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 31 毫秒
1.
A new restriction endonuclease from Streptomyces albus G.   总被引:10,自引:0,他引:10  
A restriction endonuclease, SalI, has been partially purified from Streptomyces albus G. This enzyme cleaves adenovirus-2 DNA at three sites, bacteriophage λ DNA at two sites, but does not cleave simian virus 40 DNA or φX174 DNA. It recognizes the sequence
and cuts at the sites indicated by the arrows. An endonuclease (XamI) with similar specificity has also been isolated from Xanthomonas amaranthicola.  相似文献   

2.
An additional sequence-specific endonuclease, XmaIII, has been partially purified from Xanthomonas malvacearum. XmaIII recognizes ten cleavage sites in adenovirus 2 DNA, two sites in bacteriophage lambda and no site in either simian virus 40 DNA or φX174 DNA. It recognizes the sequence
and cleaves at the sites indicated by the arrows. No other endonuclease with this particular nucleotide sequence specificity has been reported.  相似文献   

3.
A new specific endonuclease, XhoI, has been partially purified from Xanthomonas holcicola. This enzyme cleaves adenovirus-2 DNA at five sites, bacteriophage λ DNA at one site, φX174 DNA at one site, but does not cleave simian virus 40 DNA. It recognizes the sequence
and cuts at the sites indicated by the arrows. Enzymes with identical specificity have also been found in Xanthomonas papavericola and Brevibacterium luteum.  相似文献   

4.
The stimulatory effect of the aqueous extract of G. lucidum, a basidiomycetes class fungus in the APE1-enzyme-mediated processing of solitary and bistranded clustered abasic sites DNA damages is presented. Abasic sites are considered the most common type of DNA damage lesions. Our study shows enhanced activity of APE1 in the processing of abasic sites in the presence of the polysaccharides fraction of G. lucidum. Remarkable increase in the amount of single-strand breaks (SSBs) and double-strand breaks (DSBs) from solitary and bistranded clustered abasic sites respectively with APE1 in the presence of the extract was found. This trend is maintained when abasic sites in DNA oligomers are exposed to fibroblast cell extracts in the presence of the extract. While DNA conformational alteration is negligible, APE1 enzyme shows characteristic changes in the alpha helix and beta strand ratio after incubation with G. lucidum extract. The enhanced reactivity of APE1 at the molecular level in the presence of G. lucidium is attributed to this effect. This study potentially amplifies the scope of the use of G. lucidum, which was earlier shown to have only reactive oxygen species (ROS) scavenging properties with regards to DNA damage inhibition.
Graphical Abstract ?
  相似文献   

5.
6.
A specific endonuclease from Arthrobacter luteus.   总被引:20,自引:0,他引:20  
A new restriction-like endonuclease, AluI, has been partially purified from Arthrobacter luteus. This enzyme cleaves bacteriophage λ DNA, adenovirus-2 DNA and simian virus 40 DNA at many sites including all sites cleaved by the endonuclease HindIII from Haemophilus influenzae serotype d. Radioactive oligonucleotides in pancreatic DNAase digests of (5′-32P)-labelled fragments of phage λ DNA released by the action of AluI had the 5′ terminal sequence pC-T-N-. The enzyme recognises the tetranucleotide sequence
and cleaves it at the position marked by the arrows.  相似文献   

7.
8.
