Tumor suppressor genes, whose products are required for the control of cell proliferation, have been identified by their mutant phenotype of tissue overgrowth. Here we describe recent work on the molecular identification of tumor suppressor genes that function in two different cell types of the Drosophila larva: the blood cells, and the undifferentiated epithelial cells of developing imaginai discs. Mutations in the aberrant immune responses (air8) gene lead to overproduction and precocious differentiation of blood cells. This gene encodes the Drosophila homolog of human ribosomal protein S6. The mutant phenotype is consistent with a role for S6 in the control of cell proliferation, and is compatible with findings from mammalian cells where alterations in S6 expression and phosphorylation are associated with changes in cell proliferation. Mutations in the dises large (dlg) gene cause neoplastic overgrowth of imaginai discs in the larva. The mutant discs show loss of septate .junctions and of apical-basal cell polarity, and they also lose the ability to differentiate cuticular structures. The d/g protein product (DlgA) is localized at septate junctions between epithelial cells, and cDNA sequencing indicates that the gene product includes a domain with homology to guanylate kinase (GUK). Two mammalian homologs of this gene have been identified, and one of them (PSD-95/SAP90) encodes a component of synaptic densities in the brain: this protein therefore resembles the DlgA protein in being located in a specialized cell junction that functions in information transfer between cells. Mutations in the fat gene cause hyperplastic imaginai disc overgrowth, in which the overgrowing disc tissue retains its epithelial structure and its ability to differentiate. Some of the excess disc tissue is shed as vesicles suggesting a loss of cell adhesion. In support of this hypothesis, the predicted gene product shows homology to cadherins in its extracellular domain. However, the fat protein is much larger than known cadherins. As in human cancer, somatic loss of the normal alleles of tumor suppressor genes can lead to tumor formation in Drosophila; an example of this is provided by the warts (wts) locus. The wts gene was identified by the dramatic overgrowth of mitotic recombination clones that are homozygous for a wts deletion. In these clones the cuticle intrudes between epithelial cells, suggesting an alteration in cell adhesion. The study of these and other tumor suppressor genes in Drosophila is providing new evidence supporting the critical role of cell interactions and specialized apical junctions in controlling epithelial cell proliferation.

Cell proliferation is controlled by cell-autonomous mechanisms involving cyclins and cell division control genes, but il is also affected by factors originating outside the cell including both diffusible signals and contact-mediated signals from adjacent cells. Disruption of proliferation control in cancer cells is often associated with loss of cell polarity, adhesion and junctions, suggesting that the maintenance of these features is important for the operation of the signalling mechanisms controlling proliferation and differentiation.

An important source of information regarding these aspects of cell biology is provided by the identification and analysis of oncogenes and tumor suppressor genes, and their protein products. At least nine human tumor suppressor genes have now been identified by analyzing the changes in somatic cell DNA that occur during tumor development (see Bryant. 1993 for review). Structural analysis of these genes indicates (hat their protein products are involved in cell adhesion, signal transduction, and cell cycle control, although none of their products had been previously recognized in studies of normal cells. By revealing mechanisms and molecules that have not been identified in studies of normal cells, the study of tumor suppressor genes is contributing not only to our understanding of oncogenesis, but also to basic cell biology.

Recently it has become clear that Drosophila can make an important contribution in this area, because simple traditional genetic procedures can be used to identify mutations that cause overgrowth of specific tissues and thus to identify tumor suppressor genes (Gateff. 1978; Gateff. 1982; Bryant. 1993). Furthermore, molecular genetic techniques available lor Drosophila facilitate the identification and characterization of the gene products. Sequence comparison can then be used to identify human homologs as candidate tumor suppressor genes.

Mutations in most of the known Drosophila tumor suppressor genes are recessive lethals that show either hyperplastic or neoplastic overgrowth of specific tissues, followed by death of the animal during the late larval or early pupal stage. The tumor suppressor genes functioning only in the gonads arc exceptional in that the tumors form in the adult and cause sterility but not lethality. Over 50 tumor suppressor genes have been identified; they function in the embryo, in the developing gonads, in the developing brain, in developing imaginai discs (embryonic tissues that develop inside the larva and differentiate into parts of the adult fly during metamorphosis), and in the developing hematopoietic system. Fight of these genes have been cloned and characterized, and three of them show clear homology to human genes, none of which had been recognized previously as a tumor suppressor gene (Bryant. 1993). In this paper we describe recent findings on these three genes.

THE RpS6 GENE, IDENTIFIED BY THE aberrant immune responses (air8) MUTATION

The aberrant immune responses (air8) mutation was recovered in a P-elcment mutagenesis screen that led to the identification of over 20 genes in which mutations lead to melanotic tumor formation (Watson et al., 1991). The mutations arc called aberrant immune response (or air) since genetic alterations rather than foreign challenges lead to the activation of cellular defense mechanisms and the formation of melanotic tumors (Fig. 1). The air mutations fall into two classes; the phenotype of the first class (including air8) includes overgrowth of the hematopoietic organs (lymph glands) indicating that the corresponding genes function in cell proliferation control in that tissue. The second class of air mutations has no observable effect on the lymph glands, but the mutant animals develop melanotic tumors associated with specific tissues. Mutations in this class arc thought to result in “autoimmune responses” against the affected tissues (Watson et al., 1991).

The air8 mutation inhibits growth of most larval organs but causes overgrowth of the lymph glands (Fig. 2: Watson et al., 1992). II also causes overproduction of blood cells from the lymph gland (K.L.W., unpublished data) and the precocious transformation of plasmatocytcs to lamellocytes during larval life, an event that normally occurs during the pupal stage (Fig. 3; Shrestha and Gateff. 1982; Rizki and Rizki. 1984). The lamellocytes aggregate to form large melanotic tumors, and death occurs during the late larval stage (Figs 1. 3; Watson et al., 1991; Watson et al., 1992). Transplantation of air8 hemocytes into the abdomens of wild-type females leads to bloating, melanotic tumor formation and significantly more host lethality than wildtype hemocyte injections (K.L.W., unpublished data). Also, explained air8 hemocytes show’ indefinite proliferation in vitro (K.I.,W. and I). Peel, unpublished data). In contrast to its effect on the hematopoietic system, the air8 mutation appears to block cell growth in other tissues. Mitotic recombination clones of air8 mutant cells fail to survive both in the germline and developing imaginai disc cells, indicating a requirement of the normal gene product for cell survival and/or proliferation in these tissues (Watson et al., 1992).

Fig. 1.

Hemizygous air8 third-instar larvae showing the melanotic tumor phenotype. From Watson cl al. (1992).

Fig. 1.

Hemizygous air8 third-instar larvae showing the melanotic tumor phenotype. From Watson cl al. (1992).

Fig. 2.

Brain (B). imaginai discs (1), dorsal vessel (DV) and lymph gland (LG) of third-instar larvae. (A) air 8. Four pairs of hypertrophied lymph-gland lobes line the dorsal vessel in the anterior region and pericardial cells (PC) are found posterior to the lymph glands. (B) Corresponding parts of a wild-type larva al the same magnification. From Watson el al. (1991).

Fig. 2.

Brain (B). imaginai discs (1), dorsal vessel (DV) and lymph gland (LG) of third-instar larvae. (A) air 8. Four pairs of hypertrophied lymph-gland lobes line the dorsal vessel in the anterior region and pericardial cells (PC) are found posterior to the lymph glands. (B) Corresponding parts of a wild-type larva al the same magnification. From Watson el al. (1991).

Fig. 3.

