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 46 毫秒
1.
Misonidazole, and RSU 1069 and 6 of its analogues are all reported to show increased cytotoxicity towards hypoxic cells compared to oxic cells. DNA is considered to be the target through which these drugs exert their cytotoxic activity. Therefore we monitored induction of the SOS response in uvrABC excinuclease proficient and deficient strains of E. coli, under oxic and hypoxic conditions, as an indirect method of assessing the activity of these drugs towards DNA in a biological system. This was done using the SOS chromotest which utilizes E. coli strains which possess a sfiA::lacZ fusion allowing induction of the SOS response to be monitored by assaying beta-galactosidase activity. All of the drugs tested here show some induction of the SOS response in both uvrABC excinuclease proficient and deficient strains. Data shown here suggests that the uvrABC excinuclease is important in the production of a SOS induction signal from RSU 1069-induced DNA lesions and that RSU 1069 may act as a crosslinking agent. The data also shows that SOS induction activity and toxicity do not necessarily correlate and that production of a SOS induction signal may occur via a different pathway for RSU 1069 than for its analogues.  相似文献   

2.
The kinetics of SOS system induction in Escherichia coli PQ37 cells by gamma-irradiation has been studied by the SOS chromotest technique. It was shown that the synthesis of constitutive alkaline phosphatase is not immediately stopped in cells that suffered lethal damages from gamma-irradiation. The production of DNA damages inducing the SOS system was 0.021/Gy per genome. The SOS system was switched off approximately 200 min after gamma-irradiation. A correction is proposed to the calculation of the SOS system induction factor.  相似文献   

3.
The SOS response in Escherichia coli results in the coordinately induced expression of more than 40 genes which occurs when cells are treated with DNA-damaging agents. This response is dependent on RecA (coprotease), LexA (repressor), and the presence of single-stranded DNA (ssDNA). A prerequisite for SOS induction is the formation of a RecA-ssDNA filament. Depending on the DNA substrate, the RecA-ssDNA filament is produced by either RecBCD, RecFOR, or a hybrid recombination mechanism with specific enzyme activities, including helicase, exonuclease, and RecA loading. In this study we examined the role of RecA loading activity in SOS induction after UV irradiation. We performed a genetic analysis of SOS induction in strains with a mutation which eliminates RecA loading activity in the RecBCD enzyme (recB1080 allele). We found that RecA loading activity is essential for SOS induction. In the recB1080 mutant RecQ helicase is not important, whereas RecJ nuclease slightly decreases SOS induction after UV irradiation. In addition, we found that the recB1080 mutant exhibited constitutive expression of the SOS regulon. Surprisingly, this constitutive SOS expression was dependent on the RecJ protein but not on RecFOR, implying that there is a different mechanism of RecA loading for constitutive SOS expression.  相似文献   

4.
The kinetics of the SOS induction in E. coli cells of wild type and deficient in umuC gene exposed to UV and gamma-rays were analysed. In the presence of UmuC protein SOS induction was 3-5.5 times lower and delayed for about 30 minutes after both UV and gamma rays. It was shown that decrease of the SOS induction in wild type cells irradiated by UV was due to more effective elimination of the photolesions from DNA by excision repair system. UmuCD-dependent inhibition of DNA replication was discussed as a possible mechanism allowing additional time for error-free repair.  相似文献   

5.
The Escherichia coli UVM response is a recently described phenomenon in which pretreatment of cells with DNA-damaging agents such as UV or alkylating agents significantly enhances mutation fixation at a model mutagenic lesion (3,N4-ethenocytosine; epsilon C) borne on a transfected M13 single-stranded DNA genome. Since UVM is observed in delta recA cells in which SOS induction should not occur, UVM may represent a novel, SOS-independent, inducible response. Here, we have addressed two specific hypothetical mechanisms for UVM: (i) UVM results from a recA-independent pathway for the induction of SOS genes thought to play a role in induced mutagenesis, and (ii) UVM results from a polymerase switch in which M13 replication in treated cells is carried out by DNA polymerase I (or DNA polymerase II) instead of DNA polymerase III. To address these hypotheses, E. coli cells with known defects in recA, lexA, umuDC, polA, or polB were treated with UV or 1-methyl-3-nitro-1-nitrosoguanidine before transfection of M13 single-stranded DNA bearing a site-specific ethenocytosine lesion. Survival of the transfected DNA was measured as transfection efficiency, and mutagenesis at the epsilon C residue was analyzed by a quantitative multiplex DNA sequencing technology. Our results show that UVM is observable in delta recA cells, in lexA3 (noninducible SOS repressor) cells, in LexA-overproducing cells, and in delta umuDC cells. Furthermore, our data show that UVM induction occurs in the absence of detectable induction of dinD, an SOS gene. These results make it unlikely that UVM results from a recA-independent alternative induction pathway for SOS gene.  相似文献   

