JB Accepts, published online ahead of print on 2 June 2014 J. Bacteriol. doi:10.1128/JB.01494-14 Copyright © 2014, American Society for Microbiology. All Rights Reserved.

1 2 3 4 5 6 7

RecO and RecR are necessary for RecA loading in response to DNA damage and replication fork stress

8 9 10

Justin S. Lenhart1, Eileen R. Brandes1, Jeremy W. Schroeder1, Roderick J. Sorenson2, Hollis D. Showalter2 and Lyle A. Simmons1*

11 12

1

13 14 15

Department of Molecular, Cellular, and Developmental Biology, University of Michigan, Ann Arbor, MI 2

Vahlteich Medicinal Chemistry Core, College of Pharmacy, University of Michigan, Ann Arbor, MI

16 17 18 19 20 21 22 23 24 25

*Corresponding author Mailing address: 4140 Kraus Natural Science building Department of Molecular, Cellular, and Developmental Biology, University of Michigan, Ann Arbor, Michigan 48109. Phone: (734) 647-2016. E-mail: [email protected] Fax: (734) 647-0884. Running Title: RecO and RecR load RecA Keywords: RecO, RecR, RecA, DNA replication, recombination

Lenhart et al.

26 27 28

ABSTRACT RecA is central to maintaining genome integrity in bacterial cells. Despite the near

29

ubiquitous conservation of RecA in eubacteria, the pathways that facilitate RecA loading and

30

repair center assembly have remained poorly understood in Bacillus subtilis. Here, we show that

31

RecA rapidly colocalizes with the DNA polymerase complex (replisome) immediately following

32

DNA damage or damage-independent replication fork arrest. In Escherichia coli, the RecFOR

33

and RecBCD pathways serve to load RecA and the choice between these two pathways depends

34

on the type of damage under repair. We found in B. subtilis that the rapid localization of RecA to

35

repair centers is strictly dependent on RecO and RecR in response to all types of damage

36

examined including a site-specific double-stranded break and damage-independent replication

37

fork arrest. Furthermore, we provide evidence that, although RecF is not required for RecA

38

repair center formation in vivo, RecF does increase the efficiency of repair center assembly,

39

suggesting that RecF may influence the initial stages of RecA nucleation or filament extension.

40

We further identify single-stranded DNA binding protein (SSB) as an additional component

41

important for RecA repair center assembly. Truncation of the SSB C-terminus impairs the ability

42

of B. subtilis to form repair centers in response to damage and damage-independent fork arrest.

43

With these results, we conclude that the SSB-dependent recruitment of RecOR to the replisome

44

is necessary for loading and organizing RecA into repair centers in response to DNA damage and

45

replication fork arrest.

46 47

2

Lenhart et al.

48 49

INTRODUCTION DNA repair pathways are critical for genome maintenance in all living cells [for review

50

(1)]. Both prokaryotic and eukaryotic organisms are constantly exposed to endogenous and

51

exogenous sources of DNA damage that compromise genome integrity. In bacterial cells, RecA

52

is central to genome integrity and is important for strand exchange during homologous

53

recombination, stabilizing stalled replication forks and induction of the SOS transcriptional

54

response to DNA damage [for review (2-4)].

55

In the Gram-positive bacterium Bacillus subtilis, the process that recruits RecA to excess

56

ssDNA has remained unclear (5, 6). In B. subtilis and E. coli, RecA fused to green fluorescent

57

protein (GFP) forms foci in response to DNA damage (5-12). The visualization of assembled

58

RecA-GFP foci (also referred to as repair centers) provides an in vivo assay to understand the

59

events that lead to RecA loading within a living cell.

60

In E. coli, RecA is loaded through two major pathways: RecBCD and RecFOR [for

61

review (3, 13-15)]. During double-strand break repair, double-stranded ends are processed by the

62

RecBCD helicase-nuclease pathway (16). RecBCD generates a 3´ssDNA extension after

63

encountering a Chi site (17). Following end processing, RecBCD physically loads RecA onto

64

the 3´ssDNA extension (18, 19). The second loading pathway in E. coli is the RecFOR pathway

65

(20). In this pathway, RecO and RecR form a complex that is required for RecA loading onto

66

SSB coated ssDNA in vitro. The third component, RecF, can accelerate this reaction, particularly

67

on gapped DNA substrates (21-23). E. coli, RecA also colocalizes to blocked forks and it was

68

shown that RecF and RecO are important for efficient binding of RecA to stalled forks (8, 24).

