15
ATR Regulates a G2-Phase Cell-Cycle Checkpoint in Arabidopsis thaliana Kevin Culligan, a,1 Alain Tissier, b and Anne Britt a a Section of Plant Biology, University of California, Davis, California 95616 b Librophyt SAS, Centre de Cadarache, 13108 St. Paul-lez-Durance Cedex, France Ataxia telangiectasia-mutated and Rad3-related (ATR) plays a central role in cell-cycle regulation, transmitting DNA damage signals to downstream effectors of cell-cycle progression. In animals, ATR is an essential gene. Here, we find that Arabidopsis (Arabidopsis thaliana) atrÿ/ÿ mutants were viable, fertile, and phenotypically wild-type in the absence of exogenous DNA damaging agents but exhibit altered expression of AtRNR1 (ribonucleotide reductase large subunit) and alteration of some damage-induced cell-cycle checkpoints. atr mutants were hypersensitive to hydroxyurea (HU), aphidicolin, and UV-B light but only mildly sensitive to g-radiation. G2 arrest was observed in response to g-irradiation in both wild-type and atr plants, albeit with slightly different kinetics, suggesting that ATR plays a secondary role in response to double-strand breaks. G2 arrest also was observed in wild-type plants in response to aphidicolin but was defective in atr mutants, resulting in compaction of nuclei and subsequent cell death. By contrast, HU-treated wild-type and atr plants arrested in G1 and showed no obvious signs of cell death. We propose that, in plants, HU invokes a novel checkpoint responsive to low levels of deoxynucleotide triphosphates. These results demonstrate the important role of cell-cycle checkpoints in the ability of plant cells to sense and cope with problems associated with DNA replication. INTRODUCTION The detection of DNA damage is an important aspect of damage resistance. Cells do not routinely express all DNA repair mechanisms, but instead upregulate these functions as required, concomitantly slowing the cell cycle to permit time for repair before S-phase (as replication past damaged sites may be mutagenic) or before M-phase (after which point double-strand breaks and lesions in daughter strand gaps may become far more problematic to repair). Damage is generally detected by protein complexes in the form of single-stranded DNA (Crowley and Courcelle, 2002; Zou and Elledge, 2003), created when the replication fork is physically blocked by DNA lesions or during the process of DNA repair (e.g., homologous recombination, nucleotide excision repair, and mismatch repair). These damage sensors then activate downstream checkpoint mechanisms to prevent progression into the next phase of the cell cycle as well as upregulate DNA repair (Nyberg et al., 2002). Some damage sensors also are involved in the regulation of deoxynucleotide triphosphate (dNTP) pools, which may in turn influence the mode of DNA replication. One of the primary targets for dNTP pool regulation is the enzyme ribonucleotide reductase (RNR), which catalyzes the final reduction step in the production of dNTPs. Genetic studies in Saccharomyces cerevisiae suggest that the elevation of dNTP levels (through transcriptional and/or posttranslational regulation of RNR) in response to DNA damage may result in more efficient translesion DNA synthesis, enabling the cell to complete replication in spite of the presence of persisting lesions (Tanaka et al., 2000; Chabes et al., 2003; Yao et al., 2003). The proteins that sense damaged or single-stranded DNAs and trigger these elaborate responses have been extensively characterized in organisms ranging from bacteria to human (Crowley and Courcelle, 2002; Melo and Toczyski, 2002). In plants, however, the regulation of DNA replication and repair and the cell cycle in response to DNA damage is only beginning to be understood. The Schizosaccharomyces pombe Rad3 homologs (ataxia telangiectasia-mutated and Rad3-related [ATR] and ataxia telangiectasia-mutated [ATM] in mammals, Mec1 in S. cerevi- siae) are essential regulators of cell-cycle checkpoints, sensing DNA damage and/or single-stranded DNA and activating downstream effectors of cell-cycle progression and DNA repair. ATM is activated primarily by DNA double-strand breaks. ATM- deficient plants and animals are hypersensitive to g-irradiation but not to replication-blocking agents such as UV-B light, hydroxyurea (HU), or aphidicolin (Abraham, 2001; Garcia et al., 2003; K. Culligan and A. Britt, unpublished data). Furthermore, the Arabidopsis (Arabidopsis thaliana) ATM homolog is required for the transcriptional induction of repair genes in response to g-irradiation and for efficient progression through meiosis (Garcia et al., 2003), though its role in plant cell-cycle responses to DNA damage has not yet been determined. In animals, activated ATM phoshorylates downstream components of checkpoint pathways that include p53, BRCA1, NBS1, and CHK2, initiating G1-, S-, or G2-phase arrest and/or apoptosis. By contrast, ATR is activated primarily by agents that block the progression of replication forks. ATR dominant-negative 1 To whom correspondence should be addressed. E-mail kmculligan@ ucdavis.edu; fax 530-752-5410. The author responsible for distribution of materials integral to the findings presented in this article in accordance with the policy described in the Instructions for Authors (www.plantcell.org) is: Kevin M. Culligan ([email protected]). Article, publication date, and citation information can be found at www.plantcell.org/cgi/doi/10.1105/tpc.018903. This article is published in The Plant Cell Online, The Plant Cell Preview Section, which publishes manuscripts accepted for publication after they have been edited and the authors have corrected proofs, but before the final, complete issue is published online. Early posting of articles reduces normal time to publication by several weeks. The Plant Cell Preview, www.aspb.org ª 2004 American Society of Plant Biologists 1 of 14

ATR Regulates a G2Phase Cell-Cycle Checkpoint in Arabidopsis thaliana

Embed Size (px)

Citation preview

ATR Regulates a G2-Phase Cell-Cycle Checkpointin Arabidopsis thaliana

Kevin Culligan,a,1 Alain Tissier,b and Anne Britta

a Section of Plant Biology, University of California, Davis, California 95616b Librophyt SAS, Centre de Cadarache, 13108 St. Paul-lez-Durance Cedex, France

Ataxia telangiectasia-mutated and Rad3-related (ATR) plays a central role in cell-cycle regulation, transmitting DNA damage

signals to downstream effectors of cell-cycle progression. In animals, ATR is an essential gene. Here, we find that

Arabidopsis (Arabidopsis thaliana) atr�/� mutants were viable, fertile, and phenotypically wild-type in the absence of

exogenous DNA damaging agents but exhibit altered expression of AtRNR1 (ribonucleotide reductase large subunit) and

alteration of some damage-induced cell-cycle checkpoints. atr mutants were hypersensitive to hydroxyurea (HU),

aphidicolin, and UV-B light but only mildly sensitive to g-radiation. G2 arrest was observed in response to g-irradiation in

both wild-type and atr plants, albeit with slightly different kinetics, suggesting that ATR plays a secondary role in response

to double-strand breaks. G2 arrest also was observed in wild-type plants in response to aphidicolin but was defective in atr

mutants, resulting in compaction of nuclei and subsequent cell death. By contrast, HU-treated wild-type and atr plants

arrested in G1 and showed no obvious signs of cell death. We propose that, in plants, HU invokes a novel checkpoint

responsive to low levels of deoxynucleotide triphosphates. These results demonstrate the important role of cell-cycle

checkpoints in the ability of plant cells to sense and cope with problems associated with DNA replication.

INTRODUCTION

The detection of DNA damage is an important aspect of damage

resistance. Cells do not routinely express all DNA repair

mechanisms, but instead upregulate these functions as required,

concomitantly slowing the cell cycle to permit time for repair

before S-phase (as replication past damaged sites may be

mutagenic) or before M-phase (after which point double-strand

breaks and lesions in daughter strand gaps may become far

more problematic to repair). Damage is generally detected by

protein complexes in the form of single-stranded DNA (Crowley

and Courcelle, 2002; Zou and Elledge, 2003), created when the

replication fork is physically blocked by DNA lesions or during

the process of DNA repair (e.g., homologous recombination,

nucleotide excision repair, and mismatch repair). These damage

sensors then activate downstream checkpoint mechanisms to

prevent progression into the next phase of the cell cycle as well

as upregulate DNA repair (Nyberg et al., 2002).

Some damage sensors also are involved in the regulation of

deoxynucleotide triphosphate (dNTP) pools, which may in turn

influence themode of DNA replication. One of the primary targets

for dNTP pool regulation is the enzyme ribonucleotide reductase

(RNR), which catalyzes the final reduction step in the production

of dNTPs. Genetic studies in Saccharomyces cerevisiae suggest

that the elevation of dNTP levels (through transcriptional and/or

posttranslational regulation of RNR) in response to DNA damage

may result in more efficient translesion DNA synthesis, enabling

the cell to complete replication in spite of the presence of

persisting lesions (Tanaka et al., 2000; Chabes et al., 2003; Yao

et al., 2003).

