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.
The function of the MoeA protein in the biosynthesis of the molybdenum cofactor (MoCo) was analyzed in vitro, using purified His(6)-MoeA from Escherichia coli, molybdopterin (MPT) isolated from buttermilk xanthine oxidase and molybdate. The formation of MoCo was monitored by the reconstitution of nitrate reductase activity in extracts of the Neurospora crassa nit-1 mutant. Formation of MoCo from MPT and molybdate required MoeA and L-cysteine or glutathione. The reaction proceeded at micromolar molybdate levels and was time- and MoeA concentration-dependent. A physical interaction between MoeA and MPT was demonstrated by HPLC analysis of MoeA-bound MPT.  相似文献   

2.
The pterin cofactor in formate dehydrogenase isolated from Methanobacterium formicium is identified as molybdopterin guanine dinucleotide. The pterin, stabilized as the alkylated, dicarboxamidomethyl derivative, is shown to have absorption and chromatographic properties identical to those of the previously characterized authentic compound. Treatment with nucleotide pyrophosphatase produced the expected degradation products GMP and carboxyamidomethyl molybdopterin. The molybdopterin guanine dinucleotide released from the enzyme by treatment with 95% dimethyl sulfoxide is shown to be functional in the in vitro reconstitution of the cofactor-deficient nitrate reductase in extracts of the Neurospora crassa nit-1 mutant.  相似文献   

3.
The metabolic status of a patient previously characterized as deficient in sulfite oxidase was reexamined applying new methodology which has been developed to distinguish between a defect specific to the sulfite oxidase protein and sulfite oxidase deficiency which arises as a result of molybdenum cofactor deficiency. Urothione, the metabolic degradation product of the molybdenum cofactor, was undetectable in urine samples from the patient. Analysis of molybdenum cofactor levels in fibroblasts by monitoring reconstitution of apo nitrate reductase in extracts of the Neurospora crassa mutant nit-1 revealed that cells from the patient were severely depleted. Quantitation of urinary oxypurines showed that hypoxanthine and xanthine were highly elevated while uric acid remained in the normal range. These results were interpreted to indicate a severe but incomplete deficiency of the molybdenum cofactor. The presence of very low levels of active cofactor, supporting the synthesis of low levels of active sulfite oxidase and xanthine dehydrogenase, could explain the metabolic patterns of sulfur and purine products and the relatively mild clinical symptoms in this individual.  相似文献   

4.
A simple and reliable procedure of oxidation of molybdenum cofactor (MoCo) from molybdoenzymes by autoclaving samples at 120 degrees C for 20 min yielded a single predominant fluorescent species that could be quantitatively determined by reverse phase high performance liquid chromatography. This method allowed detection and quantitation of molybdopterin in cell-free extracts of the green alga Chlamydomonas reinhardtii. The MoCo oxidation product from C. reinhardtii has the same chromatographic and spectral properties as that of milk xanthine oxidase and chicken liver sulfite oxidase. The oxidized species was also detected in molybdenum cofactor mutants of Chlamydomonas reinhardtii defective at the nit-3, nit-4, nit-5/nit-6 and nit-7 loci, which strongly suggests that active molybdenum cofactor itself is not directly involved in the control of its own biosynthetic pathway.  相似文献   

5.
Two new nitrate reductase-deficient mutants from Chlamydomonas reinhardtii have been genetically and biochemically characterized. Both H1 and F23 mutants carry single recessive allelic mutations that map at a new locus designated nit-7. This locus is unlinked to the other six nit loci related to the nitrate assimilation pathway in C. reinhardtii. Both mutant alleles H1 and F23 lack an active molybdopterin cofactor, the activity of which is restored neither in vitro nor in vivo by high concentrations of molybdate. Nitrate reductase subunits in these mutants seem to assemble, although not in a stable form, in a high molecular weight complex and, as in other molybdenum cofactor-defective mutants of C. reinhardtii, they cannot reconstitute nitrate reductase activity with an active molybdenum cofactor source from extracts of ammonium-grown cells. The results suggest that nit-7 mutants are defective in molybdopterin biosynthesis. They do produce some precursor(s) that are capable of binding to nitrate reductase subunits.  相似文献   

