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.
Purified malate dehydrogenases from four species of non-sulphur purple phototrophic bacteria were examined for their heat-stability, amino acid composition and antigenic relationships. Malate dehydrogenase from Rhodospirillum rubrum, Rhodobacter capsulatus and Rhodomicrobium vannielii (which are all tetrameric proteins) had an unusually high glycine content, but the enzyme from Rhodocyclus purpureus (which is a dimer) did not. R. rubrum malate dehydrogenase was extremely heat-stable relative to the other enzymes, withstanding 65 degrees C for over 1 h with no loss of activity. By contrast, malate dehydrogenase from R. vannielii lost activity above 35 degrees C, and that from R. capsulatus above 40 degrees C. Amino acid compositional relatedness and immunological studies indicated that tetrameric phototrophic-bacterial malate dehydrogenases were highly related to one another, but only distantly related to the tetrameric enzyme from Bacillus. This suggests that, despite differences in their thermal properties, the tetrameric malate dehydrogenases of non-sulphur purple bacteria constitute a distinct biochemical class of this catalyst.  相似文献   

2.
Kinetic studies with the low-Km aldehyde reductase from ox brain.   总被引:1,自引:1,他引:0       下载免费PDF全文
Initial-rate studies of the low-Km aldehyde reductase-catalysed reduction of pyridine-3-aldehyde by NADPH gave families of parallel double-reciprocal plots, consistent with a double-displacement mechanism being obeyed. Studies on the variation of the initial velocity with the concentration of a mixture of the two substrates were also consistent with a double-displacement mechanism. In contrast, the initial-rate data indicated that a sequential mechanism was followed when NADH was used as the coenzyme. Product-inhibition studies, however, indicated that a compulsory-order mechanism was followed in which NADPH bound before pyridine-3-aldehyde with a ternary complex being formed and the release of pyrid-3-ylcarbinol before NADP+. The apparently parallel double-reciprocal plots obtained in the initial-rate studies with NADPH and pyridine-3-aldehyde were thus attributed to the apparent dissociation constant for the binary complex between the enzyme and coenzyme being finite but very low.  相似文献   

3.
NADH, (3-acetylpyridine)ADH, (3-thionicotinamide)ADH and (pyridine-3-aldehyde)ADH were labeled with deuterium by the B specific enzyme lipoyl dehydrogenase. The primary acid rearrangement products, α-O2′-6B cyclotetra-hydropyridine adenine dinucleotides, of the respective reduced coenzyme analogs were synthesized by incubation at pH 2–2.5. Analysis of the spin-spin coupling constants of the 220 MHz proton magnetic resonance spectra establishes that the structure of the ribose-tetrahydropyridine linkage of these analogs is the same as that of the primary acid product of NADH. All four specifically labeled dinucleotides possess the identical absolute configuration at the C4 position with respect to the deuterium label, establishing that the three coenzyme analogs are reduced by lipoyl dehydrogenase with the same absolute stereospecificity as NAD+. The primary acid reaction is described as a general method for the determination of the absolute configuration at the C4 position of specifically labeled β-1,3-substituted, 1,4 dihydropyridine nucleotides.  相似文献   

4.
Initial-rate measurements were made of the reduction of pyridine-3-aldehyde and p-carboxybenzaldehyde by NADPH catalyzed by pig liver aldehyde reductase I. The initial velocity analysis and product inhibition data suggest that aldehyde reductase I obeys a compulsory-order mechanism with pyridine-3-aldehyde as substrate but follows a partially random-order pathway with p-carboxybenzaldehyde. The partially random-order pathway would be operative only at high concentrations of p-carboxybenzaldehyde. In both cases, aldehydes and the corresponding alcohol substrates inhibit the enzyme at high concentration. Abortive ternary complexes are shown to be formed with pyridine-3-aldehyde and with p-carboxybenzaldehyde. Dissociation of the coenzyme from the abortive ternary complex seems only to be observed with p-carboxybenzaldehyde. This study suggests overall that an enzyme kinetic mechanism may be different, depending on whether specific interactions can occur between certain amino acid residue(s) of the protein active site and substrates. Finally, the mechanism of the inhibition of pyridine-3-aldehyde reduction by diacid derivatives is discussed.  相似文献   

