The Core Binding Factor is a heterodimeric transcription factor complex in vertebrates that is composed of a DNA binding α-subunit and a non-DNA binding β-subunit. The α-subunit is encoded by members of the Runt Domain family of proteins and the β-subunit is encoded by the CBFβ gene. In Drosophila, two genes encoding α-subunits, runt and lozenge, and two genes encoding β-subunits, Big brother and Brother, have been previously identified. Here, a sensitized genetic screen was used to isolate mutant alleles of the Big brother gene. Expression studies show that Big brother is a nuclear protein that co-localizes with both Lozenge and Runt in the eye imaginal disc. The nuclear localization and stability of Big brother protein is mediated through the formation of heterodimeric complexes between Big brother and either Lozenge or Runt. Big brother functions with Lozenge during cell fate specification in the eye, and is also required for the development of the embryonic PNS. ds-RNA-mediated genetic interference experiments show that Brother and Big brother are redundant and function together with Runt during segmentation of the embryo. These studies highlight a mechanism for transcriptional control by a Runt Domain protein and a redundant pair of partners in the specification of cell fate during development.

A fundamental problem in biology is determining how different cell types are specified during the development of a multicellular organism. This process is usually initiated by extracellular cues that activate signal transduction cascades which produce a specific cellular response. Transcription factors such as the Core Binding Factor (CBF) play a vital role in enabling such signaling cascades to regulate a set of target genes (reviewed by Ito, 1999). CBF is a heterodimer composed of a DNA binding α-subunit and a non-DNA binding β-subunit. In mammals, three genes encoding the α-subunit, Runx1, Runx2 and Runx3, and a single gene encoding the β-subunit, CBFβ (also called PEBP2β), have been identified (Bae et al., 1993; Miyoshi et al., 1991; Ogawa et al., 1993a; Ogawa et al., 1993b; Wang et al., 1993). Runx1−/− and CBFβ−/− knock-out mice show a complete block of fetal liver blood cell development (Niki et al., 1997; Okuda et al., 1996; Sasaki et al., 1996; Wang et al., 1996a; Wang et al., 1996b). The Runx2 gene is essential for osteoblast differentiation and skeletal morphogenesis, and knock-out mice are completely deficient for bone development (Ducy et al., 1997; Komori et al., 1997; Otto et al., 1997). The Runx proteins contain within their sequence a conserved 128 amino acid motif called the Runt Domain (RD) that mediates interactions with DNA and the β-subunit (Meyers et al., 1993; Ogawa et al., 1993a; Ogawa et al., 1993b; Wang et al., 1993). The β-subunit increases the affinity of Runx proteins for DNA by altering the conformation of the Runt Domain (Meyers et al., 1993; Ogawa et al., 1993a; Ogawa et al., 1993b; Tang et al., 2000; Wang et al., 1993).

In humans, the RUNX1 and CBFβ proteins have been extensively studied in the context of oncogenic forms that cause leukemia. The most frequent translocation [t(8;21)] associated with acute myeloid leukemia (AML) encodes a fusion product between RUNX1 and the protein ETO (reviewed by Downing, 1999). The fusion protein includes the RD, interacts with CBFβ, and functions as a dominant negative transcription factor. The leukemic phenotype presumably results from the accumulation of secondary mutations. Additionally, a chromosomal inversion [Inv(16)] that generates a fusion protein, CBFβ-SMMHC, between CBFβ and the smooth muscle myosin heavy chain protein, has been associated with AML (Liu et al., 1993). This fusion protein complexes with RUNX1 but is retained in the cytoplasm, and therefore disrupts transcription (Adya et al., 1998; Kanno et al., 1998; Liu et al., 1993).

The Runt Domain was named after runt (run), best known for its role as a primary pair-rule gene in Drosophila embryonic patterning and segmentation (Gergen and Wieschaus, 1986). run also plays an important role in sex determination by directly controlling the expression of the sex lethal gene (Kramer et al., 1999). Additionally, run function is required for the development of the EL neurons in the embryonic CNS (Dormand and Brand, 1998; Duffy et al., 1991). A second protein containing the RD is Lozenge (Lz), which specifies both neuronal and non-neuronal cell types in the Drosophila eye (Daga et al., 1996; Flores et al., 1998). In the non-neuronal cone cells, Lz functions combinatorially with the transcription factors downstream of the Notch (N) and Epidermal growth factor receptor (EGFR) signaling pathways to activate the expression of D-Pax2 (also known as shaven; Flores et al., 2000). Similar mechanisms also operate in the control of the prospero (pros) gene in the eye (Xu et al., 2000). Additionally, lz functions during the development of olfactory sensory organ precursors (Gupta et al., 1998) where it regulates the expression of the proneural gene amos (Goulding et al., 2000). Finally, studies have defined a role for Lz in Drosophila hematopoiesis in establishing a sub-population of blood cells called the crystal cells (Lebestky et al., 2000; Rizki and Rizki, 1980). It is interesting to note that both vertebrate and Drosophila RD proteins are involved in the control of blood cell fate.

In Drosophila, two genes encoding β-subunits, Brother (bro) and Big brother (Bgb), were first identified through homology searches (Fujioka et al., 1996; Golling et al., 1996). These are expressed in distinct but overlapping patterns during embryogenesis (Fujioka et al., 1996; Golling et al., 1996). Furthermore, Bro and Bgb have been shown to physically interact with Run and increase its affinity for DNA (Fujioka et al., 1996; Golling et al., 1996). Overexpression experiments using either Bro or Bgb suggest that such complexes can form in vivo (Li and Gergen, 1999). A more complete functional analysis of the Drosophila partner proteins was lacking because mutations in these genes were unavailable at the time.

Here we report a loss-of-function analysis of the partner proteins. A genetic screen using a sensitized background isolated a mutation in Bgb, and ds-RNA-mediated interference strategies illustrate a redundant role for Bro and Bgb during development. We further show that Bgb functions during embryonic and eye development in the context of Run and Lz and is stabilized only in the presence of these proteins.

Scanning electron microscopy

Adult flies were fastened to metal mounts with clear fingernail polish. Images were digitally acquired using the natural SEM mode of a Hitachi S-2460N scanning electron microscope with an attached Robinson detector.

Immunohistochemistry and ds-RNA experiments

A rabbit polyclonal antibody (α-Bgb) was generated against a C-terminal peptide of the predicted Bgb sequence (RDNRQDEMEAVR) (Fujioka et al., 1996; Golling et al., 1996). The antibody was affinity purified using the peptide conjugated to a Pierce Immunopure column. The purified antibody was preadsorbed against fixed cuticle from wild-type larvae and used at a final dilution of 1:150. Additionally, the following antibodies were used: α-Run (1:250; Dormand and Brand, 1998), αPros (1:10,000; Kauffmann et al., 1996), and DSHB antibodies α2B10 (1:20), αElav (1:400), αEngrailed (1:200), and mAb22C10 (1:200). Secondary antibodies (goat α-mouse-HRP, goat α-rabbit-HRP) were obtained from Jackson ImmunoResearch and used at a final dilution of 1:200. Eye discs were stained as described previously (Rogge et al., 1995).

ds-RNA experiments were performed essentially as described previously (Kennerdell and Carthew, 1998). The results from the injections are summarized in Table 1. For each experiment, the observed segmentation phenotype is shown in three independent examples in Fig. 5. A certain fraction of the injected embryos either hatched as wild type or were grossly and non-specifically affected as a result of the injection. Their numbers relative to the total injected was independent of the nature of the ds-RNA and are included in Table 1.

Genetics

EMS (methanesulfonic acid ethyl ester; 25 mM) (Lewis and Bacher, 1968) was used to mutagenize Oregon-R male flies. The mutagenized males were crossed to lzts1 virgin females and the progeny were reared at 25°C. The F1 male progeny were screened for an eye phenotype. Mutations were balanced over either SM6a or TM6b chromosomes and the enhancement was mapped using standard mapping chromosomes.

Df(3L)BgbK4 was generated by crossing y w / Y; P[1556 ry+] ry e / TM3 with y w / y w; Δ2-3 Ki / Δ2-3 Ki. Male progeny were mated to lzts1 virgins. The progeny from this cross were reared at 25°C and the males were scored for an enhanced eye phenotype. Mosaic clones were generated by crossing w / Y; BgbD FRT80B / TM6b with ey-flp w / ey-flp w; P[w+] FRT80B / P[w+] FRT80B. Mutant tissue was identified by the absence of pigment.

Mosaic clones in adult eyes were fixed and sectioned as described previously (Coyle-Thompson and Banerjee, 1993). For rescue experiments, full length Bgb cDNA (Fujioka et al., 1996) was cloned into the XbaI and NotI site of hsp70-pCaSpeR and germline transformants were generated.

