fushi tarazu


DEVELOPMENTAL BIOLOGY

Embryonic

The seven ftz stripes arise individually in a distinct, non-linear order along the anterior-posterior axis of the embryo, in even numbered stripes starting just anterior to the cephalic furrow [Images]. The indentation that will become the posterior midgut invagination forms behind the posterior most stripe. The pattern is complementary to that of even-skipped. In addition, the stripes develop differentially along the dorsal-ventral axis; most stripes emerge on the ventral side and then gradually spread dorsally until they surround the entire circumference of the embryo (Carroll, 1985, Hafen, 1984 and Yu, 1995). In addition to expression in the epidermis, ftz is transiently expressed in neuroblasts and glia of the developing CNS. In the absense of ftz CNS expression, the RP2 neuron extends a growth cone along an abnormal pathway, mimicking its sibling (RP1), suggesting a transformation in neuronal identity. evenskipped is expressed in this same neuron (Doe, 1988).

See Chris Doe's Hyper-Neuroblast map site for information on the expression of fushi tarazu in specific neuroblasts.

For more information on Drosophila neuroblast lineages, see Linking neuroblasts to their corresponding lineage, a site carried by Flybrain, an online atlas and database of the Drosophila nervous system.

Asymmetric mRNA localization targets proteins to their cytoplasmic site of function. The mechanism of apical localization of wingless and pair-rule transcripts in the Drosophila blastoderm embryo has been elucidated by directly visualizing intermediates along the entire path of transcript movement. After release from their site of transcription, mRNAs diffuse within the nucleus and are exported to all parts of the cytoplasm, regardless of their cytoplasmic destinations. Endogenous and injected apical RNAs assemble selectively into cytoplasmic particles that are transported apically along microtubules. Cytoplasmic dynein is required for correct localization of endogenous transcripts and apical movement of injected RNA particles. It is proposed that dynein-dependent movement of RNA particles is a widely deployed mechanism for mRNA localization (Wilkie, 2001).

To study the mechanism of apical localization, whether actin and/or MTs are necessary for localization of injected mRNA was tested by preinjecting cytoskeletal inhibitors 10 min before injecting the RNA. It was found that preinjection of Cytochalasin B, at concentrations that disrupt the organization of actin filaments, has no affect on Runt mRNA localization. However, a similar disruption of nuclear position has been observed in the cortical cytoplasm. In contrast, preinjection of colcemid, which destabilizes blastoderm MTs, disrupts runt, wingless, and fushi tarazu RNA localization almost entirely. It is concluded that an intact MT cytoskeleton is required for apical localization of injected RNA and that actin does not play a major role in the process. However, some minor role for actin in apical localization of RNA cannot be excluded (Wilkie, 2001).

Whether the localization of injected RNA occurs by minus end directed MT-dependent motor movement was tested by preinjecting embryos with antibodies against Drosophila cytoplasmic dynein heavy chain (dhc). Two independently raised monoclonal antibodies against dhc are each sufficient to inhibit RUN, FTZ, and WG mRNA apical localization in most, or all, embryos. Either one, the anti-dynein antibody or the colcemid injections, is sufficient to cause apical RNA to partly diffuse away from the site of injection in a similar manner to embryos injected with HB RNA alone. Injected apical RNA does not diffuse in the absence of anti-dynein antibodies or Colcemid preinjections. These results suggest that apical RNA is tethered to MTs by dynein and that dynein is required for the transport of RNA particles (Wilkie, 2001).

To further test the involvement of cytoplasmic dynein in apical transcript localization, RNA was injected into mutant cytoplasmic dynein heavy chain (Dhc64C) embryos. A marked reduction was found in the speed of movement of injected apical targeted RNAs in dynein mutants. Cytoplasmic dynein is essential for many cellular processes, so strong mutations in Dhc64C are homozygous lethal in Drosophila and cannot be studied at the blastoderm stage. Instead hypomorphic allelic combinations of Dhc64C, which are viable in trans due to intragenic complementation, were used. In two different allelic combinations of Dhc64C, injected RNA particles move at speeds 60% to 70% slower than they do in wild-type. Staining Dhc64C mutant embryos with anti-tubulin antibodies showsthat MT distribution is indistinguishable from wild-type, indicating that the reduced speed of localization is not due indirectly to a disruption of the MTs. Instead, the reduction in speed is likely to show a direct requirement for dynein in particle transport (Wilkie, 2001).

To test whether cytoplasmic dynein is also required for apical localization of endogenous transcripts, the effects of Dhc64C hypomorphic mutants and anti-dhc antibodies on the apical localization of endogenous FTZ transcripts was tested by in situ hybridization. As expected, hypomorphic Dhc64C mutants show no detectable effects on FTZ apical mRNA localization since injected RNA localizes correctly, but more slowly. In contrast, injection of anti-dhc antibody disrupts endogenous FTZ localization, leading to unlocalized stripes of ftz mRNA 20–30 min after injection. Given that FTZ mRNA has a half-life of 6 min in the blastoderm, the FTZ transcripts observed are likely to have been synthesized after the injection. It is concluded that endogenous apical mRNA localization is also dynein dependent (Wilkie, 2001).

Dynactin is a protein complex that is involved in coordinating the activities of cytoplasmic dynein, and is thought to be required for most forms of dynein-based transport. To test whether dynactin is also required for apical RNA localization, a large excess of p50/dynamitin is preinjected into embryos 10 min before injecting apically targeted RNA. p50/dynamitin causes a significant reduction in the speed of RNA particle movement. p50/dynamitin is a subunit of dynactin whose overexpression is a widely used method of disrupting the dynactin complex and demonstrating conclusively dynein-dependent motility. Dynactin is required for some cargo binding and for dynein processivity. It is concluded that apical transcript localization in the blastoderm embryo occurs by cytoplasmic dynein- and dynactin-mediated transport along MTs toward their minus ends (Wilkie, 2001).

