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.
The soluble periplasmic subunit of the formate dehydrogenase FdhA of the tetrachloroethene-reducing anaerobe Sulfurospirillum multivorans was purified to apparent homogeneity and the gene (fdhA) was identified and sequenced. The purified enzyme catalyzed the oxidation of formate with oxidized methyl viologen as electron acceptor at a specific activity of 1683 nkat/mg protein. The apparent molecular mass of the native enzyme was determined by gel filtration to be about 100 kDa, which was confirmed by the fdhA nucleotide sequence. fdhA encodes for a pre-protein that differs from the truncated mature protein by an N-terminal 35-amino-acid signal peptide containing a twin arginine motif. The amino acid sequence of FdhA revealed high sequence similarities to the larger subunits of the formate dehydrogenases of Campylobacter jejuni, Wolinella succinogenes, Escherichia coli (FdhN, FdhH, FdhO), and Methanobacterium formicicum. According to the nucleotide sequence, FdhA harbors one Fe4/S4 cluster and a selenocysteine residue as well as conserved amino acids thought to be involved in the binding of a molybdopterin guanidine dinucleotide cofactor.Abbreviations Fdh Formate dehydrogenase - PCE Tetrachloroethene  相似文献   

2.
3.
Summary Three plants, R9201 and R11301 (from cv. Maris Mink) and R12202 (from cv. Golden Promise), were selected by screening M2 populations of barley (Hordeum vulgare L.) seedlings (mutagenised with azide in the M1) for resistance to 10 mM potassium chlorate. Selections R9201 and R11301 were crossed with the wild-type cv. Maris Mink and analysis of the F2 progeny showed that one quarter lacked shoot nitrate reductase activity. These F2 plants also withered and died in the continuous presence of nitrate as sole nitrogen source. Loss of nitrate reductase activity and withering and death were due in each case to a recessive mutation in a single nuclear gene. All F1 progeny derived from selfing selection R12202 lacked shoot nitrate reductase activity and also withered and subsequently died when maintained in the continuous presence of nitrate as sole nitrogen source. All homozygous mutant plants lacked not only shoot nitrate reductase activity but also shoot xanthine dehydrogenase activity. The plants took up nitrate, and possessed wild-type or higher levels of shoot nitrite reductase activity and NADH-cytochrome c reductase activity when treated with nitrate for 18 h. We conclude that loss of shoot nitrate reductase activity, xanthine dehydrogenase activity and withering and death, in the three mutants R9201, R11301 and R12202 is due to a mutation affecting the formation of a functional molybdenum cofactor. The mutants possessed wild-type levels of molybdenum and growth in the presence of unphysiologically high levels of molybdate did not restore shoot nitrate reductase or xanthine dehydrogenase activity. The shoot molybdenum cofactor of R9201 and of R12202 is unable to reconstitute NADPH nitrate reductase activity from extracts of the Neurospora crassa nit-1 mutant and dimerise the nitrate reductase subunits present in the respective barley mutant. The shoot molybdenum cofactor of R11301 is able to effect dimerisation of the R11301 nitrate reductase subunits and can reconstitute NADPH-nitrate reductase activity up to 40% of the wild-type molybdenum cofactor levels. The molybdenum cofactor of the roots of R9201 and R11301 is also defective. Genetic analysis demonstrated that R9201, but not R11301, is allelic to R9401 and Az34 (nar-2a), two mutants previously shown to be defective in synthesis of molybdenum cofactor. The mutations in R9401 and R9201 gave partial complementation of the nar-2a gene such that heterozygotes had higher levels of extractable nitrate reductase activity than the homozygous mutants.We conclude that: (a) the nar-2 gene locus encodes a step in molybdopterin biosynthesis; (b) the mutant R11301 represents a further locus involved in the synthesis of a functional molybdenum cofactor; (c) mutant Rl2202 is also defective in molybdopterin biosynthesis; and (d) the nar-2 gene locus and the gene locus defined by R11301 govern molybdenum cofactor biosynthesis in both shoot and root.  相似文献   