6.
DNA double-strand breaks are particularly deleterious lesions that can lead to genomic instability and cell death. We investigated the SOS response to double-strand breaks in both Escherichia coli and Bacillus subtilis. In E. coli, double-strand breaks induced by ionizing radiation resulted in SOS induction in virtually every cell. E. coli strains incapable of SOS induction were sensitive to ionizing radiation. In striking contrast, we found that in B. subtilis both ionizing radiation and a site-specific double-strand break causes induction of prophage PBSX and SOS gene expression in only a small subpopulation of cells. These results show that double-strand breaks provoke global SOS induction in E. coli but not in B. subtilis. Remarkably, RecA-GFP focus formation was nearly identical following ionizing radiation challenge in both E. coli and B. subtilis, demonstrating that formation of RecA-GFP foci occurs in response to double-strand breaks but does not require or result in SOS induction in B. subtilis. Furthermore, we found that B. subtilis cells incapable of inducing SOS had near wild-type levels of survival in response to ionizing radiation. Moreover, B. subtilis RecN contributes to maintaining low levels of SOS induction during double-strand break repair. Thus, we found that the contribution of SOS induction to double-strand break repair differs substantially between E. coli and B. subtilis.  相似文献   

7.
The SOS response is an important mechanism which allows Escherichia coli cells to maintain genome integrity. Two key proteins in SOS regulation are LexA (repressor) and RecA (coprotease). The signal for SOS induction is generated at the level of a RecA filament. Depending on the type of DNA damage, a RecA filament is produced by specific activities (helicase, nuclease and RecA loading) of either RecBCD, RecF or a hybrid recombination pathway. It was recently demonstrated that RecA loading activity is essential for the induction of the SOS response after UV-irradiation. In this paper we studied the genetic requirements for SOS induction after introduction of a double-strand break (DSB) by the I-SceI endonuclease in a RecA loading deficient recB mutant (recB1080). We monitored SOS induction by assaying beta-galactosidase activity and compared induction of the response between strains having one or more inactivated mechanisms of RecA loading and their derivatives. We found that simultaneous inactivation of both RecA loading functions (in recB1080 recO double mutant) partially impairs SOS induction after introduction of a DSB. However, we found that the RecJ nuclease is essential for SOS induction after the introduction of a DSB in the recB1080 mutant. This result indicates that RecJ is needed to prepare ssDNA for subsequent loading of RecA protein. It implies that an additional type of RecA loading could exist in the cell.  相似文献   

8.
A study was made of the SOS induction of the gene sulA of Escherichia coli K12 in relation to the gene dosage of the gene recA. In experiments the sulA::lacZ fusion strain PQ37 and derivatives of PQ37 with the multi-copy plasmids pDR1453 or pBR322 were used. The SOS response was induced with nitrofurantoin, SOS induction of the gene sulA was determined on the basis of the amount of beta-galactosidase synthesized, i.e. by the SOS chromotest (Quillardet et al., 1982a). It was found in this work that cells with the plasmid pDR1453, which contain the gene recA of E. coli K12 (Sancar and Rupp, 1979), have a decreased SOS induction of the gene sulA. Cells with the plasmid pBR322 do not exhibit this decrease. Inactivation of the gene recA in the plasmid pDR1453 with preservation of the functional gene recA in the chromosome leads to a restoration of 'standard' SOS induction of the gene sulA. The results show that the amount of the gene product of the gene recA affects the SOS induction of the gene sulA.  相似文献   

