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.
An expression vector, pKGR, for the gor gene from Escherichia coli encoding glutathione reductase was constructed by subcloning of an AvaII fragment of the Clarke & Carbon bank plasmid pGR [Greer & Perham (1986) Biochemistry 25, 2736-2742] into the plasmid pKK223-3. The expression of glutathione reductase from the plasmid pKGR was found to have been successfully placed under the control of the tac promoter. Transformation of E. coli cells with this plasmid resulted in 100-200-fold increase in glutathione reductase activity in cell-free extracts. A rapid purification procedure for the enzyme, based on affinity chromatography on Procion Red HE-7B-CL-Sepharose 4B, was developed. The purified enzyme was homogeneous as judged by SDS/polyacrylamide-gel electrophoresis, and all its properties were consistent with the DNA sequence of the gene [Greer & Perham (1986) Biochemistry 25, 2736-2742] and with those previously reported for E. coli glutathione reductase [Mata, Pinto & Lopez-Barea (1984) Z. Naturforsch. C. Biosci. 39, 908-915]. These experiments have enabled an investigation of the protein chemical and mechanistic properties of the enzyme by site-directed mutagenesis.  相似文献   

2.
In order to investigate the basis of functional diversity among the pyridine nucleotide-oxidoreductases the gor gene from Pseudomonas aeruginosa PAO, which encodes glutathione reductase, was analysed. The P. aeruginosa gor gene was identified by hybridization with a short DNA sequence from the gene encoding mercuric reductase in transposon Tn501. The gene was cloned, sequenced and overexpressed in Escherichia coli. Expression of the gene enabled rescue of an E. coli gor- mutant, confirming the identity of the cloned gene. The predicted sequence of the gene product showed homology with other members of the pyridine nucleotide-disulphide oxidoreductase family, and allowed determination of positions that may be involved in substrate specificity. These predictions provided information on the relationship of sequence to function, independently of structural data used in previous studies.  相似文献   

3.
Almost all bacteria possess glutamate racemase to synthesize d-glutamate as an essential component of peptidoglycans in the cell walls. The enforced production of glutamate racemase, however, resulted in suppression of cell proliferation. In the Escherichia coli JM109/pGR3 clone, the overproducer of glutamate racemase, the copy number (i.e. replication efficiency) of plasmid DNA declined dramatically, whereas the E. coli WM335 mutant that is defective in the gene of glutamate racemase showed little genetic competency. The comparatively low and high activities for DNA supercoiling were contained in the E. coli JM109/pGR3 and WM335 cells, respectively. Furthermore, we found that the DNA gyrase of E. coli was modulated by the glutamate racemase of E. coli in the presence of UDP-N-acetylmuramyl-l-alanine, which is a peptidoglycan precursor and functions as an absolute activator for the racemase. This is the first finding of the enzyme protein participating in both d-amino acid metabolism and DNA processing.  相似文献   

4.
The cloned Escherichia coli gor gene encoding the flavoprotein glutathione reductase was placed under the control of the tac promoter in the plasmid pKK223-3, allowing expression of glutathione reductase at levels approximately 40,000 times those of untransformed cells. This greatly facilitated purification of the enzyme. By directed mutagenesis of the gor gene, His-439 was changed to glutamine (H439Q) and alanine (H439A). The tyrosine residue at position 99 was changed to phenylalanine (Y99F), and in another experiment, the H439Q and Y99F mutations were united to form the double mutant Y99FH439Q. His-439 is thought to act in the catalytic mechanism as a proton donor/acceptor in the glutathione-binding pocket. The H439Q and H439A mutants retain approximately 1% and approximately 0.3%, respectively, of the catalytic activity of the wild-type enzyme. This reinforces our previous finding [Berry et al. (1989) Biochemistry 28, 1264-1269] that direct protonation and deprotonation of the histidine residue are not essential for the reaction to occur. The retention of catalytic activity by the H439A mutant demonstrates further that a side chain capable of hydrogen bonding to a water molecule, which might then act as proton donor, also is not essential at this position. Tyr-99 is a further possible proton donor in the glutathione-binding pocket, but the Y99F mutant was essentially fully active, and the Y99FH439Q double mutant also retained approximately 1% of the wild-type specific activity.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
Directed mutagenesis of the gor gene from Escherichia coli encoding the flavoprotein glutathione reductase was used to convert the two cysteine residues that comprise its redox-active disulphide bridge to alanine (C42A) and serine (C47S) residues. A double mutant (C42AH439A) was also created in which His-439, the proton donor/acceptor in the glutathione-binding site, was additionally converted into an alanine residue. The C42A and C47S mutants were both unable to catalyse the reduction of glutathione by NADPH. The C42A mutant retained the transhydrogenase activity of the wild-type enzyme, whereas the C47S mutant was also inhibited in this reaction. These results support the view that in the catalytic mechanism of E. coli glutathione reductase, the thiolate form of Cys-42 acts as a nucleophile to initiate disulphide exchange with enzyme-bound glutathione and that the thiolate form of Cys-47 generates an essential charge-transfer complex with enzyme-bound FAD. Titration of the C42A and C42AH439A mutants indicated that the imidazole side-chain of His-439 lowered the pKa of the charge-transfer thiol (Cys-47) from 7.7 to 5.7, enhancing its ability to act as an anion at neutral pH. Several important differences between these mutants of E. coli glutathione reductase and similar mutants (or chemically modified forms) of other members of the flavoprotein disulphide oxidoreductase family were noted, but these could be explained in terms of the different redox chemistries of the enzymes concerned.  相似文献   

