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 62 毫秒
1.
A NAD-dependent, oxygen-labile alcohol dehydrogenase was purified from Desulfovibrio gigas. It was decameric, with subunits of M(r) 43,000. The best substrates were ethanol (Km, 0.15 mM) and 1-propanol (Km, 0.28 mM). N-terminal amino acid sequence analysis showed that the enzyme belongs to the same family of alcohol dehydrogenases as Zymomonas mobilis ADH2 and Bacillus methanolicus MDH.  相似文献   

2.
Alanine dehydrogenase [EC 1. 4. 1. 1] was purified to homogeneity from a crude extract of Enterobacter aerogenes ICR 0220. The enzyme had a molecular mass of about 245 kDa and consisted of six identical subunits. The enzyme showed maximal activity at about pH 10.9 for the deamination of L-alanine and at about pH 8.7 for the amination of pyruvate. The enzyme required NAD+ as a coenzyme. Analogs of NAD+, deamino-NAD+ and nicotinamide guanine dinucleotide served as coenzymes. Initial-velocity and product inhibition studies suggested that the deamination of L-alanine proceeded through a sequential ordered binary-ternary mechanism. NAD+ bound first to the enzyme, followed by L-alanine, and the products were released in the order of ammonia, pyruvate, and NADH. The Km were 0.47 mM for L-alanine, 0.16 mM for NAD+, 0.22 mM for pyruvate, 0.067 mM for NADH, and 66.7 mM for ammonia. The Km for L-alanine was the smallest in the alanine dehydrogenases studied so far. The enzyme gene was cloned into Escherichia coli JM109 cells and the nucleotides were sequenced. The deduced amino acid sequence was very similar to that of the alanine dehydrogenase from Bacillus subtilis. However, the Enterobacter enzyme has no cysteine residue. In this respect, the Enterobacter enzyme is different from other alanine dehydrogenases.  相似文献   

3.
Geranoyl-CoA carboxylase (EC 6.4.1.4) is a biotin-containing enzyme previously described in two genera of bacteria. Here we report the presence of geranoyl-CoA carboxylase in kingdom Plantae. Geranoyl-CoA carboxylase was purified 180-fold from maize leaves. The enzyme has a biotin-containing subunit of 122 kDa. The pH optimum for activity is 8.3. The apparent Km values for the substrates geranoyl-CoA, bicarbonate, and ATP are 64 +/- 5 microM, 0. 58 +/- 0.04 mM, and 8.4 +/- 0.4 microM, respectively. Subcellular fractionations indicate that geranoyl-CoA carboxylase is located in plastids. Geranoyl-CoA carboxylase activity is ubiquitous in organs of monocots and dicots and varies with development. We postulate that geranoyl-CoA carboxylase plays an important role in isoprenoid catabolism in plants, in a pathway analogous to that shown in Psuedomonas sp. In plants, this catabolic pathway would require the interaction of at least three subcellular compartments (plastids, microbodies, and mitochondria) and two biotin-containing enzymes, geranoyl-CoA carboxylase and 3-methylcrotonyl-CoA carboxylase.  相似文献   

4.
Properties and subunit structure of pig heart pyruvate dehydrogenase   总被引:1,自引:0,他引:1  
Pyruvate dehydrogenase [EC 1.2.4.1] was separated from the pyruvate dehydrogenase complex and its molecular weight was estimated to be about 150,000 by sedimentation equilibrium methods. The enzyme was dissociated into two subunits (alpha and beta), with estimated molecular weights of 41,000 (alpha) and 36,000 (beta), respectively, by polyacrylamide gel electrophoresis in sodium dodecyl sulfate. The subunits were separated by phosphocellulose column chromatography and their chemical properties were examined. The subunit structure of the pyruvate dehydrogenase was assigned as alpha2beta2. The content of right-handed alpha-helix in the enzyme molecule was estimated to be about 29 and 28% by optical rotatory dispersion and by circular dichroism, respectively. The enzyme contained no thiamine-PP, and its dehydrogenase activity was completely dependent on added thiamine-PP and partially dependent on added Mg2+ and Ca2+. The Km value of pyruvate dehydrogenase for thiamine diphosphate was estimated to be 6.5 X 10(-5) M in the presence of Mg2+ or Ca2+. The enzyme showed highly specific activity for thiamine-PP dependent oxidation of both pyruvate and alpha-ketobutyrate, but it also showed some activity with alpha-ketovalerate, alpha-ketoisocaproate, and alpha-ketoisovalerate. The pyruvate dehydrogenase activity was strongly inhibited by bivalent heavy metal ions and by sulfhydryl inhibitors; and the enzyme molecule contained 27 moles of 5,5'-dithiobis(2-nitrobenzoic acid)-reactive sulfhydryl groups and a total of 36 moles of sulfhydryl groups. The inhibitory effect of p-chloromercuribenzoate was prevented by preincubating the enzyme with thiamine-PP plus pyruvate. The structure of pyruvate dehydrogenase necessary for formation of the complex is also reported.  相似文献   

