Article
5 June 2008
Free access

The DNA replication checkpoint aids survival of plants deficient in the novel replisome factor ETG1

The EMBO Journal
(2008)
27: 1840 - 1851
Complete and accurate chromosomal DNA replication is essential for the maintenance of the genetic integrity of all organisms. Errors in replication are buffered by the activation of DNA stress checkpoints; however, in plants, the relative importance of a coordinated induction of DNA repair and cell cycle‐arresting genes in the survival of replication mutants is unknown. In a systematic screen for Arabidopsis thaliana E2F target genes, the E2F TARGET GENE 1 (ETG1) was identified as a novel evolutionarily conserved replisome factor. ETG1 was associated with the minichromosome maintenance complex and was crucial for efficient DNA replication. Plants lacking the ETG1 gene had serrated leaves due to cell cycle inhibition triggered by the DNA replication checkpoints, as shown by the transcriptional induction of DNA stress checkpoint genes. The importance of checkpoint activation was highlighted by double mutant analysis: whereas etg1 mutant plants developed relatively normally, a synthetically lethal interaction was observed between etg1 and the checkpoint mutants wee1 and atr, demonstrating that activation of a G2 cell cycle checkpoint accounts for survival of ETG1‐deficient plants.

Introduction

All eukaryotic cells replicate their nuclear DNA in a conserved manner, whereby the parent molecule is unwound and each DNA strand becomes the template for nascent DNA synthesis. For survival, DNA replication must be coordinated with growth and development. In mammals, genetic instability is a hallmark of cancer cells (Bartkova et al, 2005; Gorgoulis et al, 2005). Therefore, cells have developed signalling cascades to guard against DNA stress. DNA damage and problems of DNA replication activate the ataxia telangiectasia‐mutated (ATM) and Rad3‐related (ATR) signalling kinases that simultaneously switch on DNA repair genes and arrest cell cycle progression, allowing cells to repair DNA before entering mitosis. ATM responds specifically to double‐stranded breaks, whereas ATR primarily senses replication stress caused by blocking of the DNA replication fork progression. Also in plants, ATM and ATR concomitantly induce the expression of DNA repair genes and WEE1, which arrests cells in the G2 phase of the cell cycle (Garcia et al, 2003; Culligan et al, 2004, 2006; De Schutter et al, 2007).
Because of its essential role during development, it is not surprising that the molecular machinery controlling DNA replication is highly conserved. Despite the many years of evolution that separate animals from plants, the same E2F/DP pathway regulates entry into S phase by mediating transcriptional induction of genes required for cell cycle progression and DNA replication. The genome of Arabidopsis thaliana contains six E2Fs and two DPs (Inzé and De Veylder, 2006). Three E2F proteins (E2Fa, E2Fb, and E2Fc) bind DNA through the consensus E2F‐binding site by forming heterodimers with DP proteins (DPa and DPb). Both E2Fa and E2Fb operate as transcriptional activators, whereas E2Fc functions as a repressor (De Veylder et al, 2002; del Pozo et al, 2002; Rossignol et al, 2002; Magyar et al, 2005). The remaining Arabidopsis E2Fs (E2Fd/DEL2, E2Fe/DEL1, and E2Ff/DEL3) contain duplicated DNA‐binding domains, allowing association to consensus E2F sites in an DP‐independent manner (Kosugi and Ohashi, 2002; Ramirez‐Parra et al, 2004; Vlieghe et al, 2005).
Both in mammals and Arabidopsis, numerous E2F target genes have been identified by using microarrays, chromatin immunoprecipitations (ChIPs), and in silico analyses (Ramirez‐Parra et al, 2003; Dimova and Dyson, 2005; Vandepoele et al, 2005). These genes code for proteins that are active during DNA replication, mitosis, DNA checkpoint control, apoptosis, or differentiation. Remarkably, almost every gene encoding a protein involved in licensing for DNA replication is transcriptionally controlled by E2F transcription factors (Vandepoele et al, 2005). Licensing for DNA replication is initiated by the formation of the prereplicative complex at replication origins (Gillespie et al, 2001; Bell and Dutta, 2002; Diffley and Labib, 2002). First, the origin recognition complex (ORC) proteins bind to DNA during the early G1 to S phase of the cell division cycle. Then, the cell division cycle 6 (CDC6) protein binds to these ORC DNA sites, an event that is followed quickly by binding of chromatin licensing and DNA replication factor 1 (CDT1). Finally, the replication origins are licensed by loading the minichromosome maintenance (MCM) complex to form a prereplicative complex. The MCM complex is a heterohexamer composed of MCM2 to MCM7 and is likely a component of the helicase that unwinds DNA during replication (Tye and Sawyer, 2000; Labib and Diffley, 2001; Forsburg, 2004). Once the prereplicative complex is completely formed by the loading of the MCM complexes, the DNA is primed for replication through the action of two conserved protein kinases, the cyclin‐dependent kinase (CDK) and Cdc7–Dbf4, with the recruitment of additional replication factors as a consequence to create the preinitiation complex (Kamimura et al, 2001; Masumoto et al, 2002; Takayama et al, 2003; Kanemaki and Labib, 2006). Loading of the preinitiation complex onto the origins activates the MCM helicases and DNA polymerases, resulting in the initiation of DNA synthesis (Zou and Stillman, 2000). After activation of the origins, some of the initiation factors move with the replication forks to support the elongation step of DNA synthesis (Aparicio et al, 1997; Takayama et al, 2003; Gambus et al, 2006; Kanemaki and Labib, 2006).
By studying E2F target genes in Arabidopsis, we identified a novel E2F target gene that is directly controlled by the E2Fa and E2Fb transcription factors. Null mutants of this E2F TARGET GENE 1 (ETG1) had a slower cell cycle. Genetic analysis and transcriptional upregulation of the poly(ADP‐ribose) polymerase 2 (PARP2), WEE1, and RAD51 genes revealed that the cell cycle delay in etg1 mutant plants was due to the activation of the DNA replication checkpoint. Interestingly, the absence of a functional ETG1 allele in a wee1 or atr mutant background had a profound impact on plant development, illustrating that the DNA replication checkpoint and, more particularly, its effects on the cell cycle, aids the survival of ETG1‐deficient plants.