69

Collectively, these studies show that RecA in E. coli is loaded by the RecBCD and RecFOR

70

pathways in vitro and in vivo.

3

Lenhart et al.

71

In B. subtilis, the AddAB helicase-nuclease enzyme is required for double-stranded end

72

processing and is homologous to E. coli RecBCD (25, 26). Like RecBCD, AddAB generates a 3´

73

ssDNA extension suitable for RecA binding and filament formation (25, 26). In contrast to

74

RecBCD, B. subtilis AddAB has not been shown to physically bind or directly load RecA. B.

75

subtilis also has the RecFOR pathway (27). A notable difference is that B. subtilis RecO is

76

necessary and sufficient to load RecA onto SSB-coated ssDNA in vitro (28). Prior work shows

77

that RecA-GFP foci were largely recO and recR-independent and defects in AddAB do not

78

substantially reduce RecA-GFP foci formation in vivo (5, 29). Therefore, in B. subtilis and other

79

AddAB containing bacteria, the process responsible for loading RecA in response to DNA

80

breaks, strand gaps, and damage-independent replication fork arrest has remained unknown. This

81

is surprising considering that AddAB exists in place of RecBCD in most bacterial organisms

82

(30).

83

Here we report the proteins that are required for RecA to establish repair centers in

84

response to endogenous DNA damage, exogenous sources of DNA damage, and damage-

85

independent replication fork arrest. We use RecA-GFP focus formation as a proxy for RecA

86

loading in vivo. Importantly, we found that RecA mediator proteins RecO and RecR are

87

necessary for RecA to load and organize into repair centers in response to endogenous damage,

88

and all sources of exogenous DNA damage and replication fork arrest tested. These results

89

support the model that RecOR is critical for RecA filament nucleation on ssDNA in B. subtilis

90

and show that AddAB does not direct loading and is dependent on RecOR for RecA loading

91

following end processing. We found that RecF, another RecA mediator protein, did not

92

influence RecA repair center formation to endogenous damage. However, in response to

93

exogenous damage, RecA repair center formation was delayed, in recF deficient cells showing

4

Lenhart et al.

94

that RecF participates in RecA nucleation or enhancement of filament extension. We also found

95

that truncation of the C-terminal 35 amino acids of SSB prevents the DNA damage-dependent

96

increase in assembly of RecA repair centers, supporting a model where SSB recruits RecA-

97

mediator proteins to ssDNA for nucleation and assembly of RecA filaments. With these results,

98

we conclude that RecA loading in B. subtilis is dependent on a single loading pathway, providing

99

novel insight into the mechanism used to establish RecA repair centers in bacteria with the

100 101 102

AddAB end processing pathway. RESULTS

103

RecA colocalizes to the replisome in response to endogenous and exogenous DNA damage.

104

Previous work in E. coli and B. subtilis examined the ability of RecA-GFP to organize into foci

105

(5-12, 31). Prior experiments have tested RecA-GFP as the only source of RecA in vivo (6, 8, 9,

106

32) GFP-RecA expressed ectopically as the sole source of RecA in cells or in merodiploid cells

107

with the native recA allele intact (5, 12). Here, we examined RecA-GFP as the only source of

108

intracellular RecA, expressed at its native locus and under control of its native promoter (6). The

109

fusion allele was fully functional at low levels of damage (Figure S1 and supplemental results).

110

Unless otherwise stated, the experiments described below were performed under conditions of

111

DNA damage where the recA-GFP allele was fully functional.

112

One current model suggests that RecA predominantly localizes away from the replisome

113

in response to DNA damage and that replication is not required for repair center formation (33)

114

while other work shows that replication is required for RecA-GFP to form foci (6) and that these

115

foci form at midcell (6, 8, 12). To differentiate between these two models, we constructed a

116

strain with fluorescent fusions to RecA and the replisome marker DnaX (“Materials and

117

Methods”). Imaging of this strain showed that in untreated cells, ~75% of RecA foci colocalized

5

Lenhart et al.

118

with DnaX (replisome) foci. This observation, in consideration with prior work (6), leads us to

119

suggest that RecA initiates filament formation to establish repair centers at the replication fork

120

followed by a search for homologous DNA. Overall, this result suggests that the bulk of

121

endogenous lesions occurring in the growth conditions used here elicit recruitment of RecA to

122

the replisome (Figure 1A).

