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 546 毫秒
1.
The kinetic mechanism of NADP(+)-dependent 3 alpha-hydroxysteroid dehydrogenase and NAD(+)-dependent 3 alpha(17 beta)-hydroxysteroid dehydrogenase, purified from hamster liver cytosol, was studied in both directions. For 3 alpha-hydroxysteroid dehydrogenase, the initial velocity and product inhibition studies indicated that the enzyme reaction sequence is ordered with NADP+ binding to the free enzyme and NADPH being the last product to be released. Inhibition patterns by Cibacron blue and hexestrol, and binding studies of coenzyme and substrate are also consistent with an ordered bi bi mechanism. For 3 alpha(17 beta)-hydroxysteroid dehydrogenase, the steady-state kinetic measurements and substrate binding studies suggest a random binding pattern of the substrates and an ordered release of product; NADH is released last. However, the two enzymes transferred the pro-R-hydrogen atom of NAD(P)H to the carbonyl substrate.  相似文献   

2.
The kinetic mechanism of two major monomeric 17 beta-hydroxysteroid dehydrogenases from mouse liver cytosol was studied at pH 7 in both directions with NADP(H) and three steroid substrates: testosterone, 5 beta-androstane-3 alpha, 17 17 beta-diol, and estradiol-17 beta. In each case the reaction mechanism of the two enzymes was sequential, and inhibition patterns by-products and dead-end inhibitors were consisted with an ordered bi bi mechanism with the coenzyme binding to the free enzyme, although there was difference in affinity and maximum velocity for the steroidal substrates between the two enzymes. Binding studies of the coenzyme and substrate indicate the binding of coenzyme to the free enzyme, in which 1 mol of NADPH binds to 1 mol of each monomeric enzyme. The 4-pro-R-hydrogen atom of NADPH was transferred to the alpha-face of the steroid molecule by the two enzymes.  相似文献   

3.
The kinetic mechanism of the major sheep liver aldehyde reductase (ALR1) was studied with three aldehyde substrates: p-nitrobenzaldehyde, pyridine-3-aldehyde and D-glucuronate. In each case the enzyme mechanism was sequential and product-inhibition studies were consistent with an ordered Bi Bi mechanism, with the coenzymes binding to the free enzyme. Binding studies were used to investigate the interactions of substrates, products and inhibitors with the free enzyme. These provided evidence for the binding of D-glucuronate, L-gulonate and valproate, as well as NADP+ and NADPH. The enzyme was inhibited by high concentrations of D-glucuronate in a non-competitive manner, indicating that this substrate was able to bind to the free enzyme and to the E X NADP+ complex at elevated concentrations. Although the enzyme was inhibited by high pyridine-3-aldehyde concentrations, there was no evidence for the binding of this substrate to the free enzyme. Sheep liver ALR1 was inhibited by the ionized forms of alrestatin, sorbinil, valproate, 2-ethylhexanoate and phenobarbitone, indicating the presence of an anion-binding site similar to that described for the pig liver enzyme, which interacts with inhibitors and substrates containing a carboxy group. Sorbinil, valproate and 2-ethylhexanoate inhibited the enzyme uncompetitively at low concentrations and non-competitively at high concentrations, whereas phenobarbitone and alrestatin were non-competitive and uncompetitive inhibitors respectively. The significance of these results with respect to inhibitor and substrate binding is discussed.  相似文献   

4.
Chicken liver mitochondria were isolated in relatively pure form as indicated by electron microscopy and marker enzyme assay. The rate of respiration, respiratory control index and ADP/O ratios with several different substrates indicated that chicken liver mitochondria are more uncoupled than rat liver mitochondria. Chickens have ten-fold higher malate concentrations in liver than do rats, 2-oxoglutarate was also more abundant in chicken livers. Fasted birds had a five-fold increase in beta-hydroxybutyrate as compared with fed birds; whereas malate and lactate concentrations decreased. Fasted birds had increased levels of isocitrate dehydrogenase (NADP dependent) and lactate dehydrogenase in the cytosol, and increased malate dehydrogenase (NAD dependent), isocitrate dehydrogenase (NADP dependent) and malic enzyme activities in the mitochondria.  相似文献   

