thread


DEVELOPMENTAL BIOLOGY

Embryonic

Trophic mechanisms in which neighboring cells mutually control their survival by secreting extracellular factors play an important role in determining cell number. However, how trophic signaling suppresses cell death is still poorly understood. The survival of a subset of midline glia cells in Drosophila depends upon direct suppression of the proapoptotic protein HID via the EGF receptor/RAS/MAPK pathway. The TGFalpha-like ligand SPITZ is activated in the neurons, and glial cells compete for limited amounts of secreted SPITZ to survive. In midline glia that fail to activate the EGFR pathway, HID induces apoptosis by blocking a caspase inhibitor, Diap1/Thread. Therefore, a direct pathway linking a specific extracellular survival factor with a caspase-based death program has been established (Bergmann, 2002).

Embryonic onset of late replication requires Cdc25 down-regulation

The Drosophila midblastula transition (MBT), a major event in embryogenesis, remodels and slows the cell cycle. In the pre-MBT cycles, all genomic regions replicate simultaneously in rapid S phases that alternate with mitosis, skipping gap phases. At the MBT, down-regulation of Cdc25 phosphatase and the resulting inhibitory phosphorylation of the mitotic kinase Cdk1 create a G2 pause in interphase 14. However, an earlier change in interphase 14 is the prolongation of S phase. While the signals modifying S phase are unknown, the onset of late replication (where replication of constitutively heterochromatic satellite sequences is delayed), extends S-phase 14. Cdc25 mRNA was injected to bypass the developmentally programmed down-regulation of Cdc25 at the MBT. Introduction of either Cdc25 isoform (String or Twine) or enhanced Cdk1 activity triggered premature replication of late-replicating sequences, even after their specification, and thereby shortened S phase. Reciprocally, reduction of Cdk1 activity by knockdown of mitotic cyclins extended pre-MBT S phase. These findings suggest that high Cdc25 and Cdk1 contribute to the speed of the rapid, pre-MBT S phases and that down-regulation of these activities plays a broader role in MBT-associated changes than was previously suspected (Farrell, 2012 full text of article).

The experiments overrode the normal down-regulation of Cdc25 and Cdk1 activity at the MBT and showed that this developmental down-regulation is required for the introduction of late replication at S-phase 14 (see A model of how declining Cdc25 and Cdk1 activity results in S-phase 14 prolongation.). Increased Cdc25 or Cdk1 activity during cycle 14 abbreviates the late replication program normally active at the time. Conversely, it was found that decreasing Cdk1 activity during cycle 13 lengthens the rapid replication program that is active before the MBT. This suggests that Cdc25 and Cdk1 activity are regulating the length of S phase. Moreover, the pre-MBT S phases are unusual in having a rapid replication program and also in having high Cdc25 and Cdk1 activity during S phase. Then, cycle 14, when late replication begins in earnest, is the first cycle in which Cdc25 is effectively down-regulated and Cdk1 is inhibited by phosphorylation. Given these findings, it is proposed that this high Cdc25 and Cdk1 activity is actually the reason the pre-MBT S phases are rapid and the removal of these activities by down-regulation at the MBT is the developmental switch that lengthens S phase (Farrell, 2012).

Larval

The mechanism by which JNK signaling triggers cell death in response to TNF is poorly understood in mammals and is unknown in Drosophila. It was therefore of interest to identify the apoptotic machinery responsible for Eiger-induced cell death. Having excluded the caspase-8-like FADD/DREDD branch, focus was placed on the involvement of caspase-9, which represents another major pathway that leads to apoptosis. The key event for caspase-9 activation is its association with the protein cofactor Apaf-1 to form an active complex referred to as the apoptosome. Since many cell intrinsic insults can trigger this pathway, it has been termed the 'intrinsic death pathway'. Expression of a dominant-negative form of the Drosophila caspase-9 homolog DRONC, comprising only the CARD domain, fully blocks Eiger-induced apoptosis in a dose-dependent manner. Moreover, genetic removal of DARK, the homolog of Apaf-1, suppresses Eiger-dependent phenotypes. These results indicate that the presumptive Drosophila apoptosome is essential for the ability of Eiger to induce cell death. In agreement with this conclusion, overexpression of Thread, the Drosophila inhibitor of apoptosis protein 1 (DIAP1) blocks Eiger function. Thread/DIAP1 has been shown to bind DRONC and target it for degradation. Most instances of programmed cell death that have been analyzed in Drosophila are triggered by, and require, the genes reaper, hid, or grim, which encode small proteins that bind to and inactivate IAPs, such as Thread/DIAP1. The removal of one copy of a chromsosomal segment that includes the genes hid, grim, and reaper rescues eye ablation, and Eiger induces a strong transcriptional activation of hid and a weak activation of reaper. These results suggest, therefore, that Eiger/JNK signaling triggers DRONC by inactivating the IAPs via a transcriptional upregulation of hid (Moreno, 2002).

Ultraviolet (UV) light is absorbed by cellular proteins and DNA, promoting skin damage, aging and cancer. The UV response by cells of the Drosophila retina have been explored. The retina enters a period of heightened UV sensitivity in the young developing pupa, a stage closely associated with its period of normal developmental programmed cell death. Injury to irradiated cells include morphology changes and apoptotic cell death; these defects can be completely accounted for by DNA damage. Cell death, but not morphological changes, is blocked by the caspase inhibitor P35. Utilizing genetic and microarray data, evidence is provided for the central role of Hid expression and for Diap1 protein stability in controlling the UV response. In contrast, Reaper has no effect on UV sensitivity. Surprisingly, Dmp53 is required to protect cells from UV-mediated cell death, an effect attributed to its role in DNA repair. These in vivo results demonstrate that the cellular effects of DNA damage depend on the developmental status of the tissue (Jassim, 2003).

The major inhibitor of caspase activity in Drosophila is Diap1. Stability of Diap1 is the central point of cell death regulation in the developing retina and this is also true during UV irradiation in the retina. Genetic and microarray results further suggest that the retina requires Hid as a primary regulator of Diap1 stability during UV irradiation. Hid may represent the primary regulator of Diap1 during UV (versus ionizing) irradiation response by the fly. Alternatively, the retina utilizes Hid as its major RHG factor during its development, and this preference may simply extend to its response to UV; other tissues may exploit different Diap1 regulators that reflect their use during development (Jassim, 2003).

Together, these results identify two points of regulation during a retinal cell's response to UV irradiation. The early step involves pyrimidine dimers, and requires proper repair from factors that include XPG and p53. The second step involves activation of caspases and requires regulation of Diap1 stability; interommatidial cells utilize Hid at this step, and the remaining cells employ a different (unknown) regulator. One challenge will be to connect these two points of regulation. Multiple signaling pathways are suggested by the microarray data. These include EGFR/Ras1 signaling (a central regulator of Hid), JNK pathway signaling and TGFß pathway signaling. The role of these factors is not known, but understanding them may help to connect early and late events (Jassim, 2003).

Apoptotic cells can induce compensatory cell proliferation through the JNK and the Wingless signaling pathways: DIAP1 antagonists reaper and hid can activate the JNK pathway which in turn is required for inducing wg and cell proliferation

In many metazoans, damaged and potentially dangerous cells are rapidly eliminated by apoptosis. In Drosophila, this is often compensated for by extraproliferation of neighboring cells, which allows the organism to tolerate considerable cell death without compromising development and body size. Despite its importance, the mechanistic basis of such compensatory proliferation remains poorly understood. Apoptotic cells are shown to express the secretory factors Wingless and Decapentaplegic. When cells undergoing apoptosis were kept alive with the caspase inhibitor p35, excessive nonautonomous cell proliferation is observed. Significantly, Wg signaling is necessary and, at least in some cells, also sufficient for mitogenesis under these conditions. Finally, evidence is provided that the DIAP1 antagonists reaper and hid can activate the JNK pathway and that this pathway is required for inducing wg and cell proliferation. These findings support a model where apoptotic cells activate signaling cascades for compensatory proliferation (Ryoo, 2004).

To investigate how the inhibition of diap1 may lead to mitogen expression, attention was focused on Dronc and the Jun N-terminal Kinase (JNK) pathway. Dronc has been implicated in compensatory proliferation, and its activity can be inhibited by the expression of droncDN. In addition, the JNK signaling pathway was considered as a candidate, since its activity is known to correlate with many forms of stress-provoked apoptosis, including disruption of morphogens, cell competition, and rpr expression. In Drosophila, the JNK pathway can be effectively blocked by the expression of puckered (puc), which encodes a phosphatase that negatively regulates JNK (Ryoo, 2004).

To induce patches of undead cells, wing imaginal discs were generated with mosaic clones expressing hid and p35. 48 hr after induction, these imaginal discs contained hid-expressing clones that autonomously induced wg. Using this experimental setup, it was asked whether additional expression of either droncDN or puc would block wg induction in undead cells. When droncDN was coexpressed, a subset of the hid-expressing population was still able to induce wg. In contrast, when puc was coexpressed, wg induction by hid was almost completely blocked. These results provide evidence that the JNK pathway is required for wg induction under these conditions but fail to uncover a similar requirement for Dronc (Ryoo, 2004).

To independently investigate the role of puc and droncDN in compensatory proliferation, the size of wing discs harboring undead cells was measured and they were compared with those of the sibling controls. Under the experimental conditions, wing discs harboring hid- and p35-expressing clones were on average 53% larger than their sibling controls. Coexpression of puc within these undead clones significantly limited growth, resulting in only a small increase in wing disc size that was not statistically significant. In contrast, coexpression of droncDN did not limit growth. Wing size measurements also correlated with the degree of wg induction. The larger size of discs harboring hid- and p35-expressing cells is not due simply to extra cell survival: (1) these undead cells are derived from the normal lineage; (2) the size of wing discs expressing hid, p35, and puc serves as a control. In this case, although a large number of undead cells were generated, no significant increase in disc size was observed, in stark contrast to the discs expressing hid and p35 only. It is concluded that the JNK pathway is required for the nonautonomous growth promoting activity of the undead cells (Ryoo, 2004).

To confirm a role of puc in imaginal disc growth, rpr and p35 werecoexpressed in wild-type and puc−/+ imaginal discs. Like hid, rpr is a DIAP1 antagonist, but with a weaker cell killing activity when overexpressed in imaginal disc cells. In a puc+/+ background, a small amount of ectopic wg expression was observed, indicative of rpr's weaker DIAP1 inhibiting activity. In contrast, ectopic wg expression was strongly enhanced in puc−/+ discs. Because the puc allele used, pucE69, also acts as a lacZ reporter, JNK pathway induction could be monitored simultaneously. wg induction in undead cells correlates very well with puc-lacZ expression, with a stronger induction at the center of the wing pouch. These results further support the role of JNK in the induction of wg (Ryoo, 2004).

Next to be tested was whether the reduction of puc had an effect on apoptosis-induced cell proliferation. Whereas puc−/+ discs expressing only p35 had BrdU incorporation similar to wild-type discs, coexpression of rpr and p35 in puc−/+ led to a significant increase in BrdU incorporation. Also, the size of these discs were on average 41% larger than those coexpressing rpr and p35 in a puc+/+ background. Taken together, these results show that diap1 inhibition leads to JNK activation and that JNK activity promotes wg induction and cell proliferation (Ryoo, 2004).

To directly test if JNK signaling can activate wg and dpp expression, hepCA, a constitutively active form of hemipterous (hep), the Drosophila JNK kinase was conditionally expressed. Expression of hepCA causes induction of wg-lacZ within 22 hr and to a lesser extent also dpp-lacZ. These ß-gal-expressing cells shifted basally and were apoptotic as assayed by anti-active caspase-3 antibody labeling. Hid protein levels were also elevated in these cells. Significantly, since p35 was not use to block apoptosis in this experiment, this demonstrates that wg and dpp can be induced not only in undead cells, but also in 'real' apoptotic cells (Ryoo, 2004).

This study provides evidence that the central apoptotic regulators can control the activity of mitogenic pathways. In particular, inhibition of DIAP1, either via expression of Reaper and Hid or by mutational inactivation, leads to the induction of the putative mitogens wg and dpp. When apoptosis was initiated through DIAP1 inhibition but cells were kept alive by blocking caspases, the resulting 'undead cells' exhibited strong mitogenic activity and stimulated tissue overgrowth. Inhibiting wg signaling with a conditional TCFDN blocked cell proliferation in imaginal discs, indicating that wg has an essential mitogenic function. Finally, evidence was provided that the JNK pathway mediates mitogen expression and imaginal disc overgrowth in response to rpr and hid. Based on these results, it is proposed that apoptotic cells actively signal to induce compensatory proliferation. DIAP1 inhibits both caspases as well as dTRAF1. According to this model, when DIAP1 is inhibited in response to cellular injury, the JNK pathway is activated and wg/dpp are induced in apoptotic cells. Secretion of these factors stimulates growth of proliferation-competent neighboring cells and leads to compensatory proliferation (Ryoo, 2004).

