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.
A kinetic analysis of Drosophila melanogaster dopa decarboxylase   总被引:1,自引:0,他引:1  
The kinetic mechanism of dopa decarboxylase (3,4-dihydroxy-L-phenylalanine carboxy-lyase, EC 4.1.1.28) was investigated in Drosophila melanogaster. Based on initial velocity and product inhibition studies, an ordered reaction is proposed for dopa decarboxylase. This kinetic mechanism is interpreted in the context of measured enzyme activities and the catecholamine pools in Drosophila. The 1(2)amd gene is immediately adjacent to the gene coding for dopa decarboxylase (Ddc) and determines hypersensitivity to alpha-methyldopa in Drosophila. Dopa decarboxylase does not decarboxylate alpha-methyldopa and hence does not generate a toxic product capable of inhibiting 1(2)amd gene function. We propose that the 1(2)amd gene is involved with an unknown catecholamine pathway involving dopa but not dopamine.  相似文献   

2.
Aromatic-L-amino-acid decarboxylase activity in mouse pancreatic islets   总被引:2,自引:0,他引:2  
Aromatic-L-amino-acid decarboxylase activity has been measured in intact or homogenised pancreatic islets of ob/ob mice (Ume? ob/ob). The method used involves the trapping and measuring of the 14CO2 released from L-[1-14C]dihydroxyphenylalanine (L-dopa). Islets showed a decarboxylase activity which was dependent on pyridoxal phosphate and inhibitable by 0.1 mM benserazide or 0.1 mM alpha-monofluoromethyldopa. Maximum activity in intact islets was about 330 mmol/kg dry islet per h with an apparent Km of 3.3 mM. Islet homogenates had a Vmax of about 120 mmol/kg per h with a Km of 0.3 mM. L-5-Hydroxytryptophan, m-tyrosine and o-tyrosine interfered with the decarboxylation of L-dopa in a way that suggested a high activity also towards those substrates. L-Phenylalanine, L-tyrosine and D-glucose had no effect. At 0.05 mM L-dopa islet homogenates showed a much higher activity than homogenates of liver, kidney, or spleen. Islet uptake of L-[3H]dopa was well in excess of the decarboxylation rate and thus probably not rate-limiting. It is concluded that mouse pancreatic islets have a high activity of aromatic-L-amino-acid decarboxylase. This is in accordance with previous suggestions of a stimulatory effect of this enzyme on insulin secretion.  相似文献   

3.
The study of DOPA (3,4-dihydroxyphenylalanine) decarboxylase by steady-state methods is difficult because multiple reactions occur. The reaction with DOPA was studied at enzyme concentrations between 20 and 50 micrometer by direct observation of the bound coenzyme by using stopped-flow and conventional spectrophotometry. Four processes were observed on different time scales and three of these were attributed to stages in the decarboxylation. The fourth was attributed to an accompanying transamination that renders the enzyme inactive. It was clear that much, if not all, of the 330 nm-absorbing coenzyme present in the free enzyme plays an active part in the decarboxylation, since it is converted into 420 nm-absorbing material in the first observable step. An intermediate absorbing maximally at 390 nm is formed in a slower step. Rate and equilibrium constants have been determined and the ratio of decarboxylation to transamination was estimated to be 1200:1.  相似文献   

4.
5.
Newly deposited fertilized and unfertilized Aedes aegypti eggs are soft and white. Within a short time they darken and harden. Injection of a potent dopa decarboxylase inhibitor (dl)-3-(3,4-dihydroxyphenyl)-2-hydrazino-2-methylpropionic acid (α-MDH) into females, subsequent to a blood meal, results in oviposited eggs which are pale in colour. Moreover, such fertilized eggs do not hatch. The severity of both effects seems to be positively correlated and is dependent upon the time of α-MDH injection.Extracts of mature ovaries are capable of converting dopa to dopamine in the absence of a pretreatment with α-MDH. Mature ovaries obtained from females who had been previously injected with α-MDH could not accomplish this conversion. The inhibitor does not seem to have any effect on dopa oxidase activity and subsequent melanization. We conclude that dopamine is synthesized by blood-fed females via decarboxylation of dopa by dopa decarboxylase and propose that the normal darkening and hardening of A. aegypti eggs is a result of sclerotization.  相似文献   

