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.
Light-mediated reduction of flavoproteins with flavins as catalysts   总被引:3,自引:0,他引:3  
It has been found that small amounts of free flavins greatly accelerate the photochemical reduction of flavoproteins both to the radical and fully reduced oxidation states. This catalytic effect has been shown to be due to the rapid photochemical reduction of the free flavin to its fully reduced state, followed by its reaction with the flavoprotein to yield flavoprotein radical and by its reaction with flavoprotein radical to yield fully reduced flavoprotein. Evidence is presented that the same route may occur with flavoproteins in the absence of added flavins. In this case the photoreduction is mediated by the small equilibrium concentration of free flavin coenzyme present in a flavorprotein solution. Hence, it is suggested that flavoprotein reduction with EDTA as photosubstrate does not involve an excited state of the holoprotein, nor contact of EDTA with the enzyme, but exchange of electrons between enzyme flavin and free reduced flavin.  相似文献   

2.
Malcolm Dixon 《BBA》1971,226(2):269-284
1. The specificity of flavoproteins towards acceptors has been rather neglected, but an attempt is here made to construct a comparative table of acceptor specificities of those flavoprotein enzymes for which data exist.

2. The acceptor specificity of reduced flavin groups, when combined with apoenzyme proteins, is quite different from that of the same flavin groups in the free state (see Part II). Free flavins react very rapidly with a wide range of acceptors, but the same groups combined as flavoproteins have a severely restricted range of action.

3. There are remarkable differences between different flavoproteins. Nearly every flavoprotein fails altogether to react with at least one, and often several, of the acceptors, giving a specificity pattern which is different in each case. There seems to be no general acceptor for flavoproteins.

4. The effect of combination of a flavin with a particular apoenzyme is to inhibit specifically the reaction of the flavin with particular acceptors with which it would react very rapidly in the absence of the apoenzyme.

5. Each apoenzyme produces its own distinctive pattern of inhibitions. The degree of inhibition is often very high; the table shows over 50 cases of specific inhibitions that are essentially complete. Some of these are very difficult to explain.

6. There is no obvious parallelism between any acceptor and any other in its pattern of reactivity with a series of different flavoproteins.

7. In a few cases combination with apoenzyme specifically accelerates the reaction of the flavin with particular acceptors, so that the flavoprotein is oxidized faster than the free flavin.

8. Possible correlations are discussed between the effects of apoenzymes on the reactivity of flavins with acceptors and a number of special known features of different apoenzymes, but no adequate explanation of the differences in specificity has emerged.

9. In view of the interesting nature of the effects, a plea is made for a more intensive study of the acceptor side of flavoprotein specificity.  相似文献   


3.
The flavoprotein cholesterol oxidase (CO) from Brevibacterium sterolicum is a monomeric flavoenzyme containing one molecule of FAD cofactor covalently linked to His69. The elimination of the covalent link following the His69Ala substitution was demonstrated to result in a significant decrease in activity, in the midpoint redox potential of the flavin, and in stability with respect to the wild-type enzyme, but does not modify the overall structure of the enzyme. We used CO as a model system to dissect the changes due to the elimination of the covalent link between the flavin and the protein (by comparing the wild-type and H69A CO holoproteins) with those due to the elimination of the cofactor (by comparing the holo- and apoprotein forms of H69A CO). The apoprotein of H69A CO lacks the characteristic tertiary structure of the holoprotein and displays larger hydrophobic surfaces; its urea-induced unfolding does not occur by a simple two-state mechanism and is largely nonreversible. Minor alterations in the flavin binding region are evident between the native and the refolded proteins, and are likely responsible for the low refolding yield observed. A model for the equilibrium unfolding of H69A CO that also takes into consideration the effects of cofactor binding and dissociation, and thus may be of general significance in terms of the relationships between cofactor uptake and folding in flavoproteins, is presented.  相似文献   