It is thought that export and localization of apical mRNA in the blastoderm embryo can be divided into six distinct steps. (1) During or after completion of transcription and processing, transcripts are assembled into particles, which contain various hnRNPs and export factors, some of which may form part of the cytoplasmic localization machinery. (2) mRNA particles diffuse freely after release from the site of transcription and processing until they reach nuclear pore complexes (NPCs) on the nuclear periphery. (3) mRNA particles are exported through NPCs in all parts of the nuclear envelope. (4) The composition of the particles probably changes during export from the nucleus and in the cytoplasm to recruit dynein, dynactin, and associated proteins. (5) Particles attach to MTs and are actively transported to the apical cytoplasm. (6) Particle movement arrests in the apical cytoplasm, where they may associate with other particles and become anchored (Wilkie, 2001).

The first three steps of apical localization are thought to be common to most mRNAs, because they are essential universal processes in eukaryotic cells. However, the last three steps of the localization pathway are likely to vary among different kinds of transcripts, since the key determinant in sorting different mRNAs to their correct cytoplasmic destinations is presumably RNP particle composition in the cytoplasm. It is possible that some components required for cytoplasmic sorting are preassembled in the nucleus, as suggested by studies showing that the localization of injected FTZ mRNA depends on preincubation with the hnRNPA1 protein Squid. Indeed, a requirement for hnRNPs has also been shown for GRK mRNA localization in the oocyte, for myelin basic protein mRNA in rat oligodendrocytes, and for Vg1 transcripts in Xenopus oocytes. However, the data in this study show that injected protein-free apical RNA assembles in the cytoplasm into particles that localize correctly, arguing that all the factors needed to assemble competent localization particles can also be recruited in the cytoplasm (Wilkie, 2001).

Effects of Mutation or Deletion

ftz mutation is lethal. Comparison of the cuticular segmentation pattern of wild type and ftz mutant larvae shows the presence of denticle belts from only the odd numbered segments (Hafen, 1984).

Recently, double-stranded RNA (dsRNA) has been found to be a potent and specific inhibitor of gene activity in the nematode Caenorhabditis elegans (Fire, 1998). The potential of dsRNA to interfere with the function of genes in Drosophila, termed RNA inhibition or RNAi) has been investigated. Injection of dsRNA into embryos resulted in potent and specific interference of several genes that were tested. dsRNA corresponding to four genes with previously defined functions was introduced. dsRNA is shown to potently and specifically inhibits the activities of wg, fushi tarazu (ftz), even-skipped (eve), and tramtrack (ttk). The reception mechanism of the morphogen Wingless was determined using dsRNA. Interference of the frizzled and Drosophila frizzled 2 genes together produces defects in embryonic patterning that mimic the loss of wingless function. Interference with the function of either gene alone has no effect on patterning. Epistasis analysis indicates that frizzled and Drosophila frizzled 2 act downstream of wingless and upstream of zeste-white3 in the Wingless pathway. These results demonstrate that dsRNA interference can be used to analyze many aspects of gene function (Kennerdell, 1998).

To determine whether dsRNA-mediated interference can occur in Drosophila, RNA was synthesized in vitro, allowed to anneal, and then injected into syncytial blastoderm embryos. The ftz and eve genes were chosen for initial characterization of this method based on several criteria. Both genes are required for embryonic segmentation. Transcription of ftz and eve begins approximately 90 to 120 min after egg laying, which corresponds to a time 10 to 60 min after dsRNA injection. Although both genes function in the first few hours of embryogenesis, null mutant animals survive to the end of embryogenesis and exhibit segmentation defects in their cuticle. Finally, mutants with reduced activity of either ftz or eve produce increasingly severe phenotypes, such that a semiquantitative relationship exists between genotype and phenotype. Antisense and sense RNAs for each gene were synthesized and annealed. Injection of either ftz- or eve-annealed RNA into wild-type embryos effectively interfers with gene activity as demonstrated by cuticle phenotypes characteristic of ftz or eve mutants. In contrast, antisense or sense RNAs injected separately have an order-of-magnitude weaker interference activity than annealed RNA. When a decreasing amount of ftz-annealed RNA is injected, interference activity declines also, though interference was still detectable at the lowest dose. The abundance of each RNA strand at this dose was calculated to be about 2 million molecules per injected embryo. Assuming uniform distribution of RNA, the original injected material is diluted to about 30 molecules per cell. Thus, dsRNA is a robust inhibitor of gene activity in Drosophila, comparable in its potency to that observed in C. elegans (Kennerdell, 1998).

The phenotypes produced by ds-ftz and ds-eve RNAs are highly specific. Injected animals exhibit cuticle defects indistinguishable from ftz and eve loss-of-function mutants. The phenotypes vary significantly among individuals, possibly due to variability in the injected dose. At high doses of ds-ftz RNA, the majority of animals exhibit the null mutant phenotype. At lower doses of ds-ftz RNA, the majority of animals exhibit localized or patchy interference. This localized phenotype is consistent with loss of ftz activity. Even within a group of animals given the same dose, variation in phenotype is apparent. Some ds-eve RNA-treated animals exhibit a lawn of denticles characteristic of the known null mutant, while the remaining animals exhibit a complete pair-rule phenotype or localized pair-rule phenotype characteristic of partial loss of eve function. Since both ftz and eve are expressed in cells spanning 60% the embryo's length, the complete phenotypes observed indicate that interference can occur in cells throughout the embryo. The observed interference is at the level of gene expression. Little or no endogenous Ftz protein is observed in embryos injected with ds-ftz RNA. In contrast, embryos injected with buffer exhibit a normal pattern of Ftz protein expression (Kennerdell, 1998).