6.
Formate dehydrogenase ( FDH ) from Clostridium thermoaceticum is a known tungsten enzyme. FDH was tested for the presence of nitrogenase-type cofactor and nitrate reductase-type cofactor by the Azotobacter vinelandii UW-45 and Neurospora crassa nit-1 reconstitution assays, respectively. Tungsten formate dehydrogenase (W- FDH ), containing only a small Mo impurity, activated the nit-1 nitrate reductase extracts when molybdate was also added, but not when tungstate was added. These results show W- FDH contains the cofactor common to all known Mo-enzymes except nitrogenase. The difference between the redox chemistries of W- FDH and W-substituted sulfite oxidase appears to relate to differences in tungsten ligation other than that donated by the cofactor or to variations in the protein environment surrounding the tungsten active site.  相似文献   

7.
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   

8.
Escherichia coli trimethylamine N-oxide (TMAO) reductase I, the major enzyme among inducible TMAO reductases, was purified to homogeneity by an improved method including heat treatment, ammonium sulfate precipitation, and chromatographies on Bio-Gel A-1.5m, DEAE-cellulose, and Reactive blue-agarose. The molecular weight was estimated by gel filtration to be approximately 200,000. A single subunit peptide with a molecular weight of 95,000 was found by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. This enzyme contained 1.96 atoms of molybdenum, 0.96 atoms of iron, 1.52 atoms of zinc, and less than 0.4 atoms of acid-labile sulfur per molecular weight of 200,000. The absorption spectrum of the enzyme showed a peak at 278 nm and a shoulder at 288 nm, but no characteristic absorption was found from 350 to 700 nm. A fluorescent derivative of molybdenum cofactor was found when the enzyme was boiled with iodine in acidic solution; its fluorescence spectra were almost the same as those of the form A derivative of molybdopterin found in sulfite oxidase. The molybdenum cofactor released from heated TMAO reductase I reconstituted nitrate reductase in the extracts of Neurospora crassa mutant strain nit-1 lacking molybdenum cofactor. Thus, TMAO reductase I contains molybdopterin, which is a common constituent of some molybdenum-containing enzymes. Some kinetic properties were also determined.  相似文献   

9.
The molybdopterin cofactor from the formate dehydrogenase of Methanobacterium formicicum was studied. The cofactor was released by guanidine denaturation of homogeneous enzyme, which also released greater than 80% of the molybdenum present in the enzyme. The anoxically isolated cofactor was nonfluorescent, but after exposure to air it fluoresced with spectra similar to those of described molybdopterin cofactors. Aerobic release from acid-denatured formate dehydrogenase in the presence of I2 and potassium iodide produced a mixture of fluorescent products. Alkaline permanganate oxidation of the mixture yielded pterin-6-carboxylic acid as the only detectable fluorescent product. The results showed that the cofactor from formate dehydrogenase contained a pterin nucleus with a 6-alkyl side chain of unknown structure. Covalently bound phosphate was also present. The isolated cofactor was unable to complement the cofactor-deficient nitrate reductase of the Neurospora crassa nit-1 mutant.  相似文献   

10.
Summary Four allelic putative cnx (molybdenum-cofactor defective) cell lines (O42, P12, P31 and P47) of Nicotiana tabacum var. Xanthi, biochemically and genetically distinct from N. tabacum var. Gatersleben cnxA mutants, were examined further. Their molybdenum-cofactor could efficiently reconstitute NADPH-nitrate reductase activity from Neurospora crassa mutant nit-1 extract only in the presence of exogenous molybdenum unlike that of the wild-type cofactor which could reconstitute NADPH-nitrate reductase activity in either the absence or presence of exogenous molybdenum. Loss of cofactor activity in vivo was not due to a defect in molybdenum uptake into the cells. In vitro nitrate reductase complementation between extracts of each of these four lines and a nia mutant showed that they possessed a functional nitrate reductase haemoflavoprotein subunit. Both constitutive molybdenum cofactor and NADH cytochrome c reductase activity were derepressed in the four cell lines. These results show that the four cell lines are indeed altered at a cnx locus, called cnxB, that the defect is probably in molybdenum processing and that there is a link between synthesis of functional molybdenum cofactor and nitrate reductase aporprotein.  相似文献   