This study provides clear genetic evidence that diap1 is involved in compensatory proliferation. Overall, similar results were obtained with hypomorphic diap1 alleles (diap122-8s, diap133-1s), a null allele (diap1th5), and inactivation of diap1 by expression of Reaper and Hid. However, whereas expression of p35 effectively blocked apoptosis of diap122-8s/22-8s cells and in response to Reaper/Hid, it only partially suppressed the death of diap1th5/th5 cells. Consequently, the generation of undead cells was less efficient with the diap1th5 mutation. Moreover, these results suggest that the JNK pathway transduces the signal to activate mitogen expression and cell proliferation. Since IAPs have been shown to ubiquitylate TRAFs in both mammals and Drosophila and since no evidence was found for Dronc in growth promotion, it is attractive to speculate that JNK is regulated through direct DIAP1/TRAF1 interaction (Ryoo, 2004).

An important unresolved question is why compensatory proliferation is seen only in response to cellular injury, but not during normal developmental apoptosis. In particular, inactivation of DIAP1 by Reaper, Hid, and Grim is restricted not only to injury-provoked apoptosis, but also underlies most developmental cell deaths. One possible explanation is that activation of the JNK pathway is key to mitogenic signaling of apoptotic cells. Consistent with this idea, the JNK pathway is activated in response to tissue stress and injury, but not during developmental apoptosis. Furthermore, this study shows that JNK signaling can induce the expression of wg/dpp and nonautonomous cell proliferation. Therefore, it is possible that robust JNK activation and compensatory proliferation require the combined input of stress and apoptotic signals (Ryoo, 2004).

Effects of Mutation and RNAi

Apoptotic cell death is a mechanism by which organisms eliminate superfluous or harmful cells. Expression of the cell death regulatory protein Reaper (Rpr) in the developing Drosophila eye results in a small eye owing to excess cell death. Mutations in thread (th) are dominant enhancers of Rpr-induced cell death; th encodes a protein homologous to baculovirus inhibitors of apoptosis (IAPs). Th is called Drosophila IAP1 (DIAP1). Overexpression of DIAP1 or a related protein, DIAP2, in the eye suppresses normally occurring cell death as well as death due to overexpression of rpr or head involution defective. IAP death-preventing activity localizes to the N-terminal baculovirus IAP repeats, a motif found in both viral and cellular proteins associated with death prevention (Hay, 1995).

Drosophila Reaper (Rrp), Head involution defective (Hid), and Grim induce caspase-dependent cell death and physically interact with the cell death inhibitor DIAP1. Hid blocks Diap1's ability to inhibit caspase activity and evidence is provided suggesting that Rpr and Grim can act similarly. Based on these results, it is proposed that Rpr, Hid, and Grim promote apoptosis by disrupting productive IAP-caspase interactions and that Diap1 is required to block apoptosis-inducing caspase activity. Supporting this hypothesis, it is shown that elimination of Diap1 function results in global early embryonic cell death and a large increase in Diap1-inhibitable caspase activity and that Diap1 is still required for cell survival when expression of rpr, hid, and grim is eliminated (Wang, 1999).

Induction of apoptosis in Drosophila requires the activity of three closely linked genes: reaper, hid and grim. The proteins encoded by reaper, hid and grim activate cell death by inhibiting the anti-apoptotic activity of the Drosophila IAP1 (Diap1, also known as Thread) protein. In a genetic modifier screen, both loss-of-function and gain-of-function alleles in the endogenous diap1 gene were obtained, and the mutant proteins were functionally and biochemically characterized. Gain-of-function mutations in diap1 strongly suppress reaper-, hid- and grim-induced apoptosis. Sequence analysis of these diap1 alleles reveals that they are caused by single amino acid changes in the baculovirus IAP repeat domains of Diap1, a domain implicated in binding Reaper, Hid and Grim. Significantly, the corresponding mutant Diap1 proteins display greatly reduced binding of Reaper, Hid and Grim, indicating that Reaper, Hid and Grim kill by forming a complex with Diap1. Collectively, these data provide strong support for the idea that Reaper, Hid and Grim kill by inhibiting DIAP1's ability to antagonize caspase function (Goyal, 2000).

It is thought that the previously proposed function of IAPs upstream of reaper, hid and grim is simply an artifact of unphysiologically high levels of protein expression in heterologous systems. When IAP expression constructs are introduced into cultured cells under the control of strong promoters and at high copy numbers, the levels of proteins expressed far exceed those of the endogenous cellular IAP proteins. Under these unphysiological conditions, cellular IAPs can display properties that do not reflect their normal mechanism of action. In particular, the current results demonstrate that mutant proteins that completely lack anti-apoptotic activity in vivo can still inhibit cell death in vitro as long as they can bind to Reaper, Hid and Grim. Conversely, gain-of-function diap1 alleles that display reduced binding to Reaper, Hid and Grim have strongly increased anti-apoptotic function in vivo, but show reduced protection in heterologous cell transfection assays. These results clearly reveal the limitations of overexpression studies in cultured cells for determining the normal mechanism of action of these proteins in the cell death pathway (Goyal, 2000).

DIAP1, a Drosophila member of the IAP family of caspase inhibitors, suppresses rpr-, hid-, and grim-dependent cell death in the fly. It was reasoned that if expression of DRONC was activating the same pathway, then the GMR-DRONC eye phenotype might be sensitive to the levels of DIAP1. To test this hypothesis the amount of DIAP1 in the eye was decreased by crossing a strong loss-of-function DIAP1 point mutant, thread 5 (th5), to GMR-DRONCM flies. th5 heterozygotes are phenotypically wild type. However, flies that are heterozygous for th5, and that express GMR-DRONCM, show an enhancement of the GMR-DRONC-dependent small eye phenotype. In contrast, small eyed GMR-DRONCS flies that overexpress DIAP1 because they carry a GMR-DIAP1 transgene show a strong suppression of the small eye and pigment loss phenotypes. These observations, are consistent with the idea that DRONC activity is negatively regulated by DIAP1. However, they do not exclude the possibility that DIAP1's effects on the DRONC overexpression phenotypes are due, at least in part, to DIAP1's ability to suppress the activity of caspases such as Ice, that are activated by DRONC (Hawkins, 2000).

Reaper, Hid, and Grim are three Drosophila cell death activators that each contain a conserved NH2 -terminal Reaper-Hid-Grim (RHG) motif. The importance of the RHG motifs in Reaper and Grim have been examined for their different abilities to activate cell death during development. Analysis of chimeric R/Grim and G/Reaper proteins indicates that the Reaper and Grim RHG motifs are functionally distinct and help to determine specific cell death activation properties. A truncated GrimC protein lacking the RHG motif retains an ability to induce cell death, and unlike Grim, R/Grim, or G/Reaper, its actions are not efficiently blocked by the cell death inhibitors Diap1, Diap2, p35, or a dominant/negative Dronc caspase. Finally, a second region of sequence similarity was identified in Reaper, Hid, and Grim, that may be important for shared RHG motif-independent activities (Wing, 2001).

While the Grim-Reaper proteins do not contain defined structural domains, they each share sequence similarity in the 14 amino acids at their NH2-termini. This RHG motif is most similar between Reaper and Grim (71.4% identity), and least similar between Hid and Grim (21.4% identity). The RHG motif plays a key role in interactions between Grim-Reaper proteins and members of the Inhibitor-of-Apoptosis-Protein (IAP) family, including Drosophila Diap1 and Diap2. Like other IAPs, Diap1 and Diap2 both contain related baculovirus IAP repeat (BIR) motifs, as well as a Really Interesting New Gene (RING) finger. Diap1 is an essential cell death regulator and diap1 mutants exhibit early embryonic lethality due to massive ectopic cell death. The functions of Diap2 in regulating cell death are less clear; however, it does share a number of functional properties with Diap1. Diap1 can directly bind caspases and repress their proteolytic activities. Significantly, caspase inhibition by Diap1 is antagonized by Hid, suggesting a double-repression model where the Grim-Reaper proteins promote cell death by binding to Diaps, suppressing their ability to inhibit caspases. Recent studies have indicated that the vertebrate Diablo/SMAC protein also promotes cell death activation by binding to IAPs and suppressing their death inhibitory activities. Thus, IAP suppression may be an evolutionarily conserved cell death regulatory mechanism. In this regard, while Grim-Reaper orthologs have not been identified, the expression of each protein can induce vertebrate cells to die, implying that they may suppress vertebrate IAPs (Wing, 2001).

Diap1, like Diap2, exhibits distinct abilities to repress cell death induced by Reaper, Hid, or Grim. In the CNS midline, Diap1 more effectively blocks Grim-induced cell death than cell death induced by Reaper and Hid. In contrast, when examined in the adult eye, Diap1 is most effective at blocking Reaper-induced cell death, moderately effective against Hid, and ineffective against Grim. Similar results were obtained using the thsl gain-of-function diap1 mutant allele, which represses Reaper-induced eye cell death more effectively than death induced by Hid or Grim. Importantly, these data indicate that Diap1 has distinct, tissue-specific effects on cell death induced by Grim-Reaper proteins, and that these effects differ from those of Diap2. The basis for these functional distinctions are not yet clear. One possibility is that the associations between each Diap and Grim-Reaper protein may differ in strength, or be influenced by specific ancillary factors. Differences have been noted between Diap1 and Diap2 in their ability to bind and repress the actions of certain caspases, and Reaper, Hid, and Grim can act through different downstream caspases. Taken together, these findings suggest potentially complex functional interactions between Grim-Reaper proteins, Diaps, and caspases. It is likely that distinct activities of individual Grim-Reaper and Diap proteins provide enhanced capabilities for regulating cell death processes in different developmental and physiological contexts (Wing, 2001).

Although a homolog for CED-4/apoptotic protease-activating factor (Apaf)-1 has been described in Drosophila, its exact function and the role of the mitochondrial pathway in its activation remain unclear. The technique of RNA interference has been used to dissect apoptotic signaling pathways in Drosophila cells. Inhibition of the Drosophila CED-4/Apaf-1-related killer (ARK) homolog results in pronounced inhibition of stress-induced apoptosis, whereas loss of ARK does not protect the cells from Reaper- or Grim-induced cell death. Reduction of DIAP1 induces rapid apoptosis in these cells, whereas the inhibition of DIAP2 expression does not but results in increased sensitivity to stress-induced apoptosis; apoptosis in both cases is prevented by inhibition of ARK expression. Cells in which cytochrome c expression is decreased undergo apoptosis induced by stress stimuli, Reaper or Grim. These results demonstrate the central role of ARK in stress-induced apoptosis, which appears to act independently of cytochrome c. Apoptosis induced by Reaper or Grim can proceed via a distinct pathway, independent of ARK (Zimmermann, 2002).

Drosophila IAP1 (DIAP1) inhibits cell death to facilitate normal embryonic development. Using RNA interference it has been shown that down-regulation of DIAP1 is sufficient to induce cell death in Drosophila S2 cells. Although this cell death process is accompanied by elevated caspase activity, this activation is not essential for cell death. DIAP1 depletion-induced cell death is strongly suppressed by a reduction in the Drosophila caspase DRONC or Dark. RNA interference studies in Drosophila embryos also have demonstrated that the action of Dark is epistatic to that of DIAP1 in this cell death pathway. The cell death caused by down-regulation of DIAP1 is accelerated by overexpression of DRONC and Dark, and a caspase-inactive mutant form of DRONC can functionally substitute the wild-type DRONC in accelerating cell death. These results suggest the existence of a novel mechanism for cell death signaling in Drosophila that is mediated by DRONC and Dark (Igaki, 2002).

The observation that the pan-caspase inhibitor zD-dcb can not suppress the DIAP1 depletion-induced cell death suggests that DRONC may be able to induce cell death independent of its caspase activity. The observation that the caspase-inactive form of DRONC can functionally substitute the wild-type DRONC in accelerating DIAP1 depletion-induced cell death also supports the idea that the cell death can be mediated through non-caspase mechanisms. DRONC might have a protease-independent cell-killing activity that is activated by Dark. It is possible that DRONC is required simply as a bridging or scaffolding protein to bring other proteins together to transmit the cell death signaling. Although the possibility cannot be excluded that zD-dcb can not completely inhibit the caspase activity of DRONC, it is apparent that the mode of cell death caused by the down-regulation of DIAP1 is distinct from Reaper-induced cell death. The effects were assessed of dsRNAs synthesized from reaper, hid, grim, drob-1, and buffy/dborg-2 cDNAs on the diap1 dsRNA-induced cell death; none of them suppresses the cell death. Further in vivo analysis should help elucidate the role of the caspase-independent cell death pathway regulated by DIAP1 (Igaki, 2002).