Embryos of higher metazoans are divided into repeating units early in development. In Drosophila, the earliest segmental units to form are the parasegments. Parasegments are initially defined by alternating stripes of expression of the fushi-tarazu and even-skipped genes. How ftz and eve define the parasegment boundaries, and how parasegments are lost when ftz or eve fail to function correctly, has never been fully or properly explained. It is shown that parasegment widths are defined early by the relative levels of ftz and eve at stripe junctions. Changing these levels results in alternating wide and narrow parasegments. When shifted by 30% or more, the enlarged parasegments remain enlarged and the reduced parasegments are lost. Loss of the reduced parasegments occurs in three steps; delamination of cells from the epithelial layer, apoptosis of the delaminated cells and finally, apoptosis of inappropriate cells remaining at the surface. The establishment and maintenance of vertebrate metameres may be governed by similar processes and properties (Hughes, 2001).

Previous studies have shown that ftz and eve are the primary determinants of parasegmental boundaries and identities (even versus odd). Until quite recently, it was believed that the two genes perform these roles relatively late (stages 6-7), and that high levels and sharp anterior stripe boundaries are crucial. However, when in the right proportions, the absolute levels of ftz and eve are not particularly important. ftz and eve first define the positions of parasegment borders prior to the completion of cellularization (mid stage 5). At this time, ftz and eve stripes have a bell-shaped distribution across each stripe, and the stripes overlap with one another at their edges. It is suggested that parasegment boundaries occur at the points where stripes intersect and activity levels are equivalent. If the activity of one gene is raised while the other remains unchanged, these positions of equivalency move. The result is an alternating set of narrow and wide parasegments. These shifts become more pronounced with greater changes in activity or when both genes change in opposite directions. However, if both gene activities are increased or decreased at the same time the positions of equivalency do not change, and parasegments remain equal in width (Hughes, 2001).

It is suggested that the transition from overlapping stripe boundaries to sharp non-overlapping boundaries occurs via a combination of autoregulatory and mutually antagonistic functions. For example, if above a certain relative threshold level, ftz autoregulation dominates over repression by eve, and ftz expression rises to maximal levels while eve expression is lost. If below that relative threshold, repression by eve dominates over ftz autoregulation and ftz expression is lost while eve rises to maximal levels. The ability of ftz and eve to autoregulate and to mutually repress one another (directly or indirectly) has been well documented. Once the borders of ftz and eve stripes are established, combinatorial interactions with other segmentation gene products then determine where downstream targets such as en and wg are activated or repressed, thereby locking in the positions of the parasegment boundaries (Hughes, 2001).

ftz and eve pair-rule phenotypes have been described and explained in a number of conflicting ways. The remaining cuticle is not simply composed of every second parasegment, nor is it composed of double-width or homeotically re-transformed segments. A relative decrease in ftz or eve activity causes a decrease in width of alternate parasegments and a corresponding expansion of adjacent parasegments. The smaller parasegments are excised and the enlarged parasegments retained. Efficient deletion (greater than 90%) of the reduced parasegments occurs when they are reduced by 30% or more. Enlarged parasegments are 1.4-1.5 times wider than normal parasegments. This degree of enlargement remains the same when levels of ftz or eve are increased further or when the levels of ftz are reduced to zero (eve nulls affect all parasegments due to earlier roles). It is suggested that these maximal widths reflect the edges of stage 5 ftz and eve stripes, beyond which autoregulation cannot occur. Further expansion of these stripes may be limited by the actions of other pair-rule or gap gene products. The resulting larva is composed of half the normal number of segments, but these are 1.3-1.5 times wider than normal segments, giving an overall length that is about 65%-75% the length of a normal larva (Hughes, 2001).

Parasegments are considered to be the first 'compartments' to form within the embryo. Compartments are fields of cells that originate from a common group of founder cells and that remain defined in lineage thereafter. Cells within adjoining compartments do not mix, most likely due to differential adhesion properties. Compartments are further defined by unique gene expression patterns (e.g. ftz and eve) that respect their boundaries. Another property of compartments relevant to this study is that they are capable of sensing and modulating their size. Changes in size can be induced by injury, transplantation, irradiation, or genetic manipulation. In the case of reductions in size, compensation is most often in the form of increased cellular proliferation, and when increased in size, by programmed cell death. These studies show that parasegments can compensate for changes in size, but that this ability is relatively limited. Both reduced and enlarged parasegments showed changes in the normal numbers of apoptotic events. Dying cells are rarely seen in the ectoderm of reduced parasegments while higher than normal numbers are seen in enlarged parasegments. The numbers of dying cells and the time of onset are proportional to the degree of parasegment enlargement. These changes, however, are insufficient to compensate for the changes in widths induced in this study (Hughes, 2001).

It was also found that changes in mitotic frequency, as an alternate form of compensation, do not occur. Once established, the ratio of the number of cells per mutant parasegment, as compared to wild-type segments, remains relatively constant until cells in the reduced segments begin to delaminate. This finding agrees with those obtained previously by increasing the number of copies of the bicoid gene. Reduced parasegments in the compacted middle of the embryo fail to compensate by increasing rates of mitosis. However, these changes in width were usually subtle enough (<20%) that most segments were able to recover by reducing their rates of apoptosis. These results show that once these changes reach 30% or higher, variations in apoptotic frequencies can no longer compensate (Hughes, 2001).

One of the most novel and intriguing findings of this study was the unstable nature of reduced parasegments and the manner in which they are removed. It was found that this occurs via a three-step process. First, large patches of cells move out of the ectodermal layer. Next, they pinch off from the overlying ectoderm and then programmed cell death is initiated. Finally, the fused engrailed stripes remaining at the surface are resolved by late and sporadic apoptotic events. Although the precise spatial and temporal details of this process vary between individual embryos and different mutant backgrounds, the general trends and final consequences are the same (Hughes, 2001).

The delamination of reduced parasegment cells occurs primarily during the late stages of germ band retraction. This coincidence between reduced parasegment delamination and germ band retraction suggests the possibility that cellular movement and adhesion may play a prominent role in the delamination process. During germ band retraction, normal parasegments are reduced in width by almost half (approx. 11 cells to 7). In reduced parasegments, the corresponding decrease results in an average width of just 3 cells. This reduced width means significantly fewer contacts with other reduced parasegment cells and more contacts with the cells of neighboring parasegments. This may drive the reduced parasegment cells to increase homogeneous contacts by forming spheres, much as observed in imaginal discs when small clones of anterior compartment identity are formed in the posterior compartment (Hughes, 2001).

