首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Bacillus subtilis competence-induced RecA, SsbA, SsbB, and DprA are required to internalize and to recombine single-stranded (ss) DNA with homologous resident duplex. RecA, in the ATP·Mg2+-bound form (RecA·ATP), can nucleate and form filament onto ssDNA but is inactive to catalyze DNA recombination. We report that SsbA or SsbB bound to ssDNA blocks the RecA filament formation and fails to activate recombination. DprA facilitates RecA filamentation; however, the filaments cannot engage in DNA recombination. When ssDNA was preincubated with SsbA, but not SsbB, DprA was able to activate DNA strand exchange dependent on RecA·ATP. This work demonstrates that RecA·ATP, in concert with SsbA and DprA, catalyzes DNA strand exchange, and SsbB is an accessory factor in the reaction. In contrast, RecA·dATP efficiently catalyzes strand exchange even in the absence of single-stranded binding proteins or DprA, and addition of the accessory factors marginally improved it. We proposed that the RecA-bound nucleotide (ATP and to a lesser extent dATP) might dictate the requirement for accessory factors.  相似文献   

2.
We have investigated the structural, biochemical and cellular roles of the two single-stranded (ss) DNA-binding proteins from Bacillus subtilis, SsbA and SsbB. During transformation, SsbB localizes at the DNA entry pole where it binds and protects internalized ssDNA. The 2.8-Å resolution structure of SsbB bound to ssDNA reveals a similar overall protein architecture and ssDNA-binding surface to that of Escherichia coli SSB. SsbA, which binds ssDNA with higher affinity than SsbB, co-assembles onto SsbB-coated ssDNA and the two proteins inhibit ssDNA binding by the recombinase RecA. During chromosomal transformation, the RecA mediators RecO and DprA provide RecA access to ssDNA. Interestingly, RecO interaction with ssDNA-bound SsbA helps to dislodge both SsbA and SsbB from the DNA more efficiently than if the DNA is coated only with SsbA. Once RecA is nucleated onto the ssDNA, RecA filament elongation displaces SsbA and SsbB and enables RecA-mediated DNA strand exchange. During plasmid transformation, RecO localizes to the entry pole and catalyzes annealing of SsbA- or SsbA/SsbB-coated complementary ssDNAs to form duplex DNA with ssDNA tails. Our results provide a mechanistic framework for rationalizing the coordinated events modulated by SsbA, SsbB and RecO that are crucial for RecA-dependent chromosomal transformation and RecA-independent plasmid transformation.  相似文献   

3.
Genetic data have revealed that the absence of Bacillus subtilis RecO and one of the end-processing avenues (AddAB or RecJ) renders cells as sensitive to DNA damaging agents as the null recA, suggesting that both end-resection pathways require RecO for recombination. RecA, in the rATP·Mg2+ bound form (RecA·ATP), is inactive to catalyze DNA recombination between linear double-stranded (ds) DNA and naked complementary circular single-stranded (ss) DNA. We showed that RecA·ATP could not nucleate and/or polymerize on SsbA·ssDNA or SsbB·ssDNA complexes. RecA·ATP nucleates and polymerizes on RecO·ssDNA·SsbA complexes more efficiently than on RecO·ssDNA·SsbB complexes. Limiting SsbA concentrations were sufficient to stimulate RecA·ATP assembly on the RecO·ssDNA·SsbB complexes. RecO and SsbA are necessary and sufficient to ‘activate’ RecA·ATP to catalyze DNA strand exchange, whereas the AddAB complex, RecO alone or in concert with SsbB was not sufficient. In presence of AddAB, RecO and SsbA are still necessary for efficient RecA·ATP-mediated three-strand exchange recombination. Based on genetic and biochemical data, we proposed that SsbA and RecO (or SsbA, RecO and RecR in vivo) are crucial for RecA activation for both, AddAB and RecJ–RecQ (RecS) recombinational repair pathways.  相似文献   

