Drosophila E2F1 Has Context-Specific Pro- and Antiapoptotic Properties during Development

Share Embed


Descrição do Produto

Developmental Cell, Vol. 9, 463–475, October, 2005, Copyright ©2005 by Elsevier Inc. DOI 10.1016/j.devcel.2005.08.015

Drosophila E2F1 Has Context-Specific Pro- and Antiapoptotic Properties during Development Nam-Sung Moon,1 Maxim V. Frolov,3 Eun-Jeong Kwon,1 Luisa Di Stefano,1 Dessislava K. Dimova,4 Erick J. Morris,1 Barbie Taylor-Harding,1 Kristin White,2 and Nicholas J. Dyson1,* 1 Massachusetts General Hospital Cancer Research Center 2 Massachusetts General Hospital Cutaneous Biology Research Center Charlestown, Massachusetts 02129 3 Department of Biochemistry and Molecular Genetics University of Illinois at Chicago Chicago, Illinois 60607 4 Department of Molecular Biology and Biochemistry Rutgers University Piscataway, New Jersey 08854

Summary E2F transcription factors are generally believed to be positive regulators of apoptosis. In this study, we show that dE2F1 and dDP are important for the normal pattern of DNA damage-induced apoptosis in Drosophila wing discs. Unexpectedly, the role that E2F plays varies depending on the position of the cells within the disc. In irradiated wild-type discs, intervein cells show a high level of DNA damage-induced apoptosis, while cells within the D/V boundary are protected. In irradiated discs lacking E2F regulation, intervein cells are largely protected, but apoptotic cells are found at the D/V boundary. The protective effect of E2F at the D/V boundary is due to a spatially restricted role in the repression of hid. These loss-offunction experiments demonstrate that E2F cannot be classified simply as a pro- or antiapoptotic factor. Instead, the overall role of E2F in the damage response varies greatly and depends on the cellular context. Introduction The E2F transcription factor is best known for its ability to control the G1-to-S phase (G1/S) transition. E2F-regulated genes include cell cycle regulators, such as Cyclin E and Cdc25a, and genes that encode essential components of the DNA replication machinery, such as the MCMs and DNA polymerase-a. At these targets, E2F proteins couple gene expression to cell cycle position by recruiting repressor complexes in G0 or G1 phase of the cell cycle, and by activating transcription as cells progress from G1 into S phase. In agreement with the idea that E2F is an important regulator of cell proliferation, studies in several experimental systems have shown that the ectopic expression of E2F genes is sufficient to drive quiescent cells into S phase, while the inhibition of E2F-dependent transcription blocks *Correspondence: [email protected]

cell proliferation (reviewed in Dyson, 1998; Muller and Helin, 2000). Recent genomic studies have found that the function of the E2F transcriptional program extends further than the G1/S transition. Chromatin immunoprecipitation (ChIP) experiments show that E2F proteins bind to the promoters of genes with a diverse assortment of functions (Ren et al., 2002; Dimova et al., 2003; Cam et al., 2004). Expression profiling studies indicate that E2F proteins control the expression of a broad spectrum of genes (Ishida et al., 2001; Muller et al., 2001; Dimova et al., 2003). Hence, E2F-dependent transcription is likely to impact many aspects of cellular function. Currently the best-characterized activity of E2F, in addition to its traditional role in cell cycle progression, is its ability to induce apoptosis (reviewed in Phillips and Vousden, 2001; Evan and Vousden, 2001; Sears and Nevins, 2002; Bracken et al., 2004). Overexpressed E2F proteins activate the transcription of a large number of proapoptotic genes, including, for example, Caspase-3, -7, p19Arf, p73, and Apaf1 (Irwin et al., 2000, Stiewe and Putzer, 2000; Lissy et al., 2000; Moroni et al., 2001; Muller et al., 2001; Nahle et al., 2002; Aslanian et al., 2004; Hershko and Ginsberg, 2004). Since normal cells proliferate without suffering E2F-induced apoptosis, the proapoptotic potential of E2F is evidently held in check during development. How the switch between these activities is controlled is not well understood. Cell survival signals have been proposed to allow E2Fdriven cell proliferation by suppressing E2F-induced apoptosis (Hallstrom and Nevins, 2003; Chaussepied and Ginsberg, 2004). In addition, DNA damage signals have been suggested to specifically activate E2F1dependent transcription of proapoptotic genes. ATM/ ATR, CHK2, and P/CAF modify E2F1 after DNA damage (Lin et al., 2001; Pediconi et al., 2003; Stevens et al., 2003; Ianari et al., 2004). These changes stabilize E2F1, increase its ability to activate transcription, and allow it to preferentially bind to the promoters of some proapoptotic genes. While “activator” E2Fs (mammalian E2F1, E2F2, E2F3; Drosophila dE2F1) potently induce apoptosis when overexpressed, little is known about the normal roles of the endogenous E2F proteins in this process. DNA damage-induced apoptosis is reduced and delayed in thymocytes derived from E2f1−/− mice, suggesting that E2F1 has an important role in this cell type (Lin et al., 2001). Mutation of E2f1 or E2f3 reduces apoptosis in the central nervous system of Rb−/− mice (Tsai et al., 1998; Ziebold et al., 2001). Because E2F1 is a member of a large family of related proteins, it is possible that the collective activity of the E2F family has a general role in the regulation of apoptosis (Trimarchi and Lees, 2002). Unfortunately, such a role is difficult to investigate by using loss-of-function approaches because mice carrying mutations in multiple E2f genes have severe developmental defects. Currently, it is not feasible to genetically eliminate all E2F activity in mammals and to examine the consequences. As a result,