5.
Pancreatic islets were cultured for 24 h in the presence of 1 mM glucose, which renders islets incapable of responding to glucose with insulin release. These islets were compared to islets maintained at 20 mM glucose for 24 h. Detritiation of [2-3H]glucose and [5-3H]glucose in 1 mM glucose islets was normal, suggesting that glucose transport and phosphorylation and all enzymes of glycolysis were not down-regulated in the incapacitated islets. 14CO2 formation from [U-14C]glucose and [6-14C]glucose was inhibited up to 80% and 14CO2 from methyl succinate was inhibited up to 60%, indicating that down-regulation at (a) mitochondrial site(s) might explain the incapacitated insulin release. 14CO2 formation from [3,4-14C]glucose (which becomes [1-14C]pyruvate) was decreased, indicating that the reaction catalyzed by pyruvate dehydrogenase was down-regulated. This decrease, however, was not as large as the decreases in 14CO2 formation from [U-14C]glucose, [2-14C]glucose (which becomes [2-14C]pyruvate), or [6-14C]glucose (which becomes [3-14C]pyruvate), indicating that other reactions were also down-regulated. 14CO2 formation from [1-14C]glucose was inhibited less than that from [6-14C]glucose in the incapacitated islets (34 vs 54%) and these rates indicated that flux of glucose through the pentose phosphate pathway was increased in the incapacitated islet, such that 29% (0.4 nmol of 1.4 glucose/100 islets/90 min) was metabolized via this pathway in the incapacitated islet but only 3.4% (0.1 of 2.9 nmol glucose/100 islets/90 min) was metabolized via the pentose pathway in the 20 mM glucose islets. With rates of 14CO2 evolved from glucose labeled at C2 and C6 and from methyl succinate labeled at C1 + C4 and C2 + C3 the 14CO2 ratio formula was used to calculate the ratios of carboxylated and decarboxylated pyruvate. Roughly equal amounts of pyruvate entered the citric acid cycle by each route in islets maintained for 24 h at 1, 5, or 20 mM glucose. The results indicate that decarboxylation and carboxylation of pyruvate were about equally suppressed in incapacitated islets and that direct inhibition of reactions of the cycle was unlikely. This is consistent with evidence which indicates that down-regulation of both pyruvate carboxylase and pyruvate dehydrogenase occurs in incapacitated islets, i.e., under long-term conditions that modify amounts of enzymes (MacDonald et al., 1991, J. Biol. Chem. 266, 22392-22397).(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

6.
A previously unreported enzymatic activity is described for monomers of the beta 1 beta 1 isoenzyme of human alcohol dehydrogenase that were prepared from dimeric enzyme by freeze-thaw in liquid nitrogen. Whereas the dimeric enzyme has optimal activity at low substrate concentrations (2.5 mM ethanol, 50 microM NAD+; "low Km" activity), the monomer has its highest activity at high substrate concentrations (1.5 M ethanol, 2.5 mM NAD+; "high Km" activity). While the activity of the monomer does not appear to be saturated at 1.5 M ethanol, its maximal activity at this high ethanol concentration exceeds the Vmax of the dimer by about 3-fold. The apparent Km of NAD+ with monomers is 270 microM, and no activity could be detected with nicotinamide mononucleotide as cofactor. During gel filtration the high Km activity elutes at a lower apparent molecular weight position than the dimer. The kinetics of monomer-to-dimer reassociation are consistent with a second-order process with a rate constant of 240 M-1 s-1. The reassociation rate is markedly enhanced by the presence of NAD+. During refolding of beta 1 beta 1 following denaturation in 6 M guanidine hydrochloride, an enzyme species with high Km activity and spectral properties similar to the freeze-thaw monomer is observed, indicating that a catalytically active monomer is an intermediate in the refolding pathway. The enzymatic activity of the monomer implies that the intersubunit contacts of beta 1 beta 1 are not crucial in establishing a catalytically competent enzyme. However, the differences in specific activity and Km between monomer and dimer suggest that dimerization may serve to modulate the catalytic properties.  相似文献   