The proteins that sense damaged or single-stranded DNAs

and trigger these elaborate responses have been extensively

characterized in organisms ranging from bacteria to human

(Crowley and Courcelle, 2002; Melo and Toczyski, 2002). In

plants, however, the regulation of DNA replication and repair and

the cell cycle in response to DNA damage is only beginning to be

understood.

The Schizosaccharomyces pombe Rad3 homologs (ataxia

telangiectasia-mutated and Rad3-related [ATR] and ataxia

telangiectasia-mutated [ATM] in mammals, Mec1 in S. cerevi-

siae) are essential regulators of cell-cycle checkpoints, sensing

DNA damage and/or single-stranded DNA and activating

downstream effectors of cell-cycle progression and DNA repair.

ATM is activated primarily by DNA double-strand breaks. ATM-

deficient plants and animals are hypersensitive to g-irradiation

but not to replication-blocking agents such as UV-B light,

hydroxyurea (HU), or aphidicolin (Abraham, 2001; Garcia et al.,

2003; K. Culligan and A. Britt, unpublished data). Furthermore,

the Arabidopsis (Arabidopsis thaliana) ATM homolog is required

for the transcriptional induction of repair genes in response to

g-irradiation and for efficient progression through meiosis

(Garcia et al., 2003), though its role in plant cell-cycle responses

to DNA damage has not yet been determined. In animals,

activated ATM phoshorylates downstream components of

checkpoint pathways that include p53, BRCA1, NBS1, and

CHK2, initiating G1-, S-, or G2-phase arrest and/or apoptosis.

By contrast, ATR is activated primarily by agents that block

the progression of replication forks. ATR dominant-negative

1 To whom correspondence should be addressed. E-mail [email protected]; fax 530-752-5410.The author responsible for distribution of materials integral to thefindings presented in this article in accordance with the policy describedin the Instructions for Authors (www.plantcell.org) is: Kevin M. Culligan([email protected]).Article, publication date, and citation information can be found atwww.plantcell.org/cgi/doi/10.1105/tpc.018903.

This article is published in The Plant Cell Online, The Plant Cell Preview Section, which publishes manuscripts accepted for publication after they

have been edited and the authors have corrected proofs, but before the final, complete issue is published online. Early posting of articles reduces

normal time to publication by several weeks.

The Plant Cell Preview, www.aspb.orgª 2004 American Society of Plant Biologists 1 of 14

and conditional knockout mammalian cell lines display hyper-

sensitivity to UV-B light, HU, and aphidicolin but are also sen-

sitive to g-radiation. Thus, in animal cells, ATR is thought to

play a more generalized role in response to DNA damage than

ATM. Activated ATR phosphorylates CHK1 to initiate G2-phase

arrest but may also act at other points of the cell cycle, as

evidenced by ATR-dependent phosphorylation of BRCA1

(Tibbetts et al., 2000; Melo and Toczyski, 2002). However, the

determination of the full extent of ATR’s role in genome main-

tenance has been complicated by the fact that ATR homozygous

knockouts die early in embryogenesis (Brown and Baltimore,

2000; de Klein et al., 2000); cre/lox-induced nulls survive only

a few rounds of cell division, and the dominant-negative re-

duction-of-function cell lines have yielded conflicting results

(Cliby et al., 1998; Wright et al., 1998). Thus, a tractable genetic

system inwhich a null allele of ATR could be tolerated throughout

the development of a higher organism would be particularly

desirable.

Here, we investigate the effects of an ATR T-DNA insertion

allele in the plant Arabidopsis. The allele has a 6-kb insertion

within the kinase domain and is presumably null. We find that

the mutant plant (atr) is viable and developmentally normal in

the absence of exogenous DNA-damaging treatments but is

hypersensitive to the replication-blocking agents UV-B light,

HU, and aphidicolin. Thus, AtATR does appear to act as a

functional homolog of its mammalian counterpart. However, the

Arabidopsis atr mutants differ from mammalian mutants not

only in terms of their viability but also in their degree of sensitivity

to g-radiation.

In the course of these experiments, we also observed

a differential response to HU versus aphidicolin in wild-type

cells, with HU primarily inducing aG1 arrest, whereas aphidicolin

induced arrest in G2. This stands in contrast with the G2 arrest

induced by both agents in yeast and mammals (at concen-

trations that allowS-phase progression) and suggests that plants

may possess a novel G1 checkpoint response to low dNTP

levels. The G2 arrest in response to aphidicolin was abrogated in

atr mutants, suggesting that ATR regulates the G2 checkpoint

response to replication blocks, as it does inmammals.Moreover,

programmed cell death was observed in atrmutants in response

to aphidicolin, but not HU, under our experimental conditions.

We propose a model that suggests that the HU and aphidicolin

hypersensitivities of atrmutants are the result of its defect in the

G2 checkpoint response to blocked replication forks.

RESULTS

Identification of the ATROrtholog in Arabidopsis

ATR orthologs, including Mec1 of S. cerevisiae and Rad3 of S.

pombe, are members of a large gene family in eukaryotes that

encode Ser-Thr kinases, whose C-terminal catalytic domains

share similarity to yeast and mammalian phosphoinositide

3-kinases (Tibbetts and Abraham, 2000). We searched the

Arabidopsis genome for homologs of ATR, using the S. pombe

Rad3 C-terminal domain as a query sequence, and found

several candidate genes. Phylogenetic analysis of the corre-

sponding predicted full-length protein sequences indicated that

one was an ortholog of Target of Rapamycin (TOR), one an

ortholog of ATM, and one was an ortholog of ATR (Figure 1A).

No other candidate genes fell into these broad categories. The

ATR-orthologous sequence (previously reported in GenBank as

mRNA AtRad3 [AB040133], which we designate here as AtATR)

revealed high similarity throughout its entire length to other ATR

proteins, but its similarity to ATM and TOR sequences was

limited only to the highly conserved middle and C-terminal

regions. We confirmed the GenBank sequence and intron/exon

boundaries by isolating and sequencing cDNAs of AtATR

(�8.1 kb, ecotype Wassilewskija [Ws]) using RT-PCR, and the

resulting sequence predicted 17 exons encoding a protein of

2702 amino acids (Figure 1B).

Identification of T-DNA Insertion-Mutation Alleles

in AtATR

To study the function of ATR in Arabidopsis, we searched for

lines with T-DNA insertions within ATR. Three independent lines

were identified from different collections and were termed atr-1

(Arabidopsis Knockout Facility, ecotypeWs), atr-2 (SIGnAL data-

base, ecotype Columbia), and atr-3 (FLAG database, ecotype

Ws). atr-1 and atr-2 contain insertions in the middle of ATR, in

exons 7 and 10, respectively (Figure 1B). atr-3 contains a single,

�6-kb T-DNA insertion within the highly conserved C-terminal

kinase (PI3Kc) domain (Figures 1B to 1D), which causes a 30-bp

deletion adjacent to the catalyticAsp required for kinase function.

RT-PCR analysis of the C-terminal region of ATR confirms

that the transcript is absent in atr-3 (Figure 7). Thus, we believe

atr-3 is likely to be a null mutation and have focused the major-

ity of our characterization on this line (termed atr below).

Nonetheless, all three lines exhibit a similar recessive phenotype

(see below) that cosegregates with their respective T-DNA

insertions in ATR.

Neither AtATR nor AtATM Are Required during Normal

Somatic Development

All three atr homozygous alleles are phenotypically wild-type in

growth (of the root and shoot) and development (leaf and flower

development, seed set and viability) under standard conditions.

Thus, ATR is not essential for normal somatic development in

Arabidopsis. By contrast, ATR is an essential gene in animals;

ATR knockout mice die early in embryogenesis (Brown and

Baltimore, 2000; de Klein et al., 2000), whereas conditional

knockout human cell lines divide only a few times before dying

(Cortez et al., 2001; Brown and Baltimore, 2003). Because ATR

and ATM are paralogs known to play partially overlapping roles

in animal cells (Abraham, 2001), it is possible that ATM could

compensate for ATR deficiency in Arabidopsis. To test this

hypothesis, we established atr atm double mutants and found

no obvious phenotypic differences (in vegetative growth and

development) when compared with wild-type or single-mutant

plants, indicating that neither ATR nor ATM play an essential role

during normal vegetative (somatic) growth.

The double mutants, however, were completely sterile (Figure

2). atr mutant plants are fertile and produce normal gametes;

pollen staining for viable spores revealed no differences versus

2 of 14 The Plant Cell

Figure 1. Phylogenetic Analysis, Structure, and T-DNA Insertions of AtATR.

(A) The neighbor-joining tree of Rad3-like protein sequences was performed as described (Culligan et al., 2000). Numbers above each branch represent

the number of times the branch was found in 1000 bootstrap replicas. Arabidopsis sequences are in bold. No DNA-PKcs (DNA–protein kinase catalytic

subunit)-like sequences were identified in the Arabidopsis genome.

