tailless


DEVELOPMENTAL BIOLOGY

Embryonic

Tailless mRNA is found in the blastoderm stage [Images]. Initial activation of tll transcription is in two mirror image symmetrical caps at the poles of the embryo. This expression resolves into a posterior cap and an anterior-dorsal stripe. tll is also expressed in brain neuroblasts (Pignoni, 1990).

Early tailless expression (blastoderm stage) covers the anlage of the entire brain. Beginning approximately with the onset of gastrulation, an anterior-dorsal region with a high expression level (called HL domain) can be distinguished from a posterior-ventral domain expressing tll at a somewhat lower level. The HL domain coincides with part of the central and anterior protocerebral neurectoderm. The low expression level LL domain covers the remaining part of the protocerebral neuroectoderm. orthodenticle is expressed in a circumferential domain of the cellular blastoderm but during gastrulation becomes restricted to a domain that encompasses most of the protocerebral neurectoderm and an adjacent part of the deuterocerebral neurectoderm. All neurobasts segregating from this domain transiently express otd during stages 10 and 11. buttonhead is initially expressed in a wide domain including the anlagen of the antennal, intercalary and mandibular segments, as well as the acron. With the beginning of gastrulation, expression disappears from most of the procephalon, except for small domains of the posterior part of the deuterocerebral and tritocerebral neurectoderm and a dorsoanterior patch that partially overlaps with the dorsoanterior protocerebrum. Both the late deutocerebral and tritocerebral btd domains contain few, if any neuroblasts. empty spiracles is in an asymmetric circumferential domain of the cellular blastoderm. During gastrulation, this pattern resolves into two stripes that occupy anterior portions of the deuterocerebral neuroectoderm and the mandibular metamere, respectively. In addition, a small circular domain corresponding to the tritocerebral neurectoderm appears ventral to the deuterocerebral stripe (Younossi-Hartenstein, 1997).

Loss of tll function results in the absence of all protocerebral neuroblasts and loss of all four coherent domains of Fas II expression in the protocerebrum. Also missing is the optic lobe. orthodenticle functions in a domain that includes a large part of the protocerebrum and a smaller part of the adjacent deuterocerebrum. Loss of otd results in loss of protocerebral P1, P2 and P4 coherent domains of Fas II expression. Also missing is a nerve that carries axons from the antennal organ. In buttonhead mutation the D/T cluster is missing; consequently a cervical connection is missing that normally sends nerves to the labral sensory organ, the hypopharyngeal sensory organ and the stomatogastric nervous system (Younossi-Hartenstein, 1997).

The Drosophila brain develops from the procephalic neurogenic region of the ectoderm. About 100 neural precursor cells (neuroblasts) delaminate from this region on either side in a reproducible spatiotemporal pattern. Neuroblast maps have been prepared from different stages of the early embryo (stages 9, 10 and 11, when the entire population of neuroblasts has formed), in which about 40 molecular markers representing the expression patterns of 34 different genes are linked to individual neuroblasts. In particular, a detailed description is presented of the spatiotemporal patterns of expression in the procephalic neuroectoderm and in the neuroblast layer of the gap genes empty spiracles, hunchback, huckebein, sloppy paired 1 and tailless; the homeotic gene labial; the early eye genes dachshund, eyeless and twin of eyeless; and several other marker genes (including castor, pdm1, fasciclin 2, klumpfuss, ladybird, runt and unplugged). Based on the combination of genes expressed, each brain neuroblast acquires a unique identity, and it is possible to follow the fate of individual neuroblasts through early neurogenesis. Furthermore, despite the highly derived patterns of expression in the procephalic segments, the co-expression of specific molecular markers discloses the existence of serially homologous neuroblasts in neuromeres of the ventral nerve cord and the brain. Taking into consideration that all brain neuroblasts are now assigned to particular neuromeres and individually identified by their unique gene expression, and that the genes found to be expressed are likely candidates for controlling the development of the respective neuroblasts, these data provide a basic framework for studying the mechanisms leading to pattern and cell diversity in the Drosophila brain, and for addressing those mechanisms that make the brain different from the truncal CNS (Urbach, 2003).

The cephalic gap genes are expressed in large domains of the procephalon and play a crucial role not only in the patterning of the peripheral ectoderm, but also in regionalizing the brain primordium. The segmental organization of the Drosophila brain is based on the expression pattern of segment polarity and DV patterning genes. To see whether the cephalic gap genes respect the neuromeric boundaries segment polarity and DV patterning genes, and to provide a basis for studying their potential role in the formation or specification of brain precursor cells, the expression was studied of orthodenticle, empty spiracles, sloppy paired 1, tailless, huckebein, and hunchback in the developing head ectoderm, as well as in the entire population of identified NBs during stages 9-11 (Urbach, 2003).

tailless (tll) has been shown to be expressed in an anterior horseshoe-shaped stripe in the cellular blastoderm, which after gastrulation shows a region of high ('HL domain') and a region of low level of tll expression ('LL domain'), and at stage 9 covers most of the protocerebral neuroectoderm. Using a tll-lacZ line at stage 9 tll expression has been found in the developing brain in most protocerebral NBs (except the dorsoposterior ones). During stages 9-11 tll-lacZ expression expands in the protocerebral neuroectoderm beyond the En-positive head spot (hs). By stage 11 it is detectable in all protocerebral NBs. In addition, tll-lacZ is found in some ventral and dorsal deutocerebral NBs, indicating that tll is not exclusively confined to protocerebral progenitors (Urbach, 2003).

The development of the head and tail regions of the Drosophila embryo is dependent upon the localized polar activation of Torso (Tor), a receptor tyrosine kinase that is uniformly distributed in the membrane of the developing embryo. Trunk (Trk), the proposed ligand for Tor, is secreted as an inactive precursor into the perivitelline fluid that lies between the embryonic membrane and the vitelline membrane (VM), the inner layer of the eggshell. The spatial regulation of Trk processing is thought to be mediated by the secreted product of the torsolike (tsl) gene, which is expressed during oogenesis by a specialized population of follicle cells present at the two ends of the oocyte. Tsl protein has been shown to be specifically localized to the polar regions of the VM in laid eggs. Although Tsl can associate with nonpolar regions of the VM, the activity of polar-localized Tsl is enhanced, suggesting the existence of another spatially restricted factor acting in this pathway. The incorporation of Tsl into the VM provides a mechanism for the transfer of spatial information from the follicle cells to the developing embryo. Tsl represents the first example of an embryonic patterning determinant that is a component of the eggshell (Stevens, 2003).

Activation of Tsl may be spatially regulated through the action of gap genes. Spatial regulation of the expression of the gap genes, the first zygotic patterning genes to be expressed during embryogenesis, is determined by the activity of the three maternal pathways (anterior, posterior, and terminal) required for the development of the anterior-posterior axis of the embryo. In addition to localized maternal input, interactions between the gap gene products themselves led to further refinement of their expression domains. tll expression, for example, is specifically repressed in the segmented region of the embryo by central gap gene products such as Kr. Thus, depending on their relative levels of activity, Kr and Tll are both capable of suppressing one another's expression. This raises the possibility that centrally expressed Kr is responsible for the polar restriction of tll expression that is observed in the progeny of tsl mutant females expressing tsl from the germline. To address this question, the CBBtsl insertion was crossed into females that were mutant for all three anterior-posterior maternal pathways. Embryos produced by mothers triply mutant for bicoid (anterior), oskar (posterior), and tsl (terminal) lack all anterior-posterior patterning, express low levels of Kr uniformly along the anterior-posterior axis, and do not express tll at all. In contrast, the embryos produced by triply mutant females carrying CBBtsl did express tll in distinct polar domains, either at the anterior alone or at both poles. Consistent with this pattern of tll expression, Kr expression is specifically repressed in the corresponding polar domains. Further, these embryos also differentiate Filzkörper material, a posterior cuticular structure that requires terminal pathway activity, at one or both poles. Thus, although Tsl was distributed uniformly in their VMs, these embryos developed gap gene expression patterns and cuticular phenotypes consistent with polar activation of the Tor receptor. These findings suggest that the activity of Tsl is enhanced at, and perhaps restricted to, the polar regions of the VM; this finding implies that there is an as yet unidentified component of the terminal class pathway that is restricted to the poles and is required for the function of Tsl (Stevens, 2003).

