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.
Electron transfer reactions involving protein-protein interactions require the formation of a transient complex which brings together the two redox centres exchanging electrons. This is the case for the flavoprotein ferredoxin:NADP+ reductase (FNR) from the cyanobacterium Anabaena, an enzyme which interacts with ferredoxin in the photosynthetic pathway to receive the electrons required for NADP+ reduction. The reductase shows a concave cavity in its structure into which small proteins such as ferredoxin can fit. Flavodoxin, an FMN-containing protein that is synthesised in cyanobacteria under iron-deficient conditions, plays the same role as ferredoxin in its interaction with FNR in spite of its different structure, size and redox cofactor. There are a number of negatively charged amino acid residues on the surface of ferredoxin and flavodoxin that play a role in the electron transfer reaction with the reductase. Thus far, in only one case has charge replacement of one of the acidic residues produced an increase in the rate of electron transfer, whereas in several other cases a decrease in the rate is observed. In the most dramatic example, replacement of Glu at position 94 of Anabaena ferredoxin results in virtually the complete loss of ability to transfer electrons. Charge-reversal of positively charged amino acid residues in the reductase also produces strong effects on the rate of electron transfer. Several degrees of impairment have been observed, the most significant involving a positively charged Lys at position 75 which appears to be essential for the stability of the complex between the reductase and ferredoxin. The results presented in this paper provide a clear demonstration of the importance of electrostatic interactions on the stability of the transient complex formed during electron transfer by the proteins presently under study.  相似文献   

2.
Transhydrogenase is a proton pump. It has separate binding sites for NAD+/NADH (on domain I of the protein) and for NADP+/NADPH (on domain III). Purified, detergent-dispersed transhydrogenase from Escherichia coli catalyses the reduction of the NAD+ analogue, acetylpyridine adenine dinucleotide (AcPdAD+), by NADH at a slow rate in the absence of added NADP+ or NADPH. Although it is slow, this reaction is surprising, since transhydrogenase is generally thought to catalyse hydride transfer between NAD(H)--or its analogues and NADP(H)--or its analogues, by a ternary complex mechanism. It is shown that hydride transfer occurs between the 4A position on the nicotinamide ring of NADH and the 4A position of AcPdAD+. On the basis of the known stereospecificity of the enzyme, this eliminates the possibilities of transhydrogenation(a) from NADH in domain I to AcPdAD+ wrongly located in domain III; and (b) from NADH wrongly located in domain III to AcPdAD+ in domain I. In the presence of low concentrations of added NADP+ or NADPH, detergent-dispersed E. coli transhydrogenase catalyses the very rapid reduction of AcPdAD+ by NADH. This reaction is cyclic; it takes place via the alternate oxidation of NADPH by AcPdAD+ and the reduction of NADP+ by NADH, while the NADPH and NADP+ remain tightly bound to the enzyme. In the present work, it is shown that the rate of the cyclic reaction and the rate of reduction of AcPdAD+ by NADH in the absence of added NADP+/NADPH, have similar dependences on pH and on MgSO4 concentration and that they have a similar kinetic character. It is therefore suggested that the reduction of AcPdAD+ by NADH is actually a cyclic reaction operating, either with tightly bound NADP+/NADPH on a small fraction (< 5%) of the enzyme, or with NAD+/NADH (or AcPdAD+/AcPdADH) unnaturally occluded within the domain III site. Transhydrogenase associated with membrane vesicles (chromatophores) of Rhodospirillum rubrum also catalyses the reduction of AcPdAD+ by NADH in the absence of added NADP+/NADPH. When the chromatophores were stripped of transhydrogenase domain I, that reaction was lost in parallel with 'normal reverse' transhydrogenation (e.g., the reduction of AcPdAD+ by NADPH). The two reactions were fully recovered upon reconstitution with recombinant domain I protein. However, after repeated washing of the domain I-depleted chromatophores, reverse transhydrogenation activity (when assayed in the presence of domain I) was retained, whereas the reduction of AcPdAD+ by NADH declined in activity. Addition of low concentrations of NADP+ or NADPH always supported the same high rate of the NADH-->AcPdAD+ reaction independently of how often the membranes were washed. It is concluded that, as with the purified E. coli enzyme, the reduction of AcPdAD+ by NADH in chromatophores is a cyclic reaction involving nucleotides that are tightly bound in the domain III site of transhydrogenase. However, in the case of R. rubrum membranes it can be shown with some certainty that the bound nucleotides are NADP+ or NADPH. The data are thus adequately explained without recourse to suggestions of multiple nucleotide-binding sites on transhydrogenase.  相似文献   