4.
Naturally transformable bacteria recombine internalized ssDNA with a homologous resident duplex (chromosomal transformation) or complementary internalized ssDNAs (plasmid or viral transformation). Bacillus subtilis competence-induced DprA, RecA, SsbB, and SsbA proteins are involved in the early processing of the internalized ssDNA, with DprA physically interacting with RecA. SsbB and SsbA bind and melt secondary structures in ssDNA but limit RecA loading onto ssDNA. DprA binds to ssDNA and facilitates partial dislodging of both single-stranded binding (SSB) proteins from ssDNA. In the absence of homologous duplex DNA, DprA does not significantly increase RecA nucleation onto protein-free ssDNA. DprA facilitates RecA nucleation and filament extension onto SsbB-coated or SsbB plus SsbA-coated ssDNA. DprA facilitates RecA-mediated DNA strand exchange in the presence of both SSB proteins. DprA, which plays a crucial role in plasmid transformation, anneals complementary strands preferentially coated by SsbB to form duplex circular plasmid molecules. Our results provide a mechanistic framework for conceptualizing the coordinated events modulated by SsbB in concert with SsbA and DprA that are crucial for RecA-dependent chromosomal transformation and RecA-independent plasmid transformation.  相似文献   

5.
Bacillus subtilis RecO plays a central role in recombinational repair and genetic recombination by (i) stimulating RecA filamentation onto SsbA-coated single-stranded (ss) DNA, (ii) modulating the extent of RecA-mediated DNA strand exchange and (iii) promoting annealing of complementary DNA strands. Here, we report that RecO-mediated strand annealing is facilitated by cognate SsbA, but not by a heterologous one. Analysis of non-productive intermediates reveals that RecO interacts with SsbA-coated ssDNA, resulting in transient ternary complexes. The self-interaction of ternary complexes via RecO led to the formation of large nucleoprotein complexes. In the presence of homology, SsbA, at the nucleoprotein, removes DNA secondary structures, inhibits spontaneous strand annealing and facilitates RecO loading onto SsbA–ssDNA complex. RecO relieves SsbA inhibition of strand annealing and facilitates transient and random interactions between homologous naked ssDNA molecules. Finally, both proteins lose affinity for duplex DNA. Our results provide a mechanistic framework for rationalizing protein release and dsDNA zippering as coordinated events that are crucial for RecA-independent plasmid transformation.  相似文献   

6.
Subsaturating amounts of Bacillus subtilis SsbA, independently of the order of addition, partially inhibit the single-stranded DNA-dependent dATPase activity of RecA. This negative effect is fully overcome when a substoichiometric amount of RecO is added. SsbA added prior to RecA does not stimulate the dATP-dependent DNA strand exchange activity; however, added after RecA it enhances the extent of strand exchange. The addition of RecO stimulates RecA-mediated joint molecule formation, although it limits the accumulation of final recombination products. Thus we suggest that RecO has a dual activity: RecO acts as a RecA mediator enabling RecA to utilize SsbA-coated single-stranded DNA as a polymerization substrate and controls RecA-mediated DNA strand exchange by limiting its extent. We herein discuss the possible mechanisms of RecO involvement in the regulation of double strand break repair and genetic transformation.  相似文献   

7.
Genetic and cytological evidences suggest that Bacillus subtilis RecN acts prior to and after end-processing of DNA double-strand ends via homologous recombination, appears to participate in the assembly of a DNA repair centre and interacts with incoming single-stranded (ss) DNA during natural transformation. We have determined the architecture of RecN–ssDNA complexes by atomic force microscopy (AFM). ATP induces changes in the architecture of the RecN–ssDNA complexes and stimulates inter-complex assembly, thereby increasing the local concentration of DNA ends. The large CII and CIII complexes formed are insensitive to SsbA (counterpart of Escherichia coli SSB or eukaryotic RPA protein) addition, but RecA induces dislodging of RecN from the overhangs of duplex DNA molecules. Reciprocally, in the presence of RecN, RecA does not form large RecA–DNA networks. Based on these results, we hypothesize that in the presence of ATP, RecN tethers the 3′-ssDNA ends, and facilitates the access of RecA to the high local concentration of DNA ends. Then, the resulting RecA nucleoprotein filaments, on different ssDNA segments, might promote the simultaneous genome-wide homology search.  相似文献   