The existence of another localized factor in this pathway indicates that there are at least four levels of control that ensure the polar restriction of tll expression during embryonic development. (1) First to act is the restriction of tsl expression to a specific subpopulation of follicle cells present at the poles of the oocyte. (2) Next is the stabilization of secreted Tsl protein at the poles of the VM and its incorporation into the eggshell in an active form. (3)The facilitation of Tsl function follows, through its proposed interaction with another localized factor. (4)The final layer of control is the exclusion of tll expression from nonpolar regions through the inhibitory effects of centrally expressed gap genes. Although it has long been assumed that the spatial restriction of tsl expression was the uniquely localized element in the terminal pathway, data is presented implying the existence of another factor that enhances the activity of Tsl specifically at the poles. The function of Tsl itself is unknown, and there are currently no candidate genes encoding proteins with the enzymatic activity to bring about the proposed processing of the Trk precursor. It is likely, therefore, that the identification of this factor will greatly enhance understanding of the mechanism by which the Tor ligand is formed (Stevens, 2003).

Different combinations of gap repressors for common stripes in Anopheles and Drosophila embryos

Drosophila segmentation is governed by a well-defined gene regulation network. The evolution of this network was investigated by examining the expression profiles of a complete set of segmentation genes in the early embryos of the mosquito, Anopheles gambiae. There are numerous differences in the expression profiles as compared with Drosophila. The germline determinant Oskar is expressed in both the anterior and posterior poles of Anopheles embryos but is strictly localized within the posterior plasm of Drosophila. The gap genes hunchback and giant display inverted patterns of expression in posterior regions of Anopheles embryos, while tailless exhibits an expanded pattern as compared with Drosophila. These observations suggest that the segmentation network has undergone considerable evolutionary change in the dipterans and that similar patterns of pair-rule gene expression can be obtained with different combinations of gap repressors. The evolution of separate stripe enhancers in the eve loci of different dipterans is discussed (Goltsev, 2004).

In Drosophila, different levels of the Hunchback and Knirps gap repressor gradients define the limits of eve stripes 3, 4, 6, and 7, while Giant and Kruppel establish the borders of stripes 2 and 5. In situ hybridization probes were prepared for Anopheles orthologues of all four of these gap genes, as well as a fifth gap gene, tailless. hunchback displays a broad band of expression in the anterior half of the Anopheles embryo, encompassing both the presumptive head and thorax. This pattern is similar to that observed in Drosophila, although there are a few notable deviations: (1) there is no obvious maternal expression seen in early Anopheles embryos, whereas maternal hunchback mRNAs are strongly expressed throughout early Drosophila embryos; (2) there is a significant change in the posterior staining pattern. The Drosophila gene displays a strong posterior stripe of expression that is comparable in intensity to the anterior staining pattern. In Anopheles, this staining is significantly weaker than that of the anterior domain, and the posterior pattern is shifted anteriorly into the presumptive abdomen (Goltsev, 2004).

The Kruppel and knirps staining patterns are similar in Anopheles and Drosophila embryos. In both cases, the principal sites of expression are seen in the presumptive thorax and abdomen, respectively. However, the remaining two gap genes, giant and tailless, exhibit distinctive staining patterns. In Anopheles, giant exhibits a continuous band of staining in anterior regions, whereas the Drosophila gene is excluded from the anterior pole. Moreover, there is a prominent band of staining in the presumptive abdomen of Drosophila embryos that is not seen in Anopheles. Finally, tailless is expressed in a narrow stripe in the posterior pole of Drosophila embryos, whereas Anopheles embryos display a dynamic pattern that (transiently) extends throughout the presumptive abdomen (Goltsev, 2004).

The altered patterns of hunchback and giant expression in posterior regions raise the possibility that different combinations of gap repressors are used to establish eve stripes 5, 6, and 7 in Anopheles and Drosophila. It is unlikely that Giant establishes the posterior border of eve stripe 5 and that Hunchback delimits the posterior border of stripe 7, as seen in Drosophila. The expression profiles of additional gap genes were analyzed in an effort to identify potential repressors for these stripe borders. The most obvious candidates are huckebein and tailless, since both are expressed in the posterior pole of Drosophila embryos. No expression of huckebein was seen in early embryos, although strong staining appears after germband elongation (Goltsev, 2004).

The gap gene tailless is initially detected at the anterior and posterior poles, with roughly equivalent levels of staining at the two sites. At slightly later stages, the anterior domain is lost, and the posterior pattern expands throughout the presumptive abdomen. The tailless transcripts detected in posterior regions exhibit a graded distribution, with peak levels at the posterior pole and progressively lower levels in more anterior regions. During cellularization, staining is reduced in posterior regions and reappears near the anterior pole. This broad and dynamic staining pattern is consistent with the possibility that the Tailless repressor specifies the posterior borders of one or more posterior eve stripes (Goltsev, 2004).

Torso signaling was examined in the Anopheles embryo in an effort to understand the basis for the expanded tailless expression pattern. In Drosophila, tailless is activated by the Torso signaling pathway, which can be visualized with an antibody against diphospho (dp)-ERK. The antibody detects localized staining in the terminal regions of early Drosophila embryos. A similar staining pattern is detected in Anopheles, although staining may be somewhat broader in Anopheles than Drosophila. It is therefore conceivable that the expansion of the posterior tailless expression pattern seen in Anopheles might be due to an expanded activation of the Torso signaling pathway (Goltsev, 2004).

In Drosophila, eve stripes 6 and 7 are regulated by different concentrations of Knirps and Hunchback. Low levels of Knirps define the anterior border of stripe 7, while higher levels are needed to repress eve stripe 6. Conversely, low levels of Hunchback establish the posterior border of eve stripe 6, while higher levels regulate stripe 7. The position of the knirps expression pattern is consistent with the possibility that it defines the anterior limits of stripes 6 and 7, just as in Drosophila. However, the posterior borders of these stripes are probably not regulated by Hunchback. The expanded pattern of tailless expression seen in Anopheles might permit it to establish the posterior border of eve stripe 6 and possibly stripe 7. An alternative candidate for the posterior stripe 7 border is giant, which is expressed in a tight domain within the posterior pole. Consistent with this possibility is the observation that the posterior giant pattern comes on relatively late, and the posterior stripe 7 border is the last to form among the seven eve stripes. The reversal of the posterior hunchback and giant expression patterns, along with the expanded tailless pattern, strongly suggests that different combinations of gap repressors are used to define eve stripes 5, 6, and 7 in Drosophila and Anopheles (Goltsev, 2004).

Temporal patterning of Drosophila medulla neuroblasts controls neural fates

In the Drosophila optic lobes, the medulla processes visual information coming from inner photoreceptors R7 and R8 and from lamina neurons. It contains approximately 40,000 neurons belonging to more than 70 different types. This study describes how precise temporal patterning of neural progenitors generates these different neural types. Five transcription factors - Homothorax, Eyeless, Sloppy paired, Dichaete and Tailless - are sequentially expressed in a temporal cascade in each of the medulla neuroblasts as they age. Loss of Eyeless, Sloppy paired or Dichaete blocks further progression of the temporal sequence. Evidence is provided that this temporal sequence in neuroblasts, together with Notch-dependent binary fate choice, controls the diversification of the neuronal progeny. Although a temporal sequence of transcription factors had been identified in Drosophila embryonic neuroblasts, this work illustrates the generality of this strategy, with different sequences of transcription factors being used in different contexts (Li, 2013).

In the developing medulla, the wave of conversion of neuroepithelium into neuroblasts makes it possible to visualize neuroblasts at different temporal stages in one snapshot, with newly generated neuroblasts on the lateral edge and the oldest neuroblasts on the medial edge of the expanding crescent shaped neuroblast region. An antibody screen was conducted for transcription factors expressed in the developing medulla and five transcription factors, Hth, Ey, Slp1, D and Tll, were identified that are expressed in five consecutive stripes in neuroblasts of increasing ages, with Hth expressed in newly differentiated neuroblasts, and Tll in the oldest neuroblasts. This suggests that these transcription factors are sequentially expressed in medulla neuroblasts as they age. Neighbouring transcription factor stripes show partial overlap in neuroblasts with the exception of the D and Tll stripes, which abut each other. Previous studies have reported that Hth and Ey< were expressed in medulla neuroblasts, but they had not been implicated in controlling neuroblast temporal identities. Hth and Tll also show expression in the neuroepithelium (Li, 2013).

To address whether each neuroblast sequentially expresses the five transcription factors, their expression was examined in the neuroblast progeny. Hth, Ey and Slp1 are expressed in three different layers of neurons that correlate with birth order, that is, Hth in the first-born neurons of each lineage in the deepest layers; Ey or Slp1 in correspondingly more superficial layers, closer to the neuroblasts. This suggests that they are born sequentially in each lineage. D is expressed in two distinct populations of neurons. The more superficial population inherit D from D+ neuroblasts. D+ neurons in deeper layers (corresponding to the Hth and Ey layers) turn on D expression independently and will be discussed later. Single neuroblast clones were generated, and the expression of the transcription factors was examined in the neuroblast and its progeny. Single neuroblast clones in which the neuroblast is at the Ey+ stage include Ey+ GMCs/neurons as well as Hth+ neurons. This indicates that Ey+ neuroblasts have transited through the Hth+ stage and generated Hth+ neurons. Clones in which the neuroblast is at the D+ stage contain Slp1+ GMCs and Ey+ neurons, suggesting that D+ neuroblasts have already transited through the Slp+ and Ey+ stages. This supports the model that each medulla neuroblast sequentially expresses Hth, Ey, Slp1 and D as it ages, and sequentially produces neurons that inherit and maintain expression of the transcription factor (Li, 2013).

slp1 and slp2 are two homologous genes arranged in tandem and function redundantly in embryonic and eye development. Slp2 is expressed in the same set of medulla neuroblasts as Slp1. Slp1 and Slp2 are referred to collectively as Slp (Li, 2013).

