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 8 毫秒
1.
Many of the proteins that operate at the replication fork in Escherichia coli have been defined genetically. These include some of the subunits of the DNA polymerase III holoenzyme, the DnaB replication fork helicase, and the DnaG primase. The multiprotein primosome (which includes the DnaB and DnaG proteins), defined biochemically on the basis of its requirement during bacteriophage phi X174 complementary-strand synthesis, could serve as the helicase-primase replication machine on the lagging-strand template. In order to determine if this is the case, we have begun an investigation of the phenotypes of mutants with mutations priA, priB, and priC, which encode the primosomal proteins factor Y (protein n'), n, and n", respectively. Inactivation of priA by insertional mutagenesis resulted in the induction of the SOS response, as evinced by induction of a resident lambda prophage, extreme filamentation, and derepression of an indicator operon in which beta-galactosidase production was controlled by the dinD1 promoter. In addition, the copy numbers of resident pBR322 plasmids were reduced four- to fivefold in these strains, and production of phi X174 phage was delayed considerably. These results are discussed in the context of existing models for SOS induction and possible roles for the PriA protein at the replication fork in vivo.  相似文献   

2.
3.
Having one electron with unpaired spin, nitric oxide (NO) shows high reactivity and activates or inhibits free radical chain reactions. NO toxic and genotoxic effects appear to be the result of intracellular formation of peroxinitrite that can induce some cellular damages, including DNA strand breaks, DNA base oxidation, destruction of the key enzymes, etc. Taking into account the character of DNA damages being formed under NO activity, we proposed a formation of the SOS signal and induction the SOS DNA repair response in E. coli cells treated with NO physiological donors--DNIC and GSNO. The ability of NO donor compounds to induce the SOS DNA response in E. coli PQ37 with sfiA::lacZ operon fusion is reported here at the first time. So, the SOS DNA repair response induction is one of the function of nitric oxide.  相似文献   

4.
Abstract The 3D gene of foot-and-mouth disease virus encodes the viral RNA dependent RNA polymerase, also called virus infection associated (VIA) antigen, which is the most important serological marker of virus infection. This 3D gene from a serotype Cl virus has been cloned and overexpressed in Escherichia coli under the control of the strong lambda lytic promoters. The resulting 51 kDa recombinant protein has been shown to be immunoreactive with sera from infected animals. After induction of gene expression, an immediate and dramatic arrest of cell DNA synthesis occurs, similar to that produced by genotoxic doses of the drug mitomycin C. This effect does not occur during the production of either a truncated VIA antigen or other related and non-related viral proteins. The inhibition of DNA replication results in a subsequent induction of the host SOS DNA-repair response and in an increase of the mutation frequency in the surviving cells.  相似文献   

5.
P L Moreau 《Biochimie》1985,67(3-4):353-356
The RecA protein of Escherichia coli plays a central role in DNA repair mechanisms. When it is incubated with single-stranded DNA and a nucleoside triphosphate, the purified RecA protein acts both by promoting cleavage of the LexA protein, the repressor of the SOS genes, and by catalyzing strand exchange between a variety of DNA molecules. A model for the regulation of the activity of the RecA protein in a cell exposed to a DNA damaging treatment is proposed.  相似文献   

6.
RecF, RecO and RecR, three of the important proteins of the RecF pathway of recombination, are also needed for repair of DNA damage due to UV irradiation. recF mutants are not proficient in cleaving LexA repressor in vivo following DNA damage; therefore they show a delay of induction of the SOS response. In this communication, by measuring the in vivo levels of LexA repressor using anti-LexA antibodies, we show that recO and recR mutant strains are also not proficient in LexA cleavage reactions. In addition, we show that recO and recR mutations delay induction of β-galactosidase activity expressed from a lexA-regulated promoter following exposure of cells to UV, thus further supporting the idea that recF, recO and recR gene products are needed for induction of the SOS response.  相似文献   

