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.
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   

2.
The crystal structure of lipoamide dehydrogenase from Azotobacter vinelandii has been determined by a combination of molecular replacement and isomorphous replacement techniques yielding eventually a good-quality 2.8 A electron density map. Initially, the structure determination was attempted by molecular replacement procedures alone using a model of human glutathione reductase, which has 26% sequence identity with this bacterial dehydrogenase. The rotation function yielded the correct orientation of the model structure both when the glutathione reductase dimer and monomer were used as starting model. The translation function could not be solved, however. Consequently, data for two heavy-atom derivatives were collected using the Hamburg synchotron facilities. The derivatives had several sites in common, which was presumably a major reason why the electron density map obtained by isomorphous information alone was of poor quality. Application of solvent flattening procedures cleaned up the map considerably, however, showing clearly the outline of the lipoamide dehydrogenase dimer, which has a molecular weight of 100,000. Application of the "phased translation function", which combines the phase information of both isomorphous and molecular replacement, led to an unambiguous determination of the position of the model structure in the lipoamide dehydrogenase unit cell. The non-crystallographic 2-fold axis of the dimer was optimized by several cycles of constrained-restrained least-squares refinement and subsequently used for phase improvement by 2-fold density averaging. After ten cycles at 3.5 A, the resolution was gradually extended to 2.8 A in another 140 cycles. The 2.8 A electron density distribution obtained in this manner was of much improved quality and allowed building of an atomic model of A. vinelandii lipoamide dehydrogenase. It appears that in the orthorhombic crystals used each dimer is involved in contacts with eight surrounding dimers, leaving unexplained why the crystals are rather fragile. Contacts between subunits within one dimer, which are quite extensive, can be divided into two regions separated by a cavity. In one of the contact regions, the level of sequence identity with glutathione reductase is very low but it is quite high in the other. The folding of the polypeptide chain in each subunit is quite similar to that of glutathione reductase, as is the extended conformation of the co-enzyme FAD.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

3.
The structure of lipoamide dehydrogenase from Azotobacter vinelandii has been refined by the molecular dynamics technique to an R-factor of 19.8% at 2.2 A resolution. In the final model, the root-mean-square deviation from ideality is 0.02 A for bond lengths and 3.2 degrees for bond angles. The asymmetric unit comprises two subunits, each consisting of 466 amino acid residues and the prosthetic group FAD, plus 512 solvent molecules. The last ten amino acid residues of both chains are not visible in the electron density distribution and they are probably disordered. The operation required to superimpose the two chains forming the dimer is a rotation of exactly 180 degrees with no translation component. The final model shows the two independently refined subunits to be very similar, except for six loops located at the surface of the molecule. The structure of each subunit of the enzyme consists of four domains with the catalytic centre located at the subunit interface. The reactive disulphide bridge, 48-53, is oxidized with S gamma of Cys53 located 3.5 A away from carbon C-4a of the isoalloxazine ring. The side-chain of His450' points its N epsilon 2 towards S gamma of Cys48 and is hydrogen bonded to the carboxylate of Glu455'. The FAD is bound in an extended conformation and the isoalloxazine ring is not completely planar with an angle between the pteridine and the benzene ring of 7.3 degrees in the first subunit and of 12.1 degrees in the second one. The overall folding of lipoamide dehydrogenase is very similar to that of glutathione reductase. However, a comparison of the two enzymes, which have only 26% sequence identity, reveals significant conformational differences. These concern the tertiary as well as the quaternary structure of the two molecules. In each subunit of lipoamide dehydrogenase the NAD-binding domain and the interface domain appear to be differently oriented with respect to the FAD-binding domain by 7.1 degrees and 7.8 degrees, respectively. The interface domain contains, in addition, major changes in tertiary structure. Furthermore, the two subunits forming the dimer appear to be shifted with respect to each other by more than 4 A, when the lipoamide dehydrogenase dimer is compared with that of glutathione reductase. In spite of all these changes at the tertiary and quaternary level the active sites of the enzymes, which occur at the dimer interface, appear to be remarkably similar.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

