pannier


DEVELOPMENTAL BIOLOGY

Embryonic

pnr is localized to the dorsal 30% of the blastoderm embryo. This is the region that gives rise to amnioserosa [Images] and dorsal epidermis. Zerknüllt and Decapentaplegic are also localized to the dorsal epidermis in a broader domain. Unlike ZEN and DPP, PNR is restricted to 20-60% (anterior to posterior) of the egg length. Expression is seen in five stripes, reflecting parasegmental units. The posterior boundary expands to the proctodeum during the early phase of germ band extension [Images]. Throughout Dorsal closure, pnr declines in expression and is confined to a band of single cells on either side of the dorsal midline (Winick, 1993).

Evidence for a fruit fly hemangioblast and similarities between lymph-gland hematopoiesis in fruit fly and mammal aorta-gonadal-mesonephros mesoderm

The Drosophila lymph gland is a hematopoietic organ and, together with prospective vascular cells (cardioblasts) and excretory cells (pericardial nephrocytes), arises from the cardiogenic mesoderm. Clonal analysis provided evidence for a hemangioblast that can give rise to two daughter cells: one that differentiates into heart or aorta and another that differentiates into blood. In addition, the GATA factor gene pannier (pnr) and the homeobox gene tinman (tin), which are controlled by the convergence of Decapentaplegic (Dpp), fibroblast growth factor (FGF), Wingless (Wg) and Notch signaling, are required for the development of all cardiogenic mesoderm, including the lymph gland. An essential genetic switch differentiates between the blood or nephrocyte and vascular lineages involves the Notch pathway. Further specification occurs through specific expression of the GATA factor Serpent (Srp) in the lymph-gland primordium. These findings suggest that there is a close parallel between the molecular mechanisms functioning in the Drosophila cardiogenic mesoderm and those functioning in the mammalian aorta-gonadal-mesonephros mesoderm (Mandal, 2004).

Blood and vascular cells in the vertebrate embryo are thought to derive from oligopotent progenitor cells, called hemangioblasts, that arise in the yolk sac and in the aorta-gonadal-mesonephros (AGM) mesenchyme. A close relationship between blood and vascular progenitors is well established, but in vivo evidence that a single cell can divide to produce a blood cell and an endothelial cell is lacking in vertebrate systems. Similarly, the molecular mechanism that distinguishes between the two lineages is not well understood. To address these issues in a simple, genetically amenable system, the genetic control of hematopoiesis was analyzed in Drosophila. The results show that there are close lineage relationships between hematopoietic and vascular cells, similar to those present in the AGM of mammalian systems. Evidence is provided for conserved cassettes of transcription factors and signaling cascades that limit the pool of hemangioblastic cells and promote the blood versus vascular fate (Mandal, 2004).

In the mature Drosophila embryo, the lymph gland is formed by a paired cluster of ~20 cells flanking the aorta. The aorta and heart represent a contractile tube lined by a layer of myoepithelial vascular cells called cardioblasts. The cells flanking the aorta and heart posterior to the lymph gland are the pericardial cells, which function as excretory cells (nephrocytes). Lymph gland progenitors express the prohemocyte marker Srp and ultrastructurally resemble prohemocytes that develop at an earlier stage from the head mesoderm. Monitoring expression of the zinc-finger protein Odd-skipped (Odd) shows that the lymph gland originates from the dorsal thoracic mesoderm. Odd is expressed in segmental clusters in the dorsal mesoderm of segments T1-A6. The three thoracic Odd-positive clusters coalesce to form the lymph gland, whereas the abdominal clusters formed the pericardial nephrocytes (Mandal, 2004).

Lymph-gland progenitors, cardioblasts and pericardial cells are closely related by lineage. Labeled 'flipout' (FLP/FRT) clones were induced in embryos aged 3-4 h such that the clones contained only 2-4 cells. Of the two-cell clones, ~50% contained cardioblast and lymph-gland cells; the other clones comprised either cardioblasts or lymph-gland cells alone. Mixed clones were recovered at the late third larval stage. The finding of mixed clones indicates that the cardiogenic mesoderm of D. melanogaster contains oligopotent progenitors that, up to the final division, can give rise both to Srp-positive blood-cell progenitors that form the lymph gland and to vascular cells (Mandal, 2004).

The cardiogenic mesoderm forms part of the dorsal mesoderm, which requires the homeobox protein Tin and the GATA factor Pnr. In embryos with mutations in tin or pnr, the lymph gland was absent. Maintenance of Tin expression in the dorsal mesoderm requires the activity of at least two signaling pathways regulated by Dpp (the Drosophila homolog of transforming growth factor-ß) and Heartless (Htl; one of the D. melanogaster homologs of the FGF receptor); the dependence of cardioblast and pericardial nephrocyte development on these signaling pathways has been documented. Lymph-gland progenitors did not develop in loss-of-function dpp and htl mutants (Mandal, 2004).

Between 6 h and 8 h of development, the dorsal mesoderm splits into the cardiogenic mesoderm and the visceral mesoderm. The cardiogenic mesoderm is regulated positively by Wg and negatively by Notch. Lack of Wg signaling results in the absence of all cardiogenic lineages including lymph gland. Notch signaling has the opposite effect and restricts cardiogenic mesodermal fate. Notch is active in the dorsal mesoderm from 6 h to 10 h of development. Eliminating Notch during the first half of this interval by raising embryos homozygous with respect to the temperature-sensitive allele Nts1 at the restrictive temperature resulted in substantially more cardioblasts, pericardial cells and lymph-gland progenitors (Mandal, 2004).

Lymph-gland progenitors, cardioblasts and pericardial nephrocytes are specified in the cardiogenic mesoderm around the phase of germband retraction 8-10 h after fertilization. At this stage, Tin, which was initially expressed in the whole cardiogenic mesoderm, becomes restricted to a narrow medial compartment containing the cardioblasts. Pnr follows the same restriction. Cells located at a more lateral level in the cardiogenic mesoderm give rise to lymph-gland progenitors (in the thoracic domain) and pericardial nephrocytes (in the abdominal domain) and activate the gene odd. Slightly later, Srp is expressed in lymph-gland progenitors. As reported for the early hemocytes derived from the embryonic head, srp is centrally involved in lymph-gland specification. In srp-null embryos, Odd-expressing cells still formed a lymph gland-shaped cluster flanking the aorta, but these cells also express the pericardial marker Pericardin (Prc), suggesting that they lose some aspects of hemocyte precursor identity or gain properties of nephrocytes. As a countercorrelate, ectopic expression of Srp in the whole cardiogenic mesoderm directed by mef2-Gal4 induces pericardial cells to adopt lymph-gland fate (Mandal, 2004).

Downregulation of tin and pnr in cells in the lateral domain of the cardiogenic mesoderm is essential for lymph-gland specification. Ectopic expression of tin or pnr by twist-Gal4 (or mef2-Gal4) causes a marked reduction in the number of lymph-gland and pericardial cells. The antagonistic effect of tin on lymph-gland progenitors resembles its earlier role in the head mesoderm that gives rise to the larval blood cells; here too, ectopic expression of tin causes a reduction in the number of hemocytes (Mandal, 2004).

Inhibiting tin and upregulating odd and srp requires input from the Notch signaling pathway. A function of Notch at 6-8 h in specification of the cardiogenic mesoderm is described. Reducing Notch function between 8 h and 10 h causes an increase in the number of cardioblasts and a concomitant loss of pericardial and lymph-gland cells. Overexpressing an activated Notch construct causes a marked increase in lymph-gland size. This late requirement for Notch signaling is separable from the earlier role of Notch in restricting the overall size of the cardiogenic mesoderm. Thus, the sum total of cardioblasts and pericardial or lymph-gland cells in Nts1 embryos shifts between 8 h and 10 h and does not differ substantially from that in wild type, whereas a combined effect on cell number and cell fate is seen in embryos with a Notch deletion. In these embryos, the cardiogenic mesoderm is hyperplasic and develops as cardioblasts at the expense of lymph-gland progenitors and pericardial nephrocytes. The dual role of Notch in restricting the numbers of a pluripotent progenitor pool and in distinguishing between the progeny of these progenitors is reminiscent of the function of Notch in sense-organ development (Mandal, 2004).

Lymph-gland formation is restricted to the thoracic region by positional cues that are provided by expression of the homeobox proteins of the Antennapedia and Bithorax complex. Specifically, Ultrabithorax (Ubx), which is expressed in segments A2-A5 of the cardiogenic mesoderm, inhibits lymph-gland formation. Loss of Ubx results in the expansion of the lymph-gland fate into the abdominal segments. Conversely, overexpression of Ubx driven by mef2-Gal4 causes the transformation of lymph-gland progenitors into pericardial nephrocytes (Mandal, 2004).

These findings are suggestive of a model of lymph-gland development in Drosophila that is similar to mammalian hematopoiesis. Lymph-gland progenitors develop as part of the cardiogenic mesoderm that also gives rise to the vascular cells (aorta and heart) and to excretory cells. Similarly, progenitor cells of the blood, aorta and excretory system are closely related both molecularly and developmentally in mammals, where they form part of the AGM. Specification of the cardiogenic mesoderm requires the input of FGF and Wg signaling, as in vertebrate hematopoiesis, where the AGM region is induced in response to several converging signaling pathways including FGF, BMP and Wnt (Mandal, 2004).

The cardiogenic mesoderm in Drosophila evolves from the dorsal mesoderm and requires input from the Htl, Dpp, Wg and Notch (N) signaling pathways. The cardiogenic mesoderm then differentiates into lymph gland, vascular cells (cardioblasts) and excretory cells (pericardial nephrocytes). A subpopulation of cardioblasts and lymph-gland cells is derived from one progenitor (hemangioblast; HB). Essential for the differentiation of the cardiogenic mesoderm is the Notch-Delta (Dl)-dependent restriction of Tin and Pnr to cardioblasts and the expression of Srp in the lymph gland. In vertebrates, similar cell types are derived from a mesodermal domain called the AGM, which also requires the input of FGF, BMP and Wnt signaling. A subset of AGM-derived cells has been proposed to constitute hemangioblasts, which produce blood progenitors and endothelial cells (Mandal, 2004).

These findings show that in Drosophila, the cardiovascular and blood-cell lineages are differentiated by an antagonistic relationship between Tin or Pnr expression in the cardioblasts and Srp expression in the lymph-gland progenitors. In vertebrates, GATA factors also have a pivotal role in specifying different lineages among blood-cell progenitors, although not much is known about what differentiates between blood progenitors as a group and endothelial progenitors. The results indicate that this step is driven by input from the Notch signaling pathway. In the thoracic cardiogenic mesoderm, Notch antagonizes tin and pnr expression and aortic cardioblast formation, and promotes srp expression and the development of lymph-gland progenitors. In vertebrates, Notch signaling is also involved in both blood and vascular development. The role of Notch during AGM morphogenesis remains to be investigated (Mandal, 2004).

Cardioblasts and lymph-gland cells can arise from the division of a single cardiogenic mesodermal cell, which should be called a hemangioblast. A previous study induced clones in the cardiogenic mesoderm but used only Tin as a marker. This study also yielded mixed two-cell clones comprising a cardioblast and a nonlabeled cell, which, in light of the current findings, must be interpreted as a lymph-gland cell. Hemangioblasts have been proposed in vertebrates, although the definitive experiment in which a precursor is marked and its lineage is tracked has not been done. Blast colony-forming cells that give rise to both lineages in vitro and common markers that belong to both cell types in vivo have been identified, but direct evidence for the existence of a common precursor has not yet been found. This study, using genetic analysis of two-cell clones, establishes the existence of such a population in Drosophila. On the basis of these results, and given the conservation of the signaling and transcriptional components described here, the prediction is that many cells of the AGM in vertebrates may give rise to only blood or only vascular cells, but a number of intermixed hemangioblasts may give rise to mixed lineages. Future genetic screens aimed at finding components in early lymph-gland development will probably identify additional pathways and strategies important for vertebrate hematopoiesis (Mandal, 2004).

Larval and Adult

pannier is expressed in a dorsal stripe in the adult thorax. In the thoracic disk pannier is expressed in a broad dorsal area, and in the dorsal half of the eye-antennal disc (Hietzler, 1996).

The Bar homeobox genes function as latitudinal prepattern genes in the developing Drosophila notum. In Drosophila notum, the expression of achaete-scute proneural genes and bristle formation have been shown to be regulated by putative prepattern genes expressed longitudinally. The two Bar locus genes may belong to a different class of prepattern genes expressed latitudinally: it is suggested that the developing notum consists of checker-square- like subdomains, each governed by a different combination of prepattern genes. BarH1 and BarH2 are coexpressed in the anterior-most notal region and regulate the formation of microchaetae within the region of BarH1/BarH2 expression through activating achaete-scute. Presutural macrochaetae formation also requires Bar gene activity. Bar gene expression is restricted in dorsal and posterior regions by Decapentaplegic signaling, while the ventral limit of the expression domain of Bar genes is determined by wingless, whose expression is under the control of Decapentaplegic signaling (Sato, 1999).

