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 tryptophan residues of two forms of pig heart lipoamide dehydrogenase (LD(I) and LD(II] were investigated fluorometrically. The tryptophan contents of LD(I) and LD(II) determined by the fluorescence method were 3 mol and 2 mol per mol of FAD, respectively. These values were in good agreement with those found by the MCD method. The microenvironments of the tryptophan residues were investigated by fluorescence quenching titration with acrylamide. The tryptophan residues of both enzymes were in heterogeneous microenvironments, and CD spectra showed some differences between these microenvironments in the two enzymes. Energy transfer from tryptophan residues to bound FAD was equally efficient in the two enzymes. It seems probable that the three tryptophan residues in LD(I) are all in different microenvironments, but that two of them are in microenvironments almost identical to those of the corresponding residues in LD(II).  相似文献   

2.
The enzymatic and biochemical properties of human insulin-degrading enzyme and Escherichia coli protease III have been compared. Both enzymes were found to degrade insulin in such a way that its receptor binding activity was rapidly lost but its precipitability in trichloracetic acid was only slightly decreased. Both enzymes were also found to be inhibited by chelating agents. The bacterial enzyme, which could be purified in large amounts, was found to contain 0.6 mol of zinc per mol of enzyme but no detectable manganese. The mammalian enzyme but not the bacterial one was inhibited by a sulfhydryl alkylating agent. The two enzymes also differed in substrate specificity. The mammalian enzyme degraded insulin much better than insulin-like growth factor II, whereas the bacterial enzyme degraded them equally. The mammalian enzyme could be labeled by cross-linking to insulin = bombyxin II much greater than insulin-like growth factor I and II much greater than relaxin, while the bacterial enzyme was labeled by insulin-like growth factor II greater than insulin = insulin-like growth factor I much greater than relaxin much greater than bombyxin. Finally, sucrose gradient centrifugation and cross-linking studies both in vitro and in vivo indicated that active human enzyme partially existed as a homo- or heterodimer, whereas the bacterial enzyme was active as a monomer.  相似文献   

3.
Soybean (Glycine max (L.) Merr.) root nodules contain the enzymes of the ascorbate-glutathione cycle as an important defense against activated forms of oxygen. A key enzyme in this cycle--monodehydroascorbate reductase (MR)--was purified 646-fold and appeared as a single band on SDS-PAGE with silver or Coomassie blue staining. Purified MR contained 0.7 mol FAD/mol enzyme and had a specific activity of 288 mumol NADH oxidized.min-1.mg protein-1. The enzyme was a single subunit occurring as two isozymes (MR I and MR II) with Mr values of 39,000 and 40,000. Isoelectric focusing revealed that each isozyme consisted of two forms with pl values of 4.6 to 4.7. Ferricyanide and 2,6-dichlorophenol-indophenol were effective as electron acceptors. The purified enzyme did not possess leghemoglobin reductase activity. Inhibition by p-chloromercuribenzoate indicated the involvement of a thiol group in MR activity. The Km values were 5.6, 150, and 7 microM for NADH, NADPH, and monodehydroascorbate, respectively. The pH optimum was 8 to 9. The N-terminal sequence of 10 amino acids of MR II had little homology to known protein sequences.  相似文献   