4.
The interaction between lipoamide dehydrogenase (E3) and dihydrolipoyl transacetylase (E2p) from the pyruvate dehydrogenase complex was studied during the reconstitution of monomeric E3 apoenzymes from Azotobacter vinelandii and Pseudomonas fluorescens. The dimeric form of E3 is not only essential for catalysis but also for binding to the E2p core, because the apoenzymes as well as a monomeric holoenzyme from P. fluorescens, which can be stabilized as an intermediate at 0 degree C, do not bind to E2p. Lipoamide dehydrogenase from A. vinelandii contains a C-terminal extension of 15 amino acids with respect to glutathione reductase which is, in contrast to E3, presumably not part of a multienzyme complex. Furthermore, the last 10 amino acid residues of E3 are not visible in the electron density map of the crystal structure and are probably disordered. Therefore, the C-terminal tail of E3 might be an attractive candidate for a binding region. To probe this hypothesis, a set of deletions of this part was prepared by site-directed mutagenesis. Deletion of the last five amino acid residues did not result in significant changes. A further deletion of four amino acid residues resulted in a decrease of lipoamide activity to 5% of wild type, but the binding to E2p was unaffected. Therefore it is concluded that the C-terminus is not directly involved in binding to the E2p core. Deletion of the last 14 amino acids produced an enzyme with a high tendency to dissociate (Kd approximately 2.5 microM). This mutant binds only weakly to E2p. The diaphorase activity was still high. This indicates, together with the decreased Km for NADH, that the structure of the monomer is not appreciably changed by the mutation. Rather the orientation of the monomers with respect to each other is changed. It can be concluded that the binding region of E3 for E2p is constituted from structural parts of both monomers and binding occurs only when dimerization is complete.  相似文献   

5.
The FAD-containing enzyme lipoamide dehydrogenase (EC 1.6.4.3. NADH: lipoamide oxidoreductase) of Azotobacter vinelandii has been crystallized from polyethylene glycol solutions. The space group is P2(1)2(1)2(1) with one dimer in the asymmetric unit. The cell dimensions are: a = 64.2, b = 83.8, c = 193 A. X-ray reflections extend to at least 2.2 A resolution.  相似文献   

6.
The gene encoding lipoamide dehydrogenase from Azotobacter vinelandii has been cloned in Escherichia coli. Fragments of 9-23 kb from Azotobacter vinelandii chromosomal DNA obtained by partial digestion with Sau3A were ligated into the BamHI site of plasmid pUC9. E. coli TG2 cells were transformed with the resulting recombinant plasmids. Screening for clones which produced A. vinelandii lipoamide dehydrogenase was performed with antibodies raised against the purified enzyme. A positive colony was found which produced complete chains of lipoamide dehydrogenase as concluded form SDS gel electrophoresis of the cell-free extract, stained for protein or used for Western blotting. After subcloning of the 14.7-kb insert of this plasmid the structural gene could be located on a 3.2-kb DNA fragment. The nucleotide sequence of this subcloned fragment (3134 bp) has been determined. The protein-coding sequence of the gene consists of 1434 bp (478 codons, including the AUG start codon and the UAA stop codon). It is preceded by an intracistronic region of 85 bp and the structural gene for succinyltransferase. A putative ribosome-binding site and promoter sequence are given. The derived amino acid composition is in excellent agreement with that previously published for the isolated enzyme. The predicted relative molecular mass is 50223, including the FAD. The overall homology with the E. coli enzyme is high with 40% conserved amino acid residues. From a comparison with the three-dimensional structure of the related enzyme glutathione reductase [Rice, D. W., Schultz, G. E. & Guest, J. R. (1984) J. Mol. Biol. 174, 483-496], it appears that essential residues in all four domains have been conserved. The enzyme is strongly expressed, although expression does not depend on the vector-encoded lacZ promoter. The cloned enzyme is, in all the respects tested, identical with the native enzyme.  相似文献   

7.
The conformational stability of holo-lipoamide and apo-lipoamide dehydrogenase from Azotobacter vinelandii was studied by thermoinactivation, unfolding and limited proteolysis. The oxidized holoenzyme is thermostable, showing a melting temperature, tm = 80 degrees C. The thermal stability of the holoenzyme drastically decreases upon reduction. Unlike the oxidized and lipoamide two-electron reduced enzyme species, the NADH four-electron reduced enzyme is highly sensitive to unfolding by urea. Loss of energy transfer from Trp199 to flavin reflects the unfolding of the oxidized holoenzyme by guanidine hydrochloride. Unfolding of the monomeric apoenzyme is a rapid fully reversible process, following a simple two-state mechanism. The oxidized and two-electron reduced holoenzyme are resistant to limited proteolysis by trypsin and endoproteinase Glu-C. Upon cleavage of the apoenzyme or four-electron reduced holoenzyme by both proteases, large peptide fragments (molecular mass greater than 40 kDa) are transiently produced. Sequence studies show that limited trypsinolysis of the NADH-reduced enzyme starts mainly at the C-terminus of Arg391. In the apoenzyme, limited proteolysis by endoproteinase Glu-C starts from the C-terminus at the carboxyl ends of Glu459 and/or Glu435. From crystallographic data it is deduced that the susceptible amino acid peptide bonds are situated near the subunit interface. Thus, these bonds are inaccessible to the proteases in the dimeric enzyme and become accessible after monomerization. It is concluded that reduction of lipoamide dehydrogenase to the four-electron reduced state(s) is accompanied by conformational changes promoting subunit dissociation.  相似文献   