Bcl-2 family proteins are key regulators of apoptosis. Both pro-apoptotic and anti-apoptotic members of this family are found in mammalian cells, but only the pro-apoptotic protein Debcl has been characterized in Drosophila: Buffy, the second Drosophila Bcl-2-like protein, is a pro-survival protein. Ablation of Buffy by RNA interference leads to ectopic apoptosis, whereas overexpression of buffy results in the inhibition of developmental programmed cell death and gamma irradiation-induced apoptosis. Buffy interacts genetically and physically with Debcl to suppress Debcl-induced cell death. Genetic interactions suggest that Buffy acts downstream of Rpr, Grim and Hid, and upstream of the apical caspase Dronc. Furthermore, overexpression of buffy inhibits ectopic cell death in diap1 (th5) mutants. Taken together these data suggest that Buffy can act downstream of Rpr, Grim and Hid to block caspase-dependent cell death. Overexpression of Buffy in the embryo results in inhibition of the cell cycle, consistent with a G1/early-S phase arrest. These data suggest that Buffy is functionally similar to the mammalian pro-survival Bcl-2 family of proteins (Quinn, 2003).

Diverse domains of THREAD/DIAP1 are required to inhibit apoptosis induced by REAPER and HID in Drosophila

The proapoptotic genes reaper (rpr), grim, and head involution defective (hid) are required for virtually all embryonic apoptosis. The proteins encoded by these genes share a short region of homology at their amino termini. The Drosophila IAP homolog Thread/Diap1 (Th/Diap1) negatively regulates apoptosis during development. It has been proposed that Rpr, Grim, and Hid induce apoptosis by binding and inactivating TH/Diap1. The region of homology between the three proapoptotic proteins has been proposed to bind to the conserved BIR2 domain of TH/Diap1. An analysis of loss-of-function and gain-of-function alleles of th indicates that additional domains of Th/Diap1 are necessary to allow th to inhibit death induced by Rpr, Grim, and Hid. In addition, analysis of loss-of-function mutations demonstrates that th is necessary to block apoptosis very early in embryonic development. This may reflect a requirement to block maternally provided Rpr and Hid, or it may indicate another function of the Th/Diap1 protein (Lisi, 2000).

Several mechanisms of action have been suggested for the antiapoptotic properties of the IAP family of proteins. Among these are the binding of the Drosophila IAPs to the proapoptotic proteins Rpr, Grim, and Hid. This interaction has been demonstrated in overexpression systems, and has been proposed to involve the homologous amino-terminal 14 amino acid sequences of the apoptosis initiators with the second BIR domain of the IAPs. The data presented here suggest that this is an oversimplification. Another mechanism that has been proposed for IAP antiapoptotic activity is the direct binding and inhibition of caspases. Th/Diap1 binds to the Drosophila caspases drICE and DCP-1 and functions to inhibit their ability to induce apoptosis. Here again, this binding activity appears to rest within BIR2 (Lisi, 2000).

These physical interactions support a simple model of IAP action. In this model, IAPs act within viable cells to inhibit caspase function. The action of Rpr, Hid, and Grim interferes with the ability of IAPs to inhibit caspases, thus inducing apoptosis. On the basis of the model, the LOF mutations identified in this study would be predicted to interfere with the ability of the Th/Diap1 protein to inhibit caspase function. This is likely to be true for th109.07, which lacks most of the protein, as well as for th5 and th4, which affect conserved residues in BIR2. BIR2 is sufficient to inhibit apoptosis induced by the active form of the Drosophila caspase drICE. The th9 mutation in BIR1 suggests that this BIR is also important for the full function in caspase inhibition. Alternatively, this change in BIR1 might have long-range effects on BIR2 structure or on protein stability (Lisi, 2000).

It is interesting to note that th7, which acts as a very strong LOF mutation and seems to show some dominant-negative properties, has only the BIR1 attached to the spacer and ring domains. Thus, despite the extensive homologies between the two BIR domains of the protein, a single BIR is not sufficient for Th/Diap1 function, at least in the presence of an attached ring domain. BIR2 of Th/Diap1 and Op-IAP, as well as the single BIR of survivin, are able to inhibit apoptosis (Lisi, 2000).

Again, on the basis of the model above, the GOF mutations identified would be predicted to bind to caspases, but not to the inducers. The thSL mutation maps to a weakly conserved residue in BIR1 and does not result in increased th protein levels. This suggests that BIR1 is important for Rpr and Grim binding, but not for Hid binding, as Hid activity is unaffected in this mutation. Even in the context of overexpression in the eyes of transgenic flies, this mutant IAP retains some specificity for Rpr and Grim killing. This implies that the simple model of BIR2 binding to the conserved NH2-terminal sequences of Rpr, Grim, and Hid is not accurate, and that other residues in the protein are differentially important for Rpr and Grim, as opposed to Hid binding (Lisi, 2000).

The importance of regions outside of BIR2 for Diap1 activity is supported by the analysis of the GOF1 class of mutations, th6B and th81.03. Both of these mutations suppress Hid killing and would be predicted to inhibit Hid binding. These mutations change conserved cysteines in the ring domain to tyrosines. This suggests that the ring is important for Hid/Diap1 interaction. However, the region of Hid binding to Diap1 and Op-IAP has been mapped to BIR2, while the ring does not show any ability to bind to Hid. In addition, mutations in the ring, including those in conserved cysteines, have little effect on the ability of Op-IAP to protect against Hid killing. These data, together with the finding that both GOF1 mutations are cysteine-to-tyrosine changes, suggest that these mutations might have a novel ability to interfere with binding of Hid to BIR2. In addition, the observation that the GOF1 mutations slightly enhance Rpr and Grim killing suggests that these mutants are less potent inhibitors of caspases. This might result from weaker binding to caspases or from proteins that are slightly less stable. This second attribute would be predicted to enhance killing by any inducer that binds the IAP, but not to have an effect on Hid, which is unable to bind (Lisi, 2000 and references therein).

In conclusion, the data support a model where Rpr, Grim, and Hid interact with Th/Diap1 to induce apoptosis. Mutations that affect killing by Rpr and Grim or by Hid can be isolated, indicating that these inducers interact with Th/Diap1 in different ways. The GOF mutations that have been identified also provide useful tools to examine the roles of IAPs, rpr, grim, and hid during Drosophila development. The other Drosophila IAP homolog, DIAP2, has been shown to selectively inhibit Rpr- and Hid-induced but not Grim-induced death (Lisi, 2000).

In LOF th alleles, a developmental arrest occurs at the blastoderm stage and, subsequently, a synchronous apoptosis of all the nuclei. Earlier reports that homozygous th embryos show no ectopic apoptosis probably reflects the very early stage at which this apoptosis occurs. At this time, a direct requirement for th to block apoptosis cannot be distinguised from a requirement for th in another developmental process. This developmental defect could then result in secondary apoptosis. The latter possibility is reasonable, as many failures in development result in ectopic apoptosis. A BIR containing protein from Caenorhabditis elegans is required for cytokinesis in embryos. However, it is also possible that developmental arrest occurs as a result of the initiation of apoptosis, which is manifest only as DNA damage several hours later (Lisi, 2000).

Does this early requirement for th reflect a need to inhibit apoptosis induced by rpr, grim, and hid? Double mutants of th and Df(3L)H99, the deletion that removes rpr, grim, and hid, show a phenotype similar to th alone. This indicates that Th/Diap1 is not required to suppress zygotic Rpr, Grim, and Hid activity. However, hid and rpr mRNA can be seen in a subset of cells in the blastoderm embryo, as judged by in situ analysis. This may indicate that these gene products are supplied maternally. Th/Diap1 may be required to suppress maternally supplied Rpr, Grim, or Hid. Allelic differences in the stage at which apoptosis begins in the th mutants parallel the general ability of the alleles to inhibit apoptosis induced by Rpr, Hid, and Grim. The strong LOF alleles arrest at the blastoderm stage; the GOF1 alleles arrest much later, and the GOF2 allele is completely viable (Lisi, 2000).

Unrestrained caspase-dependent cell death caused by loss of Diap1 function requires the Drosophila Apaf-1 homolog, Dark

In mammals and Drosophila, apoptotic caspases are under positive control via the CED-4/Apaf-1/Dark adaptors and negative control via IAPs (inhibitor of apoptosis proteins). However, the in vivo genetic relationship between these opposing regulators is not known. In this study, it has been demonstrated that a dark mutation reverses catastrophic defects seen in Diap1 mutants and rescues cells specified for Diap1-regulated cell death in development and in response to genotoxic stress. dark function is required for hyperactivation of caspases that occurs in the absence of Diap1. Since the action of dark is epistatic to that of Diap1, these findings demonstrate that caspase-dependent cell death requires concurrent positive input through Apaf-1-like proteins together with disruption of IAP-caspase complexes (Rodriguez, 2002).

Mutations in dark cause a variety of developmental defects, including an enlarged larval CNS. In the absence of dark function, cells specified to die actually survive and also differentiate. The embryonic CNS midline glia is a well-studied cell lineage in the fly embryo and, though only this particular lineage was studied, it is likely that supernumerary cell types in other lineages also persist in dark mutants. Several important conclusions derive from these observations. (1) The persistence of extra cells excludes trivial explanations for reduced TUNEL labeling in dark mutants (e.g. cell death in the absence of TUNEL labeling or redirected cell fates that prevent the specification of PCD) and reinforces models that favor a fundamental role for dark in embryonic PCD. (2) As shown in C.elegans and in studies of Reaper interval Drosophila mutants, rescue from PCD can uncover a cryptic differentiation program. (3) Most important, the same midline glial cells that fail to die in dark mutants also require the induced action of reaper, grim and hid. There is a requirement for dark function in normal PCD when reaper, grim and/or hid are expressed at physiological levels. It is noteworthy that Diap1 and at least two of the cell death initiators (i.e. reaper and hid) are expressed in pools of progenitor cells, which give rise to the embryonic midline glia and also play an important role in specifying the death of these cells. Thus, although midline glia presumably receive death signals from reaper and hid leading to the disruption of Diap1-caspase interactions, these cells still fail to die and are even capable of differentiating normally in the absence of dark function. This inference supports conclusions from epistasis studies indicating that a disruption of caspase-Diap1 interactions alone is insufficient for apoptosis, and suggests that dark functions as a co-effector of cell death signaling along with reaper, grim and/or hid. It is worth noting that the mammalian ortholog of dark, Apaf-1, functions downstream from the point at which mitochondrial factors including cytochrome c are released from mitochondria. If the same holds true for the fly protein, it follows that embryonic midline glia (and perhaps other cell types) can survive and differentiate beyond this mitochondrial 'point of no return' (Rodriguez, 2002).

Additional evidence that PCD requires coordinated dark-dependent caspase activation in conjuction with the release of IAP-mediated caspase inhibition comes from tests of dark mutants under conditions of stress that elicit apoptotic responses by the embryonic cell death initiators. Like the Reaper interval mutants, dark mutants also exhibit profound failures in cell death in response to ionizing radiation. Interestingly, the dark mutation itself does not interfere with the radiation induction of reaper mRNA. Therefore the possibility that resistance to damage-induced apoptosis evident in irradiated dark mutants is caused by a disruption of upstream elements in the signaling pathway can be excluded. Instead, the results raise the possibility that dark, like reaper, may function as an important effector of Drosophila p53-mediated apoptosis. In addition, these findings reinforce an apoptotic requirement for dark (even when reaper is transcriptionally induced) and provide evidence for models where cell death signals do not converge on Diap1 alone. In future studies, it will be interesting to determine the range of damage signals that can engage dark activity, since this locus might also function in a broader range of damage signals beyond those provoked by radiation (Rodriguez, 2002).

In flies, Diap1 is thought to function as a rate-limiting brake on apoptotic cell death. If, however, Diap1 were the most proximal rate-limiting regulator of apoptosis, then the presence or absence of dark function should have no influence on Diap1-dependent effects. If, however, dark and Diap1 exert interdependent functions, then the opposite outcome is predicted. The results from these studies provide consistent and compelling evidence favoring the latter scenario since, wherever tested, the influence of Diap1 upon apoptotic signaling is heavily dependent upon intact dark function. In the darkCD4 mutant, for instance, the Diap15 mutation fails to enhance grim- and hid-induced cell killing in the eye. These experiments examined Diap1 in the heterozygous condition, but the same outcome is also true when Diap1 is tested in the homozygous state. Early in embryonic development, Diap1 homozygotes exhibit catastrophic phenotypes including morphological arrest soon after gastrulation, extensive TUNEL-positive nuclei and hyperactivation of caspases. Each of these defects is profoundly dependent upon dark function since each is dramatically reversed by the dark mutation. In darkCD4; Diap15 double mutant embryos, no evidence is found for widespread apoptosis and, in fact, the large majority of these double mutants proceed through stages of embryogenesis well beyond the point at which single Diap1 mutants arrest. Thus, loss of dark not only suppresses apoptotic deaths that would otherwise occur, but definitively reverses a profound morphogenetic arrest that ensues in the absence of Diap1. Consequently, homozygosity at dark not only prevents the onset of apoptotic markers (i.e. DNA fragmentation and/or caspase activation), but actually preserves cells that are otherwise fated to die when normal checks upon caspases are removed. Rescue from inappropriate apoptosis in this instance is consistent with what was observed in the midline glia but, rather than preserving cells fated for programmed death, rescue occurs even when signaling from IAP antagonists in the Reaper region is bypassed. A similar result, with similar implications, was obtained in the ovary, where ~90% of heteroallelic Diap16/8 mutants show abnormal degeneration of early staged egg chambers and associated sterility. Again, loss of dark function not only suppresses the pathological effect, but actually reverses these degenerative defects to the extent that half of the darkCD4; Diap16/8 double mutant females produce normal egg chambers and are fertile. Thus, in both the ovary and the embryo, loss of dark rescues functional cells from apoptosis caused by misregulated Diap1 (Rodriguez, 2002).