High resolution sequence analysis of the BgbD and Bgb9 mutants was performed using the 33P-Thermo Sequenase Radiolabeled Terminator Cycle Sequencing Kit (Amersham Pharmacia Biotech).

Transient transfections were performed using the Drosophila Expression System Kit (Invitrogen) and antibody staining was performed as described by Fehon (Fehon et al., 1990).

Isolation of dosage-sensitive enhancers

A temperature-sensitive allele of lz, lzts1, was used in a genetic screen to isolate mutations in interacting genes. At 25°C, lzts1 flies have near wild-type eyes (Fig. 1B). However, when raised at 29°C these flies have an easily visible rough eye phenotype (Fig. 1C). Even at this non-permissive temperature, lzts1 alleles have a milder eye phenotype than that due to a lz null allele (Fig. 1D). This led us to propose that the temperature-sensitive allele of lz functions at a threshold level such that any further reduction of the functional output of the Lz pathway would generate a mutant eye phenotype. In this scenario, at 25°C, a single copy loss-of-function mutation in a gene related to the Lz pathway is expected to generate an eye phenotype resembling that of a lzts1 fly raised at a higher temperature.

Wild-type male flies were mutagenized with EMS and were mated with virgin lzts1 females. Approximately 100,000 male progeny were reared at 25°C and scored for an eye phenotype. 10 flies were recovered that showed a disorganized eye phenotype. The mutations were crossed inter-se to generate 5 complementation groups: en(lz)4G/I, en(lz)3C/4C, en(lz)4F/4H, en(lz)Y, and en(lz)D/9 (Fig. 1E-I). These mutations are recessive embryonic lethal with the exception of en(lz)4G, which is homozygous viable with a disorganized eye. The en(lz)4G/I mutants were determined to be alleles of D-Pax2 based on their failure to complement the eye-specific D-Pax2spa-pol allele (Fu and Noll, 1997). The enhancer called Y maps to the left arm of the second chromosome between dp (25A2) and b(34D5), is not uncovered by any known deletions, and was not analyzed further. The en(lz)D/9 mutants were determined to be mutations in Bgb and will be discussed in detail below. The en(lz)3C/4C mutations mapped to the left arm of the third chromosome and failed to complement Df(3L)HR370. Additional complementation analysis with known mutations within the breakpoints of this deletion determined that the en(lz)3C/4C mutations are alleles of hsp83 (Hackett and Lis, 1983). The en(lz)4F/4H mutations mapped to the right arm of the third chromosome and failed to complement Df(3R)RK8-21. Complementation analysis with known mutations within the region established that en(lz)4F/4H are alleles of osa, also known as eyelid (eld) (Kennison and Tamkun, 1988; Treisman et al., 1997).

The en(lz)D/9 mutants were identified as alleles of Bgb based on several lines of evidence. First, we used a P-element excision strategy to generate a deletion belonging to the region because no chromosomal aberrations existed that eliminated this locus. A P-element located near Bgb was excised in a lzts1 sensitized background. About 20,000 lines were screened and one excision allele, en(lz)K4, was identified as a dominant enhancer of lzts1 at 25°C. This deletion completely eliminates the Bgb locus. The right breakpoint of this deletion maps within the transcribed region of an uncharacterized gene (Adams et al., 2000; Rubin et al., 2000) between Bro and Bgb. Thus, Df(3L)K4 does not eliminate any part of the Bro gene. The EMS-induced mutants, en(lz)D/9, fail to complement the lethality of Df(3L)K4 suggesting that they too carry a mutation in a gene uncovered by the deletion. Sequence analysis showed that en(lz)9 carries a mutation (A801C) in Bgb which results in a change from glutamic acid to aspartic acid in the Bgb sequence. Although non-conserved in the vertebrate protein, this amino acid is physically close to the segment that binds the α-subunit (Goger et al., 1999; Tang et al., 2000; Warren et al., 2000). More strikingly, the en(lz)D chromosome carries a mutation (G808A) in Bgb which results in a change from a conserved glycine to an arginine. This glycine in Bgb is conserved in the vertebrate CBFβ and Drosophila Bro proteins. This amino acid is also close to residues that physically interact with the Runt Domain (Goger et al., 1999; Tang et al., 2000; Warren et al., 2000). It is therefore likely that this mutation disrupts the stability of a Bgb/RD complex in vivo. Finally, flies carrying hsp70-Bgb constructs were generated using P-element mediated transformation (Rubin and Spradling, 1982) and assessed for rescue of the Bgb phenotype. Three independent hsp70-Bgb transformant lines were placed in homozygous en(lz)D and en(lz)9 genetic backgrounds. In each case, complete rescue to adult viability was obtained with the basal level of expression achieved without the application of a heat shock. Taken together, these data establish that en(lz)D and en(lz)9 carry mutations in Bgb and these lines were therefore renamed BgbD and Bgb9. The Df(3L)K4 was renamed Df(3L)BgbK4.

Expression studies

The Drosophila eye imaginal disc undergoes morphogenesis during the third larval instar. At this stage, a furrow sweeps across the disc from the posterior to the anterior. Cells entering the furrow undergo a synchronized round of cell division to give rise to a set of undifferentiated cells as well as pre-clusters of five neurons that differentiate in a stereotypical developmental order: R8, R2/R5, then R3/R4 (for a review of eye development see) (Wolff and Ready, 1993). The remaining undifferentiated cells subsequently give rise to R1/R6, R7, and the non-neuronal cone and pigment cells.

In order to determine the expression pattern of Bgb during morphogenesis of the eye disc, we generated a rabbit polyclonal antibody against a C-terminal peptide of Bgb that is not homologous with any peptide stretch of Bro. Bgb expression is seen in the basal nuclei of the undifferentiated cells immediately posterior to the furrow (Fig. 2A). Among the differentiated cells, Bgb expression is first obvious in the R8 cell (Fig. 2B). Posteriorly, Bgb is seen in the R1/R6 cells (Fig. 2B), and the R7 and cone cells (Fig. 2C). The level of Bgb is much higher in the R7 cell than in any of the other cells in the eye disc.

The expression of Bgb closely corresponds to that of Lz with some notable differences. Like Bgb, Lz is also expressed in the undifferentiated cells behind the furrow as well as in the nuclei of the differentiated R1, R6, R7 and cone cells (Flores et al., 1998). Unlike Bgb, however, Lz is not expressed in R8, nor is it expressed at a higher level in R7 than in the other cells. We predicted that Bgb is likely to function with a different RD protein in R7 and R8, and therefore we studied the expression pattern of Run in the eye. Strikingly, Run is specifically expressed in the R7 and R8 cells of the eye disc (Fig. 2D,E). Taken together, the expression pattern of Bgb correlates nicely with the combined nuclear expression patterns of Run and Lz.

The correlation between Lz and Bgb staining patterns is functionally relevant. In a lz mutant background, Bgb expression is eliminated from all cells in which Lz is normally expressed (Fig. 2F,G). Additionally, when Lz is misexpressed in R3 and R4, Bgb protein is seen in the nuclei of these cells (Fig. 2H). This misexpression is not accompanied by an increase in Bgb transcripts (not shown) which suggests that Lz does not transcriptionally regulate Bgb. Instead, as Lz and Bgb are binding partners, it is likely that Lz stabilizes the Bgb protein in cells in which they are co-expressed. We also examined the relationship between Run and Lz proteins. The expression of the Run protein is genetically downstream of Lz in the eye. In a lz mutant background, Run expression is limited to the R8 cell (Fig. 2I) and is eliminated from the presumptive R7 cell (Fig. 2J). Thus, in a lz mutant background, the only cell that expresses a RD protein is R8. This is significant since in this background, R8 is also the only cell that expresses Bgb (Fig. 2F). This suggests once again that the α-subunit, in this case Run, is able to stabilize the β-protein.

A correlation of the in vivo results described above was also seen in S2 cells. The nuclear localization of Lz is constitutive and does not require Bgb protein (Fig. 2M,N). In contrast, when Bgb alone was expressed in S2 cells, nearly all of the Bgb protein remained cytoplasmic (Fig. 2K). However, when a construct expressing wild-type lz was co-transfected, a majority of the Bgb protein was seen in the nucleus (Fig. 2L). Thus, Bgb cannot be translocated to the nucleus in the absence of Lz. The cytoplasmic protein can be detected in S2 cells because of the high level of expression. Presumably, in the in vivo situation, the cytoplasmic Bgb protein is rapidly degraded in the absence of an α-subunit.

Phenotypic analysis

The identification of Bgb mutants as enhancers of lzts1 provides genetic evidence that Bgb functions during eye development. The enhanced roughness produced by the loss of one copy of Bgb in this sensitized background is rather mild (Fig. 3A,B). However, rare escaper flies of Df(3L)BgbK4/BgbD genotype can be generated in a lzts1 background. These flies are extremely weak and rarely seen. They can only be rescued if they are dissected from their pupal cases. Such flies have a very strong adult eye phenotype (Fig. 3D) that is identical to that of the lz null allele (Fig. 3C).