3.
When electrospray ionization mass spectrometry (ESMS) was used to analyze purified bovine gamma E (gamma IVa)-crystallin, it yielded a relative molecular mass (M(r)) of 20.955 +/- 5. This mass is significantly different from that calculated from the published sequence (M(r) 20.894) (White HE et al., 1989, J Mol Biol 207:217-235). Further, ES-MS analysis of the protein after it had been reduced and carboxymethylated indicated the presence of five cysteine residues, whereas the published sequence contains six (Kilby GW et al., 1995, Eur Mass Spectrom 1:203-208). The entire protein sequence of gamma E crystallin has therefore been studied via a combination of ES-MS, ES-MS/MS, and Edman amino acid sequencing. The corrected sequence gives an M(r) of 20.955.3, which matches that obtained by ES-MS analysis of the purified native protein. The corrected sequence is also in agreement with a recent cDNA sequence obtained for a bovine gamma-crystallin by R. Hay (pers. comm.).  相似文献   

4.
Thioredoxin reductase, lipoamide dehydrogenase, and glutathione reductase are members of the pyridine nucleotide-disulfide oxidoreductase family of dimeric flavoenzymes. The mechanisms and structures of lipoamide dehydrogenase and glutathione reductase are alike irrespective of the source (subunit M(r) approximately 55,000). Although the mechanism and structure of thioredoxin reductase from Escherichia coli are distinct (M(r) approximately 35,000), this enzyme must be placed in the same family because there are significant amino acid sequence similarities with the other two enzymes, the presence of a redox-active disulfide, and the substrate specificities. Thioredoxin reductase from higher eukaryotes on the other hand has a M(r) of approximately 55,000 [Luthman, M. & Holmgren, A. (1982) Biochemistry 21, 6628-6633; Gasdaska, P. Y., Gasdaska, J. R., Cochran, S. & Powis, G. (1995) FEBS Lett 373, 5-9; Gladyshev, V. N., Jeang, K. T. & Stadtman, T.C. (1996) Proc. Natl. Acad. Sci. USA 93, 6146-6151]. Thus, the evolution of this family is highly unusual. The mechanism of thioredoxin reductase from higher eukaryotes is not known. As reported here, thioredoxin reductase from human placenta reacts with only a single molecule of NADPH, which leads to a stable intermediate similar to that observed in titrations of lipoamide dehydrogenase or glutathione reductase. Titration of thioredoxin reductase from human placenta with dithionite takes place in two spectral phases: formation of a thiolate-flavin charge transfer complex followed by reduction of the flavin, just as with lipoamide dehydrogenase or glutathione reductase. The first phase requires more than one equivalent of dithionite. This suggests that the penultimate selenocysteine [Tamura, T. & Stadtman, T.C. (1996) Proc. Natl. Acad. Sci. USA 93, 1006-1011] is in redox communication with the active site disulfide/dithiol. Nitrosoureas of the carmustine type inhibit only the NADPH reduced form of human thioredoxin reductase. These compounds are widely used as cytostatic agents, so this enzyme should be studied as a target in cancer chemotherapy. In conclusion, three lines of evidence indicate that the mechanism of human thioredoxin reductase is like the mechanisms of lipoamide dehydrogenase and glutathione reductase and differs fundamentally from the mechanism of E. coli thioredoxin reductase.  相似文献   