Tll is expressed in the oldest medulla neuroblasts. The oldest Tll+ neuroblasts show nuclear localization of Prospero (Pros), suggesting that they undergo Pros-dependent cell-cycle exit at the end of their life, as in larval nerve cord and central brain neuroblasts. Tll+ neuroblasts and their progeny express glial cells missing (gcm), and the progeny gradually turn off Tll and turn on Repo, a glial-specific marker. These cells migrate towards deeper neuronal layers and take their final position as glial cells around the medulla neuropil. Thus, Tll+ neuroblasts correspond to previously identified glioblasts between the optic lobe and central brain that express gcm and generate medulla neuropil glia. Clones in which the neuroblast is at the Tll+ stage contain Hth+ neurons and Ey+ neurons, among others, confirming that Tll+ neuroblasts represent the final temporal stage of medulla neuroblasts rather than a separate population of glioblasts. Therefore, these data clearly show that medulla neuroblasts sequentially express five transcription factors as they age. The four earlier temporal stages generate neurons that inherit and maintain the temporal transcription factor present at their birth, although a subset of neurons born during the Ey, Slp or D neuroblast stages lose expression of the neuroblast transcription factor. At the final temporal stage, neuroblasts switch to glioblasts and then exit the cell cycle (Li, 2013).

Whether cross-regulation among transcription factors of the neuroblast temporal sequence contributes to the transition from one transcription factor to the next was examined. Loss of hth or its cofactor, extradenticle (exd), does not affect the expression of Ey and subsequent progression of the neuroblast temporal sequence (Li, 2013).

ey-null mutant clones were generated using a bacterial artificial chromosome (BAC) rescue construct recombined on a chromosome containing a Flip recombinase target (FRT) site in an eyJ5.71 null background. eyJ5.71 homozygous mutant larvae were also tested. In both cases, Slp expression is lost in neuroblasts, along with neuronal progeny produced by Slp+ neuroblasts, marked by the transcription factor Twin of eyeless (Toy, see below). However, neuroblast division is not affected, and Hth remains expressed in only the youngest neuroblasts and first-born neurons. Targeted ey RNA interference (RNAi) using a Vsx-Gal4 driver that is expressed in the central region of the neuroepithelium and neuroblasts gives the same phenotype. This suggests that Ey is required to turn on the next transcription factor, Slp, but is not required to repress Hth (Li, 2013).

In clones of a deficiency mutation, slpS37A, that deletes both slp1 and slp2, neuroblasts normally transit from Hth+ to Ey+, but older neuroblasts maintain the expression of Ey and do not progress to express D or Tll, suggesting that Slp is required to repress ey and activate D (Li, 2013).

Similarly, in D mutant clones, neuroblasts are also blocked at the Slp+ stage, and do not turn on Tll, indicating that D is required to repress slp and activate tll. Finally, in tll mutant clones, D expression is not expanded into oldest neuroblasts, suggesting that tll is not required for neuroblasts to turn off D. Thus, in the medulla neuroblast temporal sequence, ey, slp and D are each required for turning on the next transcription factor. slp and D are also required for turning off the preceding transcription factor (Li, 2013).

Gain-of-function phenotypes of each gene were studied. However, misexpression of Hth, Ey, Slp1 or Slp2, or D in all neuroblasts or in large neuroblast clones is not sufficient to activate the next transcription factor or repress the previous transcription factor in neuroblasts. Only misexpressing tll in all neuroblasts is sufficient to repress D expression (Li, 2013).

In summary, cross-regulation among transcription factors is required for at least some of the transitions. No cross-regulation was observed between hth and ey. Because ey is already expressed at low levels in the neuroepithelium and in Hth+ neuroblasts, an as yet unidentified factor might gradually upregulate ey and repress hth to achieve the first transition. As tll is sufficient but not required to repress D expression, additional factors must act redundantly with Tll to repress D (Li, 2013).

The temporal sequence of neuroblasts described above could specify at least four neuron types plus glia (in fact more than ten neuron types plus glia considering that neuroblasts divide several times at each stage with overlaps between neighbouring temporal transcription factors). As this is not sufficient to generate the 70 medulla neuron types, it was asked whether another process increases diversity in the progeny neurons born from a neuroblast at a specific temporal stage. Apterous (Ap) is known to mark about half of the 70 medulla neuron types. In the larval medulla, Ap is expressed in a salt-and-pepper manner in subsets of neurons born from all temporal stages. In the progeny from Hth+ neuroblasts, all neurons seem to maintain Hth, with a subset also expressing Ap. However, only half of the neurons born from neuroblasts at other transcription factor stages maintain expression of the neuroblast transcription factor. For instance, in the progeny of Ey+ neuroblasts, Ey+ neurons are intermingled with about an equal number of Ey neurons that instead express Ap. Neuroblast clones contain intermingled Ey+ and Ap+ neurons. This is also true for the progeny of Slp+ neuroblasts: Slp1+ neurons are intermingled with Slp1 Ap+ neurons. In the progeny of D+ neuroblasts, D and Ap are co-expressed in the same neurons, and they are intermingled with neurons that express neither D nor Ap. Neurons in deeper neuronal layers (corresponding to the Ey+ and Hth+ neuron layers) also express D independently, and these neurons are Ap. The expression of Ap is stable from larval to adult stages (Li, 2013).

The intermingling of Ap+ and Ap neurons raised the possibility that asymmetric division of GMCs gives rise to one Ap+ and one Ap neuron. Two-cell clones were generated to visualize the two daughters of a GMC. In every case, one neuron is Ap+ and the other is Ap-, suggesting that asymmetric division of GMCs diversifies medulla neuron fates by controlling Ap expression (Li, 2013).

Asymmetric division of GMCs in Drosophila involves Notch (N)-dependent binary fate choice. In the developing medulla, the N pathway is involved in the transition from neuroepithelium to neuroblast, and loss of Su(H), the transcriptional effector of N signalling, leads to faster progression of neurogenesis and neuroblast formation. However, Su(H) mutant neuroblasts still follow the same transcription factor sequence and generate GMCs and neuronal progeny, allowing analysis of the effect of loss of N function on GMC progeny diversification. Notably, neurons completely lose Ap expression in Su(H) mutant clones. All mutant neurons born during the Hth+ stage still express Hth, but not Ap, suggesting that the NON daughters of Hth+ GMCs are the neurons expressing both Ap and Hth. In contrast to wild-type clones, all Su(H) mutant neurons born during the Ey+ neuroblast stage express Ey and none express Ap. Similarly, all mutant neurons born during the Slp+ neuroblast stage express Slp1 but lose Ap. These data suggest that, for Ey+ or Slp+ GMCs, the NOFF daughter maintains the neuroblast transcription factor expression, whereas the NON daughter loses this expression but expresses Ap. In the wild-type progeny born during the D+ neuroblast stage, Ap+ neurons co-express D. Both D and Ap are lost in Su(H) mutant clones in the D+ neuroblast progeny, confirming that D is transmitted to the Ap+ NON daughter of D+ GMCs. By contrast, the D+ Ap neurons in the deeper layers (corresponding to the NOFF progeny born during the Ey+ and Hth+ neuroblast stages, see above) are expanded in Su(H) mutant clones at the expense of Ap+ neurons. Therefore, the deeper layer of D expression is turned on independently in the NOFF daughters of Hth+ and Ey+ GMCs (Li, 2013).

Finally, in wild type, a considerable amount of apoptotic cells were observed dispersed among neurons, suggesting that one daughter of certain GMCs undergoes apoptosis in some of the lineages. Together these data suggest that Notch-dependent asymmetric division of GMCs further diversifies neuronal identities generated by the temporal sequence of transcription factors (Li, 2013).