7.
The morphologic and histochemical effects of 3-nitropropionic acid (NPA) were examined in cultured murine embryonal carcinoma cells. NPA caused a dose-dependent inhibition of cell proliferation of cultured murine embryonal carcinoma cells at concentrations above 1.05 mM and was lethal at 4.2 mM. Morphologic changes included gross swelling of the cells, swelling of mitochondria and accumulation of organellar debris within the cytoplasm. NPA inhibited the activity of succinate dehydrogenase but not of malate, isocitrate or glucose-6-phosphate dehydrogenases, resulting in a decrease in intracellular ATP. Although succinate dehydrogenase activity was decreased by NPA, propionic acid and its mercapto-, 2-chloro-, and 3-chloro- derivates did not affect enzyme activity. 3-Nitropropanol also inhibited succinate dehydrogenase but only at a much higher concentration than was required with NPA. The results provide evidence that cytotoxicity caused by NPA results from inhibition of succinate dehydrogenase activity leading to depression of ATP synthesis. Loss of cellular integrity is probably a direct consequence of failure of energy-dependent cell homeostatic mechanisms such as the plasma membrane Na+/K+ pump, resulting in swelling and ultimately lysis of the cell.  相似文献   

8.
Excitatory amino acids are an important cause of cell death in the hypoxic and ischaemic brain. Neuronal glutamate stores are depleted rapidly in hypoxia, but alanine production rises under such conditions and has been suggested to be a potential precursor of glutamate. To test this hypothesis, we have investigated amino acid metabolism using 13C NMR with superfused guinea pig cortical slices subjected to varying degrees of hypoxia. During severe hypoxia, brain slices metabolising 5 mM [2-(13)C]pyruvate exported [2-(13)C]alanine into the superfusion fluid. The metabolic fate of alanine during normoxia and hypoxia was tested by superfusion of brain slices with 10 mM glucose and 2 mM [2-(13)C,15N]alanine. Metabolism of exogenous alanine leads to the release of aspartate into the superfusion fluid. The pattern of labelling of aspartate indicated that it was synthesised via the glial-specific enzyme pyruvate carboxylase. 13C-labelled glutamate was produced with both normoxia and hypoxia, but concentrations were 30-fold lower than for labelled aspartate. Thus, although substantial amounts of glutamate are not synthesised from alanine in hypoxia, there is significant production of aspartate, which also may have deleterious effects as an excitatory amino acid.  相似文献   

9.
The study was aimed at evaluating changes in lens antioxidant status, glucose utilization, redox state of free cytosolic NAD(P)-couples and adenine nucleotides in rats with 6-week streptozotocin-induced diabetes, and to assess a possibility of preventing them by DL-alpha-lipoic acid. Rats were divided into control and diabetic groups treated with and without DL-alpha-lipoic acid (100 mg x kg body weight(-1) x day(-1), i.p.). The concentrations of glucose, sorbitol, fructose, myo-inositol, oxidized glutathione, glycolytic intermediates, malate, alpha-glycerophosphate, and adenine nucleotides were assayed in individual lenses spectrofluorometrically by enzymatic methods, reduced glutathione and ascorbate--colorimetrically, and taurine by HPLC. Free cytosolic NAD+:NADH and NADP+:NADPH ratios were calculated from the lactate dehydrogenase and malic enzyme systems. Sorbitol pathway metabolites were found to increase, and antioxidant concentrations were reduced in diabetic rats compared with controls. The profile of glycolytic intermediates (increase in glucose 6-phosphate and fructose 6-phosphate, decrease in fructosel,6-diphosphate, increase in dihydroxyacetone phosphate, 3-phosphoglycerate, phosphoenolpyruvate, pyruvate, and no change in lactate), and 5.9-fold increase in alpha-glycerophosphate suggest diabetes-induced inhibition of glycolysis. Free cytosolic NAD+:NADH ratios, ATP levels, ATP/ADP x inorganic phosphate (Pi), and adenylate charge were reduced in diabetic rats while free cytosolic NADP+:NADPH ratios were elevated. Diabetes-induced changes in the concentrations of antioxidants, key glycolytic intermediates, free cytosolic NAD+:NADH ratios, and energy status were partially prevented by DL-alpha-lipoic acid, while sorbitol pathway metabolites and free cytosolic NADP+:NADPH ratios remained unaffected. In conclusion, diabetes-induced impairment of lens antioxidative defense, glucose intermediary metabolism via glycolysis, energy status and redox changes are partially prevented by DL-alpha-lipoic acid. The findings support the important role of oxidative stress in lens metabolic imbalances in diabetes.  相似文献   