A targeting method to insert genes at a previously characterized genetic locus to make plant transformation and transgene expression predictable is highly desirable for plant biotechnology. We report the successful targeting of transgenes to predefined soybean (Glycine max) genome sites using the yeast FLP-FRT recombination system. First, a target DNA containing a pair of incompatible FRT sites flanking a selection gene was introduced in soybean by standard biolistic transformation. Transgenic events containing a single copy of the target were retransformed with a donor DNA, which contained the same pair of FRT sites flanking a different selection gene, and a FLP expression DNA. Precise DNA cassette exchange was achieved between the target and donor DNA via recombinase-mediated cassette exchange, so that the donor DNA was introduced at the locus previously occupied by the target DNA. The introduced donor genes expressed normally and segregated according to Mendelian laws.Plant transformation has challenges such as random integration, multiple transgene copies, and unpredictable expression. Homologous recombination (Iida and Terada, 2005; Wright et al., 2005) and DNA recombinase-mediated site-specific integration (SSI) are promising technologies to address the challenges for placing a single copy of transgenes into a precharacterized site in a plant genome.Several site-specific DNA recombination systems, such as the bacteriophage Cre-lox and the yeast FLP-FRT and R-RS, have been used in SSI studies (Ow, 2002; Groth and Calos, 2003). A common feature of these systems is that each system consists of a recombinase Cre, FLP, or R and two identical or similar palindromic recognition sites, lox, FRT, or RS. Each recognition site contains a short asymmetric spacer sequence where DNA strand exchange takes place, flanked by inverted repeat sequences where the corresponding recombinase specifically binds. If two recognition sites are located in cis on a DNA molecule, the DNA segment can be excised if flanked by two directionally oriented sites or inverted if flanked by two oppositely oriented sites. If two recognition sites are located in trans on two different DNA molecules, a reciprocal translocation can happen between the two DNA molecules or the two molecules can integrate if at least one of them is a circular DNA (Ow, 2002; Groth and Calos, 2003).Single-site SSI can integrate a circular donor DNA containing one recognition site into a similar site previously placed in a plant genome. The integrated transgene now flanked by two recognition sites is vulnerable to excision. Transient Cre expression and the use of mutant lox sites to create two less compatible sites after integration helped reduce the subsequent excision in tobacco (Nicotiana tabacum; Albert et al., 1995; Day et al., 2000). A similar approach was used to produce SSI events in rice (Oryza sativa), and the transgene was proven stably expressed over generations (Srivastava and Ow, 2001; Srivastava et al., 2004; Chawla et al., 2006). Using a promoter trap to displace a cre gene in the genome with a selection gene from the donor, approximately 2% SSI was achieved in Arabidopsis (Arabidopsis thaliana; Vergunst et al., 1998).When two recognition sites located on a linear DNA molecule are similar enough to be recognized by the same recombinase but different enough to reduce or prevent DNA recombination from happening between them, the DNA segment between the two sites may not be easily excised or inverted. When a circular DNA molecule carrying the same two incompatible sites is introduced, the circular DNA can integrate by the corresponding recombinase at either site on the linear DNA to create a collinear DNA with four recognition sites, two from the original linear DNA and two from the circular DNA. DNA excision can subsequently occur between any pair of compatible sites to restore the two original DNA molecules or to exchange the intervening DNA segments between the two DNA molecules. This process, termed recombinase-mediated cassette exchange (RMCE), can be employed to integrate transgenes directionally into predefined genome sites (Trinh and Morrison, 2000; Baer and Bode, 2001).RMCE using two oppositely oriented identical RS sites, a donor containing the R recombinase gene and a third RS site to limit random integration, resulted in cassette exchange between the donor and a previously placed target in tobacco (Nanto et al., 2005). RMCE using both the Cre-lox and FLP-FRT systems improved RMCE frequency in animal cell cultures (Lauth et al., 2002). RMCE using two directly oriented incompatible FRT sites and transiently expressed FLP recombinase achieved cassette exchange between a target previously placed in the Drosophila genome and a donor introduced as a circular DNA (Horn and Handler, 2005). A gene conversion approach involving Cre-lox- and FLP-FRT-mediated SSI, RMCE, and homologous recombination was explored in maize (Zea mays; Djukanovic et al., 2006). RMCE using two oppositely oriented incompatible lox sites and transiently expressed Cre recombinase produced single-copy RMCE plants in Arabidopsis (Louwerse et al., 2007).To develop FLP-FRT-mediated RMCE in soybean (Glycine max), we created transgenic target lines containing a hygromycin resistance gene flanked by two directly oriented incompatible FRT sites via biolistic transformation. Single-copy target lines were selected and retransformed with a donor DNA containing a chlorsulfuron resistance gene flanked by the same pair of FRT sites. An FLP expression DNA was cobombarded to transiently provide FLP recombinase. RMCE events were obtained from multiple target lines and confirmed by extensive molecular characterization.  相似文献   

9.
10.