5.
Indanol dehydrogenase was purified to apparent homogeneity from monkey liver cytosol. The enzyme was a monomer with a molecular weight of 36,000 and pI of 8.7. The amino acid composition was determined. The enzyme oxidized alicyclic alcohols including transdihydrodiols of benzene and naphthalene in the presence of both NADP+ and NAD+, and reduced several xenobiotic carbonyl compounds in the presence of NADPH, the 4-pro-R hydrogen atom of which was transferred to the substrate. The results of fluorometric binding and kinetic studies are consistent with an ordered sequential mechanism with NADP+ binding first. The enzyme was inhibited competitively versus NADP+ and uncompetitively versus 1-indanol not only by chelating agents such as 1,10-phenanthroline and 2,2'-bipyridine but also by a nonchelating isomer, 4,4'-bipyridine, which suggests hydrophobic interaction of the aromatic compounds with the enzyme, which did not contain zinc. The enzyme was also inhibited by Cibacron blue dye, synthetic estrogens, and delta 4-3-ketosteroids. The inhibition by Cibacron blue was competitive versus NADP+ and noncompetitive versus 1-indanol, whereas those by hexestrol, medroxyprogesterone acetate, and progesterone were uncompetitive versus NADP+ and competitive versus 1-indanol, corraborating the ordered addition of the coenzyme prior to 1-indanol.  相似文献   

6.
The nicotinamide adenine dinucleotide phosphate (NADP)-dependent formate dehydrogenase in Clostridium thermoaceticum used, in addition to its natural electron acceptor, methyl and benzyl viologen. The enzyme was purified to a specific activity of 34 (micromoles per minute per milligram of protein) with NADP as electron acceptor. Disc gel electrophoresis of the purified enzyme yielded two major and two minor protein bands, and during centrifugation in sucrose gradients two components of apparent molecular weights of 270,000 and 320,000 were obtained, both having formate dehydrogenase activity. The enzyme preparation catalyzed the reduction of riboflavine 5'-phosphate flavine adenine dinucleotide and methyl viologen by using reduced NADP as a source of electrons. It also had reduced NADP oxidase activity. The enzyme was strongly inhibited by cyanide and ethylenediaminetetraacetic acid. It was also inhibited by hypophosphite, an inhibition that was reversed by formate. Sulfite inhibited the activity with NADP but not with methyl viologen as acceptor. The apparent K(m) at 55 C and pH 7.5 for formate was 2.27 x 10(-4) M with NADP and 0.83 x 10(-4) with methyl viologen as acceptor. The apparent K(m) for NADP was 1.09 x 10(-4) M and for methyl viologen was 2.35 x 10(-3) M. NADP showed substrate inhibition at 5 x 10(-3) M and higher concentrations. With NADP as electron acceptor, the enzyme had a broad pH optimum between 7 and 9.5. The apparent temperature optimum was 85 C. In the absence of substrates, the enzyme was stable at 70 C but was rapidly inactivated at temperatures above 73 C. The enzyme was very sensitive to oxygen but was stabilized by thiol-iron complexes and formate.  相似文献   

7.
NAD(+)-linked and NADP(+)-linked 3 alpha-hydroxysteroid dehydrogenases were purified to homogeneity from hamster liver cytosol. The two monomeric enzymes, although having similar molecular masses of 38,000, differed from each other in pI values, activation energy and heat stability. The two proteins also gave different fragmentation patterns by gel electrophoresis after digestion with protease. The NADP(+)-linked enzyme catalysed the oxidoreduction of various 3 alpha-hydroxysteroids, whereas the NAD(+)-linked enzyme oxidized the 3 alpha-hydroxy group of pregnanes and some bile acids, and the 17 beta-hydroxy group of testosterone and androstanes. The thermal stabilities of the 3 alpha- and 17 beta-hydroxysteroid dehydrogenase activities of the NAD(+)-linked enzyme were identical, and the two enzyme activities were inhibited by mixing 17 beta- and 3 alpha-hydroxysteroid substrates, respectively. Medroxyprogesterone acetate, hexoestrol and 3 beta-hydroxysteroids competitively inhibited 3 alpha- and 17 beta-hydroxysteroid dehydrogenase activities of the enzyme. These results show that hamster liver contains a 3 alpha(17 beta)-hydroxysteroid dehydrogenase structurally and functionally distinct from 3 alpha-hydroxysteroid dehydrogenase.  相似文献   