6.
Klebsiella aerogenes adapted to a chemically-defined mineral salts medium with glucose orp-hydroxybenzoate as sole source of carbon and energy possessed constitutive decarboxylases for gentisate (2,5-dihydroxybenzoate), protocatechuate (3,4-dihydroxybenzoate) and gallate (3,4,5-trihydroxybenzoate) whose pH optima were respectively 5.9, 5.6 and 5.8. A decarboxylase for PHB was induced by PHB in both growing and resting cells; the induction was delayed or inhibited by chloramphenicol and by ultrasonic disruption of the bacteria. Crude ultrasonic preparations of PHB decarboxylase had an optimum pH of 6.0, a Michaelis constant of 4mm and an activation energy of 25,500 cal mole–1 at 28 – 38 C. All four decarboxylations proceeded without O2 and for every mole of phenolic acid decomposed one mole of CO2 and one mole of the corresponding phenol were produced. The effects of ultrasonic disruption of the bacteria suggested that permeability barriers limited the rate of decarboxylation of PHB and 2,5-DHB but not of 3,4-DHB or 3,4,5-THB. During ultrasonic disintegration PHB and 3,4-DHB decarboxylases were retained solely by insoluble centrifugeable particles, whereas 2,5-DHB and 3,4,5-THB decarboxylases were gradually released into solution.The decarboxylation of protocatechuic acid is an essential stage in the assimilation ofp-hydroxybenzoic acid byK. aerogenes, whereas the decarboxylation ofp-hydroxybenzoate itself is an injurious side reaction.We wish to thank Mr. P. J. Wragg for technical assistance.  相似文献   

7.
A method is presented for the assay, by high-performance liquid chromatography, of l-aromatic amino acid decarboxylase. The Km value of the enzyme for 3,4-dihydroxyphenylalanine (dopa) was, by this procedure, 0.16 mm, in good agreement with previous reports. When α-methyldopa was used as substrate, evidence was obtained indicating the formation of, besides α-methyldopamine, 3,4-dihydroxyphenylacetone, presumably through an internal transamination process.  相似文献   

8.
Transaminations catalysed by brain glutamate decarboxylase.   总被引:7,自引:0,他引:7       下载免费PDF全文
In addition to normal decarboxylation of glutamate to 4-aminobutyrate, glutamate decarboxylase from pig brain was shown to catalyse decarboxylation-dependent transamination of L-glutamate and direct transamination of 4-aminobutyrate with pyridoxal 5'-phosphate to yield succinic semialdehyde and pyridoxamine 5'-phosphate in a 1:1 stoichiometric ratio. Both reactions result in conversion of holoenzyme into apoenzyme. With glutamate as substrate the rates of transamination differed markedly among the three forms of the enzyme (0.008, 0.012 and 0.029% of the rate of 4-aminobutyrate production by the alpha-, beta- and gamma-forms at pH 7.2) and accounted for the differences among the forms in rates of inactivation by glutamate and 4-aminobutyrate. Rates of transamination were maximal at about pH 8 and varied in parallel with the rate constants for inactivation from pH 6.5 to 8.0. Rates of transamination of glutamate and 4-aminobutyrate were similar, suggesting that the decarboxylation step is not entirely rate-limiting in the normal mechanism. The transamination was reversible, and apoenzyme could be reconstituted to holoenzyme by reverse transamination with succinic semialdehyde and pyridoxamine 5'-phosphate. As a major route of apoenzyme formation, the transamination reaction appears to be physiologically significant and could account for the high proportion of apoenzyme in brain.  相似文献   

9.
10.
Uroporphyrinogen decarboxylase (UROD) and coproporphyrinogen oxidase (copro'gen oxidase) are two of the least well understood enzymes in the heme biosynthetic pathway. In the fifth step of the pathway, UROD converts uroporphyrinogen III to coproporphyrinogen III by the decarboxylation of the four acetic acid side chains. Copro'gen oxidase then converts coproporphyrinogen III to protoporphyrinogen IX via two sequential oxidative decarboxylations. Studies of these two enzymes are important to increase our understanding of their mechanisms. Assay comparisons of UROD and copro'gen oxidase from chicken blood hemolysates (CBH), using a newly developed micro-assay, showed that the specific activity of both enzymes is increased in the micro-assay relative to the large-scale assay. The micro-assay has distinct advantages in terms of cost, labor intensity, amount of enzyme required, and sensitivity.  相似文献   

