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 250 毫秒
1.
Pyrroline-5-carboxylic acid reductase from soybean leaves   总被引:1,自引:0,他引:1  
Pyrroline-5-carboxylic acid reductase was purified 40-fold from soybean leaves (Glycine max L. var Corsoy). The enzyme was fairly unstable, had a broad pH optimum, and was inactivated by heat and acid; NADH and NADPH both served as cofactors. It had a higher activity with NADH (about 4 ×) compared to NADPH, but a lower Km for NADPH. NADP+ inhibited both the NADH- and NADPH-dependent activity. Sulfhydryl group blocking agents reduced the activity as did the carbonyl blocking agent, NH2OH. Thiazolidine-4-carboxylic acid and phosphate inhibited the enzyme and proline inhibited only at high concentrations. ATP, GTP, and CTP were all effective inhibitors of both the NADH- and NADPH-dependent activity. Phosphorylated nucleotide inhibition was reversed by Mg2+ ions.  相似文献   

2.
Pyrroline-5-carboxylate reductase, which required reduced pyridine nucleotide and Δ′-pyrroline-5-carboxylate for proline synthesis, was isolated from pumpkin cotyledons. The enzyme was found in the soluble fraction and had a 4.5-fold greater activity with NADH than NADPH. The enzyme was inhibited by NH2OH, NADP, ATP and slightly by proline. Glutathione or pyridoxal-5-phosphate had little effect on enzyme activity. The enzyme had a pH optimum between 7·0 and 7·6 and was not inhibited by high concentrations of NADH or Δ′-pyrroline-5-carboxylate.  相似文献   

3.
UDP-Galactose 4′-epimerase was purified ca 800-fold through a multi-step procedure which included affinity chromatography using NAD+ -Agarose. Three forms of the enzyme were separated by gel-filtration but only the major form was purified. The pH optimum of the enzyme was 9.5. Exogenous NAD+ was not required for enzymic activity but its removal caused inactivation. The enzyme was unstable below pH 7.0 but stable at pH 8.0 in the presence of glycerol and at ?20° for two months. The equilibrium constant for the enzyme-catalysed reaction was 3.2 ± 0.15. The Km for UDP-galactose and UDP-glucose were 0.12 mM and 0.25 mM, respectively. The inhibition by NADH was competitive, with a Ki of 5 μM. The MW of the enzyme was 78 000; the two minor forms showed the values of 158 000 and 39 000, respectively.  相似文献   

4.
Malic enzyme (EC 1.1.1.40) converts l-malate to pyruvate and CO2 providing NADPH for metabolism especially for lipid biosynthesis in oleaginous microorganisms. However, its role in the oleaginous yeast, Yarrowia lipolytica, is unclear. We have cloned the malic enzyme gene (YALI0E18634g) from Y. lipolytica into pET28a, expressed it in Escherichia coli and purified the recombinant protein (YlME). YlME used NAD+ as the primary cofactor. Km values for NAD+ and NADP+ were 0.63 and 3.9 mM, respectively. Citrate, isocitrate and α-ketoglutaric acid (>5 mM) were inhibitory while succinate (5–15 mM) increased NADP+- but not NAD+-dependent activity. To determine if fatty acid biosynthesis could be increased in Y. lipolytica by providing additional NADPH from an NADP+-dependent malic enzyme, the malic enzyme gene (mce2) from an oleaginous fungus, Mortierella alpina, was expressed in Y. lipolytica. No significant changes occurred in lipid content or fatty acid profiles suggesting that malic enzyme is not the main source of NADPH for lipid accumulation in Y. lipolytica.  相似文献   

5.
The pyruvate dehydrogenase complex was isolated from the mitochondria of broccoli florets and shown to be similar in its reaction mechanism to the complexes from other sources. Three families of parallel lines were obtained for the initial velocity patterns, indicating a multisite ping-pong mechanism. The apparent Km values obtained were 321 ± 18, 148 ± 13, and 7.2 ± 0.51 μm for pyruvate, NAD+, and CoA, respectively. Product inhibition studies using acetyl-CoA and NADH yielded results which were in agreement with those predicted by the multisite ping-pong mechanism. Acetyl-CoA and NADH were found to be competitive inhibitors versus CoA and NAD+, respectively. All other substrate-product combinations showed uncompetitive inhibition patterns, except for acetyl-CoA versus NAD+. Among various metabolites tested, only hydroxypyruvate (Ki = 0.11 mM) and glyoxylate (Ki = 3.27 mM) were found to be capable of inhibiting the broccoli enzyme to a significant degree. Initial velocity patterns using Mg2+? or Ca2+-thiamine pyrophosphate and pyruvate as the variable substrate were found to be consistent with an equilibrium ordered mechanism where Mg? or Ca-thiamine pyrophosphate bind first, with dissociation constants of 33.8 and 3 μm, respectively. The Mg- or Ca-thiamine pyrophosphate complexes also dissociated rapidly from the enzyme complex.  相似文献   

