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 312 毫秒
1.
Glutaredoxin is essential for the glutathione (GSH)-dependent synthesis of deoxyribonucleotides by ribonucleotide reductase, and in addition, it displays a general GSH disulfide oxidoreductase activity. In Escherichia coli glutaredoxin, the active site contains a redox-active disulfide/dithiol of the sequence Cys11-Pro12-Tyr13-Cys14. In this paper, we have prepared and characterized the Cys14----Ser mutant of E. coli glutaredoxin and its mixed disulfide with glutathione. The Cys14----Ser mutant of glutaredoxin is shown to retain 38% of the GSH disulfide oxidoreductase activity of the wild-type protein with hydroxyethyl disulfide as substrate but to be completely inactive with ribonucleotide reductase, demonstrating that dithiol glutaredoxin is the hydrogen donor for ribonucleotide reductase. The covalent structure of the mixed disulfide of glutaredoxin(C14S) with GSH prepared with 15N-labeling of the protein was confirmed with nuclear magnetic resonance (NMR) spectroscopy, establishing a basis for NMR structural studies of the glutathione binding site on glutaredoxin.  相似文献   

2.
The reducing proteins glutaredoxin 3 (Grx3) and glutaredoxin 1 (Grx1) are structurally similar but exhibit different specificities toward substrates. Grx1 efficiently reduces ribonucleotide reductase and PAPS reductase, while Grx3 reduces these enzymes inefficiently or not at all. We previously described a selection for Grx3 mutants with increased activity toward substrates of Grx1 in vivo. Remarkably, we repeatedly isolated mutants with changes in only one of the amino acids of Grx3, methionine 43, converting it to either valine, leucine, or isoleucine. In this paper we present additional genetic studies and a biochemical characterization of Grx3-Met43Val, the most efficient mutant. We show that Grx3-Met43Val is able to reduce ribonucleotide reductae and PAPS reductase much more efficiently than the wild-type protein in vitro. The altered protein has an increased Vmax over that of Grx3, nearly the same Vmax as Grx1 while the Km remains high. Molecular dynamics simulations suggest that the Met43Val substitution results in changes in properties of the N-terminal cysteine of the active site leading to a considerably lower pKa. Furthermore, Grx3-Met43Val shows an 11 mV lower redox potential than the wild-type Grx3. These findings provide biochemical and structural explanations for the increased reductive efficiency of the mutant Grx3.  相似文献   

3.
Three Escherichia coli glutaredoxins catalyze GSH-disulfide oxidoreductions, but the atypical 24-kDa glutaredoxin 2 (Grx2, grxB gene), in contrast to the 9-kDa glutaredoxin 1 (Grx1, grxA gene) and glutaredoxin 3 (Grx3, grxC gene), is not a hydrogen donor for ribonucleotide reductase. To improve the understanding of glutaredoxin function, a null mutant for grxB (grxB(-)) was constructed and combined with other mutations. Null mutants for grxB or all three glutaredoxin genes were viable in rich and minimal media with little changes in their growth properties. Expression of leaderless alkaline phosphatase showed that Grx1 and Grx2 (but not Grx3) contributed in the reduction of cytosolic protein disulfides. Moreover, Grx1 could catalyze disulfide formation in the oxidizing cytosol of combined null mutants for glutathione reductase and thioredoxin 1. grxB(-) cells were more sensitive to hydrogen peroxide and other oxidants and showed increased carbonylation of intracellular proteins, particularly in the stationary phase. Significant up-regulation of catalase activity was observed in null mutants for thioredoxin 1 and the three glutaredoxins, whereas up-regulation of glutaredoxin activity was observed in catalase-deficient strains with additional defects in the thioredoxin pathway. The expression of catalases is thus interconnected with the thioredoxin/glutaredoxin pathways in the antioxidant response.  相似文献   

4.
The bacterial peroxiredoxin AhpC, a cysteine-dependent peroxidase, can be converted through a single amino acid insertion to a disulfide reductase, AhpC*, active in the glutathione and glutaredoxin pathway. Here we show that, whereas AhpC* is inactive as a peroxidase, other point mutants in AhpC can confer the in vivo disulfide reductase activity without abrogating peroxidase activity. Moreover, AhpC* and several point mutants tested in vitro exhibit an enhanced reductase activity toward mixed disulfides between glutathione and glutaredoxin (Grx-S-SG), consistent with the in vivo requirements for these components. Remarkably, this Grx-S-SG reductase activity relies not on the peroxidatic cysteine but rather on the resolving cysteine that plays only a secondary role in the peroxidase mechanism. Furthermore, putative conformational changes, which impart this unusual Grx-S-SG reductase activity, are transmissible across subunits. Thus, AhpC and potentially other peroxiredoxins in this widespread family can elaborate a new reductase function that alleviates disulfide stress.  相似文献   