The Drosophila notum is considered genetically divided into several longitudinal, side by side, domains whose boundaries are determined by pannier, wingless and iroquois expression (listed respectively from medial to lateral). To further clarify relative locations of pnr, wg and iro expression areas, third-instar larval and pupal future notum were stained with various combinations of molecular markers. In larval and pupal future notum, pnr-Gal4 is expressed medially and iro-lacZ laterally. pnr-Gal4 and iro-lacZ domains partially overlap one another, and wg-lacZ (or Wg) expression is noted in the pnr-iro overlapping region and its immediate neighbors. Bar homeobox genes may belong to an additional class of notal subdivision genes. Staining for BarH1 indicates that BarH1 is expressed latitudinally (anterior vs. posterior) in the anterior-most region of future notum and postnotum. BarH1 expression begins at early to mid third instar. Anti-Ac antibody staining and neur-lacZ expression indicates PS macrochaetae are situated in the vicinity of posterior-ventral corners of the anterior BarH1 expression domain. BarH1 and BarH2 are referred to as Bar collectively and the anterior portion of the prescutum or its precursor expressing Bar is referred to as Bar prescutum. The Bar expression domain overlaps that of pnr, wg and iro. Bar expression similar to that in wing discs is observed in haltere discs (Sato, 1999).

It is concluded that a checker-board-like subdivision of future notum is regulated by putative prepattern gene expression. Future notum may be divided into square subdomains in a checker-board-like manner, each with its own unique combinations of prepattern gene expression. Putative prepattern genes, iro and pnr, form longitudinal domains. Bar homeobox genes form the anterior-most domain. This is the first demonstration of the presence of latitudinal, front to back, prepattern genes in the notum. Bristle formation in each subdomain may be positively regulated by a region-specific combination of prepattern genes. Consistent with this, microchaetae formation in the anterolateral prescutum (the lateral Bar prescutum), where Bar and iro are coexpressed, requires the concerted action of Bar and iro (Sato, 1999).

Effects of mutation or deletion

In mutants, cells of the amnioserosa and part of the dorsal epidermis die. This causes a lethal hole in the dorsal cuticle of the larva (Ramain, 1993).

A genetic and phenotypic analysis of the gene pannier is described. Animals mutant for strong alleles die as embryos in which the cells of the amnioserosa are prematurely lost. This leads to a dorsal cuticular hole. The dorsal-most cells of the imagos are also affected: viable mutants exhibit a cleft along the dorsal midline. Pannier mRNA accumulates specifically in the dorsal-most regions of the embryo and the imaginal discs. Viable mutants and mutant combinations also affect the thoracic and head bristle patterns in a complex fashion. Only those bristles within the area of expression of pannier are affected. A large number of alleles have been studied; they reveal that pannier may have opposing effects on the expression of achaete and scute leading to a loss or a gain of bristles. Certain alleles of pannier are sensitive to the dose of extra-machrochaetae. In those parts of the epithelium where both are present, it is possible that pnr and emc function together in the same ac-sc repression pathway (Heitzler, 1996).

The regulation of cardiac gene expression by GATA zinc finger transcription factors is well documented in vertebrates. However, genetic studies in mice have failed to demonstrate a function for these proteins in cardiomyocyte specification. In Drosophila, the existence of a cardiogenic GATA factor has been implicated through the analysis of a cardial cell enhancer of the muscle differentiation gene Mef2. The GATA gene pannier is expressed in the dorsal mesoderm and required for cardial cell formation while repressing a pericardial cell fate. Ectopic expression of Pannier results in cardial cell overproduction, while co-expression of Pannier and the homeodomain protein Tinman synergistically activate cardiac gene expression and induce cardial cells. The related GATA4 protein of mice likewise functions as a cardiogenic factor in Drosophila, demonstrating an evolutionarily conserved function between Pannier and GATA4 in heart development (Gajewski, 1999).

tinman gene function is required for heart development in Drosophila. The initial programming of the cardiac lineage occurs at a time when tin is broadly expressed in the dorsal mesoderm. A subset of the tin-expressing cells will become heart precursors, appearing in 11 clusters along the dorsalmost part of the mesoderm. The Mef2 enhancer-lacZ fusion gene marks heart progenitors at stage 11 and will eventually be expressed in four pairs of cardial cells per segment of the dorsal vessel. Thus, the tin expression domain is significantly larger than the territory of heart precursor specification, suggesting the involvement of additional factors in the formation of these cells. The Mef2 heart enhancer requires the presence of at least three elements for its activity, including two Tin binding sites and one GATA sequence. The GATA gene pnr is expressed in cells of the dorsal ectoderm around the time of heart cell specification. However, there is no report of pnr transcription in the mesoderm. To investigate this possibility, embryos were stained for PNR mRNA and embryo cross-sections were examined. At late stage 10, gene expression is observed in the dorsal ectoderm of the germband-extended embryo. PNR mRNA was detected in four clusters of cells located in the dorsalmost part of the mesoderm that corresponds to the cardiogenic region. Additionally, a pnr mesodermal enhancer has been identified that directs lacZ expression in the heart-forming region, but not in the overlying ectoderm. Therefore, pnr is expressed in the cardiogenic mesoderm where it could function in cardial cell specification and the regulation of Mef2 transcription (Gajewski, 1999).

The Mef2 IIA341 enhancer is active in both cardial and ventral muscle founder cells and a GATA site is needed for the cardial cell expression. To determine if pnr function is important for enhancer activity in either of these cell types, the expression of the enhancer-lacZ fusion was examined in pnr mutant embryos. A strong reduction of reporter gene expression is observed in the dorsal mesoderm of the mutants, while normal lacZ expression is detected in the ventral muscle precursors. These results suggest Pnr is a transcriptional regulator of the Mef2 heart enhancer, consistent with the observation that the protein can bind the essential GATA sequence in an electrophoretic mobility shift assay. Alternatively, or in addition, these results could be indicative of a requirement of pnr function for the formation of the Mef2-expressing cardial cells (Gajewski, 1999).

To determine if pnr function is essential for the specification of cardial cells, wild-type and mutant embryos were stained for Mef2 protein that serves as a marker for these cells. A distinct row of cardial cells is observed in a lateral view of normal embryos at stage 13 as they migrate dorsally along the overlying ectoderm during the process of dorsal vessel formation. In contrast, these cells are greatly diminished or completely absent from the dorsal mesoderm of mutant embryos. To assess the formation of the pericardial cells in the same genetic background, embryos were stained for Eve protein, which is a marker for a subset of these cells. Eleven clusters of Eve-positive cells, each comprising three or four cells, are detected in wild-type embryos at stage 12. In contrast, an overabundance of Eve-expressing pericardial cells is observed in the pnr embryos. These results suggest that pnr function is vital to the formation of cardial cells, while simultaneously playing an important role in controlling the production of at least one pericardial cell type. Thus, forced mesodermal expression of Pnr leads to an overproduction of cardial cells and the generation of supernumerary cardial cells occurs at the expense of Eve pericardial cells (Gajewski, 1999).

Certain NK-2 class homeodomain and GATA family proteins have been shown to physically interact in their cooperative activation of gene expression in cell culture systems. To test the possibility that Tin and GATA factors could functionally interact in an embryological context, the tin, pnr and mGATA4 genes were expressed independently or in combination in Drosophila embryos. When tin, pnr or mGATA4 are expressed alone in the twi enhancer-expressing cells, the Mef2 heart enhancer is activated ectopically in the cephalic (tin) or dorsal (pnr and mGATA4) mesoderm. Since Tin is a known regulator of the Mef2 enhancer, it could be activating the Mef2 sequence in the head region through its fortuitous interaction with a co-factor normally expressed in these cells. The results are striking when both Tin and either of the GATA factors are co-expressed under the control of the twi-Gal4 driver. A cardial cell marker is now activated in both the cephalic region and throughout the dorsal and ventral trunk mesoderm. Likewise, a strong ectopic expression of the Mef2 heart enhancer is detected in ventral midline cells of the developing CNS. The data point to a combinatorial interaction of Tin and the two GATA factors in the de novo activation of the cardial cell marker in both mesodermal and non-mesodermal cells. They also suggest these genetic combinations are inducing a cardial cell fate along the ventral midline of the CNS (Gajewski, 1999).

In summary, the discovery of early heart phenotypes in pnr mutant embryos, coupled with the demonstration of uniquely conserved cardiogenic abilities of Pnr and GATA4, provide novel evidence for the function of GATA family members in the specification of a heart cell type. In an embryological context, these proteins can work with the Tin homeodomain factor to program cells into an apparent cardial fate in both mesodermal and non-mesodermal cell types. This genetic combination appears to be essential, but not necessarily sufficient, for cellular commitment to the cardiac lineage as other factors may contribute to the specification process. Additional studies using the Drosophila cardiogenic assay should prove instrumental in revealing other key members of this genetic program (Gajewski, 1999)

The pannier gene encodes a GATA transcription factor and acts in several developmental processes in Drosophila, including embryonic dorsal closure, specification of cardiac cells and bristle determination. pnr is expressed in the mediodorsal parts of thoracic and abdominal segments of embryos, larvae and adult flies. Its activity confers cells with specific adhesion properties that make them immiscible with non-expressing cells. Thus there are two genetic domains in the dorsal region of each segment: a medial (MED) region where pnr is expressed and a lateral (LAT) region where it is not. The homeobox gene iroquois is expressed in the LAT region. These regions are not formed by separate polyclones of cells, but are defined topographically (Calleja, 2000).

The domain of expression of pnr in adult flies is seen directly in pnr-Gal4/UAS-y flies or in flies or imaginal discs of the genotype pnr-Gal4/UAS-lacZ. It is restricted to a dorsal region of the head, thorax and abdomen. In the head, pnr is expressed in the dorsal region of the eye and in the head capsule and will not be considered further. In the dorsal mesothorax pnr covers a longitudinal band that occupies about 40% of the notum and extends from the dorsal midline to the medial zone. It is delineated by a straight border that is aligned with, and just lateral to, the dorsocentral bristles. In the wing disc, pnr is expressed in the region that contains the progenitor cells of the medial region of the adult notum. iro is expressed in both the wing and the thorax, but only the thoracic expression is relevant here. In the second larval instar iro is expressed in all prospective thoracic cells but, in third instar discs, it is restricted to the lateral region, indicating that there is a retraction of the iro thoracic domain during disc development. The expression of pnr and iro were compared in the thorax of late third instar discs: they are expressed in distinct and complementary subdomains that together cover the entire mesothorax. Their expression extends to both the A (notum) and the P (postnotum) mesothoracic compartments as indicated by double staining for pnr and en or iro and en. The domain of wg in the mesothorax overlaps with that of pnr and iro but is restricted to the A compartment. The expression of pnr, iro, wg and en in the haltere disc is the same as in the wing disc, indicating that the two discs share the same genetic organisation. In the abdominal segments, pnr defines a medial subdomain that occupies about 35% of each tergite. Unlike the notum, there is no morphological landmark close to the pnr expression boundary and the bristle and pigment patterns are similar on either side (Calleja, 2000).

Ectopic pnr in the wing induces MED thoracic development, indicating that pnr specifies the identity of the MED regions. Correspondingly, when pnr is removed from clones of cells in the MED domain, they sort out and apparently adopt the LAT fate. It is proposed that (1) the subdivision into MED and LAT regions is a general feature of the Drosophila body plan and (2) pnr is the principal gene responsible for this subdivision. It is argued that pnr acts like a classical selector gene but differs in that its expression is not propagated through cell divisions (Calleja, 2000).

The finding that pnr negatively regulates iro suggests that the restriction of iro expression to the LAT subdomain, and hence the appearance of two distinct subdomains, is a result of pnr activity in the MED region. In cases such as ap-gal4/UAS-pnr flies in which the restricted activity of pnr is replaced by uniform expression in all presumptive notal cells, the subdivision of the notum does not occur, instead only the MED pattern develops. Thus restriction of pnr expression to the MED region is a prerequisite for the partitioning of the notum (Calleja, 2000).

In addition to its function in partitioning, pnr also specifies the pattern of the MED region. It participates together with selector genes in a 'genetic address' that determines the identity of MED regions of the thoracic and abdominal segments. In the notum, the activity of pnr is necessary for the development of the characteristic pattern of the MED region, whereas its absence allows expression of iro and consequently formation of the LAT pattern. In the abdomen, pnr is also required for the development of the MED region. The developmental capacity of pnr and the binary mechanism in which it participates are clearly demonstrated by the ectopic expression experiments that show that Pnr induces different patterns depending on the local genetic context. In the wing blade, Pnr induces the formation of the MED notum in the anterior wing and MED postnotum in the posterior wing. This result suggests that en is contributing together with pnr to the genetic address of the MED regions. The DV compartmental segregation is not recognized by pnr; clones of pnr-expressing cells in the dorsal and ventral wing compartments differentiate the same MED-like pattern. But this is not so surprising: no DV segregation takes place within the normal domain of pnr expression (Calleja, 2000).

In the haltere disc, the response to pnr activity may be modulated by the function of Ultrabithorax, the selector gene that discriminates between haltere and wing development. In the abdomen, ectopic pnr expression in the ventral region (sternites) induces the development of a dorsal pattern (tergite). Here the response to pnr function is probably modulated by the corresponding genes of the BX-C. This transformation appears to be similar to that described for ectopic wg activity in the ventral abdomen. It is possible that Pnr may derepress wg or expand the domain of wg in the sternites and pleurae. Alternatively, the transformations induced by wg may be mediated by activation of pnr. Other features of pnr are reminiscent of classical selector genes. For example, the effect of pnr on cell affinities is a property shared by selector genes and is used to keep groups of cells from mixing during growth (Calleja, 2000).