123

With this result, we asked if different forms of DNA damage cause RecA to differentially

124

localize to or away from the replisome. Our imaging results showed that following challenge

125

with mitomycin C (MMC), an agent that forms mostly monoadducts and interstrand crosslinks

126

(34), ~65% of RecA foci colocalized with replisomes. Similarly, cells challenged with

127

phleomycin, a DNA break inducing peptide (35), showed that ~68% of RecA colocalized with

128

replisomes. We found it striking that the majority of RecA foci that assembled in response to

129

MMC and phleomycin colocalized with the replisome, supporting earlier work that replication

130

fork progression was important for RecA repair center assembly in response to a site-specific

131

DSB (6). Following challenge with UV, a treatment that would result in DNA strand gaps and

132

replication-stalling thymidine dimers, we found that foci formed immediately (within 5 minutes)

133

and ~84% of RecA were colocalized with replisomes (Figure 1). With these results, we conclude

134

that the majority of RecA repair centers localized to the replisome in response to endogenous and

135

exogenous sources of DNA damage, including phleomycin-generated DNA breaks and MMC-

136

generated inter- and intrastrand crosslinks, two types of damage that require homologous

137

recombination for repair.

138 139

RecA colocalizes to the replisome in response to damage-independent fork arrest. The

140

compound HPUra has been widely used as a tool to rapidly block replication fork progression in

6

Lenhart et al.

141

B. subtilis and other Gram-positive bacteria [e.g. (10, 36-41)]. HPUra is a replication-specific

142

class III DNA polymerase inhibitor which blocks replication within minutes (42). To study the

143

RecA response to damage-independent replication fork arrest, we chose to use this compound

144

because it has been well characterized, but with the limitation that HPUra is not commercially

145

available.

146

We began by synthesizing HPUra (see Supplemental Information). From prior literature,

147

the hydrazine congener H2-HPUra (compound 4) is described as the compound that inhibits

148

DNA synthesis [e.g. (36, 39-41)]. We show the scheme for the attempted synthesis of compound

149

4 and related congeners in Figure 2A. Briefly, we found that HPUra (compound 3) is readily

150

synthesized as described (41). However, numerous attempts to reduce HPUra to H2-HPUra using

151

sodium dithionite as described (39), or slight variations thereof, did not provide compound 4 but

152

only recovered compound 3. This led us to the conclusion that either compound 3 is active as an

153

inhibitor of DNA synthesis or compound 3 is reduced within the cell to produce compound 4. A

154

detailed description of the synthesis and structure determination for 3 and congeners is provided

155

in Supplemental Information (Figures S2-S12).

156

To test compound 3 for DNA synthesis inhibition, we synchronized replication initiation

157

using a temperature sensitive initiation mutant as described (6). We analyzed untreated cells 20

158

and 60 minutes after replication initiation and determined the position of replication forks by

159

quantitative illumina sequencing as described (43). During HPUra treatment, cells were

160

challenged with HPUra 20 minutes after replication initiation followed by analysis of DNA

161

content at 60 minutes. We show that replication forks in untreated cells progress throughout the

162

60 minute time course, while DNA replication in the HPUra treated cells arrests immediately

163

upon HPUra addition showing the same DNA content as the 20 minute untreated control. With

7

Lenhart et al.

164

these results, we conclude that, HPUra (compound 3) provides the desired effect of rapidly

165

arresting DNA synthesis in vivo (Figure 2B).

166

It has been shown that treatment of B. subtilis cells with HPUra blocks DNA synthesis

167

and triggers replication stress as measured by RecA-GFP foci formation in most cells (44). We

168

found that following HPUra treatment, RecA-GFP formed foci in most cells as rapidly as we

169

could prepare the sample for imaging. We found that 76% of cells had RecA-GFP foci at 140

170

seconds post treatment (n= 74), and ~94% of cells had RecA-GFP foci 5 minutes after HPUra

171

treatment demonstrating the rapid assembly of RecA foci following damage-independent

172

replication fork arrest (Figure 1A &C). We performed colocalization with RecA and DnaX and

173

found that ~93% of RecA foci (n=181) were colocalized with replisomes following addition of

174

HPUra to the growth medium (Figure 1C). These results show that damage-independent fork

175

arrest caused RecA to localize to nearly all replisomes in B. subtilis. This work also clarifies the

176

method for HPUra synthesis (Supplemental Information).