Developmental Cell 464

the full effect of E2F regulation on the way that cells respond to apoptotic signals is not known. Here, we have taken advantage of the relative simplicity of the Drosophila E2F family. E2F-mediated control combines the functions of both activator and repressor complexes. Flies contain just two E2F genes, de2f1 and de2f2. dE2F1, a potent transcriptional activator, appears to be functionally analogous to the mammalian “activator” E2Fs (E2F1, E2F2, and E2F3) and is controlled by RBF1 (Dynlacht et al., 1994; Ohtani and Nevins, 1994; Du and Dyson, 1999). dE2F2, by contrast, is a dedicated transcriptional repressor. dE2F2 acts with either RBF1 or RBF2 and may provide functions analogous to the mammalian “repressor” E2Fs, E2F4 and E2F5 (Frolov et al., 2001; Stevaux et al., 2002). Both dE2F1 and dE2F2 require dDP to bind to DNA. No Drosophila genes analogous to mammalian E2f6, E2f7, or E2f8 have been found. Mutations in de2f1 and rbf1 generate an imbalance in E2F regulation that disrupts the normal control of cell proliferation. Remarkably, and rather unexpectedly, the complete elimination of E2F regulation, by either the combined mutation of de2f1 and de2f2, or by mutation of dDP, has less severe consequences (Royzman et al., 1997; Frolov et al., 2001). de2f1;de2f2 and dDP mutants develop to pupal stages without major developmental defects. These mutants provide us with an opportunity to study cellular processes that require E2F in a setting in which patterns of cell proliferation appear to be relatively normal. There are several indications that the link between E2F and apoptosis is conserved in Drosophila. dE2F1, like mammalian E2F1, causes apoptosis when overexpressed and promotes the expression of proapoptotic genes such as reaper and hac-1 (Asano et al., 1996; Du et al., 1996; Zhou and Steller, 2003). Animals carrying hypomorphic alleles of de2f1 have defects in nurse cell dumping (Royzman et al., 2002), but since the complete loss-of-function allele of de2f1 causes an early block in development (Duronio et al., 1995), the full role of dE2F1 cannot be examined by using simple null mutants. Here, using dDP mutant animals as our starting point, and taking advantage of the fact that de2f2 mutants suppress the early larval lethality of de2f1 mutation, we show that E2F is an important determinant in the cellular response to DNA damage. Although E2F is generally described as a proapoptotic factor, we find that its role in vivo is more complex. In some cell types, E2F/DP proteins promote DNA damage-induced apoptosis. However, in other settings, dE2F1 is required to protect cells against DNA damage-induced apoptosis. We show that this protective function is mediated through the transcriptional regulation of hid.

Results Like many other cell types, cells in the imaginal wing discs of Drosophila third instar larvae respond to γ irradiation-induced DNA damage by imposing a cell cycle arrest. This arrest is evident in wing discs immunostained with the anti-phospho-histone H3 antibody (p-H3), an antibody that recognizes a histone modifica-

tion that is abundant in mitotic cells (Figure 1). Prior to irradiation, mitotic cells are present throughout the developing disc. Irradiated cells arrest in G2/M, in response to a DNA damage checkpoint; by 1 hr after irradiation, p-H3-positive cells are completely absent from the disc. To visualize the onset of DNA damage-induced apoptosis, irradiated discs were stained with an antibody specific for the activated form of mammalian Caspase 3 (C3). Cells with activated caspases were first evident 2 hr after irradiation, were maximal at 4 hr, and declined by 6 hr (Figure S1; see the Supplemental Data available with this article online). Interestingly, C3-stained cells were not randomly distributed in the wing disc, but arose in a pattern that was reproducible between discs and at different doses of radiation. C3 staining was highest in the wing pouch. Double staining with antibodies to proteins that are expressed in developmentally regulated patterns showed that C3-positive cells were excluded from the regions of Wingless (Wg) or Delta (Dl) expression at the D/V boundary. The D/V boundary is a characteristic zone of nonproliferating cells (ZNC) that separates the cells of the disc that will form the dorsal and ventral surfaces of the adult wing. To confirm that the C3 staining pattern reflects the distribution of dying cells, irradiated discs were stained with TUNEL or Acridine Orange. Similar results were obtained with each method (Figure 1 and data not shown). We infer that cells in the wing disc have a predetermined, differential sensitivity to DNA damage-induced apoptosis. To discover the role that E2F plays in this pattern, we examined dDP mutant discs. dDP is the only known heterodimeric partner for dE2F1 and dE2F2, and inactivation of dDP mimics the inactivation of both de2f genes. After irradiation, p-H3 staining decreased in dDP mutant wing discs in a manner that was identical to wild-type discs (Figure 1C). Hence, the lack of dDP does not impair the ability of the cells to sense or respond to DNA damage. However, the pattern of C3stained cells in dDP mutant discs differs significantly from wild-type. In dDP mutant discs, the number of C3positive cells was reduced, and these cells clustered at the center of the wing pouch in a region that overlapped the D/V boundary (Figure 1D). Similar results were obtained with TUNEL, or with Acridine Orange to mark dying cells (Figure 1D and data not shown). Time course experiments showed that apoptotic cells appeared with similar kinetics in wild-type and dDP mutant discs; the different locations of the apoptotic cells in wild-type and mutant discs persisted throughout the time course (Figure S1). Thus, the pattern of apoptosis seen in the dDP mutant discs is different from wildtype discs, and this change is not due to a delay, or acceleration, of a sequence of normal patterns. The relative sensitivity of cells to DNA damageinduced apoptosis is influenced by cell proliferation and by cell cycle position. However, in wild-type control wing discs, the patterns of cell death do not correspond to the pattern of BrdU incorporation. While BrdU is incorporated throughout much of the wing disc, the C3 staining pattern overlaps the regions of the wing disc containing presumptive intervein cells. Double staining with C3 and an antibody to dSRF (Blistered), which marks intervein cells, shows apoptotic cells lo-

dE2F1 Has Both Pro- and Antiapoptotic Functions 465

Figure 1. The Pattern of DNA Damage-Induced Cell Death Is Abnormal in dDP Mutant Wing Discs (A–D) w1118 or dDP mutant third-instar larvae were irradiated and kept at 25°C for the number of hours indicated. (A and C) Mitotic cells were visualized by staining with an anti-phospho-histone H3 antibody (p-H3). (B and D) Apoptotic cells were identified by staining with an antibody that recognizes the cleaved form of Caspase 3 (C3) or by TUNEL. Wing discs were costained with antibodies to Wg or Dl as indicated.

cated almost entirely within the dSRF-expressing regions of the wing pouch (Figure 2A). This suggests that the sensitivity to DNA damage-induced cell death is influenced by signals that govern the patterning of the wing. To test if the developmental signals involved in the cell type specification influence the sensitivity to DNA damage-induced cell death, we generated overexpression clones of Rasv12 and Notch intracellular domain (NIC) (Go et al., 1998). Activation of the Ras/Raf/ MAPK cascade is critical for the determination of wing vein cells, and Notch is subsequently activated in intervein cells adjacent to the provein region (De Celis, 2003). Rasv12-overexpressing clones were found to be resistant to DNA damage-induced cell death, whereas clones of NIC-expressing cells were highly sensitive to DNA damage-induced apoptosis (Figure 2C). The development of wing imaginal discs in dDP mutants is indistinguishable from wild-type controls. The ZNC that is seen at the D/V boundary of wild-type discs forms normally in dDP mutants (Figure 2B). The expression patterns of developmentally regulated proteins that are required for the normal specification of the wing, such as Wg, Dl, and dSRF, are identical in dDP mutant and wild-type discs (Figure S2). Unirradiated dDP mutant discs lack any significant level of cell

death, as judged by TUNEL staining, Acridine Orange staining, and C3 staining (Figure 1D and data not shown). Although dDP mutant wing discs appear to develop normally, despite lacking any dE2F/dDP heterodimers, they have very different patterns of DNA damage-induced apoptosis. From this, we infer that the differences between wild-type and mutant discs are not simply due to changes in cell cycle distribution or to changes in cell fate determination. Instead, dDP is a key determinant of the relative sensitivity of wing disc cells to DNA damage-induced apoptosis. The pattern of apoptosis in irradiated de2f1;de2f2 double mutant wing discs was identical to that in dDP wing discs (Figure 3A), confirming that the dDP mutant phenotype is indeed due to the lack of E2F activity. de2f2 null wing discs had a pattern of irradiationinduced apoptosis that was identical to wild-type (Figure 3A), suggesting that dE2F2 is not important for this phenotype. Null mutant alleles of de2f1 are lethal during early larval development, and somatic clones of de2f1 mutant are too small to be informative. Consequently, we examined the role of de2f1 by using viable hypomorphic alleles. de2f1i2 contains a premature stop codon that truncates the protein shortly after the dDP binding domain (Royzman et al., 1997). The de2f1i2-

