor, New York
:
Cold Spring Harbor Press
.
Horvitz
,
H. R.
and
Sternberg
,
P. W.
(
1991
).
Multiple intercellular signalling systems control the development of the Caenorhabditis elegans vulva
.
Nature
351
,
535
541
.
Horvitz
,
H. R.
and
Sulston
,
J. E.
(
1980
).
Isolation and genetic characterization of cell-lineage mutants of the nematode Caenorhabditis elegans
.
Genetics
96
,
435
454
.
Hoskins
,
R.
,
Hajnal
,
A. F.
,
Harp
,
S. A.
and
Kim
,
S. K.
(
1996
).
The C. elegans vulval induction gene lin-2 encodes a member of the MAGUK family of cell junction proteins
.
Development
122
,
97
111
.
Huang
,
L. S.
,
Tzou
,
P.
and
Sternberg
,
P. W.
(
1994
).
The Caenorhabditis elegans lin-15 locus encodes two negative regulators of vulval development
.
Mol. Biol. Cell
5
,
395
412
.
Huang
,
X-Y.
and
Hirsh
,
D.
(
1989
).
A second trans-spliced RNA leader sequence in the nematode Caenorhabditis elegans
.
Proc. Natl. Acad. Sci. USA
86
,
8640
8644
.
Kimble
,
J.
(
1981
).
Alterations in cell lineage following laser ablation of cells in the somatic gonad of Caenorhabditis elegans
.
Dev. Biol
.
87
,
286
300
.
Kimble
,
J.
and
Hirsh
,
D.
(
1979
).
The postembryonic cell lineages of the hermaphrodite and male gonads in Caenorhabditis elegans
.
Dev. Biol
.
70
,
396
417
.
Kimble
,
J. E.
and
White
,
J. G.
(
1981
).
On the control of germ cell development in Caenorhabditis elegans
.
Dev. Biol
.
81
,
208
219
.
Krause
,
M.
and
Hirsh
,
D.
(
1987
).
A trans-spliced leader sequence in actin mRNA in C. elegans
.
Cell
49
,
753
761
.
Lu
,
X.
and
Horvitz
,
H. R.
(
1998
).
lin-35 and lin-53, two genes that antagonize a C. elegans Ras pathway, encode proteins similar to Rb and its binding protein RbAp48
.
Cell
95
,
981
991
.
Mello
,
C.
and
Fire
,
A.
(
1995
).
DNA transformation
.
In Caenorhabditis elegans: Modern Biological Analysis of an Organism. Methods Cell Biol
.
48
(ed.
H. F.
Epstein
and
D. C.
Shakes
), pp.
451
482
.
Academic Press
,
New York
.
Meneely
,
P. M.
and
Herman
,
R. K.
(
1979
).
Lethals, steriles and deficiencies in a region of the X chromosome of Caenorhabditis elegans
.
Genetics
92
,
99
115
.
Pawson
,
T.
and
Bernstein
,
A.
(
1990
).
Receptor tyrosine kinases: genetic evidence for their role in Drosophila and mouse development
.
Trends Genet
.
6
,
350
356
.
Perrimon
,
N.
(
1994
).
Signalling pathways initiated by receptor protein tyrosine kinases in Drosophila
.
Curr. Opin. Cell Biol
.
6
,
260
266
.
Qian
,
Y.-W.
,
Wang
,
Y.-C. J.
,
Hollingsworth
Jr.,
R. E.
,
Jones
,
D.
,
Ling
,
N.
and
Lee
,
E. Y.-H. P.
(
1993
).
A retinoblastoma-binding protein related to a negative regulator of Ras in yeast
.
Nature
364
,
648
652
.
Rosenberg
,
U. B.
,
Schröder
,
C.
,
Priess
,
A.
,
Kienlin
,
A.
,
Cote
,
S.
,
Riede
,
I.
and
Jäckle
,
H.
(
1986
).
Structural homology of the product of the Drosophila Krüppel gene with Xenopus transcription factor IIIA
.
Nature
319
,
336
339
.