177 178

RecA-GFP focus formation is dependent on recO and recR. Having established that RecA-

179

GFP localized to the replisome in most cells in response to endogenous and exogenous sources

180

of damage, we asked which gene product(s) are required for RecA loading, and hence RecA-

181

GFP foci formation in vivo. We predicted that recO might contribute to RecA-GFP focus

182

formation since B. subtilis RecO helps nucleate formation of RecA onto SSB coated ssDNA in

183

vitro (28). To test recO in this capacity, we imaged and quantified RecA-GFP foci in B. subtilis

184

cells bearing a recO disruption (recO::cat). In the absence of recO, RecA-GFP foci formed in

185

less than 2.1% of cells untreated or following exogenous DNA damage (Figure 3). In addition,

186

recO is necessary for damage-independent RecA-GFP filament formation in response to fork

8

Lenhart et al.

187

arrest. This result could be explained by release of the GFP moiety via proteolytic cleavage from

188

RecA in a recO mutant. However, we found that RecA-GFP remained intact in the recO::cat

189

background (Figure S13A). We also complemented the recO mutant allele via ectopic expression

190

(Figure S13B). Together, we found that regardless of the type of damage, RecO is necessary for

191

assembly of RecA repair centers in B. subtilis (Figure 3).

192

Although B. subtilis RecO and RecR do not form a complex, they do function in the same

193

genetic pathway (28, 39). We performed the same experiment in a strain where the recR gene

194

was replaced by a cat cassette and obtained nearly identical results to those obtained in recO

195

deficient cells (Figure 3). Under all conditions examined, we found that RecA-GFP formed foci

196

in 4.5% of cells or less in cells deficient for recR. These results show that recO and recR are both

197

necessary for RecA to form foci in response to DNA damage and damage-independent fork

198

arrest (please see discussion). This result was not due to proteolytic cleavage of RecA-GFP, and

199

this phenotype was complemented by ectopic recR expression (Figure S13B and data not

200

shown). We interpret these results to mean that the RecOR pathway is necessary for RecA

201

loading in vivo.

202

RecF is the third component of the RecFOR pathway, which in E. coli is critical for the

203

repair of DNA gap substrates in vivo. This damage is the product of a single-stranded gap in

204

otherwise double stranded DNA following UV or mitomycin C exposure. To test the role of

205

RecF in RecA-GFP repair center formation, we replaced the recF locus with a cat marker

206

(recF::cat) and interestingly, we found that the percentage of cells with RecA-GFP foci was

207

mostly unaffected in untreated cells lacking recF (wild type cells untreated: 12.4%+2.6.

208

recF::cat untreated: 9.2%+2.1. recO::cat 0.7%+ 0.7 recR::cat strains untreated: 312 for each time point shown.

485 486

Figure 5. SSB is in the recOR pathway. (A) Immunoblot for RecA and DnaN in ssb3+ and

487

ssbΔ35 cells. 25 µg of soluble protein extract was applied per lane and separated via SDS PAGE,

21

Lenhart et al.

488

followed by detection with affinity purified RecA antisera in a 1:200 dilution and a 1:2000

489

dilution for anti DnaN antiserum. (B) Killing curve of cells with the indicated genetic

490

backgrounds over a range of 0 to 0.3 µM MMC. The error bars represent the standard error of

491

the mean (SEM).

492 493

Figure 6. Model for RecOR-dependent RecA loading and establishment of repair centers in

494

B. subtilis at a DSB. We propose that after AddAB catalyzes end resection to generate a 3´

495

ssDNA extension, RecOR is recruited to the large stretches of ssDNA by SSB, and then helps to

496

displace SSB and facilitate RecA loading and nucleation of RecA filament formation for

497

homology-directed repair.

22

Lenhart et al.

Table 1. Truncation of the SSB C-terminal tail eliminates RecA filament formation in response to DNA damage and replication fork stress Relevant Genotype ssb3+ ssbΔ35 ssb3+, recO::cm ssbΔ35, recO::cm ssb3+, recR::cm ssbΔ35, recR::cm

498

Exponential Growth 15.0 ± 3.0 20.0 ± 3.1 1.0 ± 1.1 8.8 ± 2.3

40 nM MMC 44.3 ± 2.5 21.5 ± 3.7 0.4 ± 0.7 9.7 ± 3.4

Condition 0.3 μM Phleomycin 46.4 ± 5.4 26.7 ± 4.0 ND ND

0.3 ± 0.6 11.0 ± 3.2

2.8 ± 1.8 11.0 ± 3.1

ND ND

40 μM HPUra

40 J/m2 UV

87.3 ± 3.8 21.6 ± 3.9 0.3 ± 0.6 9.8 ± 2.9

86.5 ± 3.5 22.6 ± 4.3 ND ND

2.9 ± 1.7 13.1 ± 2.8

ND ND

ND indicates “not determined” and error bars represent the 95% confidence interval with n> 282 cells scored for each strain and condition.