How does the neuroblast transcription factor temporal sequence, together with the Notch-dependent binary fate choice, control neuronal identities in the medulla? Transcription factor markers specifically expressed in subsets of medulla neurons, but not in neuroblasts, were examined including Brain-specific homeobox (Bsh) and Drifter (Dfr), as well as other transcription factors identified in the antibody screen, for example, Lim3 and Toy. Bsh is required and sufficient for the Mi1 cell fate, and Dfr is required for the morphogenesis of nine types of medulla neurons, including Mi10, Tm3, TmY3, Tm27 and Tm27Y (Hasegawa, 2011). Investigation were carried out to identify at which neuroblast temporal stage these neurons were born by examining co-expression with the inherited neuroblast transcription factors. Then whether the neuroblast transcription factors regulate expression of these markers and neuron fates was investigated. The results for each neuroblast stage are described below (Li, 2013).

Bsh is expressed in a subset of Hth+ neurons, suggesting that Bsh is in the NON daughter of Hth+ GMCs. Indeed, Bsh expression is lost in both Su(H) and hth mutant clones. Thus, both Notch activity and Hth are required for specifying the Mi1 fate, consistent with the previous report that Hth is required for the Mi1 fate. Ectopic expression of Hth in older neuroblasts is also sufficient to generate ectopic Bsh+ neurons, although the phenotype becomes less pronounced in later parts of the lineage. These data suggest that Hth is necessary and sufficient to specify early born neurons, but the competence to do so in response to sustained expression of Hth decreases over time. This is similar to embryonic CNS neuroblasts, where ectopic Hb is only able to specify early born neurons during a specific time window (Li, 2013).

Lim3 is expressed in all Ap progeny of both Hth+ and Ey+ neuroblasts. Toy and Dfr are expressed in subsets of neurons born from Ey+ neuroblasts, as indicated by their expression in the Ey+ neuron progeny layer. The most superficial row of Ey+ Ap neurons express Toy (and Lim3), suggesting that they are the NOFF progeny of the last-born Ey+ GMCs. Dfr is co-expressed with Ap in two or three rows of neurons that are intermingled with Ey+ neurons, suggesting that they are the NON progeny from Ey+ GMCs. In addition to these Ap+ Dfr+ neurons, Dfr is also expressed in some later-born neurons that are Ap but express another transcription factor: Dachshund (Dac), in specific sub-regions of the medulla crescent (Li, 2013).

Whether Ey in neuroblasts regulates Dfr expression in neurons was tested. As expected, Dfr-expressing neurons are lost in ey-null mutant clones, suggesting that they require Ey activity in neuroblasts, even though Ey is not maintained in Ap+ Dfr+ neurons. Furthermore, in slp mutant clones in which neuroblasts remain blocked in the Ey+ state, the Ap+ Dfr+ neuron population is expanded into later-born neurons, suggesting that the transition from Ey+ to Slp+ in neuroblasts is required for shutting off the production of Ap+ Dfr+ neurons. In addition, Ap+ Dfr+ neurons are lost in Su(H) mutant clones. Thus, Ey expression in neuroblasts and the Notch pathway together control the generation of Ap+ Dfr+ neurons (Li, 2013).

In addition to its expression with Ey in the NOFF progeny of the last-born Ey+ GMCs, Toy is also expressed in Ap+ (NON) neurons in more superficial layers generated by Slp+ and D+ neuroblasts. Consistently, in Su(H) mutant clones, an expansion of Toy+ Ey+ neurons is seen in the Ey progeny layer, followed by loss of Toy in the Slp and D progeny layer (Li, 2013).

Tests were performed to see whether Slp is required for the neuroblasts to switch from generating Toy+ Ap neurons, progeny of Ey+ neuroblasts, to generating Toy+ Ap+ neurons. Indeed, in slp mutant clones, the Toy+ Ap+ neurons largely disappear, whereas Toy+ Ap neurons expand (Li, 2013).

WAp and Toy expression was examined in specific adult neurons. OrtC1-gal4 primarily labels Tm20 and Tm5 plus a few TmY10 neurons, and these neurons express both Ap and Toy. To examine whether Slp is required for the specification of these neuron types, wild-type or slp mutant clones were generated using the mosaic analysis with a repressible cell marker (MARCM) technique by heat-shocking for 1 h at early larval stage, and the number of OrtC1-gal4-marked neurons in the adult medulla was examined. In wild-type clones, OrtC1-gal4 marks ~100 neurons per medulla. By contrast, very few neurons are marked by OrtC1-gal4 in slp mutant clones. Slp is unlikely to directly regulate the Ort promoter because Slp expression is not maintained in Ap+ Toy+ neurons. Furthermore, the expression level of OrtC1-gal4 in lamina L3 neurons is not affected by slp mutation. These data suggest that loss of Slp expression in neuroblasts strongly affects the generation of Tm20 and Tm5 neurons (Li, 2013).

In summary, these data show that the sequential expression of transcription factors in medulla neuroblasts controls the birth-order-dependent expression of different neuronal transcription factor markers, and thus the sequential generation of different neuron types (Li, 2013).

Although a temporal transcription factor sequence that patterns Drosophila nerve cord neuroblasts was reported more than a decade ago, it was not clear whether the same or a similar transcription factor sequence patterns neural progenitors in other contexts. The current identification of a novel temporal transcription factor sequence patterning the Drosophila medulla suggests that temporal patterning of neural progenitors is a common theme for generating neuronal diversity, and that different transcription factor sequences might be recruited in different contexts (Li, 2013).

There are both similarities and differences between the two neuroblast temporal sequences. In the Hb-Kr-Pdm-Cas-Grh sequence, ectopically expressing one gene is sufficient to activate the next gene, and repress the previous gene, but these cross-regulations are not necessary for the transitions, with the exception of Castor. In the Hth-Ey-Slp-D-Tll sequence, removal of Ey, Slp or D does disrupt cross-regulations necessary for temporal transitions (except the Hth-Ey transition). However, in most cases these cross-regulations are not sufficient to ensure temporal transitions, suggesting that additional timing mechanisms or factors are required (Li, 2013).

For simplicity, the medulla neuroblasts are represented as transiting through five transcription factor stages, whereas in fact the number of stages is clearly larger than five. First, neuroblasts divide more than once while expressing a given temporal transcription factor, and each GMC can have different sub-temporal identities. Furthermore, there is considerable overlap between subsequent temporal neuroblast transcription factors: neuroblasts expressing two transcription factors are likely to generate different neuron types from neuroblasts expressing either one alone (Li, 2013).

Although the complete lineage of medulla neuroblasts is still being investigated, this study shows how a novel temporal sequence of transcription factors is required to generate sequentially the diverse neurons that compose the medulla. The requirement for transcription factor sequences in the medulla and in embryonic neuroblasts suggests that this is a general mechanism for the generation of neuronal diversity. Interestingly, the mammalian orthologue of Slp1, FOXG1, acts in cortical progenitors to suppress early born cortical cell fates. Thus, transcription-factor-dependent temporal patterning of neural progenitors might be a common theme in both vertebrate and invertebrate systems (Li, 2013).

A region-specific neurogenesis mode requires migratory progenitors in the Drosophila visual system

Brain areas each generate specific neuron subtypes during development. However, underlying regional variations in neurogenesis strategies and regulatory mechanisms remain poorly understood. In Drosophila, neurons in four optic lobe ganglia originate from two neuroepithelia, the outer (OPC) and inner (IPC) proliferation centers. Using genetic manipulations, this study found that one IPC neuroepithelial domain progressively transformed into migratory progenitors that matured into neural stem cells (neuroblasts) in a second domain. Progenitors emerged by an epithelial-mesenchymal transition-like mechanism that required the Snail-family member Escargot and, in subdomains, Decapentaplegic signaling. The proneural factors Lethal of scute and Asense differentially controlled progenitor supply and maturation into neuroblasts. These switched expression from Asense to a third proneural protein, Atonal. Dichaete and Tailless mediated this transition, which was essential for generating two neuron populations at defined positions. It is proposed that this neurogenesis mode is central for setting up a new proliferative zone to facilitate spatio-temporal matching of neurogenesis and connectivity across ganglia. (Apitz, 2014).

Recent studies have distinguished three neurogenesis modes in the Drosophila CNS. First, type I neuroblasts arise from neuroepithelia and generate GMCs, which produce neuronal and glial progeny. Second, Dpn+ type II neuroblasts in the dorsomedial central brain go through a transit-amplifying Dpn+, Ase+ population, called intermediate neural precursors, which generate GMCs and postmitotic offspring. Third, lateral OPC neuroepithelial cells bypass the neuroblast stage and generate lamina precursor cells (LPCs) that divide once to produce lamina neurons. The current results provide evidence for a fourth strategy: p-IPC neuroepithelial cells give rise to progenitors that migrate to a second neurogenic domain, where they mature into type I neuroblasts. These progenitors are distinct, as they originate from the neuroepithelium, do not express markers for neuroblasts, intermediate neural precursors, GMCs or postmitotic neurons, and acquire NSC properties after completing their migration (Apitz, 2014).