4.
The binding of pyridine nucleotide to human erythrocyte glutathione reductase, an enzyme of known three-dimensional structure, requires some movement of the side chain of Tyr197. Moreover, this side chain lies very close to the isoalloxazine ring of the FAD cofactor. The analogous residue, Ile184, in the homologous enzyme Escherichia coli lipoamide dehydrogenase has been altered by site-directed mutagenesis to a tyrosine residue (I184Y) [Russell, G. C., Allison, N., Williams, C. H., Jr., & Guest, J.R. (1989) Ann. N.Y. Acad. Sci. 573, 429-431]. Characterization of the altered enzyme shows that the rate of the pyridine nucleotide half-reaction has been markedly reduced and that the spectral properties have been changed to mimic those of glutathione reductase. Therefore, Ile184 is shown to be an important residue in modulating the properties of the flavin in lipoamide dehydrogenase. Turnover in the dihydrolipoamide/NAD+ reaction is decreased by 10-fold and in the NADH/lipoamide reaction by 2-fold in I184Y lipoamide dehydrogenase. The oxidized form of I184Y shows remarkable changes in the fine structure of the visible absorption and circular dichroism spectra and also shows nearly complete quenching of FAD fluorescence. The spectral properties of the altered enzyme are thus similar to those of glutathione reductase and very different from those of wild-type lipoamide dehydrogenase. On the other hand, spectral evidence does not reveal any change in the amount of charge-transfer stabilization at the EH2 level. Stopped-flow data indicate that, in the reduction of I184Y by NADH, the first step, reduction of the flavin, is only slightly slowed but the subsequent two-electron transfer to the disulfide is markedly inhibited.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
A fluorescent probe, N-(1-anilinonaphth-4-yl)-maleimide (ANM), was specifically labeled to SH group(s) in the hydrophilic moiety of NADPH-cytochrome P-450 reductase at a ratio of 1 +/- 0.1 ANM/mol of protein. The ANM-labeled reductase and P-450 were reconstituted in phosphatidylcholine-phosphatidylethanolamine-phosphatidylserine vesicles in which all of the enzymes were functionally active. The reconstitution of the mixed-function oxidase system was found to be strongly dependent on both the lipid to protein molar ratio and phospholipid composition. The interactions of ANM-labeled reductase with P-450 in proteoliposomes were investigated by perturbation of the fluorescence of ANM. Upon incorporation of P-450 into the phospholipids vesicles (ANM-reductase/P-450/lipids identical to 1:1.4:800), a significant decrease of total fluorescence intensity and slight increase of emission anisotropy of ANM were observed. In the average fluorescence lifetime of ANM bound with reductase, an appreciable change was shown between the absence and presence of P-450 in the vesicles. These data provide clear evidence that significant molecular interactions occur between the two proteins in a membranous reconstituted system.  相似文献   

6.
The fluorescent thiol reagent N-(1-anilinonaphthyl-4)maleimide (ANM) reacts covalently with the Ca2+ ATPase moiety of fragmented sarcoplasmic reticulum in two phases as determined by the increase of fluorescence intensity and optical density at 350 nm. In the rapid phase, 5.5 nmol of ANM reacts with 1 mg of fragmented sarcoplasmic reticulum protein. Assuming that 55% of the total membrane protein is the Ca2+ ATPase, this is equivalent to 1 mol of SH/10(5) g of ATPase, designated as SH1-ANM. ANM reacts with the second SH (SH2-ANM) at a much slower rate. Reaction of ANM with both SH1-ANM and SH2-ANM produces no inhibition of phosphoenzyme (EP) formation. Upon addition of Mg . ATP in the micromolar range, at [Ca2+] = 1 microM there is an increase in the fluorescence intensity of ANM attached to SH2-ANM, while the ANM attached to SH1-ANM does not respond to Mg . ATP. Under conditions in which there is no EP formation, there is no fluorescence change. Furthermore, the enhancement of ANM fluorescence produced by Mg . ATP is reversed by ADP as it reacts with EP to form ATP. Thus, it appears that the Mg . ATP-induced fluorescence increase reflects changes of enzyme conformation produced by EP formation.  相似文献   

7.
The effect of NAD+ on lipoamide dehydrogenase from pig heart was investigated physicochemically. The observed and theoretical oxidation-reduction mid-point potentials for the oxidized lipoamide dehydrogenase (E)/two-electron-reduced lipoamide dehydrogenase (EH2) couple in the presence on NAD+ were -218 mV and -251 mV, respectively, at pH 6.0. Therefore, unexpectedly the mid-point potential of the enzyme became more positive on NAD+ binding. Decreases in the fluorescence lifetime and intensity and increase in the degree of polarization of enzyme-bound FAD were observed in the presence of NAD+. Fluorescence quenching of bound FAD by NAD+ was released by phenobarbital. The results suggest that NAD+ strengthens the intramolecular dynamic interaction between the isoalloxazine moiety and adenine moiety of bound FAD, and so alters the mid-point potential of the enzyme. These findings indicate that NAD+ acts not only as an acceptor of electrons from EH2, but also as an effector in the flavin-disulfide interaction of EH2.  相似文献   