5.
Thioredoxin and related proteins in procaryotes   总被引:15,自引:0,他引:15  
Thioredoxin is a small (Mr 12,000) ubiquitous redox protein with the conserved active site structure: -Trp-Cys-Gly-Pro-Cys-. The oxidized form (Trx-S2) contains a disulfide bridge which is reduced by NADPH and thioredoxin reductase; the reduced form [Trx(SH)2] is a powerful protein disulfide oxidoreductase. Thioredoxins have been characterized in a wide variety of prokaryotic cells, and generally show about 50% amino acid homology to Escherichia coli thioredoxin with a known three-dimensional structure. In vitro Trx-(SH)2 serves as a hydrogen donor for ribonucleotide reductase, an essential enzyme in DNA synthesis, and for enzymes reducing sulfate or methionine sulfoxide. E. coli Trx-(SH)2 is essential for phage T7 DNA replication as a subunit of T7 DNA polymerase and also for assembly of the filamentous phages f1 and M13 perhaps through its localization at the cellular plasma membrane. Some photosynthetic organisms reduce Trx-S2 by light and ferredoxin; Trx-(SH)2 is used as a disulfide reductase to regulate the activity of enzymes by thiol redox control. Thioredoxin-negative mutants (trxA) of E. coli are viable making the precise cellular physiological functions of thioredoxin unknown. Another small E. coli protein, glutaredoxin, enables GSH to be hydrogen donor for ribonucleotide reductase or PAPS reductase. Further experiments with molecular genetic techniques are required to define the relative roles of the thioredoxin and glutaredoxin systems in intracellular redox reactions.  相似文献   

6.
Abstract Thioredoxin is a small ( M r 12,000) ubiquitous redox protein with the conserved active site structure: -Trp-Cys-Gly-Pro-Cys-. The oxidized form (Trx-S2) contains a disulfide bridge which is reduced by NADPH and thioredoxin reductase; the reduced form [Trx(SH)2] is a powerful protein disulfide oxidoreductase. Thioredoxins have been characterized in a wide variety of prokaryotic cells, and generally show about 50% amino acid homology to Escherichia coli thioredoxin with a known three-dimensional structure. In vitro Trx-(SH)2 serves as a hydrogen donor for ribonucleotide reductase, an essential enzyme in DNA synthesis, and for enzymes reducing sulfate or methionine sulfoxide. E. coli Trx-(SH)2 is essential for phage T7 DNA replication as a subunit of T7 DNA polymerase and also for assembly of the filamentous phages f1 and M13 perhaps through its localization at the cellular plasma membrane. Some photosynthetic organisms reduce Trx-S2 by light and ferrodoxin; Trx-(SH)2 is used as a disulfide reductase to regulate the activity of enzymes by thiol redox control.
Thioredoxin-negative mutants ( trxA ) of E. coli are viable making the precise cellular physiological functions of thioredoxin unknown. Another small E. coli protein, glutaredoxin, enables GSH to be hydrogen donor for ribonucleotide reductase or PAPS reductase. Further experiments with molecular genetic techniques are required to define the relative roles of the thioredoxin and glutaredoxin systems in intracellular redox reactions.  相似文献   

7.
The alternate cofactor (7004 cofactor) for Escherichia coli adenosine 3'-phosphate 5'-phosphosulfate (PAPS) reductase originally discovered in an E. coli mutant (tsnC 7004) lacking thioredoxin activity has now been purified and characterized. The tryptic peptide map of the 7004 cofactor is totally different from that of thioredoxin, indicating that the two proteins are unrelated in their primary structure. The 7004 cofactor has an amino acid composition different from that of thioredoxin but similar to that of glutaredoxin, a protein required for the glutathione-dependent deoxyribonucleotide formation by ribonucleotide reductase. Thus, the 7004 cofactor could not be a mutated form of thioredoxin, as was suspected earlier. Thioredoxin but not glutaredoxin is a substrate for thioredoxin reductase, but both thioredoxin and glutaredoxin can catalyze the dithiothreitol- or glutathione-dependent reduction of PAPS. On a molar basis, the dithiothreitol-coupled cofactor activity of thioredoxin is three- to fourfold higher that that of glutaredoxin. Comparison of the cofactor activities in the glutathione-coupled and the dithiothreitol-coupled PAPS reductase reaction shows that the cofactor activity of thioredoxin in the glutathione-coupled reaction is only 23% of that observed in the dithiothreitol-coupled reaction. However, in the case of glutaredoxin, cofactor activities are approximately the same in both the dithiothreitol- and glutathione-coupled reactions.  相似文献   