Reversal of Diap1-dependent defects by dark is also evident when caspase activity is directly assayed in early embryos. In contrast to the 'hyperactivated' caspase levels detected in Diap1 single mutant lysates, Diap15; darkCD4 lysates show levels of caspase activity that are suppressed >90% relative to Diap1 mutants. The unusually high caspase activity detected in Diap1 mutant embryos is thought to reflect the action of these proteolytic enzymes unimpeded by native inhibitors. However, studies here demonstrate that removing a negative regulator alone is not sufficient to achieve caspase hyperactivation, since dark is clearly required for this unrestrained activity. It is worth mentioning that the basal DEVDase activity in double mutants is still somewhat elevated compared with control embryos. This might indicate dark-independent caspase activity due to caspase auto-activation in the absence of Diap1 or, since it is possible that the darkCD4 allele used here is not a complete null, it could reflect hypomorphic dark function. Alternatively, since a role for Diap1 in cytokinesis has not been ruled out, it is possible that this residual caspase activity ensues because of secondary developmental defects leading to cell death. Nevertheless, these enzymatic data extend the analyses to the biochemical level in a manner fully consistent with phenotypic studies (Rodriguez, 2002).

Interestingly, since Apaf-1/Dark adaptor proteins act upon procaspases while IAPs preferentially inhibit processed caspases, these results further suggest that pre-existing levels of processed caspases in most cells are probably not high enough to achieve an apoptotic threshold. Instead, a positive cell death stimulus from dark is required for the unusually high levels of caspase activation seen in Diap1-/- embryos. It is also evident that Dark-dependent 'basal' levels of DEVDase are detected in these assays several hours before the onset of embryonic PCD, and therefore authentic effector caspase (e.g. DrICE, Dcp-1) activity occurs even in the absence of overt apoptotic signals. These data indicate that constitutive levels of active effector caspases, not derived from autoproteolysis but instead promoted by Apaf-1-like adaptor proteins, may exist in many and perhaps all viable cells (Rodriguez, 2002).

Taken together, a strict Diap-1-caspase 'liberation' model does not explain sufficiently the evidence described. The action of dark is epistatic to that of Diap1, demonstrating an order of gene action whereby Dark functions either downstream or parallel to Diap1. Therefore, simple derepression of caspases via an IAP inhibitory bridge does not account adequately for epistasis between Diap1 and dark. Put another way, these results support the notion that apoptotic cell death in vivo results from the simultaneous activation of caspases by dark and the derepression of caspases by reaper, grim and/or hid. Accordingly, the findings are inconsistent with models that presume that Diap1 is the sole effector of reaper, grim and hid and that cells are 'pre-loaded' with sufficient levels of IAP-inhibited processed caspases to achieve cell killing. Instead, a 'gas and brake' model is favored whereby positive input from Apaf-1/Dark adaptors, together with removal of IAP inhibition, drives caspase activation to levels that exceed a threshold necessary for apoptosis (Rodriguez, 2002).

Role of programmed cell death in patterning the Drosophila antennal arista

Programmed cell death is a critical process for the patterning and sculpting of organs during development. The Drosophila arista, a feather-like structure at the tip of the antenna, is composed of a central core and several lateral branches. A homozygous viable mutation in the thread gene, which encodes an inhibitor of apoptosis protein, produces a branchless arista. Mutations in the proapoptotic gene hid led to numerous extra branches, suggesting that the level of cell death determines the number of branches in the arista. Consistent with this idea, it was found that thread mutants show excessive cell death restricted to the antennal imaginal disc during the middle third instar larval stage. These findings point to a narrow window of development in which regulation of programmed cell death is essential to the proper formation of the arista (Cullen, 2004).

Analysis of the th1 mutant has revealed a decrease in cell number by pupal stages, suggesting that excessive apoptosis could have occurred earlier in development. Indeed, TUNEL analysis revealed that th1 mutants show a dramatic increase in apoptosis compared to wild-type at a specific developmental timepoint, the middle third larval instar. Interestingly, caspase activity was found to be more extensive than TUNEL labeling, suggesting that caspases are activated in many antennal cells, but only a fraction succumb to apoptosis. This may indicate that there are other protectors acting downstream of caspase activation when inhibition by Thread fails. Alternatively, because this antibody detects processed effector caspases, the th1 mutant may not be able to inhibit caspase processing but may be able to inhibit enough caspase activity to prevent apoptosis. The increase in caspase activity that was observed is limited both spatially and temporally, such that by the late third larval instar, th1 discs show wild-type levels of immunolabeling. The ectopic caspase activity is also limited to the antennal portion of the eye-antennal disc, suggesting that thread activity or caspase inhibition is regulated differently in the eye and the antenna (Cullen, 2004).

One of the best-characterized activators of apoptosis in Drosophila is head involution defective or hid. Hid is thought to promote apoptosis by binding to Th, displacing it from caspases and triggering its auto-ubiquitination. hid mutants have been shown to have excessive cell numbers in the embryonic CNS and the adult eye. Here, hid mutants have numerous ectopic lateral branches in the posterior antennal arista. Mitotic clones of Df(3L)H99 dp not appear to have more branches than hid mutants alone, suggesting that hid is the primary regulator of cell death in the arista, as it is in the eye. Consistent with this idea, reaper mutants show only a mild aristal phenotype. Attempts were made to alter the amount of cell death in the arista by expression of reaper, hid, grim, or dcp-1, but high levels of expression tended to result in lethality and lower levels of expression did not produce phenotypes (Cullen, 2004).

hid mutants or H99 mosaics did not show any ectopic laterals on the anterior side of the arista, indicating that the anterior laterals could be regulated by a distinct apoptotic activator, or may be formed through an apoptosis-independent mechanism. However, because th1 mutants lack anterior laterals, and dark; th1 double mutants show ectopic anterior laterals, it is likely that an apoptotic mechanism is indeed involved. Rescue experiments indicate that a higher level of thread expression is required for anterior lateral formation, suggesting that there may be a potent apoptotic activator that can overcome low levels of Th present in the cells that give rise to the anterior laterals. Alternatively, different cohorts of caspases may be activated in the anterior and posterior cells, and the caspases in the anterior cells could require a higher level of Th for inhibition (Cullen, 2004).

The results with hid and th mutants suggest that an inhibition of cell death is required for lateral formation. This could be a direct effect, with a particular dying cell influencing the fate of a neighboring cell. Alternatively, the role for cell death could be more indirect, simply affecting the total number of cells, which in turn could determine whether a lateral will form or not. Indeed, th1 mutants have reduced cell numbers in the pupal aristae compared to wild-type, consistent with the observation of considerable apoptosis in the mid-third instar larval stage. Further support for the cell number hypothesis comes from observations of non-autonomy in H99 mitotic clones in the arista. While it was possible to detect ectopic branches in the H99 mosaics, these branches were not always marked with yellow, suggesting that they arose from heterozygous (or homozygous) yellow+ cells. Thus, the H99 clones may increase the total cell number in the developing aristae, but the specific cells that give rise to branches could be either homozygous or heterozygous for H99. How cell number influences lateral formation is unclear. It could involve lateral inhibition or lateral specification, where signaling cells induce adjacent cells to produce branches, and branch-producing cells block that fate in their neighbors. There may be a minimal number of cells required for basic support and extension of the arista; th1 mutants may have only this minimal number of cells, with no extra cells available for branch production. Alternatively, the th1 cells could be unable to produce lateral extensions due to cellular damage from insufficient caspase inhibition (Cullen, 2004).

The ectopic laterals observed in hid mutant aristae are intermediate in length and thickness compared to the long and short wild-type laterals. In addition, the normal longer branches in hid mutants are often shorter and thinner than wild-type. Since the laterals are thought to be formed as actin-rich projections of single cells, it is unclear how perturbing apoptosis could influence the length of the lateral. One possibility is that an increased cell number could lead to crowding or an overall decrease in cell size. The cell size could then influence the amount of cellular material available for the lateral projection. Alternatively, the 'undead' cells that survive abnormally may have ill-defined cell fates or lack sufficient cytoskeletal proteins to generate long lateral branches. Several caspase targets are regulators of the actin cytoskeleton, so increases in caspase activity might perturb the cytoskeleton, even though the caspase activity is not high enough to cause apoptosis. Similarly, the split laterals seen in the dark; th1 mutants could arise from cellular abnormalities (Cullen, 2004).

th1 antennal imaginal discs show increased apoptosis at a specific developmental timepoint, suggesting that regulation of th is critical in these cells. This developmental stage is characterized by rapid cell divisions and the establishment of cell fates. Key regulators of distal antennal fates are the transcription factors Distalless (Dll) and Homothorax (Hth). Coexpression of Dll and hth is sufficient to induce aristal transformations in leg, wing, head, and genital disc derivatives, accompanied by misexpression of spalt, a gene normally expressed in antennal but not leg discs. spalt and several other genes have been identified as targets of Dll and/or hth, however, most of these genes appear to be expressed in the proximal antenna, largely excluded from the presumptive arista. One exception is spineless, which is expressed in the aristal primordia during larval stages. spineless mutants show antennal to leg transformations, suggesting that its normal function is to repress leg and promote antennal fates. It remains unclear how such patterning genes could produce cell fates that are specifically susceptible to loss of Th. Perhaps these genes could directly regulate th levels transcriptionally or post-transcriptionally, and the th1 mutant may have a mutation in a corresponding region (Cullen, 2004).

The molecular nature of the th1 mutation is currently unknown. The th1 mutation behaves like a loss-of-function allele, displaying the aristal phenotype in trans to a deficiency and being rescued by a duplication for the chromosomal region. The coding sequence of the th1 allele is reported to lack any obvious mutations, although the appropriate background strain is unknown. Further investigation will be required to determine if any observed amino acid changes affect the protein function. There are three reported transcripts of th initiating from distinct promoters, but the tissue-specificity of these transcripts has not been reported in detail. Perhaps, the th1 mutation could disrupt one of the transcript variants that is primarily expressed in the presumptive arista, lowering the Th protein levels below a certain level necessary for maintaining caspase inhibition. The spontaneous nature of the th1 mutation suggests that it could be caused by the insertion of a transposable element, which could potentially disrupt specific transcripts. Future molecular analysis of the th1 mutation will contribute to the understanding of the role of cell death in patterning the antennal arista (Cullen, 2004).

The decision to undergo apoptosis appears to be regulated at the level of caspase activation, which is controlled by the IAP proteins, particularly DIAP1

In Drosophila, the APAF-1 homolog ARK is required for the activation of the initiator caspase DRONC, which in turn cleaves the effector caspases DRICE and DCP-1. While the function of ARK is important in stress-induced apoptosis in Drosophila S2 cells, since its removal completely suppresses cell death, the decision to undergo apoptosis appears to be regulated at the level of caspase activation, which is controlled by the IAP proteins, particularly DIAP1. This study further dissects the apoptotic pathways induced in Drosophila S2 cells in response to stressors and in response to knock-down of DIAP1. The induction of apoptosis is dependent in each case on expression of ARK and DRONC and surviving cells continue to proliferate. A difference was noted in the effects of silencing the executioner caspases DCP-1 and DRICE; knock-down of either or both of these have dramatic effects to sustain cell survival following depletion of DIAP1, but have only minor effects following cellular stress. These results suggest that the executioner caspases are essential for death following DIAP1 knock-down, indicating that the initiator caspase DRONC may lack executioner functions. The apparent absence of mitochondrial outer membrane permeabilization (MOMP) in Drosophila apoptosis may permit the cell to thrive when caspase activation is disrupted (Kiessling, 2006).