8.
Monodehydroascorbate reductase (EC 1.6.5.4) was purified from cucumber fruit to a homogeneous state as judged by polyacrylamide gel electrophoresis. The cucumber monodehydroascorbate reductase was a monomer with a molecular weight of 47,000. It contained 1 mol of FAD/mol of enzyme which was reduced by NAD(P)H and reoxidized by monodehydroascorbate. The enzyme had an exposed thiol group whose blockage with thiol reagents inhibited the electron transfer from NAD(P)H to the enzyme FAD. Both NADH and NADPH served as electron donors with Km values of 4.6 and 23 microM, respectively, and Vmax of 200 mol of NADH and 150 mol of NADPH oxidized mol of enzyme-1 s-1. The Km for monodehydroascorbate was 1.4 microM. The amino acid composition of the enzyme is presented. In addition to monodehydroascorbate, the enzyme catalyzed the reduction of ferricyanide and 2,6-dichloroindophenol but showed little reactivity with calf liver cytochrome b5 and horse heart cytochrome c. The kinetic data suggested a ping-pong mechanism for the monodehydroascorbate reductase-catalyzed reaction. Cucumber monodehydroascorbate reductase occurs in soluble form and can be distinguished from NADPH dehydrogenase, NADH dehydrogenase, DT diaphorase, microsome-bound NADH-cytochrome b5 reductase, and NADPH-cytochrome c reductase by its molecular weight, amino acid composition, and specificity of electron acceptors and donors.  相似文献   

9.
The time dependence of the fluorescence of tryptophanyl and flavin residues in lipoamide dehydrogenase has been investigated with single-photon decay spectroscopy. When the two FAD molecules in the enzyme were directly excited the decay could only be analyzed in a sum of two exponentials with equal amplitudes. This phenomenon was observed at 4 degrees C (tau-1 = 0.8 ns, tau-2 = 4.7 ns) and at 20 degrees C (tau-1 = 0.8 ns, tau-2 = 3.4 ns) irrespective of the emission and excitation wavelengths. This result reveals a difference in the nature of the two FAD centers. By excitation at 290 nm the fluorescence decay curves of tryptophan and FAD were obtained. The decays are analyzed in terms of energy transfer from tryptophanyl to flavin residues. The results, which are in good agreement with those obtained previously with static fluorescence methods, show that one of the two tryptophanyl residues within the subunit transfers its excitation energy to the flavin located at a distance of 1.5 nm.  相似文献   

10.
Time-resolved polarized flavin fluorescence was used to study the active site dynamics of Escherichia coli glutathione reductase (GR). Special consideration was given to the role of Tyr177, which blocks the access to the NADPH binding-site in the crystal structure of the enzyme. By comparing wild-type GR with the mutant enzymes Y177F and Y177G, a fluorescence lifetime of 7 ps that accounts for approximately 90% of the fluorescence decay could be attributed to quenching by Y177. Based on the temperature invariance for this lifetime, and the very high quenching rate, electron transfer from Y177 to the light-excited isoalloxazine part of flavin adenine dinucleotide (FAD) is proposed as the mechanism of flavin fluorescence quenching. Contrary to the mutant enzymes, wild-type GR shows a rapid fluorescence depolarization. This depolarization process is likely to originate from a transient charge transfer interaction between Y177 and the light-excited FAD, and not from internal mobility of the flavin, as has previously been proposed. Based on the fluorescence lifetime distributions, the mutants Y177F and Y177G have a more flexible protein structure than wild-type GR: in the range of 223 K to 277 K in 80% glycerol, both tyrosine mutants mimic the closely related enzyme dihydrolipoyl dehydrogenase. The fluorescence intensity decays of the GR enzymes can only be explained by the existence of multiple quenching sites in the protein. Although structural fluctuations are likely to contribute to the nonexponential decay and the probability of quenching by a specific site, the concept of conformational substates need not be invoked to explain the heterogeneous fluorescence dynamics.  相似文献   