(A) Phase contrast view of circulating hemocytes in the hemolymph of a wild-type third-instar larva. P, plasmatocylcs. Bar. 50 μm. (B) Phase contrast view of circulating hemocytes of an air8 larva including a plasmalocyle (P), lamellocyle (L), and a podocyte (Po), the latter representing an intermediate in the transformation of a plasmalocyle to a lamellocyle. Bar. 50 μm. (C) Electron micrograph showing hemocytes (H) in a lymph gland lobe of a wild-type larva. Bar. 1.5 μm. (D) Electron micrograph showing an air8 lymph gland lobe. At least three lamellocytcs. L, are encapsulating each other and many podocytes (Po) in transition to lamellocytcs can be seen. Bar. 1.5 μm. Modified from Watson et al. (1992).

Fig. 3.

(A) Phase contrast view of circulating hemocytes in the hemolymph of a wild-type third-instar larva. P, plasmatocylcs. Bar. 50 μm. (B) Phase contrast view of circulating hemocytes of an air8 larva including a plasmalocyle (P), lamellocyle (L), and a podocyte (Po), the latter representing an intermediate in the transformation of a plasmalocyle to a lamellocyle. Bar. 50 μm. (C) Electron micrograph showing hemocytes (H) in a lymph gland lobe of a wild-type larva. Bar. 1.5 μm. (D) Electron micrograph showing an air8 lymph gland lobe. At least three lamellocytcs. L, are encapsulating each other and many podocytes (Po) in transition to lamellocytcs can be seen. Bar. 1.5 μm. Modified from Watson et al. (1992).

The gene identified by the air8 mutation was the first tumor suppressor functioning in the Drosophila hematopoietic system to be characterized at the molecular level. Sequencing of air8 cDNAs from an adult female library revealed an open reading frame encoding 248 amino acids with 75% identity and 95% similarity to both human and rat ribosomal protein S6 (Fig. 4A: Watson et al., 1992). Two different P-element insertions from different melanotic tumor strains (air8 and WGI288) map to the 5’ transcribed but untranslated region of the Drosophila RpS6 gene. On northern blots, the I kb RpS6 transcript is barely detectable in air8 mutant larvae yet this transcript is abundant in wild-type animals throughout development and in air8 revenant animals that were produced by excision of the P-element (Watson et al., 1992).

Fig. 4.

(A) Sequence alignments of predicted Drosophila and human S6 ribosomal protein sequences. Alignments were generated using the IAS TA program (Pearson and Lipman. 1990). The four copies of the ten amino-acid motif that may serve as a nuclear localization signal (Chan and Wool. 1988) arc underlined. Residues that are subject to phosphorylation in the human protein arc shaded (Wettenhall et al., 1988). The consensus recognition sequence for the mitogen-activated 70K S6 kinase (Ferrari et al., 1991; flotow and Thomas. 1992) is boxed. :, identical residues; ., conservative substitutions. From Watson el al. (1992). (B) Schematic diagram of the genomic arrangement of the Drosophila RpS6 gene and its two downstream repeats 3RI and 3R2 (K. L. W. and P. J. B. unpublished data). (C) Sequence alignment of the two identical predicted repeats (Rep) with the predicted S6 amino acid sequence over the region of overlap. Conventions as in A. Note that the C-terminal tail in the repeats contains only four serine residues, compared with seven in Drosophila and human S6.

Fig. 4.

(A) Sequence alignments of predicted Drosophila and human S6 ribosomal protein sequences. Alignments were generated using the IAS TA program (Pearson and Lipman. 1990). The four copies of the ten amino-acid motif that may serve as a nuclear localization signal (Chan and Wool. 1988) arc underlined. Residues that are subject to phosphorylation in the human protein arc shaded (Wettenhall et al., 1988). The consensus recognition sequence for the mitogen-activated 70K S6 kinase (Ferrari et al., 1991; flotow and Thomas. 1992) is boxed. :, identical residues; ., conservative substitutions. From Watson el al. (1992). (B) Schematic diagram of the genomic arrangement of the Drosophila RpS6 gene and its two downstream repeats 3RI and 3R2 (K. L. W. and P. J. B. unpublished data). (C) Sequence alignment of the two identical predicted repeats (Rep) with the predicted S6 amino acid sequence over the region of overlap. Conventions as in A. Note that the C-terminal tail in the repeats contains only four serine residues, compared with seven in Drosophila and human S6.

Although mammalian S6 has not been recognized as a tumor suppressor, there are several indications that it may play a role in cell proliferation control and tumorigenesis. The human rpS6 gene has been mapped to 9p21, a region in which genetic alterations arc associated with various cancers including acute lymphoblastic leukemia and acute myelogenous leukemia (Antoine and Fried, 1992). In addition, S6 is overexpressed in a variety of tumors such as human colon carcinoma and adenomatous polyps (Pogue-Geile et al., 1991; Barnard et al., 1992). Most intriguingly, the S6 protein contains a cluster of serine residues in the carboxy terminus that shows developmentally regulated and mitogen-induced phosphorylation (Traugh and Pendergast. 1986; Sturgill and Wu. 1991); it is not yet known whether the Drosophila protein is similarly modified. In quiescent mammalian cells, most ribosomes exist as inactive monosomes and S6 is not phosphorylated (Traugh and Pendergast. 1986). However, shortly after cells arc stimulated to proliferate by treatment with scrum growth factors, insulin, tumor promoling agents, or transforming viruses, protein synthesis increases as a result of higher rates of translation initiation (Traugh and Pendergast. 1986: Ballou et al., 1988). S6 becomes multiply phosphorylated in an orderly manner aller mitogen stimulation and during development, and ribosomal subunits containing phosphorylated S6 become preferentially incorporated into polysomes (Traugh and Pendergast. 1986: Heinze et al., 1988). It has been suggested that phosphorylation of specific residues in S6 may control the state of cells by regulating the selective translation of mRNAs (Mutoh et al., 1992). The S6 protein resides in the mRNA binding domain of the 40S subunit where it might be expected that any subtle conformational and/or charge change induced by S6 phosphorylation could have a significant effect on ribosome function (Ballou et al., 1988: Traugh and Pendergast. 1986).

Two distinct families of S6 kinases, pp90rsk and pp70S6k, have been shown in vivo and in vitro to phosphorylate S6 directly via two distinct intracellular signalling pathways (Erikson. 1991). In one of these pathways. pp90rsk and S6 arc downstream targets in a serine/threonine phosphorylation cascade that has been identified in a variety of systems (Lange-Carteret al., 1993). Recently. pp70S6k function was shown to be necessary for cell cycle progression through GI and its principal target. S6, is phosphorylated during this transition (Lane et al., 1993). These results provide additional support for the hypothesis that S6 phosphorylation may regulate cell proliferation.

The growth inhibition seen in most organs of air8 animals, and the failure of mutant clones to survive in the germ line and imaginai discs, is consistent with a defect in protein synthesis. However, it is surprising to find that tumor formation in the hematopoietic system appears to occur in the absence of the RpS6 transcript and presumably S6 protein. One possible explanation is that a different protein is able to substitute for S6 in the ribosomes of hematopoietic cells, and this hypothesis is supported by the finding that the genomic DNA immediately Hanking the 3′ end of the RpS6 coding region contains two tandem repeals (3RI and 3R2) of a sequence closely related to RpS6 intron 2, exon 3 and 3′ flanking DNA (Fig. 4B). The existence of these repeats can be explained by two unequal crossover events, leading first to a duplication of part of RpS6. followed by duplication of the copy (K.L.W., unpubl.). The intron 2/exon 3 boundaries, the open reading frame and the putative polyadenylation signals are retained in 3RI and 3R2, suggesting that the repeals arc transcribed. Conceptual translation of the repeals predicts peptides showing 62.3% amino acid identity plus an additional 25.4% similarity with S6 (Fig. 4C). However, there are interesting differences between the repeats and S6 including deletions and residue changes that predict S6 isoforms with altered functions. In particular, there are residue alterations in the putative nuclear localization signals and loss of three of the C-terminal serine residues that may be phosphorylated in S6 (Fig. 4C). Preliminary results suggest that the RpS6 repeats arc transcribed (K.L.W.. unpublished data) and if these transcripts are translated, the alternative isoforms might be capable of assembling into ribosomes. Tissue-specific isoforms of ribosomal proteins have not been reported, but if they exist they could explain the tissue-specific phenotype caused by mutations in the RpS6 gene.