(B) Overview of the AtATR gene, positions of atr-1, atr-2, and atr-3 T-DNA insertions, and its corresponding cDNA structure. Numbered gray boxes

indicate exons of AtATR; the positions of the T-DNA insertions are denoted as triangles. The cDNA structure (bottom) shows the corresponding highly

conserved domains found in ATM, ATR, TOR, and FKBP12-rapamycin-associated protein sequences, including the kinase domain denoted as a black

rectangle (PI3Kc). The cDNA sequence was confirmed by isolating and sequencing independent cDNAs of AtATR (ecotype Ws) using RT-PCR. The

resulting sequence predicts 17 exons that encode a protein of 2702 amino acids.

(C) Structure of the atr-3 T-DNA insertion. Capital letters denote T-DNA insertion sequence. Corresponding amino acids are shown above the DNA

sequence. The left-border sequence encodes 10 new amino acids within the AtATR reading frame before a premature stop (boxed as TGA). The

approximate size of the entire insertion is 6 kb.

(D) Partial alignment of the kinase domain where atr-3 is inserted: 1 denotes a 30-bp deletion (equivalent to 10 amino acids) caused by the insertion, and

2 denotes the position of the catalytic Asp required for kinase function.

AtATR Regulates a G2 Checkpoint 3 of 14

the wild type, and the siliques were full (normal seed set). atm

mutant plants, however, are partially sterile because of abundant

chromosomal fragmentation during meiosis, resulting in a re-

duction in viable pollen and reduced seed set (Garcia et al.,

2003). We suggest that AtATM and AtATR play partially redun-

dant roles during meiosis in Arabidopsis.

AtATR Is Required for Full Transcriptional Induction of the

Large Subunit of RNR, AtRNR1

In the budding yeast S. cerevisiae, Mec1 (the ATR ortholog)

regulates expression of RNR at both the transcriptional and

posttranscriptional levels (Zhao et al., 2001). Null alleles ofmec1

are lethal, but this lethality is suppressed by mutations in genes

that negatively regulate RNR (Huang et al., 1998; Zhao et al.,

1998), such as the transcriptional repressor Crt1 (Huang et al.,

1998) or the inhibitor of RNR activity Sml1 (Zhao et al., 1998).

Derepression of RNR transcription in response to DNA damage

is dependent on Mec1 and the downstream kinase Dun1, which

indirectly induces an inhibitory phosphorylation of Crt1 (Elledge

et al., 1992; Zhao and Rothstein, 2002). Whether Mec1 or ATR

also play a role in the normal cell-cycle specific regulation ofRNR

transcription in unchallenged cells remains unclear.

We therefore wanted to determine if RNR transcription is

regulated in an AtATR-dependent manner in challenged and/or

unchallenged plants. We searched GenBank and identified

the Arabidopsis homolog of RNR1, the large subunit of RNR

(accession number AF092841). Transcription of RNR1 has

already been shown to be regulated in a cell-cycle dependent

manner in tobacco (Nicotiana tabacum) (Chaboute et al., 1998,

2002), and transcription is induced by DNA damage in this plant

(M.E. Chaboute, personal communication). We probed RNA gel

blots with AtRNR1 and found that AtRNR1 transcript is reduced

in atr mutants, down �40% versus the wild type under

standard conditions (Figure 3), suggesting that ATR is required

to support normal expression of RNR1. Plants treated with

chronic doses of UV-B light, sufficient to inhibit but not eliminate

growth, display a level of RNR1 transcript similar to that

observed under standard (UV-free) conditions, and again atr

plants exhibit reduced levels of RNR1 transcript. Interestingly,

plants treated with acute doses of UV-B light displayed

a transient inhibition of AtRNR1 transcript (at 2 h postirradia-

tion), followed by a return to standard condition levels at 6 h

postirradiation. This UV-induced suppression is not ATR de-

pendent, though again, the overall level of RNR1 transcript is

reduced (Figure 3).

Because we used HU (an inhibitor of RNR activity) as a DNA

replication-blocking agent (seebelow) to characterize thepheno-

typic effects of the atr mutant, we wanted to determine the

effects of HU treatment on RNR expression in whole plants.

Surprisingly, in the presence of HU,RNR1 expression is the same

in both the wild type and atr (Figure 3) with levels similar to the

wild type under standard conditions. This suggests that there

are additional ATR-independent levels of RNR1 regulation in

Arabidopsis and is consistent with previous studies showing that

HU induces RNR expression in plants (Chaboute et al., 1998).

Figure 2. Siliques Harvested from Mature Wild-Type, atr, and atm-1 (Ecotype Ws) Mutants.

Wild-type, atr�/�, and atr�/� atm1/� on average produce �50 seeds per silique. atm plants, which are partially sterile, produce a range of silique

sizes with much fewer (<10 on average) seeds. The siliques from the double atr atm line produced no seeds and were unable to outcross as males or

females, suggesting complete sterility.

4 of 14 The Plant Cell

Arabidopsis atrMutants Are Mildly Sensitive to

Acute Doses of g-Radiation but Proficient

in g-Induced G2 Arrest

Conditional mutants of ATR/Mec1 in mammalian cells and S.

cerevisiae, as well as Rad3 mutants in S. pombe, are

hypersensitive to g-radiation (Jimenez et al., 1992; Cliby et al.,

1998; Wright et al., 1998). Mutants of ATM/scTel1, including

those in Arabidopsis, are also hypersensitive to g-radiation

(Sedgwick and Boder, 1991; Garcia et al., 2003). This sensitivity

is generally interpreted as a failure of checkpoint responses to

double-strand breaks (Abraham, 2001) but may also involve

a failure to induce DNA repair genes (Garcia et al., 2003). To

determine whether ATR plays a role in g-resistance in Arabi-

dopsis, we challengedwild-type, atr, and atm lines with sublethal

doses of g-radiation. In contrast with atm mutants, atr lines

exhibited onlymild sensitivity to g-radiation. The root length of atr

was reduced by�30% versus the wild type 7 d after seeds were

g-irradiated at 200 Gray (Gy), whereas atm root length was

reduced by >95% (Figure 4A).

To determine the role of Arabidopsis ATR as a cell-cycle

checkpoint regulator in response to g-irradiation, we established

homozygous wild-type and atr mutant lines carrying a cyclinB1

promoter coupled to b-glucuronidase (PcyclinB1:GUS) reporter

construct. This construct harbors a GUS gene fused to a mitotic

destruction sequence and a cyclinB1 promoter (Colon-Carmona

et al., 1999). The GUS fusion gene is expressed upon entry into

G2 (via the cyclinB1 promoter), and the fusion protein is

degraded just before anaphase during mitosis. GUS-positive

cells stain blue in G2/early M, allowing us to visually monitor the

progression of the cell cycle in root meristems (Preuss and Britt,

2003). To determine G2-specific arrest (versus post-G2 mitotic

arrest), we further stained root tips with 49,6-diamidino-2-

phenylindole (DAPI) to visualize mitotic nuclei. In response to

400 Gy of g-radiation, atrmutants showed only a slightly altered

cell-cycle response when compared with the wild type; both

wild-type and atr meristematic cells arrest in G2 (DAPI staining

revealed no accumulation of mitotic figures) but differ slightly in

the timing of the response. For instance, at 8 h postirradiation, atr

root tips accumulated more G2 cells versus the wild type, but at

1 d postirradiation the root tips accumulated less G2 cells ver-

sus the wild type (Figure 4B). This suggests that Arabidopsis

ATR is not absolutely required for the G2 checkpoint response to

acute doses of g-radiation but does influence its rate of induction

and persistence.

Arabidopsis atrMutants Are Hypersensitive to

Replication-Blocking Agents

hsATR, scMec1, and spRad3 deficiencies are also associated

with hypersensitivity to replication blocking agents. For exam-

ple, yeast mec1 mutants fail to undergo the G2 arrest normally

induced by stalled replication and instead proceed directly into

mitosis. Consequently, mutant cells accumulate single-strand

gaps and double-strand breaks, leading to cell death (Weinert

et al., 1994). In animals, this failure to arrest induces apoptosis

(Brown and Baltimore, 2000; Cortez et al., 2001). To determine

whether ATR is required for plant cell-cycle responses to

replication blocks, we challenged wild-type and homozygous atr

mutant plants with the replication-blocking agents HU (1 mM),

aphidicolin (12 mg/mL), and UV-B light. All of these agents

inhibit replication in fundamentally different ways. HU inhibits

RNR, thus depleting available dNTPs for DNA polymerases.