4.
5.
Xanthine dehydrogenase has been purified from Pseudomonas aeruginosa cultured on a rich medium and induced with hypoxanthine. The enzyme was shown to contain FAD, iron sulfur centers and a molybdenum cofactor as prosthetic groups. Analysis of the molybdenum cofactor in this enzyme has revealed that the cofactor contains molybdopterin (MPT) rather than molybdopterin guanine dinucleotide or molybdopterin cytosine dinucleotide which have previously been identified in a number of molybdoenzymes of bacterial origin. The pterin cofactor in P.aeruginosa xanthine dehydrogenase was alkylated and the resulting product was identified as dicarboxamidomethyl molybdopterin. In addition, the pterin released from the enzyme by denaturation with guanidine-HCl was found to chromatograph on Sephadex G-15 with an apparent molecular weight of 350. These results document the first example of a bacterial enzyme with a molybdenum cofactor comprising molybdopterin and the metal only.  相似文献   

6.
Reconstitution of the apoprotein of the molybdoenzyme nitrate reductase in extracts of the Neurospora crassa mutant nit-1 with molybdenum cofactor released by denaturation of purified molybdoenzymes is efficient in the absence of exogenous MoO42? under defined conditions. Evidence is presented that this molybdate-independent reconstitution is due to transfer of intact Mo cofactor, a complex of Mo and molybdopterin (MPT), the organic constituent of the cofactor. This complex can be separated from denatured protein by gel filtration, and from excess MoO42? by reverse-phase HPLC. Sulfite oxidase, native xanthine dehydrogenase, and cyanolyzed xanthine dehydrogenase are equipotent Mo cofactor donors. Other well-studied inactive forms of xanthine dehydrogenase are also shown to be good cofactor sources. Using xanthine dehydrogenase specifically radiolabeled in the cyanolyzable sulfur, it is shown that this terminal ligand of Mo is rapidly removed from Mo cofactor under the conditions used for reconstitution.  相似文献   

7.
2-Hydroxyisonicotinate dehydrogenase from Mycobacterium sp. INA1 was purified 26-fold to apparent homogeneity. The enzyme is involved in isonicotinate degradation by Mycobacterium sp. INA1 and catalyzes the conversion of 2-hydroxyisonicotinate to 2,6-dihydroxypyridine-4-carboxylate. The purified protein exhibited a native molecular mass of 300 kDa and subunits of 97, 31 and 17 kDa, respectively, indicating an α2β2γ2 structure. The absorption spectrum of the homogeneous enzyme was characteristic for an iron/sulfur flavoprotein. 3.8 mol of iron, 3.7 mol of acid labile sulfur, 0.94 mol of FAD and 0.75 mol of molybdenum were determined per mol of protomer. The molybdenum cofactor was identified as molybdopterin cytosine dinucleotide. 2-Hydroxyisonicotinate dehydrogenase was inactivated in the presence of cyanide. According to these basic properties the protein seems to belong to the class of molybdo-iron/sulfur flavoproteins of the xanthine oxidase family.  相似文献   

8.
Summary The molybdenum cofactor of the barley mutant R9401 is not able to reconstitute NADPH nitrate reductase activity from extracts of the N. crassa nit-1 mutant nor is it able to effect dimerisation of the nitrate reductase subunits present in the R9401 mutant. Unphysiologically high levels of molybdate cannot restore nitrate reductase and xanthine dehydrogenase activity to mutant R9401 in vivo nor reactivate the Mo-co factor in vitro. The results indicate that the defect in mutant R9401 lies in the pathway leading to the formation of a functional molybdopterin moiety and that the same nuclear gene is involved in the synthesis of both shoot and root molybdenum cofactor.Abbreviations BSA bovine serum albumen - GSH glutathione (reduced) - NEM N-ethylmaleimide  相似文献   