5.
Li BH  Ma XF  Wu XD  Tian WX 《IUBMB life》2006,58(1):39-46
It was found that chlorogenic acid inhibited in vitro animal fatty acid synthase (FAS I) and the ss-ketoacyl-ACP reductase (FabG) from Escherichia coli in a concentration-dependent manner with respective IC50 of 94.8 and 88.1 microM. The results of Lineweaver-Burk plots indicated that chlorogenic acid inhibited competitively the binding of NADPH to FAS I, while left those of acetyl-CoA and malonyl-CoA unaffected. Further kinetic studies showed that chlorogenic acid blocked the activity of FAS I mainly by inhibiting the ss-ketoacyl reductase domain, which catalyzed the same reaction as that done by FabG in the fatty acid synthesis. The ss-ketoacyl reduction reactions accomplished by both FAS I and FabG required nucleotide cofactor, NADPH. Furthermore, the Lineweaver-Burk and Yonetani-Theorell analyses implicated that chlorogenic acid filled competitively in the binding-pocket of NADPH in the ss-ketoacyl reductase domain of FAS I. The similar results were also obtained from the inhibition of FabG by chlorogenic acid. As observed in these results, the inhibitions of FAS I and FabG by chlorogenic acid were highly related to the interference of the inhibitor with NADPH, which was possibly due to the similarity between chlorogenic acid and some portion of NADPH, maybe the section consisting of the two ribose groups.  相似文献   

6.
The L(+)-lactate dehydrogenase from Thermoanaerobacter ethanolicus wt was purified to a final specific activity of 598 mumol pyruvate reduced per min per mg of protein. The specific activity of the pure enzyme with L(+)-lactate was 0.79 units per mg of protein. The M(r) of the native enzyme was 134,000 containing a single subunit type of M(r) 33,500 indicating an apparent tetrameric structure. The L(+)-lactate dehydrogenase was activated by fructose 1,6-bisphosphate in a cooperative manner affecting Vmax and Km values. The activity of the enzyme was also effected by pH, pyruvate and NADH. The Km for NADH at pH 6.0 was 0.05 mM and the Vmax for pyruvate reduction at pH 6.0 was 1082 units per mg in the presence of 1 mM fructose 1,6-bisphosphate. The enzyme was inhibited by NADPH, displaying an uncompetitive pattern. This pattern indicated that NADPH was a negative modifier of the enzyme. The role of L(+)-lactate dehydrogenase in controlling the end products of fermentation is discussed.  相似文献   

7.
(3R,5R)-Clavulanic acid, a clinically used inhibitor of serine beta-lactamases, is produced by fermentation of Streptomyces clavuligerus. The early steps in clavulanic acid biosynthesis leading to the bicyclic beta-lactam intermediate (3S,5S)-clavaminic acid have been defined. However, the mechanism by which (3S,5S)-clavaminic acid is converted to the penultimate intermediate (3R,5R)-clavaldehyde is unclear. Disruption of orf15 or orf16, of the clavulanic acid biosynthesis gene cluster, blocks clavulanic acid production and leads to the accumulation of N-acetyl-glycyl-clavaminic acid and N-glycyl-clavaminic acid, suggesting that these compounds are intermediates in the pathway. Two alternative start codons have been proposed for orf17 to encode for two possible polypeptides, one of which has 92 N-terminal residues less then the other. The shorter version of orf17 was successfully expressed in Escherichia coli and purified as a monomeric protein. Sequence analyses predicting the ORF17 protein to be a member of the ATP-grasp fold superfamily were supported by soft ionization mass spectrometric analyses that demonstrated binding of ATP to the ORF17 protein. Semisynthetic clavaminic acid, prepared by in vitro reconstitution of the biosynthetic pathway from the synthetically accessible intermediate proclavaminic acid, was shown by mass spectrometric analyses to be converted to N-glycyl-clavaminic acid in the presence of ORF17, ATP, and glycine. Under the same conditions N-acetyl-glycine and clavaminic acid were not converted to N-acetyl-glycyl-clavaminic acid. The specificity of ORF17 as an N-glycyl-clavaminic acid synthetase, together with the reported accumulation of N-glycyl-clavaminic acid in orf15 and orf16 disruption mutants, suggested that N-glycyl-clavaminic acid is an intermediate in clavulanic acid biosynthesis.  相似文献   