Thus the mode of action of pnr is very similar to that of selector genes like Ubx, en or ap. The only difference between the segregation of the MED and LAT subdomains and the AP or the DV compartment segregations is the manner whereby activity of the genes is maintained. Expression of en or ap is inherited unchanged by the cellular progeny, but this is not so in the case of pnr or iro. The really critical outcome is thought to be the partitioning of groups of cells into distinct genetic subdomains. The manner in which the genes responsible for segregating the genetic subdomains, be these en, ap or pnr, maintain their activity may be of secondary importance. If pnr expression is not inherited by the cell progeny, there has to be some other mechanism to maintain activity in the appropriate region. pnr is likely to respond to threshold levels of some specific signal(s) which would probably emanate from either the midline or from the lateral margin of the dorsal field (Calleja, 2000).

pnr and iro are not the only developmental genes that define the identity of body regions by a mechanism not based on cell lineage. In the leg disc, the distinction between proximal and distal regions results from the genetic interface between Dll and hth-exd, which determines the development of the appendage in the proximodistal axis. This genetic border is not based on lineage and, as in the case of pnr in the notum, Dll activity confers specific adhesion properties to cells causing them to sort out from non-expressing cells. The distinction between external and internal analia is based on the differential activity of Dll within the domain of the Hox gene caudal, and this too is lineage independent. Recent work on the omb domain in the wing indicates that omb activity is not inherited by cells. As in the case of pnr, it has been shown that Dll and omb can induce appendage structures if expressed ectopically, suggesting that their products determine the identity of different body regions. In the abdomen, the A compartment is subdivided into anterior and posterior domains. These domains respond to Hedgehog in very different ways and yet the border between them is not colinear with a lineage boundary (Calleja, 2000 and references therein).

The process of compartmentation is an epigenetic mechanism by which groups of cells become geographically divided into subgroups that acquire characteristic and distinct genetic identities. Although compartmentation is normally associated with cell lineage segregations, the results presented here indicate that such segregations are not the defining feature of the process. The MED and LAT subdomains that are reported here are not segregated by cell lineage and yet they possess all other features associated with compartments: they originate by subdivision of groups of presumptive cells, are delimited by sharp boundaries, and are genetically specified by the combinatorial activity of a set of selector genes. The subdivision of the leg into distinct genetic domains or that of the analia may be other examples of genetic partitions without the concourse of lineage segregations. The principal difference with the MED/LAT subdivision is that the latter affects embryonic, larval as well as adult segments and is therefore a basic feature of the body plan (Calleja, 2000).

Multitype zinc-finger proteins of the class Friend of GATA/U-shaped (Ush) are known to function as transcriptional regulators of gene expression through their modulation of GATA factor activity. To better understand intrinsic properties of these proteins, the expression and function of the ush gene during Drosophila embryogenesis was investigated. ush is dynamically expressed in the embryo, including several cell types present within the mesoderm. The gene is active in the cardiogenic mesoderm, and a loss of function results in an overproduction of both cardial and pericardial cells, indicating a requirement for the gene in the formation of these distinct cardiac cell types. Conversely, ectopic expression of ush results in a decrease in the number of cardioblasts in the heart and the inhibition of a cardial cell enhancer normally regulated by the synergistic activity of the Pannier and Tinman cardiogenic factors. These findings suggest that, similar to its known function in thoracic bristle patterning, Ush functions in the control of heart cell specification through its modulation of Pannier transcriptional activity. ush is also required for mesodermal cell migration early in embryogenesis, where it shows a genetic interaction with the Heartless fibroblast growth factor receptor gene. Taken together, these results demonstrate a critical role for the Ush transcriptional regulator in several diverse processes of mesoderm differentiation and heart formation (Fossett, 2000).

The ush gene exhibits a dynamic pattern of expression during embryogenesis. Gene transcripts are first detected at high levels in the primordium of the amnioserosa at stage 5. Additional expression is observed in germ band extending embryos, in cells of the developing anterior and posterior midgut, and in hemocyte precursors present in the cephalic mesoderm. By stage 11, ush RNA is detected in the dorsal ectoderm and in precursor cells of the hemocytes and fat body. By late embryogenesis, ush expression is greatly diminished, but transcripts are still observed in the dorsal ectoderm during dorsal closure and cells within, or associated with, the central nervous system. To investigate the possible expression of ush in mesodermal cells underlying the dorsal ectoderm, cross sections of embryos at stage 11 were examined. ush RNA is detected in a changing pattern in this germ layer, initially throughout most of the mesoderm and then in subpopulations of cells, including precursors of the fat body, visceral mesoderm, and cardiogenic mesoderm. Therefore, ush is expressed in the dorsal mesoderm, where it could function in the early stages of heart formation (Fossett, 2000).

pnr mutant embryos show a loss of contractile cardial cells and an overproduction of certain nonmuscle pericardial cells in the heart-forming region. To identify a possible role for ush in these cardiogenic processes, alterations in cardiac cell production were sought in mutant embryos. The ush alleles used in this analysis were ush1 and ush2, believed to be hypomorphic mutations of the gene, and Df(2L)al, a chromosome deletion that represents a ush null mutation. The D-mef2 heart enhancer-lacZ fusion gene serves as a cardial cell marker, since it is detected in progenitors of these cells around stage 11 and thereafter in two, then four cardioblasts per hemisegment of the forming dorsal vessel. Embryos homozygous for either a point mutation or deletion of the gene show an increase in the number of cells expressing the reporter gene, as compared with the wild-type embryo. In ush1 embryos, a few hemisegments contain up to nine positive cells with an average of six cardial cells present in many clusters. In ush-deficiency embryos, a comparable increased density of cardial cells is found (Fossett, 2000).

Mef2 protein also marks cardioblasts; it is detected in the nuclei of all cardial, but not pericardial cells of the forming dorsal vessel. In wild-type embryos at stage 13, the germ band has retracted with cardioblasts migrating dorsally, separating from the dorsal somatic muscles. A lateral view at this stage shows a single row of cells that contains six stained nuclei per hemisegment. In contrast, ush mutant embryos possess supernumerary cardioblast nuclei. ush1 and ush2 embryos contain up to 12 nuclei per hemisegment with eight cells per cluster observed on average. Similar results have also been obtained with ush-deficiency embryos. Therefore, reducing or completely eliminating ush function leads to an increased production of cardial cells. Intriguingly, the ush heart phenotype uncovered by the analysis of these two markers directly contrasts the absence of cardial cells observed in pnr loss-of-function embryos (Fossett, 2000).

Production of pericardial cells was quantitated in ush mutant embryos, using Eve protein as a marker for a subset of these cells. In wild-type embryos at stage 12, there exist 11 Eve-positive clusters within the dorsal mesoderm, each containing about three cells. In contrast, the number of Eve-expressing pericardial cells increases in homozygous ush1 and ush2 embryos to an average of 5-6 per cluster. A similar increase in pericardial cell number is also observed in homozygous Df(2L)al embryos. Thus, ush gene activity is required to prevent the overproduction of this pericardial cell type, a function that has also been ascribed to the pnr gene (Fossett, 2000).

Because the loss of ush function resulted in a supernumerary cardial cell phenotype, the effect of expressing the gene throughout the mesoderm was monitored using the Gal4/UAS binary system. Mef2 was used to assess the status of cardial cells, with two contiguous rows of 52 cells present in the forming or mature dorsal vessel of wild-type embryos. In comparably staged embryos expressing ush throughout the mesoderm, a significant reduction in cardial cells is observed. The D-mef2 heart enhancer-lacZ fusion gene was used as a second marker for the cardial cells and also to assay the effect of ush expression on enhancer activity. In wild-type embryos at stage 16, the enhancer is active in eight cardial cells in most segments of the dorsal vessel. In contrast, beta-galactosidase activity is greatly diminished in the hearts of ush-expressing embryos, most likely a combination of the decrease in cardial cell number and the reduced activity of the D-mef2 cardiac enhancer. Thus, forced expression of ush has a potent negative effect on cardial cell formation and enhancer function (Fossett, 2000).

It has been shown that Pnr can function combinatorially with Tin in the regulation of the D-mef2 heart enhancer in Drosophila embryos. This synergistic activation was examined using a transient transfection assay in cultured Drosophila cells. The activation of a CAT reporter gene linked to the D-mef2 enhancer was monitored in cells transfected independently with Pnr, Tin, and Ush or with various combinations of the factors. The expression of Tin alone activated the enhancer about 2-fold above the basal level, whereas neither Pnr nor Ush affected enhancer activity. Coexpression of Pnr and Tin resulted in a synergistic activation of the enhancer to a level 5-6 times that of the basal activity, and this strong induction requires the binding of Tin to the Mef2 enhancer; a Tin DNA binding mutant, Tin (N-Q), failed to synergize with Pnr in the assay. In contrast, adding Ush as a third transfected factor significantly inhibits the synergistic activation of the enhancer by Pnr and Tin. This result demonstrates that Ush can antagonize the positive functional interaction of Pnr and Tin in the regulation of the cardial cell enhancer, which is consistent with the in vivo data (Fossett, 2000).

ush mutants contain an increased number of cardial cells. However, in about half of the embryos a disparity was noticed in the cardial cell populations, ranging from a high of 8-12 per hemisegment down to regions completely devoid of cells. This complex phenotype is observed with both ush hypomorphic and null alleles. This sporadic loss of cells from the dorsal-most part of the mesoderm is reminiscent of a htl mutant phenotype, where the absence of the encoded fibroblast growth factor receptor homolog results in an incomplete dorsal migration of mesodermal cells. In this event, cells fail to receive the ectodermal signal needed for their further commitment, resulting in a loss of dorsal mesodermal derivatives, including cardioblasts (Fossett, 2000).

To determine whether the variable absence of cardial cells in ush embryos is because of a cell migration defect, wild-type and mutant embryos were stained for Mef2 protein present in the invaginated population of mesodermal cells. In cross sections of normal embryos at stage 10, a uniform layer is observed where the dorsal-most mesodermal cells have migrated to a position adjacent to the dorsal-most ectodermal cells. In contrast, both htl and ush homozygous embryos display an irregular layer where mesodermal cells remain clustered and fail to undergo their complete dorsal migration. This result suggests ush function is required for the directional migration of the mesoderm. To investigate a potential genetic interaction of htl and ush in this process, embryos were examined that were heterozygous for mutations in each of the genes. These embryos also present a strong mesodermal migration phenotype, suggesting the two genes function in a common genetic pathway that controls this aspect of mesoderm differentiation. As was observed with homozygous ush embryos, slightly less than half of the transheterozygous embryos show a loss of cells from the dorsal mesoderm (Fossett, 2000).

It is thus thought that pnr has a dual requirement in the cardiogenic mesoderm because it is needed for the formation of cardial cells although simultaneously limiting the production of Eve-expressing pericardial cells. Based on these dissimilar phenotypes, it is postulated that Pnr might work with different combinations of factors to promote or repress the formation of cells within the distinct lineages. Recent studies have shown that Pnr and Tin act synergistically to induce cardial cells and activate gene expression, and the loss of function of either of these genes results in an absence of cardial cells. Therefore, the two work together in cardial cell specification (Fossett, 2000).

In contrast, Ush is a factor that normally suppresses cardial cell production. ush homozygous and hemizygous embryos show an increase in cardial cell number: the latter finding suggests Ush control of this cell population is dose-dependent, as is the case for Ush regulation of Pnr during sensory bristle development. Furthermore, forced expression of Ush decreases cardial cell production and D-mef2 heart enhancer activity, whereas ectopic expression of Pnr produces extra cardial cells and expands the domain of enhancer activity. Thus, Ush displays phenotypes that are in direct opposition to those of Pnr, suggesting that it can function as an antagonist of Pnr's cardiogenic activity. This conclusion is supported by the ability of Ush to inhibit the synergy of Pnr and Tin in the activation of the D-mef2 heart enhancer in cell transfection studies. As for the second cardiac phenotype, both pnr and ush are required to limit the number of Eve-expressing pericardial cells, consistent with a model in which Ush and Pnr function as corepressors in the control of these cells. To summarize, these genetic studies predict that, in the wild-type embryo, pnr is expressed and functions independent of ush in precursors of the cardioblast lineage. However, in neighboring cells that include the Eve lineage precursors, the expression and function of the two most likely overlaps. Future expression analyses of the two regulatory proteins at the resolution of single mesodermal cells will be required to elaborate on this genetic model in molecular terms (Fossett, 2000).

The GATA factor Pannier (Pnr) is required for eye and heart development in Drosophila. When U-shaped (Ush), FOG-1, FOG-2, or mutant FOG-2 is coexpressed with Pnr during these developmental processes, severe eye and heart phenotypes result, consistent with a conserved negative regulation of Pnr function. Ush appears to negatively regulate the cardiogenic function of the GATA-4 homolog Pnr, converting Pnr from a transcriptional activator to a repressor as observed during sensory bristle development. As with Ush, forced mesodermal expression of FOG-1, FOG-2, and DeltaFOG-2 (lacking a conserved motif that binds the corepressor C-terminal binding protein) also produces a diminution of cardial cells. These results demonstrate a functional conservation of the FOG proteins during Drosophila cardiogenesis, which most likely involves negative regulation of the cardiogenic activity of Pnr. In addition, forced expression of FOG proteins disrupts eye development-producing phenotypes that mimic pnr loss of function mutants, presumably by repressing Pnr activation of its downstream effector genes (Fossett, 2001).