6.
Seven independently isolated glutathione reductase-deficient (gor) Escherichia coli mutants were found to have an in vivo glutathione redox state that did not significantly differ from that of the parental strain, 98 to 99% reduced. Strains containing both a gor mutation and either a trxA mutation (thioredoxin deficient) or a trxB mutation (thioredoxin reductase deficient) were able to maintain a 94 to 96% reduced glutathione pool, suggesting that glutathione can be reduced independently of glutathione reductase and thioredoxin reductase.  相似文献   

7.
The glutathione reductase gene, gor, was cloned from the plant pathogen Xanthomonas campestris pv. phaseoli. Its gene expression and enzyme characteristics were found to be different from those of previously studied homologues. Northern blot hybridization, promoter-lacZ fusion, and enzyme assay experiments revealed that its expression, unlike in Escherichia coli, is OxyR-independent and constitutive upon oxidative stress conditions. The deduced amino acid sequence shows a unique NADPH binding motif where the most highly conserved arginine residue, which is critical for NADPH binding, is replaced by glutamine. Interestingly, a search of the available Gor amino acid sequences from various sources, including other Xanthomonas species, revealed that this replacement is specific to the genus Xanthomonas. Recombinant Gor enzyme was purified and characterized, and was found to have a novel ability to use both, NADPH and NADH, as electron donor. A gor knockout mutant was constructed and shown to have increased expression of the organic peroxide-inducible regulator gene, ohrR.  相似文献   

8.
We cloned and sequenced the glutathione reductase gene (gor) of an oxygen-tolerant Streptococcus mutans, and constructed a gor-disruption mutant by homologous recombination. The gor gene consisted of 1,350 bp, coding for a protein of 450 amino acid residues. The deduced amino acid sequence of the S. mutans gor gene product showed extensive similarity with those of glutathione reductases from prokaryotes and eukaryotes. Although the mutant could grow aerobically, it showed no growth in the presence of 2 mM diamide, a thiol-specific oxidant. In contrast, growth of the wild-type strain was not significantly inhibited by 2 mM diamide, and glutathione reductase activity was increased 2.2-fold under these conditions. In addition, the level of glutathione reductase activity in the wild-type strain was increased 3.6-fold upon exposure to air, and the elevated level of the enzyme was retained throughout the aerobic growth. Thus, glutathione reductase may be important in protection of S. mutans against oxidative stress.  相似文献   

9.
Using gene replacement and transposon Tn5 mutagenesis, an Escherichia coli ilvC panE double mutant completely lacking ketopantoate reductase activity was isolated. This E. coli double mutant was employed to isolate the E. coli panE gene by genetic complementation. The E. coli panE gene is characterized by a 912 bp coding region, which specifies a protein of 303 amino acids with a deduced molecular mass of 33.8 kD. A panE expression plasmid carrying the panE gene under the control of the tac promotor was constructed. Introduction of the panE expression plasmid into E. coli resulted in a threefold increase in ketopantoate reductase activity. It was also shown that the enhanced panE expression in E. coli K12 led to 3.5-fold increase in pantothenate excretion. Pantothenate excretion could even be more enhanced when the growth medium was supplemented with ketopantoate.  相似文献   