Bgb mutants also show defects in the development of the embryonic peripheral nervous system (Fig. 3E-J) as illustrated by staining with mAb22C10 which recognizes neurons in the central and peripheral nervous systems (Fujita et al., 1982). The axons of lateral chordotonal neurons present in the abdominal segments of wild-type embryos follow two distinct paths to the midline. While the axon of the lateral monoscolopidial chordotonal organ (lch1) follows the segmental nerve (SN) to the midline, the axons of the lateral pentascolopidial chordotonal organ (lch5) follow the intersegmental nerve (ISN) (Fig. 3G). In BgbD/BgbD mutant embryos the lch1 neuron turns toward the ISN and incorrectly follows this path to the midline (Fig. 3H). A similar phenotype with higher expressivity is seen in BgbD/Df(3L)BgbK4 embryos (Fig. 3I). The hsp70-Bgb; BgbD/BgbD rescued flies are completely viable and healthy flies, and as expected, do not show this PNS phenotype (not shown). PNS defects are also seen in runYP17/runYP17 mutant embryos (Fig. 3J), however these defects are difficult to interpret because of the additional segmentation phenotypes of run (Fig. 3M), not seen in Bgb mutants (Fig. 3L). The lack of segmentation phenotypes in Bgb could be due to a maternal contribution of the Bgb gene product. Alternatively, this could reflect a functional redundancy between the two partner proteins (see below).

Bgb and Bro function redundantly during development

The results described in Fig. 3A-D establish a role for Bgb in eye development. However, when BgbD/BgbD clones were generated in an otherwise wild-type (lz+) eye, the external phenotype of the facets (Fig. 4A) as well as the morphology of all retinal cells (Fig. 4B,C) appeared completely normal. Clones of BgbD were also examined in larval eye discs that were stained with antibodies directed against Elav, Cut and Pros. The expression pattern of these markers of neuronal and non-neuronal cells was wild type in all discs examined (not shown). Thus, all neuronal and non-neuronal cells in the eye are able to develop normally in the absence of the Bgb protein. The presence of a gene encoding the second partner protein, Bro, in the Drosophila genome, suggested to us the possibility that Bro and Bgb may function redundantly during development. This was investigated during embryonic development using the ds-RNA-mediated genetic interference method (Kennerdell and Carthew, 1998).

We injected embryos with ds-Bgb, ds-bro, or a mixture of ds-Bgb and ds-bro, and compared the resulting cuticular phenotypes with those resulting from loss of run function. run mutants are characterized by defects in segmental patterns which are visualized as deletions of denticle belts. run alleles can be ordered according to their strength by examining the denticles of the second and third abdominal segments (Gergen and Wieschaus, 1986). Weaker alleles of run tend to have only small deletions between these denticles (Fig. 5B), while stronger alleles have large deletions that are often accompanied by mirror-image duplications of denticle belts (Fig. 5C). As expected, injection of ds-run results in a cuticular phenotype (Fig. 5D-F) that is identical to the one seen with run null alleles (Fig. 5C).

Injection of ds-Bgb does not have any significant effect on the pattern of segmentation (Fig. 5G-I). In comparison, ds-Bro causes a segmentation phenotype (Fig. 5J-L) reminiscent of hypomorphic alleles of run (compare with Fig. 5B). Only a partial deletion of the naked cuticle between the second and third abdominal segments is apparent. Doubling the concentration of ds-Bro did not affect the severity of the phenotype (not shown). Strikingly, simultaneous injection of both ds-Bgb and ds-Bro resulted in embryos that have a much stronger segmentation phenotype (Fig. 5M-O), which is identical to that seen in a null allele of run (Fig. 5C), or in embryos injected with ds-run (Fig. 5D-F). These data show that although the disruption of Bgb does not generate a phenotype on its own, a role for Bgb in segmentation is revealed upon the disruption of both partner genes.

Sensitized genetic screens have proved to be powerful tools in identifying interacting proteins that participate in many different developmental pathways. A particularly impressive use of this technique in the Drosophila eye led to the identification of the mutations in the components of the RTK pathway (reviewed by Zipursky and Rubin, 1994). We have used such a screening technique to generate mutations in genes that function with lz during eye development. The identification of mutations in a direct transcriptional target of Lz, D-Pax2 (Flores et al., 2000), and the gene encoding a binding partner of Lz, Bgb, suggests that this screen is able to detect proteins whose function is directly related to that of Lz.

In our screen, two alleles of hsp83 were isolated as dominant enhancers of lzts1. Drosophila Hsp83 is a chaperone protein that has been shown to physically interact with Raf (van der Straten et al., 1997). Mutations in hsp83 were identified as downstream modifiers of the sevenless and EGFR RTK pathways (Simon et al., 1991; van der Straten et al., 1997). Recent studies have indicated an extensive collaboration between RTK pathways and Lz in the regulation of direct target genes such as D-Pax2 and pros (Flores et al., 2000; Xu et al., 2000). It is therefore likely that hsp83 strengthens the RTK signal transduction cascade that functions with Lz in the regulation of target genes. In addition, HSP90, the mammalian homolog of hsp83, has been shown to associate with a variety of different transcription factors and has also been proposed to function in nuclear transport (reviewed by Helmbrecht et al., 2000). An analysis of the relationship between Hsp83 and Lz/Bgb might provide insight into the mechanism by which this transcription factor complex is translocated to the nucleus.

The screen also uncovered two alleles of osa/eld (Kennison and Tamkun, 1988; Treisman et al., 1997), a member of the brahma (brm) complex, involved in chromatin remodeling (Collins et al., 1999; Kennison and Tamkun, 1988; Vazquez et al., 1999). The identification of osa as a dominant enhancer suggests that Lz may have a function related to chromatin remodeling. This is not surprising as other Runx family members are thought to function in this manner. For example, Runx2 binding has been implicated in the remodeling of the rat osteocalcin promoter (Javed et al., 1999). Additionally, during myeloid differentiation, Runx1 has been shown to interact with p300/CBP (Kitabayashi et al., 1998), a protein involved in histone acetylation. Further, Drosophila Run has been shown to bend DNA and is likely involved in modifying the architecture of target enhancers (Golling et al., 1996). In the eye, Lz is essential for pre-patterning an undifferentiated population of cells and preparing them to activate different target genes in response to signal transduction cascades. It is possible that this process involves remodeling of the individual enhancers through the mediation of an Osa/Lz complex. The identification of osa as a genetic modifier of lz suggests the need for future biochemical experiments to establish if such protein complexes are indeed formed during development.

In this paper we have focused on the function of the partner proteins since mutations in Bgb were identified as modifiers of lz. The similarity in the phenotype of lzts1; BgbD/Df(3L)BgbK4 mutants to the null allele of lz suggests an absolute functional requirement of the partner protein during eye development. Similarly, the ds-RNA interference results suggest that both partner proteins are able to function with Run during embryonic pattern formation.

It remains to be proved if the disorganization seen in the PNS of Bgb is attributable to Bgb function with the known RD proteins. Similar PNS defects are seen in run mutants, but these phenotypes are difficult to interpret because of the additional segmentation phenotypes that could indirectly affect PNS development. It remains possible that Bgb functions with an as yet uncharacterized RD protein in the PNS. Consistent with this explanation, a survey of the sequence of the Drosophila genome (Adams et al., 2000) reveals two additional RD proteins.

Our S2 cell expression data show that Bgb is only translocated to the nucleus in the presence of Lz. Although Bgb has a nuclear localization signal (NLS; Fujioka et al., 1996), these data suggest an additional requirement of Lz binding for its transport to the nucleus. Similar regulation of nuclear transport has been reported with Single-minded (Sim) and Tango (Tgo) heterodimers (Ward et al., 1998) as well as with Homothorax (Htx) and Extradenticle (Exd) heterodimers (Pai et al., 1998). In these examples, the localization to the nucleus of either Tgo or Exd, depends on the presence of Sim or Hth, respectively (Pai et al., 1998; Ward et al., 1998). Recent work has shown that Hth binding allows nuclear transport of Exd by simultaneously inhibiting its nuclear export signal (NES) while activating its NLS (Abu-Shaar et al., 1999; Berthelsen et al., 1999). Bgb does not have a leucine-rich sequence typically associated with an NES (Fischer et al., 1995; Wen et al., 1995); co-localization into the nucleus in this case is likely to involve an unmasking of the NLS causing its exposure to the transport machinery. Obviously, nuclear localization of both the α- and the β-subunit is a prerequisite for activation of transcription. In fact, in human AML caused by Inv(16), the CBFβ fusion protein is exclusively retained within the cytoplasm (Adya et al., 1998; Kanno et al., 1998; Liu et al., 1993).