This study further dissected the apoptotic pathways induced in Drosophila S2 cells in response to stressors and in response to knock-down of DIAP1. The induction of apoptosis is dependent on expression of the APAF-1 homolog ARK, and the initiator caspase, DRONC. Knock-down of either ARK or DRONC led not only to short term cell survival, as is also observed in mammalian cells lacking APAF-1 or caspase-9, but also to long term survival, seen as cellular accumulation as the cells continued to proliferate. This is in striking contrast to observations in mammalian cells lacking APAF-1 or caspase-9, where cells ultimately succumb to 'caspase-independent cell death' and do not proliferate. This difference is most easily explained by the difference in mitochondrial involvement: in mammals, MOMP is associated with the release of potentially toxic factors, such as AIF, endoG, Omi, and others, and with an eventual loss of mitochondrial function, any of which can contribute to death, even when downstream caspase activation is blocked or defective. The apparent absence of MOMP in Drosophila apoptosis may permit the cell to thrive when caspase activation is disrupted (Kiessling, 2006).

Studies in ARK mutants clearly demonstrate that ARK is required for cell death in vivo, since these mutants display developmental defects, including an enlarged nervous system, and resist death induced by transgenic expression of Grim. Furthermore, genetic studies revealed an epistatic relationship between ARK and DIAP1 by demonstrating that loss of ARK reverses catastrophic defects seen in DIAP1 mutants and rescues developing tissues that would otherwise die from DAIP1 inactivation. The function of ARK is required for hyperactivation of caspases which occurs in the absence of DIAP-1. One might argue that the current findings are therefore merely confirmatory. However, it should be noted that profound developmental defects are observed in mice lacking APAF-1, caspase-9, or caspase-3, which are nevertheless dispensable for stress or oncogene-induced cell death in MEFs and lymphocytes from these mice in vitro and cells of the interdigital web in vitro or in vivo. In fact, there is currently no evidence that a cell capable of proliferation can do so following MOMP, and alternative explanations of developmental defects in these knockout mice (other than survival and proliferation following MOMP) have been offered (Kiessling, 2006).

A rapid loss of the DIAP1 is observed in the S2 cells, when treated with various stressors. The full length DIAP1 protein disappears rapidly and a smaller, 27 kDa fragment accumulates over time. Interestingly, the broad spectrum caspase inhibitor zVAD-fmk does not suppress the degradation of DIAP1, but the 27 kDa cleavage product could not be detected when caspase activation was inhibited. The differences between DIAP1 degradation with or without caspase activity could be explained by the notion that the degradation of DIAP1 after treatment with apoptosis-inducing stimuli is mediated by a combination of cleavage by caspases and proteasomal degradation. Thus, the continued degradation of DIAP1 in the presence of activated caspases produces the 27 kDa fragment. It has been recently reported that caspase-dependent cleavage of DIAP1 is required for DIAP1 loss in an early stage of apoptosis and that cleavage of DIAP1 is required for degradation. Similarly, it was observed that if caspases are inhibited following apoptosis induction, DIAP1 levels remain unaltered for a number of hours, however, the inhibition of caspases does not block DIAP1 degradation at longer times (Kiessling, 2006).

While a requirement for ARK and DRONC was observed under all of the pro-apoptotic conditions in S2 cells, a difference was noted in the effects of knock-down of the executioner caspases DCP-1 and DRICE. Knock-down of either or both of these has dramatic effects to sustain cell survival following knock-down of DIAP1, but has only minor effects following cellular stress (Kiessling, 2006).

Several possible explanations were envisioned for this difference. One possibility is that knock-down of DIAP1 leads to caspase activation uniquely through permitting DRONC function to activate DCP-1 and DRICE, while stressors somehow engage other caspase activation pathways (and other caspases). This, however, is inconsistent with the observation that stress-induced apoptosis is clearly dependent on ARK and DRONC. Alternatively, it may be that stress-induced death also involves inhibition of other IAPs, such as DIAP-2 and dBRUCE, which may have a wider spectrum of effects to engage additional caspases not affected by DIAP1 alone. Previously, however, it was noted that knock-down of DIAP-2 does not trigger apoptosis, but greatly enhances susceptibility to death induced by stressors such as were used in this study. This argues that DIAP-2 function, at least, continues following such stress (such that its knock-down has an effect) and thus it is less likely to be an important explanation for the current effects (Kiessling, 2006).

The final possibility is perhaps the most interesting. The knock-down of DIAP1 leads to death, presumably through permitting low levels of ongoing (and otherwise repressed) caspase activation to function and any subsequent effect may depend on amplification, as active executioner caspases cleave and activate others. Therefore, knock-down of even one caspase in the cell may dampen this amplification so that cells survive. In contrast, the induction of apoptosis by stress may involve not only blockade of DIAP1 function (through the N-termini of Reaper, Hid, Grim and Sickle) but also another signal that amplifies caspase activation upstream of ARK. Such an upstream effect has been suggested by studies of the so-called 'GH3' region in these proteins, that appears to be required for death in Drosophila cells and can function to promote the mitochondrial pathway in vertebrate systems. Reaper, Hid, Grim, and probably Sickle are necessary for stress-induced apoptosis in Drosophila, and therefore their effects are likely to depend on ARK and ARK-DRONC interactions. Nevertheless, this line of reasoning suggests that they function not only to de-repress caspases (though blocking DIAP1), but also to do something else to bypass full dependence on DCP-1 and DRICE, perhaps by amplifying caspase activation at the level of the ARK-DRONC interaction. While speculative, this possibility is intriguing, and suggests that the induction of apoptosis in Drosophila may prove to be more complex than simple models indicate (Kiessling, 2006).

A genome-wide RNAi screen reveals multiple regulators of caspase activation

A genome-wide RNA interference screen was performed to systematically identify regulators of apoptosis induced by DNA damage in Drosophila cells. Forty-seven double- stranded RNAs were identified that target a functionally diverse set of genes, including several with a known function in promoting cell death. Further characterization uncovers 10 genes that influence caspase activation upon the removal of Drosophila inhibitor of apoptosis 1. This set includes the Drosophila initiator caspase Dronc and, surprisingly, several metabolic regulators, a candidate tumor suppressor, Charlatan, and an N-acetyltransferase, ARD1. Importantly, several of these genes show functional conservation in regulating apoptosis in mammalian cells. These data suggest a previously unappreciated fundamental connection between various cellular processes and caspase-dependent cell death (Yi, 2007).

The genes that are specifically involved in caspase-dependent cell death were classified. Substantial induction of caspase activity was observed 8 h after treatment with a topoisomerase II inhibitor, doxorubicin (dox), to induce dose-dependent cell death. Any RNAi suppressing this activity implicates the target gene in early regulation of caspase activation. In addition to dcp-1 RNAi, knockdown of dronc and jra (the Drosophila homolog of c-Jun) significantly suppressed caspase-3/7-like activity in the presence of dox, whereas the negative control, RNAi against calpain A, a calcium-dependent cysteine protease, did not affect this pathway (Yi, 2007).

This analysis was expanded to all of the genes identified in the initial RNAi screen and 20 dsRNAs were discovered that suppressed caspase activation induced by DNA damage. Interestingly, 12 of these genes were found to be epistatic to diap1 (Yi, 2007).

diap1 epistatic analysis was performed to further categorize the genes. DIAP1, the fly orthologue of the mammalian inhibitors of apoptosis proteins, is a direct inhibitor of caspases, and deficiency in DIAP1 leads to rapid caspase activation and apoptosis in vivo. Thus, apoptosis induced by the loss of DIAP1 presents an alternative apoptotic assay independent of DNA damage. Silencing of genes that regulate activation of the core apoptotic machinery may provide protection against apoptosis induced by both DNA damage and the loss of DIAP1. RNAi against dcp-1 partially suppressed cell death induced by the depletion of DIAP1 in Kc cells. Also, dronc RNAi potently protected cells against apoptosis induced by deficiency in DIAP1. Altogether, 32 of the genes confirmed from the primary screen provided significant protection against cell death induced by the silencing of DIAP1 (Yi, 2007).

Interestingly, 12 dsRNAs suppressed caspase-3/7-like activity after dox treatment and protected against cell death induced by diap1 RNAi, suggesting that these genes are required for apoptosis induced by multiple stimuli. To confirm that these genes are necessary for the full activation of caspases, it was determined whether these dsRNAs could suppress spontaneous caspase activity induced by diap1 RNAi. Maximal induction of caspase activity by diap1 RNAi was observed after 24 h, and this effect was completely suppressed by dsRNA against dcp-1. Importantly, ablating 10/12 dsRNAs resulted in the significant suppression of caspase activity compared with diap1 RNAi only (Yi, 2007).

In addition to dronc RNAi, dsRNAs targeting chn and dARD1 provided the strongest suppression of spontaneous caspase activity. Consistent with the observation that RNAi against chn protects against DNA damage-induced cell death, the mammalian orthologue neuron-restrictive silencer factor (NRSF)/RE1-silencing transcription factor (REST) was recently identified as a candidate tumor suppressor in epithelial cells (Westbrook, 2005). Previous work indicates that Chn and NRSF/REST function as a transcriptional repressor of neuronal-specific genes (Chong, 1995; Schoenherr, 1995; Tsuda, 2006), suggesting that cellular differentiation may render cells refractory to caspase activation and apoptosis. Also, several metabolic genes, CG31674, CG14740, and CG12170, were identified that may be involved in the general regulation of caspase activation. It has been demonstrated that NADPH produced by the pentose phosphate pathway regulates the activation of caspase-2 in nutrient-deprived Xenopus laevis oocytes. Together with these results, these observations provide further evidence for an intimate link between the regulation of metabolism and induction of apoptosis (Yi, 2007).

To further explore the significance of these findings, whether silencing the mammalian orthologues of the fly genes identified from the RNAi screen confers protection against dox-induced cell death was investigated in mammalian cells. A set of mammalian orthologues was selected that are believed to be nonredundant. The list includes the orthologues of dMiro, which functions as a Rho-like GTPase; dARD1, which functions as an N-acetyltransferase; CG12170, which functions as a fatty acid synthase; and Chn, which functions as a transcriptional repressor (RHOT1, hARD1, OXSM, and REST, respectively; FlyBase). In addition, Plk3, a mammalian orthologue of Polo, was tested since dsRNA targeting polo potently protected against dox treatment (Yi, 2007).

The ability of siRNAs targeting a gene of interest to protect against DNA damage was tested in HeLa cells. As a positive control, cells were transfected with siRNAs targeting Bax or Bak, two central regulators of mammalian cell death. Indeed, silencing of Bax or Bak resulted in significant protection against dox- induced cell death. It was observed that plk3 RNAi provided partial protection against dox treatment, which is consistent with previous studies implicating Plk3 in stress-induced apoptosis. Interestingly, the knockdown of hARD1 dramatically enhanced cell survival in the presence of dox to levels similar to that of Bak. This protective effect was also evident at the morphological level. In cells transfected with a nontargeting control siRNA, dox treatment resulted in typical apoptotic morphology, including cell rounding and membrane blebbing. In direct contrast, cells transfected with siRNAs against hARD1 maintained a normal and healthy morphology and continued to proliferate in the presence of dox (Yi, 2007).

To examine whether the protection provided by siRNAs targeting hARD1 and plk3 is associated with the suppression of caspase activation, caspase activity was measured in these cells treated with dox. RNAi against plk3 provided partial suppression of caspase activity, again supporting the observed protection phenotype. Interestingly, the depletion of REST resulted in some suppression of caspase activity in the presence of dox even though the protection against cell death was not statistically significant. Consistent with the viability assay, complete suppression of caspase-3/7 activity was observed in cells transfected with hARD1 siRNA. These results indicate that hARD1 is required for caspase-dependent cell death induced by DNA damage. Furthermore, all four siRNAs targeting hARD1 were individually capable of providing robust protection against cell death, strongly suggesting that these siRNAs target hARD1 specifically (Yi, 2007).

Because the silencing of hARD1 dramatically suppressed activation of the downstream caspases, whether activation of the upstream caspases in response to dox treatment is also perturbed was also examined. Remarkably, hARD1 RNAi inhibited the cleavage of caspase-2 and -9 in cells treated with dox, whereas caspase cleavage was readily detected in control cells. Thus, it is proposed that hARD1 regulates the signal transduction pathway apical to the apoptotic machinery in the DNA damage response itself or the activation of upstream caspases (Yi, 2007).

Consistent with the results of the caspase-3/7 assay, silencing of hARD1 completely inhibited the appearance of activated caspase-3 induced by dox. This assay was used for a hARD1 complementation experiment to demonstrate the proapoptotic role of hARD1 in response to DNA damage. A new siRNA pool was used, targeting the 5' untranslated region of hARD1 (5'si); this treatment inhibited caspase-3 cleavage induced by dox treatment. Furthermore, caspase-3 cleavage was observed in reconstituted hARD1 knockdown cells. Because six out of six siRNAs against hARD1 provided strong protection against DNA damage-induced apoptosis and complementation of hARD1-sensitized cells to caspase activation, it is concluded that the functional role of ARD1 for dox-induced apoptosis is evolutionally conserved from Drosophila to mammals (Yi, 2007).

