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.
The FAD binding site of rabbit liver glutathione reductase has been explored by reconstitution of the apoprotein with several FAD analogs modified in the isoalloxazine ring. The apoglutathione reductase binds the p-quinoid form of 8-mercapto-FAD, suggesting that the protein stabilizes a negative charge in the -N1-C2 = O position of the pyrimidine subnucleus. The main absorption peak in the visible spectrum of the 8-mercapto-FAD-enzyme is at 585 nm; treatment of the reconstituted protein with reducing agents of disulfide groups induces a reversible hypochromic shift of 20 nm of the peak. Thus, in 8-mercapto-FAD-glutathione reductase, the oxidation-reduction state of the active center disulfide can be monitored. The chemical reactivity toward methylmethanethiosulfonate and iodoacetamide of the 8-mercapto-FAD-enzyme shows that the flavin position 8 is freely accessible to solvent. However, position 2 is buried within the protein molecule as judged from the lack of reactivity of the 2-thio-FAD-enzyme with methylmethanethiosulfonate. Hydrogen peroxide reacts slowly with both 2-thio-FAD-enzyme and native glutathione reductase, yielding inactive enzyme with a modified spectrum; the prosthetic group is still protein bound. Differences in the active site of the rabbit liver enzyme compared to the human erythrocyte glutathione reductase are evidenced by use of FAD analogs: the peaks of reconstituted liver enzymes are shifted about 10 nm toward longer wavelengths.  相似文献   

2.
Fluorescence quenching and energy-transfer studies have been carried out to determine the position of FAD and FMN groups of NADPH-cytochrome P450 reductase and of the heme and substrate groups of cytochrome P450 with respect to the lipid/water interphase. Quenching by iodine of the fluorescence of the flavins of the reductase shows a biphasic pattern, due to the different accessibility of FAD and FMN to the solvent with Stern-Volmer constants of 7.9 x 10(-4) and 2.7 x 10(-3) mM-1, respectively. Both prosthetic groups appear to be buried within the three-dimensional structure of the native reductase, FAD more deeply embedded than FMN and with a relative contribution to the total fluorescence of flavins of 84% (FAD) and 16% (FMN). The lack of significant energy transfer (less than 5%) from FAD+FMN to the rhodamine group of the N-labeled phosphatidylethanolamine incorporated in membranes reconstituted with NADPH-cytochrome P450 reductase and phosphatidylcholine points out that both groups are located at a distance greater than 5 nm from the lipid/water interphase. Steady-state fluorescence intensity and anisotropy data obtained with native and FMN-depleted NADPH-cytochrome P450 reductase show that energy transfer between both prosthetic groups occurs in the native reductase with an efficiency of ca. 31%, consistent with a separation between these groups of 2 nm as suggested earlier by Bastiaens, P. I. H., Bonants, P. J. M., Müller, F., & Visser, A. J. W. G. [(1989) Biochemistry 28, 8416-8425] from time-resolved fluorescence anisotropy measurements.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Cytochrome P450BM-3 is a self-sufficient bacterial protein containing three naturally fused domains which bind either heme, FMN, or FAD. Resolution of protein and FMN from the isolated FMN-containing domain of cytochrome P450Betamicro-3 was accomplished using trichloroacetic acid. The apoprotein thus prepared was shown to rebind FMN to regenerate the original holoprotein as indicated by both spectroscopy and activity measurements. To better understand how the protein/flavin interaction might contribute to reactivity, the association process was studied in detail. Fluorescence quenching was used to measure a dissociation constant of the flavin-protein complex of 31 nM, comparable to FMN-containing proteins of similar reactivity and higher than that of flavodoxins. Stopped-flow kinetics were performed, and a multistep binding process was indicated, with an initial k(on) value of 1.72 x 10(5) M(-)(1) s(-)(1). Preparation of the apoprotein allowed substitution of flavin analogues for the native FMN cofactor using 8-chloro-FMN and 8-amino-FMN. Both were found to bind efficiently to the protein with only minor variations in affinity. Reductive titrations established that, as in the native FMN-containing FMN-binding domain, the 8-amino-FMN-substituted domain does not produce a stable one-electron-reduced species during titration with sodium dithionite. The 8-chloro-FMN-substituted domain, however, had sufficiently altered redox properties to form a stable red anionic semiquinone. The 8-chloro-FMN-substituted FMN-binding domain was shown in reconstituted systems to retain most of the cytochrome c reductase activity of the native domain but only a very small amount of palmitic acid hydroxylase activity. The 8-amino-FMN-substituted FMN-binding domain showed no palmitic acid hydroxylase activity and only 30% of the native cytochrome c reductase activity, demonstrating the importance of thermodynamics to the mechanism of this protein.  相似文献   