4.
The fluorescence signal of flavoproteins of rat liver mitochondria was investigated to determine the respective contributions of the various flavoenzymes. About 50% of the overall signal were found to be NAD-linked and caused by alpha-lipoamide dehydrogenase flavin (Em7.4 = -283 mV). Roughly 25% were due to a flavoprotein reducible in a non-NAD-linked reaction. This fluorescent flavoenzyme (Em7.4 = -52 mV) has been tentatively identified as a flavoprotein of the fatty-acid-oxidizing system, most probably the electron transfer flavoprotein. The remaining 25% of the signal are accounted for by flavoenzymes which are reducible by dithionite only. These flavoenzymes were not involved in the flavoprotein fluorescence alterations accompanying changes in electron flow through the respiratory chain. Contributions of other mitochondrial flavoproteins such as succinate dehydrogenase, NADH dehydrogenase, alpha-glycerophosphate dehydrogenase, proline dehydrogenase, and choline oxidase, to the overall flavin fluorescence signal of isolated rat liver mitochondria can be neglected.  相似文献   

5.
A mutagenic analysis of the amino acid residues His-104 and Cys-166, which are involved in the bi-covalent attachment of FAD to berberine bridge enzyme, was performed. Here we present a detailed biochemical characterization of the cysteine link to FAD observed in this recently discovered group of flavoproteins. The C166A mutant protein still has residual activity, but reduced to approximately 6% of the turnover rate observed for wild-type berberine bridge enzyme. A more detailed analysis of single reaction steps by stopped-flow spectrophotometry showed that the reductive half-reaction is greatly influenced by the lack of the 6-S-cysteinyl linkage, resulting in a 370-fold decrease in the rate of flavin reduction. Determination of the redox potentials for both wild type and the C166A mutein revealed that the difference in the redox potential observed can fully account for the change in the kinetic properties. The wild-type protein exhibits a midpoint potential of +132 mV, which is the highest redox potential determined for any flavoenzyme so far. Removal of the cysteine linkage to FAD in the C166A mutein leads to a redox potential of +53 mV, which is in the expected range for flavoproteins with a single covalent attachment of FAD to a His residue via its 8-alpha position. We also show that the biochemical properties of the mutein resemble that of typical flavoprotein oxidases and that deviations from this behavior observed for the wild type are due to the FAD-6-S-cysteinyl bond. In addition, rapid reaction stopped-flow experiments give no indication for a radical mechanism supporting the direct transfer of a hydride from the substrate to the cofactor.  相似文献   

6.
The lantibiotic-synthesizing flavoprotein EpiD catalyzes the oxidative decarboxylation of peptidylcysteines to peptidyl-aminoenethiols. The sequence motif responsible for flavin coenzyme binding and enzyme activity is conserved in different proteins from all kingdoms of life. Dfp proteins of eubacteria and archaebacteria and salt tolerance proteins of yeasts and plants belong to this new family of flavoproteins. The enzymatic function of all these proteins was not known, but our experiments suggested that they catalyze a similar reaction like EpiD and/or may have similar substrates and are homododecameric flavoproteins. We demonstrate that the N-terminal domain of the Escherichia coli Dfp protein catalyzes the decarboxylation of (R)-4'-phospho-N-pantothenoylcysteine to 4'-phosphopantetheine. This reaction is essential for coenzyme A biosynthesis.  相似文献   

7.
Anthranilate hydroxylase was purified from the yeast Trichosporon cutaneum. This enzyme is a simple flavoprotein which apparently does not require any additional cofactor for the conversion of anthranilate to 2,3-dihydroxybenzoate. Anthranilate hydroxylase has Mr of approximately 95,000, with subunit Mr of 50,000; it contains 2 mol of FAD/mol of enzyme. A number of compounds in addition to anthranilate serve as substrates, or effectors, for this enzyme. Oxygen-labeling experiments show that the oxygen atom at the 3-position of the product, 2,3-dihydroxybenzoate, originates from O2, while that at the 2-position is derived from H2O. A mechanism is proposed involving imine formation and hydrolysis during the reaction with the flavin hydroperoxide formed from reduced enzyme flavin and molecular oxygen. This proposal is in accord with the mechanism postulated for other flavoprotein aromatic hydroxylases.  相似文献   