THE lethal(1 )discs-large (dlg) LOCUS

Recessive lethal mutations in the dlg gene cause neoplastic overgrowth of the imaginai discs in the larva, followed by death of the animal in the early pupal stage (Woods and Bryant. 1989, 1991). The overgrowing imaginai (.lises, which are normally single-layered epithelia, become disorganized masses and lose the ability to develop into adult parts even after transplantation into normal hosts. Strong alleles disrupt epithelial structure over the entire dise whereas with weaker alleles only part of the tissue loses epithelial structure and overgrows (Figs 5, 6A). When the entire disc is affected, it can grow to at least three times the normal cell number during the extended larval period (Woods and Bryant, 1989). Normal imaginai disc cells arc highly polarized epithelial cells connected by junctional complexes including adherens and septate junctions, but in the imaginai discs of larvae carrying a dlg loss-of-function mutation, apical-basal polarity is almost completely lost and the septate junctions arc missing (D. F. W. and P. J. B., unpublished data). The brain and lymph gland also overgrow in the mutant and the larvae become bloated, but the remainder of the larval organs appear normal.

Fig. 5.

Late third-instar wing imaginai discs. (A) Wild-type. Note the Hal shape. (B) dlgXl-2, showing loss of epithelial structure over part of the disc (arrow) but normal structure elsewhere. (C) dlgXl-2 wing disc with only a small area of normal epithelial structure (arrow). From Bryant (1987).

Fig. 5.

Late third-instar wing imaginai discs. (A) Wild-type. Note the Hal shape. (B) dlgXl-2, showing loss of epithelial structure over part of the disc (arrow) but normal structure elsewhere. (C) dlgXl-2 wing disc with only a small area of normal epithelial structure (arrow). From Bryant (1987).

Fig. 6.

(A) A section of a dlgXl-2 -wing disc showing partial loss of epithelial structure (see Fig. 5b). The arrow points to a region of (he disc where the cells reniait) columnar, while the overgrowing region is highly folded and the cells are cuboidal. Bar. 5 μm. (B) Confocal section of a wild-type late third-instar wing disc stained with antibodies against dlg (green staining) and against phosphotyrosine (PY; red staining). I his optical section shows the apical surfaces of two cell layers both facing a luminal space. The dlg protein is restricted to the septate junctions and PY is highly enriched in the adherens junctions. Bar, 1 μm.

Fig. 6.

(A) A section of a dlgXl-2 -wing disc showing partial loss of epithelial structure (see Fig. 5b). The arrow points to a region of (he disc where the cells reniait) columnar, while the overgrowing region is highly folded and the cells are cuboidal. Bar. 5 μm. (B) Confocal section of a wild-type late third-instar wing disc stained with antibodies against dlg (green staining) and against phosphotyrosine (PY; red staining). I his optical section shows the apical surfaces of two cell layers both facing a luminal space. The dlg protein is restricted to the septate junctions and PY is highly enriched in the adherens junctions. Bar, 1 μm.

The dlg gene encodes a 960-amino acid protein (DlgA) that is expressed in epithelial cells, as well as other tissues including the hematopoietic organs and the nervous system, throughout development (Woods and Bryant. 1989; Woods and Bryant. 1991). Immunocytochemical analysis shows that in epithelial cells the DlgA protein is localized in a subapical belt of the lateral cell membrane (Fig. 6B), at the position of the septate junction (Woods and Bryant. 1991). Thus the DlgA protein appears to be localized in septate junctions and is required for their formation. The role of septate junctions is unknown, although they have often been considered the invertebrate equivalent of vertebrate tight junctions (Noirot-Timothee and Noirot. 1980). Tight junctions arc known to provide a transcpithelial barrier to diffusion of small molecules and to prevent diffusion of membrane components between the apical and basolateral membrane domains (Schneeberger and Lynch. 1992). Septate junctions also provide a barrier to transcpithelial diffusion of lanthanum (Green and Bergquist. 1982) and appear to separate membrane domains as indicated by the behavior of lipophilic markers (Wood, 1990).

Sequence comparisons based on the dlg cDNA sequence indicate that the encoded protein includes a series of modules each of which is also found in other proteins (Fig. 7). These domains and their possible functions are as follows:

Fig. 7.

Comparison between the predicted organization of the DlgA protein and that of its yeast and mammalian homologs. The alignments were identified by the standard BLAST algorithm (Altschul et al., 1990). Percentages shown refer to amino-acid identity with the corresponding domain of DlgA. Yeast GUK from Berger et al. (1989): PSD-95 from Cho et al. (1992); SAP90 from Kistneret al. (1993); DlgA from Woods and Bryant (1991); p55 from Ruff et al. (1991). D1-D3. DHR repeats: SH3. sre-homology 3 domains.

Fig. 7.

Comparison between the predicted organization of the DlgA protein and that of its yeast and mammalian homologs. The alignments were identified by the standard BLAST algorithm (Altschul et al., 1990). Percentages shown refer to amino-acid identity with the corresponding domain of DlgA. Yeast GUK from Berger et al. (1989): PSD-95 from Cho et al. (1992); SAP90 from Kistneret al. (1993); DlgA from Woods and Bryant (1991); p55 from Ruff et al. (1991). D1-D3. DHR repeats: SH3. sre-homology 3 domains.

DHR

There are three tandem copies of the DHR motif in the N-terminal half of the predicted DlgA protein (previously reported as a “filamentous” domain based on its sequence similarity to some filamentous proteins: Woods and Bryant. 1991). The motif was named GLGF based on an amino acid sequence present al the N-terminal end of the motif in DlgA and one of its mammalian homologs (Cho et al., 1992). However, since the GLGF sequence itself is absent from most other copies of this motif, we suggest the alternative name DHR (Discs large Homologous Region). A single copy of this domain is present in nitric oxide synthase (NOS; Cho et al., 1992), in a human cytosolic tyrosine phosphatase (Gu et al., 1991), and in the cytoplasmic domain of a human brain transmembrane protein encoded by a gene potentially involved in Friedreich ataxia (Duelos et al., 1993). Most of these proteins are localized al the cell surface, probably in association with the cortical cytoskeleton. Since DlgA and at least two of its mammalian homologs (see below) are also membrane associated, we propose that the DHR repeal is involved in directing this localization.

OPA

The OPA trinucleotide repeat is found in many genes that are developmentally regulated or show tissue-specific expression, in a variety of organisms from yeast to mammals (Wharton et al., 1985). It is translated in several reading frames to give tracts that are rich in specific amino acids. In dlii it is predicted to give rise to tracts of serines and alanines.

SH3

There is a single copy of the SH3 (src-hoinology region 3; Musacchio et al., 1992) domain in the predicted DlgA protein. Mutations in the SH3 domain of v-src can restore transforming activity in otherwise defective mutants (Dezélée et al., 1992), indicating that this domain is important for regulation of protein activity. SH3 binds a ten-amino acid proline-rich peptide motif that has been found in several signal transduction proteins (Ren et al., 1993) including one that is similar to the GTPase activating protein (GAP) associated with the ras-relatcd protein rho (Cicchetti et al., 1992). Other binding targets for SH3 include the GTP-hydrolyzing motor protein dynamin (Booker et al., 1993), and guanine nucleotide exchange factors (Olivier et al., 1993; Rozakis-Adcock et al., 1993).

PEST

The PEST domain was originally described in proteins that had a high rate of degradation in the cell and was therefore interpreted as a signal for rapid turnover (Rogers et al., 1986). However, this interpretation is open to question since the Notch protein, which contains both PEST and OPA sequences, is apparently stable for several cell cycles (Kidd et al., 1989).