8.
The in vivo flux through the oxidative branch of the pentose phosphate pathway (oxPPP) in Penicillium chrysogenum was determined during growth in glucose/ethanol carbon-limited chemostat cultures, at the same growth rate. Non-stationary 13C flux analysis was used to measure the oxPPP flux. A nearly constant oxPPP flux was found for all glucose/ethanol ratios studied. This indicates that the cytosolic NADPH supply is independent of the amount of assimilated ethanol. The cofactor assignment in the model of van Gulik et al. (Biotechnol Bioeng 68(6):602–618, 2000) was supported using the published genome annotation of P. chrysogenum. Metabolic flux analysis showed that NADPH requirements in the cytosol remain nearly the same in these experiments due to constant biomass growth. Based on the cytosolic NADPH balance, it is known that the cytosolic aldehyde dehydrogenase in P. chrysogenum is NAD +  dependent. Metabolic modeling shows that changing the NAD + -aldehyde dehydrogenase to NADP + -aldehyde dehydrogenase can increase the penicillin yield on substrate.  相似文献   

9.
Glycyrrhetic acid, derived from a main component of liquorice, was converted to 3-ketoglycyrrhetic acid reversibly by rat liver homogenates in the presence of NADPH or NADP+. Glycyrrhetic acid-oxidizing and 3-ketoglycyrrhetic acid-reducing activities were localized in microsomes among the subcellular fractions of rat liver. Glycyrrhetic acid-oxidizing activity and 3-ketoglycyrrhetic acid-reducing activities showed pH optima at 6.3 and 8.5, respectively, and required NADP+ or NAD+ and NADPH or NADH, respectively, indicating that these activities were due to glycyrrhetinate dehydrogenase. The dehydrogenase was not solubilized from the membranes by the treatment with 1 M NaCl or sonication, indicating that the enzyme is a membrane component. The dehydrogenase was solubilized with detergents such as Emalgen 913, Triton X-100 and sodium cholate, and then separated from 3β-hydroxysteroid dehydrogenase (5β-androstan-3β-ol-17-one-oxidizing activity) by butyl-Toyopearl 650 M column chromatography. Partially purified enzyme catalyzed the reversible reaction between glycyrrhetic acid and 3-ketoglycyrrhetic acid, but was inactive toward 3-epiglycyrrhetic acid and other steroids having the 3β-hydroxyl group. The enzyme required NADP+ and NADPH for the highest activities of oxidation and reduction, respectively, and NAD+ and NADH for considerable activities, similar to the results with microsomes. From these results the enzyme is defined as glycyrrhetinate dehydrogenase, being quite different from 3β-hydroxysteroid dehydrogenase of Ruminococcus sp. from human intestine, which is active for both glycyrrhetic acid and steroids having the 3β-hydroxyl group.  相似文献   

10.
ω-Hydroxyfatty acid dehydrogenase (ω-hydroxyfatty acid:NADP oxidoreductase) catalyzes the reaction ω-hydroxyfatty acid + NADP ? ω-oxofatty acid + NADPH +H+. In wound-healing potato tuber disks, the ω-oxofatty acid generated by this enzyme is further oxidized to the corresponding dicarboxylic acid by a separate enzyme, ω-oxofatty acid dehydrogenase. ω-Hydroxy acid dehydrogenase, but not ω-oxo acid dehydrogenase, was found to be induced by wounding potato tubers. ω-Hydroxy acid dehydrogenase has been purified 600-fold to near homogeneity from wound-healing potato tuber disks by a combination of gel filtration, anion-exchange, and hydroxylapatite chromatography followed by NADP-Sepharose affinity chromatography, in about 1% yield. The molecular weight and Stokes radius of this enzyme as determined by gel exclusion chromatography are 60,000 and 31 Å, respectively. Sodium dodecyl sulfate-gel electrophoresis gave a molecular weight of 31,000, indicating that the deydrogenase is a dimer with subunits of similar molecular weight. The pH optima for the reaction in the forward and reverse directions are 9.5 and 8.5, respectively, and V in the forward and reverse directions are 140 and 3200 nmol/min/mg, respectively. Apparent Km values for NADP, 16-hydroxyhexadecanoic acid, NADPH, and 16-oxohexadecanoic acid are 100, 20, 5, and 7 μm respectively. The equilibrium constant of the reaction at pH 9.5 and 30 °C is 1.4 × 10?9m. The enzyme preparation did not show any stereospecificity for hydride transfer from NADPH to 16-oxohexadecanoic acid.  相似文献   