The protein encoded by the Drosophila pair-rule gene fushi tarazu (ftz) is required for the formation of the even-numbered parasegments. The phenotypes were analyzed of ectopic expression of Ftz and Ftz protein deletions from the Tubulin alpha1 (Tubalpha1) promoter. Fusion of ftz to the Tubalpha1 promoter results in low-level ectopic expression of Ftz relative to Ftz expressed from the endogenous ftz gene. The effects of ectopic expression of four Ftz proteins, Ftz1-413 (full length wild-type Ftz), Ftzdelta257-316 (a complete deletion of the HD), Ftzdelta101-150 (a deletion that includes the major Ftz-F1 binding site) and Ftzdelta151-209 were determined. Ectopic expression of Ftz1-413, Ftzdelta257-316 and Ftzdelta101-151 did not result in an anti-ftz phenotype; however, ectopic expression of Ftz1-413, and Ftzdelta257-316 did result in a ftzUal/Rpl-like phenotype. In addition, low-level ectopic expression of Ftz1-413 and Ftzdelta257-316 rescued ftz phenotypes. This was an important observation because the even-numbered parasegment pattern of Ftz expression is considered important for normal segmentation. Therefore, the rescue of ftz phenotypes by low-level Ftz expression in all cells of the embryo suggests that the even-numbered parasegment expression pattern of Ftz is not the sole factor restricting Ftz action. Low-level ectopic expression of Ftzdelta151-209 resulted in the anti-ftz phenotype and rescued hypomorphic ftz-f1 phenotypes indicating that Ftzdelta151-209 is a hyperactive Ftz molecule. Therefore, the region encompassing amino acids 151-209 of Ftz is required in some manner for repression of Ftz activity. These results are discussed in relation to the current understanding of the mechanism of Ftz action (Argiropoulos, 2003).

A screen for genes that interact with the Drosophila pair-rule segmentation gene fushi tarazu

fushi tarazu is expressed at the blastoderm stage in seven stripes that serve to define the even-numbered parasegments. ftz encodes a DNA-binding homeodomain protein and is known to regulate genes of the segment polarity, homeotic, and pair-rule classes. Despite intensive analysis in a number of laboratories, how ftz is regulated and how it controls its targets are still poorly understood. To help understand these processes, a screen was conducted to identify dominant mutations that enhance the lethality of a ftz temperature-sensitive mutant. Twenty-six enhancers were isolated, which define 21 genes. All but one of the mutations recovered show a maternal effect in their interaction with ftz. Three of the enhancers proved to be alleles of the known ftz protein cofactor gene ftz-f1, demonstrating the efficacy of the screen. Four enhancers are alleles of Atrophin (Atro), the Drosophila homolog of the human gene responsible for the neurodegenerative disease dentatorubral-pallidoluysian atrophy. Embryos from Atro mutant germ-line mothers lack the even-numbered (ftz-dependent) engrailed stripes and show strong ftz-like segmentation defects. These defects likely result from a reduction in Even-skipped (Eve) repression ability, since Atro has been shown to function as a corepressor for Eve. In this study, evidence is presented that Atro is also a member of the trithorax group (trxG) of Hox gene regulators. Atro appears to be particularly closely related in function to the trxG gene osa, which encodes a component of the brahma chromatin remodeling complex. One additional gene was identified that causes pair-rule segmentation defects in embryos from homozygous mutant germ-line mothers. The single allele of this gene, called bek, also causes nuclear abnormalities similar to those caused by alleles of the Trithorax-like gene, which encodes the GAGA factor (Kankel, 2004).

Four of the ftz enhancers isolated in the screen proved to be alleles of Atrophin (Atro). Polyglutamine tract expansion within one of the human homologs of Atro (Atrophin-1) causes the neurodegenerative disease dentatorubral-pallidoluysian atrophy. Humans possess at least one additional Atrophin family member, Atrophin-2, which encodes a protein that can heterodimerize with Atr1. The functions of the mammalian Atrophin proteins are not well characterized. However, a role in gene repression seems likely, because Atr1 binds Eto1, a corepressor complex component, and overexpression of Atr1 can repress transcription of a reporter gene in tissue culture cells. In addition, Atr2 has been shown to interact with the histone deacetylase Hdac1. Compelling evidence has been presented that Atro also functions as a corepressor in Drosophila. eve mutations show strong dominant lethality when crossed to mothers heterozygous for Atro alleles. In the eve/+; Atro/+ embryos produced in this cross, odd-numbered en stripes are expanded, suggesting a weakening in the ability of Eve to repress paired, runt, or sloppy-paired (other pair-rule genes involved in specifying these stripes). Atro binds to the minimal repression domain of Eve, and artificial recruitment of Atro to a target gene can cause repression in vivo. A failure in the repressive activity of Eve may account for the absence of even-numbered en stripes described for embryos from Atro mutant germ-line mothers. In normal development, the even-numbered en stripes form as a result of differential repression of ftz and odd-skipped (odd) by Eve. Ftz is an activator of en, whereas Odd is a repressor. The even-numbered en stripes form where odd, but not ftz, has been repressed by Eve. If there were a failure of Eve to repress odd, zones expressing ftz but not odd would not form, and the even-numbered en stripes would not be established. Exactly this mechanism appears to be responsible for a reduction in even-numbered en stripes in mutants for the Rpd3 histone deacetylase. However, it is also possible that the even-numbered en stripes fail to appear in Atro- embryos because of a defect in the ability of Ftz to activate en. It is important to note that the odd-numbered en stripes are established almost normally in Atro mutant embryos (although they are wider than normal). These stripes are thought to be defined by differential repression of sloppy-paired, runt, and paired by Eve; the presence of these stripes in Atro- embryos indicates that Atro is not required for all repressive activities of Eve (Kankel, 2004).