8.
1. The kinetics of the thermally induced enzyme variants of the supernatant NADP-isocitrate dehydrogenase from rainbow-trout liver are investigated. 2. Fish acclimatized to 2 degrees C (cold-adapted enzyme) and 17 degrees C (warm-adapted enzyme) show different relative distributions of the three NADP-isocitrate dehydrogenase isoenzymes; this has been demonstrated with electrophoresis and electrofocusing techniques. 3. Plots of K(m) versus temperature for the cold-adapted and warm-adapted enzyme variants are complex in nature with apparent maximal enzyme-substrate affinity corresponding to the temperature at which the trout is acclimatized. Both substrates, dl-isocitrate and NADP(+), give similar curves although the magnitude of the K(m) change with temperature is much decreased in the case of NADP(+). 4. E(a) values of approx. 18kcal/mol were determined for both the cold-adapted and warm-adapted enzyme variants. 5. In an attempt to determine how velocities can be increased at low temperatures, cation, pH requirements, metabolite and enzyme concentrations were examined. 6. NAD-isocitrate dehydrogenase could not be detected in trout tissues.  相似文献   

9.
Inactive NADP-malate dehydrogenase (disulfide form) from chloroplasts of Zea mays is activated by reduced thioredoxin while the active enzyme (dithiol form) is inactivated by incubation with oxidized thioredoxin. This reductive activation of NADP-malate dehydrogenase is inhibited by over 95% in the presence of NADP and the Kd for this interaction of NADP with the inactive enzyme is about 3 microM. Other substrates of the enzyme (malate, oxaloacetate, or NADPH) do not effect the rate of enzyme activation but NADPH can reverse the inhibitory effect of NADP. It appears that NADPH (Kd = 250 microM) and NADP (Kd = 3 microM) compete for the same site, presumably the coenzyme-binding site at the active centre. Apparently the enzyme . NADP binary complex cannot be reduced by thioredoxin whereas the enzyme . NADPH complex is reduced at the same rate as is the free enzyme. Similarly the oxidative inactivation of reduced NADP-malate dehydrogenase is inhibited by up to 85% by NADP and NADPH completely reverses this inhibition. The Kd values of the active-reduced enzyme for NADP and NADPH were both estimated to be 30 microM. From these data a model was constructed which predicts how changing NADPH/NADP levels in the chloroplast might change the steady-state level of NADP-malate dehydrogenase activity. The model indicates that at any fixed ratio of reduced to oxidized thioredoxin high proportions of active NADP-malate dehydrogenase and, hence, high rates of oxaloacetate reduction, can only occur with very high NADPH/NADP ratios.  相似文献   