11.
C6-aldehydes are synthesized via lipoxygenase/hydroperoxide lyase action on polyunsaturated fatty acid (PUFA) substrates in plant leaves. The source pools and subcellular location of the processes are unknown. A close relationship is found between the composition of PUFA and the composition of C6-aldehydes. In the current study, this relationship was tested using the Arabidopsis PUFA mutant lines act1, fad2, fad3, fad5, fad6, and fad7. The results indicate that C6-aldehyde formation is influenced by the alteration of C18 PUFA levels. Mutants act1 and fad5, which are deficient in C16 unsaturated fatty acids, had wild-type levels of C6-aldehyde production. Mutants deficient in the chloroplast hexadecenoic acid/oleic acid desaturase (fad6) or hexadecadienoic acid/linoleic acid desaturase (fad7) had altered C6-aldehyde formation in a pattern similar to the changes in the PUFA. Mutations that impair phosphatidylcholine desaturase activity, such as fad2 and fad3, however, resulted in increased E-2-hexenal formation. The enzymes involved in C6-aldehyde production were partially characterized, including measurement of pH optima. The differences in C6-aldehyde formation among the fatty acid mutants of Arabidopsis appeared not to result from alteration of lipoxygenase/hydroperoxide lyase pathway enzymes. Investigation of the fatty acid composition in leaf phospholipids, glycolipids, and neutral lipids and analysis of the fatty acid composition of chloroplast and extrachloroplast lipids indicate that chloroplasts and glycolipids of chloroplasts may be the source or major source of C6-aldehyde formation in Arabidopsis leaves.  相似文献   

12.
13.
The lignin structure and enzyme activities of normal and brown-midrib (BMR-6) mutant lines of Sorghum bicolor have been compared to identify the enzyme(s) involved in the reduction of the lignin content of the mutant. The results indicate that cinnamyl-alcohol dehydrogenase (CAD) and caffeic acid O-methyltransferase are depressed in the BMR-6 line, whereas the structural modifications correspond only to a reduction of CAD activity. Apparently, the change in the Sorghum lignin content, caused by depression of CAD activity, is accompanied by the incorporation of cinnamaldehydes into the core lignin.Abbreviations CAD cinnamyl-alcohol dehydrogenase - HPLC high-performance liquid chromatography - m/z mass number - OMT caffeic acid O-methyltransferase  相似文献   

14.
Aldose reductase (EC 1.1.1.21) and aldehyde reductase II (L-hexonate dehydrogenase, EC 1.1.1.2) have been purified to homogeneity from human erythrocytes by using ion-exchange chromatography, chromatofocusing, affinity chromatography, and Sephadex gel filtration. Both enzymes are monomeric, Mr 32,500, by the criteria of the Sephadex gel filtration and polyacrylamide slab gel electrophoresis under denaturing conditions. The isoelectric pH's for aldose reductase and aldehyde reductase II were determined to be 5.47 and 5.06, respectively. Substrate specificity studies showed that aldose reductase, besides catalyzing the reduction of various aldehydes such as propionaldehyde, pyridine-3-aldehyde and glyceraldehyde, utilizes aldo-sugars such as glucose and galactose. Aldehyde reductase II, however, did not use aldo-sugars as substrate. Aldose reductase activity is expressed with either NADH or NADPH as cofactors, whereas aldehyde reductase II can utilize only NADPH. The pH optima for aldose reductase and aldehyde reductase II are 6.2 and 7.0, respectively. Both enzymes are susceptible to the inhibition by p-hydroxymercuribenzoate and N-ethylmaleimide. They are also inhibited to varying degrees by aldose reductase inhibitors such as sorbinil, alrestatin, quercetrin, tetramethylene glutaric acid, and sodium phenobarbital. The presence of 0.4 M lithium sulfate in the assay mixture is essential for the full expression of aldose reductase activity whereas it completely inhibits aldehyde reductase II. Amino acid compositions and immunological studies further show that erythrocyte aldose reductase is similar to human and bovine lens aldose reductase, and that aldehyde reductase II is similar to human liver and brain aldehyde reductase II.  相似文献   

15.
Pig heart lactate dehydrogenase (EC 1.1.1.27; H4 isoenzyme) has been subjected to the action of beef spleen cathepsin D (EC 3.4.23.5) at pH 3.9. No proteolytic effect on the tetrameric form of the dehydrogenase could be detected. On the other hand, the acid proteinase appears to act on the dissociated monomers in such a manner that they are unable to reassociate into functional tetramers. This finding could be correlated with proteolytic modification of the monomers. The extent of proteolysis depends on the temperature and duration of the incubation.  相似文献   

