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.
Three electron-transferring flavoproteins were purified to homogeneity from anaerobic, amino acid-utilizing bacteria (bacterium W6, Clostridium sporogenes, and Clostridium sticklandii), characterized, and compared with the dihydrolipoamide dehydrogenase of Eubacterium acidaminophilum. All the proteins were found to be dimers consisting of two identical subunits with a subunit Mr of about 35,000 and to contain about 1 mol of flavin adenine dinucleotide per subunit. Spectra of the oxidized proteins exhibited characteristic absorption of flavoproteins, and the reduced proteins showed an A580 indicating a neutral semiquinone. Many artificial electron acceptors, including methyl viologen, could be used with NADPH as the electron donor but not with NADH. Unlike the enzyme of E. acidaminophilum, which exhibited by itself a dihydrolipoamide dehydrogenase activity (W. Freudenberg, D. Dietrichs, H. Lebertz, and J. R. Andreesen, J. Bacteriol. 171:1346-1354, 1989), the electron-transferring flavoprotein purified from bacterium W6 reacted with lipoamide only under certain assay conditions, whereas the proteins of C. sporogenes and C. sticklandii exhibited no dihydrolipoamide dehydrogenase activity. The three homogeneous electron-transferring flavoproteins were very similar in their structural and biochemical properties to the dihydrolipoamide dehydrogenase of E. acidaminophilum and exhibited cross-reaction with antibodies raised against the latter enzyme. N-terminal sequence analysis demonstrated a high degree of homology between the dihydrolipoamide dehydrogenase of E. acidaminophilum and the electron-transferring flavoprotein of C. sporogenes to the thioredoxin reductase of Escherichia coli. Unlike these proteins, the dihydrolipoamide dehydrogenases purified from the anaerobic, glycine-utilizing bacteria Peptostreptococcus glycinophilus, Clostridium cylindrosporum, and C. sporogenes exhibited a high homology to dihydrolipoamide dehydrogenases known from other organisms.  相似文献   

2.
The glycine-utilizing bacterium Clostridium litoralis contained two enzyme systems for oxidizing dihydrolipoamide. The first one was found to be a genuine dihydrolipoamide dehydrogenase, present only in low amounts. This enzyme had the typical dimeric structure with a subunit molecular mass of about 53 kDa; however, it reacted with both NADP (Km 0.11 mM) and NAD (Km 0.5 mM). The reduction of pyridine nucleotides by dihydrolipoamide was the strongly preferred reaction. A second dihydrolipoamide-oxidizing enzyme system consisted of the interaction of two proteins, the previously described NADP(H)-dependent electron-transferring flavoprotein (D. Dietrichs, M. Meyer, B. Schmidt, and J. R. Andreesen, J. Bacteriol. 172:2088-2095, 1990) and a thioredoxin. This enzyme system was responsible for most of the dihydrolipoamide dehydrogenase activity in cell extracts. The thioredoxin did not bind to DEAE, was heat stable, and had a molecular mass of about 15 kDa. N-terminal amino acid analysis of the first 38 amino acid residues resulted in 38% homology to Escherichia coli thioredoxin and about 76% homology to a corresponding protein isolated from the physiologically close related Eubacterium acidaminophilum. The protein of the latter organism had a molecular mass of about 14 kDa and stimulated the low dihydrolipoamide dehydrogenase activity of the corresponding flavoprotein. By this interaction with NADPH-dependent flavoproteins, a new assay system for thioredoxin was established. A function of thioredoxin in glycine metabolism of some anaerobic bacteria is proposed.  相似文献   