9.
10.
The sulfate-reducing bacterimDesulfovibrio strain HDv (DSM 6830) grew faster on (S)- and on (R, S)-1,2-propanediol (μmax 0.053 h) than on (R)-propanediol (0.017 h−1) and ethanol (0.027 h−1). From (R, S)-1,2-propanediol-grown cells, an alcohol dehydrogenase was purified. The enzyme was oxygen-labile, NAD-dependent, and decameric; the subunit mol. mass was 48 kDa. The N-terminal amino acid sequence indicated similarity to alcohol dehydrogenases belonging to family III of NAD-dependent alcohol dehydrogenases, the first 21 N-terminal amino acids being identical to those of theDesulfovibrio gigas alcohol dehydrogenase. Best substrates were ethanol and propanol (K m of 0.48 and 0.33 mM, respectively). (R, S)-1,2-Propanediol was a relatively poor substrate for the enzyme, but activities in cell extracts were high enough to account for the growth rate. The enzyme showed a preference for (S)-1,2-propanediol over (R)-1,2-propanediol. Antibodies raised against the alcohol dehydrogenase ofD. gigas showed cross-reactivity with the alcohol dehydrogenase ofDesulfovibrio strain HDv and with cell extracts of six other ethanol-grown sulfate-reducing bacteria.  相似文献   

11.
Summary The structural genes (hup) of the H2 uptake hydrogenase of Rhodobacter capsulatus were isolated from a cosmid gene library of R. capsulatus DNA by hybridization with the structural genes of the H2 uptake hydrogenase of Bradyrhizobium japonicum. The R. capsulatus genes were localized on a 3.5 kb HindIII fragment. The fragment, cloned onto plasmid pAC76, restored hydrogenase activity and autotrophic growth of the R. capsulatus mutant JP91, deficient in hydrogenase activity (Hup-). The nucleotide sequence, determined by the dideoxy chain termination method, revealed the presence of two open reading frames. The gene encoding the large subunit of hydrogenase (hupL) was identified from the size of its protein product (68108 dalton) and by alignment with the NH2 amino acid protein sequence determined by Edman degradation. Upstream and separated from the large subunit by only three nucleotides was a gene encoding a 34 256 dalton polypeptide. Its amino acid sequence showed 80% identity with the small subunit of the hydrogenase of B. japonicum. The gene was identified as the structural gene of the small subunit of R. capsulatus hydrogenase (hupS). The R. capsulatus hydrogenase also showed homology, but to a lesser extent, with the hydrogenase of Desulfovibrio baculatus and D. gigas. In the R. capsulatus hydrogenase the Cys residues, (13 in the small subunit and 12 in the large subunit) were not arranged in the typical configuration found in [4Fe–4S] ferredoxins.  相似文献   

12.
When 10 strains of lactic acid bacteria were incubated with 5′-hydroxyaverantin (HAVN), a precursor of aflatoxins, seven of them converted HAVN to averufin; the same reaction is found in aflatoxin biosynthesis of aflatoxigenic fungi. These bacteria had a dehydrogenase that catalyzed the reaction from HAVN to 5′-oxoaverantin (OAVN), which was so unstable that it was easily converted to averufin. The enzyme was purified from Lactobacillus brevis IFO 12005. The molecular mass of the enzyme was 100 kDa on gel filtration chromatography and 33 kDa on SDS polyacrylamide gel electrophoresis (SDS–PAGE). The gene encoding the enzyme was cloned and sequenced. The deduced protein consisted of 249 amino acids, and its estimated molecular mass was 25,873, in agreement with that by time of flight mass spectrometry (TOF MS) analysis. Although the deduced amino acid sequence showed about 50% identity to those reported for alcohol dehydrogenases from L. brevis or L. kefir, the commercially available alcohol dehydrogenase from L. kefir did not convert HAVN to OAVN. Aspergillus parasiticus HAVN dehydrogenase showed about 25% identity in amino acid sequence with the dehydrogenase and also with these two alcohol dehydrogenases.  相似文献   