Pannier is required to establish competence for heart progenitor formation

Inductive signaling is of pivotal importance in order for developmental patterns to form. In Drosophila, the transfer of TGFß (Dpp) and Wnt (Wg) signaling information from the ectoderm to the underlying mesoderm induces cardiac-specific differentiation in the presence of Tinman, a mesoderm-specific homeobox transcription factor. Evidence that the Gata transcription factor, Pannier, and its binding partner U-shaped, also a zinc-finger protein, cooperate in the process of heart development. Loss-of-function and germ layer-specific rescue experiments suggest that pannier provides an essential function in the mesoderm for initiation of cardiac-specific expression of tinman and for specification of the heart primordium. u-shaped also promotes heart development, but unlike pannier, does so only by maintaining tinman expression in the cardiogenic region. By contrast, pan-mesodermal overexpression of pannier ectopically expands tinman expression, whereas overexpression of u-shaped inhibits cardiogenesis. Both factors are also required for maintaining dpp expression after germ band retraction in the dorsal ectoderm. Thus, it is proposed that Pannier mediates as well as maintains the cardiogenic Dpp signal. In support, it is found that manipulation of pannier activity in either germ layer affects cardiac specification, suggesting that its function is required in both the mesoderm and the ectoderm (Klinedinst, 2003).

pnr and ush are both expressed in the mesoderm at the time of cardiac mesoderm formation, in addition to their expression in the dorsal ectoderm. Mesodermal expression of pnr is restricted to the dorsal cardiogenic margin, whereas ush extends more laterally. In order to assess the requirement for pnr and ush in initiating cardiac mesoderm and cardiac cell type-specific differentiation, tin expression was examined at progressively later developmental stages in null mutants for both pnr and ush. During mid-stage 11, tin is expressed segmentally in two regions of the mesoderm. The dorsal clusters of cells correspond to the cardiac precursor cells, whereas the lateral clusters will become part of the visceral mesoderm. In same stage pnr mutant embryos, tin expression is dramatically reduced in the clusters that correspond to the cardiac precursors, indicating that cardiogenesis is not being initiated. tin expression in the visceral mesodermal clusters, as well as tin expression earlier in development, is unaffected, suggesting the heart is a focal point for pnr function, which is consistent with its cardiac-restricted expression in the mesoderm. By contrast, ush mutant embryos initially seem to exhibit normal tin expression. At later stages, when tin expression is solely restricted to the heart cells, ush mutants display a progressively more severe reduction in tin expression, approaching the phenotype of pnr mutants. Thus, both pnr and ush are required for heart-specific tin expression, although ush seems to be initially dispensable (Klinedinst, 2003).

Even though tin is initially expressed in all heart progenitors, its expression is later turned off in some specific lineages, but continues to be expressed in many myocardial and pericardial cells. To determine which heart cells are affected in pnr and ush mutants, mutant embryos were examined with various markers. eve, for example, is co-expressed with tin in 11 clusters of heart progenitors, and these lineages give rise to a subset of pericardial cells. eve expression is only moderately reduced in pnr and hardly at all in ush mutants at early as well as later stages; this is accompanied by patterning defects at progressively later stages. By contrast, the lbe-expressing heart progenitors, which produce both myocardial and pericardial cells, are dramatically reduced in pnr but less so in ush mutants. Moreover, the svp-expressing cells, which also give rise to a mixed lineage, but cease to co-express tin at later stages, are dramatically reduced in both mutants. Thus, all lineage markers assayed are reduced in both mutants, but each is affected with disproportional severity, which is consistent with the idea that the formation of each cell type has a direct requirement for pnr and ush (Klinedinst, 2003).

Both tin and pnr have been shown to be targets of Dpp signaling at stage 9/10. It is proposed that dpp is necessary again at stage 11 to activate and maintain pnr and tin expression in the cardiogenic region of the mesoderm. First, pnr is activated with the help of early stage 11 tin, which is expressed broadly throughout the dorsal mesoderm, and dpp, which is expressed in a narrow dorsal ectodermal stripe. Then, at mid-stage 11, tin is restricted to the cardiogenic region with the help of mesodermal pnr as well as continuous ectodermal Dpp signaling. Once both are activated in the cardiogenic mesoderm, they are likely to contribute to the maintenance of each other's expression, probably aided again, but only moderately, by ectodermal Dpp signaling. This interpretation is consistent with mesodermal versus ectodermal expression of dominant-negative pnrEnR and the dpp target repressor encoded by brk. They are both equally effective in reducing cardiac-specific tin when expressed in the mesoderm, but ectodermal repression is more effective when dorsal-stripe dpp at stage 11 is also affected (as in the case of ZKr-Gal4>UAS-brk, but not with ZKr-Gal4>UAS-pnrEnR) (Klinedinst, 2003).

Mesodermal overexpression of ush and co-overexpression with pnr results in a decrease in the amount of cardiac-specific tin expression, suggesting that ush may not only be required along with pnr for heart development, but also play an inhibitory role. To test this hypothesis further, pnrD4, an allele that abolishes Ush binding to Pnr was overexpressed; not only ectopic tin expression was found at early stages of cardiogenesis, but also undiminished and even increased levels of expression at later stages. A similar phenotype was observed when both pnrD4 and ush were expressed throughout the mesoderm, suggesting that ush plays an anti-cardiogenic role by antagonizing the activity of wild-type Pnr, but not that of PnrD4. It would be interesting to see if pan-mesodermal overexpression of wild-type pnr in a ush mutant background results in ectopic tin expression similar to pnrD4, or if a minimal amount of ush activity is required to maintain normal and ectopic tin expression even with forced pnr expression. Interestingly, overexpression of both pnr and tin together in the mesoderm also causes a pnrD4-like phenotype, as assayed with Hand expression, suggesting that pnr and tin collaborate during initiation and subsequent differentiation of the heart progenitors (Klinedinst, 2003).

Although in vitro the Ush-related FOG factors are primarily known for their role as transcriptional repressors, they apparently can also function as co-activators: Fog2 can synergistically activate or repress the transcriptional activity of Gata4, depending on the (cardiac) promoter and cell line used, and FOG-1 can cooperate with Gata1 to transactivate NF-E2, an erythroid cell-specific promoter. Moreover, the ventricular hypoplasia and other heart defects observed in Fog2-deficient mice suggest a deficit rather than an excess in heart development. In addition, mice with an equivalent mutation to PnrD4 knocked into the Gata4 locus, thus eliminating binding to Fog2, exhibit in many ways a similar phenotype to Fog2-deficient mice. These data are consistent with the idea that Fog2 is normally involved in promoting rather than antagonizing cardiogenesis, similar to what was found with genetic studies during Drosophila heart development (Klinedinst, 2003).

The dual role of Ush suggests that the amount of Ush may be crucial for whether it exerts its function as a an activator or repressor, perhaps by binding to different sets of co-factors in a concentration-dependent manner. Alternatively, the mode of transcriptional regulation by Ush could be stage-dependent: at stage 11, Pnr and Ush cooperate as transcriptional activators in initiating cardiac-specific tin expression and heart development, but later Ush becomes a repressor to limit the transcriptional activation of tin by Pnr (Klinedinst, 2003).

pnr and ush are initially broadly expressed in the dorsal ectoderm of the early embryo, but by germband retraction the ectodermal expression of pnr is confined to a narrow stripe of cells along the border of the amnioserosa, which overlaps with the thin dorsal dpp stripe. The early ectodermal expression of ush is restricted to the presumptive amnioserosa, and by germband extension, ush also overlaps with the dorsalmost region of the ectoderm. These patterns of expression suggest that pnr and ush may be acting in both germ layers. The genetic data, including germ layer-specific expression of wild-type and dominant-negative pnr constructs, as well as germ layer-specific rescue experiments suggest strongly that pnr and ush function is not only needed in the mesoderm, but also in the ectoderm for heart formation. The ectodermal requirement for pnr and ush in heart development is probably achieved via the maintenance of dpp expression, since dorsal stripe dpp expression diminishes in pnr and ush mutants and ectodermal interference with pnr, ush and/or dpp-signaling function compromises the normal progression of heart development (Klinedinst, 2003).

pannier and pointedP2 act sequentially to regulate heart development

The Drosophila heart consists of two major cell types: cardioblasts, which form the contractile tube of the heart; and pericardial cells, which flank the cardioblasts and are thought to filter and detoxify the blood or hemolymph of the fly. This study presents the completion of the entire cell lineage of all heart cells. Notably, a previously unappreciated distinction has been detected between the lineages of heart cells located in the posterior seven segments relative to those located more anteriorly. Using a genetic screen, the ETS-transcription factor pointed has been detected as a key regulator of cardioblast and pericardial cell fates in the posterior seven segments of the heart. In this domain, pointed promotes pericardial cell development and opposes cardioblast development. This function of pointed is carried out primarily if not exclusively by the pointedP2 isoform -- in this context, pointedP2 may act independently of Ras/MAPK pathway activity. The GATA transcription factor pannier acts early in dorsal mesoderm development to promote the development of the cardiac mesoderm and thus all heart cells. Finally, it is demonstrated that pannier acts upstream of pointed in a developmental pathway in which pannier promotes cardiac mesoderm formation, and pointed acts subsequently in this domain to distinguish between cardioblast and pericardial cell fates (Alvarez, 2003).

Pioneering work in C. elegans established the importance of elucidating cell lineages to obtain a thorough understanding of animal development. Prior work established the cell lineage of 10 out of the 16 heart cells that arise in each hemisegment of the posterior seven heart segments. To define the lineage of the remaining heart cells, the FLP/FRT lineage tracing system was used to determine the lineal relationship of the two Eve-positive pericardial cells and four Tin-positive pericardial cells that arise in each hemisegment. This system creates random clones marked by tau-lacZ reporter gene activity. Briefly, clones were induced during stage 8 just as the pan-mesodermal divisions are being completed. This allowed clones to be induced in mesodermal cells prior to the emergence of heart precursors. To identify the lineage of Eve-positive pericardial cells, embryos were double labeled for ß-galactosidase to mark clones, and with Eve to identify Eve-positive pericardial cells. To identify the lineage of Tin-positive pericardial cells, embryos were double labeled for ß-galactosidase to mark clones, and Tin to identify Tin-positive pericardial cells. In addition to the Tin-positive pericardial cells, Tin labels cardioblasts and Eve-positive pericardial cells. However, based on position and morphology one can unambiguously distinguish Tin-positive pericardial cells from Tin-positive cardioblasts and Eve-positive pericardial cells (Alvarez, 2003).

Eighteen clones were identified that contained at least one Eve-positive pericardial cell. Eleven of these clones (61.1%) consisted solely of two Eve-positive pericardial cells, six clones consisted of two Eve-positive pericardial cells and one or two nearby heart or other mesodermal cells, and one clone consisted of a single Eve-positive pericardial cell. Thus, when one Eve-positive pericardial cell is observed within a clone of two or more cells a second Eve-positive pericardial cell always exists within this clone. These data demonstrate that the two Eve-positive pericardial cells within a hemisegment are siblings and arise from an Eve-positive pericardial cell precursor (Alvarez, 2003).

Tin-positive pericardial cell clones fall into two classes: those that contained two Tin-positive pericardial cells, and those that contained one Tin-positive pericardial cell and one cardioblast. These two classes of clones arise in mutually exclusive regions of the heart. Clones that contain two Tin-positive pericardial cells arise in the posterior seven segments of the heart (this region is referred to as the posterior heart domain), whereas clones that contain one Tin-positive pericardial cell and one cardioblast arise anterior to this domain (this region is referred to as the anterior heart domain). The point of demarcation between these clonal types coincides precisely with the location of the first pair of Svp-positive cardioblasts. These data demonstrate that heart cells exhibit distinct cell lineages as a function of position along the anteroposterior axis (Alvarez, 2003).

A total of 24 Tin-positive pericardial cell clones were identified in the posterior heart domain. Fifteen of these clones consisted solely of two Tin-positive pericardial cells, eight clones consisted of two Tin-positive pericardial cells and two nearby mesodermal cells, and one clone consisted of a single Tin-positive pericardial cell. Thus, when one Tin-positive pericardial cell is observed within a clone of two or more cells, a second Tin-positive pericardial cell always exists within this clone. These data indicate that the four Tin-positive pericardial cells found in each hemisegment of the posterior domain arise from two Tin-positive pericardial cell precursors. The inability to identify any clones that contain four Tin-positive pericardial cells indicates that adjacent Tin-positive pericardial cell precursors are unlikely to share a common lineage (Alvarez, 2003).

Eighteen Tin-positive pericardial cell clones were identified in the anterior heart domain. All 18 clones consisted of one Tin-positive pericardial cell and one cardioblast. These data indicate that within this region Tin-positive pericardial cells arise from bi-potent heart precursors, each of which produces one Tin-positive pericardial cell and one cardioblast. These data also demonstrate that cardioblasts and Tin-positive pericardial cells in the anterior heart domain develop via a different cell lineage than cardioblasts and Tin-positive pericardial cells that develop in the posterior domain (Alvarez, 2003).