10.
Two independent mutants defective in glutathione reductase (EC 1.6.4.2) were isolated in an Escherichia coli K12 strain lysogenized with bacteriophage Mu. The prophage was lost (and the ability to reduce glutathione regained) by 32% of the xylose-positive transductants when T4GT7 was used as the vector, but the markers were not cotransduced by P1. Similarly, the prophage site and malA were cotransduced by T4GT7 but not by P1. The gor gene maps between min 77 and 78 on the E. coli genome, and the mutation causes no growth defect.  相似文献   

11.
A L-delta 1-pyrroline-5-carboxylate reductase activity has been detected in crude extracts of Desulfovibrio desulfuricans Norway. This P5C reductase activity is also found when a 2.5 kb D. desulfuricans DNA fragment is introduced into an Escherichia coli proC mutant. Although it restores growth of the proC mutant, the ProDd enzyme might be detrimental to the E. coli host since the plasmid carrying the cognate proDd gene is segregated at high rate by the cells but is stabilized by small deletions which lead to a loss of the P5C reductase activity.  相似文献   

12.
A negative regulator gene for synthesis of arylsulfatase in Klebsiella aerogenes was cloned. Deletion analysis showed that the regulator gene was located within a 1.6-kb cloned segment. Transfer of the plasmid, which contains the cloned fragment, into constitutive atsR mutant strains of K. aerogenes resulted in complementation of atsR; the synthesis of arylsulfatase was repressed in the presence of inorganic sulfate or cysteine, and this repression was relieved, in each case, by the addition of tyramine. The nucleotide sequence of the 1.6-kb fragment was determined. From the amino acid sequence deduced from the DNA sequence, we found two open reading frames. One of them lacked the N-terminal region but was highly homologous to the gene which codes for diadenosine tetraphosphatase (apaH) in Escherichia coli. The other open reading frame was located counterclockwise to the apaH-like gene. This gene was highly homologous to the gene which codes for dihydrofolate reductase (folA) in E. coli. We detected 30 times more activity of dihydrofolate reductase in the K. aerogenes strains carrying the plasmid, which contains the arylsulfatase regulator gene, than in the strains without plasmid. Further deletion analysis showed that the K. aerogenes folA gene is consistent with the essential region required for the repression of arylsulfatase synthesis. Transfer of a plasmid containing the E. coli folA gene into atsR mutant cells of K. aerogenes resulted in repression of the arylsulfatase synthesis. Thus, we conclude that the folA gene codes a negative regulator for the ats operon.  相似文献   

13.
Shifting the temperature from 30 to 45 degrees C in an aerobic Escherichia coli culture inhibited the expression of the antioxidant genes katG, katE, sodA, and gor. The expression was evaluated by measuring beta-galactosidase activity in E. coli strains that contained fusions of the antioxidant gene promoters with the lacZ operon. Heat shock inhibited catalase and glutathione reductase, lowered the intracellular level of glutathione, and increased its extracellular level. It also suppressed the growth of mutants deficient in the katG-encoded catalase HPI, whereas the sensitivity of the wild-type and sodA sodB mutant cells to heat shock was almost the same. In the E. coli culture adapted to growth at 42 degrees C, the content of both intracellular and extracellular glutathione was two times higher than in the culture grown at 30 degrees C. The temperature-adapted cells grown aerobically at 42 degrees C showed an increased ability to express the fused katG-lacZ genes.  相似文献   

14.
15.
Previously, we isolated a selenate- and arsenate-reducing bacterium, designated strain SF-1, from selenium-contaminated sediment and identified it as a novel species, Bacillus selenatarsenatis. B. selenatarsenatis strain SF-1 independently reduces selenate to selenite, arsenate to arsenite, and nitrate to nitrite by anaerobic respiration. To identify the genes involved in selenate reduction, 17 selenate reduction-defective mutant strains were isolated from a mutant library generated by random insertion of transposon Tn916. Tn916 was inserted into the same genome position in eight mutants, and the representative strain SF-1AM4 did not reduce selenate but did reduce nitrate and arsenate to the same extent as the wild-type strain. The disrupted gene was located in an operon composed of three genes designated srdBCA, which were predicted to encode a putative oxidoreductase complex by the BLASTX program. The plasmid vector pGEMsrdBCA, containing the srdBCA operon with its own promoter, conferred the phenotype of selenate reduction in Escherichia coli DH5α, although E. coli strains containing plasmids lacking any one or two of the open reading frames from srdBCA did not exhibit the selenate-reducing phenotype. Domain structure analysis of the deduced amino acid sequence revealed that SrdBCA had typical features of membrane-bound and molybdopterin-containing oxidoreductases. It was therefore proposed that the srdBCA operon encoded a respiratory selenate reductase complex. This is the first report of genes encoding selenate reductase in gram-positive bacteria.  相似文献   