In summary, this study used an unbiased RNAi screening platform in Drosophila cells to identify genes involved in promoting DNA damage-induced apoptosis. Forty-seven dsRNAs were isolated that suppress cell death induced by dox. These genes encode for known apoptotic regulators such as Dronc, the Drosophila orthologue of the known proapoptotic transcriptional factor c-Jun, and an ecdysone-regulated protein, Eip63F-1, thereby validating the primary screen. Furthermore, this study implicates a large class of metabolic genes that were previously not suspected to have a role in modulating caspase activation and apoptosis, such as genes involved in fatty acid biosynthesis (CG11798), amino acid/carbohydrate metabolism (CG31674), citrate metabolism (CG14740), complex carbohydrate metabolism (CG10725), and ribosome biosynthesis (CG6712). These results support the proposal that the cellular metabolic status regulates the threshold for activation of apoptosis and thus plays a critical role in the decision of a cell to live or die (Yi, 2007).

Of particular interest is the identification of ARD1. Evidence is presented that RNAi against ARD1 provides protection against cell death and leads to the suppression of caspase activation induced by DNA damage in fly cells and HeLa cells. Furthermore, deficiency in dARD1 renders fly cells resistant to the spontaneous caspase activity and cell death associated with loss of Diap1. Importantly, substantial evidence is provided that hARD1 is required for caspase activation in the presence of DNA damage in mammalian cells. Cleavage of initiator and executioner caspases are suppressed in hARD1 RNAi cells treated with dox, suggesting that hARD1 functions further upstream of caspase activation, and the complementation of hARD1 knockdown cells restores caspase-3 cleavage. These data indicate that ARD1 is necessary for DNA damage-induced apoptosis in flies and mammals (Yi, 2007).

ARD1 functions in a complex with N-acetyltransferase to catalyze the acetylation of the Nα-terminal residue of newly synthesized polypeptides and has been implicated in the regulation of heterochromatin, DNA repair, and the maintenance of genomic stability in yeast. These studies suggest that ARD1 may be involved in regulating an early step in response to DNA damage. It is anticipated that future studies will focus on determining whether ARD1 functions in similar processes in mammals. The diversity of genes identified in the screen illustrates the complex cellular integration of survival and death signals through multiple pathways (Yi, 2007).

Neuronal remodeling and apoptosis require VCP-dependent degradation of the apoptosis inhibitor DIAP1

The regulated degeneration of axons or dendrites (pruning) and neuronal apoptosis are widely used during development to determine the specificity of neuronal connections. Pruning and apoptosis often share similar mechanisms; for example, developmental dendrite pruning of Drosophila class IV dendritic arborization (da) neurons is induced by local caspase activation triggered by ubiquitin-mediated degradation of the caspase inhibitor DIAP1. This study examined the function of Valosin-containing protein (VCP), a ubiquitin-selective AAA chaperone involved in endoplasmic reticulum-associated degradation, autophagy and neurodegenerative disease, in Drosophila da neurons. Strong VCP inhibition is cell lethal, but milder inhibition interferes with dendrite pruning and developmental apoptosis. These defects are associated with impaired caspase activation and high DIAP1 levels. In cultured cells, VCP binds to DIAP1 in a ubiquitin- and BIR domain-dependent manner and facilitates its degradation. This results establish a new link between ubiquitin, dendrite pruning and the apoptosis machinery (Rumpf, 2011).

This study has analyzed the role of VCP during the development of peripheral da neurons in Drosophila. VCP inhibition was found to affect neural development and cell viability in a biphasic manner. Mild VCP inhibition causes defects in caspase activation and therefore affects apoptosis and pruning. This effect is caused by interference with the ubiquitin-dependent degradation of the caspase inhibitor DIAP1, as the data suggest that DIAP1 is a VCP substrate. By contrast, strong VCP inhibition causes severe morphological defects and cell death, probably owing to increased proteotoxic stress. A likely explanation for these seemingly contradicting phenotypes is that VCP inhibition activates pro-apoptotic signaling cascades, such as JNK signaling or the UPR. Upon strong VCP inhibition, these signals eventually override the anti-apoptotic effects through slowed DIAP1 degradation. Importantly, caspase activation can occur in the presence of DIAP1; for example, the Hid protein can displace caspases from DIAP1 and thereby activate them. Similar biphasic phenotypes have been reported for the ubiquitin-activating enzyme E1 (Uba1): hypomorphic E1 alleles support cell viability and inhibit apoptosis via stabilization of DIAP1, whereas strong loss-of-function alleles have effects on mitosis and cell viability. It is interesting to speculate whether the results could be relevant for the pathogenesis of VCP-related neurodegeneration. In fact, mild pruning defects were observed in class IV neurons upon expression of a VCP disease variant (VCP R152H). Although these defects were relatively subtle, similar defects in neuronal remodeling in humans could contribute to dementia. In addition, the results suggest that VCP mutation might induce dementia not only through stress-induced cell death but also through inhibition of other VCP-dependent neuronal processes (Rumpf, 2011).

Importantly, the work has identified VCP as a new regulator of DIAP1 degradation. An interesting question is how VCP might contribute to DIAP1 degradation. Biochemical analysis suggests that an intact BIR1 domain is a major determinant for VCP binding. Since the results suggest that VCP is not required to break up an interaction between DIAP1 and a binding partner, it is speculated that it might be required to unfold ubiquitylated DIAP1, and specifically the BIR1 domain, prior to proteasomal degradation, since it has been proposed for a GFP-based model substrate. A stable BIR1 domain, in turn, might be favorable because caspase cleavage in the DIAP1 N-terminus (after amino acid 20) exposes an N-end rule degradation signal (Rumpf, 2011).

In other VCP-dependent pathways, such as ERAD, VCP often requires adaptor proteins to perform its functions. RNAi lines directed against the VCP interactors Ufd1-like, Npl4 (CG4673), p47 and Ufd2 (CG9934) were tested for their effects on dendrite pruning but did no defects were observed with any of these lines, indicating that VCP might not need adaptors to act on DIAP1; alternatively, other, as yet unknown adaptors might be involved (Rumpf, 2011).

Taken together, this study has identified VCP as a new regulator of neuronal remodeling and developmental apoptosis, and DIAP1 was identified as the relevant substrate. These data therefore provide a new link between the ubiquitin system and apoptosis (Rumpf, 2011).

A steroid-controlled global switch in sensitivity to apoptosis during Drosophila development

Precise control over activation of the apoptotic machinery is critical for development, tissue homeostasis and disease. In Drosophila, the decision to trigger apoptosis-whether in response to developmental cues or to DNA damage-converges on transcription of inhibitor of apoptosis protein (IAP) antagonists Reaper, Hid and Grim. This study describes a parallel process that regulates the sensitivity to, rather than the execution of, apoptosis. This process establishes developmental windows that are permissive or restrictive for triggering apoptosis, where the status of cells determines their capacity to die. One switch was characterized in the sensitivity to apoptotic triggers, from restrictive to permissive, that occurs during third-instar larval (L3) development. Early L3 animals are highly resistant to induction of apoptosis by expression of IAP-antagonists, DNA-damaging agents and even knockdown of the IAP diap1. This resistance to apoptosis, however, is lost in wandering L3 animals after acquiring a heightened sensitivity to apoptotic triggers. This switch in sensitivity to death activators is mediated by a change in mechanisms available for activating endogenous caspases, from an apoptosome-independent to an apoptosome-dependent pathway. This switch in apoptotic pathways is regulated in a cell-autonomous manner by the steroid hormone ecdysone, through changes in expression of critical pro-, but not anti-, apoptotic genes. This steroid-controlled switch defines a novel, physiologically-regulated, mechanism for controlling sensitivity to apoptosis and provides new insights into the control of apoptosis during development (Kang, 2013).


REFERENCES

Alsina, B., et al. (1999). Disruption of selenoprotein biosynthesis affects cell proliferation in the imaginal discs and brain of Drosophila melanogaster. J. Cell Sci. 112: 2875-2884. 10444382

Arnt, C. R., Chiorean, M. V., Heldebrant, M. P., Gores, G. J. and Kaufmann, S. H. (2002). Synthetic Smac/DIABLO peptides enhance the effects of chemotherapeutic agents by binding XIAP and cIAP1 in situ. J. Biol. Chem. 277(46): 44236-43. 12218061

Barron, D.A. and Moberg, K. (2016). Inverse regulation of two classic Hippo pathway target genes in Drosophila by the dimerization hub protein Ctp. Sci Rep 6: 22726. PubMed ID: 26972460

Bergmann, A., Tugentman, M., Shilo, B. Z. and Steller, H. (2002). Regulation of cell number by MAPK-dependent control of apoptosis: a mechanism for trophic survival signaling. Dev. Cell. 2(2): 159-70. 11832242

Birnbaum, M. J., Clem, R. J. and Miller, L. K. (1994). An apoptosis-inhibiting gene from a nuclear polyhedrosis virus encoding a polypeptide with Cys/His sequence motifs. J. Virol. 68: 2521-2528. 8139034

Blitvich, B. J. et al. (2002). Developmental- and tissue-specific expression of an inhibitor of apoptosis protein 1 homolog from Aedes triseriatus mosquitoes. Insect Mol. Biol. 11(5):431-42. 12230542

Borensztejn, A., Boissoneau, E., Fernandez, G., Agnes, F. and Pret, A. M. (2013). JAK/STAT autocontrol of ligand-producing cell number through apoptosis. Development 140: 195-204. PubMed ID: 23222440

Bratton, S. B., et al. (2001). Recruitment, activation and retention of caspases-9 and -3 by Apaf-1 apoptosome and associated XIAP complexes. EMBO J. 20(5): 998-1009. 11230124

Chai, J., Yan, N., Huh, J. R., Wu, J. W., Li, W., Hay, B. A. and Shi, Y. (2003). Molecular mechanism of Reaper-Grim-Hid-mediated suppression of DIAP1-dependent Dronc ubiquitination. Nat Struct Biol 10: 892-898. PubMed ID: 14517550

Challa, M., et al. (2007). Drosophila Omi, a mitochondrial-localized IAP antagonist and proapoptotic serine protease. EMBO J. 26(13): 3144-56. PubMed Citation: 17557079

Chen, P., Ho, S. I., Shi, Z. and Abrams, J. M. (2004). Bifunctional killing activity encoded by conserved reaper proteins. Cell Death Differ. 11: 704-713. PubMed citation: 15002042

Chong, J. A., et al. (1995). REST: a mammalian silencer protein that restricts sodium channel gene expression to neurons. Cell 80: 949-957. PubMed Citation: 7697725

Christich, A., Kauppila, S., Chen, P., Sogame, N., Ho, S. I. and Abrams, J. M. (2002). The damage-responsive Drosophila gene sickle encodes a novel IAP binding protein similar to but distinct from reaper, grim, and hid. Curr. Biol. 12(2): 137-40. 11818065

Clavier, A., Baillet, A., Rincheval-Arnold, A., Coleno-Costes, A., Lasbleiz, C., Mignotte, B. and Guenal, I. (2014). The pro-apoptotic activity of Drosophila Rbf1 involves dE2F2-dependent downregulation of diap1 and buffy mRNA. Cell Death Dis 5: e1405. PubMed ID: 25188515

Colombani, J., Polesello, C., Josue, F. and Tapon, N. (2006). Dmp53 activates the Hippo pathway to promote cell death in response to DNA damage. Curr. Biol. 16(14): 1453-8. 16860746

Cullen, K. and McCall, K. (2004). Role of programmed cell death in patterning the Drosophila antennal arista. Dev. Biol. 275: 82-92. 15464574

dela Cruz, W. P., Friesen, P. D., Fisher, A. J.(2001). Crystal structure of baculovirus P35 reveals a novel conformational change in the reactive site loop after caspase cleavage. J. Biol. Chem. 276(35): 32933-9. 11402050

Ditzel, M., Wilson, R., Tenev, T., Zachariou, A., Paul, A., Deas, E. and Meier, P. (2003). Degradation of DIAP1 by the N-end rule pathway is essential for regulating apoptosis. Nat. Cell Biol. 5(5): 467-73. 12692559

Dorstyn, L., Read, S., Cakouros, D., Huh, J. R., Hay, B. A., and Kumar, S. (2002). The role of cytochrome c in caspase activation in Drosophila melanogaster cells. J. Cell Biol. 156: 1089-1098. 11901173

Doumanis, J., Quinn, L., Richardson, H. and Kumar, S. (2001). STRICA, a novel Drosophila melanogaster caspase with an unusual serine/threonine-rich prodomain, interacts with DIAP1 and DIAP2. Cell Death Differ. 8(4): 387-94. 11550090

Ekert, P. G., et al. (2001). DIABLO promotes apoptosis by removing MIHA/XIAP from processed caspase 9. J. Cell Biol. 152(3): 483-90. 11157976