Results

Loss of ETG1 decreases the cell division rate

Previously, we have identified 70 conserved plant E2F target genes (Vandepoele et al, 2005). This list holds 40 known regulators of DNA replication and chromatin dynamics, but also 21 genes with unknown function. To identify novel S‐phase regulatory genes, we screened the latter phenotypically with T‐DNA insertion lines. One of the insertion lines showed an endoreduplication phenotype and was designated E2F TARGET GENE 1 (ETG1; At2g40550). To address the role of the ETG1 gene in plant growth and development, we analysed the loss‐of‐function effect of ETG1 with two independent T‐DNA insertion lines. The T‐DNA was inserted into the first intron (etg1‐1; SALK_071046) or last exon (etg1‐2; SALK_145460) of the ETG1 gene (Figure 1A). ETG1 transcripts were not detected in the etg1‐1 mutant, whereas the transcript level was 80% reduced in etg1‐2, when compared to control plants (Figure 1B). In etg1 mutant seedlings, plant growth appeared macroscopically normal (Figure 1C). However, by comparing the DNA ploidy level of wild‐type leaves with that of etg1‐1 and etg1‐2 mutants, the distribution of the C values was slightly, but significantly, changed: in etg1 mutants, the population of cells with an 8C and 16C DNA ploidy level had increased, demonstrating that deficiency in ETG1 stimulated endoreduplication (Figure 1D). A similar effect was seen in root tissue (Supplementary Figure S1). The increase in the DNA ploidy level of etg1 mutants probably originated from an advanced onset of the endocycle, as illustrated by the faster increase in the population of cells with an 8C and 16C DNA content during leaf development (Supplementary Figure S2).
image
Figure 1. Molecular and phenotypic analysis of ETG1‐deficient plants. (A) Exon (boxes) and intron (lines) structure of ETG1. Coding and non‐coding regions are shown as black and white boxes, respectively. White triangles and arrows indicate T‐DNA insertion sites and primer positions used for real‐time RT–PCR analysis, respectively. (B) Real‐time RT–PCR analysis of ETG1 expression in wild‐type (WT), etg1‐1, and etg1‐2 plants. Total RNA prepared from first leaves of 9‐day‐old plants was amplified by RT–PCR. All values were normalized against the expression level of the ACTIN2 gene. (C) Seedling phenotypes of 21‐day‐old wild‐type (WT) and etg1‐1 plants. (D) Ploidy level distribution of the first leaves of 3‐week‐old wild‐type (WT), etg1‐1, and etg1‐2 plants as measured by flow cytometry. Data represent average±s.d. (n=5). *Significant statistical differences by t‐test (P<0.05) between wild type and etg1. (E) Drawing‐tube image of the first leaves of 3‐week‐old wild‐type (left) and etg1‐1 (right) plants. Bar=100 μm. (FH) Leaf growth of the first leaf pair of wild‐type (WT), etg1‐1, and etg1‐2 plants. Leaf blade area (F), epidermal cell number (G), and epidermal cell size (H) on the abaxial side of the leaf. Data represent average±s.d. (n=5). (I, J) Adult phenotype of wild‐type (WT) and etg1‐1 plants. The plants were photographed 5 weeks after germination. (J) Magnification of leaves in wild‐type (WT) and etg1‐1 plants.
When the first leaf pair from wild‐type and etg1 mutant plants at maturity was compared, the leaf blade area was almost identical for both genotypes (Figure 1F). By contrast, the average abaxial pavement cell area had increased significantly in the mutant plants (Figures 1E and G), accompanied with a decrease in cell number per leaf (Figure 1H). At the bolting stage, younger leaves showed a slightly elongated and serrated leaf phenotype (Figures 1I and J), resembling the phenotype observed in plants whose cell division is inhibited by ectopic expression of the CDK inhibitory gene KRP2 (De Veylder et al, 2001). In addition, the root growth rate of the mutant plants was significantly reduced (Supplementary Figure S3).
To study the effect of loss of ETG1 function on cell cycle progression in more detail, kinematics of leaf growth was analysed. From day 5 until day 22 after sowing, the first leaves of etg1‐1 and wild‐type plants grown side by side under the same conditions were harvested and the leaf blade area and the average cell area of the abaxial epidermal cells was measured by image analysis (De Veylder et al, 2001; Boudolf et al, 2004). The total cell number was extrapolated as the ratio of leaf blade and average cell areas. Although the leaf blade area was similar in the wild‐type and etg1‐1 plants during the whole period of leaf development (Figure 2A), the average cell area, which initially was approximately 100 μm2 in both plants, increased significantly faster in the etg1‐1 mutant (Figure 2B). The average surface area of etg1‐1 cells was 155% that of wild‐type cells at maturity (3880±320 versus 2500±255 μm2). Simultaneously, the number of epidermal cells of etg1‐1 was only approximately 60% that of the wild type (6650±530 versus 11 170±1017 cells; Figure 2C). Until day 9 after sowing, the average cell division rate for the whole leaf, calculated on the basis of the increase in cell numbers over time, were constantly lower in the etg1‐1 than in wild‐type leaves (Figure 2D). The average cell cycle duration between days 5 and 9, estimated as the inverse of the cell division rate, was significantly longer in the etg1‐1 mutant (25.3 h) than that in the wild‐type plants (21.1 h). In summary, these data illustrate that ETG1‐deficient plants suffer from a cell cycle delay, with a reduced total cell number that is offset by an enlarged cell size.
image
Figure 2. Kinematic analysis of first leaf pair of wild‐type (WT) and etg1‐1 plants. (A) Leaf blade area. (B) Average cell area on the abaxial side of the leaf. (C) Number of cells on the abaxial side of leaves. (D) Cell division rate. Leaves were harvested at the indicated time points. Data are average±s.d. (n=5).

Loss of ETG1 function causes a G2 cell cycle arrest

To pinpoint the cell cycle arrest point, we measured the ratio of 4C/2C cells of 8‐day‐old leaves by flow cytometry. At this time point, leaf cells of both genotypes divide (Figures 2C and D); consequently, 2C and 4C cells represent G1 and G2 cells, respectively. By comparing the ploidy level of wild type and etg1‐1, a significant increase in the 4C/2C cell ratio was observed (0.29±0.06 and 0.79±0.04 in wild‐type and etg1‐1 plants, respectively; Figures 3A and B). These data indicate that the G2‐to‐M transition was inhibited in the etg1‐1 mutant. Analysis of the expression levels of cell cycle marker genes by real‐time reverse transcriptase (RT)–PCR revealed that the transcript level of the G2/M‐phase‐specific CYCB1;1, CDKB1;1, and KNOLLE genes had increased in the etg1‐1 mutant, whereas those of the S‐phase‐specific histone H4 and CYCA3;1 genes was unaltered (Figure 3C). Despite the increased transcript levels of G2/M‐phase‐specific genes, a decrease in A‐type CDK activity was observed in the etg1‐1 mutant plants (Supplementary Figure S4), in agreement with the observed growth delay and extended cell cycle duration. Counting the number of root cells within the meristem in either metaphase or anaphase/telophase showed that the amount of M‐phase nuclei was approximately two‐fold lower in the etg1 mutant than that of wild‐type plants (Figures 3D–F). Combined with the observed cell cycle delay, these data indicate a prolonged G2 phase in the etg1 mutant. More particularly, induction of the CYCB1;1 and KNOLLE genes implies that the cell cycle arrest occurs around late G2.
image
Figure 3. G2 arrest caused by ETG1 depletion. (A) Ploidy‐level distribution of the first leaves of 8‐day‐old wild‐type (WT) and etg1‐1 plants as measured by flow cytometry. (B) Ratio of 4C/2C cells of the first leaves of 8‐day‐old wild‐type (WT) and etg1‐1 plants. Data represent average±s.d. (n=3). (C) Real‐time RT–PCR analysis of S phase (histone H4 and CYCA3;1) and G2/M phase (CDKB1;1, KNOLLE, and CYCB1;1) expression markers in wild‐type (white) and etg1‐1 (black) plants. Total RNA prepared from 9‐day‐old seedlings was amplified by RT–PCR. All values were normalized against the expression level of the ACTIN2 gene. (D, E) Distribution of mitotic figures in wild‐type (D) and etg1‐1 (E) plants. Metaphase and anaphase/telophase nuclei are false‐colored in red and yellow, respectively. (F) Number of mitotic cells per root tip of 3‐day‐old wild‐type (WT) and etg1‐1 seedlings. Error bars indicate s.e. (n=12); *Significant statistical differences by t‐test (P<0.05) between wild type and etg1‐1.