The Lz/Bgb complex provides an interesting example of post-translational stabilization of proteins through the formation of heterodimeric complexes. While we cannot rule out the possibility that low levels of Bgb protein remain in the cytoplasm of the cell in a lz mutant background, the likely explanation for the Bgb protein not being detectable in the absence of Lz or Run is that the β-subunit is degraded in the absence of the α-partner. Similar mechanisms involving degredation of a subunit operate in creating stable Exd/Hth and Sim/Tgo complexes. Tissue lacking Hth or Sim will cause degradation of Exd and Tgo, respectively (Pai et al., 1998; Ward et al., 1998). As an interesting contrast to our results, in mammalian systems it is the α-subunit, Runx1, that is stabilized by CBFβ. In this case, the absence of the β-partner causes a proteosome-mediated degradation of the α-subunit (Huang et al., 2001).

The initial cloning of Bro and Bgb raised the possibility that these genes might function redundantly during development. Although there is a stretch of 156 amino acids at the N terminus of Bgb that is not present in Bro (Fujioka et al., 1996), these proteins are 59% identical throughout the remainder of their sequence. Furthermore, Bro and Bgb have overlapping expression domains during embryogenesis (Fujioka et al., 1996; Golling et al., 1996). ds-RNA-mediated genetic interference experiments used here clearly show that Bro and Bgb function redundantly during development as heterodimeric partners of Run. A loss-of-function phenotype equivalent to a complete run null allele is only revealed in the absence of both Bro and Bgb.

The two partner proteins do not function redundantly in all tissues. This is highlighted by the fact that Bgb mutants have a PNS defect on their own. Thus, at least in this tissue, Bgb function is not redundant with that of Bro. This is different from redundant gene pairs such as BarH1 and BarH2 which are co-regulated in all tissues and always function together (Higashijima et al., 1992). It is also interesting to note that injection of ds-Bro generates a fairly strong segmentation phenotype, while injection of ds-Bgb does not affect segmentation patterning at all. Therefore, it is possible that in the wild-type fly, when both partners are present, Run preferentially functions with Bro. However, only in the absence of Bro, can compensation of Run function be achieved through its binding Bgb. A comparable situation exists in mice. The paralogs Hoxa3 and Hoxd3 are expressed in the same tissue, but clearly have distinct functional requirements (Chisaka and Capecchi, 1991; Condie and Capecchi, 1993). Yet, a compensating mechanism can be created in a background when one of the two genes is eliminated (Greer et al., 2000).

Redundancy is a classical and important problem in genetics. As vertebrate genomes are analyzed in increasing detail, it is becoming evident that single mutant phenotypes may be masked due to the function of an alternate redundant gene (Müller, 1999). The completion of the sequence of the Drosophila genome (Adams et al., 2000) has revealed many gene families that function redundantly. Indeed, the identification of the Rhomboid family of genes led to the observation that Rho1 and Rho3 function redundantly during EGFR signaling (Wasserman et al., 2000). Another example is sloppy paired1 (slp1) and sloppy paired2 (slp2) which were identified as functionally redundant genes based on the fact that they are expressed in the same tissue but at slightly different levels (Bellen et al., 1989; Grossniklaus et al., 1992; Wilson et al., 1989). In spite of these advances, detection of mutations in genes that function redundantly poses a difficult challenge to genetic analysis. Our data show that at least for the case in study, dosage-sensitive screens involving sensitized genetic backgrounds can be used for the purpose of identifying redundant genes. Bro and Bgb together can be considered to contribute 4 copies of the partner gene. Loss of 1 out of these 4 copies in a sensitized background (lzts1; Bgb Bro+/Bgb+ Bro+) gives rise to a detectable eye phenotype. Yet, loss of 2 copies in a wild-type background (lz+; Bgb Bro+/Bgb Bro+) does not generate a mutant phenotype. This remarkable sensitivity to dosage suggests that properly sensitized genetic screens could be used in the detection of redundant gene function.

Fig. 1.

A sensitized genetic screen for dosage-sensitive enhancers of lzts1. Scanning electron micrographs of adult eyes. Posterior is to the left and dorsal is up. All flies were reared at 25°C except for the one shown in C which was reared at 29°C. (A) The wild-type Drosophila eye has a regular array of ordered facets. (B) lzts1 flies, when reared at 25°C, have wild-type eyes. (C) When lzts1 flies are reared at 29°C, the eye appears rough and disorganized. (D) lzR1, which is a null allele of lz, gives rise to a more severe eye phenotype than lzts1 at any temperature. (E-I) Interacting mutations. When raised at 25°C, in a lzts1 background, loss of one copy of each enhancer locus causes a slight roughening of the posterior area of the adult eye. Under a light microscope the eyes also appear to be slightly glossy, suggesting a defect in lens secretion. (E) lzts1; en(lz)3C/+. (F) lzts1; en(lz)Y/+. (G) lzts1; en(lz)D/+. (H) lzts1; en(lz)4G/+. (I) lzts1; en(lz)4F/+.

Fig. 1.

A sensitized genetic screen for dosage-sensitive enhancers of lzts1. Scanning electron micrographs of adult eyes. Posterior is to the left and dorsal is up. All flies were reared at 25°C except for the one shown in C which was reared at 29°C. (A) The wild-type Drosophila eye has a regular array of ordered facets. (B) lzts1 flies, when reared at 25°C, have wild-type eyes. (C) When lzts1 flies are reared at 29°C, the eye appears rough and disorganized. (D) lzR1, which is a null allele of lz, gives rise to a more severe eye phenotype than lzts1 at any temperature. (E-I) Interacting mutations. When raised at 25°C, in a lzts1 background, loss of one copy of each enhancer locus causes a slight roughening of the posterior area of the adult eye. Under a light microscope the eyes also appear to be slightly glossy, suggesting a defect in lens secretion. (E) lzts1; en(lz)3C/+. (F) lzts1; en(lz)Y/+. (G) lzts1; en(lz)D/+. (H) lzts1; en(lz)4G/+. (I) lzts1; en(lz)4F/+.

Fig. 2.

Bgb and Run expression in the eye imaginal disc. Posterior is to the left. Arrow indicates the position of the furrow. (A-C) Wild-type eye disc stained with α-Bgb antibody. (A) At the basal level, nuclear staining is seen in the undifferentiated pool of cells. This staining begins immediately posterior to the furrow. (B) At a slightly more apical level, Bgb protein is seen in the nuclei of the R8, R1 and R6 cells. The expression of Bgb in the R8 cell initiates within 1 column of the furrow, while the expression in R1 and R6 initiates about 4 columns posterior to the furrow. (C) At the most apical focal plane, cone cells (cc) and the R7 cell express Bgb beginning about 7 columns posterior to the furrow. The R7 cell expresses Bgb at a much higher level than in any of the other cells. (D,E) Wild-type eye disc stained with α-Run antibody. (D) Run protein is detected in the nuclei of the R8 cell, initiating 1 column posterior to the furrow. Run expression was determined to be in R8 by the central location of this cell within an ommatidial cluster and through co-localization with the R8 specific marker, Boss (Kramer et al., 1991 and data not shown). (E) Run expression is seen in R7 about 7 columns posterior to the furrow. The timing of Run expression in R7 and R8 correlates well with the expression pattern of the Bgb protein in these cells. (F,G) Bgb expression in a lz77a7 background. lz77a7 is an eye-specific null allele of lz (Flores et al., 1998). (F) Bgb expression is limited to the R8 cell. (G) R7 expression of Bgb is absent. (H) Bgb expression in a lzsprite background. With this gain-of-function allele, lz is misexpressed in R3 and R4 (Daga et al., 1996). Arrowheads indicate a corresponding ectopic expression in R3 and R4 of the Bgb protein. (I-J) Run expression in a lz77a7 background. (I) Run expression in lz77a7 is seen only in the R8 cell. (J) At the slightly higher R7 focal plane, no Run staining is observed in the R7 cell. (K,L) S2 cells stained with α-Bgb. (K) In cells transiently transfected with Bgb, the majority of the Bgb protein is found in the cytoplasm. (L) When Bgb and lz are co-transfected, the majority of Bgb protein moves to the nucleus. (M,N) S2 cells stained with α-Lz. (M) In cells transiently transfected with lz cDNA, Lz protein is seen in the nucleus. (N) Cotransfection of Bgb and lz does not affect the localization of Lz protein.

Fig. 2.