8.
A proteolyzed bacteriophage (phage) might release its DNA into the environment. Here, we define the recombination functions required to resurrect an infective lytic phage from inactive environmental viral DNA in naturally competent Bacillus subtilis cells. Using phage SPP1 DNA, a model that accounts for the obtained data is proposed (i) the DNA uptake apparatus takes up environmental SPP1 DNA, fragments it, and incorporates into the cytosol different linear single-stranded (ss) DNA molecules shorter than genome-length; (ii) the SsbA-DprA mediator loads RecA onto any fragmented linear SPP1 ssDNA, but negative modulators (RecX and RecU) promote a net RecA disassembly from these ssDNAs not homologous to the host genome; (iii) single strand annealing (SSA) proteins, DprA and RecO, anneal the SsbA- or SsbB-coated complementary strands, yielding tailed SPP1 duplex intermediates; (iv) RecA polymerized on these tailed intermediates invades a homologous region in another incomplete molecule, and in concert with RecD2 helicase, reconstitutes a complete linear phage genome with redundant regions at the ends of the molecule; and (v) DprA, RecO or viral G35P SSA, may catalyze the annealing of these terminally redundant regions, alone or with the help of an exonuclease, to produce a circular unit-length duplex viral genome ready to initiate replication.  相似文献   

9.
Bacillus subtilis RecA is important for spore resistance to DNA damage, even though spores contain a single non-replicating genome. We report that inactivation of RecA or its accessory factors, RecF, RecO, RecR and RecX, drastically reduce survival of mature dormant spores to ultrahigh vacuum desiccation and ionizing radiation that induce single strand (ss) DNA nicks and double-strand breaks (DSBs). The presence of non-cleavable LexA renders spores less sensitive to DSBs, and spores impaired in DSB recognition or end-processing show sensitivities to X-rays similar to wild-type. In vitro RecA cannot compete with SsbA for nucleation onto ssDNA in the presence of ATP. RecO is sufficient, at least in vitro, to overcome SsbA inhibition and stimulate RecA polymerization on SsbA-coated ssDNA. In the presence of SsbA, RecA slightly affects DNA replication in vitro, but addition of RecO facilitates RecA-mediated inhibition of DNA synthesis. We propose that repairing of the DNA lesions generates a replication stress to germinating spores, and the RecA·ssDNA filament might act by preventing potentially dangerous forms of DNA repair occurring during replication. RecA might stabilize a stalled fork or prevent or promote dissolution of reversed forks rather than its cleavage that should require end-processing.  相似文献   

10.
To initiate homologous recombination, sequence similarity between two DNA molecules must be searched for and homology recognized. How the search for and recognition of homology occurs remains unproven. We have examined the influences of DNA topology and the polarity of RecA–single-stranded (ss)DNA filaments on the formation of synaptic complexes promoted by RecA. Using two complementary methods and various ssDNA and duplex DNA molecules as substrates, we demonstrate that topological constraints on a small circular RecA–ssDNA filament prevent it from interwinding with its duplex DNA target at the homologous region. We were unable to detect homologous pairing between a circular RecA–ssDNA filament and its relaxed or supercoiled circular duplex DNA targets. However, the formation of synaptic complexes between an invading linear RecA–ssDNA filament and covalently closed circular duplex DNAs is promoted by supercoiling of the duplex DNA. The results imply that a triplex structure formed by non-Watson–Crick hydrogen bonding is unlikely to be an intermediate in homology searching promoted by RecA. Rather, a model in which RecA-mediated homology searching requires unwinding of the duplex DNA coupled with local strand exchange is the likely mechanism. Furthermore, we show that polarity of the invading RecA–ssDNA does not affect its ability to pair and interwind with its circular target duplex DNA.  相似文献   

11.
The Rad50, Mre11 and Nbs1 complex is involved in many essential chromosomal organization processes dealing with DNA ends, including two major pathways of DNA double-strand break repair, homologous recombination and non-homologous end joining. Previous data on the structure of the human Rad50 and Mre11 (R/M) complex suggest that a common role for the protein complex in these processes is to provide a physical link between DNA ends such that they can be processed in an organized and coordinated manner. Here we describe the DNA binding properties of the R/M complex. The complex bound to both single-stranded and double-stranded DNA. Scanning force microscopy analysis of DNA binding by R/M showed the requirement for an end to form oligomeric R/M complexes, which could then migrate or transfer away from the end. The R/M complex had a lower preference for DNA substrates with 3′-overhangs compared with blunt ends or 5′-overhangs. Interestingly, ATP binding, but not hydrolysis, increased the preference of R/M binding to DNA substrates with 3′-overhangs relative to substrates with blunt ends and 5′-overhangs.  相似文献   