10.
A hybrid numerical method, which employs molecular mechanics to describe the bulk of the solvent-protein matrix and a semiempirical quantum-mechanical treatment for atoms near the reactive site, was utilized to simulate the minimum energy surface and reaction pathway for the interconversion of malate and oxaloacetate catalyzed by the enzyme malate dehydrogenase (MDH). A reaction mechanism for proton and hydride transfers associated with MDH and cofactor nicotinamide adenine dinucleotide (NAD) is deduced from the topology of the calculated energy surface. The proposed mechanism consists of (1) a sequential reaction with proton transfer preceding hydride transfer (malate to oxaloacetate direction), (2) the existence of two transition states with energy barriers of approximately 7 and 15 kcal/mol for the proton and hydride transfers, respectively, and (3) reactant (malate) and product (oxaloacetate) states that are nearly isoenergetic. Simulation analysis of the calculated energy profile shows that solvent effects due to the protein matrix dramatically alter the intrinsic reactivity of the functional groups involved in the MDH reaction, resulting in energetics similar to that found in aqueous solution. An energy decomposition analysis indicates that specific MDH residues (Arg-81, Arg-87, Asn-119, Asp-150, and Arg-153) in the vicinity of the substrate make significant energetic contributions to the stabilization of proton transfer and destabilization of hydride transfer. This suggests that these amino acids play an important role in the catalytic properties of MDH.  相似文献   

11.
Serine transhydroxymethylase, from lamb or rabbit liver, is known to catalyze slow transamination of D-alanine, but not of L-amino acids, in a tetrahydrofolate-independent reaction. Both enzymes will process the D-isomer of beta-fluoroalanine for alpha, beta-elimination of HF to yield an aminoacrylate-pyridoxal-P-enzyme intermediate. This intermediate partitions between harmless hydrolysis to pyruvate, NH4+, and active enzyme-pyridoxal-P (catalytic turnover) and suicidal enzyme alkylation by covalent modification with an average partition ratio of 40-60 turnovers/inactivation event/monomer unit of this tetrameric enzyme. Enzyme inactivation occurs with stoichiometric incorporation of radioactive label from D-[1,2-14C]fluoroalanine. Titration of enzymic cysteinyl --SH groups with 5,5'-dithiobis(2-nitrobenzoate) indicates loss of 1 --SH group on inactivation. Acid hydrolysis of radioactive-inactive enzyme confirms cysteine residue modification. Treatment of inactive enzyme with 6 M urea, then KBH4, followed by acid hydrolysis yields two radioactive compounds, lanthionine and S-carboxyhydroxyethylcysteine, in about equal amounts. The addition of tetrahydrofolate stimulates both pyruvate production and inactivation to equal extents with about a 200-fold rate acceleration at 0.5 mM tetrahydrofolate to turnover numbers of approximately 120 min-1. The Km for D-fluoroalanine is high, 10-60 mM, and this low substrate affinity suggests D-fluoroalanine will not be a useful in vivo agent for selective inactivation of liver cell serine transhydroxymethylases.  相似文献   