REFERENCES

Alonso, C. R., et al. (2001). Evolution of Ftz protein function in insects Curr. Biol. 11: 1473-1478. 11566109

Amati, B. and Gasser, S. M. (1990a). Drosophila scaffold-attached regions bind nuclear scaffolds and can function as ARS elements in both budding and fission yeasts. Mol. Cell. Biol. (10): 5442-5454. PubMed Citation: 2118998

Amati, B., et al. (1990b). Nuclear scaffold attachment stimulates, but is not essential for ARS activity in Saccharomyces cerevisiae: analysis of the Drosophila ftz SAR. EMBO J. (12): 4007-4016. PubMed Citation: 2123454

Andrioli, L. P., Digiampietri, L. A., de Barros, L. P. and Machado-Lima, A. (2012). Huckebein is part of a combinatorial repression code in the anterior blastoderm. Dev. Biol. 361(1): 177-85. PubMed Citation: 22027434

Argiropoulos, B., et al. (2003). Low-level ectopic expression of Fushi tarazu in Drosophila melanogaster results in ftzUal/Rpl-like phenotypes and rescues ftz phenotypes. Mech. Dev. 120: 1443-1453. 14654217

Baumgartner, S.,Martin, D.,Hagios, C. and Chiquet-Ehrismann, R. (1994). Tenm, a Drosophila gene related to tenascin, is a new pair-rule gene. EMBO J 13: 3728-3740. PubMed Citation: 8070401

Becker, P. B. and Wu, C. (1992). Cell-Free system for assembly of transcriptionally repressed chromatin from Drosophila embryos. Mol. Cell. Biol. 12: 2241-48. PubMed Citation: 1569951

Belozerov, V. E., Majumder, P., Shen, P. and Cai, H. N. (2003). A novel boundary element may facilitate independent gene regulation in the Antennapedia complex of Drosophila. EMBO J. 22: 3113-3121. 12805225

Berman, B. P., et al. (2004). Computational identification of developmental enhancers: conservation and function of transcription factor binding-site clusters in Drosophila melanogaster and Drosophila pseudoobscura. Genome Biol. 5(9):R61. 15345045

Biggin, M. D. and McGinnis, W. (1997). Regulation of segmentation and segmental identity by Drosophila homeoproteins: the role of DNA binding in functional activity and specificity. Development 124(22): 4425-4433. PubMed Citation: 9409661

Binari, R. and Perrimon, N. (1994). Stripe-specific regulation of pair-rule genes by hopscotch, a putative Jak family tyrosine kinase in Drosophila. Genes Dev 8: 300-12. PubMed Citation: 8314084

Bouchard, M., St-Amand, J. and Cote, S. (2000). Combinatorial activity of pair-rule proteins on the Drosophila gooseberry early enhancer. Dev. Biol. 222: 135-146. PubMed Citation: 10885752

Brown, J. L. and Wu, C. (1993). Repression of Drosophila pair-rule segmentation genes by ectopic expression of tramtrack. Development 117: 45-58. PubMed Citation: 8223261

Brown S. J., Hilgenfeld, R. B. and Denell, R.E. (1994) The beetle Tribolium castaneum has a fushi tarazu homolog expressed in stripes during segmentation. Proc. Natl. Acad. Sci. 91 (26): 12922-12926. PubMed Citation: 7809147

Brown, S. J., et al. (1997). Molecular characterization and embryonic expression of the even-skipped ortholog of Tribolium castaneum. Mech. Dev. 61 (1-2): 165-173. PubMed Citation: 9076686

Brown, S. J., Hilgenfeld, R. B. and Denell, R. E. (1994). The beetle Tribolium castaneum has a fushi tarazu homolog expressed in stripes during segmentation. Proc. Natl. Acad. Sci. 91 (26): 12922-12926. PubMed Citation: 7809147

Bullock, S. L. and Ish-Horowicz, D. (2001). Conserved signals and machinery for RNA transport in Drosophila oogenesis and embryogenesis. Nature 414(6864): 611-6. 11740552

Calhoun, V. C., Stathopoulos, A. and Levine, M. (2002). Promoter-proximal tethering elements regulate enhancer-promoter specificity in the Drosophila Antennapedia complex. Proc. Natl. Acad. Sci. 99: 9243-9247. 12093913

Calhoun, V. C. and Levine, M. (2003). Long-range enhancer-promoter interactions in the Scr-Antp interval of the Drosophila Antennapedia complex. Proc. Natl. Acad. Sci. 100(17): 9878-83. 12909726

Carroll, S.B. and Scott, M.P. (1985). Localization of Fushi tarazu protein during Drosophila embryogenesis. Cell 43: 47-57. PubMed Citation: 3000605

Chiang, C. and Beachy, C. A. (1995). Expression of a novel Toll-like gene spans the parasegment boundary and contributes to hedgehog function in the adult eye of Drosophila. Mech. Dev. 47(3): 225--239. PubMed Citation: 7848870

Choksi, S. P., et al. (2006). Prospero acts as a binary switch between self-renewal and differentiation in Drosophila neural stem cells. Dev. Cell 11(6): 775-89. Medline abstract: 17141154

Colas, J. F., et al. (1995). Drosophila 5-HT2 serotonin receptor: coexpression with fushi-tarazu during segmentation. Proc Natl Acad Sci USA 92: 5441-5445. PubMed Citation: 7777527

Colgan, J., Wampler, S. and Manley, J. L. (1993). Interaction between a transcriptional activator and transcription factor IIB in vivo. Nature 362: 549-53. PubMed Citation: 8464496

Copeland, J. W., et al. (1996). Patterning of the Drosophila embryo by a homeodomain-deleted Ftz polypeptide. Nature 379: 162-165. PubMed Citation: 8538765