3.
Five monoclonal antibodies against bacterial spores of Bacillus cereus T and Clostridium sporogenes PA3679 were developed. Two antibodies (B48 and B183) were selected for their reactivity with B. cereus T spores, two (C33 and C225) were selected for their reactivity with C. sporogenes spores, and one (D89) was selected for its reactivity with both B. cereus and C sporogenes spores. The isotypes of the antibodies were determined to be immunoglobulin G2a (IgG2a) (B48), IgG1 (B183), and IgM (C33, C225, and D89). The antibodies reacted with spores of B. cereus T, Bacillus subtilis subsp. globigii, Bacillus megaterium, Bacillus stearothermophilus, C. sporogenes, Clostridium perfringens, and Desulfotomaculum nigrificans. Antibody D89 also reacted with vegetative cells of B. cereus and C. sporogenes. Analysis of B. cereus spore extracts showed that two of the antigens with which the anti-Bacillus antibodies reacted had molecular masses of 76 kDa and approximately 250 kDa. Immunocytochemical localization indicated that antigens with which B48, B183, and D89 react are on the exosporium of the B. cereus T spore. Antibody D89 reacted with the exosporium and outer cortex of C. sporogenes spores in immunocytochemical localization studies but did not react with extracts of C. sporogenes or B. cereus spores in Western blotting. Some C. sporogenes antigens were not stable during long-term storage at -20 degrees C. Antibodies B48, B183, and D89 should prove to be useful tools for developing immunological methods for the detection of bacterial spores.  相似文献   

4.
Thermoproteus tenax possesses two different glyceraldehyde-3-phosphate dehydrogenases, one specific for NADP+ and the other for NAD+. NADP(H) inhibits the NAD+-specific enzyme competetively with respect to NAD+ whereas NAD(H) virtually does not interact with the NADP+-specific enzyme. Both enzymes represent homomeric tetramers with subunit molecular masses of 39 kDa (NADP+-specific enzyme) and 49 kDa (NAD+-specific enzyme), respectively. The NADP+-specific enzyme shows significant homology to the known glyceraldehyde-3-phosphate dehydrogenases from eubacteria and eukaryotes as indicated by partial sequencing. The enzymes are thermostable, the NADP+-specific enzyme with a half-life of 35 min at 100 degrees C, the NAD+-specific enzyme with a half-line of greater than or equal to 20 min at 100 degrees C, depending on the protein concentration. Both enzymes show conformational and functional changes at 60-70 degrees C.  相似文献   

5.
Acetaldehyde and butyraldehyde are substrates for alcohol dehydrogenase in the production of ethanol and 1-butanol by solvent-producing clostridia. A coenzyme A (CoA)-acylating aldehyde dehydrogenase (ALDH), which also converts acyl-CoA to aldehyde and CoA, has been purified under anaerobic conditions from Clostridium beijerinckii NRRL B592. The ALDH showed a native molecular weight (Mr) of 100,000 and a subunit Mr of 55,000, suggesting that ALDH is dimeric. Purified ALDH contained no alcohol dehydrogenase activity. Activities measured with acetaldehyde and butyraldehyde as alternative substrates were copurified, indicating that the same ALDH can catalyze the formation of both aldehydes for ethanol and butanol production. Based on the Km and Vmax values for acetyl-CoA and butyryl-CoA, ALDH was more effective for the production of butyraldehyde than for acetaldehyde. ALDH could use either NAD(H) or NADP(H) as the coenzyme, but the Km for NAD(H) was much lower than that for NADP(H). Kinetic data suggest a ping-pong mechanism for the reaction. ALDH was more stable in Tris buffer than in phosphate buffer. The apparent optimum pH was between 6.5 and 7 for the forward reaction (the physiological direction; aldehyde forming), and it was 9.5 or higher for the reverse reaction (acyl-CoA forming). The ratio of NAD(H)/NADP(H)-linked activities increased with decreasing pH. ALDH was O2 sensitive, but it could be protected against O2 inactivation by dithiothreitol. The O2-inactivated enzyme could be reactivated by incubating the enzyme with CoA in the presence or absence of dithiothreitol prior to assay.  相似文献   