The analysis of ten additional cardioblast clones in the anterior heart domain support a distinct cell lineage for anterior versus posterior cardioblasts. Nine clones consisted of one cardioblast and one non-Tin-expressing pericardial cell, whereas a single clone consisted of two cardioblasts. Thus, most, if not all, anterior domain cardioblasts share a sibling relationship with a pericardial cell. In addition, all anterior domain cardioblasts exhibit cell lineages distinct from posterior domain cardioblasts. Together with the lineage data on Tin-positive pericardial cells, these results support the idea that cardioblasts and Tin-positive pericardial cells in the anterior heart domain carry out distinct functions from those found in the posterior heart domain (Alvarez, 2003).

Interestingly, the lineage of the twelve cardioblasts in the anterior heart domain appears fixed with respect to whether they share a sibling relationship with a Tin-positive or Tin-negative pericardial cell. These cardioblasts were numbered 1-12 from anterior to posterior with cardioblast 12 being immediately anterior to the first Svp-positive cardioblast. Four clones were identified that contained cardioblast 12 and in each clone this cardioblast shared a sibling relationship with a Tin-negative pericardial cell. By contrast, cardioblasts 10 and 11 each share a sibling relationship with a Tin-positive pericardial cell. No multiple clones were obtained for all twelve cardioblasts; nonetheless, these data suggest a fixed relationship between the position of a cardioblast and whether its sibling pericardial cell expresses Tin. It is speculated that the differences in gene expression between different pairs of sibling cardioblasts and pericardial cells in the anterior domain may reflect functional differences between such pairs of heart cells (Alvarez, 2003).

pnt was identified as an inhibitor of cardioblast development in a screen for mutations that affect cardioblast and/or pericardial cell development. To identify genes that regulate heart development ~2000 third chromosomal lethal P element lines were screened for defects in the expression of Mef2 a protein expressed in all cardioblasts and Eve. Two P element mutations were uncovered that cause an approximate twofold increase in cardioblasts. One of these P elements [l(3)S012309] maps to cytological position 94F1-3 and was known to be an allele of pnt. To verify that lesions in pnt result in the formation of ectopic cardioblasts, the phenotype of five additional pnt alleles was assayed. Although the severity of the phenotype varies for each pnt allele, all alleles display a significant increase in cardioblast number relative to wild-type embryos. With respect to the excess cardioblast phenotype, these alleles can be grouped into an allelic series. The presence of excess cardioblasts in embryos homozygous mutant for each pnt allele indicates that pnt normally functions in heart development to repress cardioblast development (Alvarez, 2003).

The published heart phenotype of the GATA transcription factor pnr is opposite that of pnt. In pnr mutant embryos, too many pericardial cells and too few cardioblasts are thought to develop. As a first step towards examining the potential regulatory interactions between pnr and pnt, a detailed analysis was carried out of heart development in pnr mutant embryos. pnrVX6, a null allele that contains a small deletion that removes all but the N-terminal nine amino acids of pnr, was used as well as pnr1, a molecularly uncharacterized allele. In contrast to a prior study, a loss of both cardioblasts and pericardial cells was found in pnr embryos. The dorsal mesodermal phenotypes were quantified for Eve-positive pericardial cells as well as for all pericardial cells using the pan-pericardial marker Zfh1. In wild-type embryos an average of 22.7 Eve-positive pericardial cells and 61.1 Zfh1-positive pericardial cells were observed per embryo side. pnrVX6 embryos exhibit the most severe effect with an average of 9.4 and 16.9 Eve- and Zfh1-positive pericardial cells, respectively. pnrVX6/pnr1 embryos exhibit an intermediate phenotype with an average of 16.4 Eve-positive pericardial cells and 27.4 Zfh1-positive pericardial cells, while pnr1 embryos exhibit the mildest phenotype with an average of 21.2 and 37.4 Eve- and Zfh1- positive pericardial cells, respectively. A severe loss of cardioblasts and Odd-positive pericardial cells were observed in these backgrounds although these phenotypes were not quantified. The loss of cardioblasts and Odd-positive pericardial cells is most severe in pnrVX6 embryos and least severe in pnr1 embryos where short stretches of cardioblasts are still visible. These results indicate that pnr normally functions to promote the development of all heart cells. The results demonstrate that the earliest manifestation of cardiac mesoderm development is defective in pnr embryos, suggesting that the general lack of heart cells in pnr embryos arises indirectly via a defect in the specification of the cardiac mesoderm. Double mutant studies support the model that pnr acts upstream of pnt in a developmental pathway (Alvarez, 2003).

This paper indicates that pnr and pnt act sequentially to regulate heart development. pnr acts early in mesoderm development to enable the cardiac mesoderm to form. Subsequent to this event, pnt acts within the cardiac mesoderm to regulate the ability of cells to choose between the pericardial or cardioblast fate. In this context, pnt inhibits the development of the Svp-class of cardioblasts and appears to function independently of Ras/MAP kinase pathway activity (Alvarez, 2003).

The effect of pnt on heart development is restricted to the posterior seven heart segments where Svp cardioblasts normally develop. Interestingly, the lineage studies identify a clear difference in the cell lineage of cardioblasts that develop in the posterior seven heart segments versus those that develop more anteriorly. These results identify a genetic and developmental distinction between these two regions of the heart. In addition, they suggest that cells in different regions of the heart carry out different functions and that these functions are probably under homeotic gene control. Future work that addresses the physiological role of these cells in heart function and the control of their development by homeotic genes should provide a more comprehensive understanding of heart development (Alvarez, 2003).

The data suggest that PntP2 may regulate cardioblast and pericardial cell development independently of Ras/MAP kinase activity. Given that every other developmental function of pnt has been traced back to receptor tyrosine kinase/Ras signaling activity, the apparent Ras independent activity of PntP2 is puzzling. Since PntP2 is expressed broadly throughout the mesoderm, a number of models can explain the apparent Ras-independent activity of PntP2 in the heart. For example, PntP2 may not require MAP-kinase-mediated phosphorylation to carry out a subset of its function. Consistent with this, phosphorylation of PntP2 does not appear to affect its DNA-binding ability. Thus, in the absence of MAP-kinase stimulation, PntP2 is still probably able to bind target promoters alone or in complexes with other proteins. Such an activity of PntP2 could on its own regulate target gene expression by blocking the ability of other transcriptional effectors to bind to and activate target gene transcription, or through an obligate association with other proteins required to activate (or to repress) target genes (Alvarez, 2003).

A second model is that PntP2 requires MAP kinase activation but that this activity is carried out by one of the other MAP kinase pathways in Drosophila: the JNK pathway or the p38 pathway. Preliminary phenotypic analyses indicate that heart development is normal in embryos mutant for basket, the Drosophila JNK-kinase. Analysis of p38 kinase activity is presently limited because of the absence of suitable genetic backgrounds. A third possibility is that a novel Ras-dependent pathway does in fact activate PntP2 during heart development. This model is consistent with the recent identification of a novel receptor tyrosine kinase expressed in the developing visceral mesoderm. Experiments that failed to identify a pnt-like excess cardioblast phenotype upon mesodermal overexpression of a dominant-negative form of Ras argue against this model. However, Ras is maternally loaded and it is extremely difficult to eliminate all Ras activity in this manner. Thus, even though Ras-like mesodermal phenotypes were observed in these experiments, a role for Ras in regulating cardioblast number may have been missed because of differential sensitivity of different developmental pathways to partial Ras inactivation. Future work that (1) addresses the ability of MAP-kinase insensitive forms of PntP2 to regulate heart development, and (2) identifies PntP2 target genes in the heart and elucidates how PntP2 regulates such genes should help clarify the molecular basis through which PntP2 governs heart development (Alvarez, 2003).

Phenotypic analysis of pnr conflicts with a prior study that showed an increase in pericardial cells in pnr mutants. This study used Eve to identify a subset of pericardial cells in pnr1 embryos. The difference in results are attributed to use of the pnrVX6 null allele, the ability to distinguish unambiguously Eve-positive pericardial cells from Eve-positive somatic muscle progenitors, and to specific defects in dorsal closure exhibited by pnr embryos that result in the local aggregation of cells in the dorsal region of the embryo. Genetic results identify pnr1 as a hypomorphic allele and Eve-positive pericardial cell formation is found to be almost wild type in this background. In these experiments, Eve-positive pericardial cells were unambiguously identified via their co-expression of Zfh1 and it was possible to quantify precisely Eve-positive pericardial cell number in pnr1 embryos. This is important as one can observe local increases in Eve-positive mesodermal cells in pnr embryos. However, such apparent increases arise from the local aggregation of dorsal mesodermal cells in pnr1 embryos caused by defects in dorsal closure and not by an overall increase in Eve-positive mesodermal cells (Alvarez, 2003).

The genetic identification of pnr1 as a hypomorphic allele is intriguing given that molecular and expression analyses indicate the pnr1 lesion results from a premature stop codon in the middle of the first zinc finger and that the Pnr1 protein localizes predominantly to the cytoplasm. This lesion is expected to abrogate the DNA-binding ability of the Pnr protein. However, genetic experiments indicate that the Pnr1 protein retains residual activity at least with respect to heart development. These results raise the possibility that Pnr may be able to carry out some of its functions independently of DNA binding. Future work that focuses on a detailed structure function analysis of the Pnr protein should clarify whether Pnr can act independently of its DNA-binding ability in some developmental contexts (Alvarez, 2003).

The pnt allelic series indicates that pntDelta88 exhibits a milder excess cardioblast phenotype than pntS012309, pnt2, and pntRR112. This result is surprising as pntDelta88 deletes the exons pntP2 shares with pntP1 and as a result pntDelta88 is assumed to be an amorphic allele of the pnt locus. Using antisense RNA probes specific for the unique exons of pntP2, an essentially wild-type pattern of pntP2 transcription is observed in pntDelta88 mutant embryos. These data raise the possibility that the N-terminal regions of pntP2 may also retain partial activity. Studies along the lines of those suggested for Pnr should also help elucidate whether truncated forms of PntP2 retain residual activity (Alvarez, 2003).

Significant similarity exists between the embryology and molecular regulation of early heart development in Drosophila and vertebrates. In this context, the identification of a role for pnt (a member of the evolutionarily conserved ETS transcription factor family) in Drosophila heart development raises the possibility that ETS family proteins regulate vertebrate heart development. Consistent with this, ETS1 and ETS2, the two most closely related vertebrate ETS proteins to pnt, are expressed in the developing vertebrate heart -- functional studies indicate these genes regulate the expression of specific genes in the heart. However, knockout studies have not yet revealed a clear role for ETS1 or ETS2 in the morphological development or differentiation of the vertebrate heart. The existence of multiple vertebrate ETS-family members highly homologous to pnt, as well as a total of 25 ETS family members in humans suggests the possibility of functional redundancy in ETS protein function during vertebrate and mammalian heart development. Thus, a full understanding of ETS protein function during heart development awaits construction and analysis of animals multiply mutant for different ETS family members (Alvarez, 2003).

Genetic interaction of Lobe with its modifiers in dorsoventral patterning and growth of the Drosophila eye

Dorsoventral (DV) patterning is essential for growth of the Drosophila eye. Recent studies suggest that ventral is the default state of the early eye, which depends on Lobe (L) function, and that the dorsal fate is established later by the expression of the dorsal selector gene pannier (pnr). However, the mechanisms of regulatory interactions between L and dorsal genes are not well understood. For studying the mechanisms of DV patterning in the early eye disc, a dominant modifier screen was performed to identify additional genes that interact with L. The criterion of the dominant interaction was either enhancement or suppression of the L ventral eye loss phenotype. Forty-eight modifiers were identified that corresponded to 16 genes, which included fringe (fng), a gene involved in ventral eye patterning, and members of both Hedgehog (Hh) and Decapentaplegic (Dpp) signaling pathways, which promote L function in the ventral eye. Interestingly, 29% of the modifiers (6 enhancers and 9 suppressors) identified either were known to interact genetically with pnr or were members of the Wingless (Wg) pathway, which acts downstream from pnr. The detailed analysis of genetic interactions revealed that pnr and L mutually antagonize each other during second instar of larval development to restrict their functional domains in the eye. This time window coincides with the emergence of pnr expression in the eye. These results suggest that L function is regulated by multiple signaling pathways and that the mutual antagonism between L and dorsal genes is crucial for balanced eye growth (Singh, 2005).

Axial patterning plays a crucial role in organizing growth and in differentiating developing fields. To understand how the DV pattern is established in the Drosophila eye, the genetic relationships between dorsal and ventral eye genes were analyzed. A group of new genes was identified that modifies the L mutant eye phenotype not only by misexpression but also by reduced gene function (Singh, 2005).

In the early eye disc, fng is preferentially expressed in the ventral eye. The DV domain specification by Fng is also important for growth of the eye disc as its ubiquitous overexpression in the eye disc blocks eye development. Even though L and fng play important roles during ventral eye growth and patterning, the developmental interaction between the two has been unknown. This study showed that overexpression of fng can partially compensate for the loss of L gene function in the eye. This suggests that fng works either downstream or parallel to L in the growth of the ventral eye. It is possible that L and fng interact through the induction of a common target, Ser, in the eye (Singh, 2005).