8.
UDP-galactopyranose mutase is a flavoprotein which catalyses the interconversion of UDP-galactopyranose and UDP-galactofuranose. The enzyme is of interest because it provides the activated biosynthetic precursor of galactofuranose, a key cell wall component of many bacterial pathogens. The reaction mechanism of this mutase is intriguing because the anomeric oxygen forms a glycosidic bond, which means that the reaction must proceed by a novel mechanism involving ring breakage and closure. The structure of the enzyme is known, but the mechanism, although speculated on, is not resolved. The overall reaction is electrically neutral but a crypto-redox reaction is suggested by the requirement that the flavin must adopt the reduced form for activity. Herein we report a thermodynamic analysis of the enzyme's flavin cofactor with the objective of defining the system and setting parameters for possible reaction schemes. The analysis shows that the neutral semiquinone (FADH(*)) is stabilized in the presence of substrate and the fully reduced flavin is the anionic FADH(-) rather than the neutral FADH(2). The anionic FADH(-) has the potential to act as a rapid 1-electron donor/acceptor without being slowed by a coupled proton transfer and is therefore an ideal crypto-redox cofactor.  相似文献   

9.
W S Kunz 《FEBS letters》1986,195(1-2):92-96
The different flavoproteins contributing to flavin fluorescence of isolated rat liver mitochondria have distinct excitation and emission spectra. The NAD-linked flavin component was identified as alpha-lipoamide dehydrogenase, while the non-NAD-linked component was found to be electron transfer flavoprotein. The differences in excitation and emission properties of the mitochondrial flavoproteins permit selective recording of their redox state changes in isolated mitochondria.  相似文献   

10.
Yang KY  Swenson RP 《Biochemistry》2007,46(9):2298-2305
Nonresonance Raman spectroscopy has been used to investigate the protein-flavin interactions of the oxidized and anionic semiquinone states of the electron-transfer flavoprotein from the methylotrophic bacteria W3A1 (wETF) in solution. Several unique features of oxidized wETF were revealed from the Raman data. The unusually high frequency of the Raman band for the C(4)=O of the flavin suggests that hydrogen-bonding interactions with the C(4)O are very weak or nonexistent in wETF. In contrast, hydrogen bonding with the C(2)=O is one of the strongest among the flavoproteins investigated thus far. According to the crystal structure, the side-chain hydroxyl group of alphaSer254 serves as a hydrogen bond donor to the N(5) atom in the oxidized flavin cofactor in wETF. The replacement of alphaSer254 by cysteine by site-directed mutagenesis resulted in shifts in N(5)-relevant Raman bands in both the oxidized and anionic semiquinone states of the protein. These results confirm the presence of the hydrogen-bonding interaction at N(5) that is evident in the crystal structure of the oxidized protein and that it persists in the one-electron reduced state. The data suggest that these bands can serve as useful Raman markers for the N(5) interactions in both oxidation states of flavoproteins. The wETF displays unusually low frequencies of flavin ring I (o-xylene ring) relevant bands, which suggests a ring I microenvironment different from most of the other flavoproteins. As indicated by Raman data, the alphaS254C mutation changed the environment of ring I, perhaps as the consequence of changes in the mobility of the FAD domain of wETF. These unusual flavin-protein interactions may be associated with the unique redox properties of wETF.  相似文献   