11.
It is shown that most ornithine in a chicken liver homogenate is decarboxylated in the particulate fraction. This fraction, however, requires the cytosol for complete activity. The dialyzed supernatant does not activate decarboxylation of ornithine, while the supernatant is more effective when previously inactivated at 100 degrees C. The supernatant can be substituted by the intermediates of the citric acid cycle (oxaloacetate, citrate, succinate, malate), by pyruvate, and partially by ADP as well. Rotenone blocks decarboxylation suggesting that this occurs through the pathway ornithine leads to glutamic semialdehyde leads to glutamate leads to alpha-ketoglutarate, which in turn is decarboxylated. The activating metabolites would thus have a role in reoxidizing NADH, and the ketoacids also in supplying the acceptor for transamination of glutamate, and indirectly for ornithine transamination. Pyruvate and oxaloacetate do not transaminate with ornithine. Insulin promotes a marked increase of cytosol ornithine decarboxylase activity, but has little effect on decarboxylation by the particulate cellular fraction.  相似文献   

12.
LARGE doses of 3,4-dihydroxyphenylalanine (dopa) cause increased locomotor activity in rats and mice pretreated with a peripherally acting decarboxylase inhibitor1–3. The effect of dopa is enhanced in animals pretreated with a mono-amine oxidase inhibitor4–7 and reduced when the decarboxylation of dopa in the brain is inhibited1. Consequently, the increase in motor activity is thought to be due to the formation of catecholamines in the brain.  相似文献   

13.
1. When homogenates of brains from mature adult locusts (Locusta migratoria) were incubated with l-3-(3,4-dihydroxyphenyl)[3-(14)C]alanine the major radioactive metabolite was dopamine, suggesting the presence of a dopa (3,4-dihydroxyphenylalanine) decarboxylase. 2. Decarboxylation of l-dopa by this tissue, measured under optimum conditions by a radiochemical method, was 21mumol of CO(2)/h per g wet wt. Apparent decarboxylation of l-tyrosine proceeded at 0.34mumol of CO(2)/h per g wet wt. There was no detectable decarboxylation of l-tryptophan, l-histidine or l-phenylalanine. 3. Dopa decarboxylase activity was found in all major regions of the ventral nerve cord of the mature locust (range: 4-7mumol of CO(2)/h per g wet wt.) but was low or absent in thoracic peripheral nerve. 4. Marked decarboxylation of l-dopa was found in homogenates of brains of four other species of insects, and in brain and ventral nerve cord, but not in the claw nerve, of the crayfish. 5. The activity of the locust brain enzyme may be slightly lower at the time of imaginal ecdysis than during the mature period. By contrast, the dopa decarboxylase that produces dopamine as an intermediate in cuticle biosynthesis is known to be high in activity at the time of ecdysis and low in activity during the intermoult stages.  相似文献   

14.
Summary A comparative study has been made of the metabolism in several strains of Thiobacillus neapolitanus of formate, acetate, propionate, butyrate, valerate and pyruvate. Conflicting reports in the literature concerning the mechanism of pyruvate assimilation in thiobacilli have been resolved. Pyruvate undergoes decarboxylation to yield acetyl coenzyme A, which is converted to glutamate, proline and arginine via reactions of the incomplete Krebs' cycle of this organism. Pyruvate is converted also to alanine, valine, isoleucine, leucine and lysine by mechanisms like those in heterotrophs. No aspartate is formed from the C-3 of pyruvate. Removal of the C-1 of pyruvate yields carbon dioxide, which is refixed into all cell constituents. Formate is not produced by this scission reaction, as formate itself is incorporated almost exclusively into purines. Aspartate can be synthesized by the activities of phosphoenolpyruvate carboxylase and oxaloacetate-glutamate transamination. Carbon from propionate is converted principally to lipids, although some amino acid production occurs with the same distinctive labelling pattern as is found after acetate assimilation by T. neapolitanus strains C and X. Butyrate and valerate also showed some distinctive patterns of incorporation into cell constituents. Fluoropyruvate and fluoropropionate inhibited the growth of T. neapolitanus and the mechanisms of this poisoning are discussed.Generally these compounds contributed only small proportions of the total cell carbon and tended to be converted to limited numbers of cell components. The thiobacilli thus tend to conserve carbon from these compounds and not to degrade them to carbon dioxide on a large scale when growing in an otherwise autotrophic medium.  相似文献   