13.
Peroxisomes are subcellular organelles with an essentially oxidative type of metabolism. The presence in these organelles of superoxide dismutases and the generation of superoxide radicals (O2??) was first demonstrated in plant tissues and in recent years different experimental evidence has suggested the existence of cellular functions related to activated oxygen species. Some of these functions are analyzed in this work. In purified intact peroxisomes from pea (Pisum sativum L.) leaves, xanthine oxidase and urate oxidase were found to be present. The occurrence and the level of the metabolites xanthine, hypoxanthine, uric acid, and allantoin were studied in extracts of pea leaf peroxisomes by HPLC. Xanthine, uric acid, and allantoin were detected in peroxisomes. These results suggest a cellular role for leaf peroxisomes in the catabolism of purines. In peroxisomal membranes, 3 polypeptides (PMPs) with molecular masses of 18, 29 and 32 kDa, respectively, have been shown to generate superoxide radicals. These PMPs were purified from pea leaf peroxisomal membranes and characterized. While the 18- and 32-kDa PMPs use NADH as electron donor for O2?? production, the 29-kDa PMP was clearly dependent on NADPH. Very recently, the occurrence in pea leaf peroxisomes of all the enzymes of the ascorbate-glutathione cycle has been demonstrated. NADPH is required for the glutathione reductase activity of the cycle and this implies the reduction of NADP+ to NADPH. This recycling function could be carried out by the NADP-dependent glucose-6-phosphate dehydrogenase (G6PDH), 6-phosphogluconate dehydrogenase (6PGDH), and isocitrate dehydrogenase (ICDH). These 3 dehydrogenases have been demonstrated to be present in the matrix of pea leaf peroxisomes. The catabolism of purines, the superoxide-generating PMPs, the ascorbate-glutathione cycle, and the dehydrogenase-mediated recycling of NADPH, are activated oxygen roles of leaf peroxisomes that add to other functions previously known for peroxisomes from eukaryotic cells.  相似文献   

14.
DNA sequence analysis of a 3494-bp HindIII-Bc1I fragment of the Rhodobacter capsulatus nif region A revealed genes that are homologous to ORF6, nifU, nifS, nifV and nifW from Azotobacter vinelandii and Klebsiella pneumoniae. R. capsulatus nifU, which is present in two copies, encodes a novel type of NifU protein. The deduced amino acid sequences of NifUI and NifUII share homology only with the C-terminal domain of NifU from A. vinelandii and K. pneurnoniae. In contrast to nifA andnifB which are almost perfectly duplicated, the predicted amino acid sequences of the two NifU proteins showed only 39% sequence identity. Expression of the ORF6-nifU ISVW operon, which is preceded by a putative 54-dependent promoter, required the function of NifA and the nif-specific rpoN gene product encoded by nifR4. Analysis of defined insertion and deletion mutants demonstrated that only nifS was absolutely essential for nitrogen fixation in R. capsulatus. Strains carrying mutations in nifV were capable of very slow diazotrophic growth, whereas ORF6, nifU I and nifW mutants as well as a nifU I/nifUII, double mutant exhibited a Nif+ phenotype. Interestingly, R. capsulatus nifV mutants were able to reduce acetylene not only to ethylene but also to ethane under conditions preventing the expression of the alternative nitrogenase system. Homocitrate added to the growth medium repressed ethane formation and cured the NifV phenotype in R. capsulatus. Higher concentrations of homocitrate were necessary to complement the NifV phenotype of a polar nifV mutant (NifVNifW), indicating a possible role of NifW either in homocitrate transport or in the incorporation of this compound into the iron-molybdenum cofactor of nitrogenase.  相似文献   