8.
Lactococcus lactis has two essential ribonucleotide reductases for DNA biosynthesis and repair which are affected in the presence or absence of oxygen. Expression of glutaredoxin like protein (NrdH), the hydrogen donor for ribonucleotide reductase, was found to be regulated by the FNR like proteins (FlpA and FlpB). Proteomics study demonstrated that expression level of NrdH significantly decreased in the flpA and flpAB deletion mutants. The nrdH gene is located in an nrdHIEF operon and encoding the NrdEF ribonucleotide reductase, which is active under aerobic and anaerobic conditions. Regulation of expression of the nrdHIEF operons was investigated using beta-galactosidase as a reporter gene. The 588 bp fragment containing the nrdH promoter and gene cloned into the pORI vector immediately upstream of a promoterless lacZ gene. Constructed plasmid was transferred into wild type (MG1363), single mutant (flpA orflpB) and double mutant (flpAB). Aerobically, nrdH promoter activity is 15-fold higher than anaerobic expression.  相似文献   

9.
M Russel  P Model    A Holmgren 《Journal of bacteriology》1990,172(4):1923-1929
We have shown previously that Escherichia coli cells constructed to lack both thioredoxin and glutaredoxin are not viable unless they also acquire an additional mutation, which we called X. Here we show that X is a cysA mutation. Our data suggest that the inviability of a trxA grx double mutant is due to the accumulation of 3'-phosphoadenosine 5'-phosphosulfate (PAPS), an intermediate in the sulfate assimilation pathway. The presence of excess cystine at a concentration sufficient to repress the sulfate assimilation pathway obviates the need for an X mutation and prevents the lethality of a novel cys+ trxA grx double mutant designated strain A522. Mutations in genes required for PAPS synthesis (cysA or cysC) protect cells from the otherwise lethal effect of elimination of both thioredoxin and glutaredoxin even in the absence of excess cystine. Both thioredoxin and glutaredoxin have been shown to be hydrogen donors for PAPS reductase (cysH) in vitro (M. L.-S. Tsang, J. Bacteriol. 146:1059-1066, 1981), and one or the other of these compounds is presumably essential in vivo for growth on minimal medium containing sulfate as the sulfur source. The cells which lack both thioredoxin and glutaredoxin require cystine or glutathione for growth on minimal medium but maintain an active ribonucleotide reduction system. Thus, E. coli must contain a third hydrogen donor active with ribonucleotide reductase.  相似文献   

10.
Rnr4p, a novel ribonucleotide reductase small-subunit protein.   总被引:11,自引:3,他引:8       下载免费PDF全文
Ribonucleotide reductases catalyze the formation of deoxyribonucleotides by the reduction of the corresponding ribonucleotides. Eukaryotic ribonucleotide reductases are alpha2beta2 tetramers; each of the larger, alpha subunits possesses binding sites for substrate and allosteric effectors, and each of the smaller, beta subunits contains a binuclear iron complex. The iron complex interacts with a specific tyrosine residue to form a tyrosyl free radical which is essential for activity. Previous work has identified two genes in the yeast Saccharomyces cerevisiae, RNR1 and RNR3, that encode alpha subunits and one gene, RNR2, that encodes a beta subunit. Here we report the identification of a second gene from this yeast, RNR4, that encodes a protein with significant similarity to the beta-subunit proteins. The phenotype of rnr4 mutants is consistent with that expected for a defect in ribonucleotide reductase; rnr4 mutants are supersensitive to the ribonucleotide reductase inhibitor hydroxyurea and display an S-phase arrest at their restrictive temperature. rnr4 mutant extracts are deficient in ribonucleotide reductase activity, and this deficiency can be remedied by the addition of exogenous Rnr4p. As is the case for the other RNR genes, RNR4 is induced by agents that damage DNA. However, Rnr4p lacks a number of sequence elements thought to be essential for iron binding, and mutation of the critical tyrosine residue does not affect Rnr4p function. These results suggest that Rnr4p is catalytically inactive but, nonetheless, does play a role in the ribonucleotide reductase complex.  相似文献   