8.
After limited proteolysis of the dihydrolipoyl transacetylase component (E2) of Azotobacter vinelandii pyruvate dehydrogenase complex (PDC), a C-terminal domain was obtained which retained the transacetylase active site and the quaternary structure of E2 but had lost the lipoyl-containing N-terminal part of the chain and the binding sites for the peripheral components, pyruvate dehydrogenase and lipoamide dehydrogenase. The C-terminus of this domain was determined by treatment with carboxypeptidase Y and shown to be identical with the C-terminus of E2. Together with the previously determined N-terminus and the known amino acid sequence of E2, a molecular mass of 27.5 kDa was calculated. From the molecular mass of the native catalytic domain, 530 kDa, and the symmetry of the cubic structures observed on electron micrographs, a 24-meric structure is concluded instead of the 32-meric structure proposed previously. From the effect of guanidine hydrochloride on the light-scattering of intact E2 it was concluded that dissociation occurs in a two-step reaction resulting in particles with an average mass 1/6 that of the original mass before the N----D transition takes place. Cross-linking experiments with the catalytic domain indicated that the multimeric E2 is built from tetramers and that the tetramers are arranged as a dimer of dimers. A model for the quaternary structure of E2 is given, in which it is assumed that the tetrameric E2 core of PDC is formed from each of the six morphological subunits located at the lateral face of the cube. Binding of peripheral components to a site that interferes with the cubic assembly causes dissociation, resulting in the unique small PDC of A. vinelandii.  相似文献   

9.
The lpd gene encoding lipoamide dehydrogenase (dihydrolipoamide dehydrogenase; EC 1.8.1.4) was isolated from a library of Pseudomonas fluorescens DNA cloned in Escherichia coli TG2 by use of serum raised against lipoamide dehydrogenase from Azotobacter vinelandii. Large amounts (up to 15% of total cellular protein) of the P. fluorescens lipoamide dehydrogenase were produced by the E. coli clone harbouring plasmid pCJB94 with the lipoamide dehydrogenase gene. The enzyme was purified to homogeneity by a three-step procedure. The gene was subcloned from plasmid pCJB94 and the complete nucleotide sequence of the subcloned fragment (3610 bp) was determined. The derived amino acid sequence of P. fluorescens lipoamide dehydrogenase showed 84% and 42% homology when compared to the amino acid sequences of lipoamide dehydrogenase from A. vinelandii and E. coli, respectively. The lpd gene of P. fluorescens is clustered in the genome with genes for the other components of the 2-oxoglutarate dehydrogenase complex.  相似文献   

10.
Among the three closely related enzymes, lipoamide dehydrogenase, mercuric reductase, and glutathione reductase only the latter is inhibited by 2,4,6-trinitrobenzenesulfonate (TNBS). On the other hand, all three enzymes exhibit high rates of TNBS-dependent NADPH oxidation. In the case of glutathione reductase and mercuric reductase this TNBS-dependent activity displays substrate inhibition by excess of NADPH and is strongly stimulated by NADP+. The stimulation is especially pronounced with mercuric reductase, 25-fold under some conditions. Neither substrate inhibition nor stimulation by NAD+ is observed with lipoamide dehydrogenase.  相似文献   

11.
1. Glutathione reductase and lipoamide dehydrogenase are structurally and mechanistically related flavoenzymes catalyzing various one and two electron transfer reactions between NAD(P)H and substrates with different structures. 2. The two enzymes differ in their coenzyme and functional specificities. Lipoamide dehydrogenase shows higher coenzyme preference while glutathione reductase displays greater functional specificity. 3. Binding preference of the two flavoenzymes for nicotinamide coenzymes is demonstrated by 31P-NMR spectroscopy. 4. The presence of arginines in glutathione reductase which is inactivated by phenyl glyoxal, is likely to be responsible for the NADPH-activity of glutathione reductase. 5. The substrate binding sites of the two enzymes are similar, though their functional details differ. 6. The active-site histidine of glutathione reductase functions primarily as the proton donor during catalysis. While the active-site histidine of lipoamide dehydrogenase stabilizes the thiolate anion intermediate and relays a proton in the catalytic process.  相似文献   