9.
10.
Expression of several SOS functions such as induction of lambda prophage, inhibition of cell division and induction of both umuC and recA genes after UV-irradiation, nalidixic acid or mitomycin C addition was studied in an RecBC- mutant. UV-irradiation and mitomycin C induced all SOS functions studied in the RecBC- cells but at a lower level and delayed with respect to the wild-type strain. On the contrary, nalidixic acid was unable to trigger any of these SOS functions. In the RecBC- mutant, adenine only had a stimulating effect on the amplification of RecA protein synthesis following UV-irradiation. Nevertheless, in the wild-type strain the stimulating effect occurred in all SOS functions studied following UV-irradiation as well as in the amplification of RecA protein synthesis by nalidixic acid but not in the other SOS functions triggered by this compound. Furthermore, adenine produced a decrease in the mitomycin C-mediated induction of all SOS functions studied in both RecBC- and wild-type strains.  相似文献   

11.
Resistance of tumors to drugs such as cisplatin and mitomycin C (MMC) is an important factor limiting their usefulness in cancer chemotherapy. The antitumor effects of these drugs are due to the formation of bifunctional adducts in DNA, with cisplatin causing predominantly intrastrand-crosslinks and MMC causing interstrand-crosslinks. The SOS chromotest was used to study the cellular mechanisms that process DNA damage in Escherichia coli exposed to cisplatin, ultraviolet irradiation (UV) and MMC and subsequently facilitate the production of a molecular signal for induction of the SOS response. Strains used in the SOS chromotest have a fusion of lacZ with the sfiA (sulA) gene so that the amount of SOS inducing signal, which is modulated by the ability of the cell to repair DNA, is measured by assaying beta-galactosidase activity. SOS induction in a strain proficient in homologous recombination (HR) was compared with that in isogenic strains deficient in HR due to a blocked RecBC pathway caused by a recB mutation or a blocked RecFOR pathway caused by a recO mutation. The effect of cisplatin treatment in a uvrA mutant strain blocked at the first step of NER was compared with that in an isogenic strain proficient in NER. Cellular resistance was measured as percent colony forming units (cfu) for cells treated with increasing doses of cisplatin, MMC and UV relative to that in untreated control cultures. The importance of both HR pathways for resistance to these treatments was demonstrated by decreased survival in mutants with the recB mutant being more sensitive than the recO mutant. SOS induction levels were elevated in the sensitive recB strain relative to the HR proficient strain possibly due to stalled and/or distorted replication forks at crosslinks in DNA. In contrast, induction of SOS was dependent on RecFOR activity that is thought to act at daughter strand gaps in newly synthesized DNA to mediate production of the signal for SOS induction. Proficiency in NER was necessary for both survival and high levels of SOS induction in cisplatin treated cells.  相似文献   

12.
Escherichia coli responds to impairment of DNA synthesis by inducing a system of DNA repair known as the SOS response. Specific genes are derepressed through proteolytic cleavage of their repressor, the lexA gene product. Cleavage in vivo requires functional RecA protein in a role not yet understood. We used mRNA hybridization techniques to follow the rapid changes that occur with induction in cells with mutations in the recA operator or in the repressor cleavage site. These mutations allowed us to uncouple the induction of RecA protein synthesis from its role in inducing the other SOS functions. Following induction with ultraviolet light, we observed increased rates of mRNA synthesis from five SOS genes within five minutes, maximum expression ten to 20 minutes later and then a later decline to near the initial rates. The presence of a recA operator mutation did not significantly influence these kinetics, whereas induction was fully blocked by an additional mutation in the repressor cleavage site. These experiments are consistent with activation of RecA protein preceding repressor cleavage and derepression of SOS genes. The results also suggest that the timing and extent of induction of individual SOS genes may be different.  相似文献   

13.
The structural maintenance of chromosome (Smc) protein is highly conserved and involved in chromosome compaction, cohesion, and other DNA-related processes. In Bacillus subtilis, smc null mutations cause defects in DNA supercoiling, chromosome compaction, and chromosome partitioning. We investigated the effects of smc mutations on global gene expression in B. subtilis using DNA microarrays. We found that an smc null mutation caused partial induction of the SOS response, including induction of the defective prophage PBSX. Analysis of SOS and phage gene expression in single cells indicated that approximately 1% of smc mutants have fully induced SOS and PBSX gene expression while the other 99% of cells appear to have little or no expression. We found that induction of PBSX was not responsible for the chromosome partitioning or compaction defects of smc mutants. Similar inductions of the SOS response and PBSX were observed in cells depleted of topoisomerase I, an enzyme that relaxes negatively supercoiled DNA.  相似文献   