Developmental Cell 466

Figure 2. Developmental Context Influences the Cellular Sensitivity to DNA Damage-Induced Apoptosis (A and B) DNA damage-induced apoptosis occurs in intervein cells in wild-type discs, but it occurs in the D/V boundary region of dDP mutant discs. Wing discs from w1118 and dDP larvae are shown. S phase cells were visualized with BrdU. w1118 and dDP third-instar larvae were irradiated, and the patterns of cell death, pre- and postirradiation, were visualized with C3. Anti-dSRF was used to mark intervein and provein regions of the wing discs. (C) Discs containing flip-out clones of Rasv12 or Notch intracellular domain (NIC) were irradiated, and the level of cell death was visualized by C3 4 hr after irradiation. The clones are marked by GFP expression. Note that C3 staining is absent from the Rasv12-expressing cells (in clones generated in regions of the wing pouch that are normally sensitive to apoptosis) but is elevated in NIC-expressing cells (in clones generated at the periphery of the disc that do not normally undergo apoptosis).

encoded protein is able to bind to DNA, but it fails to activate transcription or to interact with RBF1. Irradiation of de2f1i2 mutant wing discs gave a pattern of DNA damage-induced apoptosis that was similar to dDP and de2f1;de2f2 discs: cell death was reduced in the intervein region but increased at the D/V boundary, when compared to wild-type (Figure 3A). This suggests that

dE2F1 determines the sensitivity to DNA damageinduced cell death in wing discs. The idea that dE2F1, a positive regulator of cell proliferation, functions in nonproliferating cells at the D/V boundary cells is supported by the observation that dE2F1 is expressed at relatively high levels in this part of the disc (Figure 3B). Further evidence of a role for dE2F1 was provided by

dE2F1 Has Both Pro- and Antiapoptotic Functions 467

Figure 3. The Pattern of Cell Death Induced by DNA Damage Is Altered by Mutations of dDP and de2f1, but Not by Mutation of de2f2 (A) de2f1;de2f2, de2f1i2, and de2f2 mutant third-instar larvae were irradiated. The patterns of cell death before and after irradiation were observed in dissected wing discs stained with C3 and anti-dSRF antibodies. (B) dE2F1 is expressed in the D/V boundary region. The pattern of dE2F1 expression was determined by comparing the staining pattern obtained with an anti-dE2F1 antibody on w1118 and de2f1;de2f2 double mutant wing discs. Note that the strongest dE2F1 signal is at the D/V boundary.

a second hypomorphic allele of de2f1. de2f1su89 contains a single point mutation in the RBF1 binding domain that compromises dE2F1’s interaction with RBF1. The de2f1su89-encoded protein is able to activate transcription, but, because it evades RBF1, it has greater transcriptional activity than wild-type dE2F1 (Weng et al., 2003). In de2f1su89 wing discs, we observed an increased level of irradiation-induced cell death in the intervein region and the appearance of apoptotic cells at the D/V boundary (Figure 4A). This indicates that the activation of E2F1-dependent transcription sensitizes cells to apoptosis in both the intervein and provein regions. The result also implies that RBF1 normally suppresses DNA damage-induced apoptosis in both regions. We tested this by examining rbf1 mutant clones.

Following irradiation, rbf1 mutant cells at the D/V boundary undergo apoptosis, while wild-type cells do not, confirming that the loss of RBF1 renders cells more susceptible to DNA damage-induced apoptosis (Figure 4B). The altered patterns of apoptosis observed in the de2f1su89 and de2f1i2 wing discs provide strong evidence that dE2F1 is an important determinant in the cellular response to DNA damage. The patterns are consistent with the idea that dE2F1/dDP-mediated activation promotes apoptosis in the intervein region of wild-type discs, and that this activity is opposed by RBF1. In contrast, at the D/V boundary, dDP, dE2F1, and RBF1 are needed to protect cells from DNA damage-induced cell death.

Developmental Cell 468

Figure 4. DNA Damage-Induced Cell Death Is Elevated When RBF1 Fails to Repress dE2F1 The effects of dE2F1 deregulation were examined by using de2f1su89, an allele with a point mutation in the RBF binding domain of dE2F1, or by using rbf1 mutant clones. (A) Third-instar de2f1su89 larvae were irradiated, and patterns of apoptosis were visualized by using C3 and anti-dSRF antibodies as described in Figure 2. (B) Irradiated mosaic wing discs containing rbf1 mutant clones. Note the elevated C3 staining in the rbf1 mutant clones (marked by the absence of GFP).

The evidence that dE2F1 sensitizes presumptive intervein cells to apoptosis conforms to the well-established view that E2F is a proapoptotic factor. At this stage of development, presumptive intervein cells are actively proliferating (see the BrdU patterns in Figure 1) and are expected to have a high level of dE2F1-driven transcription. The loss of dE2F1-mediated transcrip-

tional activation in dDP mutant discs is evident in the dramatic decrease in the expression of rnr2, a typical cell cycle-dependent dE2F1-regulated gene (Figure 5A). The decreased level of cell death in the presumptive intervein region of irradiated dDP mutant wing discs is most likely due to the decreased expression of one or more of dE2F1’s proapoptotic targets. Studies

dE2F1 Has Both Pro- and Antiapoptotic Functions 469

Figure 5. hid Expression Is Deregulated at the D/V Boundary and Is Rate Limiting for DNA Damage-Induced Cell Death in dDP Mutant Wing Discs (A) In situ hybridization was performed on wing discs from w1118 and dDP third-instar larvae. Antisense sequences from rnr2, dcp-1, and hid were used as probes. Note that hid expression is elevated in a spatially restricted manner in unirradiated dDP mutant discs, whereas expression of rnr2 and dcp-1 is reduced throughout the disc. A zone of nonproliferating cells (ZNC) is indicated by a red bar. Two different dDP wing discs are shown for comparison. Similar hid expression patterns were seen in all of the wing discs examined (19 wild-type and 32 dDP). (B) Heterozygosity for hid suppresses DNA damage-induced cell death in dDP mutant wing discs. Larvae of the indicated genotypes were irradiated, and the apoptosis was visualized as described above. Df(3L)x14 removes hid, and Df(3L)x38 removes sickle and reaper; CD4 is a mutant allele of Apaf1.