4.
The apoprotein of the FAD-containing flavoenzyme glycine oxidase from Bacillus subtilis was obtained at pH 8.5 by dialyzing the holoenzyme against 2 M KBr in 0.25 M Tris–HCl and 20% glycerol. The apoprotein of glycine oxidase shows high protein fluorescence, high exposure of hydrophobic surfaces, and low temperature stability as compared to the holoenzyme. The isolated apoprotein species is present in solution as a monomer which rapidly recovers its tertiary structure and converts into the tetrameric holoenzyme following incubation with free FAD. The reconstitution process follows a particular two-stage process; the spectral properties of the reconstituted holoenzyme were virtually indistinguishable from those observed with native glycine oxidase, while the activity was only partially (50%) recovered. The urea-induced unfolding process of glycine oxidase can be considered as a two-step (three-state) process: the presence of intermediate(s) in the unfolding process of the holoenzyme at ≈2 M urea is evident in the changes of the flavin fluorescence intensity and can be also inferred from the different urea sensitivities of the spectral probes used. On the other hand, only a single transition at ≈4.5 M urea concentration is observed for the apoprotein form. The chemical denaturation of glycine oxidase holoenzyme is partially reversible (e.g., no activity is recovered when starting the refolding from 4 M urea-denatured holoprotein). Finally, the introduction by site-directed mutagenesis of residues corresponding to those involved in the covalent link with FAD in the related flavoenzyme monomeric sarcosine oxidase failed to convert glycine oxidase into a covalent flavoprotein. These investigations show that the consequences of FAD binding for the stability and folding process distinguish glycine oxidase from enzymes active on similar compounds.  相似文献   

5.
We measured the redox potentials of frozen inactivated l-amino-acid oxidase (l-amino-acid:oxygen oxidoreductase (deaminating), EC 1.4.3.2) and inhibitor-bound (anthranilic acid) enzyme, and compared these redox properties to those of active l-amino-acid oxidase and benzoate-bound d-amino-acid oxidase (EC 1.4.3.3), respectively. The redox properties of the inactive enzyme are similar to the properties of free flavin; the potential is within 0.015 V of free flavin and no radical stabilization is seen. This corresponds to the loss of most interactions between apoprotein and flavin. In contrast, the anthranilic acid lowers the amount of radical stabilized from 85% to 35%. The potentials are still 0.150 V positive of free flavin, indicating that in the presence of inhibitor, many flavin-protein interactions remain intact. The difference between this behavior and that of d-amino-acid oxidase bound to benzoate, where the amount of radical declined from 95% to 5%, is explained on the basis of the relative tightness of binding of apoprotein to FAD. d-Amino-acid oxidase apoprotein has a relatively low Ka (106) for FAD, and benzoate has a relatively high Ka (105) for the enzyme. Therefore, the binding of benzoate increases the tightness of FAD binding to apo-d-amino-acid oxidase (1011), indicating significant changes in flavin-protein interactions. In contrast, apo-l-amino-acid oxidase binds flavin tightly (the Ka is greater than 107) and the enzyme binds to anthranilate much less tightly, with a Ka of 103. The l-amino-acid oxidase apoprotein binding to FAD is tight initially, and the binding of anthranilate changes it only slightly. Therefore, redox studies indicate that the ability of a flavoprotein to be regulated may be influenced by the strength of the interaction of flavin with the apoprotein, as well as the strength of interaction of the substrate or activator.  相似文献   