ETG1 transcript is regulated by E2Fa and E2Fb transcription factors

The ETG1 gene had originally been identified by microarray analysis as an induced transcript in plants ectopically expressing the heterodimeric E2FaDPa transcription factor (Vandepoele et al, 2005). This induction was confirmed by quantitative real‐time PCR analysis (Figure 4C). To analyse whether the ETG1 gene was directly regulated by E2F transcription factors, we looked for the presence of consensus E2F‐binding sites in its putative promoter region. Two consensus E2F‐binding elements were found, ATTCCCGC and TTTCCCGC (158 and 136 bp upstream from the putative start codon, respectively), both in a reverse orientation (Figure 4A). To address whether ETG1 is an E2F target gene in vivo, we used ChIP with antibodies against E2Fa, E2Fb, and E2Fc. Promoter fragments of the ETG1 gene were specifically amplified from the anti‐E2Fa and anti‐E2Fb immunoprecipitates. These results indicate that E2Fa and E2Fb can bind directly to the ETG1 promoter in vivo, probably participating in the regulation of its expression. By contrast, the ETG1 promoter sequences were not recovered from the immunoprecipitates with anti‐E2Fc antibodies (Figure 4B). Correspondingly, ETG1 expression levels were not altered in plants that overexpress E2Fc (Supplementary Figure S5).
image
Figure 4. Regulation of ETG1 gene expression by the E2Fa transcription factor. (A) Sequence of the ETG1 promoter showing the presence of two E2F‐binding sites at −158 (I) and −136 (II) bp from the ATG translation start codon. (B) ChIP assays in 8‐day‐old Arabidopsis plants with antibodies specific for E2Fa, E2Fb, and E2Fc. Precipitation with an anti‐HA antibody (IP HA) or without antibody (IP Ab) was used as negative controls. The relative enrichment of genomic fragments of the ETG1 promoter was estimated by semiquantitative PCR. (C) Real‐time RT–PCR analysis of ETG1 expression in wild‐type (WT) and E2Fa/DPa‐overexpressing (E2Fa/DPa OE) plants. Total RNA prepared from 6‐day‐old plants was amplified by RT–PCR. All values were normalized against the expression level of the ACTIN2 gene. (D) Histochemical localization of the GUS activity in transgenic 6‐day‐old shoot (upper) and root (bottom) apical meristems of seedlings carrying the wild‐type (WT) ETG1 promoter, or a mutant promoter with either one (ΔI or ΔII), or both (ΔI, II) E2F elements exactly deleted.
The regulation of the ETG1 promoter activity through its E2F consensus sites was further analysed with transgenic plants expressing the β‐glucuronidase (GUS) reporter gene under control of the ETG1 promoter. To define the contribution of each of the E2F‐binding sites, we deleted exactly either one (ΔI or ΔII) or both (ΔI, II) of the E2F elements. More than five independent transgenic lines were analysed per gene construct, all with identical results. In 6‐day‐old seedlings, high levels of ETG1 expression were observed in the shoot apical and root meristems (Figure 4D). This expression pattern matched that of the E2Fa and DPa genes (De Veylder et al, 2002). Deletion of either one of the E2F‐binding elements (ΔI or ΔII) led to GUS activity patterns identical to those of plants carrying the wild‐type ETG1 promoter (Figure 4D). In contrast, deletion of both E2F‐binding elements resulted in a drastic decrease in promoter activity (Figure 4D). These results suggest that E2Fa and E2Fb bind both E2F consensus elements in the ETG1 promoter and regulate its expression in dividing tissues.

ETG1 is a nuclear protein conserved in eukaryotes

The ETG1 gene encodes a protein of 598 amino‐acid residues (Supplementary Figure S6). ETG1 is a singleton in Arabidopsis. No specific amino‐acid domain could be identified with the exception of a putative nuclear localization signal, PFKKMKV (amino acids 184–190), suggesting that ETG1 resides in the nucleus. To investigate the subcellular localization of ETG1, a fusion protein of ETG1 and an enhanced green fluorescent protein (eGFP) was transiently produced in leaf epidermal cells of tobacco (Nicotiana tabacum). ETG1–eGFP fluorescence was observed in the nucleus only, illustrating that ETG1 is a nuclear protein (Figure 5A). Orthologous proteins of ETG1 were found in rice (Oryza sativa; Os01g0166800), human (Homo sapiens; C10orf119), mouse (Mus musculus; 1110007A13Rik), Xenopus laevis (CAJ81286), Drosophila melanogaster (CG3430), and fission yeast (Schizosaccharomyces pombe; SPAC1687.04) (Supplementary Figure S6), but none in budding yeast (Saccharomyces cerevisiae). Interestingly, the transcript of the human ETG1 is significantly upregulated in brain cancer, breast cancer, and seminoma (http://www.oncomine.org).
image
Figure 5. Assembly of ETG1 into the replisome. (A) Subcellular localization of ETG1. The full‐length ETG1–eGFP fusion protein is localized in the nucleus. (B) ETG1 interaction with MCM5 in yeast. Yeast PJ69‐4a cells were transformed with a plasmid encoding a fusion between the GAL4 DNA‐binding domain and ETG1 and MCM5 (GAL4‐DBD–ETG1 and GAL4‐DBD–MCM5, respectively). Yeast PJ69‐4α cells were transformed with fusions of GAL4 activation domain (AD) with ETG1, MCM5, and GUS as negative controls (GAL4‐AD–ETG1, GAL‐AD–MCM5, and GAL4‐AD–GUS, respectively). Diploid strains were spotted on plates with (+His, positive control) or without (−His) histidine. (C) BiFC assay of ETG1 interaction with MCM5 in plant. Tobacco epidermal cells were transfected with a combination of constructs encoding the indicated fusion proteins. YFPN and YFPC are the fragments containing the amino‐acid residues 1–154 and 155–238 of YFP, respectively. Arrowheads mark nuclei. (D) Subcellular localization of MCM5. The full‐length GFP–MCM5 fusion protein was localized in the nucleus and the cytoplasm. (E) DNA replication level of the wild‐type (WT) and etg1‐1 plants. Seedlings (3 day old) were soaked in BrdU solution for 3 h. Genomic DNA was extracted and the amount of BrdU incorporation was determined by ELISA with an anti‐BrdU antibody. The amount of BrdU incorporated into the wild‐type plants was arbitrarily set to 100%. Data represent average±s.d. (n=3).