Migratory progenitors arise from the p-IPC by a mechanism that shares cellular and molecular characteristics with EMT. On the basis of data on gastrulation and neural crest formation, EMT is commonly associated with cells adopting a mesenchymal state, enabling them to leave their epithelial tissue and migrate through the extracellular matrix to new locations. A recent study also reported an EMT-like process in the mammalian neocortex, whereby newborn neurons and intermediate progenitors delaminate from the ventricular neuroepithelium and radially migrate to the pial surface. This study observed that neuroepithelial cells at the p-IPC margins and migratory progenitors upregulated the Snail homolog Esg, whereas E-cad levels were decreased. Moreover, esg knockdown caused the formation of ectopic E-cad-expressing clusters adjacent to the p-IPC. Although this is a previously uncharacterized role of Drosophila esg, these findings are consistent with the requirement of two Snail transcription factors, Scratch1 and 2, and downregulation of E-cad in cortical EMT migration (Apitz, 2014).

Although TGFβ signaling is well known to induce EMT, it was unclear whether it could have such a role in the brain. Two lines of evidence are consistent with a requirement of the Drosophila family member Dpp. First, it is expressed and downstream signaling is activated in dorsal and ventral p-IPC subdomains and emerging cell streams. Second, tkv mutant cells form small neuroepithelial clusters in p-IPC vicinity. Similar to the neural crest, where distinct molecular cascades control delamination in the head and trunk, region-specific regulators may also be required in p-IPC subdomains. Because neuroblasts derived from Dpp-dependent cell streams map to defined areas in the d-IPC, this pathway could potentially couple EMT and neuron subtype specification (Apitz, 2014).

Cell migration is an essential feature of vertebrate brain development. Commonly, postmitotic immature neurons migrate from their proliferation zones to distant regions, where they further differentiate and integrate into local circuits. Examples include the radial migration of projection neurons and tangential migration of interneurons in the embryonic cortex, as well as migration of interneuron precursors in the rostral migratory stream to the olfactory bulb in adults. In contrast, IPC progenitors develop into NSCs (neuroblasts) after they migrated. A recent study found that NSCs relocating from the embryonic ventral hippocampus to the dentate gyrus act as source for adult NSCs in the subgranular zone. In addition, cerebellar granule cell precursors migrate from the rhombic lip to the external granule layer, where they proliferate during early postnatal development. The migration of neural cell types that become proliferative in a new niche could therefore constitute a more general strategy. IPC progenitors form streams of elongated, closely associated cells. Despite their different developmental state, their organization is notably similar to the neuronal chain network in the lateral walls of the subventricular zone and the rostral migratory stream in mammals, or of migratory trunk neural crest cells in chick. Further studies will need to identify the determinants directing migratory progenitors into the d-IPC (Apitz, 2014).

Several constraints could shape a neurogenesis mode that requires migratory progenitors in the larval optic lobe. The OPC is located superficially and the IPC is positioned centrally. If medulla and lobula neurons arose by neuroepithelial duplications, these new populations would need to be integrated into an ancestral visual circuit consisting of lamina and lobula plate neurons. Cellular migration may therefore be a derived feature and serve as an essential spatial adjustment of the IPC to the newly added medulla. In principle, the migratory population could consist of immature neurons. However, migratory progenitors help to establish a new superficial proliferative niche, and to align OPC and d-IPC neuroblast positions. This in turn enables the OPC and IPC to use spatially matching birth order-driven neurogenesis patterns for establishing functionally coherent connections across ganglia (Apitz, 2014).

IPC progenitors were primed to mature into neuroblasts, but were prevented to do so in cell streams. Consistently, progenitors showed weak cytoplasmic Mira expression and prematurely differentiated into neuroblasts following loss of Pcl. Although Dichaete has been shown to repress ase to maintain embryonic neuroectodermal cells in an undifferentiated state, this study did not identify such a role in the IPC. Future studies are therefore required to distinguish whether this block in neuroblast maturation is released in the d-IPC by cell-intrinsic mechanisms or locally acting signals (Apitz, 2014).

The p-IPC and d-IPC consecutively expressed three proneural factors. esg-positive p-IPC neuroepithelial cells transiently expressed L'sc as they converted into progenitors. Following arrival in the d-IPC, progenitors matured into neuroblasts, which switched bHLH protein expression from Ase to Ato. This correlated with a change in cell division orientations from toward the lamina to the optic lobe surface and the generation of two lineages, distal cells and lobula plate neurons. The progression of neuroblasts through two stages is supported by the observations that progenitors solely entered the lower d-IPC, all neuroblasts were labeled with Ase in this area, and idpp reporter gene expression in a progenitor subset persisted in both lower and upper d-IPC neuroblasts and their progeny (Apitz, 2014).

Late l'sc knockdown reduced the number of d-IPC neuroblasts and both neuron classes, whereas p-IPC formation and EMT of progenitors appeared to be unaffected. This supports the idea that l'sc promotes neuroblast formation by controlling the rate of conversion and the progenitor supply. In contrast, ase loss severely decreased the amount of lower d-IPC neuroblasts and distal cells. This revealed a central role in the maturation of progenitors into neuroblasts, endowing them with the potential to proliferate and generate a specific lineage. Although these functions are the opposite of those observed in the OPC, they align with the role of a murine Ase homolog, Achaete-scute homolog 1 (Ascl1), in the embryonic telencephalon. Ase- neuroblasts with type I proliferation patterns have not previously been described. Further underscoring the context-dependent activities of proneural bHLH factors, ato does not have the equivalent role of ase in conferring neurogenic properties to upper d-IPC neuroblasts, but acts upstream of differentiation programs controlling the projections of lobula plate neurons (Apitz, 2014).

Although Ase and Ato each regulated distinct aspects of d-IPC development, they were not required for either the transition or the extent of their expression domains. These functions were fulfilled by Dichaete and tll, whose cross-regulatory interactions were essential for the transition from Ase+ to Ato+, Dac+ expression. To link birth order and fate, temporal identity transcription factors are sequentially expressed by neuroblasts and inherited by GMCs and their progeny born during a given developmental window. Acting as the final two members of the OPC-specific series of temporal identity factors, Dichaete is required for Tll expression, whereas tll is sufficient, but not required, to inhibit Dichaete Although OPC and d-IPC neuroblasts shared the sequential expression of Dichaete and Tll, key differences include the fact that d-IPC progeny did not maintain Dichaete, that Tll was transiently expressed in newborn progeny of the upper d-IPC and was not maintained in older lineages, that Dichaete in the lower d-IPC was not required in its own expression domain for neurogenesis, and that Dichaete was required to activate tll, and tll to repress Dichaete and ase, as well as to independently upregulate Ato and Dac. Although the mechanisms that trigger the timing of the switch require further analysis, these observations support the notion that, in the d-IPC, Dichaete and tll do not function as temporal identity factors, but as switching factors between two sequential neuroblast stages. The vertebrate homologs of Dichaete and tll, Sox2 and Tlx, are essential for adult NSC maintenance and Sox2 positively regulates Tlx expression, suiggesting that core regulatory interactions between Dichaete and tll family members may be conserved (Apitz, 2014).

These studies uncovered molecular signatures for generating a migratory neural population by EMT and subsequent NSC development that are in part shared between the fly optic lobe and vertebrate cortical neurogenesis. The unexpected parallels suggest that ancestral gene regulatory cassettes imparting specific cellular properties may have been re-employed during vertebrate brain development. Analysis of p-IPC and d-IPC neurogenesis in the Drosophila optic lobe therefore opens new possibilities for systematically identifying genes regulating EMT, cell migration and sequential NSC specification (Apitz, 2014).

Effects of Mutation or Deletion

tailless mutants exhibit deletions in the terminal (anterior and posterior) domains of the embryo. There is a correlation between the presence of the posterior cap of tll expression and differentiation of a telson. In the anterior the maternal anterior system, Bicoid is required together with the terminal system and its target tailless to establish the acron (Pignoli, 1992). tailless is responsible in the tail for hindgut and Malpighian tubules and a large portion of the posterior midgut, hindgut and anal pads.

In mutants of huckebein and tailless, genes known to specify, respectively, adjacent posterior and anterior domains of the posterior midgut invagination, folded gastrulation transcription at the posterior pole does not extend as far anteriorly. The same lack of extention is evident in forkhead mutants. The double mutant huckebein tailless is the only double mutant combination of these three genes that completely eliminates the posterior midgut invagination; this combination also abolishes all expression of fog at the posterior pole. It is not clear how the anterior extent of fog transcription is delimited, since the domain of tailless expression and activity extends further to the anterior than the region of fog expression (Costa, 1994).

tailless mutations have little effect on hindsight expression; from analysis of huckebein tailless double mutants, it is clear that the only loss of Hnt protein expression in tailless mutants occurs in the region from which the Malpighian tubule primordia originate, consistent with the reported role for tll and hnt in the development of these structures. hkb mutant embryos lack Hnt protein expression in the regions from which the anterior and posterior midgut normally arise; expression remains only in the presumptive ureter of the Malpighian tubules. In hkb tll double mutant embryos, Hnt protein is not present at all in the domains that would form anterior and posterior midgut and Malpighian tubule primordia; however expression does occur in the amnioserosa. Germ-band retraction occurs in tll or hkb single mutants as well as in hkb tll double mutants, suggesting that midgut expression of Hnt is not necessary for germ-band retraction (Yip, 1997).