11.
The 10 C-terminal residues are not visible in the crystal structure of lipoamide dehydrogenase from Azotobacter vinelandii, but can be observed in the crystal structures of the lipoamide dehydrogenases from Pseudomonas putida and Pseudomonas fluorescens. In these structures, the C-terminus folds back towards the active site and is involved in interactions with the other subunit. The function of the C-terminus of lipoamide dehydrogenase from A. vinelandii was studied by deletion of 5, 9 and 14 residues, respectively. Deletion of the last 5 residues does not influence the catalytic properties and conformational stability (thermoinactivation and unfolding by guanidinium hydrochloride). Removal of 9 residues results in an enzyme (enzyme delta 9) showing decreased conformational stability and high sensitivity toward inhibition by NADH. These features are even more pronounced after deletion of 14 residues (enzyme delta 14). In addition Tyr16, conserved in all lipoamide dehydrogenases sequenced thus far, and shown from the other structures to be likely to be involved in subunit interaction, was replaced by Phe and Ser. Mutation of Tyr16 also results in a strongly increased sensitivity toward inhibition by NADH. The conformational stability of both Tyr16-mutated enzymes is comparable to enzyme delta 9. The results strongly indicate that a hydrogen bridge between tyrosine of one subunit (Tyr16 in the A. vinelandii sequence) and histidine of the other subunit (His470 in the A. vinelandii sequence), exists in the A. vinelandii enzyme. In the delta 9 and delta 14 enzymes this interaction is abolished. It is concluded that this interaction mediates the redox properties of the FAD via the conformation of the C-terminus containing residues 450-470.  相似文献   

12.
NADH peroxidase (EC 1.11.1.1) previously isolated from Streptococcus faecalis 10C1 has been crystallized. The crystal structure has been solved by multiple isomorphous replacement and solvent-flattening at 3.3 A (1 A = 0.1 nm) resolution. The enzyme forms a tetramer consisting of 4 crystallographically related subunits. The monomer chain fold is in general similar to those of glutathione reductase and lipoamide dehydrogenase. FAD binds in the same region and in a similar conformation as in glutathione reductase. The unusual cysteine-sulfenic acid participating in catalysis is located at the isoalloxazine of FAD.  相似文献   

13.
This study describes the enzymatic characterization of dihydrolipoamide dehydrogenase (DLDH) from Streptococcus pneumoniae and is the first characterization of a DLDH that carries its own substrate (a lipoic acid covalently attached to a lipoyl protein domain) within its own sequence. Full-length recombinant DLDH (rDLDH) was expressed and compared with enzyme expressed in the absence of lipoic acid (rDLDH(-LA)) or with enzyme lacking the first 112 amino acids constituting the lipoyl protein domain (rDLDH(-LIPOYL)). All three proteins contained 1 mol of FAD/mol of protein, had a higher activity for the conversion of NAD(+) to NADH than for the reaction in the reverse direction, and were unable to use NADP(+) and NADPH as substrates. The enzymes had similar substrate specificities, with the K(m) for NAD(+) being approximately 20 times higher than that for dihydrolipoamide. The kinetic pattern suggested a Ping Pong Bi Bi mechanism, which was verified by product inhibition studies. The protein expressed without lipoic acid was indistinguishable from the wild-type protein in all analyses. On the other hand, the protein without a lipoyl protein domain had a 2-3-fold higher turnover number, a lower K(I) for NADH, and a higher K(I) for lipoamide compared with the other two enzymes. The results suggest that the lipoyl protein domain (but not lipoic acid alone) plays a regulatory role in the enzymatic characteristics of pneumococcal DLDH.  相似文献   