6.
An NADH-dependent 15-ketoprostaglandin Δ13 reductase has been purified to near homogeneity from human placenta by a procedure which includes affinity chromatography on blue Sepharose. The enzyme utilizes as substrates 15-ketoprostaglandins of the E, F, A, and B series, and the reaction is experimentally irreversible. Molecular weight estimations on Sephadex G-100 and sodium dodecyl sulfate disc gel electrophoresis suggest that the enzyme is a dimer. The subunits appear to be similar in size if not identical and have a molecular weight of 35,000. The mechanism of the reaction of 15-ketoprostaglandin E2 and NADH catalyzed by this enzyme has been investigated by steady-state kinetic methods. The 13,14-dihydro-15-ketoprostaglandin product is an inhibitor of the reaction, being competitive with respect to 15-ketoprostaglandin E2 and noncompetitive with respect to NADH; NAD+ does not inhibit the reaction. NADPH and Cibacron blue 3G-A are “dead-end” inhibitors of the reaction; both act competitively with respect to NADH and noncompetitively with respect to 15-ketoprostaglandin E2. These observations are consistent with a rapid equilibrium random mechanism with the formation of an unreactive enzyme · NADH · 13,14-dihydro-15-ketoprostaglandin E2 complex. The interaction of NADPH and Cibacron blue 3G-A with the free enzyme was investigated further by fluorimetry. Both substances bind to the free enzyme and quench its fluorescence. This property was utilized to titrate the enzyme, and a value of 3.28 × 10?11 mol of binding sites/mU of enzyme was obtained.  相似文献   

7.
The NAD+-dependent isocitrate dehydrogenase from etiolated pea (Pisum sativum L.) mitochondria was purified more than 200-fold by dye-ligand binding on Matrix Gel Blue A and gel filtration on Superose 6. The enzyme was stabilized during purification by the inclusion of 20% glycerol. In crude matrix extracts, the enzyme activity eluted from Superose 6 with apparent molecular masses of 1400 ± 200, 690 ± 90, and 300 ± 50 kD. During subsequent purification steps the larger molecular mass species disappeared and an additional peak at 94 ± 16 kD was evident. The monomer for the enzyme was tentatively identified at 47 kD by sodium dodecyl-polyacrylamide gel electrophoresis. The NADP+-specific isocitrate dehydrogenase activity from mitochondria eluted from Superose 6 at 80 ± 10 kD. About half of the NAD+ and NADP+-specific enzymes remained bound to the mitochondrial membranes and was not removed by washing. The NAD+-dependent isocitrate dehydrogenase showed sigmodial kinetics in response to isocitrate (S0.5 = 0.3 mm). When the enzyme was aged at 4°C or frozen, the isocitrate response showed less allosterism, but this was partially reversed by the addition of citrate to the reaction medium. The NAD+ isocitrate dehydrogenase showed standard Michaelis-Menten kinetics toward NAD+ (Km = 0.2 mm). NADH was a competitive inhibitor (Ki = 0.2 mm) and, unexpectedly, NADPH was a noncompetitive inhibitor (Ki = 0.3 mm). The regulation by NADPH may provide a mechanism for coordination of pyridine nucleotide pools in the mitochondria.  相似文献   

8.
ATP and respiration (NADH)-driven NAD(P)+ transhydrogenase (EC 1.6.1.1) activities are low in membranes from Escherichia coli cultured on yeast extract medium (17 and 21 nmol/min × mg) but high on glucose (82 and 142 nmol/min × mg). The ATPase and respiratory activities in both cases appeared comparable. Growth of the bacteria in yeast extract medium followed by washing and replacement into a glucose medium showed that after 3 h the energy-linked and energy-independent NAD(P)+ transhydrogenase (reduction of acetylpyridine NAD+ by NADPH) activities had appeared simultaneously. Incorporation of chloramphenicol or omission of glucose in the induction medium resulted in no increase in these activities indicating that de novo protein synthesis is required for the induction of energy-linked and -independent NAD(P)+ transhydrogenase. It was found that the Km values for acetylpyridine NAD+ and NADPH for the energy-independent reaction in membranes from glucose grown cells (143 and 62 μm) were similar to those in membranes from cells grown on glucose-yeast extract (135 and 45 μm), respectively, but the maximum velocity at infinite acetyl pyridine NAD+ and NADPH increased from 353 to 2175 nmol/min × mg. Furthermore, the membrane-bound NAD(P)+ transhydrogenase in glucose-yeast extract grown cells showed substrate inhibition at high NADPH and low acetyl pyridine NAD+ levels. Further kinetic data demonstrate that the mechanism of the energy-independent NAD(P)+ transhydrogenase in E. coli is similar to that of the mitochondrial enzyme and exhibits similar responses to competitive inhibitors at the NAD+ and NADPH sites.  相似文献   