Like several other pathways, Wg signaling has multiple functions during eye development. This study identified Sgg, a serine/threonine kinase, as a modifier that suppresses the L mutant phenotype upon overexpression. Sgg is known to inhibit the Wg signaling pathway by downregulating Armadillo (Arm) via ubiquitin-mediated proteosomal degradation. Other components of the Wg signaling pathway such as pygo and dally were also identified as modifiers, which, upon overexpression in the eye, enhanced the L mutant phenotype. These results suggest that Wg signaling acts antagonistically to L function in the ventral eye. The genetic interaction of these EP lines with the L mutations represents specific enhancement rather than additive effects, since antagonists of Wg signaling were identified as suppressors, whereas members required for Wg signaling were identified as enhancers of the L mutant phenotype in the EP screen (Singh, 2005).

In this screen, it was found that the overexpression of Daughters against Dpp (Dad), an antagonist of Dpp signaling, enhances the L mutant phenotype, whereas EP insertions at hh and its receptor gene smo were identified as suppressors of the L mutant phenotype. The members of these two signaling pathways are known to be involved in eye growth and differentiation. These results raise another interesting possibility of the possible role of Hh and Dpp signaling pathways in early eye growth and patterning (Singh, 2005).

During Drosophila eye development, Hh controls progression of the furrow by inducing the expression of dpp and atonal (ato), a proneural gene responsible for R8 photoreceptor formation. In the eye, hh and dpp are involved in a positive feedback loop for the initiation and movement of the MF whereas Wg signaling acts antagonistically to Dpp signaling to block MF movement and progression. This antagonistic relation may be present even during early eye development since wg and dpp are localized to opposing regions of the undifferentiated younger eye primordia: dpp along the posterior margin and wg across the dorsal anterior region. These results suggest that the early function of Hh and Dpp signaling is to promote L-mediated ventral eye growth whereas Wg signaling acts as an antagonist (Singh, 2005).

BarH1 and BarH2 were identified as enhancers of the L mutant phenotype. BarH1 and BarH2 are a pair of homeobox proteins that express in a subset of photoreceptors and in the basal undifferentiated cells of the eye disc. B is required for the negative regulation of eye development by repressing the expression of the proneural gene ato. However, it is not known whether B plays a role in early eye growth, prior to retinal differentiation. Clonal analysis has not yet revealed evidence for B function in DV asymmetric eye patterning. However, the current data showed genetic interactions of L mutants with GOF and LOF mutants of B. The suppression of the L2/+ eye phenotype by a LOF mutation of B, which by itself has no defects in the eye, raises the possibility that B itself may not have DV asymmetric function but needs to be downregulated by L for normal growth of the early eye disc. This is also consistent with the dramatic eye reduction observed when ey-GAL4 drive B is overexpressed during early eye development. It has been shown that Wg and B have both positive and negative regulatory relationships in prepatterning of the notum. B expression is activated by Wg in the scutum whereas B represses Wg expression in the most anterior part of the notum. L mutants respond to GOF and LOF of both Wg signaling and B in a similar fashion, suggesting that Wg and B may be regulating each other positively during early eye development (Singh, 2005).

L is known to act downstream of N. In the eye, emc and h, the repressors of ato, are downregulated by N. During eye development, emc acts in collaboration with hairy (h) as the negative regulator of the morphogenetic furrow by repressing ato. Therefore, identification of emc as an antagonist of L-mediated early ventral eye growth seems possible. Interestingly, both emc and B have also been identified as modifiers of pnr (Singh, 2005).

Some of the genes that were identified as L modifiers, such as B, emc, and smo, have been well characterized, but their roles in early eye disc growth and/or DV asymmetric function have not been studied. Genes were identified involved in cell survival and growth such as disc over grown (dco), a member of the serine/threonine protein kinases family, and genes involved in vesicular trafficking, including RhoGAP68F, an ion transport such as nrv 1, and the acetyl transferase nej. It is possible that potential DV asymmetric function of these genes might have been missed by LOF analysis because of functional redundancy or these genes may be modifying the early growth function of L in the eye. More in-depth studies will be necessary to explore these possibilities. However, it is important to note that both GOF and LOF of these genes exhibit specific genetic interactions with L mutant backgrounds. In addition to the well-characterized genes, a few novel genes like EP1229 and EP1595 were identified whose functions are not known. These genes were not listed in this study as the specificity of their genetic interaction with L was not tested by using LOF mutations (Singh, 2005).

The results demonstrate that the level of pnr gene function is a crucial factor for DV patterning of the eye as increased levels of pnr gene function enhance the L mutant phenotype of ventral eye loss to no eye, whereas reduction of pnr gene function rescued the loss of the ventral eye phenotype of the L mutant. Further, the phenotypes of LOF clones of L where only the ventral cells are lost can be rescued by reducing the levels of pnr gene function. These results suggest that pnr acts antagonistically to the ventral eye growth function of L. However, it was also found that the antagonism of pnr and L is mutual. This conclusion is based on the fact that the gain-of-function phenotype of pnr in the eye is significantly enhanced when L function is reduced. These conclusions were also validated by showing that the dorsal eye enlargements associated with LOF clones of pnr can be prevented by reducing the levels of L gene function. These results suggest that optimal levels of pnr and L are necessary for DV patterning and growth of the eye. It was also found that the downstream dorsal eye selectors, Iro-C members (ara, caup, mirr) are involved in a mutually antagonistic relationship with L. These studies demonstrate that the antagonism of L holds true for key components involved in dorsal fate selection during early eye development (Singh, 2005).

The time window of the second instar of larval development was identified as the periord during which mutual antagonistic interaction of L and pnr is required for DV patterning and growth in the eye. Previously, it was shown that the pnr function in eye development is critically required during the second instar larval stage. This time window is coincident with the one that is required for the antagonism of pnr and L as shown in this study, suggesting that a major function of L in early eye development is to establish the DV domains by negatively regulating the dorsal selectors. These studies also support the physiological relevance of this mutually antagonistic interaction in DV patterning (Singh, 2005).

It is not known how L antagonizes Pnr function. One possibility is that L may be required for restricting the pnr expression domain to the dorsal margin of the eye disc. It was difficult to check whether L is cell-autonomously required for pnr repression because LOF clones of L result in the elimination of the entire or ventral eye, depending on the time when the clones are generated. Alternatively, the effect of a L mutation on pnr expression was studied. Interestingly, pnr expression, which is restricted to the dorsal eye margin in wild type eye discs, shows a nearly twofold expansion in L2/+ mutant discs. It remains to be studied whether L is required for the repression of pnr expression or for the inhibition of growth of pnr-expressing cells. On the basis of these data it is suggested that during early DV patterning, the onset of pnr expression might restrict the functional domain of L and Ser to the ventral eye. It is possible that pnr may also suppress L gene function via the Wg signaling pathway (Singh, 2005).

The results support the view that various developmental pathways cross-talk with each other to define the final form of a developing eye field. Such genes are likely to interact with both pnr and L. It is interesting to note that several pnr-interacting genes were identied as L modifiers in the screen. This illustrates the importance of the interaction of L and pnr pathways and also the efficacy of the screen. Further study of new modifiers of L may provide important clues to the mechanism of pnr-L interactions in the control of growth and/or DV patterning of the eye. Since the compound eye of Drosophila shares some similarities with the vertebrate eye and genetic machinery is highly conserved, it would be interesting to see if these antagonistic interactions between the dorsal eye selectors and the ventral eye genes play roles in the DV patterning and growth of vertebrate eyes (Singh, 2005).

The Drosophila homolog of vertebrate Islet1 is a key component in early cardiogenesis

In mouse, the LIM-homeodomain transcription factor Islet1 (Isl1) has been shown to demarcate a separate cardiac cell population that is essential for the formation of the right ventricle and the outflow tract of the heart. Whether Isl1 plays a crucial role in the early regulatory network of transcription factors that establishes a cardiac fate in mesodermal cells has not been fully resolved. This study analyzed the role of the Drosophila homolog of Isl1, tailup (tup), in cardiac specification and formation of the dorsal vessel. The early expression of Tup in the cardiac mesoderm suggests that Tup functions in cardiac specification. Indeed, tup mutants are characterized by a reduction of the essential early cardiac transcription factors Tin, Pnr and Dorsocross1-3 (Doc). Conversely, Tup expression depends on each of these cardiac factors, as well as on the early inductive signals Dpp and Wg. Genetic interactions show that tup cooperates with tin, pnr and Doc in heart cell specification. Germ layer-specific loss-of-function and rescue experiments reveal that Tup also functions in the ectoderm to regulate cardiogenesis and implicate the involvement of different LIM-domain-interacting proteins in the mesoderm and ectoderm. Gain-of-function analyses for tup and pnr suggest that a proper balance of these factors is also required for the specification of Eve-expressing pericardial cells. Since tup is required for proper cardiogenesis in an invertebrate organism, it is appropriate to include tup/Isl1 in the core set of ancestral cardiac transcription factors that govern a cardiac fate (Mann, 2009).

The specification of a subset of mesodermal cells towards a cardiac fate requires well-orchestrated interactions of a plethora of factors. Drosophila is the model system of choice to decipher the complex transcriptional network that initiates and sustains a cardiac lineage. The data place tup as an essential component in the early transcriptional network that specifies cardiac mesoderm (Mann, 2009).

After the initially broad expression domain of Tin has become restricted to the dorsal mesodermal margin, Tup expression is first seen in the cardiac mesoderm in ~10 small clusters, which co-express Eve. Slightly later, Tup is present throughout the Tin-positive cardiac mesoderm and gene expression analyses in tupisl-1, tin346, pnrVX6 and Df(3L)DocA embryos demonstrate that all four factors are required to maintain each other's expression. Additionally, analyses of cardiac gene expression in embryos that are transheterozygotic for tup and tin, pnr or Doc, showed that these factors interact genetically to specify heart cells (Mann, 2009).

Although it might be expected that Tup expression is lost in tin mutants since these embryos are devoid of heart cells, it is interesting that Tup expression in the early cell clusters is still initiated. This finding is somewhat reminiscent of the observation that the initiation of Doc expression is also independent of tin. According to the temporal appearance of Tup in the cardiac mesoderm with respect to Tin and Doc, tup is required for their maintenance rather than their initiation. By contrast, the onset of mesodermal Pnr and Tup expression appears to coincide. It was not resolved whether Tup is induced by Pnr or directly by Dpp. A direct regulation by Dpp was implicated by the reduced expression of Tup after mesodermal overexpression of UAS-brinker, which is known to bind to dpp-response elements of dpp target genes. Conversely, it was shown that dpp expression depends on tup and the present data suggest that this regulation requires pnr (Mann, 2009).

Germ layer-specific inhibition of Tup using a construct that lacks the homeodomain, but contains the two LIM domains, revealed that Tup can regulate cardiogenesis in the mesoderm as well as from the ectoderm. Since the 69B-Gal4 driver has been reported not to be strictly ectodermal, it is possible that mesodermal Tup function was also interfered with. However, the mesodermal expression of 69B-Gal4 seems to be negligible. The effect of ectodermal Tup inhibition on cardiogenesis in the mesoderm can only be explained if the function of a secreted growth factor is impaired. dpp expression was analyzed, and a slight downregulation of its transcripts was observed in embryos expressing UAS-tupδHD in the ectoderm. Since this effect might not be sufficient to account for the strong Tin phenotype, further experiments will be required to determine whether additional growth factors are affected (Mann, 2009).

To better determine the germ layer-specific contribution of Tup in cardiogenesis, attempts were made to rescue the Tin phenotype by co-expressing the full-length tup cDNA. Somewhat unexpectedly, a better rescue was obtained when both constructs were expressed in the ectoderm rather than in the mesoderm. Since the LIM domains present in tupδHD can sequester LIM-domain-binding proteins, a simple explanation for this finding is that Tup interacts with proteins that are present in the mesoderm but not in the ectoderm. It is reasonable to hypothesize that in the mesoderm the LIM domains of tupδHD not only act as a dominant-negative for Tup, but additionally for another, perhaps as yet unidentified, LIM-domain containing protein. Since it has been shown that Pnr can bind Tup through the LIM domains, it is likely that Pnr function was interfered with by overexpressing UAS-tupδHD. The requirement of the LIM domains for proper cardiac specification is shown by the reduction of Tin-expressing cells after mesodermal expression of the UAS-tupδLIM construct. Further experiments are under way to better resolve the molecular function of Tup in the different tissues (Mann, 2009).

Since the mesodermal expression of UAS-tupδHD resulted in a strong reduction of Tin-expressing cells at early stages of cardiac mesoderm formation, it was surprising to observe a rather low reduction of Dmef2-positive myocardial cells at later stages (15/16). To exclude the possibility that the twi-Gal4 driver does not sufficiently express UAS-tupδHD throughout embryogenesis, this experiment was repeated using the combined mesodermal driver twi-Gal4; 24B-Gal4. However, the phenotypes were not enhanced. A time course for Tin expression in these crosses revealed that Tin appears to recover over time. A similar phenomenon can be seen in tupisl-1 mutants, although it might not be as obvious because the mutants also lack ectodermal tup expression. In any case, the data is suggestive of a different temporal requirement for tup with respect to tin expression. It is known that tin expression depends on different transcriptional activation events. Consistent with the onset of Tup expression in the cardiac mesoderm at mid-stage 11, the earlier phases of Tin expression are unlikely to depend on Tup. Hence, the initial Tin expression at stages 8-10 is sufficient to generate a considerable number of Dmef2-positive myocardial cells at later stages (Mann, 2009).