7.
8.
Microcin B17 is a novel peptide antibiotic of low Mr (about 4000) produced by Escherichia coli strains carrying plasmid pMccB17. The action of this microcin in sensitive cells is essentially irreversible, follows single-hit kinetics, and leads to an abrupt arrest of DNA replication and, consequently, to the induction of the SOS response. RecA- and RecBC- strains are hypersensitive to microcin B17. Strains producing a non-cleavable SOS repressor (lexAl mutant) are also more sensitive than wild-type, whereas strains carrying a mutation which causes constitutive expression of the SOS response (spr-55) are less sensitive to microcin. Microcin B17 does not induce the SOS response in cells which do not have an active replication fork. The results suggest that the mode of action of this microcin is different from all other well-characterized microcins and colicins, and from other antibiotics which inhibit DNA replication.  相似文献   

9.
Constitutive expression of the SOS regulon in Escherichia coli recA730 strains leads to a mutator phenotype (SOS mutator) that is dependent on DNA polymerase V (umuDC gene product). Here we show that a significant fraction of this effect also requires DNA polymerase IV (dinB gene product).  相似文献   

10.
Chromosomal DNA is exposed to continuous damage and repair. Cells contain a number of proteins and specific DNA repair systems that help maintain its correct structure. The SOS response was the first DNA repair system described in Escherichia coli induced upon treatment of bacteria with DNA damaging agents arrest DNA replication and cell division. Induction of the SOS response involves more than forty independent SOS genes, most of which encode proteins engaged in protection, repair, replication, mutagenesis and metabolism of DNA. Under normal growth conditions the SOS genes are expressed at a basal level, which increases distinctly upon induction of the SOS response. The SOS-response has been found in many bacterial species (e.g., Salmonella typhimurium, Caulobacter crescentus, Mycobacterium tuberculosis), but not in eukaryotic cells. However, species from all kingdoms contain some SOS-like proteins taking part in DNA repair that exhibit amino acid homology and enzymatic activities related to those found in E. coli. but are not organized in an SOS system. This paper presents a brief up-to-date review describing the discovery of the SOS system, the physiology of SOS induction, methods for its determination, and the role of some SOS-induced genes.  相似文献   

11.
The SOS response is readily triggered by replication fork stalling caused by DNA damage or a dysfunctional replicative apparatus in Escherichia coli cells. E. coli dinB encodes DinB DNA polymerase and its expression is upregulated during the SOS response. DinB catalyzes translesion DNA synthesis in place of a replicative DNA polymerase III that is stalled at a DNA lesion. We showed previously that DNA replication was suppressed without exogenous DNA damage in cells overproducing DinB. In this report, we confirm that this was due to a dose-dependent inhibition of ongoing replication forks by DinB. Interestingly, the DinB-overproducing cells did not significantly induce the SOS response even though DNA replication was perturbed. RecA protein is activated by forming a nucleoprotein filament with single-stranded DNA, which leads to the onset of the SOS response. In the DinB-overproducing cells, RecA was not activated to induce the SOS response. However, the SOS response was observed after heat-inducible activation in strain recA441 (encoding a temperature-sensitive RecA) and after replication blockage in strain dnaE486 (encoding a temperature-sensitive catalytic subunit of the replicative DNA polymerase III) at a non-permissive temperature when DinB was overproduced in these cells. Furthermore, since catalytically inactive DinB could avoid the SOS response to a DinB-promoted fork block, it is unlikely that overproduced DinB takes control of primer extension and thus limits single-stranded DNA. These observations suggest that DinB possesses a feature that suppresses DNA replication but does not abolish the cell's capacity to induce the SOS response. We conclude that DinB impedes replication fork progression in a way that does not activate RecA, in contrast to obstructive DNA lesions and dysfunctional replication machinery.  相似文献   