ETG1 is a component of the replisome complex and needed for DNA replication

Coexpression patterns can reveal networks of functionally related genes and provide a deeper understanding of processes requiring multiple gene products (Stuart et al, 2003; Wei et al, 2006). To predict the ETG1 function, we searched for genes coexpressed with ETG1 by using the ATTED‐II coexpression database (Obayashi et al, 2007). This search revealed that ETG1 is highly coexpressed with genes encoding DNA replication proteins, such as the MCM family proteins (MCM2, MCM3, MCM4, MCM5, and MCM7), proliferating cell nuclear antigen proteins (PCNA1 and PCNA2), DNA primase small subunit protein, and DNA polymerase α subunits (Table I). Moreover, when searching for proteins interacting with orthologous ETG1 proteins with the BioGRID protein interactions database (http://www.thebiogrid.org/index.php), we found that the Drosophila orthologue CG3430 interacted with the MCM5 protein. An identical interaction between the Arabidopsis ETG1 and MCM5 (At2g07690) proteins was demonstrated with the yeast two‐hybrid system (Figure 5B). This protein–protein interaction was confirmed in planta by bimolecular fluorescence complementation (BiFC) experiments (Bracha‐Drori et al, 2004; Walter et al, 2004). The sequences coding for ETG1 and MCM5 were fused in frame with either the N‐terminal or C‐terminal fragment of the yellow fluorescent protein (YFP) (ETG1–YFPN or MCM5‐YFPC, respectively). Interaction between the different fusion proteins was tested by introducing them into tobacco leaf epidermal cells. YFP fluorescence was observed in the nuclei of cells transfected with ETG1–YFPN and MCM5–YFPC, demonstrating that the ETG1 protein interacted with MCM5 in the plant nucleus (Figure 5C; ETG1–YFPN and MCM5–YFPC). As expected, no fluorescence was detected when any combination with empty vectors was introduced into tobacco cells (Figure 5C; ETG1–YFPN and YFPC, YFPN and MCM5–YFPC). When the subcellular localization of MCM5 was examined in plants with MCM5–eGFP, the fusion protein MCM5–eGFP resided in both the nucleus and cytoplasm (Figure 5D).
Table 1. Genes coexpressed with the ETG1 gene
RankCorrelation coefficientLocusGene description
10.92At5g46280Minichromosome maintenance family protein 3 (MCM3)
20.90At1g44900Minichromosome maintenance family protein 2 (MCM2)
30.89At4g02060Prolifera protein (PRL)/minichromosome maintenance family protein 7 (MCM7)
40.89At5g41880DNA primase small subunit
50.88At1g07370Proliferating cell nuclear antigen 1 (PCNA1)
60.88At2g16440Minichromosome maintenance family protein 4 (MCM4)
70.87At2g07690Minichromosome maintenance family protein 5 (MCM5)
80.84At1g67630DNA polymerase α subunit B
90.84At5g67100DNA polymerase α catalytic subunit
100.83At2g29570Proliferating cell nuclear antigen 2 (PCNA2)
Top 10 ranking genes identified by ATTED‐II coexpression database.
To identify additional ETG1‐associated proteins, tandem affinity purification (TAP) was confirmed with MALDI‐TOF‐TOF‐MS‐based protein identification (Van Leene et al, 2007). Arabidopsis cell suspension cultures were stably transformed by Agrobacterium tumefaciens‐mediated cocultivation with a Pro35S:TAP‐ETG1 cassette. On the basis of the presence of the double tag, two‐step affinity purification was performed. Besides MCM5, we identified five interacting proteins, including other components of the MCM complex, namely MCM2 (At1g44900), MCM3 (At5g46280), MCM4 (At2g16440), MCM6 (At5g44635), and MCM7 (At4g02060) (Table II; Supplementary Figure S7). Combined with the subcellular localization results, these data indicate that ETG1 assembled into the replisome complex. Therefore, ETG1 depletion was expected to affect the efficiency of DNA replication. This hypothesis was substantiated by the bromodeoxyuridine (BrdU) incorporation rate that was lower in etg1‐1 mutant plants than that in the control plants (Figure 5E). Although this decrease in BrdU incorporation might be unexpected in the light of the observed increase in the endoreduplication level, the reason might be that the endocycle is far slower than the mitotic cell cycle (Beemster et al, 2005). Therefore, because of the short labelling period, the BrdU label probably refers to the mitotic S phase only.
Table 2. List of ETG1‐copurified proteins identified by MS
LocusaGene descriptionSNAPS accession numberProtein MWPeptide countSequence coverage (%)Protein score/ thresholdBest ion score/ threshold
At1g44900Minichromosome maintenance family protein 2 (MCM2)nrAt0.2_52915105 1723343936/61104/33
At5g46280Minichromosome maintenance family protein 3 (MCM3)nrAt0.2_6830286 7591423206/6174/31
At2g16440Minichromosome maintenance family protein 4 (MCM4)nrAt0.2_3361794 1682438696/61102/28
At2g07690Minichromosome maintenance family protein 5 (MCM5)nrAt0.2_1996081 5912638584/61107/31
At5g44635Minichromosome maintenance family protein 6 (MCM6)nrAt0.2_837493 4782033859/61154/30
At4g02060Prolifera protein (PRL)/ minichromosome maintenance family protein 7 (MCM7)nrAt0.2_390780 7392646987/61172/31
a
All proteins were detected in the two independent TAP experiments.

ETG1‐deficient plants activate the DNA replication checkpoint

Inhibition of DNA replication in Arabidopsis results in the simultaneous induction of DNA repair genes and the cell cycle inhibitory WEE1 gene that arrests cells in the G2 phase of the cell cycle (De Schutter et al, 2007). The decreased rate of BrdU incorporation and observed interaction of ETG1 with MCM proteins suggested that the G2 arrest noticed in ETG1‐deficient plants might be the consequence of replication checkpoint activation. The activation of a cell cycle checkpoint might be the reason for the observed decrease in CDK activity as well. To test this hypothesis, we compared the expression levels of the RAD51 (DNA repair) and WEE1 (cell cycle checkpoint) marker genes by real‐time RT–PCR in wild‐type and etg1 mutant plants. Ionizing radiation (for example, γ‐irradiation and UV light), DNA replication inhibitory drugs (for example, hydroxyurea and aphidicolin), and radiomimetic agents (such as bleomycin) are known to induce RAD51 and WEE1 expression (Chen et al, 2003; De Schutter et al, 2007). Expression of both RAD51 and WEE1 was significantly upregulated in the etg1‐1 (Figure 6A) and similarly in etg1‐2 seedlings (data not shown). Activation of the DNA stress checkpoint was confirmed by using plants that carried as transgenes promoters of PARP2 and WEE1 fused to GUS, which are markers for DNA stress and activation of the replication checkpoint, respectively (Babiychuk et al, 1998; Doucet‐Chabeaud et al, 2001; De Schutter et al, 2007). No GUS activity was observed in PARP2:GUS plants grown under non‐stress conditions (Figure 6B). By contrast, treatment of the PARP2:GUS reporter line with bleomycin resulted in a strong induction of GUS activity (Figure 6D), demonstrating the DNA stress‐inducible promoter activity. Similarly, PARP2 promoter activity was induced in an etg1‐1 background in the absence of any external DNA stress stimulus (Figure 6C). Especially, GUS activity was strongly induced in shoot apical meristem and vascular cells. Analogous results were obtained with WEE1:GUS reporter plants. In control plants, WEE1 expression was observed in the shoot apex and vascular cells (Figure 6E; De Schutter et al, 2007). This expression pattern was intensified in the etg1‐1 background (Figure 6F), confirming the real‐time RT–PCR experiments. The ETG1 gene itself was not induced under conditions that cause DNA stress, such as γ‐irradiation or UV‐B treatment (Ulm et al, 2004; Culligan et al, 2006). Additionally, etg1‐1 mutant plants were not hypersensitive to hydroxyurea or bleomycin (Supplementary Figure S8).
image
Figure 6. Expression of DNA replication checkpoint and DNA repair genes activated by loss of ETG1. (A) Real‐time RT–PCR analysis of WEE1 and RAD51 expression in wild‐type (white) and etg1‐1 (black) plants. Total RNA prepared from 9‐day‐old seedlings was amplified by RT–PCR. All values were normalized against the expression level of the ACTIN2 gene. * and **Significant statistical differences by t‐test (P<0.05 and P<0.01, respectively) between wild‐type and etg1‐1. Data represent average±s.d. (n=3). (BD) Histochemical localization of GUS activity in 6‐day‐old plants harbouring the PARP2:GUS reporter construct. (B) PARP2:GUS in a wild‐type background. (C) PARP2:GUS introgressed into etg1‐1. (D) PARP2:GUS in a wild‐type background in the presence 1 μg ml−1 bleomycin. (E, F) Histochemical localization of GUS activity in 6‐day‐old plants harbouring the WEE1:GUS reporter construct in a wild‐type background (E) and introgressed into etg1‐1 (F).