GUK

The DlgA protein contains an approximately 179 aminoacid region al the carboxyl terminus with significant sequence similarity lo the soluble enzyme guanylate kinase (GUK) of yeast (Berger et al., 1989; 36% identity) and pig (Zschocke et al., 1993; 35% identity) (Fig. 8). This enzyme transfers a phosphate group from ATP lo GMP, converting it to GDP. If the DlgA protein has GUK activity, it could regulate the production of guanine nucleotides that act as messenger molecules within the cell (Woods and Bryant. 1991). The DlgA GUK domain shows conservation or conservative substitution of all of the amino acids that have been shown to interact with the GMP substrate as well as the B motif that is thought to bind the Mg++ cofactor in the yeast enzyme (Fig. 8; Stehle and Schulz. 1990, 1992).

Fig. 8.

Alignment of the amino acid sequences of guanylate kinases with the GUK domains of p55. Dlg-A and Dlg-R proteins. The alignment was generated using the multiple alignment program OPTAL (Gorbalenya and Koonin, 1989). The consensus includes invariant amino acid residues (uppercase), and residues found in all but one sequence (lowercase); U -a bulky hydrophobic residue (I. L. V. M. F. Y, W); X -any residue. A and B motifs (see text) are boxed. Amino acid residues implicated in specific molecular interactions in yeast GUK (Stehle and Schulz. 1990. 1992) arc indicated as follows: *. binding to phosphate; !. binding to guanine ring; +. binding to magnesium; these residues and their counterparts in the other sequences are shaded. D.m. -Drosophila melanogaster. C.e. -Cacnorhahditis dedans, E.c. -Escherichia coli, S.c. -Saccharoinyces cerevisiae, H.s. -Homo sapiens. R.n. -Ranns norvégiens, S.s. -Sus seroja, VV -Vaccinia virus (In Vaccinia virus the GUK-related sequences are found in two distinct reading frames, shown by a slash, and the enzyme is probably not functional). Arg-295 of p55 was originally reported as Pro, but this was due to a sequencing error (A. Husain-Chishti. personal communication). Irom Koonin el al. (1992) with the addition of S.s. (Pig) GUK from Zschocke el al. (1993).

Fig. 8.

Alignment of the amino acid sequences of guanylate kinases with the GUK domains of p55. Dlg-A and Dlg-R proteins. The alignment was generated using the multiple alignment program OPTAL (Gorbalenya and Koonin, 1989). The consensus includes invariant amino acid residues (uppercase), and residues found in all but one sequence (lowercase); U -a bulky hydrophobic residue (I. L. V. M. F. Y, W); X -any residue. A and B motifs (see text) are boxed. Amino acid residues implicated in specific molecular interactions in yeast GUK (Stehle and Schulz. 1990. 1992) arc indicated as follows: *. binding to phosphate; !. binding to guanine ring; +. binding to magnesium; these residues and their counterparts in the other sequences are shaded. D.m. -Drosophila melanogaster. C.e. -Cacnorhahditis dedans, E.c. -Escherichia coli, S.c. -Saccharoinyces cerevisiae, H.s. -Homo sapiens. R.n. -Ranns norvégiens, S.s. -Sus seroja, VV -Vaccinia virus (In Vaccinia virus the GUK-related sequences are found in two distinct reading frames, shown by a slash, and the enzyme is probably not functional). Arg-295 of p55 was originally reported as Pro, but this was due to a sequencing error (A. Husain-Chishti. personal communication). Irom Koonin el al. (1992) with the addition of S.s. (Pig) GUK from Zschocke el al. (1993).

However, in DlgA the A motif GxxGxGK. which in the yeast enzyme binds the ATP phosphate donor, is interrupted by a three amino-acid deletion (Koonin et al., 1992). This deficiency in the ATP-binding site (a domain that is otherwise highly conserved in a wide variety of kinases) suggests that the DlgA protein may have lost the ability to bind ATP and may have acquired a novel function requiring GMP binding. Whatever the exact catalytic function of the DlgA protein, its potential involvement in guanine nucleotide metabolism or regulation suggests a relationship to signal transduction mechanisms known lo control cell proliferation, such as those mediated by pp2lras (Woods and Bryant. 1991).

The dlg gene identifies a family of related proteins -the MAGUKs

Recent findings indicate that the dlg gene has homologs in mammals and that there is at least one other related gene in

Drosophila

These genes all encode membrane-associated guanylate kinase homologs, which we abbreviate MAGUKs. Each MAGUK consists of three distinct domains (Fig. 7). (1) One or three copies of DHR. (2) an SH3 domain and (3) the guanylate kinase-like domain (GUK). The OPA and PEST domains present in the DlgA protein are not present in the other MAGUKs identified so far.

Mammalian PSD-95/SAP90

The first mammalian MAGUK was identified as a partial cDNA sequence (1779) from a collection of cDNAs made from human brain RNA (Adams et al., 1992). The partial sequence shows that 1779 is the human homolog of a gene recently identified in the rat (hat encodes a component of synaptic junctions (PSD-95/SAP90) in the adult brain (Cho et al., 1992; Kisiner et al., 1993). The predicted PSD-95/SAP90 protein shows strong homology (about 67% identity) to DlgA throughout its length and we therefore refer to it as a Dlg-R (discs large-related) protein. Unlike DlgA, which is expressed in a variety of tissues including the nervous system. PSD-95/SAP90 appears to be expressed only in the nervous system in postnatal rats. However, it is intriguing to note that, like the product of dlg PSD-95/SAP90 is localized in a specialized junction, the synaptic junction, that is involved in information transfer between cells. Also like the product of dlg, the PSD-95/SAP90 sequence shows strong conservation of the GMP-binding site but has the same deficiency in the putative ATP-binding site (Fig. 8), suggesting that its catalytic activity, if any, is similar to that of the Drosophila protein (Koonin et al., 1992). We have no information about PSD-95/SAP90 expression in embryonic stages of development, so it is still possible that it is expressed in epithelial cells and is involved in cell interactions controlling proliferation in these early stages.

Human p55

A second human MAGUK is the heavily palmitoylated protein p55, which was purified from red blood cell membrane ghosts and shows striking similarity to (he DlgA protein in the GUK region, the SH3 region, and one DHR repeal (Ruff et al., 1991; Bryant and Woods. 1992). p55 is expressed in most cell types in addition to blood cells and thus could be involved in cell growth control. We have recently cloned a Drosophila p55 homolog (D. F. W and P. J. B.. unpublished data) and the partial predicted amino acid sequence shows approximately 60% identity to that of human p55. Thus, there are two types of MAGUK in both mammals and Drosophila, suggesting that these proteins may have distinct and conserved functions. The ATP-binding site in p55 does not contain a deficiency like that seen in DlgA. but it does show a K lo R substitution (Fig. 8) that might affect enzyme activity. The p55 gene is located near the lip of the long arm of the X chromosome, between the genes encoding the blood-clotting protein factor VIII and the enzyme glucose-6-phosphate dehydrogenase (Metzen-berg and Gitschier. 1992). No tumor suppressor genes have been discovered in this chromosome region, but at least 19 genetic diseases are associated with the interval and one of them, dyskeratosis congenita, is associated with precancerous lesions (leukoplakia) and a high risk of cancer (McKay et al., 1991; Marsh et al., 1992).

Proteins interacting with MAGUK proteins

The p55 and PSD-95/SAP90 proteins were purified by sub-cellular fractionation, and in both cases several co-purifying proteins were identified (Table 1). This association may result from specific interactions between the MAGUK protein and the co-purifying molecules, or it may simply be a result of similar physico-chemical properties. The former explanation seems more likely for at least some of these proteins since they remain tightly associated with the MAGUK even after treatment with sarcosyl, a relatively harsh detergent (Cho et al., 1992). Furthermore, some of these proteins have Drosophila homologs that are found, along with DlgA, at the septate junctions (Table 1). These results suggest that the MAGUKs arc integral components of large protein networks that are restricted to specific junctional complexes or other membrane-associated structures in the cell.