Aphidicolin is an inhibitor of the replicative DNA polymerases d

and e, and UV-B light creates physical replication blocks (e.g.,

cyclobutane pyrimidine dimers) in DNA. We chose to focus our

analysis on root meristems because we and others (Wiedemeier

et al., 2002) found that both HU and aphidicolin inhibit root

growth in wild-type plants, suggesting that these agents are

absorbed by the root. We germinated wild-type and homozy-

gous atr seeds on control MS agar plates or on plates

containing either HU or aphidicolin, selecting a dose of each

agent that had mild but perceptible effects on growth of the

wild-type root. Both the wild type and atr grew at very similar

rates on control plates. On plates containing 1 mM HU, wild-

type root growth was reduced �40%, whereas the atr root

growth was reduced >95% (Figures 5A and 5D) and produced

abnormally long root hairs that were densely clustered at the

root tip. Moreover, lateral root growth initiated much earlier in

atr on HU plates compared with control plates. The shoot was

reduced in size and produced anthocyanins (stress-induced

pigments), perhaps as an indirect effect of the lack of root

growth. The same differences between the wild type and atr

(40% versus >95% growth inhibition, overproduction of root

Figure 3. RNA Gel Blot Analysis of the Wild Type and atr.

RNA samples were prepared from 7-d-old wild-type and atr plants

(whole seedlings) sown onMS agar plates plus 1mMHU, in the presence

of UV-B (chronic dose), or left untreated (n.t.). For acute doses of UV-B

light, the wild type and atr were grown for 7 d on MS agar plates and

irradiated with UV-B light (10 kJ/m2) in the dark and then harvested 2 h or

6 h after UV irradiation. Approximately 30 mg (lanes 1 to 6) or 25 mg (lanes

7 to 10) of total RNA from each sample was blotted and probed with

a putative AtRNR1 (GenBank AF092841) mRNA sequence. The AtRNR1

mRNA sequence is 2.65 kb. The bands corresponding to 4.4 and 2.4 kb

of the RNA ladder (Ambion) are shown at the left and denoted as arrows.

The ethidium bromide–stained gel is shown (bottom), and the relative

rates of expression (Rel. exp.) were determined with respect to wild-type

no-treatment control. This experiment was repeated with independently

isolated RNA samples with very similar results.

AtATR Regulates a G2 Checkpoint 5 of 14

hairs and lateral roots) were observed when germinated on

plates containing 12 mg/mL of aphidicolin or when 5-d-old seed-

lings were transferred to plates containing HU or aphidicolin,

starting �2 d after transfer (Figures 5C and 5D). Furthermore,

chronic doses of UV-B light had a similar phenotypic effect

on atr roots compared with the wild type (Figure 5B); the

mutant was also retarded (to a lesser extent at this dose) in

its root growth and formed root hairs at the tip of the root.

Thus, three treatments that block replication but differ in the

mechanism of inhibition produced similar gross phenotypic

effects in atr roots.

AtATR Regulates a G2 Checkpoint in Response to

Replication Blocks

Mec1 and ATR are required for the G2 checkpoint response to

replication blocks in fungi and mammals, respectively (Melo and

Toczyski, 2002). To determine whether ATR performs a similar

function in plants, we GUS and DAPI stained homozygous wild-

type and atr mutant lines carrying the PcyclinB1:GUS reporter

construct (described above) after growth on HU and aphidicolin.

As shown in Figure 6, the wild type and atr displayed typical

patterns of PcyclinB1:GUS expression on control plates (�10 to 20

blue cells per root tip), representing the fraction of G2/early

M cells present during normal growth. By contrast, the meri-

stematic cells of aphidicolin treated wild-type plants displayed

a substantial increase in the number of PcyclinB1:GUS-expressing

cells and perhaps in the density of their staining (Figure 6). Using

DAPI staining to visualize mitotic nuclei, we observed no

significant aphidicolin-induced accumulation of mitotic figures

in either wild-type or atr meristematic cells, thus ruling out

post-G2 mitotic arrest as a likely cause for the PcyclinB1:GUS

accumulation. This indicates that the dividing cells of wild-type

plants experiencing aphidicolin-induced replicational stress

spend a greater fraction of their cell cycle in G2 and that atr

plants are defective in this G2 arrest (Figure 6).

Figure 4. Root Growth and Cell-Cycle Arrest Phenotypes of g-Irradiated Wild-Type and atr Plants.

(A) Wild-type, atm, and atr seeds were g-irradiated at a dose of 200 Gy. The seeds were immediately germinated on MS agar plates, and root growth

was measured for 2 weeks. No-irradation controls grew as described (Garcia et al., 2003) and as in Figure 5D.

(B) Five-day-old wild-type and atr seedlings containing the PcyclinB1:GUS fusion construct (as described in the text and Figure 6) were g-irradiated with

400 Gy, and individual wild-type and atr plants were harvested at 8 h, 1 d, and 2 d time points for GUS staining (n > 12). Two representative examples

from each treatment are shown.

6 of 14 The Plant Cell

In contrast with the aphidicolin response, HU did not induce

G2-phase arrest in wild-type roots. In fact, a decline in the

fraction of G2 cells was observed from day 4 to 6 (as measured

by the lack of PcyclinB1:GUS staining; Figure 6). The cells are

presumably arresting in G1- or S-phase because no increase in

mitotic figures was observed via DAPI staining. This response

also occurred in atr root meristems but appeared to be slightly

accelerated.

To confirm the PcyclinB1:GUS analysis, we employed semi-

quantitative RT-PCR, with primers specific to the cyclinB1 gene

(which does not amplify the PcyclinB1:GUS transcript), on un-

treated, HU-, or aphidicolin-treatedwild-type and atr root tips. As

shown in Figure 7, cyclinB1 gene expression is reduced �20%

(relative to untreated control root tips) in the wild type when

grown in the presence of HU but is elevated nearly threefold in

the presence of aphidicolin, consistent with our PcyclinB1:GUS

experimental conclusion that HU induces G1- or S-phase arrest

and aphidicolin induces G2 arrest. Furthermore, aphidicolin-

treated atr root tips display no enhancement of cyclinB1

transcript in response to aphidicolin, again suggesting the lack

of G2 arrest in atr.

To distinguish between G1- or S-phase arrest in wild-type HU-

treated roots, we employed RT-PCR analysis of the S-phase

specific geneHistone H4. In plants, the expression ofHistone H4

has been shown to be strongly induced upon entry into S-phase

and maintained through G2 (Reichheld et al., 1995; Menges and

Figure 5. Phenotypes of atr Challenged with HU, Aphidicolin, or UV-B Light.

(A) Wild-type and atr seeds were germinated on MS plates or MS plus 1 mM HU and grown for 3 weeks.

(B) Wild-type and atr seeds were germinated on MS plates and grown for 3 weeks in the presence or absence of UV-B light (see Methods).

(C)Wild-typeand atr seedsweregerminatedonMSplates andgrown for 5dand then transferred toplates containing 12mg/mLof aphidicolin (Aph) for 5d.

(D) Five-day-old wild-type and atr seedlings were grown as above in the presence or absence (n.t., no treatment) of HU or aphidicolin (Aph). However, to

make a more direct comparison of the effects of HU versus aphidicolin, we employed concentrations of HU and aphidicolin that inhibited wild-type root

growth by�40% for both treatments, in this case, 1 mM HU and 12 mg/mL of aphidicolin. We measured root growth (n > 30) from each treatment for

8 d after transfer. Standard deviations are shown as error bars.

AtATR Regulates a G2 Checkpoint 7 of 14

Murray, 2002). If HU induces S-phase arrest (e.g., in wild-type

root tips), we would predict a concomitant increase in S-phase–

specific gene expression. Likewise, G1-phase arrest would

predict a concomitant decrease in S-phase–specific gene ex-

pression. As shown in Figure 7,HistoneH4 expression is reduced

in HU-treated wild-type root tips and elevated in aphidicolin-

treated wild-type root tips, suggesting G1- and G2-phase arrest,

respectively. In the atrmutant,Histone H4 expression is elevated

in the untreated root tips but is drastically reduced in the HU- and

aphidicolin-treated root tips. This suggests that the atr mutant,

although phenotypically normal in root growth, induces some S-

phase arrest (cyclinB1 is not induced) under standard growth

conditions but in the presence of replication blocking agents,

arrests and/or exits the cell cycle in G1.

Altogether, these data suggest that two independent check-

points are responsible for the arrest phenotypes observed in the

wild type versus atr: an aphidicolin-induced G2 checkpoint that

is AtATR dependent and a HU-induced G1 checkpoint.