9.
A yeast growing at 48°C was isolated from soil and the strain was identified as Cryptococcus lactativorus. The aldose reductase which the strain produced was purified 114-fold with an overall recovery of 36%. The stability of the enzyme was higher than that of other aldose reductases. The half life of the enzyme was 800 h and 14 h at 30°C and 50°C, respectively. The enzyme showed the best activity with d-xylose. l-Sorbose and d-fructose were also reduced by the enzyme. The enzyme was active with both NADPH and NADH as a conenzyme, and the activity with NADH was 1.25 times higher than that with NADPH. The Kmapp value for d-xylose was 8.6 mM and the Vmaxapp was 20.8 units/mg NADH was used as a coenzyme. The Kmapp values for NADPH and NADH were 6μM and 170 μM, respectively, when d-glucose was used as a substrate.  相似文献   

10.
Glucose-6-phosphate dehydrogenase from Leuconostoc mesenteroides utilizes either NAD+ or NADP+ as coenzyme. Kinetic studies showed that NAD+ and NADP+ interact with different enzyme forms (Olive, C., Geroch, M. E., and Levy, H. R. (1971) J. Biol. Chem.246, 2047–2057). In the present study the techniques of fluorescence quenching and fluorescence enhancement were used to investigate the interaction between Leuconostoc mesenteroides glucose-6-phosphate dehydrogenase and coenzymes. In addition, kinetic studies were performed to examine interaction between the enzyme and various coenzyme analogs. The maximum quenching of protein fluorescence is 5% for NADP+ and 50% for NAD+. The dissociation constant for NADP+, determined from fluorescence quenching measurements, is 3 μm, which is similar to the previously determined Km of 5.7 μm and Ki of 5 μm. The dissociation constant for NAD+ is 2.5 mm, which is 24 times larger than the previously determined Km of 0.106 mm. Glucose 1-phosphate, a substrate-competitive inhibitor, lowers the dissociation constant and maximum fluorescence quenching for NAD+ but not for NADP+. This suggests that glucose 6-phosphate may act similarly and thus play a role in enabling the enzyme to utilize NAD+ under physiological conditions. When NADPH binds to the enzyme its fluorescence is enhanced 2.3-fold. The enzyme was titrated with NADPH in the absence and presence of NAD+; binding of these two coenzymes is competitive. The dissociation constant for NADPH from these measurements is 24 μm; the previously determined Ki is 37.6 μm. The dissociation constant for NAD′ is 2.8 mm, in satisfactory agreement with the value obtained from protein fluorescence quenching measurements. Various compounds which resemble either the adenosine or the nicotinamide portion of the coenzyme structure are coenzyme-competitive inhibitors; 2′,5′-ADP, the most inhibitory analog tested, gives NADP+-competitive and NAD+-noncompetitive inhibition, consistent with the kinetic mechanism previously proposed. By using pairs of coenzyme-competitive inhibitors it was shown in kinetic studies that the two portions of the NAD+ structure cannot be accommodated on the enzyme simultaneously unies they are covalently linked. Fluorescence studies showed that there are both “buried” and “exposed” tryptophan residues in the enzyme structure.  相似文献   