12.
As determined by EPR, malic enzyme from pigeon liver binds Mn2+ with a half-site stoichiometry of two tight binding sites (KD=6 to 10 mum) per enzyme tetramer and at two to four weak binding sites (KD=0.43 to 1.34 mM). The activation of malic enzyme by Mn2+ at high levels of L-malate shows biphasic kinetics yielding two activator constants for Mn2+. The dissociation constants of Mn2+ for both classes of sites are of the same order as the kinetically determined activator constants of Mn2+, indicating active site binding at both classes of binding sites. The binding of Mn2+ to the tight sites enhances the paramagnetic effect of Mn2+ on 1/T1 of water protons by a factor (epsilon) of 17, while binding at the weak sites yields a smaller epsilon of 11. The coenzymes TPN and TPNH have no effects on epsilon, while the carboxylic acid substrates L-malate and pyruvate and the inhibitors D-malate and oxalate significantly decrease epsilon. TPNH causes a 38-fold tightening of binding of the substrate L-malate to the enzyme-Mn2+ complex, consistent with the previously described highly ordered kinetic scheme, but only a 2-fold tightening of binding of the competitive inhibitor D-malate. The dissociation constant of L-malate from the quaternary E-Mn2+-TPNH-L-malate complex (32 muM) agrees with the Km of L-malate (25 muM), indicating active site binding. The dissociation constants of pyruvate from the ternary E-Mn2+-pyruvate complex (12 mM) and from the quaternary E-Mn2+-TPN-pyruvate complex (20 mM) are similar to the Km of pyruvate (5 mM), also indicating active site binding and a less highly ordered kinetic scheme for the reactions of pyruvate than for those of L-malate. Analysis of the frequency dependence of 1/T1 of water protons indicates that two fast exchanging water ligands remain coordinated to Mn2+ in the binary E-Mn2+ complex. The binding of the substrates L-malate and pyruvate and of the transition state analog oxalate to the E-Mn2+ complex decrease the number of fast exchanging water ligands on Mn2+ by approximately 1, but the binding of D-malate has no significant effect on this parameter, indicating the occlusion or replacement of a water ligand of the enzyme-bound Mn2+ by a properly oriented substituent on C-2 of the substrate. Occlusion rather than replacement of a water ligand by pyruvate is established by studies of 1/T1 of 13COO- and 13CO-enriched pyruvate which indicate second sphere Mn2+ to pyruvate distances of 4.6 A (COO-) and 4.8 A (CO) in the ternary enzyme-Mn2+-pyruvate complex. Formation of the quaternary complex with TPN increases these distances by 0.8 A, indicating the participation of a second sphere enzyme-Mn2+-(H2O)-pyruvate complex in catalysis. Thus, malic enzyme, like five other enzymes which utilize metals to polarize carbonyl groups, forms a second sphere complex with its substrate.  相似文献   

13.
The flux through different segments of the tricarboxylic acid cycle was measured in rat brain synaptosomes with gas chromatography-mass spectrometry using either deuterated glutamine or [13C]aspartate. The flux between 2-oxoglutarate and oxaloacetate was estimated to be 3.14 and 4.97 nmol/min/mg protein with and without glucose, respectively. These values were 3-5-fold faster than the flux between oxaloacetate and 2-oxoglutarate (0.92 nmol/min per mg protein) measured in the presence of glucose. The pattern of intermediates labeling suggests that the overall rate-controlling reaction involves either citrate synthase or pyruvate dehydrogenase but not 2-oxoglutarate or isocitrate dehydrogenase. The enrichment in [3,3,4,4-2H4]glutamate from [2,3,3,4,4-2H5]glutamine was as rapid as in [2,3,3,4,4-2H5]glutamate, which indicates that the aspartate aminotransferase reaction is severalfold faster than the flux through the tricarboxylic acid cycle. [13C]Aspartate was rapidly converted to [13C]malate, suggesting that in intact synaptosomes aspartate entry into the mitochondrion is very slow. The finding that aspartate is taken up by mitochondria as malate, along with the observed high enrichment in [3-2H]malate (from [2,3,3,4,4-2H5]glutamine), is consistent with the substantial synaptosomal activity of the malate/aspartate shuttle.  相似文献   

14.
The prostaglandin-E2 9-reductase (PGE2 9-reductase) activity in the corpus luteum of rabbits corresponds to a cytosolic, NADPH-dependent enzyme with a molecular mass of 36 kDa. This enzyme was purified from corpora lutea on day 12 of pseudopregnancy with a 266-fold enrichment. The main purification step was affinity chromatography using Red Sepharose CL-6B. The efficiency of this column was improved by elution with 1 mM NADH prior to elution of the active fractions with 1 mM NADPH. Amino acid sequence data demonstrate that the rabbit luteal PGE2 9-reductase has to be classified as a member of the aldo-keto reductase superfamily. The enzyme revealed a wide substrate specificity comprising the reduction of aldehydes, ketones, and quinones. Apparent kinetic constants were determined using methylglyoxal, DL-glyceraldehyde, and 9,10-phenanthrenquinone as substrates. The fully purified enzyme showed two catalytic activities of particular interest: PGE2 9-reductase and 20 alpha-hydroxysteroid dehydrogenase (20 alpha-HSD) activities. The competitive inhibition of 20 alpha-HSD activity by PGE2 indicates that progesterone and PGE2 are substrates for the same enzyme. From these results, we conclude that prostaglandin and steroid metabolism are tightly linked to each other. For this reason the aldo-keto reductase could be a key enzyme in the cascade of events leading to the regression of the corpus luteum in the rabbit.  相似文献   