5.
Previous studies and the crystal structure of Anabaena PCC 7119 FNR suggest that the side chains of Arg100 and Arg264 may be directly involved in the proper NADP+/NADPH orientation for an efficient electron-transfer reaction. Protein engineering on Arg100 and Arg264 from Anabaena PCC 7119 FNR has been carried out to investigate their roles in complex formation and electron transfer to NADP+ and to ferredoxin/flavodoxin. Arg100 has been replaced with an alanine, which removes the positive charge, the long side chain, as well as the ability to form hydrogen bonds, while a charge reversal mutation has been made at Arg264 by replacing it with a glutamic acid. Results with various spectroscopic techniques indicate that the mutated proteins folded properly and that significant protein structural rearrangements did not occur. Both mutants have been kinetically characterized by steady-state as well as fast transient kinetic techniques, and the three-dimensional structure of Arg264Glu FNR has been solved. The results reported herein reveal important conceptual information about the interaction of FNR with its substrates. A critical role is confirmed for the long, positively charged side chain of Arg100. Studies on the Arg264Glu FNR mutant demonstrate that the Arg264 side chain is not critical for the nicotinamide orientation or for nicotinamide interaction with the isoalloxazine FAD moiety. However, this mutant showed altered behavior in its interaction and electron transfer with its protein partners, ferredoxin and flavodoxin.  相似文献   

6.
Thioredoxin reductase (TrxR) catalyzes the reduction of thioredoxin (Trx) by NADPH. A unique gene organization of TrxR and Trx has been found in Mycobacterium leprae, where TrxR and Trx are encoded by a single gene and, therefore, are expressed as a fusion protein (MlTrxR-Trx). This fusion enzyme is able to catalyze the reduction of thioredoxin or 5,5'-dithiobis(2-nitrobenzoic acid) or 1, 4-naphthoquinone by NADPH, though the activity is much lower than that of Escherichia coli TrxR. It has been proposed that a large conformational change is required in catalysis of E. coli TrxR. Because the reductase portion of the enzyme from M. leprae shows significant primary structure similarity with E. coli TrxR, it is possible that MlTrxR-Trx may require a similar conformational change and that the change in conformation may be affected by the tethered Trx. The reductase has been expressed without Trx attached (MlTrxR). As reported here, comparison of the steady-state and pre-steady-state kinetics of MlTrxR-Trx with those of MlTrxR suggests that the low reductase activity of the fusion enzyme is an inherent property of the reductase, and that any steric limitation caused by the attached thioredoxin in the fusion protein makes only a minor contribution to the low activity. Titration of MlTrxR-Trx and MlTrxR with 3-aminopyridine adenine dinucleotide phosphate (AADP+), an NADP(H) analogue, results in only slight quenching of FAD fluorescence, suggesting an enzyme conformation in which the binding site of AADP+ is not close to the FAD, as in one of the conformations of E. coli TrxR.  相似文献   

7.
The flavoprotein component (SiR-FP) of the sulfite reductase from Escherichia coli is an octamer containing one FAD and one FMN as cofactors per polypeptide chain. We have constructed an expression vector containing the DNA fragment encoding for the FMN-binding domain of SiR-FP. The overexpressed protein (SiR-FP23) was purified as a partially flavin-depleted polymer. It could incorporate FMN exclusively upon flavin reconstitution to reach a maximum flavin content of 1.2 per polypeptide chain. Moreover, the protein could stabilize a neutral air-stable semiquinone radical over a wide range of pHs. During photoreduction, the flavin radical accumulated first, followed by the fully reduced state. The redox potentials, determined at room temperature [E'1 (FMNH./FMN) = -130 +/- 10 mV and E'2 (FMNH2/FMNH.) = -335 +/- 10 mV], were very close to those previously reported for Salmonella typhimurium SiR-FP [Ostrowski, J., Barber, M. J., Rueger, D. C., Miller, B. E., Siegel, L. M., & Kredich, N. M. (1989) J. Biol. Chem. 264, 15796-15808]. Both the radical and fully reduced forms of SiR-FP23 were able to transfer their electrons to cytochrome c quantitatively. Altogether, the results presented herein demonstrate that the N-terminal end of E. coli SiR-FP forms the FMN-binding domain. It folds independently, thus retaining the chemical properties of the bound FMN, and provides a good model of the FAD-depleted form of native SiR-FP. Moreover, the FMN prosthetic group in SiR-FP23 and native SiR-FP is compared to that of cytochrome P450 reductase and bacterial cytochrome P450, which also contain one FAD and one FMN per polypeptide chain.  相似文献   