12.
Time-resolved polarized fluorescence spectroscopy has been applied to the bound FAD in the structurally related flavoproteins lipoamide dehydrogenase from Azotobacter vinelandii (LipDH-AV) and glutathione reductase (GR) from human erythrocytes. The fluorescence parameters as obtained from the maximum entropy analysis differ considerably in both enzymes, reflecting the unique properties of the flavin microenvironment. Three conformational substates are revealed in LipDH-AV and five in GR. Almost 90% of the population of GR molecules has a fluorescence lifetime in the order of 30 ps which originates from efficient exciplex formation with Tyr197. Equilibrium fluctuations between conformational substates are observed for LipDH-AV on a nanosecond time scale in the temperature range 277-313 K. Interconversion between conformational substates in GR is slow, indicating that large activation barriers exist between the states. In agreement with these results, a model is postulated which ascribes a role in catalysis to equilibrium fluctuations between conformational substates in GR and LipDH-AV. From time-resolved fluorescence anisotropy as a function of temperature, distinction can be made between flavin reorientational motion and interflavin energy transfer. In both proteins intersubunit energy transfer between the prosthetic groups is observed. Furthermore, it is revealed that only the flavin in glutathione reductase exhibits rapid restricted reorientational motion. Geometric information concerning the relative orientation and distance of the flavins can be extracted from the parameters describing the energy-transfer process. The obtained spatial arrangement of the flavins is in excellent agreement with crystallographic data.  相似文献   

13.
The three-dimensional structure of human class pi glutathione S-transferase from placenta (hGSTP1-1), a homodimeric enzyme, has been solved by Patterson search methods and refined at 2.8 A resolution to a final crystallographic R-factor of 19.6% (8.0 to 2.8 A resolution). Subunit folding topology, subunit overall structure and subunit association closely resembles the structure of porcine class pi glutathione S-transferase. The binding site of a competitive inhibitor, S-hexylglutathione, is analyzed and the locations of the binding regions for glutathione (G-site) and electrophilic substrates (H-site) are determined. The specific interactions between protein and the inhibitor's glutathione peptide are the same as those observed between glutathione sulfonate and the porcine isozyme. The H-site is located adjacent to the G-site, with the hexyl moiety lying above a segment (residues 8 to 10) connecting strand beta 1 and helix alpha A where it is in hydrophobic contact with Tyr7, Phe8, Val10, Val35 and Tyr106. Catalytic models are discussed on the basis of the molecular structure.  相似文献   

14.
The pyruvate dehydrogenase multienzyme complex (PDHC) catalyses the oxidative decarboxylation of pyruvate and the subsequent acetylation of coenzyme A to acetyl-CoA. Previously, limited proteolysis experiments indicated that the N-terminal region of the homodimeric pyruvate dehydrogenase (E1p) from Azotobacter vinelandii could be involved in the binding of E1p to the core protein (E2p) [Hengeveld, A. F., Westphal, A. H. & de Kok, A. (1997) Eur J. Biochem. 250, 260-268]. To further investigate this hypothesis N-terminal deletion mutants of the E1p component of Azotobacter vinelandii pyruvate dehydrogenase complex were constructed and characterized. Up to nine N-terminal amino acids could be removed from E1p without effecting the properties of the enzyme. Truncation of up to 48 amino acids did not effect the expression or folding abilities of the enzyme, but the truncated enzymes could no longer interact with E2p. The 48 amino acid deletion mutant (E1pdelta48) is catalytically fully functional: it has a Vmax value identical to that of wild-type E1p, it can reductively acetylate the lipoamide group attached to the lipoyl domain of the core enzyme (E2p) and it forms a dimeric molecule. In contrast, the S0.5 for pyruvate is decreased. A heterodimer was constructed containing one subunit of wild-type E1p and one subunit of E1pdelta48. From the observation that the heterodimer was not able to bind to E2p, it is concluded that both N-terminal domains are needed for the binding of E1p to E2p. The interactions are thought to be mainly of an electrostatic nature involving negatively charged residues on the N-terminal domains of E1p and previously identified positively charged residues on the binding and catalytic domain of E2p.  相似文献   

