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.
Monodehydroascorbate reductase (EC 1.6.5.4) was purified from cucumber fruit to a homogeneous state as judged by polyacrylamide gel electrophoresis. The cucumber monodehydroascorbate reductase was a monomer with a molecular weight of 47,000. It contained 1 mol of FAD/mol of enzyme which was reduced by NAD(P)H and reoxidized by monodehydroascorbate. The enzyme had an exposed thiol group whose blockage with thiol reagents inhibited the electron transfer from NAD(P)H to the enzyme FAD. Both NADH and NADPH served as electron donors with Km values of 4.6 and 23 microM, respectively, and Vmax of 200 mol of NADH and 150 mol of NADPH oxidized mol of enzyme-1 s-1. The Km for monodehydroascorbate was 1.4 microM. The amino acid composition of the enzyme is presented. In addition to monodehydroascorbate, the enzyme catalyzed the reduction of ferricyanide and 2,6-dichloroindophenol but showed little reactivity with calf liver cytochrome b5 and horse heart cytochrome c. The kinetic data suggested a ping-pong mechanism for the monodehydroascorbate reductase-catalyzed reaction. Cucumber monodehydroascorbate reductase occurs in soluble form and can be distinguished from NADPH dehydrogenase, NADH dehydrogenase, DT diaphorase, microsome-bound NADH-cytochrome b5 reductase, and NADPH-cytochrome c reductase by its molecular weight, amino acid composition, and specificity of electron acceptors and donors.  相似文献   

2.
Jeffers CE  Tu SC 《Biochemistry》2001,40(6):1749-1754
It is believed that the reduced FMN substrate required by luciferase from luminous bacteria is provided in vivo by NAD(P)H-FMN oxidoreductases (flavin reductases). Our earlier kinetic study indicates a direct flavin cofactor transfer from Vibrio harveyi NADPH-preferring flavin reductase P (FRP(H)) to the luciferase (L(H)) from the same bacterium in the in vitro coupled luminescence reaction. Kinetic studies were carried out in this work to characterize coupled luminescence reactions using FRP(H) and the Vibrio fischeri NAD(P)H-utilizing flavin reductase G (FRG(F)) in combination with L(H) or luciferase from V. fischeri (L(F)). Comparisons of K(m) values of reductases for flavin and pyridine nucleotide substrates in single-enzyme and luciferase-coupled assays indicate a direct transfer of reduced flavin, in contrast to free diffusion, from reductase to luciferase by all enzyme couples tested. Kinetic mechanisms were determined for the FRG(F)-L(F) and FRP(H)-L(F) coupled reactions. For these two and the FRG(F)-L(H) coupled reactions, patterns of FMN inhibition and effects of replacement of the FMN cofactor of FRP(H) and FRG(F) by 2-thioFMN were also characterized. Similar to the FRP(H)-L(H) couple, direct cofactor transfer was detected for FRG(F)-L(F) and FRP(H)-L(F). In contrast, despite the structural similarities between FRG(F) and FRP(H) and between L(F) and L(H), direct flavin product transfer was observed for the FRG(F)-L(H) couple. The mechanism of reduced flavin transfer appears to be delicately controlled by both flavin reductase and luciferase in the couple rather than unilaterally by either enzyme species.  相似文献   

3.
Biological reduction of carcinogenic chromate has been extensively studied in eukaryotic cells partly because the reduction produces stable chromium(III)-DNA adducts, which are mutagenic. Microbial reduction of chromate has been studied for bioremediation purposes, but little is known about the reduction mechanism. In eukaryotic cells chromate is mainly reduced non-enzymatically by ascorbate, which is usually absent in bacterial cells. We have characterized the reduction of chromate by a flavin reductase (Fre) from Escherichia coli with flavins. The Fre-flavin system rapidly reduced chromate, whereas chemical reduction by NADH and glutathione was very slow. Thus, enzymatic chromate reduction is likely the dominant mechanism in bacterial cells. Furthermore, the end-product was a soluble and stable Cr(III)-NAD(+) complex, instead of Cr(III) precipitate. Since intracellularly generated Cr(III) forms adducts with DNA, protein, glutathione, and ascorbate in eukaryotic cells, we suggest that the produced Cr(III) is primarily complexed to NAD(+), DNA, and other cellular components inside bacteria.  相似文献   