8.
Two soluble flavoproteins, purified from Escherichia coli cytosol and identified as flavodoxin and NADPH-flavodoxin (ferredoxin) reductase (flavodoxin reductase), have been found in combination to support the 17 alpha-hydroxylase activities of heterologously expressed bovine 17 alpha-hydroxylase cytochrome P450 (P450c17). Physical characteristics of the two flavoproteins including absorbance spectra, molecular weights, and amino-terminal sequences are identical with those reported previously for E. coli flavodoxin and flavodoxin reductase. Flavodoxin reductase, possessing FAD as a cofactor, is able to reconstitute P450c17 activities only in the presence of flavodoxin, an FMN-containing protein, and NAD(P)H. Reducing equivalents are utilized more effectively from NADPH than NADH by flavodoxin reductase. E. coli flavodoxin binds P450c17 directly and with relatively high affinity (apparent Ks approximately 0.2 microM) at low ionic strength, as evidenced by a change in spin state of the P450c17 heme iron upon titration with flavodoxin. This apparent spin shift is attenuated at moderate ionic strengths (100-200 mM KCl). In addition, bovine P450c17 binds reversibly to flavodoxin Sepharose in an ionic strength-dependent manner. These data implicate charge pairing as being important for the interaction between flavodoxin and P450c17. We propose that the amino acid sequence similarity between E. coli flavodoxin-flavodoxin reductase and the putative FMN, FAD, and NAD(P)H binding regions of cytochrome P450 reductase provides the basis for the reconstitution of P450c17 activities by this bacterial system.  相似文献   

9.
The immunostimulatory dinitrohalobenzene compound 1-chloro-2, 4-dinitrobenzene (DNCB) irreversibly inhibits mammalian thioredoxin reductase (TrxR) in the presence of NADPH, inducing an NADPH oxidase activity in the modified enzyme (Arnér, E. S. J., Bj?rnstedt, M., and Holmgren, A. (1995) J. Biol. Chem. 270, 3479-3482). Here we have further analyzed the reactivity with the enzyme of DNCB and analogues with varying immunomodulatory properties. We have also identified the reactive residues in bovine thioredoxin reductase, recently discovered to be a selenoprotein. We found that 4-vinylpyridine competed with DNCB for inactivation of TrxR, with DNCB being about 10 times more efficient, and only alkylation with DNCB but not with 4-vinylpyridine induced an NADPH oxidase activity. A number of nonsensitizing DNCB analogues neither inactivated the enzyme nor induced any NADPH oxidase activity. The NADPH oxidase activity of TrxR induced by dinitrohalobenzenes generated superoxide, as detected by reaction with epinephrine (the adrenochrome method). Addition of superoxide dismutase quenched this reaction and also stimulated the NADPH oxidase activity. By peptide analysis using mass spectrometry and Edman degradation, both the cysteine and the selenocysteine in the conserved carboxyl-terminal sequence Gly-Cys-Sec-Gly (where Sec indicates selenocysteine) were determined to be dinitrophenyl-alkylated upon incubation of native TrxR with NADPH and DNCB. A model for the interaction between TrxR and dinitrohalobenzenes is proposed, involving a functional FAD in the alkylated TrxR generating an anion nitroradical in a dinitrophenyl group, which in turn reacts with oxygen to generate superoxide. Production of reactive oxygen species and inhibited reduction of thioredoxin by the modified thioredoxin reductase after reaction with dinitrohalobenzenes may play a major role in the inflammatory reactions provoked by these compounds.  相似文献   