of mammalian E2Fs suggest that there may be many such targets; indeed, we found that expression of dcp-1, a Drosophila caspase, is reduced in unirradiated dDP mutant wing discs compared to wild-type discs (Figure 5A). Currently, it is uncertain whether dE2F1 acts directly, or indirectly, at the promoters of proapoptic genes like dcp-1. In contrast, the increased sensitivity of cells at the D/V boundary to DNA damage-induced apoptosis in dDP and de2f1;de2f2 mutants is unexpected; this indicates that the overall effect of E2F regulation protects cells from apoptosis. We looked for proapoptotic genes that are repressed by dE2F1 and misexpressed in dDP mutant discs. During these experiments, we discovered a very curious change in hid expression in dDP mutant wing discs. HID is a key regulator of cell death in Drosophila and functions by inactivating DIAP1 (reviewed in Bergmann et al., 2003). In situ hybridization showed that hid RNA is present at low levels in wild-type discs; however, hid mRNA increased, specifically, at the D/V boundary in dDP discs (Figure 5A). The pattern of elevated hid expression resembled the distribution of DNA damage-induced cell death in irradiated dDP mutant

wing discs, but it occurred prior to irradiation, raising the possibility that the sensitivity to apoptosis might be due to a localized deregulation of hid expression. To test whether the level of hid expression is functionally relevant, the patterns of DNA damage-induced cell death were examined in wing discs from wild-type and dDP mutant larvae that were either wild-type or heterozygous for hid. Reducing the gene dosage of hid dramatically decreased cell death at the D/V boundary of irradiated dDP mutant discs, but it had only a very slight effect on apoptosis in wild-type discs (Figure 5B). Similar results were obtained by using an independent hid allele, hidP05014. To determine whether the genetic interactions were specific, we examined the effects of mutant alleles of other positive apoptotic regulators that have previously been linked to E2F (Asano et al., 1996; Zhou and Steller, 2003). The pattern of DNA damage-induced apoptosis in dDP mutant discs was unaffected by introducing a single copy of the darkCD4 allele, a mutation in the Drosophila homolog of Apaf1, or a single copy of Df(3L)X38, a deletion that removes the proapoptotic genes reaper and sickle (Figure 5B). Together these results suggest that a localized increase

Developmental Cell 470

Figure 6. dE2F1 Is Recruited to Sequences Upstream of hid and Regulates hid Expression in S2 Cells (A) hid RNA levels were measured by real-time PCR on RNA samples from S2 cells treated with control, dE2F1, or dE2F2 double-stranded RNA. Averages were calculated from three independent experiments, each with triplicate samples. The error bars indicate the standard deviation of the averages. (B) The diagram shows the 5# region of the hid locus and the position of two putative E2F binding sites (open box), the best candidates in the vicinity of the transcriptional start site (−2 kb to +1 kb). Note the large intergeneic region 5# of hid. (C) ChIP shows that dE2F1 binds to genomic sequences upstream of hid. Chromatin bound dE2F1 and dE2F2 were immunoprecipitated with specific antibodies from S2 cell extracts. Preimmune rabbit serum served as a negative control. The results of quantitative PCR reactions are presented as the relative ratio between the input genomic DNA and the DNA from ChIP. The error bars indicate the standard deviation of the averages of three independent PCRs, each with triplicate samples. Primers flanking either of the two putative E2F binding sites (indicated as arrows in [B]) in the hid locus, a negative control from the rp49 gene, and a previously identified dE2F2 target gene (arp53) were used. (D) Luciferase assays were performed with three reporter constructs: a wild-type PCNA promoter (PCNAwt-luc), a PCNA promoter construct in which the E2F binding sites were mutated (PCNAE2Fmut-luc), and a construct in which the PCNA dE2F1 binding site were replaced by the hid −1.4 kb dE2F1 binding site (PCNAE2Fhid-luc). These plasmids were cotransfected in S2 cells with increasing amounts of dE2F1 expression vector (50–100 ng).

dE2F1 Has Both Pro- and Antiapoptotic Functions 471

in the expression of hid sensitizes cells at the D/V boundary of dDP wing discs to DNA damage-induced cell death. The ability of dE2F1 to repress hid expression is not limited to the wing disc. RNAi-mediated depletion of dE2F1 from S2 tissue culture cells increased the level of hid RNA, but the removal of dE2F2 had no effect (Figure 6A). ChIP was used to find out whether hid might be a direct target of dE2F1. Examination of DNA sequences upstream of the transcription start site of hid revealed two motifs with homology to consensus E2F binding sites (−165 bp and −1.4 kb), and PCR primers were designed to amplify these sequences (Figure 6B). rp49 was used as a negative control, and a previously described dE2F2 target gene, arp53, served as a positive control for dE2F2. The relative levels of input and immunoprecipitated DNA were quantified by realtime quantitative PCR (Figure 6C). A minor enrichment of genomic DNA was observed in dE2F1 immunoprecipitates when primers flanking the potential E2F binding site at −165 bp were used, but a far greater enrichment was seen with primers flanking the putative E2F binding site 1.4 kb upstream of the hid transcriptional start site. Consistent with the finding that the level of hid expression, and the pattern of DNA damage-induced apoptosis are dependent on de2f1, but not on de2f2, no enrichment of hid was found in dE2F2 ChIPs. These results suggest that hid is directly, and specifically, targeted by dE2F1. We wished to know whether the element at −1.4 kb is a bona fide E2F binding site. Reporter constructs generated by fusing small genomic fragments from the hid locus to a reporter gene failed to reproduce the regulation of the endogenous gene, and this has precluded any detailed analysis of the hid promoter. Partly because of this, and because hid mutants cannot be rescued by small genomic fragments, the hid promoter seems likely to extend over a large region (Grether et al., 1995). As an alternative approach, we asked if the hid −1.4 kb motif could function as a dE2F1-responsive element in the context of a heterologous promoter. For this, we used the PCNA promoter, a well-studied target of dE2F (Sawado et al., 1998; Thacker et al., 2003). While the wild-type PCNA reporter construct (PCNAwtluc) responds to dE2F1 overexpression, a mutant reporter construct (PCNAE2Fmut-luc) does not. When the −1.4 kb hid sequence element was inserted in place of the canonical E2F binding sites, the reporter was strongly induced by dE2F1 (Figure 6D). To confirm that this element can mediate regulation by endogenous dE2F1/RBF proteins, reporters were assayed in cells depleted of dE2F1 or RBF1 by RNAi. The reporter containing the −1.4 kb hid sequence element decreased in the absence of dE2F1, and it increased in the absence of RBF1, just like the wild-type PCNA construct (Figure 6E). Thus, this element has the hallmarks of a functional dE2F1 binding site.