Table 1.

Proteins that co-purify with p55 and PSD-9 5/ SAP90, and proteins inimunolocalized at Drosophila septate junctions

Proteins that co-purify with p55 and PSD-9 5/ SAP90, and proteins inimunolocalized at Drosophila septate junctions
Proteins that co-purify with p55 and PSD-9 5/ SAP90, and proteins inimunolocalized at Drosophila septate junctions

Most of the MAGUK-interacting proteins thus far identified appear to be part of the cytoskeletal network, perhaps reflecting the importance of these proteins in maintaining cyloarchilecture. Other associated proteins include cell adhesion molecules (FasIII; Patel et al., 1987) and signal transduction proteins (serine/threoninc kinases; Ruff et al., 1991; Cho et al., 1992). In the case of the synaptic density, where PSD-95/SAP90 is localized, the kinase activity (type II Ca2+/calmodulin-dependcnt protein kinase) has been shown to be necessary for induction of long-term potentiation of synaptic transmission (Malinow et al., 1989). There are probably additional proteins associated with the MAGUKs, including some expected to interact with the SH3 domain, whose identities and functions remain to be elucidated. Furthermore, if MAGUK proteins arc involved in transduction of extracellular signals, it seems likely that transmembrane proteins will be involved upstream of them in the signalling pathway. The identity of these elements has yet to be discovered.

THE fat LOCUS

In lethal mutations of the fat locus all of the imaginai discs show hyperplastic overgrowth, in that the overgrowing imaginai discs retain their single-layered epithelial structure and their ability to differentiate (Bryant et al., 1988; Mahoney et al., 1991). The gene potentially encodes a 5147-amino acid protein, and it is transcribed in the embryonic ectoderm and larval imaginai discs (Mahoney et al., 1991). The predicted fat gene product is a large transmembrane protein with strong sequence homology to calcium-dependent cell adhesion molecules (cadherins: Fig. 9). However, the product is much larger than typical cadherins, with 34 extracellular domains corresponding to the four domains of typical cadherins. as well as four EGF-like repeats and other cysteine-rich regions in the extracellular domain. The cytoplasmic domain shows no clear homology lo that of typical cadherins. The structure of the predicted protein suggests that cell adhesion, dependent on a cadhcrin-like molecule, is required for the cell communication that controls tissue growth. Parts of the imaginai disc epithelium are shed as vesicles in fat mutants (Bryant et al., 1988). supporting the idea that these mutations partially disrupt cell adhesion.

Fig. 9.

Comparison between the predicted organization of the Drosophila fat protein and a typical vertebrate cadherin (Mahoney et al., 1991). The predicted fat product contains 34 extracellular domains with sequence similarity to the four extracellular domains of typical cadherins. It also contains four epidermal growth factor (EGF) repeats that are absent from typical cadherins. In the predicted fat protein the First cadherin domain occurs immediately after the signal sequence, whereas in typical cadherins the signal sequence is followed by approximately 100 amino acids that arc cleaved off to produce the mature protein. There is no significant homologs between the predicted fat protein and typical cadherins in the cytoplasmic domain. C. carboxy terminus; N, amino terminus; TM, transmembrane domain. After Mahoney et al. (1991).

Fig. 9.

Comparison between the predicted organization of the Drosophila fat protein and a typical vertebrate cadherin (Mahoney et al., 1991). The predicted fat product contains 34 extracellular domains with sequence similarity to the four extracellular domains of typical cadherins. It also contains four epidermal growth factor (EGF) repeats that are absent from typical cadherins. In the predicted fat protein the First cadherin domain occurs immediately after the signal sequence, whereas in typical cadherins the signal sequence is followed by approximately 100 amino acids that arc cleaved off to produce the mature protein. There is no significant homologs between the predicted fat protein and typical cadherins in the cytoplasmic domain. C. carboxy terminus; N, amino terminus; TM, transmembrane domain. After Mahoney et al. (1991).

Cadherins have not typically been considered as tumor suppressor gene products, but recent work is suggesting strongly that the E-cadherin (= uvomorulin, homologous to L-CAM) gene located at I6q22.1 (Nall et al., 1989) may correspond to a tumor suppressor gene identified by loss of heterozygosity in this genetic region in hepatocellular, breast, and prostate carcinomas (Sato et al., 1990; Zhang et al., 1990; (artcr ci al.. 1990; Tsuda ci al., 1990; Field, 1992). E-cadherin expression is lost in head and neck and hepatocellular carcinomas (Schipper et al., 1991; Shimoyama and Hirohashi. 1991). and E-cadherin loss is correlated with invasiveness of carcinoma and epithelial cell lines (Behrens et al., 1989; Frixen et al., 1991). Non-invasive carcinoma and epithelial cell lines become invasive when treated with anli-E-cadherin antibodies (Behrens et al., 1989; Frixen et al., 1991); and invasiveness of carcinoma cell lines can be prevented by transfection with E-cadherin cDNA (Frixen et al., 1991). These results arc highly suggestive of a role for this cell adhesion molecule in tumor or metastasis suppression. The finding that mutations in the fat gene cause overgrowth supports the idea that the cadherin losses reported in cancer cells do, indeed, contribute to the malignant phenotype.

In some cases, loss of heterozygosity for a tumor suppressor gene in individual cells of a heterozygote, caused by experimentally induced mitotic recombination, leads to overgrowing clones of cells in specific tissues. This has been demonstrated for both dlg (Woods and Bryant. 1991) and fat (Mahoney et al., 1991), and the phenomenon raises the possibility of identifying new tumor suppressor genes by directly examining the effects of mutations in mitotic recombination clones. Recent advances in Drosophila genetics make screening mitotic recombination clones very efficient. The screen takes advantage of transgenic flies developed by Golic and Lindquist (1989) in which a yeast target-specific FLP recombinase gene has been placed under the control of a Drosophila heat-shock promoter and integrated into the Drosophila genome. Yeast FLP recombinase target sequences (FRTs) also have been integrated into several proximal sites on Drosophila chromosomes. Heal shocking flies carrying the FLP recombinase construct and homozygous for a FRT site leads to a high frequency of mitotic recombination al the FRT site (Golic. 1991). To use this system for screening, chromosomes arc mutagenized in the parental generation and then made homozygous in mitotic recombination clones of the F1 generation. The phenotype of the mutant clones is then examined, and chromosomes that give overgrowing clones arc recovered by breeding from the affected F1 flies.

We have recovered a deficiency for the warts gene (wts) using this screen. Mitotic recombination clones homozygous for such deficiencies produce spectacular outgrowths from the bod) surface (Fig. 10). The clones are larger than normal and tend to split into several rounded fragments, whereas control clones are intact, but irregular and elongated on both legs and wings. The large size of wts clones indicates that the constituent cells undergo more divisions than normal, and the rounded shape of mutant clones suggests that the division of mutant cells is not preferentially oriented as it appears to be in wild-type imaginai discs (Bryant and Schneiderman. 1969). The texture of the cuticle in wts clones is very unusual -each polygonal cell outline is marked by a thick ridge of cuticle (Fig. 10). The apparent intrusion of cuticle between cells suggests that there may be a failure of adhesion or other abnormality in the apical region of the epithelial cells. The wts gene has been localized lo 100A3-7 on chromosome 3 based on the cytology of a series of overlapping deficiencies that give the II/.V phenotype in mitotic recombination clones (R. W. J and P. J. B., unpublished data).

Fig. 10.