23

Lenhart et al.

499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543

REFERENCES 1.

Friedberg EC, Walker GC, Siede W, Wood RD, Schultz RA, Ellenberger T. 2006. DNA Repair and Mutagenesis: Second Edition. American Society for Microbiology, Washington, DC.

2.

Simmons LA, Foti JJ, Cohen SE, Walker GC. 2008. The SOS Regulatory Network. In Böck A, Curtiss III R, Kaper JB, Karp PD, Neidhardt FC, Nyström T, Slauch JM, C.L. S (ed.), EcoSal-Escherichia coli and Salmonella: cellular and molecular biology. ASM Press, Washington, D.C.

3.

Cox MM. 2007. Regulation of bacterial RecA protein function. Crit Rev Biochem Mol Biol 42:41-63.

4.

Little JW, Mount DW. 1982. The SOS regulatory system of Escherichia coli. Cell 29:11-22.

5.

Kidane D, Graumann PL. 2005. Dynamic formation of RecA filaments at DNA double strand break repair centers in live cells. J Cell Biol 170:357-366.

6.

Simmons LA, Grossman AD, Walker GC. 2007. Replication is required for the RecA localization response to DNA damage in Bacillus subtilis. Proc Natl Acad Sci U S A 104:1360-1365.

7.

Centore RC, Lestini R, Sandler SJ. 2008. XthA (Exonuclease III) regulates loading of RecA onto DNA substrates in log phase Escherichia coli cells. Molecular Microbiology 67:88-101.

8.

Renzette N, Gumlaw N, Nordman JT, Krieger M, Yeh SP, Long E, Centore R, Boonsombat R, Sandler SJ. 2005. Localization of RecA in Escherichia coli K-12 using RecA-GFP. Mol Microbiol 57:1074-1085.

9.

Renzette N, Gumlaw N, Sandler SJ. 2007. DinI and RecX modulate RecA-DNA structures in Escherichia coli K-12. Mol Microbiol 63:103-115.

10.

Bernard R, Marquis KA, Rudner DZ. 2010. Nucleoid occlusion prevents cell division during replication fork arrest in Bacillus subtilis. Mol Microbiol 78:866-882.

11.

Renzette N, Sandler SJ. 2008. Requirements for ATP binding and hydrolysis in RecA function in Escherichia coli. Molecular microbiology 67:1347-1359.

12.

Meile JC, Wu LJ, Ehrlich SD, Errington J, Noirot P. 2006. Systematic localisation of proteins fused to the green fluorescent protein in Bacillus subtilis: identification of new proteins at the DNA replication factory. Proteomics 6:2135-2146.

24

Lenhart et al.

544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588

13.

Clark AJ, Sandler SJ. 1994. Homologous genetic recombination: the pieces begin to fall into place. Crit. Rev. Microbiol. 20:125-142.

14.

Kowalczykowski SC, Dixon DA, Eggleston AK, Lauder SD, Rehrauer WM. 1994. Biochemistry of homologous recombination in Escherichia coli. Microbiol. Rev. 58:401465.

15.

Lenhart JS, Schroeder JW, Walsh BW, Simmons LA. 2012. DNA Repair and Genome Maintenance in Bacillus subtilis. Microbiology and molecular biology reviews : MMBR 76:530-564.

16.

Dillingham MS, Kowalczykowski SC. 2008. RecBCD enzyme and the repair of doublestranded DNA breaks. Microbiol Mol Biol Rev 72:642-671, Table of Contents.

17.

Spies M, Amitani I, Baskin RJ, Kowalczykowski SC. 2007. RecBCD enzyme switches lead motor subunits in response to chi recognition. Cell 131:694-705.

18.

Arnold DA, Kowalczykowski SC. 2000. Facilitated loading of RecA protein is essential to recombination by RecBCD enzyme. J. Biol. Chem. 275:12261-12265.

19.

Spies M, Kowalczykowski SC. 2006. The RecA binding locus of RecBCD is a general domain for recruitment of DNA strand exchange proteins. Molecular Cell 21:573-580.

20.