Why, then, isn’t hid expression strongly induced by dE2F1 in a cell cycle-dependent manner like other E2F targets? At −1.4 kb, the E2F binding site is more distal from the transcriptional start site than the sites described in most E2F-regulated promoters. To test the significance of this, we placed the hid −1.4 kb site a similar distance upstream of PCNAE2Fmut-luc. At this distance, the dE2F1 binding element failed to activate transcription from the reporter, even when dE2F1 was overexpressed (Figure 6F). Taken together these results show that hid expression is repressed by dE2F1, that dE2F1 binds to genomic sequences upstream of hid, and that, although the binding site is a functional element, it is likely to be too far removed from the transcriptional start site to provide strong activation.

Discussion The connection between DNA damage and E2F-dependent apoptosis has a special significance because E2F is deregulated in most tumor cells through lesions in the pRB pathway, and many of the commonly used treatments for cancer act by inducing DNA damage. The proapoptotic potential of E2F is well documented. Overexpression of E2F1 induces apoptosis, and a significant number of the proposed targets for E2Fs are genes with proapoptotic functions. Moreover, mammalian E2F1 is activated, specifically, in response to DNA damage. It is a less well-publicized fact that the lists of E2F-regulated genes discovered by microarray studies include many genes with antiapoptotic properties. For example, the overexpression of E2F1 was found to increase the expression of Bcl-2, TopBP1, and Grb2—genes that have been shown to suppress apoptosis in other studies (Muller et al., 2001; Chaussepied and Ginsberg, 2004; Liu et al., 2004). Here, we have taken advantage of the substantial development of dDP and de2f1;de2f2 mutant animals to examine the net contribution of E2F regulation to the DNA damage response. Because of the size of the mammalian E2F family, this type of genetic analysis would be very difficult to carry out in mammalian cells, and the results reveal how the complete elimination of E2F function influences the cellular response to DNA damage response in vivo. The results demonstrate that E2F/DP proteins are, indeed, critical determinants of the cellular response to DNA damage. Surprisingly, however, the role that E2F/DP plays is completely context dependent. In vivo, E2F/DP proteins vary from being strongly proapoptotic in some cells to being strongly antiapoptotic in others. One wonders whether the tissue culture systems that are often used to study E2F-induced apoptosis adequately reflect this complexity. What determines whether E2F makes cells sensitive or resistant to DNA damage-induced apoptosis? Our

(E) S2 cells were treated with double-stranded RNA to dE2F1, RBF1, or a control prior to the transfection of the indicated reporter plasmids. (F) Luciferase assays were performed with three reporter constructs: PCNAE2Fmut-luc, PCNAE2Fhid-luc, and a construct in which the hid −1.4 kb dE2F1 binding site was placed 1.3 kb upstream of the PCNAE2Fmut-luc start site, hidE2F-PCNAmut-luc. These reporters were cotransfected in S2 cells with 50–100 ng dE2F1 expression vector. Cells were harvested and assayed 48 hr after transfection. The error bars on the luciferase assays indicate the standard deviation of the average from duplicate experiments.

Developmental Cell 472

Figure 7. dE2F1 Proteins Determine Sensitivity to DNA Damage-Induced Cell Death in a Context-Dependent Manner Several studies have shown that activator E2Fs, like E2F1 and dE2F1, promote the transcription of proapoptotic genes. In Drosophila, dE2F1 activates the expression of the proapoptotic genes reaper and dcp-1, although, currently, it is unknown whether dE2F1 acts directly at these targets, or increases the transcription of these genes indirectly. In presumptive intervein cells, the ability of dE2F1 to activate expression of proapoptotic genes makes these cells sensitive to DNA damage-induced apoptosis. dE2F1 and dDP repress hid, most likely in association with RBF1, and limit the level of hid expression in the D/V boundary. Since dE2F1 binds to the hid promoter, hid is most likely a direct target of dE2F1 regulation, but we do not exclude the possibility that dE2F1 also has indirect effects on hid expression. In the absence of the dE2F1/dDP complex, a spatially restricted activator, “X,” activates hid transcription and promotes cell death. Because hid is a critical regulator of apoptosis, the net effect of E2F in the D/V boundary is to protect cells against apoptosis.

results suggest that a combination of factors are involved. Interestingly, the cells that are most sensitive to DNA damage-induced apoptosis in wild-type discs, and in which dE2F1/dDP is strongly proapoptotic, are actively proliferating. Conversely, the apoptotic cells seen in irradiated dDP and de2f1i2 mutant wing discs are centered on a region of nondividing cells, indicating that the antiapoptotic function of E2F occurs largely in cells that are under cell cycle arrest. However, these differential sensitivities are not simply an indirect effect of cell cycle position, because they are changed in dDP mutant discs, even though the distribution of cycling/ noncycling cells is the same as in wild-type discs. At first glance, these differences would seem to fit a model in which dE2F1/dDP complexes promote the expression of proapoptotic genes in proliferating cells, but combine with RBF1 to repress the same targets in arrested cells. Although this model is appealing, it cannot be the full explanation. Cells are actively proliferating throughout the wing disc, but dE2F1/dDP only sensitizes a spatially restricted subset of cells to DNA damage-induced apoptosis. Moreover, the ZNC does not completely correspond to the pattern of cells with activated caspases in dDP mutant discs. In addition to cell cycle-dependent fluctuations in E2F activity, clearly there must be additional, developmentally regulated signals that heighten or lessen the cellular sensitivity to dE2F1/dDP-induced apoptosis. Notch and Ras signaling pathways appear to be likely candidates for this. The cells with the highest sensitivity to DNA damageinduced apoptosis in the wild-type wing disc are situated in a region in which Notch signaling is high and Ras signaling is low. Clonal experiments show that Ras signaling protects cells against DNA damage-induced apoptosis and that Notch signaling promotes apoptosis. It has previously been shown that Ras signals can suppress HID activity by both transcriptional inhibition