11.
The l-alanine dehydrogenase (ADH) of Anabaena cylindrica has been purified 700-fold. It has a molecular weight of approximately 270000, has 6 sub-units, each of molecular weight approximately 43000, and shows activity both in the aminating and deaminating directions. The enzyme is NADH/NAD+ specific and oxaloacetate can partially substitute for pyruvate. The K m app for NAD+ is 14 M and 60 M at low and high NAD+ concentrations, respectively. The K m app for l-alanine is 0.4 mM, that for pyruvate is 0.11 mM, and that for oxaloacetate is 3.0 mM. The K m app for NH 4 + varies from 8–133 mM depending on the pH, being lowest at high pH levels (pH 8.7 or above). Alanine, serine and glycine inhibit ADH activity in the aminating direction. The enzyme is active both in heterocysts and vegetative cells and activity is higher in nitrogen-starved cultures than in N2-fixing cultures. The data suggest that although alanine is formed by the aminating activity of ADH, entry of newly fixed ammonia into organic combination does not occur primarily via ADH in N2-fixing cultures of A. cylindrica. Ammonia assimilation via ADH may be important in cultures with an excess of available nitrogen. The deaminating activity of the enzyme may be important under conditions of nitrogen-deficiency.Abbreviations ADH alanine dehydrogenase - DEAE diethylamino ethyl cellulose - EDTA ethylenediamine tetraacetic acid - GDH glutamic dehydrogenase - GS glutamine synthetase - GOT aspartate-glutamate aminotransferase - NAD+ nicotinamide adenine dinucleotide - NADH reduced nicotinamide adenine dinucleotide - NADP+ nicotinamide adenine dinucleotide phosphate - NADPH reduced nicotinamide adenine dinucleotide phosphate - SDS sodium dodecyl sulphate - Tris tris(hydroxymethyl) aminomethane  相似文献   

12.
A halophilic NAD+-dependent 2-aminobutyrate dehydrogenase (EC1.4.1.1) was purified to homogeneity from a crude extract of an extreme halophile, Halobacterium saccharovorum DSM 1137, with a 30% yield. The enzyme had a molecular mass of about 160 kDa and consisted of four identical subunits. It retained more than 70% of the activity after heating at 60 °C for 1 h and kept it at 30 °C for 8 months in the presence of 2 M NaCl. The enzyme showed maximum activity in the presence of 2 M RbCl or KCl. The enzyme required NAD+ as a coenzyme and used -2-aminobutyrate, -alanine, and -norvaline as substrates. The best substrate was -2-aminobutyrate. The optimum pH was 9.3 for the oxidative deamination of -2-aminobutyrate and 8.6 for the reductive amination of 2-ketobutyrate. The Michaelis constants were 1.2 mM for -2-aminobutyrate, 0.16 mM for NAD+, 0.012 mM for NADH, 0.78 mM for 2-ketobutyrate, and 500 mM for ammonia in the presence of 2 M KCl. The Km values for the substrates depended on the concentration of KCl, and the Km values decreased under high salt conditions.  相似文献   

13.
The autotrophic nitrifying bacterium Nitrosomonas europaea does not synthesize 2-oxoglutarate (α-ketoglutarate) dehydrogenase under aerobic conditions and so has an incomplete citric acid cycle. l-malate (S-malate) dehydrogenase (MDH) from N. europaea was predicted to show similarity to the NADP+-dependent enzymes from chloroplasts and was separated from the NAD+-dependent proteins from most other bacteria or mitochondria. MDH activity in a soluble fraction from N. europaea ATCC 19718 was measured spectrophotometrically and exhibited simple Michaelis–Menten kinetics. In the reductive direction, activity with NADH increased from pH 6.0 to 8.5 but activity with NADPH was consistently lower and decreased with pH. At pH 7.0, the K m for oxaloacetate was 20 μM; the K m for NADH was 22 μM but that for NADPH was at least 10 times higher. In the oxidative direction, activity with NAD+ increased with pH but there was very little activity with NADP+. At pH 7.0, the K m for l-malate was 5 mM and the K m for NAD+ was 24 μM. The reductive activity was quite insensitive to inhibition by l-malate but the oxidative activity was very sensitive to oxaloacetate. MDH activity was not strongly activated or inhibited by glycolytic or citric acid cycle metabolites, adenine nucleotides, NaCl concentrations, or most metal ions, but increased with temperature up to about 55 °C. The reductive activity was consistently 10–20 times higher than the oxidative activity. These results indicate that the l-malate dehydrogenase in N. europaea is similar to other NAD+-dependent MDHs (EC 1.1.1.37) but physiologically adapted for its role in a reductive biosynthetic sequence.  相似文献   