10.
Coenzyme specificity of mammalian liver D-glycerate dehydrogenase   总被引:1,自引:0,他引:1  
D-Glycerate dehydrogenase (glyoxylate reductase) was partially purified from rat liver by anion- and cation-exchange chromatography. When assayed in the direction of D-glycerate or glycolate formation, the enzyme was inhibited by high (greater than or equal to 0.5 mM), unphysiological concentrations of hydroxypyruvate or glyoxylate much more potently in the presence of NADPH than in the presence of NADH. However, the dehydrogenase displayed a much greater affinity for NADPH (Km less than 1 microM) than for NADH (Km = 48-153 microM). Furthermore, NADP was over 1000-fold more potent than NAD in inhibiting the enzyme competitively with respect to NADH. NADP also inhibited the reaction competitively with respect to NADPH whereas NAD, at concentrations of up to 10 mM had no inhibitory effect. When measured by the formation of hydroxypyruvate from D-glycerate, the enzyme also displayed a much greater affinity for NADP than for NAD. These properties indicate that liver D-glycerate dehydrogenase functions physiologically as an NADPH-specific reductase. In agreement with this conclusion, the addition of hydroxypyruvate or glyoxylate to suspensions of rat hepatocytes stimulated the pentose-phosphate pathway. The coenzyme specificity of D-glycerate dehydrogenase is discussed in relation to the biochemical findings made in D-glyceric aciduria and in primary hyperoxaluria type II (L-glyceric aciduria).  相似文献   

11.
1. An NADP+-dependent isocitrate dehydrogenase was extracted from turbot liver. The enzyme required divalent cations (Mg2+ or Mn2+) for its activity but was inhibited by high salt concentrations. 2. The enzyme had an optimum activity in the pH range between 7.5 and 9.0. The enzymic activity increased with temperature (in the range of 5 to 68 degrees C) with an Ea of 23.5 kJ/mol and a Q10 of 1.34. 3. The apparent Km values for the substrates were 6.5 microM for NADP+, 56 microM for Mg2+, 87 microM for Mn2+ and 4.2 and 73.5 microM for the complexes Mg-isocitrate and Mn-isocitrate, respectively. The physiological significance of these results is discussed. 4. The enzyme was inhibited by citrate and adenine nucleotides. The degree of inhibition depended on the relation between the concentrations and that of magnesium. A possible regulating mechanism is proposed.  相似文献   

12.
1. Hydroxyoxoglutarate was obtained by three methods: decarboxylation of oxalomalic acid, and synthesis from glyoxylate and pyruvate by using either Mg2+ or an enzyme from rat liver as catalysts. 2. The inhibitory effects of oxalomalate and hydroxyoxoglutarate upon aconitate hydratase, isocitrate dehydrogenase (NADP) and oxoglutarate dehydrogenase were investigated. 3. Oxalomalate at low concentrations (1mm) inhibited almost completely both aconitate hydratase and isocitrate dehydrogenase. Hydroxyoxoglutarate also inhibited these enzymes, but at concentrations approximately tenfold that of oxalomalate. 4. Oxalomalate and hydroxyoxoglutarate, at the higher concentrations, inhibited oxoglutarate dehydrogenase to approximately the same extent. 5. It is suggested that the ability of glyoxylate to control reaction rates in the tricarboxylic acid cycle must in some degree be due to its condensation with oxaloacetate and pyruvate to form enzyme inhibitors.  相似文献   

13.
The steady-state kinetics of the oxidative decarboxylation of 6-phosphogluconate catalysed by 6-phosphogluconate dehydrogenase from sheep liver in triethanolamine and phosphate buffers (pH 7.0) have been reinvestigated. In triethanolamine buffer the enzyme is inhibited by high NADP+ concentrations in the presence of low fixed concentrations of 6-phosphogluconate. Data are consistent with an asymmetric sequential mechanism in which NADP+ and 6-phosphogluconate bind randomly and product release is ordered. The pathway through the enzyme--6-phosphogluconate complex appears to be preferred in triethanolamine buffer. Pre-steady-state studies of the oxidative decarboxylation reaction at pH 6.0-8.0 show that hydride transfer is greater than 900 s-1. After the fast formation of NADPH in amounts equivalent to about half of the enzyme-active-centre concentration, the rate of NADPH formation is equal to the steady-state rate. Two possible interpretations are considered. Rapid fluorescence measurements of the displacement of NADPH from its complex with the enzyme at pH 6.0 and 7.0 indicate that the dissociation of NADPH is fast (greater than 800 s-1) and cannot be the rate-limiting step in oxidative decarboxylation. Coenzyme binding studies at equilibrium have been extended to include the determination of the dissociation constants for the binary complexes of enzyme with NADPH and NADP+ at pH 6.0-8.0 and the dissociation constant for NADPH in the ternary enzyme--6-phosphogluconate--NADPH complex in triethanolamine buffer, pH 7.0.  相似文献   