Sambrook
,
J.
,
Fritsch
,
E. F.
and
Maniatis
,
T.
(
1989
).
Molecular Cloning: A Laboratory Manual
.
Cold Spring Harbor, New York
:
Cold Spring Harbor Press
.
Simske
,
J. S.
,
Kaech
,
S. M.
,
Harp
,
S. A.
and
Kim
,
S. K.
(
1996
).
LET-23 receptor localization by the cell junction protein LIN-7 during C. elegans vulval induction
.
Cell
85
,
195
204
.
Sternberg
,
P. W.
and
Horvitz
,
H. R.
(
1986
).
Pattern formation during vulval development in C. elegans
.
Cell
44
,
761
772
.
Sternberg
,
P. W.
and
Horvitz
,
H. R.
(
1989
).
The combined action of two intercellular signalling pathways specifies three cell fates during vulval induction in C. elegans
.
Cell
58
,
679
693
.
Sulston
,
J. E.
and
Horvitz
,
H. R.
(
1977
).
Postembryonic cell lineages of the nematode Caenorhabditis elegans
.
Dev. Biol
.
56
,
110
156
.
Sulston
,
J. E.
and
Horvitz
,
H. R.
(
1981
).
Abnormal cell lineages in mutants of the nematode Caenorhabditis elegans
.
Dev. Biol
.
82
,
41
55
.
Sulston
,
J. E.
and
White
,
J. G.
(
1980
).
Regulation and cell autonomy during postembryonic development of Caenorhabditis elegans
.
Dev. Biol
.
78
,
577
597
.
Sulston
,
J. E.
,
Schierenberg
,
E.
,
White
,
J. G.
and
Thomson
,
J. N.
(
1983
).
The embryonic cell lineage of the nematode Caenorhabditis elegans
.
Dev. Biol
.
100
,
64
119
.
Sundaram
,
M.
and
Han
,
M.
(
1996
).
Control and integration of cell signaling pathways during C. elegans vulval development
.
BioEssays
18
,
473
480
.
Taya
,
Y.
(
1997
).
RB kinases and RB-binding proteins: new points of view
.
Trends Biochem. Sci
.
22
,
14
17
.
Tautz
,
D.
,
Lehmann
,
R.
,
Schnürch
,
H.
,
Schuh
,
R.
,
Seifert
,
E.
,
Kienlin
,
A.
,
Jones
,
K.
and
Jäckle
,
H.
(
1987
).
Finger protein of novel structure encoded by hunchback, a second member of the gap class of Drosophila segmentation genes
.
Nature
327
,
383
389
.
Thomas
,
J. H.
,
Stern
,
M. J.
and
Horvitz
,
H. R.
(
1990
).
Cell interactions coordinate the development of the Caenorhabditis elegans egg-laying system
.
Cell
62
,
1041
1052
.
White
,
J. G.
,
Southgate
,
E.
,
Thomson
,
J. N.
and
Brenner
,
S.
(
1986
).
The structure of the nervous system of Caenorhabditis elegans
.
Phil. trans. R. Soc. Lond. B
314
,
1
340
.
Wilson
,
R.
,
Ainscough
,
R.
,
Anderson
,
K.
,
Baynes
,
C.
,
Berks
,
M.
,
Bonfield
,
J.
,
Burton
,
J.
,
Connell
,
M.
,
Copsey
,
T.
,
Cooper
,
J.
and
others
(
1994
).
2.2 Mb of contiguous nucleotide sequence from chromosome III of C. elegans
.
Nature
368
,
32
38
.
Yandell
,
M. D.
,
Edgar
,
L. G.
and
Wood
,
W. B.
(
1994
).
Trimethylpsoralen induces small deletion mutations in Caenorhabditis elegans
.
Proc. Natl. Acad. Sci. USA
91
,
1381
1385
.
Yuan
,
J.
and
Horvitz
,
H. R.
(
1990
).
The Caenorhabditis elegans genes ced-3 and ced-4 act cell autonomously to cause programmed cell death
.
Dev. Biol
.
138
,
33
41
.