14.
A novel enzyme, myo-inositol-1-phosphate dehydrogenase, which catalyzes the conversion of myo-inositol 1-phosphate to ribulose 5-phosphate has been purified 84-fold from mung bean seedling employing several common techniques. The molecular weight of this purified enzyme has been recorded as 88,500 by Sephadex G-200 column chromatography, and in sodium dodecyl sulfate-polyacrylamide gel electrophoresis one protein band containing three subunits of Mr 32,000 each was discernible. Km values for NAD+ and myo-inositol 1-phosphate have been recorded as 2.8 × 10?4 and 5.0 × 10?4m, respectively. Production of NADH in myo-inositol-1-phosphate dehydrogenase reaction has also been evidenced by measurement of NADH fluorescence. Dehydrogenation and decarboxylation of myo-inositol 1-phosphate are mediated by the same enzyme. In fact, the rate of dehydrogenation corroborates with that of decarboxylation. Stoichiometry of this reaction suggests that for the production of 1 mol of ribulose 5-phosphate 2 mol of NAD+ are reduced.  相似文献   

15.
《Process Biochemistry》1999,34(4):375-381
A methylotrophic hydroxypyruvate reductase was partially purified and characterized from Methylophilus spp. using the biomimetic dye, Cibacron Blue F3FA attached to poly(HEMA-EGDMA) microspheres. The absorption capacities of the dye-affinity microspheres were determined by changing pH and the concentration of the proteins in the adsorption medium. Hydroxypyruvate reductase was desorbed from the dye-affinity support specifically with 2 mM NADH solution. The enzyme was purified 10·4-fold with 47% yield. The molecular mass and subunit molecular mass of the enzyme was estimated to be 75 kDa and 37 kDa on the basis of its mobility in polyacrylamide and SDS-polyacrylamide gels, respectively. This suggested a homogeneous dimer structure. The optimal pH was between 5·0 and 7·0, and the maximum enzyme activity was obtained at 50°C. The Km values of hydroxpyruvate reductase were 0·222 mM for hydroxpyruvate and 0·067 mM for NADH.  相似文献   

16.
Tania Bizouarn  Tina Bhakta 《BBA》2005,1708(3):404-410
Transhydrogenase (E.C. 1.6.1.1) couples the redox reaction between NAD(H) and NADP(H) to the transport of protons across a membrane. The enzyme is composed of three components. The dI and dIII components, which house the binding site for NAD(H) and NADP(H), respectively, are peripheral to the membrane, and dII spans the membrane. We have estimated dissociation constants (Kd values) for NADPH (0.87 μM), NADP+ (16 μM), NADH (50 μM), and NAD+ (100-500 μM) for intact, detergent-dispersed transhydrogenase from Escherichia coli using micro-calorimetry. This is the first complete set of dissociation constants of the physiological nucleotides for any intact transhydrogenase. The Kd values for NAD+ and NADH are similar to those previously reported with isolated dI, but the Kd values for NADP+ and NADPH are much larger than those previously reported with isolated dIII. There is negative co-operativity between the binding sites of the intact, detergent-dispersed transhydrogenase when both nucleotides are reduced or both are oxidised.  相似文献   

17.
A comparison of rat brain and liver β-hydroxybutyrate dehydrogenase (EC 1.1.1.30) has revealed that significant differences exist between the enzymes with regard to their kinetic and physical properties. In contrast to the liver enzyme, brain β-hydroxybutyrate dehydrogenase is rapidly inactivated at 46° and is unstable when stored at ?20°. The brain dehydrogenase was found to have a larger Km (apparent) for the 3-acetylpyridine analog of NAD+, and a greater energy of activation in the direction of β-hydroxybutyrate oxidation than the liver enzyme. In the reverse direction, the brain and liver dehydrogenase exhibit substrate inhibition by NADH (0.22 mM and 0.36 mM, respectively). The brain and liver β-hydroxybutyrate dehydrogenase did not differ significantly with regard to the Michaelis-Menten constants measured for NAD+ and β-hydroxybutyrate. The Km constants of brain β-hydroxybutyrate dehydrogenase for acetoacetate (0.39 mM) and NADH (0.05 mM) were lower than those determined for the liver enzyme, acetoacetate (0.73 mM) and NADH (0.35 mM) respectively. These results suggest that the β-hydroxybutyrate dehydrogenase from rat brain and liver are isozymic variants.  相似文献   