6.
Acetaldehyde and butyraldehyde are substrates for alcohol dehydrogenase in the production of ethanol and 1-butanol by solvent-producing clostridia. A coenzyme A (CoA)-acylating aldehyde dehydrogenase (ALDH), which also converts acyl-CoA to aldehyde and CoA, has been purified under anaerobic conditions from Clostridium beijerinckii NRRL B592. The ALDH showed a native molecular weight (Mr) of 100,000 and a subunit Mr of 55,000, suggesting that ALDH is dimeric. Purified ALDH contained no alcohol dehydrogenase activity. Activities measured with acetaldehyde and butyraldehyde as alternative substrates were copurified, indicating that the same ALDH can catalyze the formation of both aldehydes for ethanol and butanol production. Based on the Km and Vmax values for acetyl-CoA and butyryl-CoA, ALDH was more effective for the production of butyraldehyde than for acetaldehyde. ALDH could use either NAD(H) or NADP(H) as the coenzyme, but the Km for NAD(H) was much lower than that for NADP(H). Kinetic data suggest a ping-pong mechanism for the reaction. ALDH was more stable in Tris buffer than in phosphate buffer. The apparent optimum pH was between 6.5 and 7 for the forward reaction (the physiological direction; aldehyde forming), and it was 9.5 or higher for the reverse reaction (acyl-CoA forming). The ratio of NAD(H)/NADP(H)-linked activities increased with decreasing pH. ALDH was O2 sensitive, but it could be protected against O2 inactivation by dithiothreitol. The O2-inactivated enzyme could be reactivated by incubating the enzyme with CoA in the presence or absence of dithiothreitol prior to assay.  相似文献   

7.
Bisubstrate inhibitors, obtained by covalently linking 2-oxoglutarate with NAD+ and NADP+, were synthesized and tested for their ability to inhibit NAD+- and NADP+-dependent isocitrate dehydrogenases from pig heart mitochondria. The NADP+-dependent enzyme was specifically inhibited by the NADP oxoglutarate adduct and not by the NAD adduct. The NADP adduct was competitive with both coenzyme and substrate, isocitrate. In contrast, the NAD+-dependent enzyme was inhibited by both adducts. NAD oxoglutarate is competitive with both NAD+ and isocitrate while the NADP adduct is competitive with isocitrate but not with NAD+. Nevertheless conditions could be set up so that use of these inhibitors would be feasible for a metabolic study.  相似文献   

8.
Initial-rate studies were made of the oxidation of L-glutamate by NAD+ and NADP+ catalysed by highly purified preparations of dogfish liver glutamate dehydrogenase. With NAD+ as coenzyme the kinetics show the same features of coenzyme activation as seen with the bovine liver enzyme [Engel & Dalziel (1969) Biochem. J. 115, 621--631]. With NADP+ as coenzyme, initial rates are much slower than with NAD+, and Lineweaver--Burk plots are linear over extended ranges of substrate and coenzyme concentration. Stopped-flow studies with NADP+ as coenzyme give no evidence for the accumulation of significant concentrations of NADPH-containing complexes with the enzyme in the steady state. Protection studies against inactivation by pyridoxal 5'-phosphate indicate that NAD+ and NADP+ give the same degree of protection in the presence of sodium glutarate. The results are used to deduce information about the mechanism of glutamate oxidation by the enzyme. Initial-rate studies of the reductive amination of 2-oxoglutarate by NADH and NADPH catalysed by dogfish liver glutamate dehydrogenase showed that the kinetic features of the reaction are very similar with both coenzymes, but reactions with NADH are much faster. The data show that a number of possible mechanisms for the reaction may be discarded, including the compulsory mechanism (previously proposed for the enzyme) in which the sequence of binding is NAD(P)H, NH4+ and 2-oxoglutarate. The kinetic data suggest either a rapid-equilibrium random mechanism or the compulsory mechanism with the binding sequence NH4+, NAD(P)H, 2-oxoglutarate. However, binding studies and protection studies indicate that coenzyme and 2-oxoglutarate do bind to the free enzyme.  相似文献   

9.
The NADP analog and NAD diphosphate were tested for the coenzyme or inhibiting activity toward various dehydrogenases. These NAD derivatives showed little or no activity of as coenzymes for most of dehydrogenases tested. Only glyceraldehyde 3-phosphate dehydrogenase reduced the NADP analog under the high concentration of enzyme system. These NAD derivatives showed no inhibiting effect toward the reduction or oxidation of pyridine coenzymes.  相似文献   