Aphidicolin, but Not HU, Induces Cell Death in atr

To further characterize the different cellular responses to HU and

aphidicolin, we DAPI stained wild-type and atr root tips (time

points and treatments as above) to observe the integrity of the

nuclei. As seen in Figure 8A, wild-type roots had very organized

meristems with round and slightly diffuse nuclei, and they

retained this morphology on HU and aphidicolin. However, atr

exhibited different cellular morphologies on both HU and aphidi-

colin. The cell files of atr meristems grown on HU were dis-

organized and the cells enlarged, but the nuclei were similar in

appearance to the wild type (round and slightly diffuse). All cells

up to the root cap had a defined nucleus (Figure 8A). The

meristems of atr grown on aphidicolin also were disorganized,

but the appearance of the nuclei differed from that observed in

either thewild type (all conditions) or atr grown onHU. Starting on

day 4, the meristematic nuclei near the tip became condensed,

and DAPI staining was reduced relative to atr grown on HU or the

wild type (all conditions). By day 6, most cells in the meristematic

zone of aphidicolin-treated roots had completely lost their nuclei,

leaving only a blue haze (Figures 8A and 8B). This indicates that

atr mutant cells undergo nuclear degradation (cell death) when

treatedwith aphidicolin, but not HU, at dose levels of both agents

that inhibit wild-type growth by�40%. Although some cell death

also can be seen at higher doses of HU (data not shown), these

higher doses inhibited wild-type growth >95%. This suggests that

(1) the aphidicolin-induced AtATR-dependent G2 arrest plays

a critical role in protecting the genome, and (2) that additional

factors, such as the G1 arrest induced by HU, may rescue atr

meristems from massive genomic instability induced by pro-

gression throughM-phase in the absence of complete replication.

DISCUSSION

ATR plays a central role in the cell’s response to DNA damage by

activating cell-cycle checkpoints that allow time for DNA repair.

We have shown that Arabidopsis atr mutants are viable and

normal under standard growth conditions but are hypersensi-

tive to replication blocks and mildly sensitive to g-radiation.

Furthermore, we have shown that AtATR is required for the

G2-phase checkpoint in response to replication blocks but is

not absolutely required in the G2-arrest response to double-

strand breaks (using acute doses of g-radiation). The differences

seen in response to HU, aphidicolin, and g-irradiation suggest

that additional checkpoints also play important roles in the cell-

cycle and cell-death phenotypes observed.

AtATR Plays a Relatively Minor Role in g-Resistance

At 400Gyof g-radiation, the atrmutant is similar to thewild type in

root growth and its ability to recover from g-radiation treatments.

By contrast, atm is not able to recover from g-radiation doses as

Figure 6. GUS Staining of Root Tips.

Five-day-old wild-type and atr-3 seedlings (carrying PcyclinB1:GUS) were transferred to aphidicolin (Aph), or HU, or control (n.t.) plates, and root growth

was measured every other day after transfer for 8 d (Figure 5D). Individual wild-type and atr plants were harvested at each time point for GUS staining of

root tips. Wild-type and atr were grown for 5 d on MS plates and then transferred to control plates or to plates containing 1 mM HU or 12 mg/mL of

aphidicolin. These experiments were repeated several times with very similar results; thus, representative pictures are shown. GUS stained root tips

from 2, 4, and 6 d after transfer are shown. At least 12 individual plants were harvested and stained at each treatment and time point. Two representative

examples are shown for each treatment.

8 of 14 The Plant Cell

low as 80Gy (Garcia et al., 2003). This suggests that AtATR plays

a relatively minor role in g-resistance versus AtATM in plants.

Human cell lines expressing a dominant-negative kinase mutant

of ATR have been used to characterize the relative contributions

of ATR and ATM to g-resistance (Cliby et al., 1998; Wright et al.,

1998). These experiments have clearly demonstrated that ATR is

involved in g-resistance, but independent studies employing

these approaches have yielded inconclusive results in the relative

contributions of ATR and ATM to g-resistance (Jimenez et al.,

1992; Cliby et al., 1998; Wright et al., 1998).

The cell-cycle arrest phenotypes of atr in comparison to the

wild type suggest that the mild g-sensitivity of atr is probably not

an absolute effect of an abrogatedcheckpoint inG2; both thewild

type and atr display an accumulation of GUSwithin themeristem

consistent with G2 arrest, although with slightly different kinet-

ics. Although we do not yet know if AtATM contributes to the

g-induced G2 checkpoint, it is possible that AtATR and AtATM

playpartially redundant roles in this response. In this case, AtATM

may partially compensate for AtATR deficiency throughout all or

part of the cell cycle. Inmammalian cells, experiments employing

conditional knockout (null) alleles of ATR have demonstrated that

G2 arrest is abrogated in response to g-radiation, indicating that

ATR is necessary for the G2 checkpoint in response to double-

strand breaks (Brown and Baltimore, 2003). However, these

experiments used synchronized cells, irradiating cells during S/

G2-phase and then assaying for progression into M-phase.

Because the acute doses of g-radiation used here create double-

strand breaks in all phases of the cell cycle, it is possible that ATR

isprimarily activatedduring a specificphaseof thecell cycle (e.g.,

S-phase) in response to double-strand breaks (for review, see

Abraham, 2001). In this scenario, only a fraction of the

meristematic cells (most are in G1) with DNA damage would

invoke ATR-dependent G2 arrest, such as cells progressing

through S-phase. G1 cells that acquire damage might be

activated forG2 arrest through anATR-independentmechanism,

progress through S-phase, and then arrest in G2-phase. It will be

interesting to determine the cell-cycle responses to acute and

chronic doses of g-radiation in atm and atr atm double mutant

plants in comparison to atr. Nonetheless, our preliminary results

suggest that plants may regulate the G2-phase checkpoint in

response to g-radiation differently than animals.

AtATR Responses to Replication Blocks

It is clear from both yeast and animal studies that ATR plays

a critical role in the G2 checkpoint response to replication blocks.

Many of these studies have employed HU and/or aphidicolin as

model agents for replication blocks. HUor aphidicolin treatment of

wild-type yeast and animals cells (at concentrations that allow S-

phase progression) results in G2 arrest. In ATR-deficient mam-

malian cells, progression into mitosis (without G2 delay) leads to

caspase-independent chromosomal fragmentation, which in turn

induces an apoptotic response (Brown and Baltimore, 2000). We

have shown that atr plants are hypersensitive to the growth-

inhibitory effects of HU and aphidicolin. The G2 arrest (as mea-

sured by PcyclinB1:GUS) in wild-type root meristems on aphidicolin

and the concomitant lack of G2 arrest in atr root meristems indi-

cates that ATR is required for the G2-phase checkpoint response

to replication blocks. The lackofwild-typeG2arrest in response to

HU, however,was surprising. Because the fundamental difference

betweenHUand aphidicolin in blockingS-phase replication is that

HU-treated cells are reduced in dNTPs, this implies that the G1

arrest is activated bya novel checkpoint responsive to lowdNTPs.

Previous studies in human cells demonstrated a p53-dependent

G1 arrest in response to low ribonucleotide levels, but not

deoxynucleotide levels (Linke et al., 1996). The response to low

dNTPsmay help to prevent plant cells from initiatingS-phaseDNA

replication in the absence of sufficient dNTP pools. Because

phosphate is a relatively immobile and limiting nutrient in soil (Abel

et al., 2002), this checkpoint could also represent a plant-specific

response that helps direct root growth away from phosphate-

depleted regions of soil, eliminating growth of root meristems in

phosphate-depleted regions, while promoting the formation of

other lateral root meristems elsewhere along the root.

Figure 7. Semiquantitative RT-PCR Analysis of G2-Phase and S-Phase–

Specific Genes.

Wild-type and atr plants were grown as described in Figure 6. Root tips

(n > 200) were excised (1 to 2 mm from the root cap) from wild-type and

atr plants 4-d after transfer for RNA extraction, cDNA preparation, and

RT-PCR analysis. Twenty-five microliters of each RT-PCR sample was

loaded on an ethidium bromide–stained agarose gel, and representative

gels are shown. All RT-PCR products were between 400 bp and 850 bp

in size. The RT-PCR product of eIF4A (eukaryotic initiation factor) was

employed as a standard for RT-PCR amplification. Numbers below each

gel band denote its relative expression (first normalized to eIF4A) to the

wild-type untreated (n.t.) sample. Standard deviations for two replicates

are shown in parentheses. Aph, aphidicolin.

AtATR Regulates a G2 Checkpoint 9 of 14

Figure 8. DAPI-Stained Root Tips from Wild-Type and atr Grown as in Figure 6.

(A) Shown are root tips day 4 and day 6 after transfer.

(B) Close-up of 6 d after transfer to aphidicolin wild-type (left) and atr (right) DAPI-stained root tips.