6.
Biosynthesis of ferredoxin-NADP+ reductase in higher plants was investigated in relation with the mechanism of formation of the holoenzyme. The putative precursor of the flavoprotein, obtained after cell-free translation on a wheat germ extract primed with poly(A)-rich mRNA, was able to spontaneously bind free FAD, rendering a functional prereductase. The newly synthesized preholoenzyme showed diaphorase and cytochrome c reductase activities, an apparent molecular mass of 45 kDa, and contained FAD as the only flavin cofactor. It gave a positive reaction towards antisera against mature ferredoxin-NADP+ reductase. On the other hand, intracellular distribution of flavin-synthesizing enzymes indicates that FAD formation occurs in the cytoplasm; that is, in the same compartment as the site of reductase synthesis. On the basis of the preceding data a model is presented for the biosynthesis of the enzyme in vivo, involving conjugation of the apoprotein with FAD in the cytoplasm, followed by transport of the preholoreductase across the chloroplast envelope to reach its final destiny in the thylakoid membrane.  相似文献   

7.
The repression-derepression control of Nostoc muscorum nitrate reductase was studied with regard to the Mo-cofactor and apoprotein levels. It was found that the synthesis of Mo-cofactor is constitutive but the apoprotein is subject to the repression-derepression control. In NH4+ medium apoprotein synthesis was repressed and in N2 and NO3? media apoprotein synthesis was derepressed. The apoprotein levels were similar in NO3? and N2 media; however, the nitrate reductase activity was lower in N2 medium due to lower Mo-cofactor activity. The lower Mo-cofactor activity in N2-fixing conditions as compared to that in non-N2-fixing conditions was consistent with the earlier view that the Mo-cofactor of nitrate reductase may be a precursor for FeMo-cofactor of nitrogenase.  相似文献   

8.
cDNA of Capsicum annuum Yolo Wonder (paprika) has been prepared from total cellular RNA, and the complete gene encoding paprika ferredoxin-NADP(+) reductase (pFNR) precursor was sequenced and cloned from this cDNA. Fusion to a T7 promoter allowed expression in Escherichia coli. Both native and recombinant pFNR were purified to homogeneity and crystallized. The crystal structure of pFNR has been solved by Patterson search techniques using the structure of spinach ferredoxin-NADP(+) reductase as search model. The structure was refined at 2.5-A resolution to a crystallographic R-factor of 19.8% (R(free) = 26.5%). The overall structure of pFNR is similar to other members of the ferredoxin-NADP(+) reductase family, the major differences concern a long loop (residues 167-177) that forms part of the FAD binding site and some of the variable loops in surface regions. The different orientation of the FAD binding loop leads to a tighter interaction between pFNR and the adenine moiety of FAD. The physiological redox partners [2Fe-2S]-ferredoxin I and NADP(+) were modeled into the native structure of pFNR. The complexes reveal a protein-protein interaction site that is consistent with existing biochemical data and imply possible orientations for the side chain of tyrosine 362, which has to be displaced by the nicotinamide moiety of NADP(+) upon binding. A reasonable electron transfer pathway could be deduced from the modeled structures of the complexes.  相似文献   