11.
We present the first report on characterization of the covalent flavinylation site in flavoprotein pyranose 2-oxidase. Pyranose 2-oxidase from the basidiomycete fungus Trametes multicolor, catalyzing C-2/C-3 oxidation of several monosaccharides, shows typical absorption maxima of flavoproteins at 456, 345, and 275 nm. No release of flavin was observed after protein denaturation, indicating covalent attachment of the cofactor. The flavopeptide fragment resulting from tryptic/chymotryptic digestion of the purified enzyme was isolated by anion-exchange and reversed-phase high-performance liquid chromatography. The flavin type, attachment site, and mode of its linkage were determined by mass spectrometry and nuclear magnetic resonance (NMR) spectroscopy of the intact flavopeptide, without its prior enzymatic degradation to the central aminoacyl moiety. Mass spectrometry identified the attached flavin as flavin adenine dinucleotide (FAD). Post-source decay analysis revealed that the flavin is covalently bound to histidine residue in the peptide STHW, consistent with the results of N-terminal amino acid sequencing by Edman degradation. The type of the aminoacyl flavin covalent link was determined by NMR spectroscopy, resulting in the structure 8alpha-(N(3)-histidyl)-FAD.  相似文献   

12.
To understand flavoprotein mechanisms and reactivity, biochemical and biophysical methods are usually employed, and differences between wild-type and mutated proteins with altered primary structures are placed under specific consideration. Alternatively, the cofactor can be modified, and modified flavoproteins can be studied accordingly. Here we present an efficient and general method for modifying the cofactor of flavoproteins in vivo. The modified cofactor is incorporated into apoprotein during protein biosynthesis in a riboflavin-auxotrophic Escherichia coli strain, which expresses a bacterial riboflavin transporter to import flavins from the medium. This system was used to introduce roseoflavin into the riboflavin-binding protein dodecin and into microbial blue-light photoreceptors of the BLUF (blue-light sensors using FAD) and LOV (light oxygen voltage) families. The modified photoreceptors showed absorption and fluorescence different from those of proteins carrying their natural cofactor or chromophores in solution, but did not show any photochemical reaction as implied by former physiological studies.  相似文献   

13.
A number of enzymes require flavin for their catalytic activity, although the reaction catalyzed involves no redox reaction. The best studied of these enigmatic nonredox flavoproteins are the acetohydroxy acid synthases (AHAS), which catalyze early steps in the synthesis of branched-chain amino acids in bacteria, yeasts, and plants. Previously, work from our laboratory showed strong amino acid sequence homology between these enzymes and Escherichia coli pyruvate oxidase, a classical flavoprotein dehydrogenase that catalyzes the decarboxylation of pyruvate to acetate. We have now shown this homology (i) to also be present in the DNA sequences and (ii) to represent functional homology in that pyruvate oxidase has AHAS activity and a protein consisting of the amino-terminal half of pyruvate oxidase and the carboxy-terminal half of E. coli AHAS I allows native E. coli AHAS I to function without added flavin. The hybrid protein contains tightly bound flavin, which is essential for the flavin substitution activity. These data, together with the sequence homologies and identical cofactors and substrates, led us to propose that the AHAS enzymes are descended from pyruvate oxidase (or a similar protein) and, thus, that the flavin requirement of the AHAS enzymes is a vestigial remnant, which may have been conserved to play a structural rather than a chemical function.  相似文献   

14.
Representative examples of the various classes of flavoproteins have been converted to their apoprotein forms and the native flavin replaced by 8-mercapto-FMN or 8-mercapto-FAD. The spectral and catalytic properties of the modified enzymes are characteristically different from one group to another; the results suggest that flavin interactions at positions N(1) or N(5) of the flavin chromophore have profound influences on the properties of the flavoprotein. 1. The 8-thiolate anion form of 8-mercaptoflavin has an absorption maximum in the region 520 to 550 nm epsilon approximately 30 mM-1 cm-1). This form is retained on binding to flavoproteins whose physiological reactions involve obligatory one-electron transfers (e.g. flavodoxin, NADPH-cytochrome P-450 reductase). In the native form these enzymes stabilize the blue neutral radical of the flavin. A radical form of 8-mercaptoflavin is also stabilized by these proteins. 2. The p-quinoid form of 8-mercaptoflavin has an absorption maximum in the range 560 to 600 nm (epsilon approximately 30 mM-1 cm-1). This form is stabilized on binding to flavoproteins of the dehydrogenase-oxidase class (e.g. glucose oxidase, D-amino acid oxidase, lactate oxidase, Old Yellow Enzyme). These same enzymes in their native flavin form stabilize the red semiquinone, and have a pronounced reactivity with sulfite to form flavin N(5)-sulfite adducts. These properties of the native enzyme, including the ability to react with nitroalkane carbanions, are not exhibited by the 8-mercaptoflavoproteins. 3. A group of flavoenzymes fails to conform strictly to the above classification, exhibiting some properties of both classes. These include the examples of flavoprotein hydroxylases and transhydrogenases studied. 4. The riboflavin-binding protein of hen egg whites binds 8-mercaptoriboflavin preferentially in the unionized state, resulting in a shift in pK from 3.8 with free 8-mercaptoriboflavin to greater than or equal to 9.0 with the protein-bound form.  相似文献   