10 of 14 The Plant Cell

The cell death observed in atr in response to aphidicolin, but

not HU, suggests that theG1 (dNTP-dependent) checkpoint may

rescue atr from cell death. However, atr mutants are hypersen-

sitive to the effects of HU on the growth of the root, a fact that

may appear paradoxical given the hypothesis that this line is

proficient in the G1 arrest induced by this agent. For this reason,

we remind the reader that the HU-treated wild-type roots are

growing, though at a slightly reduced rate. In other words, the

cells are progressing through the entire cell cycle, although they

are spending a relatively longer fraction of that cycle in G1. We

thus propose the following model for Arabidopsis cell-cycle

checkpoint regulation in response to HU. HU treatment induces

two independent checkpoints in Arabidopsis: one in G1 in

response to low dNTPs (AtATR independent) and one in G2 in

response to replication blocks (AtATR dependent). HU-treated

wild-type and atr meristematic cells arrest in G1 in response to

low dNTP pools, thus retarding their entry into S-phase. The G1

checkpoint is eventually relieved after enough dNTPs accumu-

late to permit entry into S-phase, and some fraction of cells leak

through into S-phase. The resulting S-phase–induced depletion

of dNTPs in the presence of HU then results in a requirement for

AtATR-dependent G2 checkpoint activation. The inability of

atr to activate this G2 checkpoint results in defective DNA

replication and meristematic failure.

By contrast, aphidicolin-treated atr cells do not invoke the G1

checkpoint because they have normal levels of dNTPs and enter

S-phase without delay. atr cells then fail to arrest in G2 and

proceed into mitosis with incompletely replicated genomes,

perhaps to a much larger degree than HU-treated atr cells. This

results in amajority of the daughter G1 cells having an intolerable

amount of single-strand gaps and double-strand breaks, leading

to meristematic failure and cell death.

The cell death seen in atr was unexpected. Mutants in DNA

repair/maintenance genes in Arabidopsis, such asDNA ligase IV,

ERCC1, and telomerase, do not exhibit obvious cell death, even

in response to treatment with DNA damaging agents (Riha et al.,

2001; Friesner and Britt, 2003; Hefner et al., 2003). A prevailing

hypothesis is that plants, which have no p53 homolog, lack DNA

damage–induced cell death (apoptosis). We believe the atr

response to aphidicolin might be an example of programmed cell

death, as indicated by the fact that all cells in the root tip die,

including normally nondividing cells (e.g., epidermal cells), sug-

gesting that the signal for the cell death is diffusible. Alternatively,

it is possible that the cells surrounding the dying meristem are

recruited to reconstitute themissingmeristem, and this induction

of the cell cycle induces their death. In this case, death might

simply be the direct result of the chromosomal fragmentation

induced by progression through M-phase with an incompletely

replicated genome. In any event, it is clear that replication-

induced checkpoints play an important role in protecting cells

from the consequences of replicational stress.

AtATR-Dependent and AtATR-Independent Regulation

of AtRNR1 Transcription

The regulation of RNR in yeast is well defined and involves

transcriptional and posttranslational control (i.e., inhibitory

protein–protein interactions, modification of protein localization,

and allosteric regulation). For instance, the yeast RNR1 (large

subunit) gene is expressed during vegetative growth, upregu-

lated in late G1 and early S, and further upregulated in response

to DNA damage (Elledge et al., 1992). The expression patterns of

several RNR genes have been characterized in tobacco,

including an RNR1-like gene (Chaboute et al., 1998, 2002). The

reduced level of RNR1 transcript (Figure 3) we observed in atr

versus the wild type, in the absence of DNA damaging agents or

with chronic UV-B light, suggests that RNR1 is transcriptionally

maintained in Arabidopsis through both AtATR-dependent and

AtATR-independent mechanisms. Our analysis employing HU

on whole seedlings (Figure 3) further indicate that RNR1 is

regulated transcriptionally, and independently of AtATR, via

a feedback-type mechanism that boosts RNR1 transcripts when

dNTPs are low.

The rapid reduction of AtRNR1 transcript in response to acute

doses of UV-B in wild-type and atr plants suggests an additional

modeof transcriptional regulation forAtRNR1 in response toDNA

damage that is independent ofAtATR. In yeast,RNRupregulation

in response toDNAdamage is rapid, occurringwithin 30min after

treatment (Elledge and Davis, 1990; Elledge, 1993). It was thus

surprising that under our experimental conditions, an acute dose

of UV-B light caused a rapid reduction of AtRNR1 transcript

within 2 h after UV-B treatment, followed by a return to no-

treatment control levels within 6 h after UV-B treatment. This

response could represent a process bywhich inhibition ofAtRNR

expression leads to low dNTP levels and a concomitant arrest of

the cell cycle during DNA damage, possibly through the G1

checkpoint described above. A decrease in RNR expression

might also be expected to inhibit the potentially mutagenic

actions of bypass polymerases, which require high levels of

dNTPs for activity (Johnson et al., 2001; Chabes et al., 2003). One

plausible model for this mode of regulation would be that

immediately after experiencing acute DNA damage, cells initially

try to eliminate as much of the damage as possible via error-free

pathways. Initial downregulation of dNTPs would inhibit the

inherently mutagenic bypass polymerases and contribute to the

arrest of the cell cycle. Return to normal levels of dNTPs then

allows the cell cycle to progress and enables the cell to utilize

bypass polymerases as needed to complete replication. This

model is suggestive of the SOS response in Escherichia coli,

wherein error-free repair mechanisms are induced early, and

error-prone translesion synthesis is induced later (Goodman and

Woodgate, 2000; Rangaragan et al., 2002) in response to acute

induction of DNA damage.

AtATR Is Required for the Completion of Meiosis in the

Absence of AtATM

Given the partial sterility of atm mutant plants and nonsterility of

atr, the observation of complete sterility of the atr atm double

mutant suggests that ATR can partially complement ATM func-

tion during meiosis. Although it is not clear whether ATR has

adistinctmeiotic function or if ATR is simply filling in for the lack of

ATM (or vice versa), our experiments do at the very least suggest

that ATR is present and functional during meiosis in Arabidopsis.

In mammalian cells, in situ localization studies have shown that

both ATM and ATR are localized to meiotic chromosomes

AtATR Regulates a G2 Checkpoint 11 of 14

(Keegan et al., 1996;Moens et al., 1999). Further studies using atr

and atm single and double mutant plants will allow us to better

understand the role of ATR in plant meiosis, possibly leading to

additional insights into ATR function during animal meiosis.

Conclusions

We have shown that plants encode an ortholog of ATR and that

its role in G2 checkpoint regulation is similar in many respects to

its yeast and mammalian counterparts. The presence of ATM

and other downstream effectors of the cell cycle and DNA repair,

including BRCA1 (Lafarge and Montane, 2003) and a putative

CHK1 (S.B. Preuss, K. Culligan, and A. Britt, unpublished data)

for example, further suggest that G2 cell-cycle checkpoint

responses to DNA damage and replication blocks are regulated

in a similar fashion in plants and animals. This is not to say that

plant responses to DNA damage and replication blocks are

identical to that of animals; perhaps the most significant

difference between plants and animals is the lack of a p53-

dependent programmed cell death response to persisting DNA

damage. Plants may encode an analogous pathway for the

degradation of hopelessly compromised cells, as evidenced by

the cell death observed in atr, but this is probably a response to

massive disruption of DNA replication coupled to uninhibited

progression through the cell cycle. Perhaps because plants do

not die from cancer, a cell death response tominor levels of DNA

damage (such as those that occur spontaneously and persist in

repair defective mutants) would be counteradaptive. Because of

this insensitivity (versus animals) to persisting DNA damage,

plants may ultimately prove to be a robust model system for the

study of many fundamental processes in cell-cycle regulation

and DNA repair pathways in eukaryotes (for review, see Hays,

2002).

METHODS

Isolation of Mutants

The atr-1 allele (ecotypeWs) was isolated from the Arabidopsis Knockout

Facility as described (Krysan et al., 1999). The atr-2 allele (ecotype

Columbia; SALK_032841) was identified using the Salk SIGnAL Web site

(http://signal.salk.edu), and the seeds obtained from the ABRC. The atr-3

allele (ecotype Ws) was identified using the FLAG database (Genoplante;

http://genoplante-info.infobiogen.fr/ ).

Growth of Arabidopsis

For standard growth conditions, seeds were sown on 13 MS (GIBCO,

Cleveland, OH) phytagel (Sigma, St. Louis, MO) agar, pH 5.8, and grown

under cool-white lamps filtered through Mylar at an intensity of 100 to

150 mmol/m2/s with a 24-h day light cycle at 218C. HU (Sigma; 500 mM

stock in water) and aphidicolin (ICN Biomedicals, Basingstoke, UK;

20 mg/mL stock in DMSO) were added to MS plates (as above) to con-

centrations described in the text and figure legends. For experiments

using aphidicolin, all control and HU plates contained an equivalent

amount (0.05%) of DMSO to allow direct comparison.