and at the level of posttranslational modification (Bergmann et al., 1998; Kurada and White, 1998), but precisely how Notch signals affect E2F-dependent apoptosis is unclear. Future studies are needed to discover whether the Notch- and Ras-mediated signals alter the program of E2F transcription, or whether these pathways converge with E2F on the apoptotic machinery. One of the major difficulties in studying the biological functions of E2F is that E2F complexes affect the expression of a large number of genes and can act in a variety of different ways. It is difficult to assess the overall role of E2F regulation in a given process by studying an individual E2F gene, or a single E2F target. The rate-limiting targets for E2F function most likely vary from context to context, and they may not always be the usual suspects. In the D/V boundary of the developing wing disc, in which E2F/DP complexes protect from DNA damage-induced apoptosis, E2F/DP proteins are needed specifically to limit the expression of hid. Remarkably, the loss of E2F/DP leads to an upregulation of hid in this one part of the disc. This change occurs prior to irradiation, and it alters the cellular sensitivity to DNA damage. We found no apoptosis in unirradiated dDP mutant wing discs, implying that the change in hid expression in the dDP mutant wing disc is not, by itself, sufficient to induce apoptosis. The elevated hid expression is clearly important, because reducing the gene dosage of hid almost completely eliminated DNA damage-induced apoptosis in dDP mutant discs, but not in wild-type discs. What is the connection between dE2F1 and hid? Since dE2F1 binds to sequences upstream of the hid transcription start site, the transcription of hid is most likely reduced by the direct action of E2F complexes. Previous studies in mammalian cells have shown that E2F1 can directly repress transcription of some E2F1specific targets (Croxton et al., 2002a, 2002b), although

dE2F1 Has Both Pro- and Antiapoptotic Functions 473

the mechanisms underlying these effects are not well understood. The dE2F1 binding site upstream of hid has two interesting features that may be significant. First, unlike most dE2F-regulated promoters that have been examined to date, this binding site is bound specifically by dE2F1, but not by dE2F2 (Dimova et al., 2003). This specificity fits with the genetic evidence that de2f1, rather than de2f2, is important for protection from DNA damage-induced apoptosis, and it may explain why hid is not generally repressed by dE2F2 complexes. A second curious feature is that the dE2F1 binding site upstream of hid is surprisingly distal from the transcription start site. In most E2F-induced promoters, E2F binding sites are typically within 500 bp of the transcriptional start site. The position of the E2F1 binding site in the hid promoter, at −1.4 kb, may be part of the reason why hid differs from other dE2F1 targets and is not activated in a cell cycle-dependent manner. We note that the pattern of hid expression that sensitizes cells to apoptosis in dDP mutants occurs in the absence of E2F regulation; therefore, dE2F1 does not directly contribute to the pattern itself, but it presumably serves to interfere with another transcription factor (Factor X in Figure 7) that is specifically active within this region. As the hid promoter is largely uncharacterized, we do not exclude the possibility that dE2F1 may act through additional sites or that it may also repress hid expression via an indirect mechanism. In order to test this model, it will be necessary to identify the factors that control hid expression in vivo. Figure 7 illustrates a simple model for the contextspecific effects of dE2F1. In the intervein region, dE2F1 increases the expression of proapoptotic genes. In doing so, dE2F1 helps set a level of sensitivity for DNA damage-induced apoptosis, and this threshold is reduced when dE2F1 or dDP are removed. At the D/V boundary, dE2F1/dDP complexes are also needed, most likely in conjunction with RBF, to limit the expression of hid. When E2F regulation is removed, the increase in hid expression outweighs the changes in expression of other E2F targets, making cells more sensitive to apoptosis. If E2F’s contribution to the DNA damage response varies in mammalian cells as much as it does in Drosophila, then this would have implications for the use of general E2F inhibitors that are currently under development. These results suggest that a global inhibitor of E2F activity, or even a specific inhibitor of activator E2Fs, may have the unintended consequence of making some normal cell types very sensitive to DNA damage-induced apoptosis. Experimental Procedures Drosophila Strains dDP mutants: dDPa2/dDPa4 de2f1;de2f2 double mutants: de2f1729;de2f276q1/de2f191;de2f2g5 de2f2 mutants: de2f276q1/de2f2329;p[mpp6] de2f1i2 mutants: de2f1i2/de2f1729 and de2f1su89 To distinguish genotypes at the third instar larval stage, stocks

were maintained with balancers containing act-GFP or Tubby larval markers. To generate rbf1 mutant clones, the rbf1D14 mutant chromosome was recombined with a FRT19A chromosome. rbf1D14, FRT19A/FM7 virgin females were crossed with GFPubi,FRT19A; enGal4,UAS-FLP males. rbf1D14 mutant clones were identified in wing discs by the lack of a GFP signal. A hid deletion allele, X14, and a P element insertion allele, hidP05014, were used to generate flies heterozygous for hid. darkCD4 (Rodriguez et al., 1999) was recombined with dDPa4 to test the effects of hac-1 haplo-insufficiency; Df(3L)X38 removes reaper and sickle. Irradiation Wandering third instar larvae were exposed to 40 Gy of γ-ray by using Cs-137 as the source of irradiation. After 4 hr at 25°C, wing discs were dissected, fixed, and stained. In each experiment, a minimum of ten larvae was analyzed. Similar effects were seen in each set of discs examined. BrdU Labeling and Acridine Orange Staining Wing discs were dissected from third instar larvae into Schneider’s medium and incubated in medium containing 0.2 mg/ml BrdU for 30 min at room temperature. Discs were fixed with 4% formaldehyde for 30 min at 4°C, and DNA was denatured with 2 N HCl. BrdU-incorporated nuclei were visualized by immunocytochemistry with anti-BrdU (Becton Dickinson). For Acridine Orange staining, wing discs were dissected in PBS, incubated in PBS plus 1 ␮g/ml Acridine Orange for 5 min, washed with PBS three times for 5 min, and visualized by fluorescence microscopy. Immunocytochemistry and In Situ Hybridization Antibodies were obtained from the following sources: anti-Wingless (Developmental Studies Hybridoma Bank, DSHB), anti-p-H3 (Up-State Bio), anti-C3 (Cell Signaling), anti-Delta (DSHB), antidSRF (Active Motive), anti-dE2F1 (a gift from Carole Seum). In each figure, we have shown discs that represent the typical response of the genotype. The immunostaining protocol is provided in Supplemental Data. For in situ hybridization, samples were prepared as described (Du, 2000), and DIG-labeled antisense RNA probes were detected by an anti-DIG antibody conjugated with alkaline phosphatase. RNAi Treatment and Quantitative PCR RNAi treatments of S2 cells were performed as described previously (Dimova et al., 2003; Frolov et al., 2003). After treatment, total RNA was prepared with TRIzol (Invitrogen). Reverse transcription PCR (RT-PCR) was performed with PE Applied Biosystems Taq Man Reverse Transcription reagents, according to the manufacturer’s instructions. Real-time PCR was performed with an ABI prism 7700 Sequence Detection system. Relative levels of specific mRNAs were determined by using SYBR Green I detection chemistry (Applied Biosystems; Foster City, CA) and were quantified by using the comparative CT method. Rp49 was used for normalization. Primers (see Supplemental Data) were designed with Primer Express 1.0 software (Applied Biosystems; Foster City, CA). ChIP DNA sequences were retrieved from FlyBase (http://flybase.bio. indiana.edu/), and potential E2F binding sites were located by using PERL script with the sequence NWTSSCSS. ChIP was performed as described in Dimova et al., 2003 and Frolov et al., 2003. Coprecipitated DNA sequences were detected by real-time quantitative PCR. Reporter Assays PCNA wild-type promoter reporter plasmid (PCNAwt-luc) and a mutated control (PCNAE2Fmut-luc) have been described previously (Sawado et al., 1998). PCNA-hidE2F-luc, in which the hid E2F site replaces E2F sites in the PCNA promoter, was generated by site-directed mutagenesis. hidE2F-PCNAmut-luc was generated by PCR of lambda phage sequences with a primer carrying the hid E2F binding site. Amplified sequences were inserted 5# to PCNAE2Fmut-luc. Transient transfection assays were carried out