11.
Cell-free extracts of Proteus mirabilis were able to reconstitute NADPH-dependent assimilatory nitrate reductase in crude extracts of the Neurospora crassa mutant strain nit-1, lacking molybdenum cofactor. Molybdenum cofactor was formed in the cytoplasm of the bacterium even in the presence of oxygen during growth though under these conditions no molybdo enzymes are formed. As a consequence no cofactor could be released by acid treatment from membranes of cells grown aerobically. The amount of cofactor released from membranes of cells grown anaerobically under various conditions was proportional to the amount of molybdo enzymes formed. During growth in the presence of tungstate a cofactor, which lacks molybdenum, was found in the cytoplasm. For detection of this so-called demolybdo cofactor the presence of molybdate during reconstitution was essential. Moreover, the cytoplasmic cofactor pool in cells grown in the presence of tungstate appeared to be two to three times higher than in cells grown under similar conditions without tungstate. After anaerobic growth in the presence of tungstate, the inactive demolybdo reductases were shown to contain partly no cofactor and partly a demolybdo cofactor. The P. mirabilis chlorate resistant mutant S 556 did not contain molybdenum cofactor. In two other chl-mutants the cofactor activity was the same as in the wild type.  相似文献   

12.
We have previously shown that Escherichia coli MoeA and MogA are required in vivo for the final step of molybdenum cofactor biosynthesis, the addition of the molybdenum atom to the dithiolene of molybdopterin. MoeA was also shown to facilitate the addition of molybdenum in an assay using crude extracts from E. coli moeA(-) cells. The experiments detailed in this report utilized an in vitro assay for MoeA-mediated molybdenum ligation to de novo synthesized molybdopterin using only purified components and monitoring the reconstitution of human aposulfite oxidase. In this assay, maximum activation was achieved by delaying the addition of aposulfite oxidase to allow for adequate molybdenum coordination to occur. Tungsten, which substitutes for molybdenum in hyperthermophilic organisms, could also be ligated to molybdopterin using this system, though not as efficiently as molybdenum. Addition of thiol compounds to the assay inhibited activity. Addition of MogA also inhibited the reaction. However, in the presence of ATP and magnesium, addition of MogA to the assay increased the level of aposulfite oxidase reconstitution beyond that observed with MoeA alone. This effect was not observed in the absence of MoeA. The results presented here demonstrate that MoeA is responsible for mediating molybdenum ligation to molybdopterin, whereas MogA stimulates this activity in an ATP-dependent manner.  相似文献   

13.
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   

14.
The mob genes of several bacteria have been implicated in the conversion of molybdopterin to molybdopterin guanine dinucleotide. The mob locus of Rhodobacter sphaeroides WS8 comprises three genes, mobABC. Chromosomal in-frame deletions in each of the mob genes have been constructed. The mobA mutant strain has inactive DMSO reductase and periplasmic nitrate reductase activities (both molybdopterin guanine dinucleotide-requiring enzymes), but the activity of xanthine dehydrogenase, a molybdopterin enzyme, is unaffected. The inability of a mobA mutant to synthesise molybdopterin guanine dinucleotide is confirmed by analysis of cell extracts of the mobA strain for molybdenum cofactor forms following iodine oxidation. Mutations in mobB and mobC are not impaired for molybdoenzyme activities and accumulate wild-type levels of molybdopterin and molybdopterin guanine dinucleotide, indicating they are not compromised in molybdenum cofactor synthesis. In the mobA mutant strain, the inactive DMSO reductase is found in the periplasm, suggesting that molybdenum cofactor insertion is not necessarily a pre-requisite for export.  相似文献   

15.
There were significant differences in the contents of molybdenum cofactor (Mo-co), both in a low-molecular-mass form (free Mo-co) and in a protein-bound form, in seeds of sevenVicia faba genotypes. Low-molecular-mass Mo-co species present in the extracts were detected by their ability to reactivate, through a dialysis membrane, aponitrate reductase from theNeurospora crassa nit-1 mutant. In extracts of all genotypes tested, the amount of Mo-co capable of directly reactivating nitrate reductase of theN. crassa nit-1 mutant was always much higher than that of low-molecular-mass Moco. These data cannot be explained by considering, as traditionally, that Mo-co detected directly, i.e. without any previous treatment for its release from Mo-coproteins, corresponds to free low-molecular mass Mo-co. A protein which bound Mo-co was purified to electrophoretic homogeneity. This protein consisted of a single 70-kDa polypeptide chain and carried a Mo-co that could be efficiently released when in contact with aponitrate reductase.Abbreviations CP carrier protein - Mo-co molybdenum cofactor - NR nitrate reductase - XO xanthine oxidase  相似文献   