Farrell, J. A., Shermoen, A. W., Yuan, K. and O'Farrell, P. H. (2012). Embryonic onset of late replication requires Cdc25 down-regulation. Genes Dev. 26(7): 714-25. PubMed Citation: 22431511

Fraire-Zamora, J. J., Tosi, S., Solon, J. and Casanova, J. (2021). Control of hormone-driven organ disassembly by ECM remodeling and Yorkie-dependent apoptosis. Curr Biol. PubMed ID: 34666006

Fraser, A. G., James, C., Evan, G. I,, Hengartner, M. O. (1999). Caenorhabditis elegans inhibitor of apoptosis protein (IAP) homolog BIR-1 plays a conserved role in cytokinesis. Curr. Biol. 9(6): 292-301. 10209096

Freel, C. D., et al. (2008). Mitochondrial localization of Reaper to promote inhibitors of apoptosis protein degradation conferred by GH3 domain-lipid interactions. J. Biol. Chem. 283(1): 367-79. PubMed citation: 17998202

Geisbrecht, E. R. and Montell, D. J. (2004). A role for Drosophila IAP1-mediated caspase inhibition in Rac-dependent cell migration. Cell 118(1): 111-25. 15242648

Goyal, L., McCall, K., Agapite, J., Hartwieg, E., and Steller, H. (2000). EMBO J. 19: 589-597 Induction of apoptosis by Drosophila reaper, hid and grim through inhibition of IAP function. 10675328

Graves, H. K., Woodfield, S. E., Yang, C. C., Halder, G. and Bergmann, A. (2012). Notch signaling activates Yorkie non-cell autonomously in Drosophila. PLoS One 7: e37615. PubMed ID: 22679484

Hanson, A. J., et al. (2012). XIAP monoubiquitylates Groucho/TLE to promote canonical Wnt signaling. Mol. Cell 45(5): 619-28. PubMed Citation: 22304967

Harlin, H., Birkey Reffey, S., Duckett, C. S., Lindsten, T., and Thompson, C. B. (2001). Characterization of XIAP-deficient mice. Mol. Cell. Biol. 21: 3604-3608. 11313486

Harvey, K. F., Pfleger, C. M. and Hariharan, A. K. (2003). The Drosophila Mst ortholog, hippo, restricts growth and cell proliferation and promotes apoptosis. Cell 114: 457-467. 12941274

Hawkins, C. J., Yoo, S. J., Peterson, E. P., Wang, S. L., Vernooy, S. Y., and Hay, B. A. (2000). The Drosophila caspase DRONC cleaves following glutamate or aspartate and is regulated by DIAP1, HID, and GRIM. J. Biol. Chem. 275: 27084-27093. 10825159

Hay, B. A., Wassarman, D. A. and Rubin, G. M. (1995). Drosophila homologs of baculovirus inhibitor of apoptosis proteins function to block cell death. Cell 83(7): 1253-62. 8548811

Hays, R., Wickline, L. and Cagan, R. (2002). Morgue mediates apoptosis in the Drosophila melanogaster retina by promoting degradation of DIAP1. Nat. Cell Biol. 4(6): 425-31. 12021768

Hegde, R., et al. (2001). Identification of Omi/HtrA2 as a mitochondrial apoptotic serine protease that disrupts inhibitor of apoptosis protein-caspase interaction. J. Biol. Chem. 277: 432-438. PubMed Citation: 11606597

Holcik, M., Gordon, B. W. and Korneluk, R. G. (2003). The internal ribosome entry site-mediated translation of antiapoptotic protein XIAP is modulated by the heterogeneous nuclear ribonucleoproteins C1 and C2. Mol. Cell. Biol. 2003 23(1):280-8. 12482981

Holley, C. L., Olson, M. R., Colon-Ramos, D. A. and Kornbluth, S. (2002). Reaper eliminates IAP proteins through stimulated IAP degradation and generalized translational inhibition. Nat. Cell Biol. 4: 439-444. 12021770

Hozak, R. R., Manji, G. A. and Friesen, P. D. (2000). The BIR motifs mediate dominant interference and oligomerization of inhibitor of apoptosis Op-IAP. Mol. Cell. Biol. 20(5): 1877-85. 10669762

Hu, S. and Yang, X. (2003). Cellular inhibitor of apoptosis 1 and 2 are ubiquitin ligases for the apoptosis inducer Smac/DIABLO. J. Biol. Chem. 278(12): 10055-60. 12525502

Huang, Q., Tang, X., Wang, G., Fan, Y., Ray, L., Bergmann, A., Belenkaya, T. Y., Ling, X., Yan, D., Lin, Y., Ye, X., Shi, W., Zhou, X., Lu, F., Qu, J. and Lin, X. (2014). Ubr3 E3 ligase regulates apoptosis by controlling the activity of DIAP1 in Drosophila. Cell Death Differ 21(12):1961-70. PubMed ID: 25146930

Huang, X., Shi, L., Cao, J., He, F., Li, R., Zhang, Y., Miao, S., Jin, L., Qu, J., Li, Z. and Lin, X. (2014). The sterile 20-like kinase tao controls tissue homeostasis by regulating the hippo pathway in Drosophila adult midgut. J Genet Genomics 41: 429-438. PubMed ID: 25160975

Hunter, A. M. et al. (2003). A novel ubiquitin fusion system bypasses the mitochondria and generates biologically active Smac/DIABLO. J. Biol. Chem. 278(9): 7494-9. 12511567

Igaki, T., Yamamoto-Goto, Y., Tokushige, N., Kanda, H. and Miura, M. (2002). Down-regulation of DIAP1 triggers a novel Drosophila cell death pathway mediated by Dark and DRONC. J. Biol. Chem. 277: 23103-23106. 12011068

Janiesch P. C., et al. (2007). The ubiquitin-selective chaperone CDC-48/p97 links myosin assembly to human myopathy. Nat. Cell Biol. 9: 379-390. PubMed Citation: 17369820

Jassim, O. W., Fink, J. L. and Cagan, R. L. (2003). Dmp53 protects the Drosophila retina during a developmentally regulated DNA damage response. EMBO J. 22: 5622-5632. 14532134

Jia, J., Zhang, W., Wang, B., Trinko, R., and Jiang, J.(2003). The Drosophila Ste20 family kinase dMST functions as a tumor suppressor by restricting cell proliferation and promoting apoptosis. Genes Dev. 17: 2514-2519. 14561774

Jin, S., M. Kalkum, M. Overholtzer, A. Stoffel, B. T. Chait, and A. J. Levine. (2003). CIAP1 and the serine protease HTRA2 are involved in a novel p53-dependent apoptosis pathway in mammals. Genes Dev. 17: 359-367. 12569127

Kang, Y. and Bashirullah, A. (2013). A steroid-controlled global switch in sensitivity to apoptosis during Drosophila development. Dev Biol 386(1): 34-41. PubMed ID: 24333635

Kawamura, K., et al. (2003). Survivin acts as an antiapoptotic factor during the development of mouse preimplantation embryos. Dev Biol. 256(2): 331-41. 12679106

Kiessling, S. and Green, D. R. (2006). Cell survival and proliferation in Drosophila S2 cells following apoptotic stress in the absence of the APAF-1 homolog, ARK, or downstream caspases. Apoptosis 11(4): 497-507. 16532275

Kominsky, D. J., Bickel, R. J. and Tyler, K. L. (2002). Reovirus-induced apoptosis requires mitochondrial release of Smac/DIABLO and involves reduction of cellular inhibitor of apoptosis protein levels. J. Virol. 76(22): 11414-24. 12388702

Kuo, C. T., Jan, L. Y. and Jan, Y. N. (2005). Dendrite-specific remodeling of Drosophila sensory neurons requires matrix metalloproteases, ubiquitin-proteasome, and ecdysone signaling, Proc. Natl. Acad. Sci. 102: 15230-15235. Medline abstract: 16210248

Kuo, C. T., Zhu, S., Younger, S., Jan, L. Y. and Jan, Y. N. (2006). Identification of E2/E3 ubiquitinating enzymes and caspase activity regulating Drosophila sensory neuron dendrite pruning. Neuron 51(3): 283-90. Medline abstract: 16880123

Kuranaga, E., Kanuka, H., Igaki, T., Sawamoto, K., Ichijo, H., Okano, H. and Miura, M. (2002). Reaper-mediated inhibition of DIAP1-induced DTRAF1 degradation results in activation of JNK in Drosophila. Nat Cell Biol. 4(9): 705-10. 12198495

Kuranaga, E., et al. (2006). Drosophila IKK-related kinase regulates nonapoptotic function of caspases via degradation of IAPs. Cell 126(3): 583-96. Medline abstract: 16887178

Lai, Z.-C., et al. (2005). Control of cell proliferation and apoptosis by Mob as tumor suppressor Mats. Cell 12: 675-685. 15766530

Lamb, J. A., et al. (2003). JunD mediates survival signaling by the JNK signal transduction pathway. Molec. Cell 11: 1479-1489. 12820962

Leulier, F., et al. (2006). The Drosophila IAP DIAP2 functions in innate immunity and is essential to resist gram-negative bacterial infection. Mol. Cell. Biol. 26(21): 7821-31. Medline abstract: 16894030

Li, X., Wang, J. and Shi, Y. (2011). Structural mechanisms of DIAP1 auto-inhibition and DIAP1-mediated inhibition of drICE. Nat. Commun. 2: 408. PubMed Citation: 21811237

Lisi, S., Mazzon, I., and White, K. (2000). Diverse domains of THREAD/DIAP1 are required to inhibit apoptosis induced by REAPER and HID in Drosophila. Genetics 154(2): 669-78. 10655220

Luque, L. E., Grape, K. P. and Junker, M. (2002). A highly conserved arginine is critical for the functional folding of inhibitor of apoptosis (IAP) BIR domains. Biochemistry 41(46): 13663-71. 12427028

Martins, L. M., et al. (2001). The serine protease Omi/HtrA2 regulates apoptosis by binding XIAP through a Reaper-like motif. J. Biol. Chem. 277: 439-444. PubMed Citation: 11602612

Mané-Padrós, D., et al. (2010). The hormonal pathway controlling cell death during metamorphosis in a hemimetabolous insect. Dev. Biol. 346(1): 150-60. PubMed Citation: 20638378

Martins, L. M., et al. (2002). The serine protease Omi/HtrA2 regulates apoptosis by binding XIAP through a reaper-like motif. J. Biol. Chem. 277(1): 439-44. 11602612

Martins, L. M., et al. (2003). Binding specificity and regulation of the serine protease and PDZ domains of HtrA2/Omi. J. Biol. Chem. 278(49): 49417-27. 14512424

Meier, P., Silke, J., Leevers, S. J., and Evan, G. I. (2000). The Drosophila caspase DRONC is regulated by DIAP1. EMBO J. 19: 598-611. 10675329

Moreno, E., Yan, M. and Basler, K. (2002). Evolution of TNF signaling mechanisms: JNK-dependent apoptosis triggered by Eiger, the Drosophila homolog of the TNF superfamily. Curr. Biol. 12: 1263-1268. 12176339

Morey, M., Corominas, M. and Serras, F. (2003). DIAP1 suppresses ROS-induced apoptosis caused by impairment of the selD/sps1 homolog in Drosophila. J. of Cell Sci. 116: 4597-4604. 14576353

Muro, I., Hay, B. A. and Clem, R. J. (2002). The Drosophila DIAP1 protein is required to prevent accumulation of a continuously generated, processed form of the apical caspase DRONC. J. Biol. Chem. 51: 49644-49650. 12397080

Olson, M. R., Holley, C. L., Yoo, S. J., Huh, J. R., Hay. B. A. and Kornbluth, S. (2003a). Reaper is regulated by IAP-mediated ubiquitination. J. Biol. Chem. 278(6): 4028-34. 12446669

Olson, M. R. et al. (2003b). A GH3-like domain in reaper is required for mitochondrial localization and induction of IAP degradation. J. Biol. Chem. 278: 44758-44768. PubMed citation: 12917412

Oshima, K., et al. (2006). IKK epsilon regulates F actin assembly and interacts with Drosophila IAP1 in cellular morphogenesis. Curr. Biol. 16(15): 1531-7. Medline abstract: 16887350

Pantalacci, S., Tapon, N. and Leopold, P. (2003). The Salvador partner Hippo promotes apoptosis and cell-cycle exit in Drosophila. Nat Cell Biol. 5(10): 921-7. 14502295

Quinn, L., Coombe, M., Mills, K., Daish, T., Colussi, P., Kumar, S. and Richardson, H. (2003). Buffy, a Drosophila Bcl-2 protein, has anti-apoptotic and cell cycle inhibitory functions. EMBO J. 22(14): 3568-79. 12853472

Recasens-Alvarez, C., Ferreira, A. and Milán, M. (2017). JAK/STAT controls organ size and fate specification by regulating morphogen production and signalling. Nat Commun 8: 13815. PubMed ID: 28045022