Developmental Cell 474

as described in Dynlacht et al., 1994. Luciferase and β-galactosidase were assayed by standard methods. Supplemental Data Supplemental Data including oligonucleotide sequences used in this paper and a detailed immunostaining protocol are available at http://www.developmentalcell.com/cgi/content/full/9/4/463/DC1. Acknowledgments We thank W. Du and R. Duronio for de2f1su89 and de2f2329;p[mpp6], G. Hurlbut for UAS-Rasv12, and S. Fre for UAS-Nic alleles. We thank A. Veraksa for anti-Dl and R. Patel for the hid in situ probe. This study was supported by National Institutes of Health grants PO1CA095281 and GM53203 (N.D.) and GM55568 (K.W.). N.-S.M. is a recipient of a Canadian Institute of Health Research fellowship. M.V.F. was a Leukemia & Lymphoma Society special fellow, and D.K.D. was a recipient of a fellowship from the Massachusetts General Hospital Fund for Medical Discovery. Received: March 16, 2005 Revised: July 19, 2005 Accepted: August 25, 2005 Published: October 3, 2005 References Asano, M., Nevins, J.R., and Wharton, R.P. (1996). Ectopic E2F expression induces S phase and apoptosis in Drosophila imaginal discs. Genes Dev. 10, 1422–1432. Aslanian, A., Iaquinta, P.J., Verona, R., and Lees, J.A. (2004). Repression of the Arf tumor suppressor by E2F3 is required for normal cell cycle kinetics. Genes Dev. 18, 1413–1422. Bergmann, A., Agapite, J., McCall, K., and Steller, H. (1998). The Drosophila gene hid is a direct molecular target of Ras-dependent survival signaling. Cell 95, 331–341. Bergmann, A., Yang, A.Y., and Srivastava, M. (2003). Regulators of IAP function: coming to grips with the grim reaper. Curr. Opin. Cell Biol. 15, 717–724. Bracken, A.P., Ciro, M., Cocito, A., and Helin, K. (2004). E2F target genes: unraveling the biology. Trends Biochem. Sci. 29, 409–417. Cam, H., Balciunaite, E., Blais, A., Spektor, A., Scarpulla, R.C., Young, R., Kluger, Y., and Dynlacht, B.D. (2004). A common set of gene regulatory networks links metabolism and growth inhibition. Mol. Cell 16, 399–411. Chaussepied, M., and Ginsberg, D. (2004). Transcriptional regulation of AKT activation by E2F. Mol. Cell 16, 831–837. Croxton, R., Ma, Y., and Cress, W.D. (2002a). Differences in DNA binding properties between E2F1 and E2F4 specify repression of the Mcl-1 promoter. Oncogene 21, 1563–1570. Croxton, R., Ma, Y., Song, L., Haura, E.B., and Cress, W.D. (2002b). Direct repression of the Mcl-1 promoter by E2F1. Oncogene 21, 1359–1369. De Celis, J.F. (2003). Pattern formation in the Drosophila wing: the development of the veins. Bioessays 25, 443–451. Dimova, D.K., Stevaux, O., Frolov, M.V., and Dyson, N.J. (2003). Cell cycle-dependent and cell cycle-independent control of transcription by the Drosophila E2F/RB pathway. Genes Dev. 17, 2308–2320. Du, W. (2000). Suppression of the rbf null mutants by a de2f1 allele that lacks transactivation domain. Development 127, 367–379. Du, W., and Dyson, N. (1999). The role of RBF in the introduction of G1 regulation during Drosophila embryogenesis. EMBO J. 18, 916–925. Du, W., Xie, J.E., and Dyson, N. (1996). Ectopic expression of dE2F and dDP induces cell proliferation and death in the Drosophila eye. EMBO J. 15, 3684–3692. Duronio, R.J., O’Farrell, P.H., Xie, J.E., Brook, A., and Dyson, N. (1995). The transcription factor E2F is required for S phase during Drosophila embryogenesis. Genes Dev. 9, 1445–1455. Dynlacht, B.D., Brook, A., Dembski, M., Yenush, L., and Dyson, N.