Bgb and Run expression in the eye imaginal disc. Posterior is to the left. Arrow indicates the position of the furrow. (A-C) Wild-type eye disc stained with α-Bgb antibody. (A) At the basal level, nuclear staining is seen in the undifferentiated pool of cells. This staining begins immediately posterior to the furrow. (B) At a slightly more apical level, Bgb protein is seen in the nuclei of the R8, R1 and R6 cells. The expression of Bgb in the R8 cell initiates within 1 column of the furrow, while the expression in R1 and R6 initiates about 4 columns posterior to the furrow. (C) At the most apical focal plane, cone cells (cc) and the R7 cell express Bgb beginning about 7 columns posterior to the furrow. The R7 cell expresses Bgb at a much higher level than in any of the other cells. (D,E) Wild-type eye disc stained with α-Run antibody. (D) Run protein is detected in the nuclei of the R8 cell, initiating 1 column posterior to the furrow. Run expression was determined to be in R8 by the central location of this cell within an ommatidial cluster and through co-localization with the R8 specific marker, Boss (Kramer et al., 1991 and data not shown). (E) Run expression is seen in R7 about 7 columns posterior to the furrow. The timing of Run expression in R7 and R8 correlates well with the expression pattern of the Bgb protein in these cells. (F,G) Bgb expression in a lz77a7 background. lz77a7 is an eye-specific null allele of lz (Flores et al., 1998). (F) Bgb expression is limited to the R8 cell. (G) R7 expression of Bgb is absent. (H) Bgb expression in a lzsprite background. With this gain-of-function allele, lz is misexpressed in R3 and R4 (Daga et al., 1996). Arrowheads indicate a corresponding ectopic expression in R3 and R4 of the Bgb protein. (I-J) Run expression in a lz77a7 background. (I) Run expression in lz77a7 is seen only in the R8 cell. (J) At the slightly higher R7 focal plane, no Run staining is observed in the R7 cell. (K,L) S2 cells stained with α-Bgb. (K) In cells transiently transfected with Bgb, the majority of the Bgb protein is found in the cytoplasm. (L) When Bgb and lz are co-transfected, the majority of Bgb protein moves to the nucleus. (M,N) S2 cells stained with α-Lz. (M) In cells transiently transfected with lz cDNA, Lz protein is seen in the nucleus. (N) Cotransfection of Bgb and lz does not affect the localization of Lz protein.

Fig. 3.

Bgb phenotypes. (A-D) SEMs of adult eyes. Posterior is to the left and dorsal is up. All flies were reared at 25oC. (A) lzts1; BgbD/+. The BgbD mutant enhances lzts1 on the posterior side of the eye (compare with Fig. 1B). (B) lzts1; Df(3L)BgbK4/+. The BgbK4 deletion enhances the lzts1 eye phenotype. (C) lzR1. The lz null eye is very smooth and does not show any ommatidial organization. (D) lzts1; Df(3L)BgbK4/BgbD. Rare escapers of this genotype have a phenotype resembling a lz null eye. (E-M) Embryonic phenotype of Bgb and run mutants. Anterior is to the left and dorsal is up. Embryos were stained with mAb22C10 (E-J) or α-Engrailed antibody (K-M). (E) Wild-type stage-15 embryo. mAb22C10 recognizes neurons of the chordotonal organs (box). These are easily identified on the basis of their stereotypical arrangement and position in abdominal segments A1-A7. (F) A schematic diagram of the lateral chordotonal neurons based on that by Campos-Ortega and Hartenstein (1997). Within each wild-type cluster (WT panel) the axon of the single lateral chordotonal neuron, lch1, pioneers the segmental nerve (SN) to its target in the midline. The remaining five lateral chordotonal neurons, lch5, follow the intersegmental nerve (ISN) to their target in the midline. In Bgb mutant embryos (Bgb panel), the axon of the lch1 neuron incorrectly follows the ISN to the midline. In this and subsequent panels an arrowhead indicates the absence of the lch1 axon in its normal position. This axon does not join the SN, but instead aberrantly turns toward the ISN and follows this incorrect path to the midline. The ventral sensilla, vp5, has an apical projection (asterisk in F and H to distinguish this structure from the axon of the lch1 neuron). (G,H) Lateral chordotonal neurons of (G) wild-type and (H) BgbD/BgbD stage-15 embryos. (G) The axons of the lch1 neurons correctly follow the SN to the midline (arrow). (H) In the mutants, axons of the lch1 neurons do not follow the SN (arrowheads), but incorrectly project towards the lch5 and follow the ISN to the midline. In this genetic background the phenotype is not fully expressed and only 2 of the 3 lateral chordotonal clusters shown have this defect. (I) Lateral chordotonal neurons of BgbD/Df(3L)BgbK4 stage-15 embryo. In these mutants, the axons of the lch1 neurons aberrantly project towards the lch5 and do not follow the SN to the midline (arrowheads). This phenotype is similar to, but stronger than, that seen in (H). (J) Lateral chordotonal neurons of runYP17/runYP17 stage-15 embryo. Only two sets of lateral chordotonal neurons are seen within the region corresponding to the previous panels because of a lack of segments in this mutant (Gergen and Wieschaus, 1986). The axon of one of the lch1 neurons shown misprojects and does not follow the SN to the midline (arrowhead). (K) Wild-type stage-11 embryo. Engrailed protein is expressed as a 14 stripe pattern in the posterior compartment of each segment. (L) BgbD/BgbD stage-11 embryo. Engrailed expression is as seen in wild type. (M) runYP17/runYP17 stage-11 embryo. These embryos show a segmentation phenotype, not seen in Bgb mutants.

Fig. 3.

Bgb phenotypes. (A-D) SEMs of adult eyes. Posterior is to the left and dorsal is up. All flies were reared at 25oC. (A) lzts1; BgbD/+. The BgbD mutant enhances lzts1 on the posterior side of the eye (compare with Fig. 1B). (B) lzts1; Df(3L)BgbK4/+. The BgbK4 deletion enhances the lzts1 eye phenotype. (C) lzR1. The lz null eye is very smooth and does not show any ommatidial organization. (D) lzts1; Df(3L)BgbK4/BgbD. Rare escapers of this genotype have a phenotype resembling a lz null eye. (E-M) Embryonic phenotype of Bgb and run mutants. Anterior is to the left and dorsal is up. Embryos were stained with mAb22C10 (E-J) or α-Engrailed antibody (K-M). (E) Wild-type stage-15 embryo. mAb22C10 recognizes neurons of the chordotonal organs (box). These are easily identified on the basis of their stereotypical arrangement and position in abdominal segments A1-A7. (F) A schematic diagram of the lateral chordotonal neurons based on that by Campos-Ortega and Hartenstein (1997). Within each wild-type cluster (WT panel) the axon of the single lateral chordotonal neuron, lch1, pioneers the segmental nerve (SN) to its target in the midline. The remaining five lateral chordotonal neurons, lch5, follow the intersegmental nerve (ISN) to their target in the midline. In Bgb mutant embryos (Bgb panel), the axon of the lch1 neuron incorrectly follows the ISN to the midline. In this and subsequent panels an arrowhead indicates the absence of the lch1 axon in its normal position. This axon does not join the SN, but instead aberrantly turns toward the ISN and follows this incorrect path to the midline. The ventral sensilla, vp5, has an apical projection (asterisk in F and H to distinguish this structure from the axon of the lch1 neuron). (G,H) Lateral chordotonal neurons of (G) wild-type and (H) BgbD/BgbD stage-15 embryos. (G) The axons of the lch1 neurons correctly follow the SN to the midline (arrow). (H) In the mutants, axons of the lch1 neurons do not follow the SN (arrowheads), but incorrectly project towards the lch5 and follow the ISN to the midline. In this genetic background the phenotype is not fully expressed and only 2 of the 3 lateral chordotonal clusters shown have this defect. (I) Lateral chordotonal neurons of BgbD/Df(3L)BgbK4 stage-15 embryo. In these mutants, the axons of the lch1 neurons aberrantly project towards the lch5 and do not follow the SN to the midline (arrowheads). This phenotype is similar to, but stronger than, that seen in (H). (J) Lateral chordotonal neurons of runYP17/runYP17 stage-15 embryo. Only two sets of lateral chordotonal neurons are seen within the region corresponding to the previous panels because of a lack of segments in this mutant (Gergen and Wieschaus, 1986). The axon of one of the lch1 neurons shown misprojects and does not follow the SN to the midline (arrowhead). (K) Wild-type stage-11 embryo. Engrailed protein is expressed as a 14 stripe pattern in the posterior compartment of each segment. (L) BgbD/BgbD stage-11 embryo. Engrailed expression is as seen in wild type. (M) runYP17/runYP17 stage-11 embryo. These embryos show a segmentation phenotype, not seen in Bgb mutants.

Fig. 4.

Lack of Bgb eye phenotype in a lz+ background. (A) Somatic clone of BgbD/BgbD tissue in the eye. The mutant tissue was first identified at low magnification using the difference in pigmentation between the mutant and the wild-type tissue. This boundary was subsequently marked in the SEM with the black dotted line. No defects are evident in the external structure of the facets within the clone. (B,C) Plastic section through a somatic clone of BgbD/BgbD tissue. (B) In this clone, marked by the lack of pigment granules, the R1-6 and R7 cells (1-7) appear completely normal. (C) In this deeper section of the same clone, R8 cells (8) are seen to develop normally.