These analyses further implicate that Tup might act as a transcriptional activator or repressor depending on the cellular context and on the factors with which it is co-expressed. This is most strikingly observed with respect to the Odd-expressing pericardial and lymph gland cells. In tup mutants, Odd-positive cells are missing in both organs. A similar phenotype is seen when Tup is overexpressed in the mesoderm using the twi-Gal4 driver. The loss of Odd-expressing cells in lymph glands is reminiscent of the phenotype observed in tup mutants, although it is less severe. This differential occurrence of the phenotype indicates that tup can differentially regulate factors involved in cardiogenesis versus lymph gland development. This is substantiated by the finding that mesodermal overexpression of tup results in an increase in Hand expression in the lymph glands, while Hand expression throughout the dorsal vessel is only mildly affected. Despite the loss of Odd-positive cells after early mesodermal tup overexpression, Tup is required in the pericardial and lymph gland cells at later stages to maintain Odd expression. Moreover, overexpressing tup in the pericardial cell lineage yields additional Odd-expressing pericardial cells and rescues Odd expression in the lymph glands (Mann, 2009).

To obtain more insight into possible functional interactions with other cardiac transcription factors, tup was overexpressed in combination with pnrD4. The latter is a highly active variant of wild-type pnr that contains an amino acid substitution in the N-terminal zinc finger, which abolishes binding of Ush to Pnr. Mesodermal overexpression of pnrD4 results in robust ectopic activation of Tin and embryos co-overexpressing tup and pnrD4 exhibit the same phenotype. Most likely, a possible influence of Tup on Pnr activity, regardless of whether it is positive or negative, is concealed by the strong gain-of-function pnr allele. However, analysis of Eve expression does provide insight into possible regulatory interactions between Tup and Pnr. Mesodermal overexpression of each factor alone yields opposing phenotypes, and when both factors are co-overexpressed PnrD4 can efficiently counteract Tup activity and prevent the overspecification of Eve cells. Vice versa, Tup can, although only moderately, counteract the effect of PnrD4. It has been shown that during patterning of the thorax, Tup can antagonize the proneural activity of Pnr by forming a heterodimer, and that the physical interaction between Pnr and Tup is mediated by the two zinc fingers of Pnr. Hence, the somewhat weak, but possibly antagonistic, function of Tup towards PnrD4 in Eve-positive cell specification could be due to the amino acid substitution encoded in the pnrD4 allele, which might weaken the interaction between the two factors, as compared with wild-type Pnr. Overexpression of a Tup construct that lacks both LIM domains did not result in expanded Eve-positive clusters, which strongly suggests that the effect of Pnr on Tup activity, as seen when both factors are co-expressed, requires the presence of the LIM domains (Mann, 2009).

In summary, these data demonstrate the crucial role of tup in the proper specification of cardiac mesoderm in an invertebrate organism. Therefore, tup/Isl1 should be added to the core set of ancestral cardiac transcription factors. Consequently, this implicates that the evolution of the vertebrate four-chambered heart does not necessarily require the acquisition of a novel network of cardiac transcription factors. At least, it is unlikely that tup/Isl1 is part of a regulatory network separate from that of tin/Nkx2.5, pnr/Gata4 and Doc/Tbx5/6 because it is an essential factor for the formation of the simple linear heart tube in the fly (Mann, 2009).

Spire, an actin nucleation factor, regulates cell division during Drosophila heart development

The Drosophila dorsal vessel is a beneficial model system for studying the regulation of early heart development. Spire (Spir), an actin-nucleation factor, regulates actin dynamics in many developmental processes, such as cell shape determination, intracellular transport, and locomotion. Through protein expression pattern analysis, this study demonstrates that the absence of spir function affects cell division in Myocyte enhancer factor 2-, Tinman (Tin)-, Even-skipped- and Seven up (Svp)-positive heart cells. In addition, genetic interaction analysis shows that spir functionally interacts with Dorsocross, tin, and pannier to properly specify the cardiac fate. Furthermore, through visualization of double heterozygous embryos, it was determined that spir cooperates with CycA for heart cell specification and division. Finally, when comparing the spir mutant phenotype with that of a CycA mutant, the results suggest that most Svp-positive progenitors in spir mutant embryos cannot undergo full cell division at cell cycle 15, and that Tin-positive progenitors are arrested at cell cycle 16 as double-nucleated cells. It is concluded that Spir plays a crucial role in controlling dorsal vessel formation and has a function in cell division during heart tube morphogenesis (Xu, 2012).

Proper dorsal vessel morphogenesis is critically dependent upon intercellular signaling and the regulation of gene expression. Although great progress has been made in the study of heart development, it is not known exactly how many genes and pathways are involved in this cardiogenic process or how many of these factors cooperate together. Previous genetic screens have identified genes that play roles in the specification, morphogenesis, and differentiation of the heart, including mastermind and tup. The current sensitized screen has also proved to be an efficient method to find additional factors in this process, suggesting that much remains to be learned about the molecular components involved in correct dorsal vessel formation (Xu, 2012).

Spir is required for the proper timing of cytoplasmic streaming in Drosophila, and loss of spir leads to premature microtubule-dependent fast cytoplasmic streaming during oogenesis, the loss of oocyte polarity, and female sterility. Even though it is known that spir is an important actin filament nucleation factor, the findings are the first report to describe a function of spir for cell division. Through phenotypic analysis of the spir mutant phenotype, it was found that many cardioblast nuclei are partially or completely divided. However, the cytoplasm is not divided in the absence of spir, which is consistent with the function of spir in cytoplasmic movement. Thirteen rapid nuclear division cycles without cell division initiate Drosophila embryo development, followed by three waves of cell division. The first wave of cell division occurs in the mesoderm at cell cycle 14. After this initial division, cells migrate, spread dorsally and undergo a second round of cell division at cell cycle 15. The third wave of cell division in the mesoderm occurs at the end of germband extension during cell cycle 16. There are two different types of cardioblast precursor cells: one type divides into two identical Tin-positive cardioblasts (TC), and the other type divides into one Svp-positive cardioblast (SC) and one Svp-positive pericardial cell (SPC). Based on the comparison of CycA and spir mutant phenotypes, a tentative cell division model is proposed to demonstrate spir function in determining cardiac cell fate (see A cell division model of spir function during heart development). In a wild-type background, one Svp-positive super progenitor (SSP) divides into two Svp-positive progenitors (SP), then each of these cells divides into one SPC and one SC. For Tin-positive super progenitors (TSP), after each divides into two Tin-positive progenitors (TP), each TP further divides into two identical TCs. In the current model, division from the super progenitor to progenitors takes place at cell cycle 15, and division from progenitors to full differentiated heart cells occurs at cell cycle 16. In CycA mutants, mitosis 16 is blocked such that both SPs and TPs stop cell division. This results in the two SPs assuming a myocardial fate. Thus the number of SCs remains normal, but half of the TCs are missing in the CycA mutants. The data suggest that in spir mutant embryos, most of the SPs fail to undergo full cell division at cycle 15 resulting in a SPC fate with paired nuclei. A subset of these cells are able to undergo the 15th cell division but are arrested at cycle 16 as double-nucleated cells which exhibit both Svp and Mef2 staining, characteristic of the SCs seen in the CycA mutants. Similarly, for TPs, cycle 16 was also blocked such that it resulted in two double-nucleated cells. In summary, Spir affects mitosis 16 for Tin-positive cell division and both mitosis 15 and 16 for Svp-positive cell division (Xu, 2012).

Antibody staining suggests that Spir is expressed ubiquitously before stages 12-13 and is located in both nuclei and cytoplasm. After cell cycle 16 cell division stops, occurring during stage 10-11. The expression of Spir in the cytoplasm then decreases gradually. At stage 15, the staining pattern shows mostly nucleus expression with some cytoplasmic expression and by stage 16 the nuclei become distinct indicating nucleus staining only. It is hypothesized that expression of Spir decreases in the cytoplasm but remains constant in the nuclei when cell division halts (Xu, 2012).

The genetic analysis of spir, Doc, pnr and tin suggests that these factors may regulate each other during dorsal vessel formation, and especially significant is the interaction between spir and pnr. Pnr is a GATA class transcription factor expressed in both the dorsal ectoderm and dorsal mesoderm, where it is required for cardiac cell specification. Proper dorsal vessel formation is inhibited in pnr loss-of-function embryos due to failure in the specification of the cardiac progenitors. In spir mutants, the expression pattern of Pnr remains normal. However, Spir is over-expressed in the cardiac mesoderm in pnr mutants, suggesting that Pnr may repress the expression of the spir (Xu, 2012).

In conclusion, Spir is a newly-identified factor functioning in cell division during dorsal vessel formation. Tin-, Eve- and Svp-positive heart cells are all affected in the absence of spir. Also, spir expression depends on the transcription factors Doc, tin and pnr. Genetic interaction data also show that spir cooperates with CycA in heart cell division (Xu, 2012).


REFERENCES

Akasaka, T., et al. (2006). The ATP-sensitive potassium (KATP) channel-encoded dSUR gene is required for Drosophila heart function and is regulated by tinman. Proc. Natl. Acad. Sci. 103: 11999-12004. PubMed ID: 16882722

Aldaz, S., Morata, G. and Azpiazu, N. (2003). The Pax-homeobox gene eyegone is involved in the subdivision of the thorax of Drosophila. Development 130: 4473-4482. PubMed ID: 12900462

Alvarez, A. D., Shi, W., Wilson, B. A. and Skeath, J. B. (2003). pannier and pointedP2 act sequentially to regulate Drosophila heart development. Development 130: 3015-3026. PubMed ID: 12756183

Ashe, H. L., Mannervik, M. and Levine, M. (2000). Dpp signaling thresholds in the dorsal ectoderm of the Drosophila embryo. Development 127: 3305-3312. PubMed ID: 10887086

Asmar, J., Biryukova, I. and Heitzler, P. (2008). Drosophila dLMO-PA isoform acts as an early activator of achaete/scute proneural expression. Dev. Biol. 316(2): 487-97. PubMed ID: 18329012

Ayyar S., et al. (2007). NF-kappaB/Rel-mediated regulation of the neural fate in Drosophila. PLoS ONE 2: e1178. PubMed ID: 18000549

Ayyar, S., et al. (2010). An arthropod cis-regulatory element functioning in sensory organ precursor development dates back to the Cambrian. BMC Biol. 8: 127. PubMed ID: 20868489

Bejarano, F., Luque, C. M., Herranz, H., Sorrosal, G., Rafel, N., et al., (2008). A gain-of-function suppressor screen for genes involved in dorsal-ventral boundary formation in the Drosophila wing. Genetics 178: 307-323. PubMed ID: 18202376

Biryukova, I. and Heitzler, P. (2005). The Drosophila LIM-homeodomain protein Islet antagonizes proneural cell specification in the peripheral nervous system. Dev. Biol. 288: 559-570. PubMed ID: 16259974

Briegel, K., et al. (1996). Regulation and function of transcription factor GATA-1 during red blood cell differentiation. Development 122: 3839-3850. PubMed ID: 9012505

Bronstein, R., et al. (2011a). Transcriptional regulation by CHIP/LDB complexes. PLoS Genet. 6(8): e1001063. PubMed ID: 20730086

Calleja, M., et al. (1996). Visualization of gene expression in living adult Drosophila. Science 274: 252-255. PubMed ID: 8824191

Calleja, M., et al. (2000). Generation of medial and lateral dorsal body domains by the pannier gene of Drosophila. Development 127: 3971-3980. PubMed ID: 10952895

Chen, L., et al. (2002). Ssdp proteins interact with the LIM-domain-binding protein Ldb1 to regulate development. Proc. Natl. Acad. Sci. 99: 14320-14325. PubMed ID: 12381786

Culi, J. and Modolell, J. (1998). Proneural gene self-stimulation in neural precursors: an essential mechanism for sense organ development that is regulated by Notch signaling. Genes Dev. 12: 2036-2047. PubMed ID: 9649507

Eivers E., et al. (2009). Mad is required for wingless signaling in wing development and segment patterning in Drosophila. PLoS ONE 4: e6543. PubMed ID: 19657393

Fossett, N., et al. (2000). The multitype zinc-finger protein U-shaped functions in heart cell specification in the Drosophila embryo. Proc. Natl. Acad. Sci. 97: 7348-7353. PubMed ID: 10861002

Fossett, N., et al. (2001). The Friend of GATA proteins U-shaped, FOG-1, and FOG-2 function as negative regulators of blood, heart, and eye development in Drosophila. Proc. Natl. Acad. Sci. 98: 7342-7347. PubMed ID: 11404479

Fromental-Ramain, C., Vanolst, L., Delaporte, C. and Ramain, P. (2008). pannier encodes two structurally related isoforms that are differentially expressed during Drosophila development and display distinct functions during thorax patterning. Mech. Dev. 125(1-2): 43-57. PubMed ID: 18042352

Gajewski, K., et al. (2001). Pannier is a transcriptional target and partner of Tinman during Drosophila cardiogenesis. Dev. Bio. 233: 425-436. PubMed ID: 11336505

Garcia-Garcia, M. J., et al. (1999). Different contributions of pannier and wingless to the patterning of the dorsal mesothorax of Drosophila. Development 126: 3523-3532. PubMed ID: 10409499