4.
Chorismate synthase (CS) catalyses the conversion of 5- enol pyruvylshikimate 3-phosphate (EPSP) to form chorismate, which is the last common intermediate in the synthesis of the three aromatic amino acids phenylalanine, tyrosine and tryptophan. Despite the overall redox-neutral reaction, catalysis has an absolute requirement for reduced flavin. In the fungus Neurospora crassa , a flavin reductase (FR) activity able to generate reduced flavin mononucleotide in the presence of NADPH is an intrinsic feature of a bifunctional CS. In all bacterial and plant species investigated to date, purified CSs lack an FR activity and are correspondingly 8–10 kDa smaller than the N. crassa CS (on the basis of SDS–PAGE). The cloning of N. crassa CS and subsequent characterization of the purified heterologously expressed enzyme indicates that, surprisingly, the FR probably resides within a region conserved amongst both mono- and bifunctional CSs and is not related to non-homologous sequences which contribute to the larger molecular mass of the N. crassa CS. This information directed this work towards the smaller Saccharomyces cerevisiae CS, the sequence of which was known, although the protein has not been extensively characterized biochemically. Here the characterization of the S. cerevisiae CS is reported in more detail and it is shown that the protein is also bifunctional. With this knowledge, S. cerevisiae could be used as a genetic system for studying the physiological consequences of bifunctionality. The phylogenetic relationship amongst known CSs is discussed.  相似文献   

5.
6.
Cytochrome b(5) reductase (cb5r) catalyzes the transfer of reducing equivalents from NADH to cytochrome b(5). Utilizing an efficient heterologous expression system that produces a histidine-tagged form of the hydrophilic, diaphorase domain of the enzyme, site-directed mutagenesis has been used to generate cb5r mutants with substitutions at position 91 in the primary sequence. Arginine 91 is an important residue in binding the FAD prosthetic group and part of a conserved "RxY(T)(S)xx(S)(N)" sequence motif that is omnipresent in the "ferredoxin:NADP(+) reductase" family of flavoproteins. Arginine 91 was replaced with K, L, A, P, D, Q, and H residues, respectively, and all the mutant proteins purified to homogeneity. Individual mutants were expressed with variable efficiency and all exhibited molecular masses of approximately 32 kDa. With the exception of R91H, all the mutants retained visible absorption spectra typical of a flavoprotein, the former being produced as an apoprotein. Visible absorption spectra of R91A, L, and P were red shifted with maxima at 458 nm, while CD spectra indicated an altered FAD environment for all the mutants except R91K. Fluorescence spectra showed a reduced degree of intrinsic flavin fluorescence quenching for the R91K, A, and P, mutants, while thermal stability studies suggested all the mutants, except R91K, were somewhat less stable than the wild-type domain. Initial-rate kinetic measurements demonstrated that the mutants exhibited decreased NADH:ferricyanide reductase activity with the R91P mutant retaining the lowest activity, corresponding to a k(cat) of 283 s(-1) and a K(NADH)(m) of 105 microM, when compared to the wild-type domain (k(cat) = 800 s(-1) K(NADH)(m) = 6 microM). These results demonstrate that R91 is not essential for FAD binding in cb5r; however, mutation of R91 perturbs the flavin environment and alters both diaphorase substrate recognition and utilization.  相似文献   

7.
8.
Time-resolved flavin fluorescence anisotropy studies on glutathione reductase (GR) have revealed a remarkable new phenomenon: wild-type GR displays a rapid process of fluorescence depolarization, that is absent in mutant enzymes lacking a nearby tyrosine residue that blocks the NADPH-binding cleft. Fluorescence lifetime data, however, have shown a more rigid active-site structure for wild-type GR than for the tyrosine mutants. These results suggest that the rapid depolarization in wild-type GR originates from an interaction with the flavin-shielding tyrosine, and not from restricted reorientational motion of the flavin. A novel mechanism of fluorescence depolarization is proposed that involves a transient charge-transfer complex between the tyrosine and the light-excited flavin, with a concomitant change in the direction of the emission dipole moment of the flavin. This interaction is likely to result from side-chain relaxation of the tyrosine in the minor fraction of enzyme molecules in which this residue is in an unsuitable position for immediate fluorescence quenching at the moment of excitation. Support for this mechanism is provided by binding studies with NADP+ and 2'P-5'ADP-ribose that can intercalate between the flavin and tyrosine and/or block the latter. Fluorescence depolarization analyses as a function of temperature and viscosity confirm the dynamic nature of the process. A comparison with fluorescence depolarization effects in a related flavoenzyme indicates that this mechanism of flavin fluorescence depolarization is more generally applicable.  相似文献   