15.
In the yeast Saccharomyces cerevisiae, choline kinase (ATP:choline phosphotransferase, EC 2.7.1.32) is the product of the CKI gene. Choline kinase catalyzes the committed step in the synthesis of phosphatidylcholine by the CDP-choline pathway. The yeast enzyme was overexpressed 106-fold in Sf-9 insect cells and purified 71.2-fold to homogeneity from the cytosolic fraction by chromatography with concanavalin A, Affi-Gel Blue, and Mono Q. The N-terminal amino acid sequence of purified choline kinase matched perfectly with the deduced sequence of the CKI gene. The minimum subunit molecular mass (73 kDa) of purified choline kinase was in good agreement with the predicted size (66.3 kDa) of the CKI gene product. Native choline kinase existed in oligomeric structures of dimers, tetramers, and octomers. The amounts of the tetrameric and octomeric forms increased in the presence of the substrate ATP. Antibodies were raised against the purified enzyme and were used to identify choline kinase in insect cells and in S. cerevisiae. Maximum choline kinase activity was dependent on Mg2+ ions (10 mM) at pH 9.5 and at 30 degrees C. The equilibrium constant (0.2) for the reaction indicated that the reverse reaction was favored in vitro. The activation energy for the reaction was 6.26 kcal/mol, and the enzyme was labile above 30 degrees C. Choline kinase exhibited saturation kinetics with respect to choline and positive cooperative kinetics with respect to ATP (n = 1.4-2.3). Results of the kinetic experiments indicated that the enzyme catalyzes a sequential Bi Bi reaction. The Vmax for the reaction was 138.7 micromol/min/mg, and the Km values for choline and ATP were 0.27 mM and 90 microM, respectively. The turnover number per choline kinase subunit was 153 s-1. Ethanolamine was a poor substrate for the purified choline kinase, and it was also poor inhibitor of choline kinase activity. ADP inhibited choline kinase activity (IC50 = 0.32 mM) in a positive cooperative manner (n = 1.5), and the mechanism of inhibition with respect to ATP and choline was complex. The regulation of choline kinase activity by ATP and ADP may be physiologically relevant.  相似文献   

16.
Kinetic methods have been used to determine the interrelationship between HCO-3, pyruvate and acetyl-CoA and their effect on pigeon kidney pyruvate carboxylase (pyruvate: CO2 ligase [ADP], EC 6.4.1.1). HCO-3 shows a negative co-operative effect (biphasic kinetics with two different Km values). Pyruvate influences the attachment of HCO-3 to this enzyme. The same has been shown for acetyl-CoA. Contrary to the results of other investigators no co-operative effect was seen with pyruvate even at different concentrations of acetyl-CoA. HCO-3 itself shows hardly any effect on the homotropic positive co-operativity (sigmoidal kinetics) of acetyl-CoA. The negative co-operative effect of HCO-3 could not be removed even at saturating concentrations of pyruvate and/or acetyl-CoA, which is also supported by the n and Rs values. The results of this communication bring out differences between pigeon kidney pyruvate carboxylase and pyruvate carboxylase from other sources. It is also suggested that there may be different allosteric and regulatory sites for acetyl-CoA, HCO-3 and pyruvate.  相似文献   

17.
Previous studies have capitalized on ordered kinetic mechanisms in the design of biospecific affinity chromatographic methods for highly efficient purifications and mechanistic studies of enzymes. The most direct tactic has been the use of immobilised analogues of the following, usually enzyme-specific substrates, e.g., lactate/pyruvate in the case of lactate dehydrogenase for which NAD+ is the leading substrate. Such immobilised specific substrates are, however, often difficult or impossible to synthesise. The locking-on strategy reverses the tactic by using the more accessible immobilised leading substrate, immobilised NAD+, as adsorbent with soluble analogues of the enzyme-specific ligands (e.g., lactate in the case of lactate dehydrogenase) providing a substantial reinforcement of biospecific adsorption sufficient to effect adsorptive selection of an enzyme from a group of enzymes such as the NAD(+)-specific enzymes. The value of this approach is demonstrated using model studies with lactate dehydrogenase (LDH, EC 1.1.1.27), alcohol dehydrogenase (ADH, EC 1.1.1.1), glutamate dehydrogenase (GDH, EC 1.4.1.3) and malate dehydrogenase (MDH, EC 1.1.1.37). Purification of bovine liver GDH in high yield from crude extracts is described using the tactic.  相似文献   