11.
We measured both pyridine nucleotide levels and ribonucleotide reductase-specific activity in Yoshida ascites hepatoma cells as a function of growth in vivo and during recruitment from non-cycling to cycling state in vitro. Oxidized nicotinamide adenine dinucleotide (NAD+) and reduced nicotinamide adenine dinucleotide (NADP) levels remained unchanged during tumour growth, while NADP+ and reduced nicotinamide adenine dinucleotide phosphate (NADPH) levels were very high in exponentially growing cells and markedly decreased in the resting phase. Ribonucleotide reductase activity paralleled NADP(H) (NADP+ plus NADPH) intracellular content. The concomitant increase in both NADP(H) levels and ribonucleotide reductase activity was also observed during G1-S transition in vitro. Cells treated with hydroxyurea showed a comparable correlation between the pool size of NADP(H) and ribonucleotide reductase activity. On the basis of these findings, we suggest that fluctuations in NADP(H) levels and ribonucleotide reductase activity might play a critical role in cell cycle regulation.  相似文献   

12.
Cyclic AMP arrests T lymphocytes in the G1 phase of the cell cycle, and prolonged exposure results in cytolysis. Both of these effects require cyclic AMP-dependent protein kinase. We recently observed that some S49 mouse T lymphoma cell lines selected for hydroxyurea resistance were not arrested in G1 by cyclic AMP. Further analysis revealed that these cell lines were cyclic AMP-dependent protein kinase deficient, and conversely, other cyclic AMP-dependent protein kinase deficient cell lines not selected for hydroxyurea resistance were two- to threefold more hydroxyurea resistant. However, hydroxyurea is a specific inhibitor of ribonucleotide reductase and does not inhibit this kinase. We subsequently showed that cyclic AMP-dependent protein kinase will phosphorylate the M2 but not the M1 subunit of ribonucleotide reductase in vitro, and this phosphorylation will diminish CDP reductase activity. In vivo phosphorylation of M2 occurred under conditions similar to those that generate cell cycle arrest. We conclude that the M2 subunit of ribonucleotide reductase can be a target of cyclic AMP-dependent protein kinase. The phosphorylated enzyme has diminished activity, and this may play a role in cyclic AMP-induced lymphocyte cell cycle arrest.  相似文献   

13.
The large subunit of ribonucleotide reductase from Escherichia coli contains redox-active cysteine residues. In separate experiments, five conserved and 2 nonconserved cysteine residues were substituted with alanines by oligonucleotide-directed mutagenesis. The activities of the mutant proteins were determined in the presence of three different reductants: thioredoxin, glutaredoxin, or dithiothreitol. The results indicate two different classes of redox-active cysteines in ribonucleotide reductase: 1) C-terminal Cys-754 and Cys-759 responsible for the interaction with thioredoxin and glutaredoxin; and 2) Cys-225 and Cys-439 located at the nucleotide-binding site. Our classification of redox-active cysteines differs from the location of the active site cysteines in E. coli ribonucleotide reductase suggested previously (Lin, A.-N. I., Ashley, G. W., and Stubbe, J. (1987) Biochemistry 26, 6905-6909).  相似文献   

14.
The effect of mutations in the genes encoding glutathione, glutaredoxin, thioredoxin, and thioredoxin reductase on the response of growing Escherichia coli to oxidative stress was studied. The gshA mutants defective in glutathione synthesis had the lowest resistance to high doses of H2O2, whereas the trxB mutants defective in thioredoxin reductase synthesis had the highest resistance to this oxidant, exceeding that of the parent strain. Among the studied mutants, the trxB cells demonstrated the highest basic levels of catalase activity and intracellular glutathione; they were able to rapidly reach the normal GSH level after oxidative stress. At the same time, these bacteria showed high frequency of induced mutations. The expression of the katG and sulA genes suggests that, having different sensitivity to high oxidant concentrations, the studied mutants differ primarily in their ability to induce the antioxidant genes of the OxyR and SOS regulons.  相似文献   