The mobilization proteins of the broad-host-range plasmid R1162 can initiate conjugative transfer of a plasmid from a 19-bp locus that is partially degenerate in sequence. Such loci are likely to appear by chance in the bacterial chromosome and could act as cryptic sites for transfer of chromosomal DNA when R1162 is present. The R1162-dependent transfer of chromosomal DNA, initiated from one such potential site in Pectobacterium atrosepticum, is shown here. A second active site was identified in Escherichia coli, where it is also shown that large amounts of DNA are transferred. This transfer probably reflects the combined activity of the multiple cryptic origins in the chromosome. Transfer of chromosomal DNA due to the presence of a plasmid in the cytoplasm describes a previously unrecognized potential for the exchange of bacterial DNA.The discovery over 50 years ago of bacterial mating by Lederberg and Cavalli-Sforza (summarized by Cavalli-Sforza [10]) was a major step in the development of modern microbial genetics. It was later realized that chromosomal DNA transfer was due to the integration of a plasmid, the F factor. This plasmid is able to effect its own transfer and, in the integrated state, chromosomal DNA as well. A unique site on the plasmid, the origin of transfer (oriT), is essential for this process. A complex of plasmid- and host-encoded proteins assemble at oriT (15); the F-encoded relaxase then cleaves one of the DNA strands, forming a covalent, protein-DNA intermediate (27) that is delivered to the type IV secretion apparatus, also encoded by F (24). The possibility of low-level transfer initiating from chromosomal sites was raised early in the history of conjugation (12). However, it is now clear that the exacting architecture of the oriT DNA-protein complex, both for the F factor and related oriTs, results in a very high degree of sequence and structural specificity, and no secondary origins in the chromosome have been identified.The broad-host-range plasmid R1162 (RSF1010), like the F factor, is also transferred from cell to cell. Although the mechanism of transfer is similar for the two plasmids, the oriTs are structurally very different. In prior studies we have determined that the R1162 oriT is small, structurally simple, and able to accommodate base changes at different positions without a complete loss of function (5, 21). As a result of this relaxed specificity, the R1162 mobilization (Mob) proteins can activate the related but different oriT of another plasmid, pSC101 (28).The R1162 oriT is shown in Fig. Fig.1.1. The R1162 relaxase MobA interacts with the core region, highly conserved in the R1162 Mob family (5), and the adjacent, inner arm of the inverted repeat (23). The protein forms at nic a tyrosyl phosphodiester linkage with the 5′ end of the DNA strand (30), which is then unwound from its complement as the protein-DNA complex is passed into the recipient cell. Circular plasmid DNA is reformed by a reverse of the initial protein-DNA transesterification, with the relaxase now binding to the 3′ end of the transferred DNA. This binding requires the complete inverted repeat, which probably forms a hairpin loop to recreate the double-stranded character of the relaxase binding site on duplex DNA.Open in a separate windowFIG. 1.Base sequence of R1162 oriT on the nicked strand. The relaxase cleaves at nic; subsequent transfer is 5′ to 3′ so that most of oriT is transferred last. Below is the consensus sequence of DNA active for initiation of transfer.An important feature of the steps in DNA processing at the R1162 oriT is that an inverted repeat is not required for the initiation of transfer or passage of DNA into a new cell. As a result of this and the permissiveness of the relaxase to base changes within oriT, different sites in the chromosome that are not part of a plasmid oriT might nevertheless be capable of initiating transfer when cloned into a plasmid. We previously identified one such site in the chromosome of Erwinia carotovora subsp. atroseptica (now Pectobacterium atrosepticum) (21). In addition, by testing libraries of oriTs with one or more mutations for relaxase-induced nicking, we found that a large population of DNAs with different sequences was potentially active (21). The consensus sequence for activity, derived from these studies, is shown in Fig. Fig.11.I show here that ectopic sites on the bacterial chromosome, active for the initiation of transfer of plasmid DNA, can also serve for the transfer of chromosomal DNA when the Mob proteins are provided in trans. Thus, the presence of R1162 in the cell can result in a cryptic sexuality in bacteria, resulting in the transfer of chromosomal genes. Although such events are likely to be infrequent, the broad-host-range of R1162 and its close relatives, with their ability to reside stably in the cytoplasms of many different bacteria, could be another source of gene exchange for a variety of different species.  相似文献   

11.