Activation of the DNA replication checkpoint aids the survival of ETG1‐deficient plants

DNA replication stress caused by blocking of the replication fork is mainly sensed by the ATR kinase (Culligan et al, 2004). Previously, we have demonstrated that the WEE1 gene is transcriptionally activated upon replication stress in an ATR‐dependent manner and transiently arrests cells in the G2 phase, allowing them to finalize DNA replication before proceeding into mitosis (De Schutter et al, 2007). When the increased cell cycle duration observed in the etg1 mutant plants is assumed to result from the activation of the replication checkpoint, ETG1 deficiency is expected to have a dramatic impact on the development of plants that are unable to arrest their cell cycle in response to DNA stress. To test this hypothesis, double mutants were constructed between etg1‐1 and two DNA stress checkpoint mutants, atr‐2 and wee1‐1. WEE1‐deficient plants fail to arrest their cell cycle when DNA replication is perturbed, but DNA repair genes are still induced (De Schutter et al, 2007; T Cools, DI and LDV, unpublished data). In contrast, atr mutant plants fail both to arrest their cell cycle and to induce repair genes (Culligan et al, 2004, 2006). Both wee1‐1 and atr‐2 mutants were hypersensitive to replication‐blocking or DNA‐damaging drugs, but were viable and developed normally in the absence of exogenous DNA stress treatments (Figures 7A–D). By contrast, etg1‐1 wee1‐1 and etg1‐1 atr‐2 double mutant plants had a dwarf phenotype under non‐stress conditions, illustrating a synthetic interaction between ETG1 and WEE1, or ATR (Figures 7E–H). Scanning electron microscopy revealed severe growth suppression (Figures 7I–K). Especially, the size of trichomes was reduced in the double mutants (Figures 7O–Q). No significant difference in leaf epidermal cell shape was observed in etg1‐1 wee1‐1 double mutants, whereas cells lost their jigsaw‐like shape in etg1‐1 atr‐2 plants (Figures 7L–N). The double mutants arrested at an early stage of development. These results unequivocally indicate that the activation of the DNA replication checkpoint in etg1 mutant plants is essential for their survival.
image
Figure 7. Activation of DNA replication stress checkpoint by ETG1 depletion. Seedling phenotypes of 21‐day‐old wild‐type (Col‐0) (A), etg1‐1 (B), wee1‐1 (C), atr‐2 (D), etg1‐1 wee1‐1 (E, F) and etg1‐1 atr‐2 (G, H) grown on MS plate. (F, H) Magnification of etg1‐1 wee1‐1 and etg1‐1 atr‐2 plants, respectively. Scanning electron micrographs of 14‐day‐old whole seedlings (IK), leaf epidermal cells (LN), and trichomes (OQ). (I, L, O) Wild‐type, (J, M, P), etg1‐1 wee1‐1, and (K, N, Q) etg1‐1 atr‐2. Bars=5 mm (A–E, G), 1 mm (F, H), 500 μm (I–K), 50 μm (L–N, P, Q), and 100 μm (O).

Discussion

ETG1 represents a novel component of the replisome

We have identified the ETG1 protein as an important replication factor. We propose that the ETG1 protein is a component of the replication fork that is necessary to maintain genome stability during S phase and that deficiency of ETG1 might generate DNA damage and genome instability (Figure 8). ETG1, originally discovered as a transcript induced in E2Fa‐DPa‐overexpressing plants, was demonstrated to be a direct target gene of the activating E2Fa/E2Fb transcription factors. Two E2F‐binding sites in the ETG1 promoter redundantly control ETG1 expression. Interestingly, not only is the ETG1 gene highly conserved among species but also E2F‐binding sites can be found in the promoter region of ETG1 orthologues of rice, Xenopus, Drosophila, mouse, and human. Moreover, in E2F1‐depleted cells, transcripts of the Drosophila ETG1 orthologue (CG3430) are decreased (Dimova et al, 2003). Thus, besides the gene product, also the transcriptional control mechanisms that drive ETG1 expression are conserved, suggesting that E2F‐dependent regulation of the ETG1 expression and of its orthologues is highly important during the replication process.
image
Figure 8. Model for ETG1 function. (A) Control of the ETG1 transcription by E2Fa and E2Fb transcription factors in late G1 phase. ETG1 protein assembles into the DNA replication complex and aids DNA replication during S phase. (B) ETG1 depletion causing DNA replication stress and activating the ATR kinase that induces the expression of DNA repair genes, such as RAD51 and PARP2, and the checkpoint kinase WEE1. WEE1 arrests cells in the G2 phase of the cell cycle.
Our data show that ETG1 is part of the replisome. First, etg1 mutants display a clear replication phenotype. Second, both ETG1 and its Drosophila orthologue (CG3430) associate with MCM proteins that bind to the replication origin and travel along the DNA with the remainder of the replication machinery (Kearsey and Labib, 1998; Tye, 1999; Labib et al, 2000). Moreover, while this work was in progress, the human orthologous ETG1 has been reported to bind chromatin in a cell cycle phase‐dependent manner with the highest affinity for DNA during G1/S and S (Sakwe et al, 2007). When the subcellular localization of MCM5 in plants was examined with an MCM5–eGFP fusion protein, MCM5–eGFP was found to reside in both the nucleus and cytoplasm. In yeast, the subcellular localization of the MCM5 protein changes with the cell cycle: it is nuclear between the end of M phase and G1‐to‐S transition, but is cytoplasmic in other phases of the cell cycle (Hennessy et al, 1991). Its dual localization in the nucleus and cytoplasm indicates that a similar situation might hold true for the Arabidopsis MCM5 protein. However, by BiFC experiments, we observed that MCM5 and ETG1 interacted only in nuclei and not in the cytoplasm, suggesting an interaction solely when the MCM proteins dock to the replication origins.

ETG1 deficiency activates a DNA replication checkpoint that accounts for survival

ETG1‐deficient plants do not only show a replication defect but also a prolonged G2 phase because of the activation of the DNA replication checkpoint, as illustrated by the upregulation of the DNA repair genes PARP2 and RAD51 and the cell cycle checkpoint gene WEE1. Despite the prolonged cell cycle duration, resulting in a strongly reduced total cell number, ETG1 single mutants had a relatively minor developmental phenotype. In contrast, in an atr or wee1 mutant background, ETG1 deficiency has a huge impact on plant growth with dwarfism, improper cell differentiation, and a developmental arrest as a consequence. These results illustrate that ATR and WEE1 make an essential contribution to the survival of plants without ETG1 function. Previously, lack of ATM or ATR activity has been demonstrated to emphasize the phenotypic defects of telomerase‐deficient (tert) mutants (Vespa et al, 2005). Whereas in a wild‐type background, the tert mutants display growth and developmental phenotypes from the sixth generation onwards, tert atm and tert atr double mutants have misshapen leaves, no apical dominance, delayed flowering, and significantly reduced fertility in earlier generations. These data indicate that the activation of the ATM and ATR pathways is required for survival of plants suffering genomic instability. However, it is impossible to attribute the attenuation of the phenotype by functional ATM and ATR to the induction of DNA repair genes, or the ability to arrest cell cycle progression. As wee1 mutant plants are defective in checkpoint activation but are still able to induce DNA repair genes (De Schutter et al, 2007; T Cools, DI, and LDV, unpublished data), the severe phenotype of the etg1 wee1 double mutants indicates that the activation of the G2 checkpoint contributes mainly to the survival process. Nevertheless, compared to the etg1 wee1 mutant, the observed growth defects were slightly more outspoken for the etg1 atr double mutant, illustrating that the phenotypes resulting from the inability to arrest the cell cycle are partly offset by DNA repair.