15.
Ribonucleoside diphosphate reductase (EC1.17.4.1) was previously characterized in exponentially growing mouse L cells selectively permeabilized to small molecules by treatment with dextran sulfate (Kucera and Paulus, 1982b). This characterization has now been extended to cells in specific phases of the cell cycle and in transition between cell cycle phases, with activity studied both in situ (permeabilized cells) and in cell extracts. Cells at various stages in the cell cycle were obtained by unit-gravity sedimentation employing a commercially available reorienting chamber device, by G1 arrest induced by isoleucine limitation, and by metaphase arrest induced by Colcemid. G1 cells from both cycling and noncycling populations had negligible levels of ribonucleotide reductase activity as measured by CDP reduction both in situ and in extracts. When G1 arrested cells were allowed to progress to S phase, ribonucleotide reductase activity increased in parallel with [3H]thymidine incorporation into DNA. Ribonucleotide reductase activity in extracts increased at a somewhat greater rate than in situ activity. S phase ribonucleotide reductase activity measured in situ resembled the previously characterized activity in exponentially growing cells with respect to an absolute dependence on ATP or its analogs as positive allosteric effector, sensitivity to the negative allosteric effector dATP, and low susceptibility to stimulation by NADPH, dithiothreitol, and FeCl3. Disruption of permeabilized cells caused reductase activity to become highly dependent on the presence of both dithiothreitol and FeCl3. As synchronized cultures progressed from S into G2/M phase, no significant change in ribonucleotide reductase activity was seen. On the other hand, when cells that had been arrested in metaphase by Colcemid were allowed to resume cell cycle traversal by removing the drug, in situ ribonucleotide reductase activity decreased by 75% within 2.5 h. This decrease seemed to be a late mitotic event, since it was not correlated with the percentage of cells entering G1 phase. The cause of a subsequent slight increase of in situ ribonucleotide reductase activity is not clear. Parallel measurements of ribonucleotide reductase activity in cell extracts indicated also an initial decline accompanied by increasing dependence on added dithiols and FeCl3, followed by complete activity loss. Our results suggest a cell cycle pattern of ribonucleotide reductase activity that involves negligible levels in G1 phase, a progressive increase of activity upon entry into S phase paralleling overall DNA synthesis, continued retention of significant ribonucleotide reductase activity well into the metaphase period of mitosis, and a very rapid decline in activity during the later phases of mitosis. The periods of increase and decrease of ribonucleotide reductase activity were accompanied by modulation of the properties of the enzyme as indicated by differential changes in enzyme activity measured in situ and in extracts.  相似文献   

16.
Trypanosoma brucei, the causative agent of African sleeping sickness, synthesizes deoxyribonucleotides via a classical eukaryotic class I ribonucleotide reductase. The unique thiol metabolism of trypanosomatids in which the nearly ubiquitous glutathione reductase is replaced by a trypanothione reductase prompted us to study the nature of thiols providing reducing equivalents for the parasite synthesis of DNA precursors. Here we show that the dithiol trypanothione (bis(glutathionyl)spermidine), in contrast to glutathione, is a direct reductant of T. brucei ribonucleotide reductase with a K(m) value of 2 mm. This is the first example of a natural low molecular mass thiol directly delivering reducing equivalents for ribonucleotide reduction. At submillimolar concentrations, the reaction is strongly accelerated by tryparedoxin, a 16-kDa parasite protein with a WCPPC active site motif. The K(m) value of T. brucei ribonucleotide reductase for T. brucei tryparedoxin is about 4 micrometer. The disulfide form of trypanothione is a powerful inhibitor of the tryparedoxin-mediated reaction that may represent a physiological regulation of deoxyribonucleotide synthesis by the redox state of the cell. The trypanothione/tryparedoxin system is a new system providing electrons for a class I ribonucleotide reductase, in addition to the well known thioredoxin and glutaredoxin systems described in other organisms.  相似文献   