12.
The repair of single-stranded gaps in duplex DNA by homologous recombination requires the proteins of the RecF pathway. The assembly of RecA protein onto gapped DNA (gDNA) that is complexed with the single-stranded DNA-binding protein is accelerated by the RecF, RecO, and RecR (RecFOR) proteins. Here, we show the RecFOR proteins specifically target RecA protein to gDNA even in the presence of a thousand-fold excess of single-stranded DNA (ssDNA). The binding constant of RecF protein, in the presence of the RecOR proteins, to the junction of ssDNA and dsDNA within a gap is 1–2 nm, suggesting that a few RecF molecules in the cell are sufficient to recognize gDNA. We also found that the nucleation of a RecA filament on gDNA in the presence of the RecFOR proteins occurs at a faster rate than filament elongation, resulting in a RecA nucleoprotein filament on ssDNA for 1000–2000 nucleotides downstream (5′ → 3′) of the junction with duplex DNA. Thus, RecA loading by RecFOR is localized to a region close to a junction. RecFOR proteins also recognize RNA at the 5′-end of an RNA-DNA junction within an ssDNA gap, which is compatible with their role in the repair of lagging strand gaps at stalled replication forks.  相似文献   

13.
Cytological and genetic evidence suggests that the Bacillus subtilis DNA uptake machinery localizes at a single cell pole and takes up single-stranded (ss) DNA. The integration of homologous donor DNA into the recipient chromosome requires RecA, while plasmid establishment, which is independent of RecA, requires at least RecO and RecU. RecA and RecN colocalize at the polar DNA uptake machinery, from which RecA forms filamentous structures, termed threads, in the presence of chromosomal DNA. We show that the transformation of chromosomal and of plasmid DNA follows distinct pathways. In the absence of DNA, RecU accumulated at a single cell pole in competent cells, dependent on RecA. Upon addition of any kind of DNA, RecA formed highly dynamic thread structures, which rapidly grew and shrank, and RecU dissipated from the pole. RecO visibly accumulated at the cell pole only upon addition of plasmid DNA, and, to a lesser degree, of phage DNA, but not of chromosomal DNA. RecO accumulation was weakly influenced by RecN, but not by RecA. RecO annealed ssDNA complexed with SsbA in vitro, independent of any nucleotide cofactor. The DNA end-joining Ku protein was also found to play a role in viral and plasmid transformation. On the other hand, transfection with SPP1 phage DNA required functions from both chromosomal and plasmid transformation pathways. The findings show that competent bacterial cells possess a dynamic DNA recombination machinery that responds in a differential manner depending if entering DNA shows homology with recipient DNA or has self-annealing potential. Transformation with chromosomal DNA only requires RecA, which forms dynamic filamentous structures that may mediate homology search and DNA strand invasion. Establishment of circular plasmid DNA requires accumulation of RecO at the competence pole, most likely mediating single-strand annealing, and RecU, which possibly down-regulates RecA. Transfection with SPP1 viral DNA follows an intermediate route that contains functions from both chromosomal and plasmid transformation pathways.  相似文献   

14.
15.
Biochemical analysis of human Dna2   总被引:1,自引:1,他引:0  
Yeast Dna2 helicase/nuclease is essential for DNA replication and assists FEN1 nuclease in processing a subset of Okazaki fragments that have long single-stranded 5′ flaps. It is also involved in the maintenance of telomeres. DNA2 is a gene conserved in eukaryotes, and a putative human ortholog of yeast DNA2 (ScDNA2) has been identified. Little is known about the role of human DNA2 (hDNA2), although complementation experiments have shown that it can function in yeast to replace ScDNA2. We have now characterized the biochemical properties of hDna2. Recombinant hDna2 has single-stranded DNA-dependent ATPase and DNA helicase activity. It also has 5′–3′ nuclease activity with preference for single-stranded 5′ flaps adjacent to a duplex DNA region. The nuclease activity is stimulated by RPA and suppressed by steric hindrance at the 5′ end. Moreover, hDna2 shows strong 3′–5′ nuclease activity. This activity cleaves single-stranded DNA in a fork structure and, like the 5′–3′ activity, is suppressed by steric hindrance at the 3′-end, suggesting that the 3′–5′ nuclease requires a 3′ single-stranded end for activation. These biochemical specificities are very similar to those of the ScDna2 protein, but suggest that the 3′–5′ nuclease activity may be more important than previously thought.  相似文献   

16.
A new technique of PCR hot start using oligonucleotide primers with a stem–loop structure is developed here. The molecular beacon oligonucleotide structure without any chromophore addition to the ends was used. The 3′-end sequence of the primers was complementary to the target and five or six nucleotides complementary to the 3′-end were added to the 5′-end. During preparation of the reaction mixture and initial heating, the oligonucleotide has a stem–loop structure and cannot serve as an effective primer for DNA polymerase. After heating to the annealing temperature it acquires a linear structure and primer extension can begin.  相似文献   