18.
The transport of the nephrotoxic mycotoxin ochratoxin A across the renal peritubular membrane was examined in suspensions of rabbit renal proximal tubules. Ochratoxin A transport across the peritubular membrane was a high-affinity, low-capacity carrier-mediated process with a Jmax value of 0.12 +/- 0.4 nmol/mg of protein/min and a Km value of 1.4 +/- 0.1 microM. The apparent Michaelis constants for inhibition of [3H]para-aminohippurate (PAH) uptake by ochratoxin A inhibition was 1.5 microM, which is similar to the Km value for ochratoxin A uptake in tubule suspensions and suggests that ochratoxin A could be a substrate for the organic anion pathway. The capacity and affinity for peritubular ochratoxin A transport were 40-fold lower and > 100-fold greater, respectively, than those measured for the peritubular uptake of [3H]PAH in tubule suspensions. A concentration of 2.5 mM PAH, which reduced the uptake of [3H]PAH by 90%, reduced ochratoxin A uptake by only 40% to 50%, whereas probenecid concentrations of 0.6 to 2 mM reduced ochratoxin A accumulation in tubule suspensions up to approximately 80% to 90%. This probenecid-sensitive, PAH-insensitive uptake of ochratoxin A suggested that at least one mediated pathway other than the organic anion transporter was involved in the peritubular uptake of this mycotoxin. A 2 mM concentration of the fatty acid octanoate and 1.5 mM concentration of the nonsteroidal anti-inflammatory agent piroxicam were as effective as probenecid in blocking ochratoxin A uptake. The apparent Ki values for inhibition of ochratoxin A uptake by probenecid, piroxicam and octanoate were 30.5 +/- 7.9, 23.2 +/- 10.4 and 81.5 +/- 8.7 microM, respectively. The ability of octanoic acid to inhibit ochratoxin A transport to the same extent as probenecid and a greater extent than PAH suggests that a separate fatty acid transport pathway may be involved in the accumulation of ochratoxin A by suspensions of rabbit renal proximal tubules.  相似文献   

19.
1. Phenylpyruvate inhibits pyruvate plus malate oxidation in human and rat skeletal muscle mitochondria in state 3 and in the uncoupled state, it has, however, no effect in state 4. 2. Inhibition by phenylpyruvate of pyruvate oxidation by intact uncoupled rat muscle mitochondria was competitive, with the Ki value about 0.18 mM. 3. It is suggested that the inhibition of pyruvate oxidation is due to the action of phenylpyruvate on muscle pyruvate dehydrogenase, and is the principal cause of the elevated concentration of pyruvate and lactate in blood plasma of phenylketonuric patients.  相似文献   

20.
At 9 mM glucose, experimental results show that mitochondrial phosphate depletion (induced by glucose phosphorylation, catalyzed by mitochondrial hexokinase) reduces the activities of the respiratory chain, oxidative phosphorylation, and glutaminase. Consequently, the 14C-lactate oxidation to 14CO2 is lowered in the presence of glucose. The fall of ATP level triggers a high aerobic glycolysis by deinhibiting fructose-6-P kinase. NADH, generated by enhanced glyceraldehyde-3-P dehydrogenase activity, increases the reducing power. Moreover, the lactate dehydrogenase (LDH) system is shifted toward lactate formation, while NAD+ is regenerated and the oligomycin-inhibited ATP production is replaced by the iodoacetate-inhibited ATP production. From 14CO2 production and lactate accumulation it is calculated that about 60% of 14C-glucose which disappears is channelled into extraglycolytic reactions. On the contrary, 82% of glucose below l mM is metabolized through non-glycolytic reactions. The pyruvate kinase-M2 (PK-M2) inhibition does not limit the glycolytic flow from 9 mM glucose, but it may cause sustained gluconeogenesis.  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