9.
A new method is described for the large-scale reversible dissociation of flavoproteins into apoprotein and prosthetic group using hydrophobic-interaction chromatography. Lipoamide dehydrogenase from Azotobacter vinelandii and butyryl-CoA dehydrogenase from Megasphaera elsdenii are selected to demonstrate the usefulness of the method. In contrast to conventional methods, homogeneous preparations of apoproteins in high yields are obtained. The apoproteins show high reconstitutability. The holoenzymes are bound to phenyl-Sepharose CL-4B at neutral pH in the presence of ammonium sulfate. FAD is subsequently removed at pH 3.5-4.0 by addition of high concentrations of KBr. Large amounts of apoenzymes (200-500 mg), showing negligible residual activity, are eluted at neutral pH in the presence of 50% ethylene glycol. The holoenzyme of lipoamide dehydrogenase can be reconstituted while the apoprotein is still bound to the column or the apoenzyme can be isolated in the free state. In both cases the yield and degree of reconstitution of holoenzyme is more than 90% of starting material. Apo-lipoamide-dehydrogenase exists mainly as a monomer in solution and reassociates to the native dimeric structure in the presence of FAD. The apoenzyme is stable for a long period of time when kept in 50% ethylene glycol at -18 degrees C. Steady-state fluorescence-polarization measurements of protein-bound FAD indicate that reconstituted lipoamide dehydrogenase possesses a high stability which is governed by the low dissociation rate constant of the apoenzyme-FAD complex. The holoenzyme of butyryl-CoA dehydrogenase cannot be reconstituted when the apoenzyme is bound to the column. However, stable apoprotein can be isolated in the free state yielding 50-80% of starting material, depending on the immobilization conditions. The coenzyme A ligand present in native holoenzyme is removed during apoprotein preparation. The apoenzyme is relatively stable when kept in 50% ethylene glycol at -18 degrees C. From kinetic and gel filtration experiments it is concluded that the reconstitution reaction of butyryl-CoA dehydrogenase is governed by both the pH-dependent hydrodynamic properties of apoenzyme and the pH-dependent stability of reconstituted enzyme. At pH 7, the apoenzyme is in equilibrium between dimeric and tetrameric forms and reassociates to a native-like tetrameric structure in the presence of FAD. The stability of reconstituted enzyme is strongly influenced by the presence of CoA ligands as shown by fluorescence-polarization measurements. The degree of reconstitution of butyryl-CoA dehydrogenase is more than 80% of the original specific activity under certain conditions.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

10.
H Masui  M Satoh    T Satoh 《Journal of bacteriology》1994,176(6):1624-1629
Spheroplasts prepared from a molybdenum cofactor-deficient mutant of Rhodobacter sphaeroides f. sp. denitrificans secreted dimethyl sulfoxide (DMSO) reductase which had no molybdenum cofactor and therefore no activity, whereas those from wild-type cells secreted the active reductase. The inactive DMSO reductase proteins were separated by nondenaturing electrophoresis into two forms: form I, with the same mobility as the native enzyme, and form II, with slower mobility. Both forms had the same mobility on denaturing gel. Form I and active DMSO reductase had the same profile on gel filtration chromatography. Form II was eluted a little faster than the native enzyme, suggesting that DMSO reductase form II was not an aggregated form but a compactly folded form very similar to the native enzyme. Form II was digested by trypsin and denatured with urea, whereas form I was unaffected, like native DMSO reductase. These results suggested that form II was a partially unfolded but compactly folded apoprotein of DMSO reductase.  相似文献   