The wts phenotype is shown in mitotic recombination clones (arrows) homozygous for Df(3R)AI77der20, induced by gamma irradiation. (‘Iones arc marked by homozygous recessive Ki+, which produces normal (long, straight) bristles in a background of Kis heterozygous cells that produce short, curved bristles. Polygonal ridges of cuticle surround cells within clones on the leg (A) and on the wing (B), where the unusual rounded shape of the clones is also apparent (R. W. J. and P. J. B., unpublished data).

Fig. 10.

The wts phenotype is shown in mitotic recombination clones (arrows) homozygous for Df(3R)AI77der20, induced by gamma irradiation. (‘Iones arc marked by homozygous recessive Ki+, which produces normal (long, straight) bristles in a background of Kis heterozygous cells that produce short, curved bristles. Polygonal ridges of cuticle surround cells within clones on the leg (A) and on the wing (B), where the unusual rounded shape of the clones is also apparent (R. W. J. and P. J. B., unpublished data).

The mitotic recombination screen is efficient because it allows immediate detection of candidate chromosomes in Ft flies without the necessity of breeding each Ft individual as is required in other screens for lethal mutations. The screen is ideally suited to the identification of tumor suppressor genes that function in imaginai discs.

The preceding examples show’ that the Drosophila model system can be used effectively to identify and characterize tumor suppressor genes that have human homologs. Perhaps not surprisingly, these genes encode proteins at various positions in the cell which might act at different points in the pathways of signalling and signal transduction that control cell proliferation. In the case of epithelial cells, a requirement for cell adhesion in proliferation control is suggested by the overgrowth caused by mutations affecting the fat transmembrane protein and possibly the fat membrane-associated protein. This requirement is consistent with experimental work indicating that contactdependent cell interactions control proliferation in epithelial tissue (Bryant. 1987). The involvement of guanine-nucleotide mediated signal transduction is suggested by the nature of the predicted protein produced by the dlg locus. In the hematopoietic system, little experimental work has been done on the mechanisms controlling cell proliferation and differentiation, but the evidence from the RpS6 gene suggests that phosphorylation cascades and translational control mechanisms will turn out to be important in these cells. The cloning and characterization of more tumor suppressor genes in Drosophila seems likely to provide a wealth of new information on how cell proliferation is controlled by cell signalling in a variety of tissues and organs.

The authors’ research is supported by grants from the National Institutes of Health and the National science Foundation.