10.
Vertebrate ferredoxins are 12-14-kDa iron-sulfur proteins, some of which transfer electrons to mitochondrial cytochrome P450s. The function of many of these cytochrome P450s is to catalyze stereospecific hydroxylation of endogenous steroids. As part of our interest in the kidney mitochondrial 1 alpha-hydroxylation of 25-hydroxyvitamin D3, we have constructed an expression plasmid coding for a fusion protein containing the chick kidney ferredoxin. We subcloned chick kidney ferredoxin cDNA, obtained from our vitamin D-deficient chick kidney library by polymerase chain reaction (Brandt, M. E., Gabrik, A. H., and Vickery, L. E. (1991) Gene (Amst.) 97, 113-117) into Qiagen's pQE9, which contains an N-terminal 6xHis tag (peptide sequence for 6 adjacent histidines present in the recombinant proteins). The coding sequence was preceded by a factor Xa cleavage site. The resulting plasmid, pQTcFdx, was overexpressed in Escherichia coli, and the soluble fusion protein was purified from the cell lysate in one step by Ni(II)-nitrilotriacetic acid-agarose chromatography. We obtained 7-10 mg of greater than 99% homogeneous fusion protein from a 1-liter culture and 4-6 mg of mature ferredoxin cleaved by factor Xa. The fusion protein possessed an absorption spectrum and an electron paramagnetic resonance spectrum quantitatively indistinguishable from those published for ferredoxin purified from adrenal glands and placenta or expressed in E. coli with another vector. The fusion protein was active in supporting the 1 alpha-hydroxylation of 25-hydroxyvitamin D3 in a reconstitution assay of a solubilized, partially purified preparation of cytochrome P450 from vitamin D-deficient chick kidney. We conclude that the procedure described here is an efficient way to produce and purify vertebrate ferredoxin; the [2Fe-2S] cofactor is assembled in vivo and effectively incorporated into the fusion protein in E. coli; slight alterations at the N terminus do not alter incorporation of the [2Fe-2S] cofactor or the biological activity of ferredoxin, and post-translational modifications, such as phosphorylation, are not an absolute requirement for ferredoxin electron transporting activity. The recombinant ferredoxin can be used for physical studies and other structure-function studies.  相似文献   

11.
Aquacobalamin reductase (NADPH), which catalyzes the reduction of aquacobalamin to cob(II)alamin in the synthesis of cobalamin coenzymes, has already been purified from mitochondria of Euglena gracilis and partly characterized. Here, the enzyme was further characterized to clarify its enzymatic properties. The enzyme reduced 2 mol of aquacobalamin per mole of NADPH and had NADPH diaphorase-like activity. The 16 amino acid residues at the NH2-terminal of the enzyme were identical with those of the NADPH diaphorase domain of pyruvate: NADP+ oxidoreductase, which is involved in Euglena wax ester fermentation. Peptide mapping of the aquacobalamin reductase showed that elution during C-18 reversed-phase high-performance liquid chromatography was identical to that of the NADPH diaphorase domain. Immunoblotting indicated that the Euglena aquacobalamin reductase had a higher molecular weight (166,000) in the intact mitochondria than the purified enzyme (65,000), and that the molecular weights of the native and purified enzyme were identical with those of the subunit and the NADPH diaphorase domain, respectively. These results showed that the aquacobalamin reductase isolated earlier was the NADPH diaphorase domain, cleaved by trypsin during preparation of the mitochondrial homogenate from the native enzyme. Purified pyruvate:NADP+ oxidoreductase also had the activity of aquacobalamin reductase, which suggests that the enzyme in Euglena mitochondria has more than one function in the synthesis of cobalamin co-enzymes.  相似文献   

12.
The basic nonhistone phosphoprotein 110/8.4 (M.W. X 10(-3)/pI) was found in 0.35 M NaCl nuclear extracts of four tumor tissues, i.e., fast-growing Novikoff hepatoma, Morris hepatoma 3924A, HeLa cells, and Namalwa cells; it was also found in fetal rat liver. This protein was not detected in normal or regenerating liver and thus may represent an "oncofetal" protein of potential interest as a cancer "marker." Protein 110/8.4 was purified approximately 4000- to 5000-fold under nondenaturing condition from 0.35 M NaCl nuclear extracts of Novikoff hepatoma cells or Namalwa cells by ammonium sulfate fractionation, calcium phosphate gel treatment, and phosphocellulose chromatography. Sodium dodecyl sulfate:polyacrylamide gel electrophoresis of the purified native protein revealed a single polypeptide chain with a molecular weight of approximately 110,000. The pI of the protein was estimated to be 8.4 by nonequilibrium pH gradient electrophoresis in 9 M urea; accordingly, this protein was designated 110/8.4. Amino acid analysis showed that Protein 110/8.4 had an acidic:basic amino acid ratio of 1.25 and a high lysine and serine content; approximately 20% of the serine residues were found to be phosphorylated. Hydrazinolysis indicated that the carboxyl-terminal amino acid was serine; the amino terminus appeared to be blocked. Binding of Protein 110/8.4 to DNA was studied by the nitrocellulose filter assay. High-affinity binding occurred at ionic strength equal to or below 0.15 M.  相似文献   