Core, N., et al. (1997). Transcriptional regulation of the Drosophila homeotic gene teashirt by the homeodomain protein Fushi tarazu. Mech. Dev. 68(1-2): 157-172. PubMed Citation: 9431813

Crispino, J. D., et al. (1996). Cis-acting elements distinct from the 5' splice site promote U1-independent pre-mRNA splicing. RNA 2: 664-673. PubMed Citation: 8756409

Dearolf, C. R., Topol, J. and Parker, C. S. (1989). The caudal gene product is a direct activator of fushi tarazu transcription during Drosophila embryogenesis. Nature 341: 340-3. PubMed Citation: 2571934

Delanoue, R. and Davis, I. (2005). Dynein anchors its mRNA cargo after apical transport in the Drosophila blastoderm embryo. Cell 122: 97-106. 16009136

Deshpande, G., et al. (1999). Novel functions of nanos in downregulating mitosis and transcription during the development of the Drosophila germline. Cell 99: 271-281.

Doe, C.Q., et al. (1988). Expression and function of the segmentation gene fushi tarazu during Drosophila neurogenesis. Science 239: 170-175

Doe, C.Q., Chu-LaGraff, Q., Wright, D.M. and Scott, M.P. (1991). The prospero gene specifies cell fates in the Drosophila central nervous system. Cell 65: 451-464

Dong, J., et al. (1998). A phosphorylation site in the Ftz homeodomain is required for activity. EMBO J. 17: 2308-2318

Dréan, B. S.-Le, et al. (1998). Dynamic changes in the functions of Odd-skipped during early Drosophila embryogenesis. Development 125: 4851-4861

Erkner, A., et al. (2002). Grunge, related to human Atrophin-like proteins, has multiple functions in Drosophila development. Development 129: 1119-1129. 11874908

Fang, M., Ren, H., Liu, J., Cadigan, K. M., Patel, S. R. and Dressler, G. R. (2009). Drosophila ptip is essential for anterior/posterior patterning in development and interacts with the PcG and trxG pathways. Development 136(11): 1929-38. PubMed Citation: 19429789

Fire, A., Xu, S., Montgomery, M., Kostas, S.A., Driver, S.E., and Mello, C.C. (1998). Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature 391: 806-811. PubMed Citation: 9486653

Florence, B., et al. (1997). Ftz-F1 is a cofactor in Ftz activation of the Drosophila engrailed gene. Development 124: 839-847. PubMed Citation: 9043065

Francis-Lang, H., Davis, I., and Ish-Horowicz, D. (1996). Asymmetric localization of Drosophila pair-rule transcripts from displaced nuclei: evidence for directional nuclear export. EMBO J. 15: 640-649. PubMed Citation: 8599947

Gabilondo, H., et al. (2011). A targeted genetic screen identifies crucial players in the specification of the Drosophila abdominal Capaergic neurons. Mech. Dev. 128(3-4): 208-21. PubMed Citation: 21236339

Gindhart, J. G., Jr., King, A. N. and Kaufman, T. C. (1995). Characterization of the cis-regulatory region of the Drosophila homeotic gene Sex combs reduced. Genetics 139: 781-795. 7713432

Guichet, A., et al. (1997). The nuclear receptor homolog Ftz-F1 and the homeodomain protein Ftz are mutually dependent cofactors. Nature 385: 548-552

Hafen, E., Kuriowa, A. and Gehring, W.J. (1984). Spatial distribution of the transcripts from the segmentation gene fushi tarazu during Drosophila embryonic development. Cell 37: 833-841

Han, W., Yu, Y., Altan, N. and Pick, L. (1993). Multiple proteins interact with the fushi tarazu proximal enhancer. Mol. Cell. Biol. 13: 5549-59

Han, W., et al. (1998). A binding site for multiple transcriptional activators in the fushi tarazu proximal enhancer is essential for gene expression in vivo. Mol. Cell. Biol. 18(6): 3384-3394

Hillman, M. A. and Gecz, J. (2001). Fragile XE-associated familial mental retardation protein 2 (FMR2) acts as a potent transcription activator. J. Hum. Genet. 46: 251-259. PubMed Citation: 11355014

Hiromi, Y. and Gehring, W. J. (1987). Regulation and function of the Drosophila segmentation gene fushi tarazu. Cell 50: 963-74

Hou, H. Y., et al. (2009). Stripy Ftz target genes are coordinately regulated by Ftz-F1. Dev. Biol. 335(2): 442-53. PubMed Citation: 19679121

Hughes, C. L. and Kaufman, T. C. (2002). Exploring the myriapod body plan: expression patterns of the ten Hox genes in a centipede. Development 129: 1225-1238. 11874918

Hughes, S. C. and Krause, H. M. (2001). Establishment and maintenance of parasegmental compartments. Development 128: 1109-1118. 11245576

Jiménez, G., Pinchin, S. M. and Ish-Horowicz, D. (1996). In vivo interactions of the Drosophila Hairy and Runt transcriptional repressors with target promoters. EMBO J. 7088-7098

John, A., Smith, S. T. and Jaynes, J. B. (1995). Inserting the Ftz homeodomain into engrailed creates a dominant transcriptional repressor that specifically turns off Ftz target genes in vivo. Development 121: 1801-1813

Kankel, M. W., Duncan, D. M. and Duncan, I. (2004). A screen for genes that interact with the Drosophila pair-rule segmentation gene fushi tarazu. Genetics 168(1): 161-80. 15454535

Katsani, K. R., Hajibagheri, M. A. and Verrijzer, C. P. (1999). Co-operative DNA binding by GAGA transcription factor requires the conserved BTB/POZ domain and reorganizes promoter topology. EMBO J. 18(3): 698-708

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

Kosman, D. and Small, S. (1997). Concentration-dependent patterning by an ectopic expression domain of the Drosophila gap gene knirps. Development 124: 1343-1354

Krause, H.M. and Gehring, W.J. (1989). Stage-specific phosphorylation of the fushi tarazu protein during Drosophila development. EMBO J 8: 1197-1204