DNA Bending by AraC: a Negative Mutant   总被引:1,自引:1,他引:0       下载免费PDF全文
  相似文献   

12.
In addition to their capacity for template-directed 5′ → 3′ DNA synthesis at the polymerase (pol) site, DNA polymerases have a separate 3′ → 5′ exonuclease (exo) editing activity that is involved in assuring the fidelity of DNA replication. Upon misincorporation of an incorrect nucleotide residue, the 3′ terminus of the primer strand at the primer-template (P/T) junction is preferentially transferred to the exo site, where the faulty residue is excised, allowing the shortened primer to rebind to the template strand at the pol site and incorporate the correct dNTP. Here we describe the conformational changes that occur in the primer strand as it shuttles between the pol and exo sites of replication-competent Klenow and Klentaq DNA polymerase complexes in solution and use these conformational changes to measure the equilibrium distribution of the primer between these sites for P/T DNA constructs carrying both matched and mismatched primer termini. To this end, we have measured the fluorescence and circular dichroism spectra at wavelengths of >300 nm for conformational probes comprising pairs of 2-aminopurine bases site-specifically replacing adenine bases at various positions in the primer strand of P/T DNA constructs bound to DNA polymerases. Control experiments that compare primer conformations with available x-ray structures confirm the validity of this approach. These distributions and the conformational changes in the P/T DNA that occur during template-directed DNA synthesis in solution illuminate some of the mechanisms used by DNA polymerases to assure the fidelity of DNA synthesis.Escherichia coli DNA polymerase (DNAP)2 I is a single subunit polymerase that is organized into three functional domains: an N-terminal domain that is associated with 5′ → 3′ exonuclease activity, an intermediate domain that carries the 3′ → 5′ proofreading activity, and a C-terminal domain that is associated with the 5′ → 3′ template-directed polymerization activity. An important role of DNAP I is to remove the RNA primers of the Okazaki fragments formed during lagging strand DNA synthesis in E. coli replication and to fill in the resulting gaps by template-directed DNA synthesis (1). An N-terminal deletion mutant of DNAP I, known as the “large fragment” or Klenow form of the enzyme, contains only the polymerase (pol) and the 3′ → 5′ exonuclease (exo) domains. The Klenow polymerase has served and continues to serve as an excellent model system for isolating and defining general structure-function relationships in polymerases and in the supporting machinery of DNA replication.The main function of the 3′ → 5′ exonuclease activity of DNAP I is to remove misincorporated nucleotide residues from the 3′-end of the primer (2), thus contributing significantly to the overall fidelity of DNA replication (3). Contrary to initial expectations, crystallographic studies showed that the pol and exo active sites are quite far apart in replication polymerases, about 30 Å in Klenow (4). As a consequence, the ability of polymerases to “shuttle” the 3′-end of the primer strand efficiently between the pol and the exo sites in order to rectify misincorporation events during polymerization is critical to maintaining the overall accuracy of template-directed replication. Elucidation of the mechanisms of this shuttling and determination of the factors that control the rates (and equilibria) of the active site switching reaction will certainly increase our understanding of fidelity control by DNA polymerases.An early crystallographic study of the Klenow polymerase complexed with fully paired primer-template (P/T) DNA revealed that 3–4 nt of the 3′-primer terminus had been unwound from the template stand and partitioned into the exo site and that an extended single-stranded DNA (ssDNA) binding pocket of the exo site appeared to make position-specific hydrophobic contacts with the unstacked bases at the 3′-end of the primer (4). A separate crystallographic study of an editing complex confirmed that an ssDNA fragment 4 nt in length was bound at the exo site in the same conformation as seen for the single-stranded 3′-primer sequence unwound from P/T DNA (5). A structure of Klenow polymerase with the DNA bound at the pol site has not yet been reported, although such structures have been obtained for other homologous polymerases, including Klentaq (the “large fragment” of Thermus aquaticus (Taq) DNAP), Bacillus stearothermophilus (Bst) “large fragment” polymerase, and the T7 DNAP (68), all of which are members of the polymerase family that includes Klenow.The amino acid residues involved in the binding of DNA at the pol site in these polymerases (determined from co-crystal structures) and those of Klenow (determined by site-directed mutagenesis