Fig. 4.

Lack of Bgb eye phenotype in a lz+ background. (A) Somatic clone of BgbD/BgbD tissue in the eye. The mutant tissue was first identified at low magnification using the difference in pigmentation between the mutant and the wild-type tissue. This boundary was subsequently marked in the SEM with the black dotted line. No defects are evident in the external structure of the facets within the clone. (B,C) Plastic section through a somatic clone of BgbD/BgbD tissue. (B) In this clone, marked by the lack of pigment granules, the R1-6 and R7 cells (1-7) appear completely normal. (C) In this deeper section of the same clone, R8 cells (8) are seen to develop normally.

Fig. 5.

ds-RNA-mediated genetic interference. Cuticle preparations showing denticle belts corresponding to abdominal segments A1-A8 and thoracic segment T3. Anterior is to the left. (A) Wild-type injected with buffer only. T3 and A1-A8 denticle belts are clearly visible. (B) runYP17. This hypomorphic allele produces a weak segmentation phenotype. A deletion of the lateral edges between the A2/A3 denticle belts results in the curved appearance of these belts. Deletion of the naked cuticle between the A4/A5 and the A6/A7 denticles is also evident. (C) runXD106. This null allele produces a severe segmentation phenotype. Large deletions between the A2/A3, A4/A5, and A6/A7 denticle belts are apparent. These deletions result in mirror image duplications of the corresponding denticles. (D-F) Three independent examples of wild-type embryos injected with double stranded RNA corresponding to the run gene (ds-run). In each example there is a complete deletion of the naked cuticle between A2/A3, A4/A5, and A6/A7 which results in a mirror image duplication of the remaining denticles. This is the same phenotype as that seen in null alleles of run (compare with C). (G-I) Three independent examples of wild-type embryos injected with ds-Bgb. No perceptible alteration of the segmentation pattern is evident. These animals are indistinguishable from those injected with buffer (compare with A). (J-L) Three independent examples of wild-type embryos injected with ds-Bro. Defects in the segmentation pattern are similar to those seen in hypomorphic run mutants (compare with B). (M-O) Three independent examples of wild-type embryos injected with a mixture of ds-Bgb and ds-Bro. An extremely strong segmentation phenotype identical to that seen with null alleles of run (compare with C) and wild-type embryos injected with ds-run (compare with D-E) can be seen. Thus, although injection of ds-Bgb did not affect segmentation, and injection of ds-Bro produced a mild segmentation defect, injection of a mixture of both ds-Bgb and ds-Bro had a synergistic effect on the pattern of segmentation.

Fig. 5.

ds-RNA-mediated genetic interference. Cuticle preparations showing denticle belts corresponding to abdominal segments A1-A8 and thoracic segment T3. Anterior is to the left. (A) Wild-type injected with buffer only. T3 and A1-A8 denticle belts are clearly visible. (B) runYP17. This hypomorphic allele produces a weak segmentation phenotype. A deletion of the lateral edges between the A2/A3 denticle belts results in the curved appearance of these belts. Deletion of the naked cuticle between the A4/A5 and the A6/A7 denticles is also evident. (C) runXD106. This null allele produces a severe segmentation phenotype. Large deletions between the A2/A3, A4/A5, and A6/A7 denticle belts are apparent. These deletions result in mirror image duplications of the corresponding denticles. (D-F) Three independent examples of wild-type embryos injected with double stranded RNA corresponding to the run gene (ds-run). In each example there is a complete deletion of the naked cuticle between A2/A3, A4/A5, and A6/A7 which results in a mirror image duplication of the remaining denticles. This is the same phenotype as that seen in null alleles of run (compare with C). (G-I) Three independent examples of wild-type embryos injected with ds-Bgb. No perceptible alteration of the segmentation pattern is evident. These animals are indistinguishable from those injected with buffer (compare with A). (J-L) Three independent examples of wild-type embryos injected with ds-Bro. Defects in the segmentation pattern are similar to those seen in hypomorphic run mutants (compare with B). (M-O) Three independent examples of wild-type embryos injected with a mixture of ds-Bgb and ds-Bro. An extremely strong segmentation phenotype identical to that seen with null alleles of run (compare with C) and wild-type embryos injected with ds-run (compare with D-E) can be seen. Thus, although injection of ds-Bgb did not affect segmentation, and injection of ds-Bro produced a mild segmentation defect, injection of a mixture of both ds-Bgb and ds-Bro had a synergistic effect on the pattern of segmentation.

Table 1.
graphic
graphic

We are grateful to Y. Ito for sharing results prior to publication. We appreciate the helpful comments from P. Gergen. We thank A. Brand, R. Carthew, I. Duncan, P. Gergen, K. Matthews, N. Perrimon, G. Rubin, J. Treisman, and the Bloomington stock center, for fly stocks and antibodies. We thank V. Hartenstein for help in analyzing the PNS phenotype. We thank the members of the Banerjee laboratory for thoughtful discussions and comments on the manuscript. We thank Rahul Warrior for helpful suggestions about 33P sequencing. We acknowledge the work performed by Karin Dumstrei in the isolation of the Df(3L)BgbK4 mutant and we are grateful for help from Hui Wang in generating the germline transformants. The 2B10, ELAV, 4D9 anti-engrailed/invected, and 22C10 antibodies were obtained from the University of Iowa Developmental Studies Hybridoma Bank developed under the auspices of the NICHD. This work was supported by an NIH grant to U. B. (2RO1EY08152) and a US Public Health Service National Research Service Award to J. K. (GM07185) and T. L. (GM07185).