15.
M Akhtar  D E Stevenson  D Gani 《Biochemistry》1990,29(33):7648-7660
L-Methionine decarboxylase from Dryopteris filix-mas catalyzes the decarboxylation of L-methionine and a range of straight- and branched-chain L-amino acids to give the corresponding amine products. The deuterium solvent isotope effects for the decarboxylation of (2S)-methionine are D(V/K) = 6.5 and DV = 2.3, for (2S)-valine are D(V/K) = 1.9 and DV = 2.6, and for (2S)-leucine are D(V/K) = 2.5 and DV = 1.0 at pL 5.5. At pL 6.0 and above, where the value of kcat for all of the substrates is low, the solvent isotope effects on Vmax for methionine are 1.1-1.2 whereas the effects on V/K remain unchanged, indicating that the solvent-sensitive transition state occurs before the first irreversible step, carbon dioxide desorption. The enzyme also catalyzes an abortive decarboxylation-transamination reaction in which the coenzyme is converted to pyridoxamine phosphate [Stevenson, D. E., Akhtar, M., & Gani, D. (1990a) Biochemistry (first paper of three in this issue)]. At very high concentration, the product amine can promote transamination of the coenzyme. However, the reaction occurs infrequently and does not influence the partitioning between decarboxylation and substrate-mediated abortive transamination under steady-state turnover conditions. The partition ratio, normal catalytic versus abortive events, can be determined from the amount of substrate consumed by a known amount of enzyme at infinite time, and the rate of inactivation can be determined by measuring the decrease in enzyme activity with respect to time. For methionine, the values of Km as determined from double-reciprocal plots of concentration versus inactivation rate are the same as those calculated from initial catalytic (decarboxylation) rate data, indicating that a single common intermediate partitions between product formation and slow transamination. The partition ratio is sensitive to changes in pH and is also dependent upon the structure of the substrate; methionine causes less frequent inactivation than either valine or leucine. The pH dependence of the partition ratio with methionine as substrate is very similar to that for V/K. Both curves show a sharp increase at approximately pH 6.25, indicating that a catalytic group on the enzyme simultaneously suppresses the abortive reaction and enhances physiological reaction in its unprotonated state. Experiments conducted in deuterium oxide allowed the solvent isotope effects for the partition ratio and the abortive reaction to be determined.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

16.
Peroxisomes from mung bean hypocotyl (Vigna radiata L.) degrade 2-oxoisocaproate, the transamination product of leucine, via isobutyryl-CoA and propionyl-CoA to acetyl-CoA. The methyl group at the C-3 position forms a barrier to β-oxidation. This barrier is overcome in the peroxisomes by several enzymatic steps. Senecioate (3-methylcrotonate), 2-hydroxyisovalerate, and 2-oxoisovalerate were detected as free acid intermediates. Senecioate, formed from 3-methylcrotonyl-CoA, is transformed by enzymatic hydrolysis to 2-hydroxyisovalerate. 2-Hydroxyisovalerate is then oxidized to 2-oxoisovalerate in an H2O2-producing reaction, exhibiting 1:1 stoichiometry of the products, by a 2-hydroxyacid oxidase which is different from the peroxisomal marker enzyme glycollate oxidase. 2-oxoisovalerate is activated by an NAD-dependent oxidative decarboxylation to isobutyryl-CoA. Accumulation of 2-oxoisovalerate in the presence of arsenite, an inhibitor of oxidative decarboxylations, is a feature of this latter pathway of degradation of isovaleryl-CoA or senecioate. It is concluded that the barrier caused by the methyl group of 2-oxoisocaproate is surmounted in higher plant peroxisomes in a manner different to that in mammalian mitochondria.  相似文献   