15.
The xanthine oxidases and dehydrogenases are among the most conserved enzymes in all living kingdoms. They contain the molybdopterin cofactor Moco. We show here that in the fungi, in addition to xanthine dehydrogenase, a completely different enzyme is able to catalyse the oxidation of xanthine to uric acid. In Aspergillus nidulans this enzyme is coded by the xanA gene. We have cloned the xanA gene and determined its sequence. A deletion of the gene has the same phenotype as the previously known xanA1 miss-sense mutation. Homologues of xanA exist only in the fungal kingdom. We have inactivated the cognate gene of Schizosaccharomyces pombe and this results in strongly impaired xanthine utilization as a nitrogen source. We have shown that the Neurospora crassa homologue is functionally equivalent to xanA. The enzyme coded by xanA is an alpha-ketoglutarate- and Fe(II)-dependent dioxygenase which shares a number of properties with other enzymes of this group. This work shows that only in the fungal kingdom, an alternative mechanism of xanthine oxidation, not involving Moco, has evolved using the dioxygenase scaffold.  相似文献   

16.
A thermostable homodimeric isocitrate dehydrogenase from the hyperthermophilic sulfate-reducing archaeon Archaeoglobus fulgidus was purified and characterized. The mol. mass of the isocitrate dehydrogenase subunit was 42 kDa as determined by SDS-PAGE. Following separation by SDS-PAGE, A. fulgidus isocitrate dehydrogenase could be renatured and detected in situ by activity staining. The enzyme showed dual coenzyme specificity with a high preference for NADP+. Optimal temperature for activity was 90° C or above, and a half-life of 22 min was found for the enzyme when incubated at 90° C in a 50 mM Tricine-KOH buffer (pH 8.0). Based on the N-terminal amino acid sequence, the gene encoding the isocitrate dehydrogenase was cloned. DNA sequencing identified the icd gene as an open reading frame encoding a protein of 412 amino acids with a molecular mass corresponding to that determined for the purified enzyme. The deduced amino acid sequence closely resembled that of the isocitrate dehydrogenase from the archaeon Caldococcus noboribetus (59% identity) and bacterial isocitrate dehydrogenases, with 57% identity with isocitrate dehydrogenase from Escherichia coli. All the amino acid residues directly contacting substrate and coenzyme (except Ile-320) in E. coli isocitrate dehydrogenase are conserved in the enzyme from A. fulgidus. The primary structure of A. fulgidus isocitrate dehydrogenase confirmes the presence of Bacteria-type isocitrate dehydrogenases among Archaea. Multiple alignment of all the available amino acid sequences of di- and multimeric isocitrate dehydrogenases from the three domains of life shows that they can be divided into three distinct phylogenetic groups. Received: 6 February 1997 / Accepted: 12 June 1997  相似文献   

17.
Xanthine dehydrogenase (XDH) from Pseudomonas putida 86, which was induced 65-fold by growth on hypoxanthine, was purified to homogeneity. It catalyzes the oxidation of hypoxanthine, xanthine, purine, and some aromatic aldehydes, using NAD+ as the preferred electron acceptor. In the hypoxanthine:NAD+ assay, the specific activity of purified XDH was 26.7 U (mg protein)−1. Its activity with ferricyanide and dioxygen was 58% and 4%, respectively, relative to the activity observed with NAD+. XDH from P. putida 86 consists of 91.0 kDa and 46.2 kDa subunits presumably forming an α4β4 structure and contains the same set of redox-active centers as eukaryotic XDHs. After reduction of the enzyme with xanthine, electron paramagnetic resonance (EPR) signals of the neutral FAD semiquinone radical and the Mo(V) rapid signal were observed at 77 K. Resonances from FeSI and FeSII were detected at 15 K. Whereas the observable g factors for FeSII resemble those of other molybdenum hydroxylases, the FeSI center in contrast to most other known FeSI centers has nearly axial symmetry. The EPR features of the redox-active centers of P. putida XDH are very similar to those of eukaryotic XDHs/xanthine oxidases, suggesting that the environment of each center and their functionality are analogous in these enzymes. The midpoint potentials determined for the molybdenum, FeSI and FAD redox couples are close to each other and resemble those of the corresponding centers in eukaryotic XDHs.  相似文献   