14.
From Trypanosoma cruzi, the causative agent of Chagas' disease, a lipoamide dehydrogenase was isolated. The enzyme, an FAD-cystine oxidoreductase, shares many physical and chemical properties with T. cruzi trypanothione reductase, the key enzyme of the parasite's thiol metabolism. 1. From 60 g epimastigotic T. cruzi cells, 2.7 mg lipoamide dehydrogenase was extracted. The flavoenzyme was purified 3000-fold to homogeneity with an overall yield of 26%. 2. The enzyme is a dimer with a subunit Mr of 55,000. With 1 mM lipoamide (Km approximately 5 mM) and 100 microM NADH (Km = 23 microM), the specific activity at pH 7.0 is 297 U/mg. 3. With excess NADH, the enzyme is reduced to the EH2.NADH complex and, by addition of lipoamide, it is reoxidized, indicating that it can cycle between the oxidized state E and the two-electron-reduced state, EH2. 4. As shown by N-terminal sequencing of the enzyme, 21 out of 30 positions are identical with those of pig heart and human liver lipoamide dehydrogenase. The sequenced section comprises the GGGPGG stretch, which represents the binding site for the pyrophosphate moiety of FAD. 5. After reduction of Eox to the two-electron-reduced state, the enzyme is specifically inhibited by the nitrosourea drug 1,3-bis(2-chloroethyl)-1-nitrosourea (Carmustine), presumably by carbamoylation at one of the nascent active-site thiols. 6. Polyclonal rabbit antibodies raised against T. cruzi lipoamide dehydrogenase and trypanothione reductase are specific for the respective enzyme, as shown by immunoblots of the pure proteins and of cell extracts.  相似文献   

15.
Lipoamide dehydrogenase, a component of the bovine adrenal ketoglutarate dehydrogenase complex, catalyzes the oxidation of NADH by p-quinones and ferricyanide. The kinetics of oxidation obey the ping-pong mechanism. At pH 7.0, the constants for the active center oxidation by quinones (kox) are equal to 1.1 X 10(4)-5.3 X 10(5) M-1s-1 and increase as the acceptor potential rises. The values of kox for quinones change insignificantly within the pH range of 7.7-5.0, whereas that for ferricyanide increases 10-fold with a decrease of pH from 7.0 to 5.0. The value of the catalytic constant for the enzyme (kcat) reaches its maximum at pH 5.5. The quinones interact with the thiol groups of lipoamide dehydrogenase by inhibiting the fluorescence of FAD and diaphorease activity. The reaction is catalyzed by a basic amino acid (pK 6.7) within the composition of the enzyme.  相似文献   

16.
A temperature-dependent change in the microenvironment of the coenzyme, FAD, of D-amino acid oxidase was investigated by means of steady-state and picosecond time-resolved fluorescence spectroscopy. Relative emission quantum yields from FAD bound to D-amino acid oxidase revealed the temperature transition when concentration of the enzyme was lowered. The observed fluorescence decay curves were well described with four-exponential decay functions. The amplitude of the shortest lifetime (tau 0), approximately 25 ps, was always negative, which indicates that the fluorescence of D-amino acid oxidase at approximately 520 nm appears after a metastable state of the excited isoalloxazine decays. The other components with positive amplitudes were assigned to dimer or associated forms of the enzyme, monomer, and free FAD dissociated from the enzyme. Ethalpy and entropy changes of intermediate states in the quenching processes were evaluated according to the absolute rate theory. The temperature transition was much more pronounced in the monomer than in the dimer or associated forms of the enzyme.  相似文献   

17.
Wu X  Takahashi M  Chen SG  Monnier VM 《Biochemistry》2000,39(6):1515-1521
Amadoriases are a novel class of FAD enzymes which catalyze the oxidative deglycation of glycated amino acids to yield corresponding amino acids, glucosone, and H(2)O(2). We previously reported the purification and characterization of two amadoriase isoenzymes from Aspergillus fumigatus and the molecular cloning of amadoriase II. To identify the primary structure of amadoriase I, we prepared a cDNA library from Aspergillus fumigatus and isolated a clone using a probe amplified by polymerase chain reaction with primers designed according to the partial amino acid sequences from peptide mapping. The primary structure of the enzyme deduced from the nucleotide sequence comprises 445 amino acid residues. The enzyme contains 1 mol of FAD as a cofactor, which is covalently linked to Cys342, as determined by mutagenesis analysis, matrix-assisted laser desorption/ionization time-of-flight mass spectrometry, and electrospray ionization-collisional-activated dissociation tandem mass spectrometry. Sequence alignment studies show that amadoriase I has 22% homology with monomeric sarcosine oxidase in which FAD is also linked to a homologous Cys residue. Amadoriases are of potential importance as tools for uncoupling hyperglycemia and glycation reactions that are thought to play a role in diabetic complications.  相似文献   