16.
The effect of thioctic acid on functioning of the antioxidant glutathione-dependent system and activity of enzymes, supplying this system with NADPH, were studied under conditions of toxic hepatitis in rats. A decrease in the glutathione reductase and glutathione peroxidase activities towards normal levels was observed in animals with toxic hepatitis after administration of thioctic acid. Administration of thioctic acid under conditions of toxic hepatitis caused a decrease in the NADPH-dependent isocitrate dehydrogenase and glucose-6-phosphate dehydrogenase activities; this evidently reflects lowered requirements in the NADPH supply for operation of the glutathione-dependent system. Thus, these studies have shown that thioctic acid may serve as a factor regulating the extent of the oxidative stress development and the state of the glutathione antioxidant system.  相似文献   

17.
The general acyl-CoA dehydrogenase from pig liver mitochondria has been crystallized in a form suitable for detailed three-dimensional x-ray structure analysis. Crystals grown from Tris buffer and polyethylene glycol solution diffract to high resolution and have the space group C2221, a = 128.2, b = 136.1, and c = 106.3 A. A measured crystal density of 1.178 g/cm3 and a crystal volume/unit of molecular mass, Vm = 2.5 A3/dalton, suggest that the asymmetric unit contains two monomers of the tetrameric dehydrogenase molecule.  相似文献   

18.
Kinetic and physicochemical properties of hamster liver diacetyl reductase have been examined. The results of kinetic studies on the reduction of diacetyl and NADPH to acetoin and NADP+ suggest that the reaction follows an Ordered Bi Bi mechanism in which NADPH binds first before diacetyl. The enzyme is a tetrameric glycoprotein of single subunits of a molecular weight of 23,500 with a sedimentation coefficient of 6.0S. The enzyme does not contain Zn, Cu, or Fe. The amino acid composition revealed an unusually low proportion of proline residues (0.9%). p-Chloromercuriphenylsulfonate and phenylglyoxal inactivated the enzyme, but the presence of NADPH prevented the loss of activity due to thiol and arginine modification. The enzyme transferred the pro 4S hydrogen atom of NADPH to the substrate and the binding of the enzyme to NADPH resulted in a red shift of the ultraviolet absorption spectrum of the cofactor.  相似文献   

19.
Dissociation, denaturation, and deactivation of aldolase from rabbit muscle in the acid pH range have been investigated using sedimentation analysis, fluorescence, circular dichroism, and activity tests. Under comparable experimental conditions the pH-dependent profiles of deactivation and denaturation parallel the dissociation of the enzyme. In the range of dissociation at pH4-5tetramers and monomers are in equilibrium. Intrinsic chromophores and far-ultraviolet circular dichroism suggest the transition to be a complex multistep process. At pH approximately 2.3 the enzyme is split into its fully inactive monomers which still contain some residual secondary structure. After reassociation under optimum conditions (0.2 M phosphate buffer pH 7.6, 1 mM EDTA, 0.1 mM dithiothreitol, 0 degrees C, enzyme concentration 0.4-59 mug/ml) up to 95% enzymic activity is recovered which belongs to a renatured tetrameric species indistinguishable from the native enzyme by all available biochemical and physicochemical criteria.  相似文献   

20.
Some physical and chemical properties of the monomeric NADP+-dependent aldehyde reductase (previously called TPN-L-hexonate dehydrogenase or D-glucuronate reductase) from pig kidney have been examined. The amino acid composition has been determined. Four of the five thiol groups react with p-mercuribenzoate at pH 7, with no resulting loss of catalytic activity. High concentrations of p-mercuribenzoate cause complete enzyme inhibition, which can be partly reversed by addition of aldehyde reductase is low (9%, estimated from the ellipticity at 208 nm), and 70 to 80% of the tyrosine and tryptophan residues aare buried within the molecule. One molecule of NADPH binds to the enzyme (Kp equal 25 muM), causing a blue shift and enhancement of the coenzyme fluorescence, and suggesting that the environment of the active site is hydrophobic. In the reduction of D-glyceraldehyde, catalyzed by aldehyde reductase, the pro-4R "A" hydrogen of NADPH attacks the re face of the carbonyl group. This stereospecificity is the same as in the reductions of D-glyceraldehyde and acetaldehyde effected by rabbit muscle dehydrogenase and liver alcohol dehydrogenase, respectively.  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