Endoreduplication might compensate for reductions in cell number caused by activation of the DNA stress checkpoint

Interestingly, at maturity, the population of cells with a high DNA ploidy level was larger in etg1 mutant than in control plants, suggesting that the replisome complex of endoreduplicating cells differs from that found in mitotically dividing cells and that ETG1 is not required for the endoreduplication cycle. Previously, DNA double‐stranded breaks caused by depletion of the chromatin assembly factor 1 have been found to be accompanied with extra endoreduplication cycles (Endo et al, 2006; Exner et al, 2006; Ramirez‐Parra and Gutierrez, 2007). Additionally, wild‐type plants treated with the DNA‐damaging drug zeocin significantly increased the DNA ploidy level (Ramirez‐Parra and Gutierrez, 2007). The mechanisms that link the maintenance of chromosomal stability with the DNA ploidy level under normal and stressed growth conditions remain unclear. Previously, we have found that the level of A‐type CDK activity determines whether a plant cell divides or endoreduplicates (Verkest et al, 2005). Therefore, the extra endoreduplication cycles under replication stress might be an indirect effect from the activation of the WEE1 expression, inactivating the A‐type CDK by tyrosine phosphorylation (De Schutter et al, 2007). The decrease in CDK activity might prevent entry into mitosis, but still allow S phase to be re‐initiated, and, thus, re‐replication of the genome. The size of cells has often been found to correlate with their DNA content (Melaragno et al, 1993; Folkers et al, 1997; Traas et al, 1998; Sugimoto‐Shirasu and Roberts, 2003). Therefore, in endoreduplicating species the coupling of checkpoint activation with initiation of the endocycle might represent a mechanism by which a reduction in cell number by replication stress is compensated by cell enlargement. Such a coupling might correspond with a survival mechanism allowing plants to achieve a critical biomass required for reproduction, possibly explaining the evolutionary success of the endoreduplication programme among angiosperms.

Materials and methods

Flow cytometric analyses

Leaves were chopped with a razor blade in 300 μl of buffer (45 mM MgCl2, 30 mM sodium citrate, 20 mM 3‐[N‐morpholino]propanesulphonic acid, pH 7, and 1% Triton X‐100). To the supernatants, 1 μl of 4,6‐diamidino‐2phenylindole (DAPI) from a stock of 1 mg/ml was added, which was filtered over a 30‐μm mesh. The nuclei were analysed with the CyFlow flow cytometer and the FloMax software (Partec, Münster, Germany). At least three biological and two technical replicates were used for each sample analysed.

Determination of the mitotic index

Roots were fixed in a solution of formaldehyde, ethanol, and acetic acid (2:17:1) for 12 h at 4°C, washed twice in water, and mounted under cover slips. The samples were crushed, snap‐frozen with liquid nitrogen to remove the cover slip, and mounted in Vectashield (Vector Laboratories) containing 1 μg/ml DAPI. The roots were analysed for mitotic stages with a Zeiss Axiovert fluorescence microscope.

TAP analysis

TAP experiments were carried out according to Van Leene et al (2007). In short, the ETG1‐coding sequence was cloned by recombination into the pKNTAP vector, generating a Pro35S:TAP‐ETG1 cassette (pKNTAPETG1). Arabidopsis cell suspension cultures were stably transformed by Agrobacterium‐mediated cocultivation with pKNTAPETG1. Transformed Arabidopsis cells were selected and transferred to a liquid medium for upscaling. Expression levels of TAP‐tagged proteins were checked by protein blotting with an anti‐calmodulin‐binding protein antibody (data not shown). In the first round of affinity purification, protein extracts of 15 g plant material were incubated with an IgG resin. Bound complexes were released and eluted from the resin by tag cleavage with TEV protease. In the second affinity step on a calmodulin agarose column, co‐eluting non‐interacting proteins and the TEV protease were removed with the flow‐through. Finally, both the ETG1 bait and interacting proteins were eluted from the calmodulin agarose through EGTA‐mediated removal of calcium. Eluted proteins were separated on 4–12% NuPAGE gels, excised, and analysed by MALDI‐TOF‐TOF‐MS as described (Van Leene et al, 2007). To increase the stringency of the data set, contaminating proteins due to experimental background as determined by Van Leene et al (2007) were systematically subtracted from the lists of co‐purified proteins.

BrdU ELISA

Seedlings (3 day old) grown on MS agar plates were incubated in the labelling solution with 10 μM BrdU (Roche) at room temperature for 3 h. After treatment, the genomic DNA was extracted with a DNeasy® Plant Mini Kit (Qiagen). The amount of incorporated BrdU was determined by ELISA with an anti‐BrdU‐POD antibody (Roche). Three biological and two technical replicates were used at each time point for ELISA. Here, 50 μl of the extracted DNA (0.2 μg/ml) was placed in each well. The ELISA procedure was chiefly that of the 5‐bromo‐2′‐deoxyuridine Labeling and Detection Kit III (Roche).

Supplementary data

Supplementary data are available at The EMBO Journal Online (http://www.embojournal.org).

Acknowledgements

We thank Tarik Gaamouche, Amandine Radziejwoski, Shinya Takahashi, Yoshihiro Hase, Eveline Van De Slijke, Jelle Van Leene, Geert Persiau, Ann Pharazyn, and Dominique Eeckhout for technical support, all members of the cell cycle group for fruitful discussions and suggestions, Dr Crisanto Gutierrez for the E2Fc antibody, the Arabidopsis Biological Research Center for providing T‐DNA insertion lines, and Martine De Cock for help in preparing the paper. This study was supported by grants from the European Union SY‐STEM (MRTN‐GT‐2004‐005336) and the Institute for the Promotion of Innovation through Science and Technology (‘Generisch Basisonderzoek aan de Universiteiten’ (GBOU) no. 20193 and 20176). TL and TY are indebted to the Institute for the Promotion of Innovation by Science and Technology in Flanders and the Ministry of Education, Science, Sports and Culture for a predoctoral fellowship and a Grant‐in‐Aid for Young Scientists ((B), 19770044), respectively. LDV is a postdoctoral fellow of the Research Foundation‐Flanders.