9.
N Valkova  R Szittner  E A Meighen 《Biochemistry》1999,38(42):13820-13828
Bacterial luciferases (LuxAB) can be readily classed as slow or fast decay luciferases based on their rates of luminescence decay in a single turnover assay. Luciferases from Vibrio harveyi and Xenorhabdus (Photorhabdus) luminescens have slow decay rates, and those from the Photobacterium genus, such as P. (Vibrio) fischeri, P. phosphoreum, and P. leiognathi, have rapid decay rates. By generation of an X. luminescens-based chimeric luciferase with a 67 amino acid substitution from P. phosphoreum LuxA in the central region of the LuxA subunit, the "slow" X. luminescens luciferase was converted into a chimeric luciferase, LuxA(1)B, with a significantly more rapid decay rate. Two other chimeras with P. phosphoreum sequences substituted closer to the carboxyl terminal of LuxA, LuxA(2)B and LuxA(3)B, retained the characteristic slow decay rates of X. luminescens luciferase but had weaker interactions with both reduced and oxidized flavins, implicating the carboxyl-terminal regions in flavin binding. The dependence of the luminescence decay on concentration and type of fatty aldehyde indicated that the decay rate of "fast" luciferases arose due to a high dissociation constant (K(a)) for aldehyde (A) coupled with the rapid decay of the resultant aldehyde-free complex via a dark pathway. The decay rate of luminescence (k(T)) was related to the decanal concentration by the equation: k(T) = (k(L)A + k(D)K(a))/(K(a) + A), showing that the rate constant for luminescence decay is equal to the decay rate via the dark- (k(D)) and light-emitting (k(L)) pathways at low and high aldehyde concentrations, respectively. These results strongly implicate the central region in LuxA(1)B as critical in differentiating between "slow" and "fast" luciferases and show that this distinction is primarily due to differences in aldehyde affinity and in the decomposition of the luciferase-flavin-oxygen intermediate.  相似文献   

10.
p-Hydroxyphenylacetate hydroxylase (HPAH) from Acinetobacter baumannii catalyzes the hydroxylation of p-hydroxyphenylacetate (HPA) to form 3,4-dihydroxyphenylacetate (DHPA). HPAH is composed of two proteins: a flavin mononucleotide (FMN) reductase (C1) and an oxygenase (C2). C1 catalyzes the reduction of FMN by NADH to generate reduced FMN (FMNH-) for use by C2 in the hydroxylation reaction. C1 is unique among the flavin reductases in that the substrate HPA stimulates the rates of both the reduction of FMN and release of FMNH- from the enzyme. This study quantitatively shows the kinetics of how the C1-bound FMN can be reduced and released to be used efficiently as the substrate for the C2 reaction; additional FMN is not necessary. Reactions in which O2 is rapidly mixed with solutions containing C1-FMNH- and C2 are very similar to those in which solutions containing O2 are mixed with one containing the C2-FMNH- complex. This suggests that in a mixture of the two proteins FMNH- binds more tightly to C2 and has already been completely transferred to C2 before it reacts with oxygen. Rate constants for the transfer of FMNH- from C1 to C2 were found to be 0.35 and >or=74 s-1 in the absence and presence of HPA, respectively. The reduction of cytochrome c by FMNH- was also used to measure the dissociation rate of FMNH- from C1. In the absence of HPA, FMNH- dissociates from C1 at 0.35 s-1, while with HPA present it dissociates at 80 s-1; these are the same rates as those for the transfer from C1 to C2. Therefore, the dissociation of FMNH- from C1 is rate-limiting in the intermolecular transfer of FMNH- from C1 to C2, and this process is regulated by the presence of HPA. This regulation avoids the production of H2O2 in the absence of HPA. Our findings indicate that no protein-protein interactions between C1 and C2 are necessary for efficient transfer of FMNH- between the proteins; transfer can occur by a rapid-diffusion process, with the rate-limiting step being the release of FMNH- from C1.  相似文献   

11.
12.
13.
The enzyme responsible for iodide salvage in the thyroid, iodotyrosine deiodinase, was solubilized from porcine thyroid microsomes by limited proteolysis with trypsin. The resulting protein retained deiodinase activity and was purified using anion exchange, dye, and hydrophobic chromatography successively. Peptide sequencing of the final isolate identified the gene responsible for the deiodinase. The amino acid sequence of the porcine enzyme is highly homologous to corresponding genes in a variety of mammals including humans, and the mouse gene was expressed in human embryonic kidney 293 cells to confirm its identity. The amino acid sequence of the deiodinase suggests the presence of three domains. The N-terminal domain provides a membrane anchor. The intermediate domain contains the highest sequence variability and lacks homology to structural motifs available in the common databases. The C-terminal domain is highly conserved and resembles bacterial enzymes of the NADH oxidase/flavin reductase superfamily. A three-dimensional model of the deiodinase based on the coordinates of the minor nitroreductase of Escherichia coli indicates that a Cys common to all of the mammal sequences is located adjacent to bound FMN. However, the deiodinase is not structurally related to other known flavoproteins containing redox-active cysteines or the iodothyronine deiodinases containing an active site selenocysteine.  相似文献   