16.
Experiments were performed to determine whether defects in molybdenum cofactor metabolism were responsible for the pleiotropic loss of the molybdoenzymes nitrate reductase and formate dehydrogenase in chl mutants of Escherichia coli. In wild-type E. coli, molybdenum cofactor activity was present in both the soluble and membrane-associated fractions when the cells were grown either aerobically or anaerobically, with and without nitrate. Molybdenum cofactor in the soluble fraction decreased when the membrane-bound nitrate reductase and formate dehydrogenase were induced. In the chl mutants, molybdenum cofactor activity was found in the soluble fraction of chlA, chlB, chlC, chlD, chlE, and chlG, but only chlB, chlC, chlD, and chlG expressed cofactor activity in the membrane fraction. The defect in the chlA mutants which prevented incorporation of the soluble cofactor into the membrane also caused the soluble cofactor to be defective in its ability to bind molybdenum. This cofactor was not active in the absence of molybdate, and it required at least threefold more molybdate than did the wild type in the Neurospora crassa nit-1 complementation assay. However, the cofactor from the chlA strain mediated the dimerization of the nit-1 subunits in the presence and absence of molybdate to yield the 7.9S dimer. Growth of chlA mutants in medium with increased molybdate did not repair the defect in the chlA cofactor nor restore the molybdoenzyme activities. Thus, molybdenum cofactor was synthesized in all the chl mutants, but additional processing steps may be missing in chlA and chlE mutants for proper insertion of cofactor in the membrane.  相似文献   

17.
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   

18.
Extracts of Aspergillus nidulans wild type (bi-1) and the nitrate reductase mutant niaD-17 were active in the in vitro restoration of NADPH-dependent nitrate reductase when mixed with extracts of Neurospora crassa, nit-1. Among the A. nidulans cnx nitrate reductase mutants tested, only the molybdenum repair mutant, cnxE-14 grown in the presence of 10-minus 3 M Na2 MoO4 was active in the restoration assay. Aspergillus extracts contained an inhibitor(s) which was measured by the decrease in NADPH-dependent nitrate reductase formed when extracts of Rhodospirillum rubrum and N. crassa, nit-1 were incubated at room temperature. The inhibition by extracts of A. nidulans, bi-1, cnxE-14, cnxG-4 and cnxH-3 was a linear function of time and a logarithmic function of the protein concentration in the extract. The molybdenum content of N. crassa wild type and nit-1 mycelia were found to be similar, containing approx. 10 mu g molybdenum/mg dry mycelium. The NADPH-dependent cytochrome c reductase associated with nitrate reductase was purified from both strains. The NADPH-dependent cytochrome c reductase associated with nitrate reductase was purified from both strains. The enzyme purified from wild-type N. crassa contained more than 1 mol of molybdenum per mol of enzyme, whereas the enzyme purified from nit-1 contained negligible amounts of molybdenum.  相似文献   

19.
All molybdenum enzymes except nitrogenase contain a common molybdenum cofactor, whose organic moiety is a novel pterin called molybdopterin (MPT). To assist in elucidating the biosynthetic pathway of MPT, two MPT-deficient mutants of Escherichia coli K-12 were isolated. They lacked activities of the molybdenum enzymes nitrate reductase and formate dehydrogenase, did not reconstitute apo nitrate reductase from a Neurospora crassa nit-1 strain, and did not yield form A, a derivative of MPT. By P1 mapping, these two mutations mapped to chlA and chlE, loci previously postulated but never definitely shown to be involved in MPT biosynthesis. The two new mutations are in different genetic complementation groups from previously isolated chlA and chlE mutations and have been designated as chlM and chlN (closely linked to chlA and chlE, respectively). The reported presence of Mo cofactor activity in the chlA1 strain is shown to be due to in vitro synthesis of MPT through complementation between a trypsin-sensitive macromolecule from the chlA1 strain and a low-molecular-weight compound from the nit-l strain.  相似文献   

20.
All molybdoenzyme activities are absent in chlB mutants because of their inability to synthesize molybdopterin guanine dinucleotide, which together with molybdate constitutes the molybdenum cofactor in Escherichia coli. The chlB mutants are able to synthesize molybdopterin. We have previously shown that the inactive nitrate reductase present in a chlB mutant can be activated in a process requiring protein FA and a heat-stable low-molecular-weight substance. We show here that purified nitrate reductase from the soluble fraction of a chlB mutant can be partially activated in a process that requires protein FA, GTP, and an additional protein termed factor X. It appears that the molybdopterin present in the nitrate reductase of a chlB mutant is converted to molybdopterin guanine dinucleotide during activation. The activation is absolutely dependent upon both protein FA and factor X. Factor X activity is present in chlA, chlB, chlE, and chlG mutants.  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