12.
Expression of the site-specific adenine methylase HhaII (GmeANTC, where me is methyl) or PstI (CTGCmeAG) induced the SOS DNA repair response in Escherichia coli. In contrast, expression of methylases indigenous to E. coli either did not induce SOS (EcoRI [GAmeATTC] or induced SOS to a lesser extent (dam [GmeATC]). Recognition of adenine-methylated DNA required the product of a previously undescribed gene, which we named mrr (methylated adenine recognition and restriction). We suggest that mrr encodes an endonuclease that cleaves DNA containing N6-methyladenine and that DNA double-strand breaks induce the SOS response. Cytosine methylases foreign to E. coli (MspI [meCCGG], HaeIII [GGmeCC], BamHI [GGATmeCC], HhaI [GmeCGC], BsuRI [GGmeCC], and M.Spr) also induced SOS, whereas one indigenous to E. coli (EcoRII [CmeCA/TGG]) did not. SOS induction by cytosine methylation required the rglB locus, which encodes an endonuclease that cleaves DNA containing 5-hydroxymethyl- or 5-methylcytosine (E. A. Raleigh and G. Wilson, Proc. Natl. Acad. Sci. USA 83:9070-9074, 1986).  相似文献   

13.
Survival and induction of the SOS system by 5-azacytidine, an analog of cytidine, were studied in Escherichia coli K-12. This compound did not produce any effect on the viability of dcm and dam dcm mutants. Furthermore, recA430 and lexA1 strains (both mutations interfere with LexA repressor cleavage but not recombination proficiency) were more resistant than the wild-type strain of E. coli K-12. In contrast, recBC and recA13 mutants were more sensitive to 5-azacytidine than the wild type. Transient exposure of E. coli to 5-azacytidine for 60 min induced both recA-dependent inhibition of cell division and induction of lambda prophage in Dcm+ strains but not in Dcm- mutants. Expression of both functions was dependent on recBC exonuclease. On the other hand, 5-azacytidine was unable to trigger the induction of umuCD and mucB genes and no amplification of RecA protein synthesis in either Dcm+ or Dcm- strains was observed. These last results are in agreement with previously reported data suggesting that there is a discrimination in the expression of the several SOS functions and that some SOS genes may be induced without amplification of RecA protein synthesis.  相似文献   

14.
We have investigated the possible role of Escherichia coli DNA polymerase (Pol) I in chromosomal replication fidelity. This was done by substituting the chromosomal polA gene by the polAexo variant containing an inactivated 3′→5′ exonuclease, which serves as a proofreader for this enzyme's misinsertion errors. Using this strain, activities of Pol I during DNA replication might be detectable as increases in the bacterial mutation rate. Using a series of defined lacZ reversion alleles in two orientations on the chromosome as markers for mutagenesis, 1.5‐ to 4‐fold increases in mutant frequencies were observed. In general, these increases were largest for lac orientations favouring events during lagging strand DNA replication. Further analysis of these effects in strains affected in other E. coli DNA replication functions indicated that this polAexo mutator effect is best explained by an effect that is additive compared with other error‐producing events at the replication fork. No evidence was found that Pol I participates in the polymerase switching between Pol II, III and IV at the fork. Instead, our data suggest that the additional errors produced by polAexo are created during the maturation of Okazaki fragments in the lagging strand.  相似文献   

15.
Replication of plasmid pBR322 in Escherichia coli cells normally requires RNA synthesis and thus is sensitive to rifampin, an inhibitor of RNA polymerase. In cells induced for the SOS response, however, derivatives of pBR322 were found to replicate in the presence of rifampin. This rifampin-resistant replication of pBR322 requires the insertion of certain sequences of DNA. The replication depends on recF+ and DNA polymerase I.  相似文献   