studies (9, 10)) are highly conserved, suggesting that a similar DNA binding mode at the pol site may apply to all of the DNAP I polymerases. The crystal structure of Klenow revealed that the polymerization domain has a shape reminiscent of a right hand in which the palm, fingers, and thumb domains form the DNA-binding crevice. Structural studies with various DNAP I polymerases in the presence of P/T DNA constructs yielded an “open” binary complex, whereas the addition of the next correct dNTP (as a chain-terminating dideoxy-NTP) resulted in the formation of a catalytically competent “closed” ternary complex (68). In the latter complex, the 3′-primer terminus was base-paired with the template DNA, and the templating base was poised for incorporation of the next correct nucleotide. These structures showed that the conformation of the DNA primer terminus bound at the pol site is markedly different from that of the “frayed open” primer observed at the exo site in Klenow (4, 5).Although crystallographic studies have provided a wealth of information about the conformations of the DNA substrates bound at the active sites of DNAP, replication itself is a dynamic process (reviewed in Ref. 11), and it is critical to be able to distinguish between various forms of DNA-polymerase complexes in solution in order to fully understand the mechanistic details of the replication process. A solution approach used by Millar and co-workers (reviewed in Ref. 12) for studying the conformation of DNA in these complexes involved measuring the time-resolved fluorescence anisotropy properties of a dansyl fluorophore attached to a DNA base located 8 bp upstream of the P/T DNA junction. The changes in the lifetime of the fluorophore, which appeared to depend mostly on the local environment occupied by the probe within the protein (i.e. buried versus partially exposed), were correlated with specific binding conformations of the primer to provide an estimate of the fractional occupancy of the pol and the exo sites. Reha-Krantz and co-workers (13) more recently used a related approach, here involving the monitoring of changes in the fluorescent lifetimes of a single 2-aminopurine (2-AP) base (a fluorescent analogue of adenine) site-specifically substituted in the template strand at the P/T junction, to make similar fractional occupancy measurements. However, we note that structural interpretations of these fluorescence experiments relied heavily on the available crystal structures, and it remained to be shown directly that the 3′-end of the primer in P/T DNA constructs assumes the same distribution of conformations when bound to the protein in solution.To get around this problem, as well as to directly investigate the conformations of the primer DNA in both active sites of the Klenow and Klentaq polymerases, we have used a novel CD spectroscopic approach to characterize the solution conformations of primer DNA bound to Klenow and Klentaq DNAPs. Previously, we had shown that CD spectroscopy, in conjunction with fluorescence measurements, can be used to examine changes in local DNA and RNA conformations at 2-AP dimer probes inserted at specified positions within the nucleic acid frameworks of a variety of macromolecular machines functioning in solution (1416). 2-AP is a structural isomer of adenine that forms base pairs with thymine in DNA (and uridine in RNA), and the substitution of 2-AP for adenine in such bp does not significantly perturb the structure or stability of the resultant double helix. Furthermore, when these probes are used as dimer pairs, the CD spectrum primarily reflects the interaction of the transition dipoles of the two probes themselves and thus the local conformation of the DNA at those positions within the P/T DNA. The characteristic CD and fluorescence signals for 2-AP probes in nucleic acids occur at wavelengths of >300 nm, a spectral region in which the protein and the canonical nucleic acid components of the “macromolecular machines of gene expression” are otherwise transparent. In this study, we have examined the binding of Klenow and Klentaq polymerases to P/T DNA constructs that were designed to be comparable with the nucleic acid components of functioning replication complexes. By examining the low energy CD spectra of site-specifically placed 2-AP probes, we have been able to characterize base conformations at defined positions within the DNA to reveal conformational features of specific DNA bases bound at and near both the pol and the exo active sites of these polymerases. These measurements, in that they directly reflect the actual conformations of the DNA chains bound within the active sites of the functioning polymerase, have also provided a direct means to estimate the equilibrium distributions of primer ends between the two active sites for various P/T DNA constructs.  相似文献   

13.