Lall, S., et al. (1999). Squid hnRNP protein promotes apical cytoplasmic transport and localization of Drosophila pair-rule transcripts. Cell 98(2): 171-80

Laughon, A. and Scott, M.P. (1984). Sequence of a Drosophila segmentation gene: protein structure homology with the DNA-binding proteins. Nature 310: 25-31

Lavorgna, G., Ueda, H. Clos, J. and Wu, C. (1991). FTZ-F1, a steroid hormone receptor-like protein implicated in the activation of fushi tarazu. Science 252: 848-51

Lawrence, P. A. and Pick, L. (1998). How does the fushi tarazu gene activate engrailed in the Drosophila embryo? Dev. Genet. 23(1): 28-34

Li, M. and Arnosti, D. N. (2011). Long- and short-range transcriptional repressors induce distinct chromatin states on repressed genes. Curr. Biol. 21: 406-12. PubMed Citation: 21353562

Li, M., Ma, Z., Liu, J. K., Roy, S., Patel, S. K., Lane, D. C. and Cai, H. N. (2015). An organizational hub of developmentally regulated chromatin loops in the Drosophila Antennapedia complex. Mol Cell Biol 35(23):4018-29. PubMed ID: 26391952

Liang, Z. and Biggin, M. D. (1998). Eve and Ftz regulate a wide array of genes in blastoderm embryos: the selector homeoproteins directly or indirectly regulate most genes in Drosophila. Development 125(22): 4471-4482

Löhr, U., Yussa, M. and Pick, L. (2001). Drosophila fushi tarazu: a gene on the border of homeotic function. Curr. Biol. 11: 1403-1412. 11566098

Macias, A., Pelaz, S. and Morata, G., (1994). Genetic factors controlling the expression of the abdominal-A gene of Drosophila within its domain. Mech Dev 46: 15-25

Majumder, P., et al. (2009). Diverse transcription influences can be insulated by the Drosophila SF1 chromatin boundary. Nucleic Acids Res. 37(13): 4227-33. PubMed Citation: 19435880

Martinez, C. A., and Arnosti, D. N. (2008). Spreading of a corepressor linked to action of long-range repressor hairy. Mol. Cell. Biol. 28: 2792-2802. PubMed Citation: 18285466

Mullen, J. R. and DiNardo, S. (1995). Establishing parasegments in Drosophila embryos: roles of odd-skipped and naked genes. Dev. Biol. 169: 295-308

Muller, J. and Bienz, M. (1992). Sharp anterior boundary of homeotic gene expression conferred by the fushi tarazu protein. EMBO J 11: 3653-61

Nambu, P. A. and Nambu, J. R. (1996). The Drosophila fish-hook gene encodes a HMG domain protein essential for segmentation and CNS development. Development 122: 3467-75

Nasiadka, A. and Krause, H. M. (1999). Kinetic analysis of segmentation gene interactions in Drosophila embryos. Development 126: 1515-1526

Nasiadka, A., Grill, A. and Krause, H. M. (2000). Mechanisms regulating target gene selection by the homeodomain-containing protein Fushi tarazu. Development 127: 2965-2976.

Ohno, C. K., Ueda, H. and Petkovich, M. (1994) The Drosophila nuclear receptors FTZ-F1 alpha and FTZ-F1 beta compete as monomers for binding to a site in the fushi tarazu gene. Mol Cell Biol 14: 3166-75

Ohtsuki, S., Levine, M. and Cai, H. N. (1998). Different core promoters possess distinct regulatory activities in the Drosophila embryo. Genes Dev. 12: 547-556.

Okada, M. and Hirose, S. (1998). Chromatin remodeling mediated by Drosophila GAGA factor and ISWI activates fushi tarazu gene transcription in vitro. Mol. Cell. Biol. 18(5): 2455-2461

Park, J. M., et al. (2001). Drosophila mediator complex is broadly utilized by diverse gene-specific transcription factors at different types of core promoters. Mol. Cell. Bio. 21: 2312-2323. 11259581

Patel, S. R., Kim, D., Levitan, I. and Dressler, G. R. (2007). The BRCT-domain containing protein PTIP links PAX2 to a histone H3, lysine 4 methyltransferase complex. Dev. Cell 13: 580-592. PubMed Citation: 17925232

Pattatucci, A. M., Otteson, D. C. and Kaufman. T. C. (1990a). A functional and structural analysis of the Sex combs reduced locus of Drosophila melanogaster. Genetics 129: 423-41

Pattatucci, A. M. and Kaufman, T. C. (1990b). The homeotic gene Sex combs reduced of Drosophila melanogaster is differentially regulated in the embryonic and imaginal stages of development. Genetics 129: 443-61

Payankaulam, S., and Arnosti, D. N. (2009). Groucho corepressor functions as a cofactor for the Knirps short-range transcriptional repressor. Proc. Natl. Acad. Sci. 106: 17314-17319. PubMed Citation: 19805071

Pick, L., Schier, A., Affolter, M., Schmidt-Glenewinkel, T. and Gehring, W.J. (1990). Analysis of the ftz upstream element: germ layer-specific enhancers are independently autoregulated. Genes Dev. 4: 1224--1239

Poortinga, G., Watanabe, M. and Parkhurst, S. M. (1998). Drosophila CtBP: a Hairy-interacting protein required for embryonic segmentation and hairy-mediated transcriptional repression. EMBO J. 17(7): 2067-2078

Prazak, L., Fujioka. M. and Gergen, J. P. (2010). Non-additive interactions involving two distinct elements mediate sloppy-paired regulation by pair-rule transcription factors. Dev. Biol. 344: 1048-1059. PubMed Citation: 20435028

Pritchard, D.K. and Schubiger, G. (1996). Activation of transcription in Drosophila embryos is a gradual process mediated by the nucleocytoplasmic ratio. Genes Dev. 10: 1131-42. PubMed Citation: 8654928

Qi, D., Bergman, M., Aihara, H., Nibu, Y., and Mannervik, M. (2008). Drosophila Ebi mediates Snail-dependent transcriptional repression through HDAC3-induced histone deacetylation. EMBO J. 27: 898-909. PubMed Citation: 18309295

Riedl, A. and Jacobs-Lorena, M. (1996). Determinants of Drosophila fushi tarazu mRNA instability. Mol. Cell. Biol. 16: 3047-3053

Reinitz, J. and Levine, M. (1990). Control of the initiation of homeotic gene expression by the gap genes giant and tailless in Drosophila. Dev Biol 140: 57-72

Riedl, A. and Jacobs-Lorena, M. (1996). Determinants of Drosophila fushi tarazu mRNA instability. Mol. Cell Biol. 16: 3047-53.