11.
Human glutathione reductase (NADPH + GSSG + H+ in equilibrium with NADP+ + 2 GSH) is a suitable enzyme for correlating spectroscopic properties and chemical reactivities of protein-bound FAD analogues with structural data. FAD, the prosthetic group of the enzyme, was replaced by FAD analogues, which were modified at the positions 8, 1, 2, 4, 5 and 6, respectively, of the isoalloxazine ring. When compared with a value of 100% for native glutathione reductase, the specific activities of most enzyme species ranged from 40% to 17%, in the order of the prosthetic groups 8-mercapto-FAD greater than 8-azido-FAD = 8-F-FAD = 8-C1-FAD greater than 4-thio-FAD = 1-deaza-FAD greater than 2-thio-FAD. The enzymic activities indicate a correct orientation of the bound analogues. The enzyme species containing 5-deaza-FAD and 6-OH-FAD, respectively, had no more glutathione reductase activity than the FAD-free apoenzyme. 5-Deaza-FAD X glutathione reductase was crystallized for X-ray diffraction analysis. Detailed studies were focussed on position 8 of the flavin. 8-Cl-FAD X glutathione reductase and 8-F-FAD X glutathione reductase reacted only poorly with HS- to give 8-mercapto-FAD X glutathione reductase, which suggests that the region around Val61 hinders the halogen anion from leaving the tetrahedral intermediate. Other experiments showed that position 8 is accessible to certain solvent-borne reagents. 8-Mercapto-FAD X glutathione reductase, for instance, reacted readily and stoichiometrically with the thiol reagent methylmethanethiosulfonate. 8-Mercapto-FAD X glutathione reductase does not exhibit a long wavelength charge transfer absorption band upon reduction, as it is the case for the 2-electron-reduced FAD-containing enzyme. This behaviour indicates that the charge transfer interaction between flavin and the thiolate of Cys63 in the native enzyme is not per se essential for catalysis. The absorption spectrum of the blue anionic 8-mercapto-FAD bound to glutathione reductase suggests that the protein concurs to the stabilization of a negative charge in the pyrimidine subnucleus. In light of the protein structure this effect is attributed to the dipole moment of alpha-helix 338-354 which starts out close to the N(1)/C(2)/O(2 alpha) region of the flavin. 1-Deaza-FAD binds as tightly as FAD to the apoenzyme. The resulting holoenzyme was found to be enzymically active but structurally unstable. In this respect 1-deaza-FAD . glutathione reductase mimics the properties of the enzyme species found in inborn glutathione reductase deficiency.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

12.
A riboflavin-binding hexamerin isolated from pupal hemolymph of Hyalophora cecropia has a native Mr of 510,000, subunit Mr of 85,000, and a 5% carbohydrate content. An intrachain cross-link was confirmed in protease limit digests. Ellman titration confirmed the presence of a sulfhydryl group, which is needed for this linkage. Though Cu2+ is known to promote the linkage, heavy metals were not detected in the isolate. Heat denaturation released ligand with the absorbency, fluorescence spectra, and chromatographic behavior of riboflavin. Binding resulted in substantial quenching of the fluorescence of both the isoalloxazine in riboflavin and of aromatic groups in the apoprotein. Kinetic analysis indicated a KD of 2.5 × 10?7 M for riboflavin, 1.3 × 10?7 M for lumiflavin, and greater than 1 × 10?6 M for FMN and FAD. Over four moles of flavin were bound per mole of hexamerin. The amount of riboflavin in pupal hemolymph is sufficient to occupy only 2–3 of these sites. Riboflavin is also associated with lipophorin and vitellogenin, but the molar ratios after protein isolation were low. On a standard laboratory diet, riboflavin is in great excess, but most of it is apparently excreted before the apoprotein first appears in the hemolymph, just before wandering. The concentration of riboflavin-binding hexamerin rises to 15–30 mg/ml in pupae; relative to other hexamerins, very little is stored in the fat body. All of the apoprotein and 75% of riboflavin disappear from the hemolymph during adult development. An amount of flavin at least equal to that stored in pupal hemolymph is transferred to the eggs formed during this period. © 1994 Wiley-Liss, Inc.  相似文献   

13.
Summary

The role of the prosthetic groups (FAD and FMN) of NADPH-cytochrome P450 reductase (P450 reductase)in 3-hydroxyanthranilamide (3-OH An.Amide)-catalyzed, NADPH-dependent superoxide anion (O2-) production via the reductase was examined using the native and FMN-depleted preparations of P450 reductase which was partially purified from rat liver microsomes. NADPH-dependent O2-production by the FMN-depleted preparation was about 10% of that by the native preparation. 3-OH An. Amide-catalyzed, NADPH-dependent O2-production by the FMN-depleted preparation was less than 10% of that by the native preparation. FMN supplementation returned O2-production to near normal. We observed the same results for NADPH oxidation and hydrogen peroxide formation. O2-production, NADPH oxidation, and hydrogen peroxide formation were inhibited by native superoxide dismutase (SOD), but not by boiled, denatured SOD. These results indicate that the prosthetic groups, especially FMN, of P450 reductase play a critical role in 3-OH An.Amide-catalyzed, NADPH-dependent O2-production via the reductase.  相似文献   