Abu-Shaar, M., Ryoo, H. D. and Mann, R. S. (
1999
). Control of the nuclear localization of Extradenticle by competing nuclear import and export signals.
Genes Dev
.
13
,
935
-945.
Adams, M. D., Celniker, S. E., Holt, R. A., Evans, C. A., Gocayne, J. D., Amanatides, P. G., Scherer, S. E., Li, P. W., Hoskins, R. A., Galle, R. F. et al. (
2000
). The genome sequence of Drosophila melanogaster.
Science
287
,
2185
-2195.
Adya, N., Stacy, T., Speck, N. A. and Liu, P. P. (
1998
). The leukemic protein core binding factor beta (CBFbeta)-smooth-muscle myosin heavy chain sequesters CBFalpha2 into cytoskeletal filaments and aggregates.
Mol. Cell. Biol
.
18
,
7432
-7443.
Bae, S. C., Yamaguchi-Iwai, Y., Ogawa, E., Maruyama, M., Inuzuka, M., Kagoshima, H., Shigesada, K., Satake, M. and Ito, Y. (
1993
). Isolation of PEBP2 alpha B cDNA representing the mouse homolog of human acute myeloid leukemia gene, AML1.
Oncogene
8
,
809
-814.
Bellen, H. J., O’Kane, C. J., Wilson, C., Grossniklaus, U., Pearson, R. K. and Gehring, W. J. (
1989
). P-element-mediated enhancer detection: a versatile method to study development in Drosophila.
Genes Dev
.
3
,
1288
-1300.
Berthelsen, J., Kilstrup-Nielsen, C., Blasi, F., Mavilio, F. and Zappavigna, V. (
1999
). The subcellular localization of PBX1 and EXD proteins depends on nuclear import and export signals and is modulated by association with PREP1 and HTH.
Genes Dev
.
13
,
946
-953.
Campos-Ortega, J. A. and Hartenstein, V. (
1997
). The Embryonic Development of Drosophila melanogaster. Berlin, New York: Springer.
Chisaka, O. and Capecchi, M. R. (
1991
). Regionally restricted developmental defects resulting from targeted disruption of the mouse homeobox gene hox-1.5 [see comments].
Nature
350
,
473
-479.
Collins, R. T., Furukawa, T., Tanese, N. and Treisman, J. E. (
1999
). Osa associates with the Brahma chromatin remodeling complex and promotes the activation of some target genes.
EMBO J
.
18
,
7029
-7040.
Condie, B. G. and Capecchi, M. R. (
1993
). Mice homozygous for a targeted disruption of Hoxd-3 (Hox-4.1) exhibit anterior transformations of the first and second cervical vertebrae, the atlas and the axis.
Development
119
,
579
-595.
Coyle-Thompson, C. A. and Banerjee, U. (
1993
). The strawberry notch gene functions with Notch in common developmental pathways.
Development
119
,
377
-395.
Daga, A., Karlovich, C. A., Dumstrei, K. and Banerjee, U. (
1996
). Patterning of cells in the Drosophila eye by Lozenge, which shares homologous domains with AML1.
Genes Dev
.
10
,
1194
-1205.
Dormand, E. L. and Brand, A. H. (
1998
). Runt determines cell fates in the Drosophila embryonic CNS.
Development
125
,
1659
-1667.
Downing, J. R. (
1999
). The AML1-ETO chimaeric transcription factor in acute myeloid leukaemia: biology and clinical significance.
Brit. J. Haematol
.
106
,
296
-308.
Ducy, P., Zhang, R., Geoffroy, V., Ridall, A. L. and Karsenty, G. (
1997
). Osf2/Cbfa1: a transcriptional activator of osteoblast differentiation [see comments].
Cell
89
,
747
-754.
Duffy, J. B., Kania, M. A. and Gergen, J. P. (
1991
). Expression and function of the Drosophila gene runt in early stages of neural development.
Development
113
,
1223
-1230.
Fehon, R. G., Kooh, P. J., Rebay, I., Regan, C. L., Xu, T., Muskavitch, M. A. and Artavanis-Tsakonas, S. (
1990
). Molecular interactions between the protein products of the neurogenic loci Notch and Delta, two EGF-homologous genes in Drosophila.
Cell
61
,
523
-534.
Fischer, U., Huber, J., Boelens, W. C., Mattaj, I. W. and Lührmann, R. (
1995
). The HIV-1 Rev activation domain is a nuclear export signal that accesses an export pathway used by specific cellular RNAs.
Cell
82
,
475
-483.
Flores, G. V., Daga, A., Kalhor, H. R. and Banerjee, U. (
1998
). Lozenge is expressed in pluripotent precursor cells and patterns multiple cell types in the Drosophila eye through the control of cell-specific transcription factors.
Development
125
,
3681
-3687.
Flores, G. V., Duan, H., Yan, H., Nagaraj, R., Fu, W., Zou, Y., Noll, M. and Banerjee, U. (
2000
). Combinatorial signaling in the specification of unique cell fates.
Cell
103
,
75
-85.
Fu, W. and Noll, M. (
1997
). The Pax2 homolog sparkling is required for development of cone and pigment cells in the Drosophila eye [see comments].
Genes Dev
.
11
,
2066
-2078.
Fujioka, M., Yusibova, G. L., Sackerson, C. M., Tillib, S., Mazo, A., Satake, M. and Goto, T. (
1996
). Runt domain partner proteins enhance DNA binding and transcriptional repression in cultured Drosophila cells.
Genes Cells
1
,
741
-754.
Fujita, S. C., Zipursky, S. L., Benzer, S., Ferrús, A. and Shotwell, S. L. (
1982
). Monoclonal antibodies against the Drosophila nervous system.
Proc. Natl Acad. Sci. USA
79
,
7929
-7933.
Gergen, J. P. and Wieschaus, E. (
1986
). Dosage requirements for runt in the segmentation of Drosophila embryos.
Cell
45
,
289
-299.
Goger, M., Gupta, V., Kim, W. Y., Shigesada, K., Ito, Y. and Werner, M. H. (
1999
). Molecular insights into PEBP2/CBF beta-SMMHC associated acute leukemia revealed from the structure of PEBP2/CBF beta.
Nat. Struct. Biol
.
6
,
620
-623.
Golling, G., Li, L., Pepling, M., Stebbins, M. and Gergen, J. P. (
1996
). Drosophila homologs of the proto-oncogene product PEBP2/CBF beta regulate the DNA-binding properties of Runt.
Mol. Cell. Biol
.
16
,
932
-942.
Goulding, S. E., zur Lage, P. and Jarman, A. P. (
2000
). amos, a proneural gene for Drosophila olfactory sense organs that is regulated by lozenge [see comments].
Neuron
25
,
69
-78.
Greer, J. M., Puetz, J., Thomas, K. R. and Capecchi, M. R. (
2000
). Maintenance of functional equivalence during paralogous Hox gene evolution [see comments].
Nature
403
,
661
-665.
Grossniklaus, U., Pearson, R. K. and Gehring, W. J. (
1992
). The Drosophila sloppy paired locus encodes two proteins involved in segmentation that show homology to mammalian transcription factors.
Genes Dev
.
6
,
1030
-1051.
Gupta, B. P., Flores, G. V., Banerjee, U. and Rodrigues, V. (
1998
). Patterning an epidermal field: Drosophila lozenge, a member of the AML-1/Runt family of transcription factors, specifies olfactory sense organ type in a dose-dependent manner.
Dev. Biol
.
203
,
400
-411.
Hackett, R. W. and Lis, J. T. (
1983
). Localization of the hsp83 transcript within a 3292 nucleotide sequence from the 63B heat shock locus of D. melanogaster.
Nucl. Acids Res
.
11
,
7011
-7030.
Helmbrecht, K., Zeise, E. and Rensing, L. (
2000
). Chaperones in cell cycle regulation and mitogenic signal transduction: a review.
Cell Prolif
.
33
,
341
-365.
Higashijima, S., Michiue, T., Emori, Y. and Saigo, K. (
1992
). Subtype determination of Drosophila embryonic external sensory organs by redundant homeo box genes BarH1 and BarH2.
Genes Dev
.
6
,
1005
-1018.
Huang, G., Shigesada, K., Ito, K., Wee, H. J., Yokomizo, T. and Ito, Y. (
2001
). Dimerization with PEBP2beta protects RUNX1/AML1 from ubiquitin-proteasome-mediated degradation.
EMBO J
.
20
,
723
-733.
Ito, Y. (
1999
). Molecular basis of tissue-specific gene expression mediated by the runt domain transcription factor PEBP2/CBF.
Genes Cells
4
,
685
-696.
Javed, A., Gutierrez, S., Montecino, M., van Wijnen, A. J., Stein, J. L., Stein, G. S. and Lian, J. B. (
1999
). Multiple Cbfa/AML sites in the rat osteocalcin promoter are required for basal and vitamin D-responsive transcription and contribute to chromatin organization.
Mol. Cell. Biol
.
19
,
7491
-7500.
Kanno, Y., Kanno, T., Sakakura, C., Bae, S. C. and Ito, Y. (
1998
). Cytoplasmic sequestration of the polyomavirus enhancer binding protein 2 (PEBP2)/core binding factor alpha (CBFalpha) subunit by the leukemia-related PEBP2/CBFbeta-SMMHC fusion protein inhibits PEBP2/CBF-mediated transactivation.
Mol. Cell. Biol
.
18
,
4252
-4261.
Kauffmann, R. C., Li, S., Gallagher, P. A., Zhang, J. and Carthew, R. W. (
1996
). Ras1 signaling and transcriptional competence in the R7 cell of Drosophila.
Genes Dev
.
10
,
2167
-2178.
Kennerdell, J. R. and Carthew, R. W. (
1998
). Use of dsRNA-mediated genetic interference to demonstrate that frizzled and frizzled 2 act in the wingless pathway.
Cell
95
,
1017
-1026.
Kennison, J. A. and Tamkun, J. W. (
1988
). Dosage-dependent modifiers of polycomb and antennapedia mutations in Drosophila.
Proc. Natl. Acad. Sci. USA
85
,
8136
-8140.
Kitabayashi, I., Yokoyama, A., Shimizu, K. and Ohki, M. (
1998
). Interaction and functional cooperation of the leukemia-associated factors AML1 and p300 in myeloid cell differentiation.