Upon DNA replication stress, stalled DNA replication forks serve as a platform to recruit many signaling proteins, leading to the activation of the DNA replication checkpoint. Activation of Rad53, a key effector kinase in the budding yeast Saccharomyces cerevisiae, is essential for stabilizing DNA replication forks during replication stress. Using an activity-based assay for Rad53, we found that Mrc1, a replication fork-associated protein, cooperates with Mec1 to activate Rad53 directly. Reconstitution of Rad53 activation using purified Mec1 and Mrc1 showed that the addition of Mrc1 stimulated a more than 70-fold increase in the ability of Mec1 to activate Rad53. Instead of increasing the catalytic activity of Mec1, Mrc1 was found to facilitate the phosphorylation of Rad53 by Mec1 via promotion of a stronger enzyme-substrate interaction between them. Further, the conserved C-terminal domain of Mrc1 was found to be required for Rad53 activation. These results thus provide insights into the role of the adaptor protein Mrc1 in activating Rad53 in the DNA replication checkpoint.Faithful replication of the genome is important for the survival of all organisms. During DNA replication, replication stress can arise from a variety of situations, including intrinsic errors made by DNA polymerases, difficulties in replicating repeated DNA sequences, and failures to repair damaged DNA caused by either endogenous oxidative agents or exogenous mutagens such as UV light and DNA-damaging chemicals (13). In eukaryotes, there is an evolutionarily conserved DNA replication checkpoint that becomes activated in response to DNA replication stress. It helps to stabilize DNA replication forks, block late replication origin firing, and delay mitosis and ultimately helps recovery from stalled replication forks after DNA repair (47). Defects in the DNA replication checkpoint could result in elevated genomic instabilities, cancer development, or cell death (8, 9).Aside from replicating the genome, the DNA replication forks also provide a platform to assemble many signaling proteins that function in the DNA replication checkpoint. In the budding yeast Saccharomyces cerevisiae, Mec1, an ortholog of human ATR,2 is a phosphoinositide 3-kinase-like kinase (PIKK) involved in sensing stalled DNA replication forks. Mec1 forms a protein complex with Ddc2 (ortholog of human ATRIP). The Mec1-Ddc2 complex is recruited to stalled replication forks through replication protein A (RPA)-coated single-stranded DNA (10, 11). The Mec3-Rad17-Ddc1 complex, a proliferating cell nuclear antigen (PCNA)-like checkpoint clamp and ortholog of the human 9-1-1 complex, was shown to be loaded onto the single- and double-stranded DNA junction of the stalled replication forks by the clamp loader Rad24-RFC complex (12). Once loaded, the Mec3-Rad17-Ddc1 complex stimulates Mec1 kinase activity (13). Dbp11 and its homolog TopBP1 in vertebrates are known components of the replication machinery (14). In addition to regulating the initiation of DNA replication, they were found to play a role in the DNA replication checkpoint (1517). They interact with the 9-1-1 complex and directly stimulate Mec1/ATR activity in vitro (1820). Thus, the assembly of multiple protein complexes at stalled DNA replication forks appears to facilitate activation of the DNA replication checkpoint (13, 18).Mrc1 (for mediator of replication checkpoint) was originally identified to be important for cells to respond to hydroxyurea in S. cerevisiae and Schizosaccharomyces pombe (21, 22). Mrc1 is a component of the DNA replisome and travels with the replication forks along chromosome during DNA synthesis (2325). Deletion of MRC1 causes defects in DNA replication, indicating its role in the normal progression of DNA replication (23). Interestingly, when DNA replication is blocked by hydroxyurea, Mrc1 undergoes Mec1- and Rad3 (S. pombe ortholog of Mec1)-dependent phosphorylation (21, 22). In S. cerevisiae, mutations of Mrc1 at the (S/T)Q sites, which are consensus phosphorylation sites of the Mec1/ATR family kinases, abolishes hydroxyurea-induced Mrc1 phosphorylation in vivo, suggesting a direct phosphorylation of Mrc1 by Mec1 (21, 22).Rad53 and Cds1, homologs of human Chk2, are the major effector kinases in the DNA replication checkpoints in S. cerevisiae and S. pombe, respectively. Activation of Rad53 is a hallmark of DNA replication checkpoint activation and is important for the maintenance of DNA replication forks in response to DNA replication stress (5, 6). Thus, it is important to understand how Rad53 activity is