13.
Pyruvate ferredoxin oxidoreductase (POR) from the hyperthermophilic archaeon Pyrococcus furiosus (Pf) catalyzes the final oxidative step in carbohydrate fermentation in which pyruvate is oxidized to acetyl-CoA and CO2, coupled to the reduction of ferredoxin (Fd). POR is composed of two 'catalytic units' of molecular mass approximately 120 kDa. Each unit consists of four subunits, alpha beta gamma delta, with masses of approximately 44, 36, 20, and 12 kDa, respectively, and contains at least two [4Fe-4S] clusters. The precise mechanism of catalysis and the role of the individual subunits are not known. The gene encoding the delta-subunit of Pf POR has been expressed in E. coli, and the protein was purified after reconstitution with iron and sulfide. The reconstituted delta-subunit (recPOR-delta) is monomeric with a mass of 11 879 +/- 1.2 Da as determined by mass spectrometry, in agreement with that predicted from the gene sequence. Purified recPOR-delta contains 8 Fe mol/mol and remained intact when incubated at 85 degreesC for 2 h, as judged by its visible absorption properties. The reduced form of the protein exhibited an EPR spectrum characteristic of two, spin-spin interacting [4Fe-4S]1+ clusters. When compared with the EPR properties of the reduced holoenzyme, the latter was shown to contain a third [4Fe-4S]1+ cluster in addition to the two within the delta-subunit. The reduction potential of the two 4Fe clusters in isolated recPOR-delta (-403 +/- 8 mV at pH 8.0 and 24 degreesC) decreased linearly with temperature (-1.55 mV/ degreesC) up to 82 degreesC. RecPOR-delta replaced Pf Fd as an in vitro electron carrier for two oxidoreductases from Pf, POR and Fd:NADP oxidoreductase, and the POR holoenzyme displayed a higher apparent affinity for its own subunit (apparent Km = 1.0 microM at 80 degreesC) than for Fd (apparent Km = 4.4 microM). The molecular and spectroscopic properties and amino acid sequence of the isolated delta-subunit suggest that it evolved from an 8Fe-type Fd by the addition of approximately 40 residues at the N-terminus, and that this extension enabled it to interact with additional subunits within POR.  相似文献   

14.
Hemoglobins (Hb), which have the important task of delivering molecular oxygen by facilitating its reversible binding to the heme, are now thought to have evolved in all groups of organisms including prokaryotes, fungi, plants and animals. Our recent finding of a light-inducible chloroplastic Hb in the green unicellular alga Chlamydomonas eugametos has further extend this idea, while raising questions about the function that an Hb could play in a high oxygen environment such as in the chloroplast. In order to understand the role played by this new Hb, we have undertaken its biochemical characterization. To facilitate the characterization of Chlamydomonas Hb, which represents less than 0.01% of the soluble protein in the green alga, the protein has been expressed in Escherichia coli and purified to apparent homogeneity. The purified recombinant protein possesses a non-covalently bound iron-protoporphyrin IX heme. The oxy form of the recombinant Hb. purified directly from bacterial cells, is very stable, with a measured half-life of 7 days at pH 8 and has an ultraviolet/visible spectrum similar to those of the related cytoplasmic Hbs of the ciliated protozoa Paramecium and Tetrahymena and of the cyanobacterium Nostoc commune. In contrast to what has been reported for oxymyoglobins and oxyhemoglobins, the dioxygen molecule bound to the L1637 Hb can be reduced by the electron-transfer mediator phenazine methosulfate in the presence of NADPH, indicating that the heme pocket of Chlamydomonas Hb may be more accessible to small molecules. With regard to this we found that when the small reducing agent sodium dithionite is used to reduce the met form, it must be removed anaerobically from the Hb prior to oxygenation of the protein to stably produce the oxy form. Otherwise, the oxy form is obtained readily from the met form under an oxygenic atmosphere when ferredoxin and ferredoxin NADP+ reductase are used to enzymically reduce the Hb. Finally, the spectra of the deoxy and met forms were unusual, the heme being partly low-spin at physiological pH. These results confirm the existence of a reversible oxygen-binding protein in the chloroplast of C. eugametos. The unusual spectral and biochemical properties of the protein may reflect a specialized function for this Hb.  相似文献   