Supporting Information

References

Aparicio OM, Weinstein DM, Bell SP (1997) Components and dynamics of DNA replication complexes in S. cerevisiae: redistribution of MCM proteins and Cdc45p during S phase. Cell 91: 59–69
Babiychuk E, Cottrill PB, Storozhenko S, Fuangthong M, Chen Y, O'Farrell MK, Van Montagu M, Inzé D, Kushnir S (1998) Higher plants possess two structurally different poly(ADP‐ribose) polymerases. Plant J 15: 635–645
Bartkova J, Hořejší Z, Koed K, Krämer A, Tort F, Zieger K, Guldberg P, Sehested M, Nesland JM, Lukas C, ∅rntoft T, Lukas J, Bartek J (2005) DNA damage response as a candidate anti‐cancer barrier in early human tumorigenesis. Nature 434: 864–870
Beemster GTS, De Veylder L, Vercruysse S, West G, Rombaut D, Van Hummelen P, Galichet A, Gruissem W, Inzé D, Vuylsteke M (2005) Genome‐wide analysis of gene expression profiles associated with cell cycle transitions in growing organs of Arabidopsis. Plant Physiol 138: 734–743
Bell SP, Dutta A (2002) DNA replication in eukaryotic cells. Annu Rev Biochem 71: 333–374
Boudolf V, Barrôco R, de Almeida Engler J, Verkest A, Beeckman T, Naudts M, Inzé D, De Veylder L (2004) B1‐type cyclin‐dependent kinases are essential for the formation of stomatal complexes in Arabidopsis thaliana. Plant Cell 16: 945–955
Bracha‐Drori K, Shichrur K, Katz A, Oliva M, Angelovici R, Yalovsky S, Ohad N (2004) Detection of protein–protein interactions in plants using bimolecular fluorescence complementation. Plant J 40: 419–427
Chen I‐P, Haehnel U, Altschmied L, Schubert I, Puchta H (2003) The transcriptional response of Arabidopsis to genotoxic stress—a high‐density colony array study (HDCA). Plant J 35: 771–786
Culligan K, Tissier A, Britt A (2004) ATR regulates a G2‐phase cell‐cycle checkpoint in Arabidopsis thaliana. Plant Cell 16: 1091–1104
Culligan KM, Robertson CE, Foreman P, Doerner P, Britt AB (2006) ATR and ATM play both distinct and additive roles in response to ionizing radiation. Plant J 48: 947–961
De Schutter K, Joubès J, Cools T, Verkest A, Corellou F, Babiychuk E, Van Der Schueren E, Beeckman T, Kushnir S, Inzé D, De Veylder L (2007) Arabidopsis WEE1 kinase controls cell cycle arrest in response to activation of the DNA integrity checkpoint. Plant Cell 19: 211–225
De Veylder L, Beeckman T, Beemster GTS, de Almeida Engler J, Ormenese S, Maes S, Naudts M, Van Der Schueren E, Jacqmard A, Engler G, Inzé D (2002) Control of proliferation, endoreduplication and differentiation by the Arabidopsis E2Fa/DPa transcription factor. EMBO J 21: 1360–1368
De Veylder L, Beeckman T, Beemster GTS, Krols L, Terras F, Landrieu I, Van Der Schueren E, Maes S, Naudts M, Inzé D (2001) Functional analysis of cyclin‐dependent kinase inhibitors of Arabidopsis. Plant Cell 13: 1653–1667
del Pozo JC, Boniotti MB, Gutierrez C (2002) Arabidopsis E2Fc functions in cell division and is degraded by the ubiquitin‐SCFAtSKP2 pathway in response to light. Plant Cell 14: 3057–3071
Diffley JFX, Labib K (2002) The chromosome replication cycle. J Cell Sci 115: 869–872
Dimova DK, Dyson NJ (2005) The E2F transcriptional network: old acquaintances with new faces. Oncogene 24: 2810–2826
Dimova DK, Stevaux O, Frolov MV, Dyson NJ (2003) Cell cycle‐dependent and cell cycle‐independent control of transcription by the Drosophila E2F/RB pathway. Genes Dev 17: 2308–2320
Doucet‐Chabeaud G, Godon C, Brutesco C, de Murcia G, Kazmaier M (2001) Ionising radiation induces the expression of PARP‐1 and PARP‐2 genes in Arabidopsis. Mol Genet Genomics 265: 954–963
Endo M, Ishikawa Y, Osakabe K, Nakayama S, Kaya H, Araki T, Shibahara K, Abe K, Ichikawa H, Valentine L, Hohn B, Toki S (2006) Increased frequency of homologous recombination and T‐DNA integration in Arabidopsis CAF‐1 mutants. EMBO J 25: 5579–5590
Exner V, Taranto P, Schönrock N, Gruissem W, Hennig L (2006) Chromatin assembly factor CAF‐1 is required for cellular differentiation during plant development. Development 133: 4163–4172
Folkers U, Berger J, Hülskamp M (1997) Cell morphogenesis of trichomes in Arabidopsis: differential control of primary and secondary branching by branch initiation regulators and cell growth. Development 124: 3779–3786
Forsburg SL (2004) Eukaryotic MCM proteins: beyond replication initiation. Microbiol Mol Biol Rev 68: 109–131
Gambus A, Jones RC, Sanchez‐Diaz A, Kanemaki M, van Deursen F, Edmondson RD, Labib K (2006) GINS maintains association of Cdc45 with MCM in replisome progression complexes at eukaryotic DNA replication forks. Nat Cell Biol 8: 358–366
Garcia V, Bruchet H, Camescasse D, Granier F, Bouchez D, Tissier A (2003) AtATM is essential for meiosis and the somatic response to DNA damage in plants. Plant Cell 15: 119–132
Gillespie PJ, Li A, Blow JJ (2001) Reconstitution of licensed replication origins on Xenopus sperm nuclei using purified proteins. BMC Biochem 2: 15.1–15.11
Gorgoulis VG, Vassiliou L‐VF, Karakaidos P, Zacharatos P, Kotsinas A, Liloglou T, Venere M, DiTullio Jr RA, Kastrinakis NG, Levy B, Kletsas D, Yoneta A, Herlyn M, Kittas C, Halazonetis TD (2005) Activation of the DNA damage checkpoint and genomic instability in human precancerous lesions. Nature 434: 907–913
Hennessy KM, Clark CD, Botstein D (1991) Subcellular localization of yeast CDC46 varies with the cell cycle. Genes Dev 4: 2252–2263
Inzé D, De Veylder L (2006) Cell cycle regulation in plant development. Annu Rev Genet 40: 77–105
Kamimura Y, Tak Y‐S, Sugino A, Araki H (2001) Sld3, which interacts with Cdc45 (Sld4), functions for chromosomal DNA replication in Saccharomyces cerevisiae. EMBO J 20: 2097–2107
Kanemaki M, Labib K (2006) Distinct roles for Sld3 and GINS during establishment and progression of eukaryotic DNA replication forks. EMBO J 25: 1753–1763
Kearsey SE, Labib K (1998) MCM proteins: evolution, properties, and role in DNA replication. Biochim Biophys Acta 1398: 113–136
Kosugi S, Ohashi Y (2002) E2Ls, E2F‐like repressors of Arabidopsis that bind to E2F sites in a monomeric form. J Biol Chem 277: 16553–16558
Labib K, Diffley JFX (2001) Is the MCM2–7 complex the eukaryotic DNA replication fork helicase? Curr Opin Genet Dev 11: 64–70
Labib K, Tercero JA, Diffley JFX (2000) Uninterrupted MCM2–7 function required for DNA replication fork progression. Science 288: 1643–1647
Magyar Z, De Veylder L, Atanassova A, Bakó L, Inzé D, Bögre L (2005) The role of the Arabidopsis E2FB transcription factor in regulating auxin‐dependent cell division. Plant Cell 17: 2527–2541
Masumoto H, Muramatsu S, Kamimura Y, Araki H (2002) S‐Cdk‐dependent phosphorylation of Sld2 essential for chromosomal DNA replication in budding yeast. Nature 415: 651–655
Melaragno JE, Mehrotra B, Coleman AW (1993) Relationship between endopolyploidy and cell size in epidermal tissue of Arabidopsis. Plant Cell 5: 1661–1668
Obayashi T, Kinoshita K, Nakai K, Shibaoka M, Hayashi S, Saeki M, Shibata D, Saito K, Ohta H (2007) ATTED‐II: a database of co‐expressed genes and cis elements for identifying co‐regulated gene groups in Arabidopsis. Nucleic Acids Res 35: D863–D869
Ramirez‐Parra E, Fründt C, Gutierrez C (2003) A genome‐wide identification of E2F‐regulated genes in Arabidopsis. Plant J 33: 801–811
Ramirez‐Parra E, Gutierrez C (2007) E2F regulates FASCIATA1, a chromatin assembly gene whose loss switches on the endocycle and activates gene expression by changing the epigenetic status. Plant Physiol 144: 105–120
Ramirez‐Parra E, López‐Matas MA, Fründt C, Gutierrez C (2004) Role of an atypical E2F transcription factor in the control of Arabidopsis cell growth and differentiation. Plant Cell 16: 2350–2363
Rossignol P, Stevens R, Perennes C, Jasinski S, Cella R, Tremousaygue D, Bergounioux C (2002) AtE2F‐a and AtDP‐a, members of the E2F family of transcription factors, induce Arabidopsis leaf cells to re‐enter S phase. Mol Genet Genomics 266: 995–1003
Sakwe AM, Nguyen T, Athanasopoulos V, Shire K, Frappier L (2007) Identification and characterization of a novel component of the human minichromosome maintenance complex. Mol Cell Biol 27: 3044–3055
Stuart JM, Segal E, Koller D, Kim SK (2003) A gene‐coexpression network for global discovery of conserved genetic modules. Science 302: 249–255
Sugimoto‐Shirasu K, Roberts K (2003) ‘Big it up’: endoreduplication and cell‐size control in plants. Curr Opin Plant Biol 6: 544–553
Takayama Y, Kamimura Y, Okawa M, Muramatsu S, Sugino A, Araki H (2003) GINS, a novel multiprotein complex required for chromosomal DNA replication in budding yeast. Genes Dev 17: 1153–1165
Traas J, Hülskamp M, Gendreau E, Höfte H (1998) Endoreduplication and development: rule without dividing? Curr Opin Plant Biol 1: 498–503
Tye BK (1999) MCM proteins in DNA replication. Annu Rev Biochem 68: 649–686
Tye BK, Sawyer S (2000) The hexameric eukaryotic MCM helicase: building symmetry from nonidentical parts. J Biol Chem 275: 34833–34836
Ulm R, Baumann A, Oravecz A, Máté Z, Ádám É, Oakeley EJ, Schäfer E, Nagy F (2004) Genome‐wide analysis of gene expression reveals function of the bZIP transcription factor HY5 in the UV‐B response of Arabidopsis. Proc Natl Acad Sci USA 101: 1397–1402
Van Leene J, Eeckhout D, Stals H, Persiau G, Van De Slijke E, Van Isterdael G, Laukens K, Remmerie N, Abdelkrim A, Pharazyn A, Van Onckelen H, Inzé D, Witters E, De Jaeger G (2007) Tandem affinity purification of cell cycle protein complexes from Arabidopsis cell suspension cultures. Mol Cell Proteomics 6: 1226–1238
Vandepoele K, Vlieghe K, Florquin K, Hennig L, Beemster GTS, Gruissem W, Van de Peer Y, Inzé D, De Veylder L (2005) Genome‐wide identification of potential plant E2F target genes. Plant Physiol 139: 316–328
Verkest A, de O Manes C‐L, Vercruysse S, Maes S, Van Der Schueren E, Beeckman T, Genschik P, Kuiper M, Inzé D, De Veylder L (2005) The cyclin‐dependent kinase inhibitor KRP2 controls the onset of the endoreduplication cycle during Arabidopsis leaf development through inhibition of mitotic CDKA;1 kinase complexes. Plant Cell 17: 1723–1736
Vespa L, Couvillion M, Spangler E, Shippen DE (2005) ATM and ATR make distinct contributions to chromosome end protection and the maintenance of telomeric DNA in Arabidopsis. Genes Dev 19: 2111–2115
Vlieghe K, Boudolf V, Beemster GTS, Maes S, Magyar Z, Atanassova A, de Almeida Engler J, De Groodt R, Inzé D, De Veylder L (2005) The DP‐E2F‐like DEL1 gene controls the endocycle in Arabidopsis thaliana. Curr Biol 15: 59–63
Walter M, Chaban C, Schütze K, Batistic O, Weckermann K, Näke C, Blazevic D, Grefen C, Schumacher K, Oecking C, Harter K, Kudla J (2004) Visualization of protein interactions in living plant cells using bimolecular fluorescence complementation. Plant J 40: 428–438
Wei H, Persson S, Mehta T, Srinivasasainagendra V, Chen L, Page GP, Somerville C, Loraine A (2006) Transcriptional coordination of the metabolic network in Arabidopsis. Plant Physiol 142: 762–774
Zou L, Stillman B (2000) Assembly of a complex containing Cdc45p, replication protein A, and Mcm2p at replication origins controlled by S‐phase cyclin‐dependent kinases and Cdc7p–Dbf4p kinase. Mol Cell Biol 20: 3086–3096