17.
Telomeres are DNA repeated sequences that associate with shelterin proteins and protect the ends of eukaryotic chromosomes. Human telomeres are composed of 5′TTAGGG repeats and ends with a 3′ single-stranded tail, called G-overhang, that can be specifically bound by the shelterin protein hPOT1 (human Protection of Telomeres 1). In vitro studies have shown that the telomeric G-strand can fold into stable contiguous G-quadruplexes (G4). In the present study we investigated how hPOT1, in complex with its shelterin partner TPP1, binds to telomeric sequences structured into contiguous G4 in potassium solutions. We observed that binding of multiple hPOT1–TPP1 preferentially proceeds from 3′ toward 5′. We explain this directionality in terms of two factors: (i) the preference of hPOT1–TPP1 for the binding site situated at the 3′ end of a telomeric sequence and (ii) the cooperative binding displayed by hPOT1–TPP1 in potassium. By comparing binding in K+ and in Li+, we demonstrate that this cooperative behaviour does not stem from protein-protein interactions, but from structuring of the telomeric DNA substrate into contiguous G4 in potassium. Our study suggests that POT1-TPP1, in physiological conditions, might preferentially cover the telomeric G-overhang starting from the 3′-end and proceeding toward 5′.  相似文献   

18.
19.
The annotated whole-genome sequence of Mycobacterium tuberculosis revealed the presence of a putative recD gene; however, the biochemical characteristics of its encoded protein product (MtRecD) remain largely unknown. Here, we show that MtRecD exists in solution as a stable homodimer. Protein-DNA binding assays revealed that MtRecD binds efficiently to single-stranded DNA and linear duplexes containing 5′ overhangs relative to the 3′ overhangs but not to blunt-ended duplex. Furthermore, MtRecD bound more robustly to a variety of Y-shaped DNA structures having ≥18-nucleotide overhangs but not to a similar substrate containing 5-nucleotide overhangs. MtRecD formed more salt-tolerant complexes with Y-shaped structures compared with linear duplex having 3′ overhangs. The intrinsic ATPase activity of MtRecD was stimulated by single-stranded DNA. Site-specific mutagenesis of Lys-179 in motif I abolished the ATPase activity of MtRecD. Interestingly, although MtRecD-catalyzed unwinding showed a markedly higher preference for duplex substrates with 5′ overhangs, it could also catalyze significant unwinding of substrates containing 3′ overhangs. These results support the notion that MtRecD is a bipolar helicase with strong 5′ → 3′ and weak 3′ → 5′ unwinding activities. The extent of unwinding of Y-shaped DNA structures was ∼3-fold lower compared with duplexes with 5′ overhangs. Notably, direct interaction between MtRecD and its cognate RecA led to inhibition of DNA strand exchange promoted by RecA. Altogether, these studies provide the first detailed characterization of MtRecD and present important insights into the type of DNA structure the enzyme is likely to act upon during the processes of DNA repair or homologous recombination.  相似文献   

20.
A family of multiple autonomously replicating sequences (ARSs) which are located at several chromosomal ends of Hansenula polymorpha DL-1 has been identified and characterized. Genomic Southern blotting with an ARS, HARS36, originating from the end of a chromosome, as a probe showed several homologues in the genome of H. polymorpha. Nucleotide sequences of the three fragments obtained by a selective cloning for chromosomal ends were nearly identical to that of HARS36. All three fragments harbored an ARS motif and ended with 18 to 23 identical repetitions of 5′-GGGTGGCG-3′ which resemble the telomeric repeat sequence in other eukaryotes. Transformation of H. polymorpha with nonlinearized plasmids containing the newly obtained telomeric ARSs almost exclusively resulted in the targeted integration of a single copy or multiple tandem copies of the plasmid into the chromosomes. The sensitivity to exonuclease Bal31 digestion of the common DNA fragment in all integrants confirmed the telomeric origin of HARS36 homologues, suggesting that several chromosomal ends, if not all of them, consisted of the same ARS motif and highly conserved sequences observed in HARS36. Even though the frequencies of targeted recombination were varied among the ends of the chromosomes, the overall frequency was over 96%. The results suggested that the integration of the plasmids containing telemeric ARSs occurred largely through homologous recombination at the telomeric repeats, which serve as high-frequency recombination targets.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号