Adams
,
M. D.
,
Dubnick
,
M.
,
Kerlavage
,
A. R.
,
Moreno
,
R.
,
Kelley
,
J. M.
,
Utterback
,
T. R.
,
Nagle
,
J. W.
,
Fields
,
C.
and
Venter
,.
J. C
(
1992
).
Sequence identification of 2.375 human brain genes
.
Nature
355
,
632
634
.
Altschul
,
S.F.
,
Gish
,
W.
,
Miller
,
W.
,
Myers
,
E. W.
and
Lipman
,
D. J.
(
1990
).
Basic local alignment search tool
.
J. Malee. Biol
.
215
,
403
410
.
Antoine
,
M.
and
Fried
,
M.
(
1992
).
The organization of the intron containing human S6 ribosomal protein (rpS6) gene and determination of its location at chromosome 9p21
.
Hum. Mol. Genet
.
1
,
565
570
.
Ballou
,
L. M.
, .
Jeno
,
P.
, and
Thomas
,
G.
(
1988
).
Control of S6 phosphorylation during the mitogenic response
.
Adv. Exp Med. Biol
231
,
445
452
.
Barnard
,
G. F.
,
Staniunas
,
R. J.
,
Bao
,
S.
,
Mafune
,
K. -i.
,
Steele
,
G. D.
,
Gollan
,.
J. L
, and
Chen
,
L. B.
(
1992
).
Increased expression of human ribosomal phosphoprotein P0 messenger RNA in hepatocellular carcinoma and colon carcinoma
.
Cancer Res.
52
,
3067
3072
.
Behrens
,.
J.
,
Mareel
,
M
M.
Van Roy
,
F. M
and
Birchmeier
,
W.
(
1989
).
Dissecting tumor cell invasion: epithelial cells acquire invasive properties alter the loss of uvomorulin-mediated cell-cell adhesion
.
J. cell Biol
.
108
,
2435
2447
.
Berger
,
A.
,
Schiltz
,
E.
and
Schulz
,
G. E.
(
1989
)
Guanylate kinase from Saccharomyces cerevisiae. Isolation and characterization, crystallization and preliminary X-ray analysis, amino acid sequence and comparison with adenylate kinases, Eur
.
J. Biochem
.
184
,
433
443
.
Booker
,
G. W.
,
Gout
,
I.
,
Downing
,
A. K.
,
Driscoll
,
P.C.
,
Boyd
,
J.
,
Waterfield
,
M. D
, and
Campbell
,
I.D.
(
1993
).
Solution structure and ligand-binding site of the SH3 domain of the p85α subunit of phosphatidylinositol 3-kinase
.
Cell
73
,
813
822
.
Bryant
,
P. J.
(
1987
).
Experimental and genetic analysis of growth and cell proliferation in Drosophila imaginai discs
.
In Genetic Regulation of Development
(ed.
W. F.
Loomis
), pp.
339
372
.
New york
:
Alan R. Liss
,
Bryant
,
P. J.
,
Huettner
,
B.
,
Held
,
L. I.
, Jr
.,
Ryerse
,
J.
and
Szidonya
,
J.
(
1988
).
Mutations at the fat locus interfere with cell proliferation control and epithelial morphogenesis in Drosophila
.
Dev. Biol
.
129
,
541
554
.
Bryant
,
P. J.
(
1993
).
Towards the cellular functions of tumour suppressors
.
Trends Cell Biol
.
3
,
31
35
.
Bryant
,
P. J.
and
Schneiderman
,
H. A.
(
1969
).
Cell lineage, growth, and determination in the imaginai leg discs of Drosophila tnelanogasler
.
Dev, Biol
.
20
,
263
290
.
Bryant
,
P. J.
and
Woods
,
D. F.
(
1992
).
A major palmitoylated membrane protein of human erythrocytes show s homology to yeast guanylate kinase and to the product of a Drosophila tumor suppressor gene
.
cell
68
,
621
622
.
Carter
,
B. S.
,
Ewing
,
C. M.
Ward
,
W. S.
,
Treiger
,
B. F.
,
Aalders
,
T. W.
,
Schalken
,
J. A.
,
Epstein
,
J. I.
, and
Isaacs
,
W. B.
(
1990
).
Allelic loss of chromosomes 16q and 10q in human prostate cancer
.
Proc. Natl. Acad. Sei. USA
87
,
8751
8755
.
Chan
,
y.-L.
and
Wool
,
I. G.
(
1988
).
The primary structure of rat ribosomal protein S6
.
J. Biol. Chem
.
263
,
2891
2896
.
Cho
,
K.-O.
,
Hunt
,
C. A.
, and
Kennedy
,
M B.
(
1992
).
The rat brain postsynaptic density fraction contains a homolog of the Drosophila discslarge tumor suppressor protein
.
Neuron
9
,
929
942
.
Cicchetti
,
P.
,
Mayer
,
B. J.
,
Thiel
,
G.
and
Baltimore
,
D.
(
1992
).
Identification of a protein that binds to the SH3 region of Abl and is similar to Bcr and GAP-rho
.
Science
257
,
803
806
.
Dezéléc
.
P.
,
Barnier
,
J. V.
,
Hampe
,
A.
,
Laugier
,
D.
,
Marx
,
M.
Galibert
,
F.
and
Calothy
,
G.
(
1992
).
Small deletion in v-scr SH3 domain of a transformation defective mutant of Rous sarcoma virus restores wild type transforming properties
.
Virol
.
189
,
556
567
.
Duelos
,
F.
,
Boschert
,
U.
,
Sirugo
,
G.
,
Mandel
,
J. L.
Hen
,
R.
and
Koenig
,
M.
(
1993
).
Gene in the region of the Friedreich ataxia locus encodes a putative transmembrane protein expressed in the nervous system
.
Proc. Natl. Acad. Sei. U. S. A
.
90
.
109
113
.
Erikson
.
R. L.
(
1991
).
Structure, expression, and regulation of protein kinases involved in the phosphorylation of ribosomal protein S6
.
J. Biol. Chem
.
266
,
6007
6010
.
Ferrari
,
S.
,
Bandi
,
H. R.
,
Hofsteenge
,.
J.
,
Bussian
,
B. M
and
Thomas
,
G.
(
1991
).
Mitogen-activated 70K S6 kinase. Identification of in vitro 40 S ribosomal S6 phosphorylation sites
.
J. Biol Chem
.
266
,
22770
22775
.
Field
,
J. K.
(
1992
).
Oncogenes and tumour-suppressor genes in squamous cell carcinoma of the head and neck
.
Eur.J. Cancer /B
.
28B
.
67
76
.
Flotow
,
H.
and
Thomas
,
G.
(
1992
).
Substrate recognition determinants of the mitogen-activated 70K S6 kinase from rat liver
.
J. Biol Chem
.
267
,
3074
3078
.
Frixen
,
L. H
,
Behrens
,
J.
,
Sachs
,
M
,
Eberle
,
G.
,
Voss
,
B.
,
Warda
,
A.
,
Lochner
,
D
, and
Birchmeier
,
W.
(
1991
).
I -cadherin-mediated cell-cell adhesion prevents invasiveness of human carcinoma cells
.
J. Cell Biol
.
113
,
173
185
.
Gateff
.
E.
(
1978
).
Malignant neoplasms of genetic origin in Drosophila melanogaster
.
Science
200
,
1448
1459
.
Gateff
,
E.
(
1982
).
Cancer, genes, and development: the Drosophila case
.
Adv. Cancer Res
.
37
,
33
74
.
Golic
,
K. G.
(
1991
).
Site-specific recombination between homologous chromosomes in Drosophila
.
Science
252
,
958
961
.
Golic
,
K. G.
and
Lindquist
,
S.
(
1989
).
The FLP recombinase of yeast catalyzes site-specific recombination in the Drosophila genome
.
Cell
59
,
499
509
.
Gorbalenya
,
A. E.
and
Koonin
,
E. V.
(
1989
)
Viral proteins containing the purine NTP-binding sequence pattern
.
Nucl. Acids Res
17
,
8413
8440
.
Green
,
C. R.
and
Bergquist
,
P. R.
(
1982
).
Phylogenetic relationships within the Invertebrata in relation to the structure of septate junctions and the development of ‘occluding’ junctional types
.
J. Cell Sci
.
53
,
279
305
.
Gu
,
M X.
,
york
,.
J. D.
,
Warshawsky
,
I.
and
Majerus
,
P. W
(
1991
).
Identification, cloning, and expression of a cytosolic megakaryocyte protein-tyrosine-phosphatase with sequence homology to cytoskeletal protein 4.1
.
Proc Natl. Acad. Sci. USA
88
,
5867
5871
.
Heinze
,
H.
Arnold
,
H H.
Fischer
,
D.
and
Kruppa
,.
J.
(
1988
).
The primary structure of the human ribosomal protein S6 derived from a cloned cDNA
.
J. Biol. Chem
.
263
,
4139
4144
.
Kidd
,
S.
,
Baylies
,
M K.
,
Gasic
,
G. P.
, and
young
,
M. W.
(
1989
).
Structure and distribution of the Notch protein in developing Drosophila
.
Genes Dev
.
3
,
1113
1129
.
Kistner
,
U.
,
Wenzel
,
B. M.
veh
.
R. W.
,
Cases-Langhoff
,
C.
,
Garner
,
A. M.
Appeltauer
,
U
,
Voss
,
B.
,
Gundelfinger
E. D.
and
Garner
,
C. C.
(
1993
).
SAP90. a rat presynaptic protein related to the product of the Drosophila tumor suppressor gene dig-A
J. Biol. Chem
.
268
,
4580
4583
.
Koonin
,
E. V.
,
Woods
,
D. F.
and
Bryant
,
P. J.
(
1992
).
dlg-R proteins: modified guanylate kinases
.
Nature Genet.
2
,
256
257
.
Lane
,
H. A.
,
Fernandez
,
A.
,
Lamb
,
N. J. C.
and
Thomas
,
G.
(
1993
).
p70s6k function is essential for G1 progression
.
Nature
363
,
170
172
.
Lange-Carter
,
C. A.
,
Pleiman
,
C. M.
Gardner
,
A. M.
Blunter
,
K. J.
, and
Johnson
,
G. L.
(
1993
)
A divergence in the MAP kinase regulatory network defined by MEK kinase and Ral
.
Science
260
,
315
319
.
Mahoney
,
P. A.
,
Weber
,
U.
Onofrechuk
,
P.
,
Biessmann
,
H.
Bryant
,
P.J.
and
Goodman
,
C. S.
(
1991
)