15.
Aromatic ketone reductase activity of microsomes showed a unique cofactor requirement: Addition of NADP and glucose-6-phosphate was as effective as that of an artificial NADPH generating system, whereas NADPH alone served as a cofactor less efficiently. Microsomal aromatic ketone reductase, purified partially from guinea pig liver microsomes after solubilization with Triton X-100, reduced 5 beta-dihydrotestosterone, aromatic aldehydes, and ketones with NADPH as a cofactor. However, addition of hexose-6-phosphate dehydrogenase, purified from the same source, as an NADPH generator produced about 2 times higher activity than that of yeast glucose-6-phosphate dehydrogenase or NADPH alone.  相似文献   

16.
In Azotobacter vinelandii, deletion of the fdxA gene, which encodes ferredoxin I (FdI), leads to activation of the expression of the fpr gene, which encodes NADPH-ferredoxin reductase (FPR). In order to investigate the relationship of these two proteins further, the interactions of the two purified proteins have been examined. AvFdI forms a specific 1:1 cross-linked complex with AvFPR through ionic interactions formed between the Lys residues of FPR and Asp/Glu residues of FdI. The Lys in FPR has been identified as Lys258, a residue that forms a salt bridge with one of the phosphate oxygens of FAD in the absence of FdI. UV-Vis and circular dichroism data show that on binding FdI, the spectrum of the FPR flavin is hyperchromatic and red-shifted, confirming the interaction region close to the FAD. Cytochrome c reductase assays and electron paramagnetic resonance data show that electron transfer between the two proteins is pH-dependent and that the [3Fe-4S]+ cluster of FdI is specifically reduced by NADPH via FPR, suggesting that the [3Fe-4S] cluster is near FAD in the complex. To further investigate the FPR:FdI interaction, the electrostatic potentials for each protein were calculated. Strongly negative regions around the [3Fe-4S] cluster of FdI are electrostatically complementary with a strongly positive region overlaying the FAD of FPR, centered on Lys258. These proposed interactions of FdI with FPR are consistent with cross-linking, peptide mapping, spectroscopic, and electron transfer data and strongly support the suggestion that the two proteins are physiological redox partners.  相似文献   

17.
18.
Cell-free extracts of Methanobacterium thermoautotrophicum were found to contain high activities of the following oxidoreductases (at 60 degrees C): pyruvate dehydrogenase (coenzyme A acetylating), 275 nmol/min per mg of protein; alpha-ketoglutarate dehydrogenase (coenzyme A acylating), 100 nmol/min per mg; fumarate reductase, 360 nmol/min per mg; malate dehydrogenase, 240 nmol/min per mg; and glyceraldehyde-3-phosphate dehydrogenase, 100 nmol/min per mg. The kinetic properties (apparent V(max) and K(M) values), pH optimum, temperature dependence of the rate, and specificity for electron acceptors/donors of the different oxidoreductases were examined. Pyruvate dehydrogenase and alpha-ketoglutarate dehydrogenase were shown to be two separate enzymes specific for factor 420 rather than for nicotinamide adenine dinucleotide (NAD), NADP, or ferredoxin as the electron acceptor. Both activities catalyzed the reduction of methyl viologen with the respective alpha-ketoacid and a coenzyme A-dependent exchange between the carboxyl group of the alpha-ketoacid and CO(2). The data indicate that the two enzymes are similar to pyruvate synthase and alpha-ketoglutarate synthase, respectively. Fumarate reductase was found in the soluble cell fraction. This enzyme activity coupled with reduced benzyl viologen as the electron donor, but reduced factor 420, NADH, or NADPH was not effective. The cells did not contain menaquinone, thus excluding this compound as the physiological electron donor for fumarate reduction. NAD was the preferred coenzyme for malate dehydrogenase, whereas NADP was preferred for glyceraldehyde-3-phosphate dehydrogenase. The organism also possessed a factor 420-dependent hydrogenase and a factor 420-linked NADP reductase. The involvement of the described oxidoreductases in cell carbon synthesis is discussed.  相似文献   