15.
Epidermin from Staphylococcus epidermidis Tü3298 is an antimicrobial peptide of the lantibiotic family that contains, amongst other unusual amino acids, S:-[(Z:)- 2-aminovinyl]-D-cysteine. This residue is introduced by post-translational modification of the ribosomally synthesized precursor EpiA. Modification starts with the oxidative decarboxylation of its C-terminal cysteine by the flavoprotein EpiD generating a reactive (Z:)-enethiol intermediate. We have determined the crystal structures of EpiD and EpiD H67N in complex with the substrate pentapeptide DSYTC at 2.5 A resolution. Rossmann-type monomers build up a dodecamer of 23 point symmetry with trimers disposed at the vertices of a tetrahedron. Oligomer formation is essential for binding of flavin mononucleotide and substrate, which is buried by an otherwise disordered substrate recognition clamp. A pocket for the tyrosine residue of the substrate peptide is formed by an induced fit mechanism. The substrate contacts flavin mononucleotide only via Cys-Sgamma, suggesting its oxidation as the initial step. A thioaldehyde intermediate could undergo spontaneous decarboxylation. The unusual substrate recognition mode and the type of chemical reaction performed provide insight into a novel family of flavoproteins.  相似文献   

16.
Riboflavin (vitamin B(2)) serves as the precursor for FMN and FAD in almost all organisms that utilize the redox-active isoalloxazine ring system as a coenzyme in enzymatic reactions. The role of flavin, however, is not limited to redox processes, as ~ 10% of flavin-dependent enzymes catalyze nonredox reactions. Moreover, the flavin cofactor is also widely used as a signaling and sensing molecule in biological processes such as phototropism and nitrogen fixation. Here, we present a study of 374 flavin-dependent proteins analyzed with regard to their function, structure and distribution among 22 archaeal, eubacterial, protozoan and eukaryotic genomes. More than 90% of flavin-dependent enzymes are oxidoreductases, and the remaining enzymes are classified as transferases (4.3%), lyases (2.9%), isomerases (1.4%) and ligases (0.4%). The majority of enzymes utilize FAD (75%) rather than FMN (25%), and bind the cofactor noncovalently (90%). High-resolution structures are available for about half of the flavoproteins. FAD-containing proteins predominantly bind the cofactor in a Rossmann fold (~ 50%), whereas FMN-containing proteins preferably adopt a (βα)(8)-(TIM)-barrel-like or flavodoxin-like fold. The number of genes encoding flavin-dependent proteins varies greatly in the genomes analyzed, and covers a range from ~ 0.1% to 3.5% of the predicted genes. It appears that some species depend heavily on flavin-dependent oxidoreductases for degradation or biosynthesis, whereas others have minimized their flavoprotein arsenal. An understanding of 'flavin-intensive' lifestyles, such as in the human pathogen Mycobacterium tuberculosis, may result in valuable new intervention strategies that target either riboflavin biosynthesis or uptake.  相似文献   