14.
A gene has been constructed coding for a unique fusion protein, NADH:cytochrome c reductase, that comprises the soluble heme-containing domain of rat hepatic cytochrome b(5) as the amino-terminal portion of the protein and the soluble flavin-containing domain of rat hepatic cytochrome b(5) reductase as the carboxyl terminus. The gene has been expressed in Escherichia coli resulting in the highly efficient production of a functional hybrid hemoflavoprotein which has been purified to homogeneity by a combination of ammonium sulfate precipitation, affinity chromatography on 5'-ADP agarose, and size-exclusion chromatography. The purified protein exhibited a molecular mass of approximately 46 kDa by polyacrylamide gel electrophoresis and 40,875 Da, for the apoprotein, using mass spectrometry which also confirmed the presence of both heme and FAD prosthetic groups. The fusion protein showed immunological cross-reactivity with both anti-rat cytochrome b(5) and anti-rat cytochrome b(5) reductase antibodies indicating the conservation of antigenic determinants from both native domains. Spectroscopic analysis indicated the fusion protein contained both a b-type cytochrome and flavin chromophors with properties identical to those of the native proteins. Amino-terminal and internal amino acid sequencing confirmed the identity of peptides derived from both the heme- and flavin-binding domains with sequences identical to the deduced amino acid sequence. The isolated fusion protein retained NADH:ferricyanide reductase activity (k(cat) = 8.00 x 10(2) s(-1), K(NADH)(m) = 4 microM, K(FeCN(6))(m) = 11 microM) comparable to that of that of native NADH:cytochrome b(5) reductase and also exhibited both NADH:cytochrome c reductase activity (k(cat) = 2.17 x 10(2) s(-1), K(NADH)(m) = 2 microM, K(FeCN(6))(m) = 11 microM, K(Cyt.c)(m) = 1 microM) and NADH:methemoglobin reductase activity (k(cat) = 4.40 x 10(-1) s(-1), K(NADH)(m) = 3 microM, K(mHb)(m) = 47 microM), the latter two activities indicating efficient electron transfer from FAD to heme and retention of physiological function. This work represents the first successful bacterial expression of a soluble, catalytically competent, rat hepatic cytochrome b(5)-cytochrome b(5) reductase fusion protein that retains the functional properties characteristic of the individual heme and flavin domain.  相似文献   

15.
The turnover of catalase apoprotein and catalase heme was studied in cotyledons of sunflower (Helianthus annuus L.) seedlings by density labeling of apoprotein and radioactive labeling of heme moieties. The heavy isotope (50% 2H2O) and the radioactive isotope ([14C]5-aminolevulinic acid) were applied either during growth in the dark (day 0-2.5) or in the light (day 2.5 and 5). Following isopycnic centrifugation of catalase purified from cotyledons of 5-day-old seedlings, superimposition curve fitting was used to determine the amounts of radioactive heme moieties in native and density-labeled catalase. Data from these determinations indicated that turnover of catalase heme and apoprotein essentially was coordinate. Only small amounts of heme groups were recycled into newly synthesized apoprotein during growth in the light, and no evidence was found for an exchange of heme groups in apoprotein moieties. It followed from these observations that degradation of catalase apoprotein was slightly faster than that of catalase heme. A degradation constant for catalase apoprotein of 0.263 per day was determined from the data on heme recycling and the degradation constant of catalase heme determined previously to be 0.205 per day (R Eising, B Gerhardt [1987] Plant Physiol 84: 225-232).  相似文献   