19.
Thermal unfolding of dodecameric manganese glutamine synthetase (622,000 M(r)) at pH 7 and approximately 0.02 ionic strength occurs in two observable steps: a small reversible transition (Tm approximately 42 degrees C; delta H approximately equal to 0.9 J/g) followed by a large irreversible transition (Tm approximately 81 degrees C; delta H approximately equal to 23.4 J/g) in which secondary structure is lost and soluble aggregates form. Secondary structure, hydrophobicity, and oligomeric structure of the equilibrium intermediate are the same as for the native protein, whereas some aromatic residues are more exposed. Urea (3 M) destabilizes the dodecamer (with a tertiary structure similar to that without urea at 55 degrees C) and inhibits aggregation accompanying unfolding at < or = 0.2 mg protein/mL. With increasing temperature (30-70 degrees C) or incubation times at 25 degrees C (5-35 h) in 3 M urea, only dodecamer and unfolded monomer are detected. In addition, the loss in enzyme secondary structure is pseudo-first-order (t1/2 = 1,030 s at 20.0 degrees C in 4.5 M urea). Differential scanning calorimetry of the enzyme in 3 M urea shows one endotherm (Tmax approximately 64 degrees C; delta H = 17 +/- 2 J/g). The enthalpy change for dissociation and unfolding agrees with that determined by urea titrations by isothermal calorimetry (delta H = 57 +/- 15 J/g; Zolkiewski M, Nosworthy NJ, Ginsburg A, 1995, Protein Sci 4: 1544-1552), after correcting for the binding of urea to protein sites exposed during unfolding (-42 J/g). Refolding and assembly to active enzyme occurs upon dilution of urea after thermal unfolding.  相似文献   

20.
NAD kinase is a Ca2+/calmodulin (CaM)-dependent enzyme capable of converting cellular NAD to NADP. The enzyme purified from pea seedlings can be activated by highly conserved soybean CaM, SCaM-1, but not by the divergent soybean CaM isoform, SCaM-4 (Lee, S. H., Kim, J. C., Lee, M. S., Heo, W. D., Seo, H. Y., Yoon, H. W., Hong, J. C., Lee, S. Y., Bahk, J. D., Hwang, I., and Cho, M. J. (1995) J. Biol. Chem. 270, 21806-21812). To determine which domains were responsible for this differential activation of NAD kinase, a series of chimeric SCaMs were generated by exchanging functional domains between SCaM-4 and SCaM-1. SCaM-4111, a chimeric SCaM-1 that contains the first domain of SCaM-4, was severely impaired (only 40% of maximal) in its ability to activate NAD kinase. SCaM-1444, a chimeric SCaM-4 that contains the first domain of SCaM-1 exhibited nearly full ( approximately 70%) activation of NAD kinase. Only chimeras containing domain I of SCaM-1 produced greater than half-maximal activation of NAD kinase. To define the amino acid residue(s) in domain I that were responsible for this differential activation, seven single residue substitution mutants of SCaM-1 were generated and tested for NAD kinase activation. Among these mutants, only K30E and G40D showed greatly reduced NAD kinase activation. Also a double residue substitution mutant, K30E/G40D, containing these two mutations in combination was severely impaired in its NAD kinase-activating potential, reaching only 20% of maximal activation. Furthermore, a triple mutation, K30E/M36I/G40D, completely abolished NAD kinase activation. Thus, our data suggest that domain I of CaM plays a key role in the differential activation of NAD kinase exhibited by SCaM-1 and SCaM-4. Further, the residues Lys30 and Glu40 of SCaM-1 are critical for this function.  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