10.
The yeast Candida parapsilosis possesses two routes of electron transfer from exogenous NAD(P)H to oxygen. Electrons are transferred either to the classical cytochrome pathway at the level of ubiquinone through an NAD(P)H dehydrogenase, or to an alternative pathway at the level of cytochrome c through another NAD(P)H dehydrogenase which is insensitive to antimycin A. Analyses of mitoplasts obtained by digitonin/osmotic shock treatment of mitochondria purified on a sucrose gradient indicated that the NADH and NADPH dehydrogenases serving the alternative route were located on the mitochondrial inner membrane. The dehydrogenases could be differentiated by their pH optima and their sensitivity to amytal, butanedione and mersalyl. No transhydrogenase activity occurred between the dehydrogenases, although NADH oxidation was inhibited by NADP+ and butanedione. Studies of the effect of NADP+ on NADH oxidation showed that the NADH:ubiquinone oxidoreductase had Michaelis-Menten kinetics and was inhibited by NADP+, whereas the alternative NADH dehydrogenase had allosteric properties (NADH is a negative effector and is displaced from its regulatory site by NAD+ or NADP+).  相似文献   

11.
Two of the three metabolic subtypes of species utilizing C4-pathway photosynthesis are defined by high activities of either NADP malic enzyme (NADP malic enzyme type) or a coenzyme A (CoA)- and acetyl-CoA-activated NAD malic enzyme (NAD malic enzyme type). These enzymes function to decarboxylate malate as an integral part of the photosynthetic process. Leaves of NADP malic enzyme-type species also contain significant NAD-dependent malic enzyme activity. The purpose of the present study was to examine the nature and photosynthetic role of this activity. With Zea mays, this NAD-dependent activity was found to vary widely in fresh leaf extracts. Incubating extracts at 25 °C resulted in a disproportionate increase in NAD activity so that the final ratio of NADP to NAD activity was always about 5. Strong evidence was provided that the NADP and NAD malic enzyme activities in Z. mays extracts were catalyzed by the same enzyme. These activities remained associated during purification and were coincident after polyacrylamide gel electrophoresis. The pH optimum for NAD-dependent activity was about 7.1, compared with 8.3 for NADP malic enzyme activity. Other properties of the NAD-dependent activity are described, a particularly notable feature being the inhibition of this activity by less than 1 μm NADP and NADPH. Evidence is provided that the NADP malic enzyme of several other NADP malic enzyme-type C4 species also has associated activity toward NAD. We concluded that the NAD-dependent malic enzyme activity would have no significant function in photosynthesis.  相似文献   

12.
Various kinetic approaches were carried out to investigate kinetic attributes for the dual coenzyme activities of mitochondrial aldehyde dehydrogenase from rat liver. The enzyme catalyses NAD(+)- and NADP(+)-dependent oxidations of ethanal by an ordered bi-bi mechanism with NAD(P)+ as the first reactant bound and NAD(P)H as the last product released. The two coenzymes presumably interact with the kinetically identical site. NAD+ forms the dynamic binary complex with the enzyme, while the enzyme-NAD(P)H complex formation is associated with conformation change(s). A stopped-flow burst of NAD(P)H formation, followed by a slower steady-state turnover, suggests that either the deacylation or the release of NAD(P)H is rate limiting. Although NADP+ is reduced by a faster burst rate, NAD+ is slightly favored as the coenzyme by virtue of its marginally faster turnover rate.  相似文献   

13.
Carbonyl compounds such as alpha-ketoglutarate, pyruvate, oxaloacetate, butyraldehyde, acetaldehyde or acetone react with NAD or NADP to give adducts. Binding studies of adducts to dehydrogenases are performed by means of ultraviolet differential spectroscopy, circular dichroism and spectrofluorimetry. The dehydrogenases show a high degree of binding specificity toward the adducts which contain their specific oxidized substrate and their specific coenzyme. The high selectivity of the dehydrogenases for adducts is evidenced by binding studies of NAD(P)-pyruvate and NAD(P)-alpha-ketoglutarate adducts on glutamate dehydrogenase at pH 7.6 and 8.9. Evidence is presented showing that adducts bind to the active site of the enzymes.  相似文献   

14.
Two alcohol dehydrogenases (alcohol: NAD oxidoreductase, EC 1.1.1.1 and alcohol: NADP oxidoreductase, EC 1.1.1.2) were partially purified from extracts of strawberry seeds by conventional methods. Some of physical, chemical and kinetic properties of the enzymes are described. On the basis of gel filtration, the molecular weights were estimated to be approximately 78,000 for NAD-dependent enzyme and 82,000 for NADP-dependent enzyme. Thiol-reacting compounds inhibited both enzymes. NAD-dependent alcohol dehydrogenase reacted only with aliphatic alcohols and aldehydes, while aromatic and terpene alcohols and aldehydes were the better substrates for NADP-dependent alcohol dehydrogenase than aliphatic alcohols and aldehydes.  相似文献   