14.
Bovine liver D-glycerate dehydrogenase (D-glycerate:NAD (NADP) oxidoreductase, EC 1.1.1.29) adapts its kinetic behaviour to a sequential mechanism. The presence of NaCl causes an appreciable variation in the Km and V values. relative to the both substrates in the hydroxypyruvate/D-glycerate dehydrogenase/NADH system, which does not happen in the D-glycerate/D-glycerate dehydrogenase/NAD system. The former system is inhibited by high concentrations of NaCl and activated by low salt concentrations. The hydroxypyruvate concentration causing substrate inhibition increases as the concentration of NaCl increases; excess NADH inhibition is independent of the salt concentration. The variation of the initial rates of both systems, in the presence of chlorides having monovalent and divalent cations, or sodium halides, Na2SO4 and NaNO3 (at constant ionic strength) suggests that the anions have a specific action on the enzyme. An increase in the NaCl concentration causes a displacement of the optimum D-glycerate dehydrogenase pH (with hydroxypyruvate and NADH as substrates) towards the acid area. The enzyme stability, at varying pH, varies with the salt concentration.  相似文献   

15.
Alcohol dehydrogenase (E. C. 1.1.1.1) from Thermoanaerobium brockii at 25 degrees C and at 65 degrees C is more active with secondary than primary alcohols. The enzyme utilizes NADP and NADPH as cosubstrates better than NAD and NADH. The maximum velocities (V(m)) for secondary alcohols at 65 degrees C are 10 to 100 times higher than those at 25 degrees C, whereas the K(m) values are more comparable.At both 25 degrees C and 65 degrees C the substrate analogue 1,1,1,3,3,3-hexafluoro-2-propanol inhibited the oxidation of alcohol competitively with respect to cyclopentanol, and uncompetitively with respect to NADP. Dimethylsulfoxide inhibited the reduction of cyclopentanone competitively with respect to cyclopentanone, and uncompetitively with respect to NADPH. As a product inhibitor, NADP was competitive with respect to NADPH. These results demonstrate that the enzyme binds the nucleotide and then the alcohol or ketone to form a ternary complex which is converted to a product ternary complex that releases product and nucleotide in that order.At 25 degrees C, all aldehydes and ketones examined inhibited the enzyme at concentrations above their Michaelis constants. The substrate inhibition by cyclopentanone was incomplete, and it was uncompetitive with respect to NADPH. Furthermore, cyclopentanone as a product inhibitor showed intercept-linear, slope-parabolic inhibition with respect to cyclopentanol. These results indicate that cyclopentanone binds to the enzyme-NADP complex at high concentrations. The resulting ternary complex slowly dissociates NADP and cyclopentanone.At 65 degrees C, all of the secondary alcohols, with the exception of cyclohexanol, show substrate activation at high concentration. Experiments in which NADP was the variable substrate and cyclopentanol as the constant-variable substrate over a wide range of concentrations gave double reciprocal plots in which the intercepts showed substrate activation and the slopes showed substrate inhibition. These results indicate that the secondary alcohols bind to the enzyme-NADPH complex at high concentrations and that the resulting ternary complex dissociates NADPH faster than the enzyme-NADPH complex. (c) 1993 John Wiley & Sons, Inc.  相似文献   