Information & Authors

Information

Published In

The EMBO Journal cover image
Read More
The EMBO Journal
Vol. 27 | No. 13
9 July 2008
Table of contents
Pages: 1840 - 1851

Submission history

Received: 13 February 2008
Accepted: 7 May 2008
Published online: 5 June 2008
Published in issue: 9 July 2008

Permissions

Request permissions for this article.

Keywords

  1. Arabidopsis
  2. DNA replication
  3. DNA replication checkpoint
  4. ETG1
  5. MCM proteins

Authors

Affiliations

Naoki Takahashi
Department of Plant Systems Biology, Flanders Institute for Biotechnology Gent Belgium
Department of Molecular Genetics, Ghent University Gent Belgium
Tim Lammens
Department of Plant Systems Biology, Flanders Institute for Biotechnology Gent Belgium
Department of Molecular Genetics, Ghent University Gent Belgium
Véronique Boudolf
Department of Plant Systems Biology, Flanders Institute for Biotechnology Gent Belgium
Department of Molecular Genetics, Ghent University Gent Belgium
Sara Maes
Department of Plant Systems Biology, Flanders Institute for Biotechnology Gent Belgium
Department of Molecular Genetics, Ghent University Gent Belgium
Takeshi Yoshizumi
Plant Functional Genomics Research Group, RIKEN Plant Science Center Yokohama Japan
Geert De Jaeger
Department of Plant Systems Biology, Flanders Institute for Biotechnology Gent Belgium
Department of Molecular Genetics, Ghent University Gent Belgium
Erwin Witters
Department of Chemistry, Centre for Proteome Analysis and Mass Spectrometry, University of Antwerp Antwerpen Belgium
Flemish Institute for Technological Research (VITO) Mol Belgium
Dirk Inzé
Department of Plant Systems Biology, Flanders Institute for Biotechnology Gent Belgium
Department of Molecular Genetics, Ghent University Gent Belgium
Lieven De Veylder [email protected]
Department of Plant Systems Biology, Flanders Institute for Biotechnology Gent Belgium
Department of Molecular Genetics, Ghent University Gent Belgium
Corresponding author. Department of Plant Systems Biology, VIB, Ghent University, Technologiepark 927, B‐9052 Gent, Belgium. Tel.: +32 9 3313800; Fax +32 9 3313809; E-mail: [email protected]

Metrics & Citations

Metrics

Citations

Download Citations

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Select your manager software from the list below and click Download.

Citing Literature

  • Comparison of the Effects of UV-C Light in the Form of Flash or Continuous Exposure: A Transcriptomic Analysis on Arabidopsis thaliana L., International Journal of Molecular Sciences, 10.3390/ijms252413718, 25, 24, (13718), (2024).
  • A Role for USP7 in DNA Replication, Molecular and Cellular Biology, 10.1128/MCB.00639-13, 34, 1, (132-145), (2023).
  • Leaf-size control beyond transcription factors: Compensatory mechanisms, Frontiers in Plant Science, 10.3389/fpls.2022.1024945, 13, (2023).
  • SNF1 ‐related protein kinase 1 represses Arabidopsis growth through post‐translational modification of E2Fa in response to energy stress , New Phytologist, 10.1111/nph.18597, 237, 3, (823-839), (2022).
  • A negative feedback loop of TOR signaling balances growth and stress-response trade-offs in plants, Cell Reports, 10.1016/j.celrep.2022.110631, 39, 1, (110631), (2022).
  • Cell cycle checkpoint control in response to DNA damage by environmental stresses, The Plant Journal, 10.1111/tpj.15567, 109, 3, (490-507), (2021).
  • Arabidopsis Target of Rapamycin Coordinates With Transcriptional and Epigenetic Machinery to Regulate Thermotolerance, Frontiers in Plant Science, 10.3389/fpls.2021.741965, 12, (2021).
  • RETRACTED ARTICLE: SOG1 transcription factor promotes the onset of endoreduplication under salinity stress in Arabidopsis, Scientific Reports, 10.1038/s41598-021-91293-1, 11, 1, (2021).
  • Overexpression Populus d-Type Cyclin Gene PsnCYCD1;1 Influences Cell Division and Produces Curved Leaf in Arabidopsis thaliana, International Journal of Molecular Sciences, 10.3390/ijms22115837, 22, 11, (5837), (2021).
  • The Arabidopsis GRAS-type SCL28 transcription factor controls the mitotic cell cycle and division plane orientation , Proceedings of the National Academy of Sciences, 10.1073/pnas.2005256118, 118, 6, (2021).

View Options

View options

PDF

View PDF

Figures

Tables

Media

Share

Share

Copy the content Link

Share on social media