UV-B and g-Irradiation of Arabidopsis

For chronic doses of UV-B light, plants were sown on MS plates and

grown with a 24-h light cycle under cool-white lamps, supplemented with

UV-B light using a mediumwave UV-B lamp (Spectronics, Westbury, NY;

peak intensity 312 nm) filtered through three sheets of cellulose acetate

(Golden State Plastics, Sacramento, CA; thickness 0.12 mm) to eliminate

the UV-C (<280 nm) fraction of emitted light. The cellulose acetate was

replaced every 3 d. The dose rate of UV-B was 2.5 J/s/m2. For acute

doses of UV-B, 5-d-old seedlings were irradiated with a UV trans-

illuminator (Fisher, Houston, TX), peak output at 305 nm, filtered through

a single cellulose acetate sheet. UV-B intensities were measured with

a UV-B radiometer with a peak response at 310 nm and a 50% response

at 280 and 340 nm (UV Products, San Gabriel, CA). For g-radiation

experiments, Arabidopsis seeds and plants were irradiated using a Cs137

source (Institute of Toxicology and Environmental Health, University of

California, Davis). Seeds were imbibed in water at 48C for 3 d, irradiated,

and then immediately placed on MS agar plates for germination and

placed in the growth chamber; plants were grown on MS plates (for 5 d),

irradiated, and then immediately returned to the growth chamber.

Preparation of RNA and RNA Gel Blotting

Total RNA was prepared from 7-d-old seedlings using an RNeasy plant

mini kit (Qiagen, Valencia, CA). RNA was blotted onto positively charged

membranes (BriteStar-Plus; Ambion, Austin, TX) and probed with a32P radioactively labeled AtRNR1 (GenBank AF092841) cDNA fragment.

This 761-bp fragment was generated with PCR primers RNR1 (59-AAG-

AATGGAGTGAGGAACTCTC-39) and RNR3 (59-CTCAAGCAACATTAA-

GCATTAG-39). The resulting bands were visualized on a Storm 860

PhosphorImager and quantified using ImageQuant software (Molecular

Dynamics, Sunnyvale, CA).

Semiquantitative RT-PCR

Wild-type and atr Arabidopsis plants were sown on MS agar plates,

grown for 5 d, then transferred to MS plates (n.t.), MS plates plus 1 mM

HU (1HU), or MS plates plus 12 mg/mL of aphidicolin (1Aph) for 4 d of

additional growth. All plants were verified for their respective pheno-

types described above. Approximately 1 to 2 mm of root tips, from >200

individual plants of each treatment, were excised and pooled together

for RNA extraction (RNeasy mini-prep; Qiagen). RNA samples were

DNase treated (DNase set; Qiagen) and quantified. To produce cDNA

for RT-PCR, 2 mg of total RNA from each condition described above

was reverse transcribed, employing random hexamer primers, with

a Superscript first-strand synthesis kit (Stratagene, La Jolla, CA) in a 40-

mL reaction according to the manufacturers protocol. RT-PCR was

performed in 50 mL reactions using Ex-Taq polymerase (Takara Bio,

Otsu, Shiga, Japan), 2 mL of cDNA, and the following gene-specific

primers: eIF4A-1 (59-CTCTCGCAATCTTCGCTCTTCTCTTT-39), eIF4A-5

(59-TCATAGATCTGGTCCTTGAAAC-39), CycB1-2 (59-GTCGCTTTCT-

TCTTAGTAGCCTTCT-39), CycB1-6 (59-GGCCTCCATTCACTCTCAA-

CAG-39), H4-5 (59-ATGTCTGGTCGTGGAAAGGGAG-39), H4-8 (59-AC\

CAAATTGCGTGTTTCCATTG-39), cATR-1 (59-CAGCGCCCAAAGAAGA-

TCATTC-39), and cATR-3 (59-GGCCCGCTGAGCATGTGGGTTC-39). All

primers (except Histone H4) spanned intron-exon boundaries to

eliminate the possibility of amplifying contaminating genomic DNA.

PCR-cycle parameters for all primer pairs were 20 s at 948C, 20 s at

628C, and 45 s at 728C. Primer pairs were first tested for log-linear

amplification using 18, 21, 24, 27, and 30 cycles of PCR. The optimized

number of cycles for each primer pair was: eIF4A, 23 cycles; cycB1, 27

cycles; Histone H4, 29 cycles; and ATR, 27 cycles. PCR products of

ethidium bromide–stained gels were quantified using ImageQuant

software as above.

GUS and DAPI Staining

GUS staining was performed as described (Jefferson et al., 1987; Preuss

and Britt, 2003). DAPI staining was performed by fixing roots in a solution

12 of 14 The Plant Cell

of ethanol and acetic acid (3:1). The roots were then washed two times in

PBS, pH 7.4, and stained in PBS containing 2.5mg/mL of DAPI (Molecular

Probes, Eugene, OR) for 20 min. The roots were then washed in PBS two

times and mounted on slides in 50% glycerol. Slides were visualized on

a Zeiss Axiophot microscope, and photographs were taken with a Zeiss

MC100 camera. DAPI staining was visualized with a standard UV

fluorescence filter set.

Sequence data from this article have been deposited with the EMBL/

GenBank data libraries under accession numbers AB040133 and

AF092841.

ACKNOWLEDGMENTS

We thank Ying Peng for assistance in isolating the atr atm double mutant

and Cheryl Whistler, Joanna Friesner, and Eli Hefner for critical reading of

the manuscript. We also thank the Arabidopsis Knockout Facility

(University of Wisconsin, Madison) and the SIGnAL (Salk Institute) and

FLAG databases for providing T-DNA knockout resources. This work was

supported by a USDA postdoctoral fellowship (2001-01860) to K.M.C.

and a National Science Foundation Grant (MCB-9983142) to A.B.B.

Received October 30, 2003; accepted February 13, 2004.

REFERENCES

Abel, S., Ticconi, C., and Delatorre, C. (2002). Phosphate sensing in

higher plants. Physiol. Plant 115, 1–8.

Abraham, R. (2001). Cell cycle checkpoint signaling through the ATM

and ATR kinases. Genes Dev. 15, 2177–2196.

Brown, E., and Baltimore, D. (2003). Essential and dispensable roles of

ATR in cell cycle arrest and genome maintenance. Genes Dev. 17,

615–628.

Brown, E.J., and Baltimore, D. (2000). ATR disruption leads to

chromosome fragmentation and early embryonic lethality. Genes

Dev. 15, 397–402.

Chabes, A., Clement, B., Domkin, V., Zhao, X., Rothstein, R., and

Thelander, L. (2003). Survival of DNA damage in yeast directly

depends on increased dNTP levels allowed by relaxed feedback

inhibition of ribonucleotide reductase. Cell 112, 391–401.

Chaboute, M., Combettes, B., Clement, B., Gigot, C., and Philipps,

G. (1998). Molecular characterization of tobacco ribonucleotide

reductase RNR1 and RNR2. Plant Mol. Biol. 38, 797–806.

Chaboute, M.E., Clement, B., and Philipps, G. (2002). S phase and

meristem-specific expression of the tobacco RNR1b gene is medi-

ated by an E2F element located in the 59 leader sequence. J. Biol.

Chem. 277, 17845–17851.

Cliby, W., Roberts, C., Cimprich, K., Stringer, C., Lamb, J.,

Schreiber, S., and Friend, S. (1998). Overexpression of a kinase-

inactive ATR protein causes sensitivity to DNA-damaging agents and

defects in cell cycle checkpoints. EMBO J. 17, 159–169.

Colon-Carmona, A., You, R., Haimovitch-Gal, T., and Doerner, P.

(1999). Spatio-temporal analysis of mitotic activity with a labile cyclin-

GUS fusion protein. Plant J. 20, 503–508.

Cortez, D., Guntuku, S., Qin, J., and Elledge, S.J. (2001). ATR and

ATRIP: Partners in checkpoint signaling. Science 294, 1713–1716.

Crowley, D.J., and Courcelle, J. (2002). Answering the call: Coping

with DNA damage at the most inopportune time. J. Biomed.

Biotechnol. 2, 66–74.

Culligan, K.M., Meyer-Gauen, G., Lyons-Weiler, J., and Hays, J.B.

(2000). Evolutionary origin, diversification and specialization of

eukaryotic MutS homolog mismatch repair proteins. Nucleic Acids

Res. 28, 463–471.

de Klein, A., Muijtjens, M., van Os, R., Verhoeven, Y., Smit, B., Carr,

A., Lehmann, A., and Hoeijmakers, J. (2000). Targeted disruption of

the cell-cycle checkpoint gene ATR leads to early embryonic lethality

in mice. Curr. Biol. 10, 479–482.

Elledge, S., Zhou, Z., and Allen, J.B. (1992). Ribonucleotide re-

ductase: Regulation, regulation, regulation. Trends Biochem. Sci. 17,

119–123.

Elledge, S.J. (1993). DNA damage and cell cycle regulation of

ribonucleotide reductase. Bioessays 15, 333–339.

Elledge, S.J., and Davis, R.W. (1990). Two genes differentially

regulated in the cell cycle and by DNA-damaging agents encode

alternative regulatory subunits of ribonucleotide reductase. Genes

Dev. 4, 740–751.