16.
A recA-like gene was isolated from a gene library of Lactococcus lactis subsp. lactis by intergeneric complementation of an E. coli recA mutant. A plasmid was obtained which fully complemented the RecA response to DNA damaging agents and UV inducibility of prophage, but not P1 plating efficiency in an E. coli recA mutant. The cloned DNA fragment also partially complemented the rec mutation in Lc. lactis MMS36. Hybridization studies showed that there was no detectable sequence homology between the recA gene of E. coli and Lc. lactis subsp. lactis chromosomal DNA.  相似文献   

17.
A mutant of Escherichia coli that contains essentially no detectable glutathione has been isolated. The mutant contains a very low level of the enzyme glutathione synthetase and accumulates lambda-glutamyl cysteine at a concentration approximately equal to the level of glutathione found in its parent. No significant differences in growth were observed between the mutant and its parent. However, the activity of at least one enzyme was found to be affected by the absence of glutathione; the specific activity of the B1 subunit of ribonucleoside diphosphate reductase was greatly reduced. The possibility that the decreased B1 activity is due to a mutation in the structural gene coding for B1 or its regulatory gene could be eliminated. This suggests that one role of glutathione in the cell is to maintain at least this one protein in an active state. We propose the designation gshB for the gene coding for glutathione synthetase.  相似文献   

18.
A new strategy for molecular cloning in the cyanobacterium Anacystis nidulans R-2 is described. This strategy involved the use of a transposon and was developed for the cloning of a gene encoding methionine biosynthesis. A met::Tn901 mutant was isolated. Chromosomal DNA fragments were cloned in the Escherichia coli plasmid vector pACYC184. A recombinant plasmid carrying the inactivated met::Tn901 gene was selected after transformation to E. coli. The cloned met::Tn901 DNA fragment was used as a probe to select the corresponding A. nidulans R-2 wild-type met gene from a gene library prepared in E. coli, using the newly constructed shuttle cosmid vector pPUC29. When transformed into A. nidulans Met- mutants, this cloned gene allowed the mutants to grow prototrophically.  相似文献   

19.
Acetohydroxyacid isomeroreductase (AHAIR) is the shared second enzyme in the biosynthetic pathways leading to isoleucine and valine. AHAIR is encoded by the ilvC gene in bacteria. A 1,544-bp fragment of genomic DNA containing the ilvC gene was cloned from the cyanobacterium Synechocystis sp. strain PCC 6803, and the complete nucleotide sequence was determined. The identity of the gene was established by comparison of the nucleotide and derived peptide sequences with those of other ilvC genes. The highest degree of sequence similarity was found with the ilvC gene from Rhizobium meliloti. The isolated Synechocystis ilvC gene complemented an Escherichia coli ilvC mutant lacking AHAIR activity. The expressed Synechocystis gene encodes a protein that has a molecular mass of 35.7 kDa and that has AHAIR activity in an in vitro assay. Polyclonal antibodies raised against purified Synechocystis AHAIR produced a single band on a Western blot (immunoblot) of a Synechocystis cell extract and detected the protein in an extract of an E. coli ilvC mutant strain that was transformed with a plasmid containing the Synechocystis ilvC gene. The antibody did not react with an extract of an E. coli ilvC mutant strain that was transformed with a control plasmid lacking the Synechocystis ilvC gene or with an extract of an E. coli IlvC+ control strain.  相似文献   

20.
Bacillus subtilis plasmid pGR71 is a promoter-probe shuttle vector derived from pUB110. The expression of the cat gene on pGR71 in B. subtilis requires the insertion of a Bacillus promoter and a ribosomal binding site (RBS) into the HindIII cloning site immediately upstream from the cat gene. A recombinant plasmid of pGR71, named pGR71-369, was obtained by a spontaneous deletion of a fragment containing most of the inserted HindIII fragment and the replication origin necessary for multiplication in Escherichia coli. The expression of the cat gene in B. subtilis cells carrying this plasmid was inducible by heat. Nucleotide sequence analysis of the upstream region of the cat gene, deletion analysis, and dot blot hybridization analysis of mRNA in various conditions revealed that the cat gene was expressed by heat-inducible translational coupling and that the regulatory region of heat inducibility was present in the upstream region of the cat gene.  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