Mammalian cell culture studies have shown that several members of the nuclear receptor super family such as glucocorticoid receptor, retinoic acid receptor and thyroid hormone receptor can repress the activity of AP-1 proteins (referring to Drosophila Kayak and Jun) by a mechanism that does not require the nuclear receptor to bind to DNA directly, but that is otherwise poorly understood. Several aspects of nuclear receptor function are believed to rely on this inhibitory mechanism, which is referred to as transrepression. This study presents evidence that nuclear receptor-mediated transrepression of AP-1 occurs in Drosophila melanogaster. In two different developmental situations, embryonic dorsal closure and wing development, several nuclear receptors, including Seven up, Tailless, and Eagle antagonize AP-1. The inhibitory interactions with nuclear receptors are integrated with other modes of AP-1 regulation, such as mitogen-activated protein kinase signaling. A potential role of nuclear receptors in setting a threshold of AP-1 activity required for the manifestation of a cellular response is discussed (Gritzan, 2002).

The best understood AP-1-dependent process in Drosophila development is a coordinated cell sheet movement known as dorsal closure. During DC, lateral epidermal cells migrate dorsally and close the epidermis on the dorsal side of the embryo. Failure to undergo DC results in a characteristic dorsal open phenotype, the cuticle of affected embryos displays a dorsal hole. Mutations in genes encoding the Drosophila homologs of JNKK, (JNK, Jun and Fos) all give rise to similar dorsal open phenotypes. Thus, it is thought that DC requires activation of Jun/Fos heterodimers by a JNK-type MAPK cascade. Embryos homozygous for kay1, a fos null allele are devoid of zygotic Fos activity and DC fails. A large dorsal hole forms and the cuticle collapses. In an embryo homozygous kay2, a hypomorphic fos-allele, AP-1 activity is reduced but not eliminated. Correspondingly, the DC phenotype is weaker. The embryo displays a small dorso-anterior hole (Gritzan, 2002).

To test whether Drosophila NRs can antagonize AP-1, a variety of AP-1 constructs were in the embryonic epidermis. Interestingly, expression of some, but not all, NRs tested result in DC phenotypes of different strengths. Expression of Svp in the dorsal epidermis under the control of pnrGal4 results in a DC phenotype reminiscent of that of kay2 homozygotes. This finding is consistent with a suppression of AP-1 activity by Svp. Similarly, expression of Tll under the control of a heat shock promoter causes a weak dorsal open phenotype. The differentiation of ventral cells does not seem to be disturbed by Tll expression since the pattern of denticles in this part of the epidermis appears grossly normal. Thus, Tll expression specifically affects the dorsal epidermis where AP-1 activity is required. The expression of Knrl under the control of pnrGal4 elicits stronger DC phenotypes with the dorsal hole frequently extending over several segments (Gritzan, 2002).

Does modulation of AP-1 activity by NRs occur only in situations where AP-1 is regulated by JNK or does this type of regulation also operate in different contexts? A function for Fos downstream of ERK has been demonstrated in the differentiation of wing veins. Extra vein material can result from elevated levels of ERK, as in flies carrying a gain-of-function allele of the rolled gene, which encodes Drosophila ERK. This allele, called rolledSevenmaker (rlSem), encodes a form of ERK with increased resistance to inactivation by dephosphorylation. Expression of a dominant-negative form of Fos in the wings of rlSem flies results in loss of ectopic vein material. Conversely, overexpression of Fos enhances the extra-vein phenotype caused by rlSem. 32B Gal4, UAS Sem flies express the RlSem form of ERK in the wing from a UAS-driven transgene. As a consequence of elevated levels of ERK activity, these animals develop ectopic wing vein material. Reducing fos gene dosage in this system strongly suppresses the vein phenotype, consistent with the proposed role of Fos as an ERK effector. Thus, 32B Gal4 UAS Sem flies provide a suitable system to examine how genetic manipulations of AP-1 activity affect vein differentiation. To investigate a potential role of the Drosophila NRs in this process, one copy of kni, eg, tll or svp was removed in 32B Gal4, UAS Sem flies. Reducing kni function does not influence the vein phenotype. However, heterozygosity for any of the other three receptors tested reproducibly leads to a mild enhancement of the ectopic vein differentiation. As an unambiguously scoreable criterion to statistically evaluate phenotypic effects, the presence of ectopic vein material posterior to L5 was chosen. This area of the wing is relatively resistant to the formation of extra vein material. Quantitative analysis clearly shows that whereas the formation of extra vein material posterior to L5 in 32B Gal4 UAS Sem flies is suppressed by reducing fos activity, it is enhanced by a reduction of eg, svp or tll function. These data suggest that all three NRs antagonize AP-1 activity in wing vein differentiation, conceivably in a redundant manner (Gritzan, 2002).

It is speculated that the modulation of AP-1 activity by NRs contributes to what has recently been termed signal consolidation. Cells have to place a value on incoming signals (e.g. EGF-induced ERK activity) such that they are either answered by a biological response (e.g. the execution of a transcriptional program) or disregarded as noise. It is proposed that the modulation of AP-1 activity by NRs facilitates the interpretation of the EGF signal in wing vein differentiation by defining a threshold of ERK activity. Cells in which ERK activity does not reach this threshold do not mount an AP-1-dependent transcriptional response to the EGF signal. When transrepressional control is impaired (as in the svp, tll double mutant clones) the threshold is lowered and more cells than appropriate interpret EGF-induced ERK activity as a consolidated signal. This leads to the formation of ectopic vein material. This model is supported by the finding that the ectopic vein tissue observed in clones of tll and svp mutant tissue did arise close to the position of the endogenous veins and not randomly throughout the clonal area. Thus, the regulation of AP-1 by NRs appears to convey cell-intrinsic information (Gritzan, 2002).


REFERENCES

Apitz, H. and Salecker, I. (2015). A region-specific neurogenesis mode requires migratory progenitors in the Drosophila visual system. Nat Neurosci 18: 46-55. PubMed ID: 25501037

Bodofsky, S., Liberatore, K., Pioppo, L., Lapadula, D., Thompson, L., Birnbaum, S., McClung, G., Kartik, A., Clever, S. and Wightman, B. (2018). A tissue-specific enhancer of the C. elegans nhr-67/tailless gene drives coordinated expression in uterine stem cells and the differentiated anchor cell. Gene Expr Patterns 30: 71-81. PubMed ID: 30404043

Chen, R., Deng, X. and Zhu, S. (2022). The Ets protein Pointed P1 represses Asense expression in type II neuroblasts by activating Tailless. PLoS Genet 18(1): e1009928. PubMed ID: 35100262

Chen, Y. C., et al. (2009). The Torso signaling pathway modulates a dual transcriptional switch to regulate tailless expression. Nucleic Acids Res. 37(4): 1061-72. PubMed Citation: 19129218

Chen, Y.-J., et al. (2002). Tramtrack69 is required for the early repression of tailless expression. Mech. Dev. 116: 75-83. 12128207

Cinnamon, E., et al. (2004). Capicua integrates input from two maternal systems in Drosophila terminal patterning. EMBO J. 23: 4571-4582. 15510215

Cockerill, K.A., Billin, A.N. and Poole, S.J. (1993). Regulation of expression domains and effects of ectopic expression reveal gap gene-like properties of the linked pdm genes of Drosophila. Mech. Dev. 41: 139-153. PubMed Citation: 8518192

Costa, M., Wilson, E. T. and Wieschaus, E. (1994). A putative cell signal encoded by the folded gastrulation gene coordinates cell shape changes during Drosophila gastrulation. Cell 76 (6): 1075-1089. PubMed Citation: 8137424

Curt, J. R., Yaghmaeian Salmani, B. and Thor, S. (2019). Anterior CNS expansion driven by brain transcription factors. Elife 8. PubMed ID: 31271353

Daniel, A., Dumstrei, K., Lengyel, J. A. and Hartenstein, V. (1999). The control of cell fate in the embryonic visual system by atonal, tailless and EGFR signaling. Development 126: 2945-2954. PubMed Citation: 10357938

de las Heras, J. M. and Casanova, J. (2006). Spatially distinct downregulation of Capicua repression and Tailless activation by the Torso RTK pathway in the Drosophila embryo. Mech. Dev. 123(6): 481-6. 16753285