17.
Cells infected with varicella-zoster virus (VZV) express a viral ribonucleotide reductase which is distinct from that present in uninfected cells. VZV open reading frames 18 and 19 (ORF18 and ORF19) are homologous to the herpes simplex virus type 1 genes encoding the small and large subunits of ribonucleotide reductase, respectively. We generated recombinant VZV by transfecting cultured cells with four overlapping cosmid DNAs. To construct a virus lacking ribonucleotide reductase, we deleted 97% of VZV ORF19 from one of the cosmids. Transfection of this cosmid with the other parental cosmids yielded a VZV mutant with a 2.3-kbp deletion confirmed by Southern blot analysis. Virus-specific ribonucleotide reductase activity was not detected in cells infected with VZV lacking ORF19. Infection of melanoma cells with ORF19-deleted VZV resulted in plaques smaller than those produced by infection with the parental VZV. The mutant virus also exhibited a growth rate slightly slower than that of the parental virus. Chemical inhibition of the VZV ribonucleotide reductase has been shown to potentiate the anti-VZV activity of acyclovir. Similarly, the concentration of acyclovir required to inhibit plaque formation by 50% was threefold lower for the VZV ribonucleotide reductase deletion mutants than for parental virus. We conclude that the VZV ribonucleotide reductase large subunit is not essential for virus infection in vitro; however, deletion of the gene impairs the growth of VZV in cell culture and renders the virus more susceptible to inhibition by acyclovir.  相似文献   

18.
DNA-replication fork displacement rates were measured in mouse S49 lymphosarcoma cell lines and in derivatives of those cell lines. One of the derivatives lacks dCMP deaminase activity and two others bear defined mutations in ribonucleotide reductase. We also examined a revertant cell line that was selected from one of the ribonucleotide reductase mutants and has regained normal ribonucleotide reductase activity. Our results show a correlation between decreased fork-displacement rates and alterations in ribonucleotide reductase, suggesting a possible involvement of this enzyme in the replication apparatus.  相似文献   

19.
Two heat-sensitive (reversibly arrested in G1 phase at 39.5 degrees C, multiplying at 33 degrees C) and two cold-sensitive (reversibly arrested in G1 phase at 33 degrees C, multiplying at 39.5 degrees C) cell-cycle mutants of the P-815-X2 murine mastocytoma line were tested for ribonucleotide reductase activity, using cells made permeable to nucleotides. After transfer of the heat-sensitive mutant cells to 39.5 degrees C, ribonucleotide reductase activity, similar to thymidine kinase (Schneider, E., Müller, B. and Schindler, R. (1983) Biochim. Biophys. Acta 741, 77-85), but unlike DNA polymerase alpha (Schneider, E., Müller, B. and Schindler, R. (1985) Biochim. Biophys. Acta 825, 375-383), decreased rapidly and in parallel with numbers of cells in S phase, whereas in the cold-sensitive mutant cells brought to 33 degrees C, ribonucleotide reductase activity decreased approx. 8 h later than numbers of DNA-synthesizing cells. When arrested heat- or cold-sensitive mutant cells were returned to the permissive temperature, ribonucleotide reductase activities, similar to DNA polymerase alpha and to thymidine kinase in heat-sensitive mutants, increased essentially in parallel with reentry of cells into S phase, whereas the increase in thymidine kinase activity in the cold-sensitive mutants was previously shown to occur approx. one cell-cycle time later. This indicates that ribonucleotide reductase and thymidine kinase are coordinately expressed in the heat-sensitive, but independently regulated in the cold-sensitive mutants.  相似文献   

20.
We have constructed a plasmid, pAHOB1, with a 482-b AluI fragment containing the Escherichia coli glutaredoxin gene (grx) cloned under lambda PL promoter control. Growth of E. coli N4830/pAHOB1 cells at 30 degrees C followed by heat induction at 40 degrees C for 5 h resulted in expression of glutaredoxin as 20% of the soluble E. coli protein. Methods for the preparation of gram amounts of glutaredoxin and 5 mM solutions suitable for NMR studies were developed. About 10% of the glutaredoxin activity showed an unexpected higher isoelectric point and was isolated by DEAE-cellulose chromatography at pH 6.0. Sequence analysis demonstrated that this novel form (grx-90) contained five additional N-terminal residues (Met-Arg-Arg-Glu-Ile) added to the glutaredoxin molecule with 85 residues (grx-85). Grx-90 originates from an alternative ATG initiation codon present 5' of the previously identified translation start site on the grx gene in E. coli. Despite the highly charged N-terminal extension, grx-90 showed full activity as a GSH-disulfide oxidoreductase and the same apparent Km value (0.14 microM) as glutaredoxin in GSH-dependent reduction of CDP by ribonucleotide reductase. Both grx-90 and grx-85 showed identical competition curves in radioimmunoassays. The presence of grx-90 was also demonstrated in log-phase E. coli C600 cells as 5 to 10% of total glutaredoxin by immunological techniques. The molar extinction coefficient of native glutaredoxin (12,500 M-1 cm-1 at 280 nm) was 15% higher than expected from its content of one Trp and four Tyr residues.  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