14.
Russell TR  Demeler B  Tu SC 《Biochemistry》2004,43(6):1580-1590
The homodimeric NADH:flavin oxidoreductase from Aminobacter aminovorans is an NADH-specific flavin reductase herein designated FRD(Aa). FRD(Aa) was characterized with respect to purification yields, thermal stability, isoelectric point, molar absorption coefficient, and effects of phosphate buffer strength and pH on activity. Evidence from this work favors the classification of FRD(Aa) as a flavin cofactor-utilizing class I flavin reductase. The isolated native FRD(Aa) contained about 0.5 bound riboflavin-5'-phosphate (FMN) per enzyme monomer, but one bound flavin cofactor per monomer was obtainable in the presence of excess FMN or riboflavin. In addition, FRD(Aa) holoenzyme also utilized FMN, riboflavin, or FAD as a substrate. Steady-state kinetic results of substrate titrations, dead-end inhibition by AMP and lumichrome, and product inhibition by NAD(+) indicated an ordered sequential mechanism with NADH as the first binding substrate and reduced FMN as the first leaving product. This is contrary to the ping-pong mechanism shown by other class I flavin reductases. The FMN bound to the native FRD(Aa) can be fully reduced by NADH and subsequently reoxidized by oxygen. No NADH binding was detected using 90 microM FRD(Aa) apoenzyme and 300 microM NADH. All results favor the interpretation that the bound FMN was a cofactor rather than a substrate. It is highly unusual that a flavin reductase using a sequential mechanism would require a flavin cofactor to facilitate redox exchange between NADH and a flavin substrate. FRD(Aa) exhibited a monomer-dimer equilibrium with a K(d) of 2.7 microM. Similarities and differences between FRD(Aa) and certain flavin reductases are discussed.  相似文献   

15.
Free reduced flavins are involved in a variety of biological functions. They are generated from NAD(P)H by flavin reductase via co-factor flavin bound to the enzyme. Although recent findings on the structure and function of flavin reductase provide new information about co-factor FAD and substrate NAD, there have been no reports on the substrate flavin binding site. Here we report the structure of TTHA0420 from Thermus thermophilus HB8, which belongs to flavin reductase, and describe the dual binding mode of the substrate and co-factor flavins. We also report that TTHA0420 has not only the flavin reductase motif GDH but also a specific motif YGG in C terminus as well as Phe-41 and Arg-11, which are conserved in its subclass. From the structure, these motifs are important for the substrate flavin binding. On the contrary, the C terminus is stacked on the NADH binding site, apparently to block NADH binding to the active site. To identify the function of the C-terminal region, we designed and expressed a mutant TTHA0420 enzyme in which the C-terminal five residues were deleted (TTHA0420-ΔC5). Notably, the activity of TTHA0420-ΔC5 was about 10 times higher than that of the wild-type enzyme at 20-40 °C. Our findings suggest that the C-terminal region of TTHA0420 may regulate the alternative binding of NADH and substrate flavin to the enzyme.  相似文献   