Riley, G. R., et al. (1991). Positive and negative control of the Antennapedia promoter P2. Dev Suppl 1: 177-85

Russell, S. R. H., et al. (1996). The Dichaete gene of Drosophila melanogaster encodes a SOX-domain protein required for embryonic segmentation. Development 122: 3669-76

Schier, A. F. and Gehring, W. J. (1992). Direct homeodomain-DNA interaction in the autoregulation of the fushi tarazu gene. Nature 356: 804-7

Schroeder, M. D., Greer, C. and Gaul, U. (2011). How to make stripes: deciphering the transition from non-periodic to periodic patterns in Drosophila segmentation. Development 138(14): 3067-78. PubMed Citation: 21693522

Schwartz, C. J. E., et al. (2001). FTZ-Factor1 and Fushi tarazu interact via conserved nuclear receptor and coactivator motifs. EMBO J. 20: 510-519. 11157757

Skeath, J. B., et al. (1992). Gene regulation in two dimensions: the proneural achaete-scute genes are controlled by combinations of axis-patterning genes through a common intergenic control region. Genes and Dev. 6:2606-19.

Suzuki, T., et al. (2001). Segmentation gene product Fushi tarazu is an LXXLL motif-dependent coactivator for orphan receptor FTZ-F1. Proc. Natl. Acad. Sci. 98: 12403-12408. 11592991

Swantek, D. and Gergen, J. P. (2004). Ftz modulates Runt-dependent activation and repression of segment-polarity gene transcription. Development 131: 2281-2290. 15102703

Takemaru, Ki., et al. (1997). Multiprotein bridging factor 1 (MBF1) is an evolutionarily conserved transcriptional coactivator that connects a regulatory factor and TATA element-binding protein. Proc. Natl. Acad. Sci. 94(14): 7251-7256.

Tang, A. H., et al. (2001). Transcriptional regulation of cytoskeletal functions and segmentation by a novel maternal pair-rule gene, lilliputian. Development 128(5): 801-813. PubMed Citation: 11171404

Telford, M. J. (2000). Evidence for the derivation of the Drosophila fushi tarazu gene from a Hox gene orthologous to lophotrochozoan Lox5. Curr. Biol. 10: 349-352

Topol, J., et al. (1991). Synthetic oligonucleotides recreate Drosophila fushi tarazu zebra-stripe expression. Genes Dev. 5(5): 855-867

Tsai, C. and Gergen. P. (1995). Pair-rule expression of the Drosophila fushi tarazu gene: a nuclear receptor response element mediates the opposing regulatory effects of runt and hairy. Development 121: 453-462

Vander Zwan, C. J., et al. (2003). A DNA-binding-independent pathway of repression by the Drosophila Runt protein. Blood Cells Mol. Dis. 30: 207-222. PubMed Citation: 12732185

Vanderzwan-Butler, C. J., Prazak, L. M. and Gergen, J. P. (2006). The HMG-box protein Lilliputian is required for Runt-dependent activation of the pair-rule gene fushi-tarazu. Dev. Biol. 301(2): 350-60. PubMed Citation: 17137570

Wilkie, G. S. and Davis, I. (2001). Drosophila wingless and pair-rule transcripts localize apically by Dynein-mediated transport of RNA particles. Cell 105: 209-219. 11336671

Wilson, D. S., et al. (1996). Conservation and diversification in homeodomain-DNA interactions: A comparative genetic analysis. Proc. Natl. Acad. Sci. 93: 6886-91

Yokoshi, M., Segawa, K. and Fukaya, T. (2020). Visualizing the role of boundary elements in enhancer-promoter communication. Mol Cell 78(2):224-235. PubMed ID: 32109364

Yu, Y.and Pick, L. (1995). Non-periodic cues generate seven ftz stripes in the Drosophila embryo. Mech. Dev. 50: 163-175

Yu, Y., et al. (1997). The nuclear receptor Ftz-F1 is a cofactor for the Drosophila homeodomain protein Ftz. Nature 385: 552-555

Yu, Y., et al. (1999). A double interaction screen identifies positive and negative ftz gene regulators and Ftz-interacting proteins. Mech. Dev. 83(1-2): 95-105

Yussa, M., et al. (2001). The nuclear receptor Ftz-F1 and homeodomain protein Ftz interact through evolutionarily conserved protein domains. Mech. Dev. 39-53. 11520662

Zhao, K., et al. (1996). Binding of matrix attachment regions to nuclear lamin is mediated by the rod domain and depends on the lamin polymerization state. FEBS Lett. 380(1-2): 161-164

Zhao, C., et al. (2000). Target selectivity of Bicoid is dependent on nonconsensus site recognition and protein-protein interaction. Mol. Cell. Bio. 29: 8112-8123


fushi tarazu: Biological Overview | Evolutionary Homologs | Transcriptional regulation | Targets of activity | Protein interactions | mRNA Transport | Developmental Biology | Effects of Mutation

date revised: 15 April 2020

Home page: The Interactive Fly © 1997 Thomas B. Brody, Ph.D.

The Interactive Fly resides on the
Society for Developmental Biology's Web server.