16.
Role of DNA polymerase II in repair replication in Escherichia coli   总被引:11,自引:0,他引:11  
  相似文献   

17.
ATP hydrolysis during SOS induction in Escherichia coli.   总被引:2,自引:1,他引:1  
Changes in cellular ATP concentration during SOS induction in strains of Escherichia coli with different levels of RecA and LexA proteins were studied. UV irradiation of RecA+ strains induced a twofold increase in the ATP concentration around the first 20 min, followed by a decrease to the values of nonirradiated cells. On the other hand, mutants defective in RecA protein or with either deficient RecA protease activity or cleavage-resistant LexA repressor did not show any decrease, suggesting that ATP consumption is related to LexA repressor hydrolysis. Furthermore, strains presenting a constitutive synthesis of RecA protein showed the same changes in ATP concentration as the wild-type strain. Likewise, the presence in a RecA+ strain of a LexA(Def) protein, which is defective in its capacity for binding specifically to SOS operators, did not disturb the changes in ATP when compared with the LexA+ RecA+ strain. Moreover, after UV irradiation, a LexA(Def) RecA- double mutant showed an important increase in ATP concentration, which remained elevated for at least 120 min after UV treatment.  相似文献   

18.
Inhibition of DNA replication with hydroxyurea during thymine starvation of Escherichia coli shows that active DNA synthesis is not required for thymineless death (TLD). Hydroxyurea experiments and thymine starvation of lexA3 and uvrA DNA repair mutants rule out unbalanced growth, the SOS response, and nucleotide excision repair as explanations for TLD.  相似文献   

19.
DNA polymerase I (PolI) functions both in nucleotide excision repair (NER) and in the processing of Okazaki fragments that are generated on the lagging strand during DNA replication. Escherichia coli cells completely lacking the PolI enzyme are viable as long as they are grown on minimal medium. Here we show that viability is fully dependent on the presence of functional UvrA, UvrB, and UvrD (helicase II) proteins but does not require UvrC. In contrast, delta polA cells grow even better when the uvrC gene has been deleted. Apparently UvrA, UvrB, and UvrD are needed in a replication backup system that replaces the PolI function, and UvrC interferes with this alternative replication pathway. With specific mutants of UvrC we could show that the inhibitory effect of this protein is related to its catalytic activity that on damaged DNA is responsible for the 3' incision reaction. Specific mutants of UvrA and UvrB were also studied for their capacity to support the PolI-independent replication. Deletion of the UvrC-binding domain of UvrB resulted in a phenotype similar to that caused by deletion of the uvrC gene, showing that the inhibitory incision activity of UvrC is mediated via binding to UvrB. A mutation in the N-terminal zinc finger domain of UvrA does not affect NER in vivo or in vitro. The same mutation, however, does give inviability in combination with the delta polA mutation. Apparently the N-terminal zinc-binding domain of UvrA has specifically evolved for a function outside DNA repair. A model for the function of the UvrA, UvrB, and UvrD proteins in the alternative replication pathway is discussed.  相似文献   

20.
Summary Previous studies have shown that transformation of Escherichia coli by plasmid DNA modified in vitro by carcinogens leads to RecA-dependant recombination between homologous plasmid and chromosomal DNA sequences. The mechanism of this recombination has now been studied using recombination-deficient mutants, and the influence of induction of the SOS response on the level of recombination investigated. Plasmid pNO1523, containing the str + operon (Sms), has been modified in vitro by either irradiation with UV light, or by reaction with (±) trans-benzo(a)pyrene-7,8-dihydrodiol-9,10-epoxide (BPDE) and used to transform streptomycin-resistant hosts. The formation of Ampr transformants which also carry streptomycin resistance was used as a measure of the level of recombination between plasmid and chromosomal DNA. Transformation of recB and recC mutants produced no change in the level of recombination while in the recF mutant a significant decrease was observed compared to the wild type host. Thermal induction of the SOS response in tif-1 and tif-1 umuC mutants followed by transformation led to a four-fold increase in recombination in both cases. The results suggest that the streptomycin-resistant transformants arise exclusively via a recombinational pathway which is largely dependant on the recF gene product, and that this pathway is influenced by induction of the SOS response. These results are discussed in terms of the mechanism of this recombination.  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