controlled. Interestingly, mutation of all the (S/T)Q sites of Mrc1 not only abolishes the phosphorylation of Mrc1 by Mec1 but also compromises hydroxyurea-induced Rad53 activation in S. cerevisiae (21). Similarly, mutation of the TQ sites of Mrc1 in S. pombe was shown to abolish the binding between Cds1 and Mrc1 as well as Cds1 activation (22). Further, mutation of specific TQ sites of Mrc1 in S. pombe abolishes its binding to Cds1 in vitro and the activation of Cds1 in vivo (26). Thus, Mec1/Rad3-dependent phosphorylation of Mrc1 is responsible for Mrc1 binding to Rad53/Cds1, which is essential for Rad53/Cds1 activation.An intriguing property of the Chk2 family kinases is their ability to undergo autophosphorylation and activation in the absence of other proteins in vitro (27, 28). First, autophosphorylation of a conserved threonine residue in the activation loop of Chk2 family kinase was found to be an essential part of their activation processes (26, 2931). Second, a direct and trans-phosphorylation of the N-terminal TQ sites of the Chk2 family kinases by the Mec1/ATR family kinases is also important for their activation in vivo. Analogous to the requirement of N-terminal TQ site phosphorylation of Chk2 by ATR in human (32), the activation of Rad53/Cds1 in vivo requires phosphorylation of TQ sites in their N termini by Mec1/Rad3 (33, 34).Considering that Mec1, Mrc1, and many other proteins are recruited at stalled DNA replication forks and have been shown to be involved in DNA replication checkpoint activation, a key question remains unresolved: what is the minimal system that is capable of activating Rad53 directly? Given the direct physical interaction between Mrc1 and Rad53 and the requirement of Mrc1 and Mec1 in vivo, it is likely that they both play a role in Rad53 activation. Furthermore, what is the molecular mechanism of Rad53 activation by its upstream activators? To address these questions, a faithful reconstitution of the activation of Rad53 using purified proteins is necessary. In this study, we developed an activity-based assay consisting of the Dun1 kinase, a downstream substrate of Rad53, and Sml1, as a substrate of Dun1, to quantitatively measure the activity of Rad53. Using this coupled kinase assay from Rad53 to Dun1 and then to Sml1, we screened for Mrc1 and its associated factors to see whether they could directly activate Rad53 in vitro. Our results showed that Mec1 and Mrc1 collaborate to constitute a minimal system in direct activation of Rad53.  相似文献   

14.
15.
16.
17.
18.
19.
Abasic (AP) sites are very frequent and dangerous DNA lesions. Their ability to block the advancement of a replication fork has been always viewed as a consequence of their inhibitory effect on the DNA synthetic activity of replicative DNA polymerases (DNA pols). Here we show that AP sites can also affect the strand displacement activity of the lagging strand DNA pol δ, thus preventing proper Okazaki fragment maturation. This block can be overcome through a polymerase switch, involving the combined physical and functional interaction of DNA pol β and Flap endonuclease 1. Our data identify a previously unnoticed deleterious effect of the AP site lesion on normal cell metabolism and suggest the existence of a novel repair pathway that might be important in preventing replication fork stalling.Loss of purine and pyrimidine bases is a significant source of DNA damage in prokaryotic and eukaryotic organisms. Abasic (apurinic and apyrimidinic) lesions occur spontaneously in DNA; in eukaryotes it has been estimated that about 104 depurination and 102 depyrimidation events occur per genome per day. An equally important source of abasic DNA lesions results from the action of DNA glycosylases, such as uracil glycosylase, which excises uracil arising primarily from spontaneous deamination of cytosines (1). Although most AP sites are removed by the base excision repair (BER)5 pathway, a small fraction of lesions persists, and DNA with AP lesions presents a strong block to DNA synthesis by replicative DNA polymerases (DNA pols) (2, 3). Several studies have been performed to address the effects of AP sites on the template DNA strand on the synthetic activity of a variety of DNA pols. The major replicative enzyme of eukaryotic cells, DNA pol δ, was shown to be able to bypass an AP lesion, but only in the presence of the auxiliary factor proliferating cell nuclear antigen (PCNA) and at a very reduced catalytic efficiency if compared with an undamaged DNA template (4). On the other hand, the family X DNA pols β and λ were shown to bypass an AP site