18.
Mitochondria from the parasitic helminth, Hymenolepis diminuta, catalyzed both NADPH:NAD+ and NADH:NADP+ transhydrogenase reactions which were demonstrable employing the appropriate acetylpyridine nucleotide derivative as the hydride ion acceptor. Thionicotinamide NAD+ would not serve as the oxidant in the former reaction. Under the assay conditions employed, neither reaction was energy linked, and the NADPH:NAD+ system was approximately five times more active than the NADH:NADP+ system. The NADH:NADP+ reaction was inhibited by phosphate and imidazole buffers, EDTA, and adenyl nucleotides, while the NADPH:NAD+ reaction was inhibited only slightly by imidazole and unaffected by EDTA and adenyl nucleotides. Enzyme coupling techniques revealed that both transhydrogenase systems functioned when the appropriate physiological pyridine nucleotide was the hydride ion acceptor. An NADH:NAD+ transhydrogenase system, which was unaffected by EDTA, or adenyl nucleotides, also was demonstrable in the mitochondria of H. diminuta. Saturation kinetics indicated that the NADH:NAD+ reaction was the product of an independent enzyme system. Mitochondria derived from another parasitic helminth, Ascaris lumbricoides, catalyzed only a single transhydrogenase reaction, i.e., the NADH:NAD+ activity. Transhydrogenase systems from both parasites were essentially membrane bound and localized on the inner mitochondrial membrane. Physiologically, the NADPH:NAD+ transhydrogenase of H. diminuta may serve to couple the intramitochondrial metabolism of malate (via an NADP linked “malic” enzyme) to the anaerobic NADH-dependent ATP-generating fumarate reductase system. In A. lumbricoides, where the intramitochondrial metabolism of malate depends on an NAD-linked “malic” enzyme which is localized primarily in the intermembrane space, the NADH:NAD+ transhydrogenase activity may serve physiologically in the translocation of hydride ions across the inner membrane to the anaerobic energy-generating fumarate reductase system.  相似文献   

19.
Human tissue from uninvolved liver of cancer patients was fractionated using differential centrifugation and characterized for 11βHSD enzyme activity against corticosterone, dehydrocorticosterone, 7α- and 7β-hydroxy-dehydroepiandrosterone, and 7-oxo-dehydroepiandrosterone. An enzyme activity was observed in nuclear protein fractions that utilized either NADP+ or NAD+, but not NADPH and NADH, as pyridine nucleotide cofactor with Km values of 12 ± 2 and 390 ± 2 μM, compared to the Km for microsomal 11βHSD1 of 43 ± 8 and 264 ± 24 μM, respectively. The Km for corticosterone in the NADP+-dependent nuclear oxidation reaction was 102 ± 16 nM, compared to 4.3 ± 0.8 μM for 11βHSD1. The Kcat values for nuclear activity with NADP+ was 1687 nmol/min/mg/μmol, compared to 755 nmol/min/mg/μmol for microsomal 11βHSD1 activity. Inhibitors of 11βHSD1 decreased both nuclear and microsomal enzyme activities, suggesting that the nuclear activity may be due to an enzyme similar to 11βHSD Type 1 and 2.  相似文献   

20.
Procedures are described for isolating highly purified porcine liver pyruvate and α-ketoglutarate dehydrogenase complexes. Rabbit serum stabilized these enzyme complexes in mitochondrial extracts, apparently by inhibiting lysosomal proteases. The complexes were purified by a three-step procedure involving fractionation with polyethylene glycol, pelleting through 12.5% sucrose, and a second fractionation under altered conditions with polyethylene glycol. Sedimentation equilibrium studies gave a molecular weight of 7.2 × 106 for the liver pyruvate dehydrogenase complex. Kinetic parameters are presented for the reaction catalyzed by the pyruvate dehydrogenase complex and for the regulatory reactions catalyzed by the pyruvate dehydrogenase kinase and pyruvate dehydrogenase phosphatase. For the overall catalytic reaction, the competitive Ki to Km ratio for NADH versus NAD+ and acetyl CoA versus CoA were 4.7 and 5.2, respectively. Near maximal stimulations of pyruvate dehydrogenase kinase by NADH and acetyl CoA were observed at NADH:NAD+ and acetyl CoA:CoA ratios of 0.15 and 0.5, respectively. The much lower ratios required for enhanced inactivation of the complex by pyruvate dehydrogenase kinase than for product inhibition indicate that the level of activity of the regulatory enzyme is not directly determined by the relative affinity of substrates and products of catalytic sites in the pyruvate dehydrogenase complex. In the pyruvate dehydrogenase kinase reaction, K+ and NH+4 decreased the Km for ATP and the competitive inhibition constants for ADP and (β,γ-methylene)adenosine triphosphate. Thiamine pyrophosphate strongly inhibited kinase activity. A high concentration of ADP did not alter the degree of inhibition by thiamine pyrophosphate nor did it increase the concentration of thiamine pyrophosphate required for half-maximal inhibition.  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