Umezu K, Kolodner RD. 1994. Protein interactions in genetic recombination in Escherichia coli. Interactions involving RecO and RecR overcome the inhibition of RecA by single-stranded DNA-binding protein. J. Biol. Chem. 269:30005-30013.

21.

Morimatsu K, Wu Y, Kowalczykowski SC. 2012. RecFOR proteins target RecA protein to a DNA gap with either DNA or RNA at the 5' terminus: implication for repair of stalled replication forks. The Journal of Biological Chemistry 287:35621-35630.

22.

Ryzhikov M, Koroleva O, Postnov D, Tran A, Korolev S. 2011. Mechanism of RecO recruitment to DNA by single-stranded DNA binding protein. Nucleic Acids Res 39:6305-6314.

23.

Umezu K, Chi N-W, Kolodner RD. 1993. Biochemical interactions of the Escherichia coli RecF, RecO, and RecR proteins with RecA protein and single-stranded binding protein. Proc. Natl. Acad. Sci. U.S.A. 90:3875-3879.

24.

Lia G, Rigato A, Long E, Chagneau C, Le Masson M, Allemand JF, Michel B. 2013. RecA-promoted, RecFOR-independent progressive disassembly of replisomes stalled by helicase inactivation. Molecular Cell 49:547-557.

25.

Chedin F, Ehrlich SD, Kowalczykowski SC. 2000. The Bacillus subtilis AddAB helicase/nuclease is regulated by its cognate Chi sequence in vitro. J Mol Biol 298:7-20.

25

Lenhart et al.

589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632

26.

Saikrishnan K, Yeeles JT, Gilhooly NS, Krajewski WW, Dillingham MS, Wigley DB. 2012. Insights into Chi recognition from the structure of an AddAB-type helicasenuclease complex. The EMBO journal 31:1568-1578.

27.

Fernandez S, Kobayashi Y, Ogasawara N, Alonso JC. 1999. Analysis of the Bacillus subtilis recO gene: RecO forms part of the RecFLOR function. Mol Gen Genet 261:567573.

28.

Manfredi C, Carrasco B, Ayora S, Alonso JC. 2008. Bacillus subtilis RecO nucleates RecA onto SsbA-coated single-stranded DNA. J Biol Chem 283:24837-24847.

29.

Sanchez H, Kidane D, Castillo Cozar M, Graumann PL, Alonso JC. 2006. Recruitment of Bacillus subtilis RecN to DNA double-strand breaks in the absence of DNA end processing. J Bacteriol 188:353-360.

30.

Cromie GA. 2009. Phylogenetic ubiquity and shuffling of the bacterial RecBCD and AddAB recombination complexes. J Bacteriol 191:5076-5084.

31.

Lesterlin C, Ball G, Schermelleh L, Sherratt DJ. 2013. RecA bundles mediate homology pairing between distant sisters during DNA break repair. Nature 10.1038/nature12868.

32.

Simmons LA, Goranov AI, Kobayashi H, Davies BW, Yuan DS, Grossman AD, Walker GC. 2009. Comparison of responses to double-strand breaks between Escherichia coli and Bacillus subtilis reveals different requirements for SOS induction. J Bacteriol 191:1152-1161.

33.

Kidane D, Sanchez H, Alonso JC, Graumann PL. 2004. Visualization of DNA doublestrand break repair in live bacteria reveals dynamic recruitment of Bacillus subtilis RecF, RecO and RecN proteins to distinct sites on the nucleoids. Mol Microbiol 52:1627-1639.

34.

Dronkert ML, Kanaar R. 2001. Repair of DNA interstrand cross-links. Mutat Res 486:217-247.

35.

Sleigh MJ. 1976. The mechanism of DNA breakage by phleomycin in vitro. Nucleic Acids Res 3:891-901.

36.

Gass KB, Low RL, Cozzarelli NR. 1973. Inhibition of a DNA polymerase from Bacillus subtilis by hydroxyphenylazopyrimidines. Proceedings of the National Academy of Sciences of the United States of America 70:103-107.

37.

Goranov AI, Katz L, Breier AM, Burge CB, Grossman AD. 2005. A transcriptional response to replication status mediated by the conserved bacterial replication protein DnaA. Proc Natl Acad Sci U S A 102:12932-12937.

26

Lenhart et al.

633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677

38.

Goranov AI, Kuester-Schoeck E, Wang JD, Grossman AD. 2006. Characterization of the global transcriptional responses to different types of DNA damage and disruption of replication in Bacillus subtilis. J Bacteriol 188:5595-5605.