The fat tumor suppressor gene in Drosophila encodes a novel member of the cadherin gene superfamily
.
Cell
67
,
853
868
.
Malinow
,
R.
,
Schulman
,
H.
and
Tsien
,
R. W.
(
1989
).
Inhibition of postsynaptic PKC or CaMKII blocks induction but not expression of LTP
.
Science
245
,
862
866
.
Marsh
,
J. C. W.
,
Will
,
A. J.
,
Hows
,
J. M.
Sartori
,
P.
,
Darbyshire
,
P. J.
,
Williamson
,
P.J.
,
Oscier
,
D. G.
,
Dexter
,
T. M
and
Testa
,
G.
(
1992
)
“Stem cell” origin of the hematopoietic defect in dyskeratosis congenita
.
Blood
79
,
3138
3144
.
McKay
,
G. S.
,
Ogden
,
G. R.
and
Chisholm
,
D. M
(
1991
).
Lingual hyperkeratosis in dyskeratosis congenita: ultrastructural findings
.
J. Oral. Pathol. Med
.
20
,
196
199
.
Metzenberg
,
A.
and
Gitschier
,
J.
(
1992
).
The gene encoding the palmitoylated erythrocyte membrane protein. p55. originates at the CpG island 3′ to the factor VIII gene
.
Hum. Malee. Genet
.
1
,
97
101
.
Musacchio
,
A.
,
Gibson
,
T.
,
Lehto
,
V
,-P. and
Saraste
,
M.
(
1992
).
SH3 An abundant protein domain in search of a function
.
J FEBS Lett
.
307
,
55
61
.
Mutoh
,
T.
,
Rudkin
,
B. B.
and
Guroff
,
G.
(
1992
).
Differential responses of the phosphorylation of ribosomal protein S6 to nerve growth factor and epidermal growth factor in PC12 cells
.
J. Neuroehem
.
58
(
1
),
175
185
.
Natt
,
E.
,
Magenis
,
R. E.
,
Zimmer
,
J.
,
Mansouri
,
A.
and
Scherer
,
G.
(
1989
).
Regional assignment of the human loci for uvomorulin (UVO) and chymotrypsinogen B (CTRB) with the help of two overlapping deletions on the long arm of chromosome 16
.
Cvtogenet. Cell Genet
.
50
,
145
148
.
Noirot-Timothee
,
C.
and
Noirot
,
C.
(
1980
).
Septate and scalarifonn junctions in arthropods
.
Im Rev. Cytol
.
63
,
97
140
.
Olivier
,
J. P.
,
Raabe
,
T.
,
Henkemeyer
,
M.
Dickson
,
B.
,
Mbamahi
,
G.
,
Margolis
,
B.
,
Schlessinger
,
J.
,
Halen
,
E.
and
Pawson
,
T.
(
1993
).
A Drosophila SH2-SH3 adaptor protein implicated in coupling the sevenless tyrosine kinase to an activator of Ras guanine nucleotide exchange
.
Sos. Cell
73
,
179
191
.
Patel
,
N. H.
Snow
,
P. M
and
Goodman
,
C. S.
(
1987
).
Characterization and cloning of faseielin III: a glycoprotein expressed on a subset of neurons and axon pathways in Drosophila
.
Cell
48
,
975
988
.
Pearson
,
W. R.
and
Lipman
,
D. J.
(
1990
).
Improved tools for biological sequence comparison
.
Proc. Natl. Acad. Sei. USA
85
,
2444
2448
.
Pogue-Geile
,
K.
,
Griser
,
J. R.
,
Shu
,
M.
Miller
,
C.
,
Wool
,
I. G.
,
Meisler
,
A. I.
and
Pipas
,
J. M
(
1991
).
Ribosomal protein genes are overexpressed in human colorectal cancer: isolation of a cDNA clone encoding the human S3 ribosomal protein
.
Mol. Cell Biol
.
11
,
3842
3849
.
Ren
,
R.
,
Mayer
,
B. J.
,
Cicchetti
,
P.
and
Baltimore
,
D.
(
1993
).
Identification of a ten-amino acid proline-rich SH3 binding sue
.
Science
259
,
1157
1161
.
Rizki
,
T. M.
and
Rizki
,
R. M.
(
1984
).
The cellular defense system of Drosophila nielanoyaster
.
In Insect Ultraslructure
(ed.
R. C.
King
), pp.
579
603
.
Plenum Publishing Corporation
.
Rogers
,
S.
,
Wells
,
R.
and
Rechsteincr
,
M.
(
1986
).
Amino acid sequences common to rapidly degraded proteins: the PEST hypothesis
.
Science
234
,
364
36X
.
Rozakis-Adcock
,
M.
,
Fernley
,
R.
,
Wade
,
J.
Pawson
,
T.
, and
Bowtell
,
D.
(
1993
).
The SH2 and SH3 domains of mammalian Grb2 couple the EGI receptor Io the Ras activator mSos 1
.
Nature
363
,
83
85
.
Ruff
,
P.
,
Speicher
,
D.W.
and
Husain-Chishti
,
A.
(
1991
).
Molecular identification of a major pahnitoylated erythrocyte membrane protein containing the sre homology 3 motif
.
Proc. Natl. Acad. Sei. USA
88
,
6595
6599
.
Sato
,
T.
,
Tanigaini
,
A.
,
yamakawa
,
K.
,
Akiyama
,
F.
,
Kastnni
,
F.
,
Sakamoto
, and
Nakamura
,
y.
(
1990
).
Allelotype of breast cancer: cumulative allele losses promote tumor progression in primary breast cancer
.
Cancer Res
.
50
,
7184
7189
.
Schipper
,
J. H.
,
Frixcn
,
U. H.
,
Behrens
,
J.
,
Unger
,
A.
,
Jahnke
,
K.
and
Birchineier
,
W.
(
1991
).
E-cadherin expression in squamous cell carcinomas of head and neck: Inverse correlation with tumor dedifferentiation and lymph node metastasis
.
Cancer Res
.
51
,
6328
6337
.
Schneeberger
,
E. E.
and
Lynch
,
R. D.
(
1992
).
Structure, function, and regulation of cellular light junctions
.
Am. J. Physiol. Limy Cell. Mol. Physiol
.
262
,
1647
1661
.
Shimoyama
,
y.
and
Hirohashi
,
S.
(
1991
).
Cadherin intercellular adhesion molecule in hepatocellular carcinomas: Loss of E-cadherin expression in an undifferentiated carcinoma
.
Cancer Lett
.
57
,
131
135
.
Shrestha
,
R.
and
Gateff
,
E.
(
1982
).
infrastructure and cytochemistry of the cell types in the larval hematopoietic organs and hemolymph of Drosophila nielanoyaster
.
Dev. Growth Diff
.
24
,
65
82
.
Stehle
,
T.
and
Schulz
,
G. E.
(
1990
).
Three-dimensional structure of the complex of guanylate kinase from yeast with its substrate GMP
.
J. Motee. Biol
.
211
,
249
254
.
Stehle
,
T.
and
Schulz
,
G. E.
(
1992
).
Relined structure of the complex between guanylate kinase and its substrate GMP at 2.0 A resolution
.
J. Molec. Biol
.
224
,
1127
1141
.
Sturgill
,
T. W.
and
Wu
,
J.
(
1991
).
Recent progress in characterization of protein kinase cascades for phosphorylation of ribosomal protein S6
.
Riochim. Riophys. Acta
.
1092
,
350
357
.
Traugh
,
J. A.
and
Pendergast
,
A. M.
(
1986
).
Regulation of protein synthesis by phosphorylation of ribosomal protein S6 and aminoacyl-tRNA synthetases
.
Prog. Nucl. Acid. Res. Molec. Biol
.
33
,
195
230
.
Tsuda
,
H
,
Zhang
,
W. D.
,
Shimosato
,
y.
,
yokota
,
J.
,
Tenida
,
M.
,
Sugimnra
,
T.
,
Miyamura
,
T.
and
Hirohashi
,
S.
(
1990
).
Allele loss on chromosome 16 associated with progression of human hepatocellular carcinoma
.
Proc. Natl. Acad. Sei. USA
87
,
6791
6794
.
Watson
,
K. L.
,
Johnson
,
T. K.
and
Denell
,
R. E.
(
1991
).
Lethah (1) aberrant immune response mutations leading to melanotic tumor formation in Drosophila nielanoyaster
.
Dev. Genet
.
12
,
173
187
.
Watson
,
K. L.
,
Konrad
,
K. D.
,
Woods
,
D. F.
and
Bryant
,
P. J.
(
1992
).
Drosophila homolog of the human S6 ribosomal protein is required for tumor suppression in the hematopoietic system
.
Proc. Natl. Acad. Sei. USA
89
,
11302
11306
.
Wettenhall
,
R. E.
,
Nick
,
IL P.
and
Lithgow
,
T.
(
1988
).
Primary structure of mammalian ribosomal protein S6
.
biochemistry
27
,
170
177
.
Wharton
,
K. A.
,
yedvobnick
,
B.
,
Finnerty
,
V.
and
Artavanis-Tsakonas
,
S.
(
1985
).
opa: A novel family of transcribed repeats shared by the Notch locus and other developmentally regulated loci in D. melanogaster
.
Cell
40
,
55
62
.
Wood
,
R. L.
(
1990
).
The septate junction limits mobility of lipophilic markers in plasma membranes of Hydra vulyaris (atlenitata)
.
Cell Tissue Res
.
259
,
61
66
.
Woods
.
D. F.
and
Bryant
,
P. J.
(
1989
).
Molecular cloning of the lethal(l)discs larye-1 oncogene of Drosophila
.
Develop. Biol
.
134
,
222
235
.
Woods
,
D. F.
and
Bryant
,
P. J.
(
1991
).
The discs-large tumor suppressor gene of Drosophila encodes a guanylate kinase homolog localized at septate junctions
.
Cell
66
,
451
464
.
Zhang
,
W. D.
,
Hirohashi
,
S.
,
Tsuda
,
H
,
Shimosato
,
y.
,
yokota
,
J.
,
Terada
,
M.
and
Sugimura
,
T.
(
1990
).
Frequent loss of heterozygosity on chromosomes 16 and 4 in human hepatocellular carcinoma
..
Jpn. J. Cancer Res
.
81
,
108
111
.
Zschocke
,
P. D.
,
Schiltz
,
E.
and
Schulz
,
G. E.
(
1993
).
Purification and sequence determination of guanylate kinase from pig brain
.
Eur. J. Biochem
.
213
,
263
269
.