17.
Phe(1395) stacks parallel to the FAD isoalloxazine ring in neuronal nitric-oxide synthase (nNOS) and is representative of conserved aromatic amino acids found in structurally related flavoproteins. This laboratory previously showed that Phe(1395) was required to obtain the electron transfer properties and calmodulin (CaM) response normally observed in wild-type nNOS. Here we characterized the F1395S mutant of the nNOS flavoprotein domain (nNOSr) regarding its physical properties, NADP(+) binding characteristics, flavin reduction kinetics, steady-state and pre-steady-state cytochrome c reduction kinetics, and ability to shield its FMN cofactor in response to CaM or NADP(H) binding. F1395S nNOSr bound NADP(+) with 65% more of the nicotinamide ring in a productive conformation with FAD for hydride transfer and had an 8-fold slower rate of NADP(+) dissociation. CaM stimulated the rates of NADPH-dependent flavin reduction in wild-type nNOSr but not in the F1395S mutant, which had flavin reduction kinetics similar to those of CaM-free wild-type nNOSr. CaM-free F1395S nNOSr lacked repression of cytochrome c reductase activity that is typically observed in nNOSr. The combined results from pre-steady-state and EPR experiments revealed that this was associated with a lesser degree of FMN shielding in the NADP(+)-bound state as compared with wild type. We conclude that Phe(1395) regulates nNOSr catalysis in two ways. It facilitates NADP(+) release to prevent this step from being rate-limiting, and it enables NADP(H) to properly regulate a conformational equilibrium involving the FMN subdomain that controls reactivity of the FMN cofactor in electron transfer.  相似文献   

18.
Many flavoproteins catalyze the oxidation of primary and secondary amines, with the transfer of a hydride equivalent from a carbon-nitrogen bond to the flavin cofactor. Most of these amine oxidases can be classified into two structural families, the D-amino acid oxidase/sarcosine oxidase family and the monoamine oxidase family. This review discusses the present understanding of the mechanisms of amine and amino acid oxidation by flavoproteins, focusing on these two structural families.  相似文献   

19.
The VAO flavoprotein family is a rapidly growing family of oxidoreductases that favor the covalent binding of the FAD cofactor. In this review we report on the catalytic properties of some newly discovered VAO family members and their mode of flavin binding. Covalent binding of the flavin is a self-catalytic post-translational modification primarily taking place in oxidases. Covalent flavinylation increases the redox potential of the cofactor and thus its oxidation power. Recent findings have revealed that some members of the VAO family anchor the flavin via a dual covalent linkage (6-S-cysteinyl-8α-N1-histidyl FAD). Some VAO-type aldonolactone oxidoreductases favor the non-covalent binding of the flavin cofactor. These enzymes act as dehydrogenases, using cytochrome c as electron acceptor.  相似文献   

20.
p-Hydroxyphenylacetate (HPA) hydroxylase (HPAH) was purified from Acinetobacter baumannii and shown to be a two-protein component enzyme. The small component (C1) is the reductase enzyme with a subunit molecular mass of 32 kDa. C1 alone catalyses HPA-stimulated NADH oxidation without hydroxylation of HPA. C1 is a flavoprotein with FMN as a native cofactor but can also bind to FAD. The large component (C2) is the hydroxylase component that hydroxylates HPA in the presence of C1. C2 is a tetrameric enzyme with a subunit molecular mass of 50 kDa and apparently contains no redox centre. FMN, FAD, or riboflavin could be used as coenzymes for hydroxylase activity with FMN showing the highest activity. Our data demonstrated that C2 alone was capable of utilizing reduced FMN to form the product 3,4-dihydroxyphenylacetate. Mixing reduced flavin with C2 also resulted in the formation of a flavin intermediate that resembled a C(4a)-substituted flavin species indicating that the reaction mechanism of the enzyme proceeded via C(4a)-substituted flavin intermediates. Based on the available evidence, we conclude that the reaction mechanism of HPAH from A. baumannii is similar to that of bacterial luciferase. The enzyme uses a luciferase-like mechanism and reduced flavin (FMNH2, FADH2, or reduced riboflavin) to catalyse the hydroxylation of aromatic compounds, which are usually catalysed by FAD-associated aromatic hydroxylases.  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