39.

Mackenzie JM, Neville MM, Wright GE, Brown NC. 1973. Hydroxyphenylazopyrimidines: characterization of the active forms and their inhibitory action on a DNA polymerase from Bacillus subtilis. Proceedings of the National Academy of Sciences of the United States of America 70:512-516.

40.

Neville MM, Brown NC. 1972. Inhibition of a discrete bacterial DNA polymerase by 6(p-hydroxyphenylazo)-uracil and 6-(p-hydroxyphenylazo-)-isocytosine. Nature: New biology 240:80-82.

41.

Wright GE, Brown NC. 1974. Synthesis of 6-(phenylhydrazino)uracils and their inhibition of a replication-specific deoxyribonucleic acid polymerase. J Med Chem 17:1277-1282.

42.

Brown NC. 1970. 6-(p-hydroxyphenylazo)-uracil: a selective inhibitor of host DNA replication in phage-infected Bacillus subtilis. Proceedings of the National Academy of Sciences of the United States of America 67:1454-1461.

43.

Walsh BW, Bolz SA, Wessel SR, Schroeder JW, Keck JL, Simmons LA. 2014. RecD2 Helicase Limits Replication Fork Stress in Bacillus subtilis. Journal of Bacteriology 196:1359-1368.

44.

Wang JD, Sanders GM, Grossman AD. 2007. Nutritional control of elongation of DNA replication by (p)ppGpp. Cell 128:865-875.

45.

Bell JC, Plank JL, Dombrowski CC, Kowalczykowski SC. 2012. Direct imaging of RecA nucleation and growth on single molecules of SSB-coated ssDNA. Nature 491:274-278.

46.

Costes A, Lecointe F, McGovern S, Quevillon-Cheruel S, Polard P. 2010. The Cterminal domain of the bacterial SSB protein acts as a DNA maintenance hub at active chromosome replication forks. PLoS Genet 6:e1001238.

47.

Shereda RD, Kozlov AG, Lohman TM, Cox MM, Keck JL. 2008. SSB as an organizer/mobilizer of genome maintenance complexes. Crit Rev Biochem Mol Biol 43:289-318.

48.

Chase JW, L'Italien JJ, Murphy JB, Spicer EK, Williams KR. 1984. Characterization of the Escherichia coli SSB-113 mutant single-stranded DNA-binding protein. Cloning of the gene, DNA and protein sequence analysis, high pressure liquid chromatography peptide mapping, and DNA-binding studies. The Journal of Biological Chemistry 259:805-814.

27

Lenhart et al.

678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721

49.

Curth U, Genschel J, Urbanke C, Greipel J. 1996. In vitro and in vivo function of the C-terminus of Escherichia coli single-stranded DNA binding protein. Nucleic Acids Research 24:2706-2711.

50.

Johnson BF. 1977. Genetic mapping of the lexC-113 mutation. Molecular & general genetics : MGG 157:91-97.

51.

Lecointe F, Serena C, Velten M, Costes A, McGovern S, Meile JC, Errington J, Ehrlich SD, Noirot P, Polard P. 2007. Anticipating chromosomal replication fork arrest: SSB targets repair DNA helicases to active forks. Embo J 26:4239-4251.

52.

Antony E, Weiland E, Yuan Q, Manhart CM, Nguyen B, Kozlov AG, McHenry CS, Lohman TM. 2013. Multiple C-terminal tails within a single E. coli SSB homotetramer coordinate DNA replication and repair. Journal of Molecular Biology 425:4802-4819.

53.

Gupta R, Ryzhikov M, Koroleva O, Unciuleac M, Shuman S, Korolev S, Glickman MS. 2013. A dual role for mycobacterial RecO in RecA-dependent homologous recombination and RecA-independent single-strand annealing. Nucleic Acids Research 41:2284-2295.

54.

Lu D, Keck JL. 2008. Structural basis of Escherichia coli single-stranded DNA-binding protein stimulation of exonuclease I. Proceedings of the National Academy of Sciences of the United States of America 105:9169-9174.

55.

Shereda RD, Bernstein DA, Keck JL. 2007. A central role for SSB in Escherichia coli RecQ DNA helicase function. J Biol Chem 282:19247-19258.

56.

Hardwood CR, Cutting SM. 1990. Molecular Biological Methods for Bacillus. John Wiley & Sons, Chichester.

57.