18.
E. Fernández  J. Cárdenas 《Planta》1981,153(3):254-257
Wild-type Chlamydomonas reinhardii cells have xanthine dehydrogenase activity when grown with nitrate, nitrite, urea, or amino acid media. Mutant strains 102, 104, and 307 of Chlamydomonas, lacking both xanthine dehydrogenase and nitrate reductase activities, were incapable of restoring the NADPH-nitrate reductase activity of the mutant nit-1 of Neurospora crassa, whereas wild type cells and mutants 203 and 305 had xanthine dehydrogenase and were able to reconstitute the nitrate reductase activity of nit-1 of Neurospora. Therefore, it is concluded that in Chlamydomonas a common cofactor is shared by xanthine dehydrogenase and nitrate reductase. Xanthine dehydrogenase is repressed by ammonia and seems to be inessential for growth of Chlamydomonas.  相似文献   

19.
Methanobacterium thermoautotrophicum (strain Marburg) was found to grow on media supplemented with tungstate rather than with molybdate. The Archaeon then synthesized a tungsten iron-sulfur isoenzyme of formylmethanofuran dehydrogenase. The isoenzyme was purified to apparent homogeneity and shown to be composed of four different subunits of apparent molecular masses 65 kDa, 53 kDa, 31 kDa, and 15 kDa and to contain per mol 0.4 mol tungsten, <0.05 mol molybdenum, 8 mol non-heme iron, 8 mol acid-labile sulfur and molybdopterin guanine dinucleotide. Its molecular and catalytic properties were significantly different from those of the molybdenum isoenzyme characterized previously. The two isoenzymes also differed in their metal specificity: the active molybdenum isoenzyme was only synthesized when molybdenum was available during growth whereas the active tungsten isoenzyme was also generated during growth of the cells on molybdate medium. Under the latter conditions the tungsten isoenzyme was synthesized containing molybdenum rather than tungsten.Abbreviations MFR methanofuran - CHO-MFR N-formylmethanofuran - MGD molybdopterin guanine dinucleotide - MAD molybdopterin adenine dinucleotide - MHD molybdopterin hypoxanthine dinucleotide - FPLC fast protein liquid chromatography - SDS/PAGE sodium dodecylsulfate/polyacrylamide gel electrophoresis - ICP-MS inductively coupled plasma mass spectrometry  相似文献   

20.
Toluene and related aromatic compounds are anaerobically degraded by the denitrifying bacterium Thauera sp. strain K172 via oxidation to benzoyl-CoA. The postulated initial step is methylhydroxylation of toluene to benzyl alcohol, which is either a free or enzyme-bound intermediate. Cells grown with toluene or benzyl alcohol contained benzyl alcohol dehydrogenase, which is possibly the second enzyme in the proposed pathway. The enzyme was purified from benzyl-alcohol-grown cells and characterized. It has many properties in common with benzyl alcohol dehydrogenase from Acinetobacter and Pseudomonas species. The enzyme was active as a homotetramer of 160kDa, with subunits of 40kDa. It was NAD+-specific, had an alkaline pH optimum, and was inhibited by thiol-blocking agents. No evidence for a bound cofactor was obtained. Various benzyl alcohol analogues served as substrates, whereas non-aromatic alcohols were not oxidized. The N-terminal amino acid sequence indicates that the enzyme belongs to the class of long-chain Zn2+-dependent alcohol dehydrogenases, although it appears not to contain a metal ion that can be removed by complexing agents.Dedicated to Prof. Achim Trebst  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