Reuveny, A., Elhanany, H. and Volk, T. (2009). Enhanced sensitivity of midline glial cells to apoptosis is achieved by HOW(L)-dependent repression of Diap1. Mech. Dev. 126: 30-41. PubMed Citation: 18984040

Rodriguez, A., Chen, P., Oliver, H. and Abrams, J. M. (2002). Unrestrained caspase-dependent cell death caused by loss of Diap1 function requires the Drosophila Apaf-1 homolog, Dark. EMBO J. 21: 2189-2197. 11980716

Rumpf, S., Lee, S. B., Jan, L. Y. and Jan, Y. N. (2011). Neuronal remodeling and apoptosis require VCP-dependent degradation of the apoptosis inhibitor DIAP1. Development 138(6): 1153-60. PubMed Citation: 21343367

Ryoo, H. D., Bergmann, A., Gonen, H., Ciechanover, A. and Steller, H. (2002). Regulation of Drosophila IAP1 degradation and apoptosis by reaper and ubcD1. Nat. Cell Biol. 4(6): 432-8. Medline abstract: 12021769

Ryoo, H. D., Gorenc, T. and Steller, H. (2004). Apoptotic cells can induce compensatory cell proliferation through the JNK and the Wingless signaling pathways. Dev. Cell 7: 491-501. 15469838

Salvesen, G. S., and Duckett, C. S. (2002). IAP proteins: blocking the road to death's door. Nat. Rev. Mol. Cell. Biol. 3: 401-410. 12042762

Santhanam, A., Peng, W. H., Yu, Y. T., Sang, T. K., Chen, G. C. and Meng, T. C. (2014). Ecdysone-induced receptor tyrosine phosphatase PTP52F regulates Drosophila midgut histolysis by enhancement of autophagy and apoptosis. Mol Cell Biol 34: 1594-1606. PubMed ID: 24550005

Schoenherr, C. J. and Anderson, D. J. (1995). The neuron-restrictive silencer factor (NRSF): a coordinate repressor of multiple neuron-specific genes. Science 267: 1360-1363. PubMed Citation: 7871435

Schreader, B. A., Wang, Y. and Nambu, J. R. (2003). Drosophila morgue and the intersection between protein ubiquitination and programmed cell death. Apoptosis. 8(2): 129-39. 12766473

Schmollinger, J. C. et al. (2003). Melanoma inhibitor of apoptosis protein (ML-IAP) is a target for immune-mediated tumor destruction. Proc. Natl. Acad. Sci. 100(6): 3398-403. 12626761

Schuberth C. and Buchberger A. (2008). UBX domain proteins: major regulators of the AAA ATPase Cdc48/p97. Cell. Mol. Life Sci. 65: 2360-2371. PubMed Citation: 18438607

Scott, F. L., et al. (2005). XIAP inhibits caspase-3 and -7 using two binding sites: evolutionarily conserved mechanism of IAPs. EMBO J. 24: 645-655. 15650747

Seo, T. W., Lee, J. S. and Yoo, S. J. (2014). Cellular Inhibitor of apoptosis protein 1 ubiquitinates endonuclease G but does not affect endonuclease G-mediated cell death. Biochem Biophys Res Commun 451(4): 644-9. PubMed ID: 25139236

Shiozaki, E. N., et al. (2003). Mechanism of XIAP-mediated inhibition of caspase-9. Mol. Cell 11(2): 519-27. 12620238

Sierra, J., et al. (2006). The APC tumor suppressor counteracts beta-catenin activation and H3K4 methylation at Wnt target genes. Genes Dev. 20: 586-600. PubMed Citation: 16510874

Song, Z., Yao, X. and Wu, M. (2003). Direct interaction between survivin and Smac/DIABLO is essential for the anti-apoptotic activity of survivin during taxol-induced apoptosis. J. Biol. Chem. 278(25): 23130-40. 12660240

Speliotes, E. K., Uren, A., Vaux, D., Horvitz, H. R. (2000). The survivin-like C. elegans BIR-1 protein acts with the Aurora-like kinase AIR-2 to affect chromosomes and the spindle midzone. Mol. Cell 6(2): 211-23. 10983970

Srinivasula, S. M., et al. (2001). A conserved XIAP-interaction motif in caspase-9 and Smac/DIABLO regulates caspase activity and apoptosis. Nature 410(6824): 112-6. 11242052

Srinivasula, S. M., et al. (2003). Inhibitor of apoptosis proteins are substrates for the mitochondrial serine protease Omi/HtrA2. J. Biol. Chem. 278(34): 31469-72. 12835328

Suzuki, Y., et al. (2001). A serine protease, HtrA2, is released from the mitochondria and interacts with XIAP, inducing cell death. Mol. Cell 8(3): 613-21. PubMed Citation: 11583623

Tamm, I., et al. (2003). Peptides targeting caspase inhibitors. J. Biol. Chem. 278(16): 14401-5. 12538646

Tenev, T., et al. (2002). Jafrac2 is an IAP antagonist that promotes cell death by liberating Dronc from DIAP1. EMBO J. 21: 5118-5129. 12356728

Tapon, N., et al. (2002). salvador promotes both cell cycle exit and apoptosis in Drosophila and is mutated in human cancer cell lines. Cell 110: 467-478. 12202036

Treier, M. Seufert, W. and Jentsch, S. (1992). Drosophila UbcD1 encodes a highly conserved ubiquitin-conjugating enzyme involved in selective protein degradation. EMBO J. 11: 367-372. Medline abstract: 1310935

Tsogtbaatar, O., Won, J. H., Kim, G. W., Han, J. H., Bae, Y. K. and Cho, K. O. (2019). An ADAMTS Sol narae is required for cell survival in Drosophila. Sci Rep 9(1): 1270. PubMed ID: 30718556

Tsuda, L., et al. (2006). An NRSF/REST-like repressor downstream of Ebi/SMRTER/Su(H) regulates eye development in Drosophila. EMBO J. 25(13): 3191-202. PubMed Citation: 16763555

Udan, R. S., Kango-Singh, M., Nolo, R., Tao, C. and Halder, G. (2003). Hippo promotes proliferation arrest and apoptosis in the Salvador/Warts pathway. Nat. Cell Biol. 5(10): 914-20. 14502294

Vassilev, A., Kaneko, K. J., Shu, H., Zhao, Y. and DePamphilis, M. L. (2001). TEAD/TEF transcription factors utilize the activation domain of YAP65, a Src/Yes-associated protein localized in the cytoplasm. Genes Dev. 15: 1229-1241. PubMed Citation: 11358867

Verhagen, A. M., et al. (2000). Identification of DIABLO, a mammalian protein that promotes apoptosis by binding to and antagonizing IAP proteins. Cell 102(1): 43-53. 10929712

Verhagen, A. M., et al. (2002). HtrA2 promotes cell death through its serine protease activity and its ability to antagonize inhibitor of apoptosis proteins. J. Biol. Chem. 277(1): 445-54. 11604410

Vernooy, S. Y., Chow, V., Su, J., Verbrugghe, K., Yang, J., Cole, S., Olson, M. R. and Hay, B. A.(2002). Drosophila Bruce can potently suppress Rpr- and Grim-dependent but not Hid-dependent cell death. Curr. Biol. 12(13): 1164-8. 1212162

Vishal, K., Bawa, S., Brooks, D., Bauman, K. and Geisbrecht, E. R. (2018). Thin is required for cell death in the Drosophila abdominal muscles by targeting DIAP1. Cell Death Dis 9(7): 740. PubMed ID: 29970915

Wang, S. L., Hawkins, C. J., Yoo, S. J., Muller, H.-A. J., and Hay, B. A. (1999). The Drosophila caspase inhibitor DIAP1 is essential for cell survival and is negatively regulated by HID. Cell 98: 453-463. 10481910

Westbrook, T. F., et al. (2005). A genetic screen for candidate tumor suppressors identifies REST. Cell 121: 837-848. PubMed Citation: 15960972

Williams, D.W. and Truman, J. W. (2005). Cellular mechanisms of dendrite pruning in Drosophila: insights from in vivo time-lapse of remodeling dendritic arborizing sensory neurons. Development 132: 3631-3642. Medline abstract: 16033801

Wilson, R., Goyal, L., Ditzel, M., Zachariou, A., Baker, D. A., Agapite, J., Steller, H., and Meier, P. (2002). The DIAP1 RING finger mediates ubiquitination of Dronc and is indispensable for regulating apoptosis. Nat. Cell Biol. 4: 445-450. 12021771

Wing, J. P., Schwartz, L. M. and Nambu, J. R. (2001). The RHG motifs of Drosophila Reaper and Grim are important for their distinct cell death-inducing abilities. Mech. Dev. 102: 193-203. 11287192

Wing, J. P., et al. (2002). Drosophila Morgue is an F box/ubiquitin conjugase domain protein important for grim-reaper mediated apoptosis. Nat. Cell Biol. 4(6): 451-456. 12021772

Wright, C. W. and Clem, R. J. (2002). Sequence requirements for Hid binding and apoptosis regulation in the baculovirus inhibitor of apoptosis Op-IAP. Hid binds Op-IAP in a manner similar to Smac binding of XIAP. J. Biol. Chem. 277(4): 2454-62. 11717313

Wu, J. W., et al. (2001). Structural analysis of a functional DIAP1 fragment bound to Grim and Hid peptides. Mol. Cell 8: 95-104. 11511363

Wu, S., Huang, J., Dong, J., and Pan, D. (2003). hippo encodes a Ste-20 family protein kinase that restricts cell proliferation and promotes apoptosis in conjunction with salvador and warts. Cell 114: 445-456. 12941273

Wu, S., Liu, Y., Zheng, Y., Dong, J. and Pan, D. (2008). The TEAD/TEF family protein Scalloped mediates transcriptional output of the Hippo growth-regulatory pathway. Dev. Cell 14(3): 388-98. PubMed Citation: 18258486

Yang, C. S., Sinenko, S. A., Thomenius, M. J., Robeson, A. C., Freel, C. D., Horn, S. R. and Kornbluth, S. (2013). The deubiquitinating enzyme DUBAI stabilizes DIAP1 to suppress Drosophila apoptosis. Cell Death Differ 21(4): 604-11. PubMed ID: 24362437

Yang, Q. H., Church-Hajduk, R., Ren, J., Newton, M. L. and Du, C. (2003). Omi/HtrA2 catalytic cleavage of inhibitor of apoptosis (IAP) irreversibly inactivates IAPs and facilitates caspase activity in apoptosis. Genes Dev. 2003 17(12): 1487-96. 12815069

Yang, Y., Fang, S., Jensen, J. P., Weissman, A. M., and Ashwell, J. D. (2000). Science 288: 874-877. 10797013

Yi, C. H., et al. (2007). A genome-wide RNAi screen reveals multiple regulators of caspase activation. J. Cell Biol. 179: 619-626. PubMed Citation: 17998402

Yoo, S. J., Huh, J. R., Muro, I., Yu, H., Wang, L., Wang, S. L., Feldman, R. M. R., Clem, R. J., Müller, H.-A. J., and Hay, B. A. (2002). Hid, Rpr and Grim negatively regulate DIAP1 levels through distinct mechanisms. Nat. Cell Biol. 4: 416-424. 12021767

Yoo, S. J. (2005). Grim stimulates Diap1 poly-ubiquitination by binding to UbcD1. Mol. Cells 20(3): 446-51. 16404163

Yu, S. Y., Yoo, S. J., Yang, L., Zapata, C., Srinivasan, A., Hay, B.A. and Baker, N. E. (2002). A pathway of signals regulating effector and initiator caspases in the developing Drosophila eye. Development 129: 3269-3278. Medline abstract: 12070100

Vandergaast, R., Mitchell, J. K., Byers, N. M. and Friesen, P. D. (2015). Insect Inhibitor of Apoptosis IAPs are negatively regulated by signal-induced N-terminal degrons absent within viral IAPs. J Virol 89(8):4481-93. PubMed ID: 25653450

Zhang, L., et al. (2008). The TEAD/TEF family of transcription factor Scalloped mediates Hippo signaling in organ size control. Dev. Cell 14: 377-387. PubMed Citation: 18258485

Zimmermann, K. C., Ricci, J. E., Droin, N. M. and Green, D. R. (2002). The role of ARK in stress-induced apoptosis in Drosophila cells. J. Cell Biol. 156(6): 1077-87. 11901172

Zoog, S. J., et al. (2002). Baculovirus apoptotic suppressor P49 is a substrate inhibitor of initiator caspases resistant to P35 in vivo. EMBO J. 21(19): 5130-40. 12356729

Zou, H., et al. (2003). Regulation of the Apaf-1/caspase-9 apoptosome by caspase-3 and XIAP. J. Biol. Chem. 278(10): 8091-8. 12506111


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

date revised: 30 April 2021

Home page: The Interactive Fly © 2003 Thomas B. Brody, Ph.D.

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