16.
The flavin prosthetic group (FAD) of p-hydroxybenzoate hydroxylase from Pseudomonas fluorescens was replaced by a stereochemical analog, which is spontaneously formed from natural FAD in alcohol oxidases from methylotrophic yeasts. Reconstitution of p-hydroxybenzoate hydroxylase from apoprotein and modified FAD is a rapid process complete within seconds. Crystals of the enzyme-substrate complex of modified FAD-containing p-hydroxybenzoate hydroxylase diffract to 2.1 A resolution. The crystal structure provides direct evidence for the presence of an arabityl sugar chain in the modified form of FAD. The isoalloxazine ring of the arabinoflavin adenine dinucleotide (a-FAD) is located in a cleft outside the active site as recently observed in several other p-hydroxybenzoate hydroxylase complexes. Like the native enzyme, a-FAD-containing p-hydroxybenzoate hydroxylase preferentially binds the phenolate form of the substrate (pKo = 7.2). The substrate acts as an effector highly stimulating the rate of enzyme reduction by NADPH (kred > 500 s-1). The oxidative part of the catalytic cycle of a-FAD-containing p-hydroxybenzoate hydroxylase differs from native enzyme. Partial uncoupling of hydroxylation results in the formation of about 0.3 mol of 3,4-dihydroxybenzoate and 0.7 mol of hydrogen peroxide per mol NADPH oxidized. It is proposed that flavin motion in p-hydroxybenzoate hydroxylase is important for efficient reduction and that the flavin "out" conformation is associated with the oxidase activity.  相似文献   

17.
Quaternary structure and composition of squash NADH:nitrate reductase   总被引:6,自引:0,他引:6  
NADH:nitrate reductase (EC 1.6.6.1) was isolated from squash cotyledons (Cucurbita maxima L.) by a combination of Blue Sepharose and zinc-chelate affinity chromatographies followed by gel filtration on Bio-Gel A-1.5m. These preparations gave a single protein staining band (Mr = 115,000) on sodium dodecyl sulfate gel electrophoresis, indicating that the enzyme is homogeneous. The native Mr of nitrate reductase was found to be 230,000, with a minor form of Mr = 420,000 also occurring. These results indicate that the native nitrate reductase is a homodimer of Mr = 115,000 subunits. Acidic amino acids predominate over basic amino acids, as shown both by the amino acid composition of the enzyme and an isoelectric point for nitrate reductase of 5.7. The homogeneous nitrate reductase had a UV/visible spectrum typical of a b-type cytochrome. The enzyme was found to contain one each of flavin (as FAD), heme iron, molybdenum, and Mo-pterin/Mr = 115,000 subunit. A model is proposed for squash nitrate reductase in which two Mr = 115,000 subunits are joined to made the native enzyme. Each subunit contains 1 eq of FAD, cytochrome b, and molybdenum/Mo-pterin.  相似文献   