Deshpande, G., Calhoun, G. and Schedl, P. (2004). Overlapping mechanisms function to establish transcriptional quiescence in the embryonic Drosophila germline. Development 131: 1247-1257. 14960492

Diaz, R. J., et al. (1996). Graded effect of tailless on posterior gut development: molecular basis of an allelic series of a nuclear receptor gene. Mech. Dev. 54: 119-130. PubMed Citation: 8808411

Firth, L., et al. (2000). Identification of genomic regions that interact with a viable allele of the Drosophila protein tyrosine phosphatase Corkscrew. Genetics 156: 733-748. PubMed Citation: 11014820

Finley, K. D., et al. (1998). dissatisfaction encodes a Tailless-like nuclear receptor expressed in a subset of CNS neurons controlling Drosophila sexual behavior. Neuron 21(6): 1363-74. PubMed Citation: 9883729

Furriols, M., Sprenger, F. and Casanova, J. (1996). Variation in the number of activated torso receptors correlates with differential gene expression. Development 122: 2313-7. PubMed Citation: 8681811

Gaul, U. and Weigel, D. (1991). Regulation of Kruppel expression in the anlage of the Malpighian tubules in the Drosophila embryo. Mech. Dev. 33: 57-67. PubMed Citation: 1982922

Ghiglione, C., Perrimon, N. and Perkins, L. A. (1999). Quantitative variations in the level of MAPK activity control patterning of the embryonic termini in Drosophila. Dev. Biol. 205(1): 181-93. PubMed Citation: 9882506

Goltsev, Y., et al. (2004). Different combinations of gap repressors for common stripes in Anopheles and Drosophila embryos. Dev. Biol. 275: 435-446. 15501229

Goriely, A., et al. (1996). A functional homologue of goosecoid in Drosophila. Development 122: 1641-1650

Greenwood, S. and Struhl, G. (1997). Different levels of Ras activity can specify distinct transcriptional and morphological consequences in early Drosophila embryos. Development 124(23): 4879-4886. PubMed Citation: 9428424

Gritzan, U., Weiss, C., Brennecke, J. and Bohmann, D. (2002). Transrepression of AP-1 by nuclear receptors in Drosophila. Mech. Dev. 115: 91-100. 12049770

Gursky, V. V., Jaeger, J., Kozlov, K. N., Reinitz, J. and Samsonov, A. M. (2004). Pattern formation and nuclear divisions are uncoupled in Drosophila segmentation: Comparison of spatially discrete and continuous models. Physica D 197: 286-302. Full text: Gursky, 2004

Gutjahr, T., Frei, E. and Noll, M. (1993). Complex regulation of early paired expression: initial activation by gap genes and pattern modulation by pair-rule genes. Development 117(2): 609-623. PubMed Citation: 8330531

Hader, T., et al. (1998). Activation of posterior pair-rule stripe expression in response to maternal caudal and zygotic knirps activities. Mech. Dev. 71(1-2): 177-186. PubMed Citation: 9507113

Haecker, A., et al. (2007). Drosophila brakeless interacts with atrophin and is required for tailless-mediated transcriptional repression in early embryos. PLoS Biol. 2007 Jun;5(6):e145. PubMed citation: 17503969

Hakes, A. E. and Brand, A. H. (2020). Tailless/TLX reverts intermediate neural progenitors to stem cells driving tumourigenesis via repression of asense/ASCL1. Elife 9 pii: e53377. PubMed ID: 32073402

Hartmann, B., Reichert, H. and Walldorf, U. (2001). Interaction of gap genes in the Drosophila head: tailless regulates expression of empty spiracles in early embryonic patterning and brain development. Mech. Dev. 109(2): 161-72. 11731230

Hasegawa, E., Kitada, Y., Kaido, M., Takayama, R., Awasaki, T., Tabata, T. and Sato, M. (2011). Concentric zones, cell migration and neuronal circuits in the Drosophila visual center. Development 138: 983-993. PubMed ID: 21303851

Hollemann, T., Bellefroid, E. and Pieler, T. (1998). The Xenopus homologue of the Drosophila gene tailless has a function in early eye development. Development 125(13): 2425-2432. PubMed Citation: 9609825

Hou, X.S., Chou, T.B., Melnick, M.B. and Perrimon, N. (1995). The torso receptor tyrosine kinase can activate Raf in a Ras-independent pathway. Cell 81(1): 63-71. PubMed Citation: 7720074

Huelskamp, M. and Tautz, D. (1991). Gap genes and gradients--the logic behind the gaps. BioEssays 13: 261-268. PubMed Citation: 1679987

Jaeger, J., et al. (2004a). Dynamic control of positional information in the early Drosophila embryo. Nature 430: 368-371. 15254541

Jaeger, J., et al. (2004b). Dynamical analysis of regulatory interactions in the gap gene system of Drosophila melanogaster. Genetics 167: 1721-1737. 15342511

Jimenez, G., et al. (2000). Relief of gene repression by torso RTK signaling: role of capicua in Drosophila terminal and dorsoventral patterning. Genes Dev. 14(2): 224-31. PubMed Citation: 10652276

Kato, M. and Sternberg, P. W. (2009). The C. elegans tailless/Tlx homolog nhr-67 regulates a stage-specific program of linker cell migration in male gonadogenesis. Development 136(23): 3907-15. PubMed Citation: 19906858

Kispert, A., Herrmann, B.B., Leptin, M. and Reuter, R. (1994). Homologs of the mouse Brachyury gene are involved in the specification of posterior terminal structures in Drosophila, Tribolium and Locusta. Genes Dev. 8(18): 2137-50. PubMed Citation: 7958884

Klingler, M. and Gergen, J. P. (1993). Regulation of runt transcription by Drosophila segmentation genes. Mech Dev 43: 3-19. PubMed Citation: 8240970

Kurusu, M., Maruyama, Y., Adachi, Y., Okabe, M., Suzuki, E. and Furukubo-Tokunaga, K. (2009). A conserved nuclear receptor, Tailless, is required for efficient proliferation and prolonged maintenance of mushroom body progenitors in the Drosophila brain. Dev Biol 326(1): 224-236. PubMed ID: 19084514

La Rosee, A., et al. (1997). Mechanism and Bicoid-dependent control of hairy stripe 7 expression in the posterior region of the Drosophila embryo. EMBO J. 16(14): 4403-4411. PubMed Citation: 9250684

Leatherman, J. L., Levin, L., Boero, J. and Jongens, J. A. (2002). germ cell-less acts to repress transcription during the establishment of the Drosophila germ cell lineage. Curr. Biol. 12: 1681-1685. 12361572

Li, X., Erclik, T., Bertet, C., Chen, Z., Voutev, R., Venkatesh, S., Morante, J., Celik, A. and Desplan, C. (2013). Temporal patterning of Drosophila medulla neuroblasts controls neural fates. Nature 498: 456-462. PubMed ID: 23783517

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 Citation: 22513375

Liaw, G. J. and Lengyel, J. A. (1993). Control of tailless expression by bicoid, dorsal and synergistically interacting terminal system regulatory elements. Mech Dev 40: 47-61

Liaw, G.J., Rudolph, K.M., Huang, J.D., Dubnicoff, T., Courey, A.J. and Lengyel, J. (1995). The torso response element binds GAGA and NTF-1/Elf-1, and regulates tailless by relief of repression. Genes Dev. 9: 3163-3176

Liu, H.-K. et al (2008). The nuclear receptor tailless is required for neurogenesis in the adult subventricular zone. Genes Dev. 22: 2473-2478. PubMed Citation: 18794344

Liu, H. K., et al. (2010). The nuclear receptor tailless induces long-term neural stem cell expansion and brain tumor initiation. Genes Dev. 24(7): 683-95. PubMed Citation: 20360385

Liu, X. and Lengyel. J. A. (2000). Drosophila arc encodes a novel adherens junction-associated PDZ domain protein required for wing and eye development. Dev. Biol. 221: 419-434

Lloyd, T. E., et al. (2002). Hrs regulates endosome membrane invagination and tyrosine kinase receptor signaling in Drosophila. Cell 108: 261-26. 11832215

Lowe, C. J., et al. (2003). Anteroposterior patterning in hemichordates and the origins of the chordate nervous system. Cell 113: 853-865. 12837244

Margolis, J. S., et al. (1995). Posterior stripe expression of hunchback is driven from two promoters by a common enhancer element. Development 121: 3067-3077

Mishra, A. K., Bernardo-Garcia, F. J., Fritsch, C., Humberg, T. H., Egger, B. and Sprecher, S. G. (2018). Patterning mechanisms diversify neuroepithelial domains in the Drosophila optic placode. PLoS Genet 14(4): e1007353. PubMed ID: 29677185

Mlodzik, M. and Gehring, W.J. (1987) Hierarchy of the genetic interactions that specify the anteroposterior segmentation pattern of the Drosophila embryo as monitored by caudal protein expression. Development 101: 421-435

Mohler, J. (1995). Spatial regulation of segment polarity gene expression in the anterior terminal region of the Drosophila blastoderm embryo. Mech Dev 50: 151-161

Monaghan, A.P., Grau, E., Boack, D. and Schutz, G.(1995). The mouse homolog of the orphan nuclear receptor tailless is expressed in the developing forebrain. Development 121: 839-853

Monaghan, A. P., et al. (1997). Defective limbic system in mice lacking the tailless gene. Nature 390(6659): 515-517

Much, J. W., et al. (2000). The fax-1 nuclear hormone receptor regulates axon pathfinding and neurotransmitter expression. Development 127(4): 703-12.