Gilleard, J. S., et al. (1999). ELT-3: A Caenorhabditis elegans GATA factor expressed in the embryonic epidermis during morphogenesis. Dev. Biol. 208(2): 265-280. PubMed ID: 10191044

Ghazi, A., Paul, L. and VijayRaghavan, K. (2003). Prepattern genes and signaling molecules regulate stripe expression to specify Drosophila flight muscle attachment sites. Mech. Dev. 120: 519-528. PubMed ID: 12782269

Haenlin, M., et al. (1997). Transcriptional activity of Pannier is regulated negatively by heterodimerization of the GATA DNA-binding domain with a cofactor encoded by the u-shaped gene of Drosophila. Genes Dev. 11(22): 3096-3108

Han, Z. and Olson, E. N. (2005). Hand is a direct target of Tinman and GATA factors during Drosophila cardiogenesis and hematopoiesis. Development 132: 3525-3536. PubMed ID: 15975941

He, A., Kong, S. W., Ma, Q., and Pu, W. T. (2011). Co-occupancy by multiple cardiac transcription factors identifies transcriptional enhancers active in heart. Proc. Natl. Acad. Sci. 108: 5632-5637. PubMed ID: 21415370

Heitzler, P., et al. (1996). A genetic analysis of pannier, a gene necessary for viability of dorsal tissues and bristle positioning in Drosophila. Genetics 143(3): 1271-1286. PubMed ID: 8807299

Heitzler, P., Vanolst, L., Biryukova, I. and Ramain, P. (2003). Enhancer-promoter communication mediated by Chip during Pannier-driven proneural patterning is regulated by Osa. Genes Dev. 17: 591-596. PubMed ID: 12629041

Herranz, H. and Morata, G. (2001). The functions of pannier during Drosophila embryogenesis. Development 128: 4837-4846. PubMed ID: 11731463

Immarigeon, C., Bernat-Fabre, S., Auge, B., Faucher, C., Gobert, V., Haenlin, M., Waltzer, L., Payet, A., Cribbs, D. L., Bourbon, H. G. and Boube, M. (2019). Drosophila Mediator subunit Med1 is required for GATA-dependent developmental processes: divergent binding interfaces for conserved coactivator functions. Mol Cell Biol. PubMed ID: 30670567

Immarigeon, C., Bernat-Fabre, S., Guillou, E., Verger, A., Prince, E., Benmedjahed, M. A., Payet, A., Couralet, M., Monte, D., Villeret, V., Bourbon, H. M. and Boube, M. (2020). Mediator complex subunit Med19 binds directly GATA transcription factors and is required with Med1 for GATA-driven gene regulation in vivo. J Biol Chem. PubMed ID: 32737196

Jazwinska, A., Rushlow, C. and Roth, S. (1999b). The role of brinker in mediating the graded response to Dpp in early Drosophila embryos. Development 126(15): 3323-3334. PubMed ID: 10393112

Johnson, A. N., Mokalled, M. H., Haden, T. N. and Olson, E. N. (2011). JAK/Stat signaling regulates heart precursor diversification in Drosophila. Development 138: 4627-4638. PubMed ID: 21965617

Junion, G., et al. (2012). A transcription factor collective defines cardiac cell fate and reflects lineage history. Cell 148(3): 473-86. PubMed ID: 22304916

Klinedinst, S. L. and Bodmer, R. (2003). Gata factor Pannier is required to establish competence for heart progenitor formation. Development 130: 3027-3038. PubMed ID: 12756184

Kornhauser, J. M., et al. (1994). Temporal and spatial changes in GATA transcription factor expression are coincident with development of the chicken optic tectum. Brain Res. Mol. Brain Res. 23(1-2): 100-110. PubMed ID: 8028475

Liang, H. L., Xu, M., Chuang, Y. C. and Rushlow, C. (2012). Response to the BMP gradient requires highly combinatorial inputs from multiple patterning systems in the Drosophila embryo. Development. 139(11): 1956-64. PubMed ID: 22513375

Letizia, A., Barrio, R. and Campuzano, S. (2007). Antagonistic and cooperative actions of the EGFR and Dpp pathways on the iroquois genes regulate Drosophila mesothorax specification and patterning. Development 134(7): 1337-46. PubMed ID: 17329358

Lo, P. C. and Frasch, M. (2001). A role for the COUP-TF-related gene seven-up in the diversification of cardioblast identities in the dorsal vessel of Drosophila. Mech. Dev. 104: 49-60. PubMed ID: 15922573

Lossky, M. and Wesick, P.C. (1995). Regulation of Drosophila yolk protein genes by an ovary-specific GATA factor. Mol. Cell Biol. 15: 6943-52. PubMed ID: 8524261

Ma, G. T., et al. (1997). GATA-2 and GATA-3 regulate trophoblast-specific gene expression in vivo. Development 124, 907-914 (1997). PubMed ID: 9043071

Mandal, L., Banerjee, U. and Hartenstein, V. (2004). Evidence for a fruit fly hemangioblast and similarities between lymph-gland hematopoiesis in fruit fly and mammal aorta-gonadal-mesonephros mesoderm. Nat. Genet. 36: 1019-1023. PubMed ID: 15286786

Mann, T., Bodmer, R. and Pandur, P. (2009). The Drosophila homolog of vertebrate Islet1 is a key component in early cardiogenesis. Development 136(2): 317-26. PubMed ID: 19088091

Marcellini, S. and Simpson, P. (2006). Two or four bristles: Functional evolution of an Enhancer of scute in Drosophilidae. PLoS Biol. 4(12): e386. PubMed ID: 17105353

Maurel-Zaffran, C. and Treisman, J. E. (2000). pannier acts upstream of wingless to direct dorsal eye disc development in Drosophila. Development 127: 1007-1016. PubMed ID: 10662640

Mirzoyan, Z. and Pandur, P. (2013). The Iroquois Complex Is Required in the Dorsal Mesoderm to Ensure Normal Heart Development in Drosophila. PLoS One 8: e76498. PubMed ID: 24086746

Miskolczi-McCallum, C. M., Scavetta, R. J., Svendsen, P. C., Soanes, K. H. and Brook, W. J. (2005). The Drosophila melanogaster T-box genes midline and H15 are conserved regulators of heart development. Dev. Biol. 2005 278(2): 459-72. PubMed ID: 15680363

Nasonkin, I., Alikasifoglu, A., Ambrose, C., Cahill, P., Cheng, M., Sarniak, A., Egan, M. and Thomas, P. (1999). A novel sulfonylurea receptor family member expressed in the embryonic Drosophila dorsal vessel and tracheal system. J. Biol. Chem. 274: 29420-24925. PubMed ID: 10506204

Negre, B. and Simpson, P. (2009). Evolution of the achaete-scute complex in insects: convergent duplication of proneural genes. Trends Genet. 25: 147-152. PubMed ID: 19285745

Oros, S. M., Tare, M., Kango-Singh, M. and Singh, A. (2011). Dorsal eye selector pannier (pnr) suppresses the eye fate to define dorsal margin of the Drosophila eye. Dev. Biol. 346(2): 258-71. PubMed ID: 20691679

Page, B. D., et al. (1997). ELT-1, a GATA-like transcription factor, is required for epidermal cell fates in Caenorhabditis elegans embryos. Genes Dev. 11:1651-1661. PubMed ID: 9224715

Partington, G. A., et al. (1997). GATA-2 is a maternal transcription factor present in Xenopus oocytes as a nuclear complex which is maintained throughout early development. Dev. Biol. 181: 144-155. PubMed ID: 9013926

Pereira, P. S., Pinho, S., Johnson, K., Couso, J. P. and Casares, F. (2006). A 3' cis-regulatory region controls wingless expression in the Drosophila eye and leg primordia. Dev. Dyn. 235: 225-234. PubMed ID: 16261625

Qian, L., Liu, J. and Bodmer R. (2005). Neuromancer Tbx20-related genes (H15/midline) promote cell fate specification and morphogenesis of the Drosophila heart. Dev. Biol. 279(2): 509-24. PubMed ID: 15733676

Ramain, P., et al. (1993). pannier, a negative regulator of achatae and scute in Drosophila, encodes a zinc finger protein with homology to the vertebrate transcription factor GATA-1. Development 119: 1277-1291. PubMed ID: 7916679

Ramain, P., et al. (2000). Interactions between Chip and the Achaete/Scute-Daughterless heterodimers are required for Pannier-driven proneural patterning. Mol. Cell 6: 781-790. PubMed ID: 11090617

Rehorn, K.-P., et al. (1996). A molecular aspect of hematopoiesis and endoderm development common to vertebrates and Drosophila. Development 122: 4023-4031 . PubMed ID: 9012522

Reim, I., Mohler, J. and Frasch, M. (2005a). Tbx20-related genes, mid and H15, are required for tinman expression, proper patterning, and normal differentiation of cardioblasts in Drosophila. Mech. Dev. 122: 1056-1069. PubMed ID: 15922573

Reim, I. and Frasch, M. (2005b). The Dorsocross T-box genes are key components of the regulatory network controlling early cardiogenesis in Drosophila. Development 132: 4911-4925. PubMed ID: 16221729

Reiter, J. F., et al. (1999). Gata5 is required for the development of the heart and endoderm in zebrafish. Genes. Dev. 13: 2983-2995. PubMed ID: 10580005

Rusten, T. E., et al. (2002). The role of TGFß signaling in the formation of the dorsal nervous system is conserved between Drosophila and chordates. Development 129: 3575-3584. PubMed ID: 12117808

Sato, M, et al. (1999). Bar homeobox genes are latitudinal prepattern genes in the developing Drosophila notum whose expression is regulated by the concerted functions of decapentaplegic and wingless. Development 126: 1457-1466. PubMed ID: 10068639

Sato, M. and Saigo, K., et al. (2000). Involvement of pannier and u-shaped in regulation of Decapentaplegic-dependent wingless expression in developing Drosophila notum. Mech. Dev. 93: 127-138. PubMed ID: 10781946

Schlesinger, J., et al. (2011). The cardiac transcription network modulated by Gata4, Mef2a, Nkx2.5, Srf, histone modifications, and microRNAs. PLoS Genet. 7: e1001313. PubMed ID: 21379568

Singh, A. and Choi, K.-W. (2003). Initial state of the Drosophila eye before dorsoventral specification is equivalent to ventral. Development 130: 6351-6360. PubMed ID: 14623824

Singh, A., Chan, J., Chern, J. J., Choi, K. W. (2005). Genetic interaction of Lobe with its modifiers in dorsoventral patterning and growth of the Drosophila eye. Genetics 171(1): 169-183. PubMed ID: 15976174

Tokusumi, T., et al. (2007). U-shaped protein domains required for repression of cardiac gene expression in Drosophila. Differentiation 75: 166-174. PubMed ID: 17316386

Tomoyasu, Y., Ueno, N. and Nakamura, M. (2000). The Decapentaplegic morphogen gradient regulates the notal wingless expression through induction of pannier and u-shaped in Drosophila. Mech. Dev. 96: 37-49. PubMed ID: 10940623

Torres-Vazquez, J., et al. (2001). The transcription factor Schnurri plays a dual role in mediating Dpp signaling during embryogenesis. Development 128: 1657-1670. PubMed ID: 11290303

Valadez-Graham, V., et al. (2012). XNP/dATRX interacts with DREF in the chromatin to regulate gene expression. Nucleic Acids Res. 40(4): 1460-74. PubMed ID: 22021382

van Meyel, D. J., Thomas, J. B. and Agulnick, A. D. (2003). Ssdp proteins bind to LIM-interacting co-factors and regulate the activity of LIM-homeodomain protein complexes in vivo. Development 130: 1915-1925. PubMed ID: 12642495

Vanolst, L., Fromental-Ramain, C. and Ramain, P. (2005). Toutatis, a TIP5-related protein, positively regulates Pannier function during Drosophila neural development. Development 132(19): 4327-38. 16141224

Winick, J. et al. (1993). A GATA family transcription factor is expressed along the embryonic dorso-ventral axis in Drosophila melanogaster . Development 119: 1055-1065. PubMed ID: 7916677

Wülbeck, C. and Simpson, P. (2002). The expression of pannier and achaete-scute homologues in a mosquito suggests an ancient role of pannier as a selector gene in the regulation of the dorsal body pattern. Development 129: 3861-3871. PubMed ID: 12135924

Xu, P., et al. (2012). Spire, an actin nucleation factor, regulates cell division during Drosophila heart development. PLoS ONE 7(1): e30565. PubMed ID: 22276214

Xu, R. H., et al. (1997). Differential regulation of neurogenesis by the two Xenopus GATA-1 genes. Mol. Cell. Biol. 17: 436-443. PubMed ID: 8972224

Yang, M., Hatton-Ellis, E. and Simpson, P. (2012). The kinase Sgg modulates temporal development of macrochaetes in Drosophila by phosphorylation of Scute and Pannier. Development 139(2): 325-34. PubMed ID: 22159580

Zenvirt, S., Nevo-Caspi, Y., Rencus-Lazar, S. and Segal, D. (2008). Drosophila LIM-only is a positive regulator of transcription during thoracic bristle development. Genetics 179(4): 1989-99. PubMed ID: 18689881


pannier: Biological Overview | Evolutionary Homologs | Regulation | Developmental Biology | Effects of Mutation

date revised: 10 December 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.