17.
Tyrosinase usually catalyzes the conversion of monophenols to o-diphenols and the oxidation of o-diphenols to the corresponding quinones. However, when 3,4-dihydroxymandelic acid was provided as the substrate, 3,4-dihydroxybenzaldehyde was produced. These results led to the proposal that tyrosinase catalyzes an unusual oxidative decarboxylation of this substrate (Sugumaran, M. (1986) Biochemistry 25, 4489-4492). However, 3,4-dihydroxybenzaldehyde is also obtained through the oxidation of 3,4-dihydroxymandelic acid by sodium periodate and on a mercury electrode. These results led to the proposal that tyrosinase catalyzes the oxidation of the substrate into o-quinone, which reacts immediately with a molecule of substrate, oxidizing it and through decarboxylation generates an intermediate (quinone methide) which transforms into 3,4-dihydroxybenzaldehyde; simultaneously, the original o-quinone is reduced to 3,4-dihydroxymandelic acid.  相似文献   

18.
[1-14C]-Labelled 2-oxoacid analogues of leucine, valine and phenylalanine were used to study the metabolism of these 2-oxoacids in the brain, kidney and heart muscle of rats. By following the 14CO2 release during 30–60 min of incubation at 37°C the decarboxylation rate was determined and measurement of the 14C-incorporation into the corresponding amino acid yielded the transamination rate. From these rates, decarboxylation/transamination ratios could be calculated which are indicative for the metabolic fate of the 2-oxoacid in the various organs. The results obtained show that all three tissues are capable of utilizing the 2-oxoacid analogues of leucine, valine and phenylalanine, however, to a different extent: kidney > heart muscle > brain. The decarboxylation/transamination ratios reveal that the branched-chain 2-oxoacids are predominantly decarboxylated in kidney and heart muscle while in brain they are mainly transaminated. The ratios calculated for phenylpyruvate in all tissues are within 0.19 and 0.36, indicating that this 2-oxoacid is preferentially transaminated. The results are discussed with respect to possible dietary alterations of enzymes involved in 2-oxoacid metabolism in order to improve transamination of these compounds.  相似文献   

19.
Mau YH  Wang WY 《Plant physiology》1988,86(3):793-797
The first committed intermediate of chlorophyll biosynthesis, δ-aminolevulinic acid (ALA), is synthesized from glutamate in the plant cell. The last step of ALA synthesis is a transamination reaction which converts glutamate-1-semialdehyde (GSA) to ALA. The mechanism of the transamination was examined by using glutamate, specifically labeled with either 1-13C or 15N, as substrate for ALA synthesis. After incubating with crude enzymes extracted from Chlamydomonas reinhardtii, the distribution of labels in purified ALA molecules was examined by nuclear magnetic resonance spectroscopy and gas chromatography-mass spectrometry. We found that both isotopes were present in the same ALA molecule. We interpret the results to mean that intermolecular transamination occurs during the conversion of GSA to ALA.  相似文献   

20.
In animals and plants, uroporphyrinogen decarboxylase catalyzes the stepwise decarboxylations of uroporphyrinogen, the precursor of heme and chlorophyll. To better understand its metabolic roles, we characterized the enzyme purified to electrophoretic homogeneity (about 11,000-fold) from human erythrocytes by a novel uroporphyrin-sepharose affinity chromatographic method. Native polyacrylamide disc gel electrophoresis of the purified enzyme preparation showed two bands detected by staining either for protein or with uroporphyrin-I. Each individual protein eluted from the gel when subjected to re-electrophoresis on SDS-polyacrylamide gel, appeared as a single protein band with molecular masses of approximately 54,000 and approximately 35,000 daltons respectively. Both proteins were able to catalyze all four decarboxylation steps, though the ratios of enzyme activity using octa-, hepta-, hexa- to pentacarboxylic porphyrinogen substrates were distinctly different. Also, their kinetic analysis with heptacarboxylic porphyrinogen-I substrate provided distinctly different apparent Michaelis constants. This provides the first evidence that decarboxylations of uroporphyrinogen to coproporphyrinogen are catalyzed by two isoenzymes.  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