Nakamura, A., Amikura, R., Mukai, M., Kobayashi, S. and Lasko, P. (1996). Requirement for a noncoding RNA in Drosophila polar granules for germ cell establishment. Science 274: 2075-2079. 8953037

Ochoa-Espinosa, A., Yucel, G., Kaplan, L., Pare, A., Pura, N., Oberstein, A., Papatsenko, D. and Small S. (2005). The role of binding site cluster strength in Bicoid-dependent patterning in Drosophila. Proc. Natl. Acad. Sci. 102(14): 4960-5. 15793007

Parouch, Z., Finley, R.L., Kidd, T., Wainwright, S.M., Ingham, P.W., Brent, R. and Ish-Horowicz, D. (1994). Groucho is required for Drosophila neurogenesis, segmentation, and sex determination and interacts directly with hairy-related bHLH proteins. Cell 79(5): 805-815

Paroush, Z., Wainwright, S. M. and Ish-Horowicz, D. (1997). Torso signalling regulates terminal patterning in Drosophila by antagonising Groucho-mediated repression. Development 124(19): 3827-3834

Perkins, T. J., Jaeger, J., Reinitz, J. and Glass, L. (2006). Reverse engineering the gap gene network of Drosophila melanogaster. PLoS Comput. Biol. 2(5): e51. 16710449

Pignoni, F., Baldarelli, R.M., Steingrimsson, E., Diaz, R.J., Patopoutian, A., Merriam, J.R. and Lengyel, J.A. (1990). The Drosophila gene tailless is expressed at the embryonic termini and is a member of the steroid receptor superfamily. Cell 62: 151-163. PubMed Citation: 2364433

Pignoni, F., Steingrimsson, E. and Lengyel, J.A. (1992). bicoid and the terminal system activate tailless expression in the early Drosophila embryo. Development 115: 239-251. PubMed Citation: 1638983

Qian, S., Capovilla, M. and Pirrotta, Vl. (1993). Molecular mechanisms of pattern formatin by the BRE enhancer of the Ubx gene. EMBO J. 12(10): 3865-3877. PubMed Citation: 8404855

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

Rivera-Pomar, R. and Jäckle, H. (1996). From gradients to stripes in Drosophila embryogenesis: Filling in the gaps. Trends Genet 12: 478-483. 8973159

Rives-Quinto, N., Komori, H., Ostgaard, C. M., Janssens, D. H., Kondo, S., Dai, Q., Moore, A. W. and Lee, C. Y. (2020). Sequential activation of transcriptional repressors promotes progenitor commitment by silencing stem cell identity genes. Elife 9. PubMed ID: 33241994

Rudolph, K. M., et al. (1997). Complex regulatory region mediating tailless expression in early embryonic patterning and brain development. Development 124(21): 4297-4308. PubMed Citation: 9334278

Sanchez, L. and Thieffry, D. (2001). A logical analysis of the gap gene system. J. Theor. Biol. 211: 115-141. 11419955

Sarin, S., Antonio, C., Tursun, B. and Hobert, O. (2009). The C. elegans Tailless/TLX transcription factor nhr-67 controls neuronal identity and left/right asymmetric fate diversification. Development 136(17): 2933-44. PubMed Citation: 19641012

Schroder, R., et al. (2000). Conserved and divergent aspects of terminal patterning in the beetle Tribolium castaneum. Proc. Natl. Acad. Sci. 97: 6591-6596. PubMed Citation: 10823887

Seydoux, G. and Dunn, M. A. (1997). Transcriptionally repressed germ cells lack a subpopulation of phosphorylated RNA polymerase II in early embryos of Caenorhabditis elegans and Drosophila melanogaster. Development 124: 2191-2201. 9187145

Sluder, A. E., Lindbloom, T. and Ruvkun, G. (1997). The Caenorhabditis elegans orphan nuclear hormone receptor gene rhr-2 functions in early embryonic development. Dev. Biol. 184: 303-319. PubMed Citation: 9133437

Stevens, L. M., et al. (2003). The Drosophila embryonic patterning determinant Torsolike is a component of the eggshell. Curr. Biol. 13: 1058-1063. 12814553

Tsai, C. C., Kramer, S. G. and Gergen, J. P. (1998). Pair-rule gene runt restricts orthodenticle expression to the presumptive head of the Drosophila embryo. Dev. Genet. 23(1): 35-44. PubMed Citation: 9706692

Tsuda, L., Inoue, Y.H., Yoo, M.A., Mizuno, M., Hata, M., Lim, Y.M., Adachi, T., Ryo, H., Masamune, Y. and Nishida, Y. (1993). A protein kinase similar to MAP kinase activator acts downstream of the raf kinase in Drosophila. Cell 72(3): 407-414. PubMed Citation: 8381718

Tsurumi, A., et al. (2011). STAT is an essential activator of the zygotic genome in the early Drosophila embryo. PLoS Genet. 7(5): e1002086. PubMed Citation: 21637778

Tuckfield, A., et al. (2002). Binding of the RING polycomb proteins to specific target genes in complex with the grainyhead-like family of developmental transcription factors. Mol. Cell. Biol. 22: 1936-1946. 11865070

Urbach, R. and Technau, G. M. (2003). Molecular markers for identified neuroblasts in the developing brain of Drosophila. Development 130: 3621-3637. 12835380

Van Doren, M., Williamson, A. L. and Lehmann, R. (1998). Regulation of zygotic gene expression in Drosophila primordial germ cells. Curr. Biol. 8: 243-246. 9501989

Wang, L. and Coulter, D. E. (1996). bowel, an odd-skipped homolog, functions in the terminal pathway during Drosophila embryogenesis. EMBO J. 15: 3182-3196. 8670819

Wang, L., et al. (2006). Histone deacetylase-associating Atrophin proteins are nuclear receptor corepressors. Genes Dev. 20: 525-530. 16481466

Wimmer, E. A., et al., (1995). Trans- and cis-acting requirements for blastodermal expression of the head gap gene buttonhead. Mech. Dev. 53: 235-245. PubMed Citation: 8562425

Wu, L. H. and Lengyel, J. A. (1998). Role of caudal in hindgut specification and gastrulation suggests homology between Drosophila amnioproctodeal invagination and vertebrate blastopore. Development 125: 2433-2442. PubMed Citation: 9609826

Yaghmaeian Salmani, B., Monedero Cobeta, I., Rakar, J., Bauer, S., Curt, J. R., Starkenberg, A. and Thor, S. (2018). Evolutionarily conserved anterior expansion of the central nervous system promoted by a common PcG-Hox program. Development 145(7). PubMed ID: 29530878

Yip, M.L, Lamka, M. L. and Lipshitz, H. D. (1997). Control of germ-band retraction in Drosophila by the zinc-finger protein HINDSIGHT. Development 124 (11): 2129-2141. PubMed Citation: 9187140

Younossi-Hartenstein, A., et al. (1997). Control of early neurogenesis in the Drosophila brain by the head gap genes tll, otd, ems, and btd. Dev. Biol 182: 270-283. PubMed Citation: 9070327

Yu, R. T., et al. (1994). Relationship between Drosophila gap gene tailless and a vertebrate nuclear receptor Tlx. Nature 370: 375-379. PubMed Citation: 8047143

Yu, R. T., et al. (2000). The orphan nuclear receptor Tlx regulates Pax2 and is essential for vision. Proc. Natl. Acad. Sci. 97: 2621-2625. PubMed Citation: 10706625

Zhang, C. L., Zou. Y., Yu, R. T., Gage, F. H. and Evans, R. M. (2006). Nuclear receptor TLX prevents retinal dystrophy and recruits the corepressor atrophin1. Genes Dev. 20: 1308-1320. PubMed citation


tailless: Biological Overview | Evolutionary Homologs | Regulation | Targets of Activity | Developmental Biology | Effects of Mutation

date revised: 15 April 2020

Home page: The Interactive Fly © 2020 Thomas B. Brody, Ph.D.

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