18.
The flavoprotein lipoamide dehydrogenase was purified, by an improved method, from commercial baker's yeast about 700-fold to apparent homogeneity with 50-80% yield. The enzyme had a specific activity of 730-900 U/mg (about twice the value of preparations described previously). The holoenzyme, but not the apoenzyme, possessed very high stability against proteolysis, heat, and urea treatment and could be reassociated, with fair yield, with the other components of yeast pyruvate dehydrogenase complex to give the active multienzyme complex. The apoenzyme was reactivated when incubated with FAD but not FMN. As other lipoamide dehydrogenases, the yeast enzyme was found to possess diaphorase activity catalysing the oxidation of NADH with various artificial electron acceptors. Km values were 0.48 mM for dihydrolipoamide and 0.15 mM for NAD. NADH was a competitive inhibitor with respect to NAD (Ki 31 microM). The native enzyme (Mr 117000) was composed of two apparently identical subunits (Mr 56000), each containing 0.96 FAD residues and one cystine bridge. The amino acid composition differed from bacterial and mammalian lipoamide dehydrogenases with respect to the content of Asx, Glx, Gly, Val, and Cys. The lipoamide dehydrogenases of baker's and brewer's yeast were immunologically identical but no cross-reaction with mammalian lipoamide dehydrogenases was found.  相似文献   

19.
The dissociation of pyruvate oxidase (PO) caused by pressure up to 220 MPa at various conditions was explored by measuring the intrinsic fluorescence spectra and polarization. At 5 degrees C and pH 7.6 the standard volume change (deltaV0) and free energy upon dissociation of the enzyme is -220 ml/mol and 29.83 kCal/mol, respectively. It was found that FAD was irreversibly removed during the pressure-dissociation of the enzyme. A much smaller standard volume change (-153 ml/mol) and lower free energy (24.92 kCal/mol) of apo-pyruvate oxidase (apo-PO) compared with the native enzyme indicated that FAD played very important role in stabilizing the enzyme and significantly influenced the standard volume change. The substrate pyruvic acid can significantly stabilize the enzyme against pressure in spite the standard volume for the enzyme in this case has a big increase relative to the native enzyme. The comparison of the intrinsic fluorescence of the native and the activated enzyme obtained by limited proteolysis indicated that the physical separation of alpha-peptide from the enzyme only occurred when the subunits were dissociated from each other under pressure.  相似文献   

20.
Lipoamide dehydrogenase (NADH:lipoamide oxidoreductase EC 1.6.4.3) has been isolated from Ascaris suum muscle mitochondria. This activity has been purified to apparent homogeneity from both the pyruvate dehydrogenase complex and from 150,000g mitochondrial supernatants which were devoid of pyruvate dehydrogenase complex activity. The enzymes from both sources exhibited similar kinetic, catalytic, and regulatory properties and appear to be identical as judged by polyacrylamide gel electrophoresis. The native enzyme acts as a dimer, containing 2 mol of FAD, and has a subunit molecular weight of 54,000, as estimated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and gel chromatography. The enzyme also possesses substantial NADH:NAD+ transhydrogenase activity. Heat denaturation and differential solubilization experiments imply that the transhydrogenase activity previously reported is, in fact, associated with the lipoamide dehydrogenase moiety of the Ascaris pyruvate dehydrogenase complex. Whether or not this activity functions physiologically in hydride ion translocation, as previously suggested, remains to be demonstrated.  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