16.
Regulation of the pentose phosphate cycle   总被引:25,自引:12,他引:13       下载免费PDF全文
1. A search was made for mechanisms which may exert a ;fine' control of the glucose 6-phosphate dehydrogenase reaction in rat liver, the rate-limiting step of the oxidative pentose phosphate cycle. 2. The glucose 6-phosphate dehydrogenase reaction is expected to go virtually to completion because the primary product (6-phosphogluconate lactone) is rapidly hydrolysed and the equilibrium of the joint dehydrogenase and lactonase reactions is in favour of virtually complete formation of phosphogluconate. However, the reaction does not go to completion, because glucose 6-phosphate dehydrogenase is inhibited by NADPH (Neglein & Haas, 1935). 3. Measurements of the inhibition (which is competitive with NADP(+)) show that at physiological concentrations of free NADP(+) and free NADPH the enzyme is almost completely inhibited. This indicates that the regulation of the enzyme activity is a matter of de-inhibition. 4. Among over 100 cell constituents tested only GSSG and AMP counteracted the inhibition by NADPH; only GSSG was highly effective at concentrations that may be taken to occur physiologically. 5. The effect of GSSG was not due to the GSSG reductase activity of liver extracts, because under the test conditions the activity of this enzyme was very weak, and complete inhibition of the reductase by Zn(2+) did not abolish the GSSG effect. 6. Preincubation of the enzyme preparation with GSSG in the presence of Mg(2+) and NADP(+) before the addition of glucose 6-phosphate and NADPH much increased the GSSG effect. 7. Dialysis of liver extracts and purification of glucose 6-phosphate dehydrogenase abolished the GSSG effect, indicating the participation of a cofactor in the action of GSSG. 8. The cofactor removed by dialysis or purification is very unstable. The cofactor could be separated from glucose 6-phosphate dehydrogenase by ultrafiltration of liver homogenates. Some properties of the cofactor are described. 9. The hypothesis that GSSG exerts a fine control of the pentose phosphate cycle by counteracting the NADPH inhibition of glucose 6-phosphate dehydrogenase is discussed.  相似文献   

17.
Dihydrodiol dehydrogenase activity was detected in the cytosol of various mouse tissues, among which kidney exhibited high specific activity comparable to the value for liver. The enzyme activity in the kidney cytosol was resolved into one major and three minor peaks by Q-Sepharose chromatography: one minor form cross-reacted immunologically with hepatic 3 alpha-hydroxysteroid dehydrogenase and another with aldehyde reductase. The other minor form was partially purified and the major form was purified to homogeneity. These two forms, although different in their charges, were monomeric proteins with the same molecular weight of 39,000 and had similar catalytic properties. They oxidized cis-benzene dihydrodiol and alicyclic alcohols as well as trans-dihydrodiols of benzene and naphthalene in the presence of NADP+ or NAD+, and reduced several xenobiotic aldehydes and ketones with NAD(P)H as a cofactor. The enzymes also catalyzed the oxidation of 3 alpha-hydroxysteroids and epitestosterone, and the reduction of 3- and 17-ketosteroids, showing much lower Km values (10(-7)-10(-6) M) for the steroids than for the xenobiotic alcohols. The results of mixed substrate experiments, heat stability, and activity staining on polyacrylamide gel electrophoresis suggested that, in the two enzymes, both dihydrodiol dehydrogenase and 3(17)alpha-hydroxysteroid dehydrogenase activities reside on a single enzyme protein. Thus, dihydrodiol dehydrogenase existed in four forms in mouse kidney cytosol, and the two forms distinct from the hepatic enzymes may be identical to 3(17)alpha-hydroxysteroid dehydrogenases.  相似文献   

18.
20-Hydroxyleukotriene B4 was converted by rat liver homogenates in the presence of NAD+ to a more polar product on reverse-phase high-performance liquid chromatography. The product was identified as 20-carboxyleukotriene B4 by straight-phase high performance liquid chromatography, ultraviolet spectrophotometry and gas chromatography-mass spectrometry. The oxidative activity of the homogenates was located in the cytosol with an optimal pH of 8.0. The activity was dependent on NAD+, and NADP+ could not substitute for NAD+. 1 mol of 20-carboxyleukotriene B4 was formed with the reduction of 2 mol of NAD+. The reaction was inhibited by pyrazole and 4-methylpyrazole, inhibitors of alcohol dehydrogenase, and by various alcohols, such as ethanol, 12-hydroxylaurate, and 20-hydroxyprostaglandin E1. Disulfiram, an inhibitor of aldehyde dehydrogenase, also inhibited the activity. These results suggest that two discrete steps catalyzed by different enzymes, alcohol dehydrogenase and aldehyde dehydrogenase, are involved in the oxidation of 20-hydroxyleukotriene B4 in rat liver cytosol. The enzyme system seems to be different from that of human neutrophils.  相似文献   