Smith BT, Grossman AD, Walker GC. 2001. Visualization of mismatch repair in bacterial cells. Mol. Cell 8:1197-1206.

58.

Lemon KP, Grossman AD. 2000. Movement of replicating DNA through a stationary replisome. Mol. Cell 6:1321-1330.

59.

Dupes NM, Walsh BW, Klocko AD, Lenhart JS, Peterson HL, Gessert DA, Pavlick CE, Simmons LA. 2010. Mutations in the Bacillus subtilis beta clamp that separate its roles in DNA replication from mismatch repair. J Bacteriol 192:3452-3463.

60.

Klocko AD, Schroeder JW, Walsh BW, Lenhart JS, Evans ML, Simmons LA. 2011. Mismatch repair causes the dynamic release of an essential DNA polymerase from the replication fork. Mol Microbiol 82:648-663.

28

Lenhart et al.

722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740

61.

Klocko AD, Crafton KM, Walsh BW, Lenhart JS, Simmons LA. 2010. Imaging mismatch repair and cellular responses to DNA damage in Bacillus subtilis. J Vis Exp 36:1-4.

62.

Davies BW, Bogard RW, Dupes NM, Gerstenfeld TA, Simmons LA, Mekalanos JJ. 2011. DNA damage and reactive nitrogen species are barriers to Vibrio cholerae colonization of the infant mouse intestine. PLoS Pathog 7:e1001295.

63.

Kobayashi H, Simmons LA, Yuan DS, Broughton WJ, Walker GC. 2008. Multiple Ku orthologues mediate DNA non-homologous end-joining in the free-living form and during chronic infection of Sinorhizobium meliloti. Mol Microbiol 67:350-363.

64.

Schroeder JW, Simmons LA. 2013. Complete Genome Sequence of Bacillus subtilis Strain PY79. Genome Announcements 1(6). pii: e01085-13. doi: 10.1128/genomeA.01085-13

65.

Li H, Durbin R. 2009. Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics 25:1754-1760.

29

A

Colocalization of RecA-GFP with DnaX-mCherry % Colocalization (+/- 95% CI)

Condition

Number of Cells Analyzed

Endogenous (untreated)

74.8 (8.4)

103

40 nM Mitomycin C

65.0 (6.7)

197

84.3 (5.8)

153

400 nM Phleomycin

67.9 (6.3)

212

40 μM HPUra

93.3 (3.3)

223

UV 40 Joules/m

2

Replisome

RecA-GFP

Overlay

B

C

Figure 1.

Phleomycin

Phleomycin

Phleomycin

HPUra

HPUra

HPUra

A

B

0.0

log2(fold enrichment) 0.2 0.4

0.6

60 minutes 20 minutes HPUra

2

Figure 2.

3

0 1 genome position (Mb)

2

Percent of cells with RecA-GFP foci

100 75 50 25 0 WT

Genotype Mitomycin C Phleomycin HPUra UV

Figure 3.

-

+ -

- - -+- -- + -- - +-

recO::cat

+ -

- - +- - +- -+

recR::cat

-

recF::cat

+- - - - +- - - - +- - - -+-

+- - - +- - - +- - -+

A

C

Percentage of cells with RecA-GFP foci

100

80 WT

70 80

60 50

60

40 40

30

recF::cat

20

20

10 0

recO::cat recR::cat

0 0

10

20

30

40

Time (min)

B α-RecA α-SSB recF

Figure 4.

+ -

50

60

0

20

40

60

80

100

120

Time (min) after I-SceI expression

α-DnaN

Relative Survival

α-RecA

bΔ 3

B

1

ss

ss

b3 +

5

A

0.1

0.01

0.001

0.0001

0

0.1 0.2 0.3 Mitomycin C (µM) ssbΔ35

ssb3+

Δ35, recO::cat

ssb3+, recO::cat

Δ35, recR::cat

ssb3+, recR::cat ssb3+, recA::spec

Figure 5.

1) 3´ ssDNA bound by SSB tetramers recruits the RecOR proteins

SSB RecO RecA RecF RecR

2) RecF and RecOR nucleate RecA on ssDNA by displacing SSB

filament growth

3) RecOR facilitate RecA filament growth

Figure 6.

RecO and RecR are necessary for RecA loading in response to DNA damage and replication fork stress.

RecA is central to maintaining genome integrity in bacterial cells. Despite the near-ubiquitous conservation of RecA in eubacteria, the pathways that ...
3MB Sizes 0 Downloads 3 Views