EMBO J
.
17
,
2994
-3004.
Komori, T., Yagi, H., Nomura, S., Yamaguchi, A., Sasaki, K., Deguchi, K., Shimizu, Y., Bronson, R. T., Gao, Y. H., Inada, M. et al. (
1997
). Targeted disruption of Cbfa1 results in a complete lack of bone formation owing to maturational arrest of osteoblasts [see comments].
Cell
89
,
755
-764.
Kramer, H., Cagan, R. L. and Zipursky, S. L. (
1991
). Interaction of bride of sevenless membrane-bound ligand and the sevenless tyrosine-kinase receptor [see comments].
Nature
352
,
207
-212.
Kramer, S. G., Jinks, T. M., Schedl, P. and Gergen, J. P. (
1999
). Direct activation of Sex-lethal transcription by the Drosophila runt protein.
Development
126
,
191
-200.
Lebestky, T., Chang, T., Hartenstein, V. and Banerjee, U. (
2000
). Specification of Drosophila hematopoietic lineage by conserved transcription factors.
Science
288
,
146
-149.
Lewis, E. B. and Bacher, F. (
1968
). Methods of feeding ethyl methane sulfonate (EMS) to Drosophila males.
Dros. Inf. Serv
.
43
,
193
-194.
Li, L. H. and Gergen, J. P. (
1999
). Differential interactions between Brother proteins and Runt domain proteins in the Drosophila embryo and eye.
Development
126
,
3313
-3322.
Liu, P., Tarle, S. A., Hajra, A., Claxton, D. F., Marlton, P., Freedman, M., Siciliano, M. J. and Collins, F. S. (
1993
). Fusion between transcription factor CBF beta/PEBP2 beta and a myosin heavy chain in acute myeloid leukemia.
Science
261
,
1041
-1044.
Meyers, S., Downing, J. R. and Hiebert, S. W. (
1993
). Identification of AML-1 and the (8;21) translocation protein (AML-1/ETO) as sequence-specific DNA-binding proteins: the runt homology domain is required for DNA binding and protein-protein interactions.
Mol. Cell. Biol
.
13
,
6336
-6345.
Miyoshi, H., Shimizu, K., Kozu, T., Maseki, N., Kaneko, Y. and Ohki, M. (
1991
). t(8;21) breakpoints on chromosome 21 in acute myeloid leukemia are clustered within a limited region of a single gene, AML1.
Proc. Natl. Acad. Sci. USA
88
,
10431
-10434.
Müller, U. (
1999
). Ten years of gene targeting: targeted mouse mutants, from vector design to phenotype analysis.
Mech. Dev
.
82
,
3
-21.
Niki, M., Okada, H., Takano, H., Kuno, J., Tani, K., Hibino, H., Asano, S., Ito, Y., Satake, M. and Noda, T. (
1997
). Hematopoiesis in the fetal liver is impaired by targeted mutagenesis of a gene encoding a non-DNA binding subunit of the transcription factor, polyomavirus enhancer binding protein 2/core binding factor.
Proc. Natl. Acad. Sci. USA
94
,
5697
-5702.
Ogawa, E., Inuzuka, M., Maruyama, M., Satake, M., Naito-Fujimoto, M., Ito, Y. and Shigesada, K. (
1993
a). Molecular cloning and characterization of PEBP2 beta, the heterodimeric partner of a novel Drosophila runt-related DNA binding protein PEBP2 alpha.
Virology
194
,
314
-331.
Ogawa, E., Maruyama, M., Kagoshima, H., Inuzuka, M., Lu, J., Satake, M., Shigesada, K. and Ito, Y. (
1993
b). PEBP2/PEA2 represents a family of transcription factors homologous to the products of the Drosophila runt gene and the human AML1 gene.
Proc. Natl. Acad. Sci. USA
90
,
6859
-6863.
Okuda, T., van Deursen, J., Hiebert, S. W., Grosveld, G. and Downing, J. R. (
1996
). AML1, the target of multiple chromosomal translocations in human leukemia, is essential for normal fetal liver hematopoiesis.
Cell
84
,
321
-330.
Otto, F., Thornell, A. P., Crompton, T., Denzel, A., Gilmour, K. C., Rosewell, I. R., Stamp, G. W., Beddington, R. S., Mundlos, S., Olsen, B. R. et al. (
1997
). Cbfa1, a candidate gene for cleidocranial dysplasia syndrome, is essential for osteoblast differentiation and bone development [see comments].
Cell
89
,
765
-771.
Pai, C. Y., Kuo, T. S., Jaw, T. J., Kurant, E., Chen, C. T., Bessarab, D. A., Salzberg, A. and Sun, Y. H. (
1998
). The Homothorax homeoprotein activates the nuclear localization of another homeoprotein, extradenticle, and suppresses eye development in Drosophila.
Genes Dev
.
12
,
435
-446.
Rizki, T. M. and Rizki, R. M. (
1980
). A mutant affecting the crystal cells in Drosophila melanogaster.
Roux’s Arch. Dev. Biol
.
188
.
Rogge, R., Green, P. J., Urano, J., Horn-Saban, S., Mlodzik, M., Shilo, B. Z., Hartenstein, V. and Banerjee, U. (
1995
). The role of yan in mediating the choice between cell division and differentiation.
Development
121
,
3947
-3958.
Rubin, G. M., Hong, L., Brokstein, P., Evans-Holm, M., Frise, E., Stapleton, M. and Harvey, D. A. (
2000
). A Drosophila complementary DNA resource.
Science
287
,
2222
-2224.
Rubin, G. M. and Spradling, A. C. (
1982
). Genetic transformation of Drosophila with transposable element vectors.
Science
218
,
348
-353.
Sasaki, K., Yagi, H., Bronson, R. T., Tominaga, K., Matsunashi, T., Deguchi, K., Tani, Y., Kishimoto, T. and Komori, T. (
1996
). Absence of fetal liver hematopoiesis in mice deficient in transcriptional coactivator core binding factor beta.
Proc. Natl. Acad. Sci. USA
93
,
12359
-12363.
Simon, M. A., Bowtell, D. D., Dodson, G. S., Laverty, T. R. and Rubin, G. M. (
1991
). Ras1 and a putative guanine nucleotide exchange factor perform crucial steps in signaling by the sevenless protein tyrosine kinase.
Cell
67
,
701
-716.
Tang, Y. Y., Crute, B. E., Kelley, J. J., Huang, X., Yan, J., Shi, J., Hartman, K. L., Laue, T. M., Speck, N. A. and Bushweller, J. H. (
2000
). Biophysical characterization of interactions between the core binding factor alpha and beta subunits and DNA.
FEBS Lett
.
470
,
167
-172.
Treisman, J. E., Luk, A., Rubin, G. M. and Heberlein, U. (
1997
). eyelid antagonizes wingless signaling during Drosophila development and has homology to the Bright family of DNA-binding proteins.
Genes Dev
.
11
,
1949
-1962.
van der Straten, A., Rommel, C., Dickson, B. and Hafen, E. (
1997
). The heat shock protein 83 (Hsp83) is required for Raf-mediated signalling in Drosophila.
EMBO J
.
16
,
1961
-1969.
Vazquez, M., Moore, L. and Kennison, J. A. (
1999
). The trithorax group gene osa encodes an ARID-domain protein that genetically interacts with the brahma chromatin-remodeling factor to regulate transcription.
Development
126
,
733
-742.
Wang, Q., Stacy, T., Binder, M., Marin-Padilla, M., Sharpe, A. H. and Speck, N. A. (
1996
a). Disruption of the Cbfa2 gene causes necrosis and hemorrhaging in the central nervous system and blocks definitive hematopoiesis.
Proc. Natl. Acad. Sci. USA
93
,
3444
-3449.
Wang, Q., Stacy, T., Miller, J. D., Lewis, A. F., Gu, T. L., Huang, X., Bushweller, J. H., Bories, J. C., Alt, F. W., Ryan, G. et al. (
1996
b). The CBFbeta subunit is essential for CBFalpha2 (AML1) function in vivo.
Cell
87
,
697
-708.
Wang, S., Wang, Q., Crute, B. E., Melnikova, I. N., Keller, S. R. and Speck, N. A. (
1993
). Cloning and characterization of subunits of the T-cell receptor and murine leukemia virus enhancer core-binding factor.
Mol. Cell. Biol
.
13
,
3324
-3339.
Ward, M. P., Mosher, J. T. and Crews, S. T. (
1998
). Regulation of bHLH-PAS protein subcellular localization during Drosophila embryogenesis.
Development
125
,
1599
-1608.
Warren, A. J., Bravo, J., Williams, R. L. and Rabbitts, T. H. (
2000
). Structural basis for the heterodimeric interaction between the acute leukaemia-associated transcription factors AML1 and CBFbeta.
EMBO J
.
19
,
3004
-3015.
Wasserman, J. D., Urban, S. and Freeman, M. (
2000
). A family of rhomboid-like genes: Drosophila rhomboid-1 and roughoid/rhomboid-3 cooperate to activate EGF receptor signaling.
Genes Dev
.
14
,
1651
-1663.
Wen, W., Meinkoth, J. L., Tsien, R. Y. and Taylor, S. S. (
1995
). Identification of a signal for rapid export of proteins from the nucleus.
Cell
82
,
463
-473.
Wilson, C., Pearson, R. K., Bellen, H. J., O’Kane, C. J., Grossniklaus, U. and Gehring, W. J. (
1989
). P-element-mediated enhancer detection: an efficient method for isolating and characterizing developmentally regulated genes in Drosophila.
Genes Dev
.
3
,
1301
-1313.
Wolff, T. and Ready, D. F. (
1993
). Pattern formation in the Drosophila retina. In The Development of Drosophila melanogaster (ed. M. Bate and A. Martinez Arias), pp. 1277-1325. Plainview, N.Y.: Cold Spring Harbor Laboratory Press.
Xu, C., Kauffmann, R. C., Zhang, J., Kladny, S. and Carthew, R. W. (
2000
). Overlapping activators and repressors delimit transcriptional response to receptor tyrosine kinase signals in the Drosophila eye.
Cell
103
,
87
-97.
Zipursky, S. L. and Rubin, G. M. (
1994
). Determination of neuronal cell fate: lessons from the R7 neuron of Drosophila.
Ann. Rev.Neurosci
.
17
,
373
-397.