14.
We have transfected SOS-induced and uninduced cells of a uvrA6 strain of Escherichia coli with single-stranded M13mp7-based vectors that carried a single trans-syn T-T cyclobutane dimer at a unique site. Unlike constructs carrying the cis-syn isomer of this lesion, these vectors could be replicated with modest efficiency (14%) in the absence of SOS induction and therefore provided an opportunity to measure directly the influence of such induction on error rate and mutation spectrum. We found that translesion synthesis in the absence of SOS induction was remarkably accurate; only 4% of the replicated bacteriophage contained mutations, which were exclusively targeted single T deletions. In SOS-induced cells, error frequency increased to 11% and the resulting mutations included targeted substitutions and near-targeted single base additions, as well as the T deletions. Replication efficiency was 29% in these conditions. SOS induction therefore leads not only to an enhanced capacity to replicate damaged DNA but also to a marked change in mutation frequency and spectrum.  相似文献   

15.
We investigated the dynamics of the SOS response induction and the frequency of reversions induced by the monofunctional alkylating compound N-methyl-N'-nitro-N-nitrosoguanidine in Escherichia coli cells exposed to osmotic stress for 1 h. During the stress treatment of the wild-type cultures adapted and not adapted to the alkylating agent, the maximum SOS response values and induced reversion frequencies were recorded twice. The SOS response values and induced reversion frequencies remained unchanged during the whole period after attaining the maximum values in adapted and nonadapted cells carrying a mutation in the excision repair gene. Presumably, the SOS mutagenesis mechanisms are turned on in the cells with an inactivated excision repair system earlier than in wild-type cells.  相似文献   

16.
In natural environments, bacteria are frequently exposed to sub‐lethal levels of DNA damage, which leads to the induction of a stress response (the SOS response in Escherichia coli). Natural environments also vary in nutrient availability, resulting in distinct physiological changes in bacteria, which may have direct implications on their capacity to repair their chromosomes. Here, we evaluated the impact of varying the nutrient availability on the expression of the SOS response induced by chronic sub‐lethal DNA damage in E. coli. We found heterogeneous expression of the SOS regulon at the single‐cell level in all growth conditions. Surprisingly, we observed a larger fraction of high SOS‐induced cells in slow growth as compared with fast growth, despite a higher rate of SOS induction in fast growth. The result can be explained by the dynamic balance between the rate of SOS induction and the division rates of cells exposed to DNA damage. Taken together, our data illustrate how cell division and physiology come together to produce growth‐dependent heterogeneity in the DNA damage response.  相似文献   

17.
We investigated the dynamics of the SOS response induction and the frequency of reversions induced by the monofunctional alkylating compound N-methyl-N′-nitro-N-nitrosoguanidine in Escherichia coli cells exposed to osmotic stress for 1 h. During the stress treatment of the wild-type cultures adapted and not adapted to the alkylating agent, the maximum SOS response values and induced reversion frequencies were recorded twice. The SOS response values and induced reversion frequencies remained unchanged during the whole period after attaining the maximum values in adapted and nonadapted cells carrying a mutation in the excision repair gene. Presumably, the SOS mutagenesis mechanisms are turned on in the cells with an inactivated excision repair system earlier than in wild-type cells.  相似文献   

18.
A knock-out mutant of the dinR gene that encodes the SOS regulon repressor in Bacillus subtilis was constructed. The yneA, yneB and ynzC genes transcribed divergently from the dinR gene were strongly induced in mutant cells. Northern hybridization analyses revealed that these genes collectively form an operon and belong to the SOS regulon. The simultaneous deletion of dinR and yneA suppressed the filamentous phenotype of the dinR mutant. Furthermore, although yneA is suppressed in the wild-type cell in the absence of SOS induction, artificial expression of the YneA protein using an IPTG-inducible promoter resulted in cell elongation. Disruption of yneA significantly reduced cell elongation after the induction of the SOS response by mitomycin C in dinR+ cells. These results indicate that the YneA protein is responsible for cell division suppression during the SOS response in B. subtilis. Localization of the FtsZ protein to the cell division site was reduced in dinR-disrupted or yneA-expressing cells, further suggesting that the YneA protein suppresses cell division through the suppression of FtsZ ring formation. Interestingly, the B. subtilis YneA protein is structurally and phylogenetically unrelated to its functional counterpart in Escherichia coli, SulA.  相似文献   

19.
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   

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