18.
NADPH-cytochrome c reductase has been isolated from a top-fermenting ale yeast, Saccharomyces cerevisiae (Narragansett strain), after ca. a 240-fold purification over the initial extract of an acetone powder, with a final specific activity (at pH 7.6, 30 °C) of ca. 150 μmol cytochrome c reduced min?1mg?1 protein. The preparation appears to be homogeneous by the criteria of: sedimentation velocity; electrophoresis on cellulose acetate in buffers above neutrality; and by polyacrylamide gel electrophoresis. Although the reductase appeared to partially separate into species “A” and “B” on DEAE-cellulose at pH 8.8, the two species have proven to be indistinguishable electrophoretically (above pH 8) and by sedimentation. By sedimentation equilibrium at 20 °C, a molecular weight of ca. 6.8 (± 0.4) × 104 was obtained with use of a V?20 ° = 0.741 calculated from its amino acid composition. After disruption in 4 m guanidinium chloride- 10 mm dithioerythritol- 1 mm EDTA, pH 6.4 at 20 °C, an M?r of 3.4 (± 0.1) × 104 resulted, which points to a subunit structure of two polypeptide chains per mole. Confirmatory evidence of the two-subunit structure with similar, if not identical, polypeptide chains was obtained by polyacrylamide gel electrophoresis in dodecyl-sulfate, after disruption in 4 m urea and 2% sodium dodecyl sulfate, and yielded a subunit molecular weight of ca. 4 × 104. Sulfhydryl group titration with 4,4′-dithiodipyridine under acidic conditions revealed one sulfhydryl group per monomer, which apparently is necessary for the catalytic reduction of cytochrome c. NADPH, as well as FAD, protects this-SH group from reaction with 5,5′-dithiobis (2-nitrobenzoate). The visible absorption spectrum of the oxidized enzyme (as prepared) has absorption maxima at 383 and 455 nm, typical of a flavoprotein. Flavin analysis (after dissociation by thermal denaturation of the “A” protein) conducted fluorometrically, revealed the presence of 2.0 mol of FAD per 70,000 g, in confirmation of the deduced subunit structure. The identity of the FAD dissociated from either “A” or “B” protein was confirmed by recombination with apo-d-amino acid oxidase and by thin-layer chromatography. A kinetic approach was used to estimate the dissociation constant for either FAD or FMN (which also yields a catalytically active enzyme) to the apoprotein reductase at 30 °C and pH 7.6 (0.05 m phosphate) and yielded values of 4.7 × 10?8m for FAD and 4.4 × 10?8m for FMN.  相似文献   

19.
NADH:nitrate reductase (EC 1.6.6.1) from Chlorella vulgaris has been purified 640-fold with an over-all yield of 26% by a combination of protamine sulfate fractionation, ammonium sulfate fractionation, gel chromatography, density gradient centrifugation, and DEAE-chromatography. The purified enzyme is stable for more than 2 months when stored at minus 20 degrees in phosphate buffer (pH 6.9) containing 40% (v/v) glycerol. After the initial steps of the purification, a constant ratio of NADH:nitrate reductase activity to NADH:cytochrome c reductase and reduced methyl viologen:nitrate reductase activities was observed. One band of protein was detected after polyacrylamide gel electrophoresis of the purified enzyme. This band also gave a positive stain for heme, NADH dehydrogenase, and reduced methyl viologen:nitrate reductase. One band, corresponding to a molecular weight of 100, 000, was detected after sodium dodecyl sulfate polyacrylamide gel electrophoresis. The enzyme contains FAD, heme, and molybdenum in a 1:1:0.8 ratio. One "cyanide binding site" per molybdenum was found. No non-heme-iron or labile sulfide was detected. From a dry weight determination of the purified enzyme, a minimal molecular weight of 152, 000 per molecule of heme or FAD was calculated. An s20, w of 9.7 S for nitrate reductase was found by the use of sucrose density gradient centrifugation and a Stokes radius of 89 A was estimated by gel filtration techniques. From these values, and the assumption that the partial specific volume is 0.725 cc/g, a molecular weight of 356, 000 was estimated for the native enzyme. These data suggest that the native enzyme contains a minimum of 2 molecules each of FAD, heme, and molybdenum and is composed of at least three subunits.  相似文献   

20.
Treatment by p-hydroxymercuribenzoate of nitrate reductase from spinach leaves causes the disappearance of NADH-diaphorase activity and the appearance of an FAD-requirement for the inactivation by NAD(P)H of FNH2-nitrate reductase. The diaphorase activity of the treated preparation is not affected by incubation with FAD or the addition of this nucleotide to the assay mixture. Conversely, filtration of the native preparation through a column of Sepharose 6B produces the appearance of an FAD-stimulation of the diaphorase activity, but no effect of FAD on the NADH-inactivation was observed. These differences between the FAD-requirement of NADH-diaphorase activity and NADH-inactivation agree with the postulated independence of the two processes.  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