15.
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   

16.
The gene encoding the dihydrolipoyltransacetylase component (E2) of the pyruvate dehydrogenase complex from Azotobacter vinelandii has been cloned in Escherichia coli. A plasmid containing a 2.8-kbp insert of A. vinelandii chromosomal DNA was obtained and its nucleotide sequence determined. The gene comprises 1911 base pairs, 637 codons excluding the initiation codon GUG and stop codon UGA. It is preceded by the gene encoding the pyruvate dehydrogenase component (E1) of pyruvate dehydrogenase complex and by an intercistronic region of 11 base pairs containing a good ribosome binding site. The gene is followed downstream by a strong terminating sequence. The relative molecular mass (64913), amino acid composition and N-terminal sequence are in good agreement with information obtained from studies on the purified enzyme. Approximately the first half of the gene codes for the lipoyl domain. Three very homologous sequences are present, which are translated in three almost identical units, alternated with non-homologous regions which are very rich in alanyl and prolyl residues. The N-terminus of the catalytic domain is sited at residue 381. Between the lipoyl domain and the catalytic domain, a region of about 50 residues is found containing many charged amino acid residues. This region is characterized as a hinge region and is involved in the binding of the pyruvate dehydrogenase and lipoamide dehydrogenase components. The homology with the dihydrolipoyltransacetylase from E. coli is high: 50% amino acid residues are identical.  相似文献   

17.
Fifty years after a role of vanadium in biological fixation was proposed, it was shown that in addition to their well-characterized molybdendum nitrogenases, Azotobacter chroococcum and Azotobacter vinelandii both have a genetically distinct nitrogenase system in which the conventional molybdoprotein is replaced by a vanadoprotein. Both Mo-nitrogenases and V-nitrogenases have similar requirements for activity: MgATP, a low potential reductant and the absence of oxygen. The genes encoding the V-nitrogenase are expressed only under conditions of Mo-deficiency. V-Nitrogenase of A.chroococcum is made up of a tetrameric VFe protein (Mr 210,000) with an alpha 2 beta 2 structure containing two V atoms, 23 Fe atoms and 20 acid-labile sulphide atoms per tetramer, and a dimeric Fe protein (Mr 64,000) with a gamma 2 structure containing four Fe atoms and four acid-labile sulphide atoms per dimer. Vanadium K-edge X-ray absorption spectroscopy indicates that V in the VFe protein, like Mo in MoFe protein, has S, Fe and possibly O as nearest neighbours. A vanadium- and iron-containing cofactor (FeVaco) can be extracted from the VFe protein and will restore C2H2 reductase, but no nitrogenase activity, to the inactive MoFe protein accumulated by mutants unable to synthesize the molybdenum- and iron-containing co-factor of Mo-nitrogenase. The products of C2H2 reduction by the hybrid protein (C2H6 as well as C2H4) are a characteristic of the VFe protein and provide evidence that FeVaco is, or forms part of the active site of V-nitrogenase.  相似文献   

18.
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   

19.
The pyruvate-dehydrogenase complex from Azotobacter vinelandii.   总被引:1,自引:0,他引:1  
The pyruvate dehydrogenase complex from Axotobacter vinelandii was isolated in a five-step procedure. The minimum molecular weight of the pure complex is 600,000, as based on an FAD content of 1.6 nmol-mg protein-1. The molecular weight is 1.0-1.2 X 10(6), indicating 1 mole of lipoamide dehydrogenase dimer per complex molecule. Sodium dodecylsulphate gel electrophoretical patterns show that apart from pyruvate dehydrogenase (Mr89,000) and lipoamide dehydrogenase (Mrmonomer 56,000) two active transacetylase isoenzymes are present with molecular weight on the gel 82,000 and 59,000 but probably actually lower. The pure complex has a specific activity of the pyruvate-NAD+ reductase (overall) reaction of 10 units-mg protein-1 at 25 degrees C. The partial reactions have the following specific activities in units-mg protein-1 at 25 degrees C under standard conditions: pyruvate-K3Fe(CN)6 reductase 0.14, transacetylase 3.6 and lipoamide dehydrogenase 2.9. The properties of this complex are compared with those from other sources. NADPH reduced the FAD of lipoamide dehydrogenase as well in the complex as in the free form. NADP+ cannot be used as electron acceptor. Under aerobic conditios pyruvate oxidase reaction, dependent on Mg2+ and thiamine pyrophosphate, converts pyruvate into CO2 and acetate; V is 0.2 mumol 02-min-1-mg-1, Km(pyruvate)0.3 mM. The kinetics of this reaction shows a linear 1/velocity-1/[pyruvate] plot. K3Fe(CN)6 competes with the oxidase reaction. The oxidase activity is stimulated by AMP and sulphate and is inhibited by acetyl-CoA. The partially purified enzyme contains considerable phosphotransacetylase activity. The pure complex does not contain this activity. The physiological significance of this activity is discussed.  相似文献   

20.
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   

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