but in a very mutagenic way (5). Recent genetic evidence in Saccharomyces cerevisiae cells showed that DNA pol δ is the enzyme replicating the lagging strand (6). According to the current model for Okazaki fragment synthesis (79), the action of DNA pol δ is not only critical for the extension of the newly synthesized Okazaki fragment but also for the displacement of an RNA/DNA segment of about 30 nucleotides on the pre-existing downstream Okazaki fragment to create an intermediate Flap structure that is the target for the subsequent action of the Dna2 endonuclease and the Flap endonuclease 1 (Fen-1). This process has the advantage of removing the entire RNA/DNA hybrid fragment synthesized by the DNA pol α/primase, potentially containing nucleotide misincorporations caused by the lack of a proofreading exonuclease activity of DNA pol α/primase. This results in a more accurate copy synthesized by DNA pol δ. The intrinsic strand displacement activity of DNA pol δ, in conjunction with Fen-1, PCNA, and replication protein A (RP-A), has been also proposed to be essential for the S phase-specific long patch BER pathway (10, 11). Although it is clear that an AP site on the template strand is a strong block for DNA pol δ-dependent synthesis on single-stranded DNA, the functional consequences of such a lesion on the ability of DNA pol δ to carry on strand displacement synthesis have never been investigated so far. Given the high frequency of spontaneous hydrolysis and/or cytidine deamination events, any detrimental effect of an AP site on the strand displacement activity of DNA pol δ might have important consequences both for lagging strand DNA synthesis and for long patch BER. In this work, we addressed this issue by constructing a series of synthetic gapped DNA templates with a single AP site at different positions with respect to the downstream primer to be displaced by DNA pol δ (see Fig. 1A). We show that an AP site immediately upstream of a single- to double-strand DNA junction constitutes a strong block to the strand displacement activity of DNA pol δ, even in the presence of RP-A and PCNA. Such a block could be resolved only through a “polymerase switch” involving the concerted physical and functional interaction of DNA pol β and Fen-1. The closely related DNA pol λ could only partially substitute for DNA pol β. Based on our data, we propose that stalling of a replication fork by an AP site not only is a consequence of its ability to inhibit nucleotide incorporation by the replicative DNA pols but can also stem from its effects on strand displacement during Okazaki fragment maturation. In summary, our data suggest the existence of a novel repair pathway that might be important in preventing replication fork stalling and identify a previously unnoticed deleterious effect of the AP site lesion on normal cell metabolism.Open in a separate windowFIGURE 1.An abasic site immediately upstream of a double-stranded DNA region inhibits the strand displacement activity of DNA polymerase δ. The reactions were performed as described under “Experimental Procedures.” A, schematic representation of the various DNA templates used. The size of the resulting gaps is indicated in nt. The position of the AP site on the 100-mer template strand is indicated relative to the 3′ end. Base pairs in the vicinity of the lesion are indicated by dashes. The size of the gaps (35–38 nt) is consistent with the size of ssDNA covered by a single RP-A molecule, which has to be released during Okazaki fragment synthesis when the DNA pol is approaching the 5′-end of the downstream fragment. When the AP site is covered by the downstream terminator oligonucleotide (Gap-3 and Gap-1 templates) the nucleotide placed on the opposite strand is C to mimic the situation generated by spontaneous loss of a guanine or excision of an oxidized guanine, whereas when the AP site is covered by the primer (nicked AP template), the nucleotide placed on the opposite strand is A to mimic the most frequent incorporation event occurring opposite an AP site. B, human PCNA was titrated in the presence of 15 nm (lanes 2–4 and 10–12) or 30 nm (lanes 6–8 and 14–16) recombinant human four subunit DNA pol δ, on a linear control (lanes 1–8) or a 38-nt gap control (lanes 9–16) template. Lanes 1, 5, 9, and 13, control reactions in the absence of PCNA. C, human PCNA was titrated in the presence of 60 nm DNA pol δ, on a linear AP (lanes 2–4) or 38-nt gap AP (lanes 6–9) template. Lanes 1 and 5, control reactions in the absence of PCNA.  相似文献   

20.
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