Friesner, J.D., and Britt, A.B. (2003). Ku80- and DNA ligase IV-deficient

plants are sensitive to ionizing radiation and defective in T-DNA

integration. Plant J. 34, 427–440.

Garcia, V., Bruchet, H., Camescasse, D., Fabienne, G., Bouchez,

D.L., and Tissier, A. (2003). AtATM is essential for meiosis

and the somatic response to DNA damage in plants. Plant Cell 15,

119–132.

Goodman, M.F., and Woodgate, R. (2000). The biochemical basis and

in vivo regulation of SOS-induced mutagenesis promoted by

Escherichia coli DNA polymerase V (UmuD’2C). Cold Spring Harb.

Symp. Quant. Biol. 65, 31–40.

Hays, J. (2002). Arabidopsis thaliana, a versatile model system for study

of eukaryotic genome-maintenance functions. DNA Repair 1,

579–600.

Hefner, E.A., Preuss, S.B., and Britt, A.B. (2003). Arabidopsis mutants

sensitive to gamma radiation include the homolog of the human repair

gene ERCC1. J. Exp. Bot. 54, 669–680.

Huang, M., Zhou, Z., and Elledge, S. (1998). The DNA replication and

damage checkpoint pathways induce transcription by inhibition of the

Crt1 repressor. Cell 94, 595–605.

Jefferson, R., Kavanagh, T.A., and Bevan, M.W. (1987). GUS fusions:

Beta-glucuronidase as a sensitive and versatile gene fusion marker in

higher plants. EMBO J. 6, 3901–3907.

Jimenez, G., Yucel, J., Rowley, R., and Subramani, S. (1992). The

rad31 gene of Schizosaccharomyces pombe is involved in multiple

checkpoint functions and in DNA repair. Proc. Natl. Acad. Sci. USA

89, 4952–4956.

Johnson, R., Haracska, L., Prakash, S., and Prakash, L. (2001). Role

of DNA polymerase zeta in the bypass of a (6-4) TT photoproduct.

Mol. Cell. Biol. 21, 3558–3563.

Keegan, K.S., et al. (1996). The Atr and Atm protein kinases associate

with different sites along meiotically pairing chromosomes. Genes

Dev. 10, 2423–2437.

Krysan, P.J., Young, J.C., and Sussman, M.R. (1999). T-DNA as an

insertional mutagen in Arabidopsis. Plant Cell 11, 2283–2290.

Lafarge, S., and Montane, M. (2003). Characterization of the

Arabidopsis thaliana ortholog of the human breast cancer suscepti-

bility gene 1:AtBRCA1, strongly induced by X-rays. Nucleic Acids

Res. 31, 1148–1155.

Linke, S., Clarkin, K., Di Leonardo, A., Tsou, A., and Wahl, G. (1996).

A reversible, p53-dependent G0/G1 cell cycle arrest induced by

ribonucleotide depletion in the absence of detectable DNA damage.

Genes Dev. 10, 934–947.

Melo, J., and Toczyski, D. (2002). A unified view of the DNA-damage

checkpoint. Curr. Opin. Cell Biol. 14, 237–245.

Menges, M., and Murray, J.A. (2002). Synchronous Arabidopsis

suspension cultures for analysis of cell-cycle gene activity. Plant J.

30, 203–212.

AtATR Regulates a G2 Checkpoint 13 of 14

Moens, P.B., Tarsounas, M., Morita, T., Habu, T., Rottinghaus, S.,

Freire, R., Jackson, S., Barlow, C., and Wynshaw-Boris, A. (1999).

The association of ATR protein with mouse meiotic chromosome

cores. Chromosoma 108, 95–102.

Nyberg, K.A., Michelson, R.J., Putnam, C.W., and Weinert, T.A.

(2002). Toward maintaining the genome: DNA damage and replication

checkpoints. Annu. Rev. Genet. 36, 617–656.

Preuss, S.B., and Britt, A.B. (2003). A DNA damage induced cell cycle

checkpoint in Arabidopsis. Genetics 164, 323–334.

Rangaragan, S., Woodgate, R., and Goodman, M.F. (2002). Replica-

tion restart in UV-irradiated Escherichia coli involving pols II, III, V,

PriA, RecA and RecFOR proteins. Mol. Microbiol. 43, 617–628.

Reichheld, J., Sonobe, S., Clement, B., Chaubet, N., and Gigot, C.

(1995). Cell cycle-regulated histone gene expression in synchronized

plant cells. Plant J. 7, 245–252.

Riha, K., McKnight, T., Griffing, L., and Shippen, D. (2001). Living with

genome instability: Plant responses to telomere dysfunction. Science

291, 1797–1800.

Sedgwick, P., and Boder, E. (1991). Ataxia-telangiectasia. In Hered-

itary Neuropathies and Spinocerebellar Atrophies, J.M.B.V. deJong,

ed (New York: Elsevier Science Publishing), pp. 347–423.

Tanaka, H., Arakawa, H., Yamaguchi, T., Shiraishi, K., Fukuda, S.,

Matsui, K., Takei, Y., and Nakamura, Y. (2000). A ribonucleotide

reductase gene involved in a p53-dependent cell-cycle checkpoint for

DNA damage. Nature 404, 42–49.

Tibbetts, R.S., and Abraham, R.T. (2000). PI3K-related kinases: Roles

in cell cycle regulation and DNA damage responses. In Signalling

Networks and Cell Cycle Control: The Molecular Basis of Cancer and

Other Diseases, J. Gutkind, ed (Totowa, NJ: Humana Press), pp.

267–301.

Tibbetts, R.S., Cortez, D., Brumbaugh, K.M., Scully, R., Livingston,

D., Elledge, S.J., and Abraham, R.T. (2000). Functional interactions

between BRCA1 and the checkpoint kinase ATR during genotoxic

stress. Genes Dev. 14, 2989–3002.

Weinert, T.A., Kiser, G.L., and Hartwell, L.H. (1994). Mitotic

checkpoint genes in budding yeast and the dependence of mitosis

on DNA replication and repair. Genes Dev. 8, 652–665.

Wiedemeier, A., Judy-March, J., Hocart, C., Wasteneys, G.O.,

Williamson, R., and Baskin, T. (2002). Mutant alleles of Arabidopsis

RADIALLY SWOLLEN 4 and 7 reduce growth anisotropy without

altering the transverse orientation of cortical microtubules or cellulose

microfibrils. Development 129, 4821–4830.

Wright, J., Keegan, K., Herendeen, D., Bentley, N., Carr, A.,

Hoekstra, M., and Concannon, P. (1998). Protein kinase mutants

of human ATR increase sensitivity to UV and ionizing radiation and

abrogate cell cycle checkpoint control. Proc. Natl. Acad. Sci. USA 95,

7445–7450.

Yao, R., Zhang, Z., An, X., Bucci, B., Perlstein, D.L., Stubbe, J., and

Huang, M. (2003). Subcellular localization of yeast ribonucleotide

reductase regulated by the DNA replication and damage checkpoint

pathways. Proc. Natl. Acad. Sci. USA 100, 6628–6633.

Zhao, X., Chabes, A., Domkin, V., Thelander, L., and Rothstein, R.

(2001). The ribonucleotide reductase inhibitor Sml1 is a new target of

the Mec1/Rad53 kinase cascade during growth and in response to

DNA damage. EMBO J. 20, 3544–3553.

Zhao, X., Muller, E.G., and Rothstein, R. (1998). A suppressor of two

essential checkpoint genes identifies a novel protein that negatively

affects dNTP pools. Mol. Cell 2, 329–340.

Zhao, X., and Rothstein, R. (2002). The Dun1 checkpoint kinase

phosphorylates and regulates the ribonucleotide reductase inhibitor

Sml1. Proc. Natl. Acad. Sci. USA 99, 3746–3751.

Zou, L., and Elledge, S.J. (2003). Sensing DNA damage through ATRIP

recognition of RPA-ssDNA complexes. Science 300, 1542–1548.

14 of 14 The Plant Cell

DOI 10.1105/tpc.018903; originally published online April 9, 2004;Plant Cell

Kevin Culligan, Alain Tissier and Anne Britt Arabidopsis thalianaATR Regulates a G2-Phase Cell-Cycle Checkpoint in

 This information is current as of September 5, 2014

 

Permissions https://www.copyright.com/ccc/openurl.do?sid=pd_hw1532298X&issn=1532298X&WT.mc_id=pd_hw1532298X

eTOCs http://www.plantcell.org/cgi/alerts/ctmain

Sign up for eTOCs at:

CiteTrack Alerts http://www.plantcell.org/cgi/alerts/ctmain

Sign up for CiteTrack Alerts at:

Subscription Information http://www.aspb.org/publications/subscriptions.cfm

is available at:Plant Physiology and The Plant CellSubscription Information for

ADVANCING THE SCIENCE OF PLANT BIOLOGY © American Society of Plant Biologists