19.
1. N10-Formyltetrahydrofolate dehydrogenase was purified to homogeneity from rat liver with a specific activity of 0.7--0.8 unit/mg at 25 degrees C. The enzyme is a tetramer (Mw = 413,000) composed of four similar, if not identical, substrate addition and give the Km values as 4.5 micron [(-)-N10-formyltetrahydrofolate] and 0.92 micron (NADP+) at pH 7.0. Tetrahydrofolate acts as a potent product inhibitor [Ki = 7 micron for the (-)-isomer] which is competitive with respect to N10-formyltetrahydrofolate and non-competitive with respect to NADP+. 3. Product inhibition by NADPH could not be demonstrated. This coenzyme activates N10-formyltetrahydrofolate dehydrogenase when added at concentrations, and in a ratio with NADP+, consistent with those present in rat liver in vivo. No effect of methionine, ethionine or their S-adenosyl derivatives could be demonstrated on the activity of the enzyme. 4. Hydrolysis of N10-formyltetrahydrofolate is catalysed by rat liver N10-formyltetrahydrofolate dehydrogenase at 21% of the rate of CO2 formation based on comparison of apparent Vmax. values. The Km for (-)-N10-folate is a non-competitive inhibitor of this reaction with respect to N10-formyltetrahydrofolate, with a mean Ki of 21.5 micron for the (-)-isomer. NAD+ increases the maximal rate of N10-formyltetrahydrofolate hydrolysis without affecting the Km for this substrate and decreases inhibition by tetrahydrofolate. The activator constant for NAD+ is obtained as 0.35 mM. 5. Formiminoglutamate, a product of liver histidine metabolism which accumulates in conditions of excess histidine load, is a potent inhibitor of rat liver pyruvate carboxylase, with 50% inhibition being observed at a concentration of 2.8 mM, but has no detectable effect on the activity of rat liver cytosol phosphoenolpyruvate carboxykinase measured in the direction of oxaloacetate synthesis. We propose that the observed inhibition of pyruvate carboxylase by formiminoglutamate may account in part for the toxic effect of excess histidine.  相似文献   

20.
Dihydrouracil dehydrogenase (NADP+) (EC 1.3.1.2) was partially purified from the cytosol fraction of rat liver and fractionated by disc gel electrophoresis. A major and minor band were visualized by staining for enzyme activity. The substrate specificity of these bands was investigated. It was found that both bands were two to three times more active with dihydrothymine as substrate than with dihydrouracil in the presence of NADP+ and the optimum pH of 7.4. Mitochondrial fractions containing most of the NADH-dependent uracil reductase of rat liver cells were fractionated by centrifugation in sucrose density gradients. Two procedures involving linear or discontinuous gradients were used. By both, good separation of NADH- and NADPH- dependent reductases was achieved. Marker enzyme studies supported the view that the NADH-dependent enzyme is located principally in mitochondria whereas the NADPH-dependent enzyme is mainly in plasma and endoplasmic reticulum membranes. For the NADH-dependent reductase the apparent Km for thymine at pH 7.4 was 1.39 times that found for uracil whereas for the NADPH-dependent enzyme the apparent Km values were similar for the two substrates at this pH. Dihydrouracil was the principal product isolated by paper chromatography from the reaction mixture containing a partially purified fraction of mitochondria, uracil and NADH at pH 7.4. This fraction also catalyzed the formation of radioactive carbon dioxide from [2-14C]uracil. The proportion of CO2 formed by the mitochondria was about 10% of that formed by the original homogenate.  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