16.
The flavoenzyme thioredoxin reductase from Escherichia coli contains an oxidation-reduction active disulfide made up of Cys135 and Cys138. Mutations changing each Cys residue to a Ser residue have been effected (Prongay, A. J., engelke, D. R., and Williams, C. H., Jr. (1989) J. Biol. Chem. 264, 2656-2664). The FAD prosthetic group of each altered thioredoxin reductase has been replaced with 1-deaza-FAD (a flavin analog with carbon substituted for nitrogen at position 1), 4-thio-FAD (a flavin analog with sulfur substituted for oxygen at position 4), and 6-thiocyanato-FAD. 1-Deaza-FAD-TRR(Cys135,Ser138) has absorbance and fluorescence spectral properties similar to the oxidized form of wild type apothioredoxin reductase reconstituted with 1-deaza-FAD. The absorbance spectrum of 1-deaza-FAD-TRR(Ser135,Cys138) is similar to the spectrum of the two-electron reduced form of wild type apothioredoxin reductase reconstituted with 1-deaza-FAD, indicating that it is a mixture of two species (O'Donnell, M. E., and Williams, C. H., Jr. (1984) J. Biol. Chem. 259, 2243-2251). The spectrum of one of these species of 1-deaza-FAD-TRR(Ser135,Cys138) resembles the spectrum of oxidized 1-deaza-FAD bound to wild type apothioredoxin reductase. The other species has an absorbance spectrum with a single peak at 400 nm (epsilon 400 = 11,100 M-1 cm-1) and resembles the spectrum of a thiolate adduct at the C4a position of the 1-deaza-FAD. The equilibrium between these species is pH-dependent, with a maximum of 50% C4a-adduct formation at low pH, and is linked to pK alpha values at 8.2 and 9.3. The absorbance spectrum of 4-thio-FAD-TRR(Cys135,Ser138) resembles the spectrum of the unbound 4-thio-FAD, whereas 4-thio-FAD-TRR(Ser135,Cys138) has a spectrum indicative of a mixture of 4-thio-FAD and FAD, suggesting a reaction between the 4-position of the flavin and Cys138. The binding of 6-thiocyanato-FAD to the apoprotein of the mutated enzymes showed no evidence for a reaction between the thiols and the group at the 6-position of the flavin.  相似文献   

17.
As in the case of the succinate and sarcosine dehydrogenases of liver mitochondria, the flavin prosthetic group of the bacterial sarcosine dehydrogenase can be released from the enzyme by proteolytic digestion with trypsin and chymotrypsin. The flavin, isolated in the dinucleotide form and covalently bound to a peptide fragment, is converted to the mononucleotide and purified by sequential chromatography on Sephadex G-25, DEAE-Sephadex A-25, followed by preparative paper chromatography and high voltage electrophoresis.The absorption maxima of the purified flavin at pH 7 are found at 268, 350, and 447 nm, with 268:447 nm and 350:447 nm ratios of 3.08 and 0.79, respectively. The fluorescence excitation and emission maxima, 450 and 530 nm, respectively, are similar to those of flavin mononucleotide. The fluorescence of the flavin-peptide is maximal at pH 3.0–3.1.Amino acid analysis of the flavin-peptide (riboflavin form) gave the following molar ratios of amino acids to flavin: Lys(1), Asp(2), Thr(1), Ser(1), Glu(1), Gly(1), and Ala(1). Aspartic acid was the N-terminal amino acid. Upon more extensive hydrolysis, histidine was obtained in 71–84% yields. Employing aminopeptidase M, the partial sequence of amino acids in the flavin-peptide was determined to be as follows: Flavin
-Asp-Lys-Ser-Glu-Gly-His-(Asp,Ala,Thr)-Evidence is presented that the isoalloxazine ring is linked covalently via its 8 α-methyl group to N-3 of histidine.  相似文献   

18.
19.
The functions of quinone reductase 2 have eluded researchers for decades even though a genetic polymorphism is associated with various neurological disorders. Employing enzymatic studies using adrenochrome as a substrate, we show that quinone reductase 2 is specific for the reduction of adrenochrome, whereas quinone reductase 1 shows no activity. We also solved the crystal structure of quinone reductase 2 in complexes with dopamine and adrenochrome, two compounds that are structurally related to catecholamine quinones. Detailed structural analyses delineate the mechanism of quinone reductase 2 specificity toward catechol quinones in comparison with quinone reductase 1; a side-chain rotational difference between quinone reductase 1 and quinone reductase 2 of a single residue, phenylalanine 106, determines the specificity of enzymatic activities. These results infer functional differences between two homologous enzymes and indicate that quinone reductase 2 could play important roles in the regulation of catecholamine oxidation processes that may be involved in the etiology of Parkinson disease.  相似文献   

20.
Ketopantoyl lactone reductase is a conjugated polyketone reductase   总被引:1,自引:0,他引:1  
Ketopantoyl lactone reductase (EC 1.1.1.168) of Saccharomyces cerevisiae was found to catalyze the reduction of a variety of natural and unnatural conjugated polyketone compounds and quinones, such as isatin, ninhydrin, camphorquinone and beta-naphthoquinone in the presence of NADPH. 5-Bromoisatin is the best substrate for the enzyme (Km = 3.1 mM; Vmax = 650 mumol/min/mg). The enzyme is inhibited by quercetin, and several polyketones. These results suggest that ketopantoyl lactone reductase is a carbonyl reductase which specifically catalyzes the reduction of conjugated polyketones.  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