15.
Clostridial glutamate dehydrogenase mutants, designed to accommodate the 2'-phosphate of disfavoured NADPH, showed the expected large specificity shifts with NAD(P)H. Puzzlingly, similar assays with oxidized cofactors initially revealed little improvement with NADP(+) , although rates with NAD(+) were markedly diminished. This article reveals that the enzyme's discrimination in favour of NAD(+) and against NADP(+) had been greatly underestimated and has indeed been abated by a factor of >?16,000 by the mutagenesis. Initially, stopped-flow studies of the wild-type enzyme showed a burst increase of A(340) with NADP(+) but not NAD(+), with amplitude depending on the concentration of the coenzyme, rather than enzyme. Amplitude also varied with the commercial source of the NADP(+). FPLC, HPLC and mass spectrometry identified NAD(+) contamination ranging from 0.04 to 0.37% in different commercial samples. It is now clear that apparent rates of NADP(+) utilization mainly reflected the reduction of contaminating NAD(+), creating an entirely false view of the initial coenzyme specificity and also of the effects of mutagenesis. Purification of the NADP(+) eliminated the burst. With freshly purified NADP(+), the NAD(+) : NADP(+) activity ratio under standard conditions, previously estimated as 300 : 1, is 11,000. The catalytic efficiency ratio is even higher at 80,000. Retested with pure cofactor, mutants showed marked specificity shifts in the expected direction, for example, 16 200 fold change in catalytic efficiency ratio for the mutant F238S/P262S, confirming that the key structural determinants of specificity have been successfully identified. Of wider significance, these results underline that, without purification, even the best commercial coenzyme preparations are inadequate for such studies.  相似文献   

16.
A soluble NAD+-linked isocitrate dehydrogenase has been isolated from Crithidia fasciculata. The enzyme was purified 128-fold, almost to homogeneity, and was highly specific for NAD+ as the coenzyme. There is also a cytoplasmic NADP+-linked and a mitochondrial isocitrate dehydrogenase in the organism. Studies of the physical and kinetic properties of the soluble NAD+-isocitrate dehydrogenase from this organism showed that it resembled microbial NADP+-isocitrate dehydrogenases in general, all of which are cytoplasmic enzymes. The enzyme appeared not to be related to other NAD+-isocitrate dehydrogenases, which are found in the mitochondria of eukaryotic cells. The molecular weight of the soluble NAD+-isocitrate dehydrogenase was 105,000 which is within the range of the values for microbial NADP+-isocitrate dehydrogenases. Similar to the NADP+-isocitrate dehydrogenase in this organism, the enzyme was inhibited in a concerted manner by glyoxalate plus oxalacetate. Kinetic analysis revealed that Mn2+ was involved in the binding of isocitrate to the enzyme. Inhibition of the NAD+-linked isocitrate dehydrogenase by p-chloromercuribenzoate could be prevented by prior incubation of the enzyme with both Mn2+ and isocitrate; however, neither ion alone conferred protection. Free isocitrate, free Mn2+, and the Mn2+-isocitrate complex could all bind to the enzyme. Four different mechanisms with respect to the binding of isocitrate to the enzyme were tested. Of these, the formation of the active enzyme-Mn2+-isocitrate complex from (a) the random binding of Mn2+, isocitrate, and the Mn2+-isocitrate complex, or (b) the binding of Mn2+-isocitrate with free Mn2+ and isocitrate acting as dead-end competitors were both in agreement with these data.  相似文献   