(1994). DNA-binding and trans-activation properties of Drosophila E2F and DP proteins. Proc. Natl. Acad. Sci. USA 91, 6359–6363. Dyson, N. (1998). The regulation of E2F by pRB-family proteins. Genes Dev. 12, 2245–2262. Evan, G.I., and Vousden, K.H. (2001). Proliferation, cell cycle and apoptosis in cancer. Nature 411, 342–348. Frolov, M.V., Huen, D.S., Stevaux, O., Dimova, D., Balczarek-Strang, K., Elsdon, M., and Dyson, N.J. (2001). Functional antagonism between E2F family members. Genes Dev. 15, 2146–2160. Frolov, M.V., Stevaux, O., Moon, N.S., Dimova, D., Kwon, E.J., Morris, E.J., and Dyson, N.J. (2003). G1 cyclin-dependent kinases are insufficient to reverse dE2F2-mediated repression. Genes Dev. 17, 723–728. Go, M.J., Eastman, D.S., and Artavanis-Tsakonas, S. (1998). Cell proliferation control by Notch signaling in Drosophila development. Development 125, 2031–2040. Grether, M.E., Abrams, J.M., Agapite, J., White, K., and Steller, H. (1995). The head involution defective gene of Drosophila melanogaster functions in programmed cell death. Genes Dev. 9, 1694– 1708. Hallstrom, T.C., and Nevins, J.R. (2003). Specificity in the activation and control of transcription factor E2F-dependent apoptosis. Proc. Natl. Acad. Sci. USA 100, 10848–10853. Hershko, T., and Ginsberg, D. (2004). Up-regulation of Bcl-2 homology 3 (BH3)-only proteins by E2F1 mediates apoptosis. J. Biol. Chem. 279, 8627–8634. Ianari, A., Gallo, R., Palma, M., Alesse, E., and Gulino, A. (2004). Specific role for p300/CREB-binding protein-associated factor activity in E2F1 stabilization in response to DNA damage. J. Biol. Chem. 279, 30830–30835. Irwin, M., Marin, M.C., Phillips, A.C., Seelan, R.S., Smith, D.I., Liu, W., Flores, E.R., Tsai, K.Y., Jacks, T., Vousden, K.H., and Kaelin, W.G., Jr. (2000). Role for the p53 homologue p73 in E2F-1-induced apoptosis. Nature 407, 645–648. Ishida, S., Huang, E., Zuzan, H., Spang, R., Leone, G., West, M., and Nevins, J.R. (2001). Role for E2F in control of both DNA replication and mitotic functions as revealed from DNA microarray analysis. Mol. Cell. Biol. 21, 4684–4699. Kurada, P., and White, K. (1998). Ras promotes cell survival in Drosophila by downregulating hid expression. Cell 95, 319–329. Lin, W.C., Lin, F.T., and Nevins, J.R. (2001). Selective induction of E2F1 in response to DNA damage, mediated by ATM-dependent phosphorylation. Genes Dev. 15, 1833–1844. Lissy, N.A., Davis, P.K., Irwin, M., Kaelin, W.G., and Dowdy, S.F. (2000). A common E2F-1 and p73 pathway mediates cell death induced by TCR activation. Nature 407, 642–645. Liu, K., Luo, Y., Lin, F.T., and Lin, W.C. (2004). TopBP1 recruits Brg1/ Brm to repress E2F1-induced apoptosis, a novel pRb-independent and E2F1-specific control for cell survival. Genes Dev. 18, 673–686. Moroni, M.C., Hickman, E.S., Denchi, E.L., Caprara, G., Colli, E., Cecconi, F., Muller, H., and Helin, K. (2001). Apaf-1 is a transcriptional target for E2F and p53. Nat. Cell Biol. 3, 552–558. Muller, H., and Helin, K. (2000). The E2F transcription factors: key regulators of cell proliferation. Biochim. Biophys. Acta 1470, M1–12. Muller, H., Bracken, A.P., Vernell, R., Moroni, M.C., Christians, F., Grassilli, E., Prosperini, E., Vigo, E., Oliner, J.D., and Helin, K. (2001). E2Fs regulate the expression of genes involved in differentiation, development, proliferation, and apoptosis. Genes Dev. 15, 267– 285. Nahle, Z., Polakoff, J., Davuluri, R.V., McCurrach, M.E., Jacobson, M.D., Narita, M., Zhang, M.Q., Lazebnik, Y., Bar-Sagi, D., and Lowe, S.W. (2002). Direct coupling of the cell cycle and cell death machinery by E2F. Nat. Cell Biol. 4, 859–864. Ohtani, K., and Nevins, J.R. (1994). Functional properties of a Drosophila homolog of the E2F1 gene. Mol. Cell. Biol. 14, 1603–1612. Pediconi, N., Ianari, A., Costanzo, A., Belloni, L., Gallo, R., Cimino, L., Porcellini, A., Screpanti, I., Balsano, C., Alesse, E., et al. (2003). Differential regulation of E2F1 apoptotic target genes in response to DNA damage. Nat. Cell Biol. 5, 552–558.

dE2F1 Has Both Pro- and Antiapoptotic Functions 475

Phillips, A.C., and Vousden, K.H. (2001). E2F-1 induced apoptosis. Apoptosis 6, 173–182. Ren, B., Cam, H., Takahashi, Y., Volkert, T., Terragni, J., Young, R.A., and Dynlacht, B.D. (2002). E2F integrates cell cycle progression with DNA repair, replication, and G(2)/M checkpoints. Genes Dev. 16, 245–256. Rodriguez, A., Oliver, H., Zou, H., Chen, P., Wang, X., and Abrams, J.M. (1999). Dark is a Drosophila homologue of Apaf-1/CED-4 and functions in an evolutionarily conserved death pathway. Nat. Cell Biol. 1, 272–279. Royzman, I., Hayashi-Hagihara, A., Dej, K.J., Bosco, G., Lee, J.Y., and Orr-Weaver, T.L. (2002). The E2F cell cycle regulator is required for Drosophila nurse cell DNA replication and apoptosis. Mech. Dev. 119, 225–237. Royzman, I., Whittaker, A.J., and Orr-Weaver, T.L. (1997). Mutations in Drosophila DP and E2F distinguish G1-S progression from an associated transcriptional program. Genes Dev. 11, 1999–2011. Sawado, T., Yamaguchi, M., Nishimoto, Y., Ohno, K., Sakaguchi, K., and Matsukage, A. (1998). dE2F2, a novel E2F-family transcription factor in Drosophila melanogaster. Biochem. Biophys. Res. Commun. 251, 409–415. Sears, R.C., and Nevins, J.R. (2002). Signaling networks that link cell proliferation and cell fate. J. Biol. Chem. 277, 11617–11620. Stevaux, O., Dimova, D., Frolov, M.V., Taylor-Harding, B., Morris, E., and Dyson, N. (2002). Distinct mechanisms of E2F regulation by Drosophila RBF1 and RBF2. EMBO J. 21, 4927–4937. Stevens, C., Smith, L., and La Thangue, N.B. (2003). Chk2 activates E2F-1 in response to DNA damage. Nat. Cell Biol. 5, 401–409. Stiewe, T., and Putzer, B.M. (2000). Role of the p53-homologue p73 in E2F1-induced apoptosis. Nat. Genet. 26, 464–469. Thacker, S.A., Bonnette, P.C., and Duronio, R.J. (2003). The contribution of E2F-regulated transcription to Drosophila PCNA gene function. Curr. Biol. 13, 53–58. Trimarchi, J.M., and Lees, J.A. (2002). Sibling rivalry in the E2F family. Nat. Rev. Mol. Cell Biol. 3, 11–20. Tsai, K.Y., Hu, Y., Macleod, K.F., Crowley, D., Yamasaki, L., and Jacks, T. (1998). Mutation of E2f-1 suppresses apoptosis and inappropriate S phase entry and extends survival of Rb-deficient mouse embryos. Mol. Cell 2, 293–304. Weng, L., Zhu, C., Xu, J., and Du, W. (2003). Critical role of active repression by E2F and Rb proteins in endoreplication during Drosophila development. EMBO J. 22, 3865–3875. Zhou, L., and Steller, H. (2003). Distinct pathways mediate UVinduced apoptosis in Drosophila embryos. Dev. Cell 4, 599–605. Ziebold, U., Reza, T., Caron, A., and Lees, J.A. (2001). E2F3 contributes both to the inappropriate proliferation and to the apoptosis arising in Rb mutant embryos. Genes Dev. 15, 386–391.

Lihat lebih banyak...

Comentários

Copyright © 2017 DADOSPDF Inc.