17.
1. Skeletal muscle mitochondrial NAD(P)-dependent malic enzyme [EC 1.1.1. 39, L-malate:NAD+ oxidoreductase (decarboxylating)] from herring could use both coenzymes, NAD and NADP, in a similar manner. 2. The coenzyme preference of mitochondrial NAD(P)-dependent malic enzyme was probed using dual wavelength spectroscopy and pairing the natural coenzymes, NAD or NADP with their respective thionicotinamide analogues, s-NADP or s-NAD, that have absorbance maxima in reduced forms at 400 nm. 3. s-NAD and s-NADP were found to be good alternate substrates for NAD(P)-dependent malic enzyme, the apparent Km values for the thioderivatives were similar to those of the corresponding natural coenzymes. 4. ATP produced greater inhibition of the NAD or s-NAD linked reactions than of the NADP or s-NADP-linked reactions of skeletal muscle mitochondrial NAD(P)-dependent malic enzyme. 5. At 5 mM malate concentration and in the presence of 2 mM ATP the NADP-linked reaction is favoured and the activity ratios, V(s-NADP)/V(NAD) or V(NADP)/V(s-NAD), are 6 and 26, respectively.  相似文献   

18.
The amino acid sequence is reported for CNBr and tryptic peptide fragments of the NAD(+)-dependent glutamate dehydrogenase of Clostridium symbiosum. Together with the N-terminal sequence, these make up about 75% of the total sequence. The sequence shows extensive similarity with that of the NADP(+)-dependent glutamate dehydrogenase of Escherichia coli (52% identical residues out of the 332 compared) allowing confident placing of the peptide fragments within the overall sequence. This demonstrated sequence similarity with the E. coli enzyme, despite different coenzyme specificity, is much greater than the similarity (31% identities) between the GDH's of C. symbiosum and Peptostreptococcus asaccharolyticus, both NAD(+)-linked. The evolutionary implications are discussed. In the 'fingerprint' region of the nucleotide binding fold the sequence Gly X Gly X X Ala is found, rather than Gly X Gly X X Gly. The sequence found here has previously been associated with NADP+ specificity and its finding in a strictly NAD(+)-dependent enzyme requires closer examination of the function of this structural motif.  相似文献   

19.
Sharkey MA  Gori A  Capone M  Engel PC 《The FEBS journal》2012,279(17):3003-3009
Active-site mutants of glutamate dehydrogenase from Clostridium?symbiosum have been designed and constructed and the effects on coenzyme preference evaluated by detailed kinetic measurements. The triple mutant F238S/P262S/D263K shows complete reversal in coenzyme selectivity from NAD(H) to NADP(H) with retention of high levels of catalytic activity for the new coenzyme. For oxidized coenzymes, k(cat) /K(m) ratios of the wild-type and triple mutant enzyme indicate a shift in preference of approximately 1.6?×?10(7) -fold, from ~?80?000-fold in favour of NAD(+) to ~?200-fold in favour of NADP(+) . For reduced coenzymes the corresponding figure is 1.7?×?10(4) -fold, from ~?1000-fold in favour of NADH to ~?17-fold in favour of NADPH. A fourth mutation (N290G), previously identified as having a potential bearing on coenzyme specificity, did not engender any further shift in preference when incorporated into the triple mutant, despite having a significant effect when expressed as a single mutant.  相似文献   

20.
An isocitrate dehydrogenase able to function with either NADP or NAD as coenzyme was purified to homogeneity from cell-free extracts of the purple photosynthetic eubacterium Rhodomicrobium vannielii using a rapid two-step procedure involving dye-ligand affinity chromatography. The enzyme was obtained in 60% yield with specific activities of 23 U.mg protein-1 (NADP-linked reaction) and 18.5 U.mg protein-1 (NAD-linked reaction). The purified enzyme was monomeric and migrated with an approximate Mr of 75,000-80,000 on both SDS/PAGE and non-denaturing PAGE. Affinity constants (Km values) of 2.5 microM for NADP and 0.77 mM for NAD and values for kcat/Km of 981,200 min-1.mM-1 (NADP) and 2455 min-1.mM-1 (NAD) indicated a greater specificity for NADP compared to NAD. A number of metabolites were examined for possible differential regulatory effects on the NADP- and NAD-linked reactions, using a dual-wavelength assay. Oxaloacetate was found to be an effective inhibitor of both reactions and the enzyme was also sensitive to concerted inhibition by glyoxylate and oxaloacetate. The amino-acid composition and the identity of 39 residues at the N-terminus were determined and compared to other isocitrate dehydrogenases. The results suggested a relationship between the Rm. vannielii enzyme and the monomeric isocitrate dehydrogenase isoenzyme II from Vibrio ABE-1.  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