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 15 毫秒
1.
The extent of hydroxylation, transamination, and decarboxylation in the metabolism of excess phenylalanine in vivo has been examined by measuring the amount of radioactivity from [14C]phenylalanine that is converted to 14CO2 and urinary metabolites. Transamination and direct decarboxylation represent only 6% of total phenylalanine metabolism. The major aromatic metabolites in the urine after phenylalanine loading are phenylacetylglycine, phenylacetic acid, phenylpyruvate, and phenylalanine. A small but significant portion (1.5%) of phenylalanine is degraded to nonaromatic compounds. The maximum phenylalanine oxidation in vivo is approximately 75%24 h at saturating concentrations of phenylalanine; thus, the major route of degradation of phenylalanine in the rat, even when intake is high, is via formation of acetoacetic acid and fumaric acid.  相似文献   

2.
《Insect Biochemistry》1989,19(4):401-407
DOPA quinone imine conversion factor (QICF) was partially purified from pharate pupal cuticles of the tobacco hornworm [Manduca sexta (L.)] and its properties were examined. QICF was stable at alkaline pH rather than acidic pH. This was a thermolabile factor and its activity was lost on incubation at 50–60°C for 22 min. QICF effectively catalyzed the decolorization of l-dopachrome and l-dopachrome methyl ester optimally at pH 6–7 but its catalytic efficiency for decolorization of dopamine chrome was quite low. This factor did not act on chromes derived from d-DOPA, dl-α-methyldopa, N-methyl dopamine and norepinephrine. No catalytic activity was found in QICF for decarboxylation of 5,6-dihydroxyindole-2-carboxylic acid. In the presence of tyrosinase, QICF accelerated melanochrome formation. From the results, it is suggested that QICF catalyzes a decarboxylation and/or rearrangement of dopachrome during the conversion of dopachrome to 5,6-dihydroxyindole, and this factor might play an important role in the tyrosinase-mediated rapid melanization requisite for insect larva.  相似文献   

3.
We have measured the 13C kinetic isotope effect at pH 4.0, 5.0, 6.0, and 6.5 and in D2O at pD 5.0 and the rate of D-H exchange of the alpha and beta protons of aspartic acid in D2O at pD 5.0 for the reaction catalyzed by the enzyme aspartate beta-decarboxylase from Alcaligenes faecalis. The 13C kinetic isotope effect, with a value of 1.0099 +/- 0.0002 at pH 5.0, is less than the intrinsic isotope effect for the decarboxylation step, indicating that the decarboxylation step is not entirely rate limiting. We have been able to estimate probable values of the relative free energies of the transition states of the enzymatic reaction up to and including the decarboxylation step from the 13C kinetic isotope effect and the rate of D-H exchange of alpha-H. The pH dependence of the kinetic isotope effect reflects the pKa of the pyridine nitrogen of the coenzyme pyridoxal 5'-phosphate but not that of the imine nitrogen. A mechanism is proposed for the exchange of aspartate beta-H that is consistent with the stereochemistry suggested earlier.  相似文献   

4.
Phosphoenolpyruvate carboxykinase of chicken liver cytosol was purified to homogeneity by procedures including affinity chromatography with GTP as a ligand. The purified enzyme showed a molecular weight of 68,000 on gel electrophoresis in the presence of dodecyl sulfate. Comparative studies on this enzyme and its isozyme purified from chicken liver mitochondria were performed. As regards amino acid composition, the cytosolic enzyme was quite different from the mitochondrial enzyme, but was rather similar to rat liver cytosolic phosphoenolpyruvate carboxykinase. Specific activities of the cytosolic enzyme were 30-100% higher than those of the mitochondrial enzyme for oxaloacetate-CO2 exchange, oxaloacetate decarboxylation, and phosphoenolpyruvate carboxylation reactions, though the relative rates of the activities were similar, decreasing in the order given. Apparent Michaelis constants for oxaloacetate in the oxaloacetate decarboxylation reaction were 11.6 and 17.9 microM for the cytosolic and the mitochondrial enzyme, respectively, but the values for GTP, GDP, phosphoenolpyruvate, and CO2 in the oxaloacetate decarboxylation and phosphoenolpyruvate carboxylation reactions were 1.3-2.2 times higher for the cytosolic enzyme than for the mitochondrial enzyme. Thus, the fundamental catalytic properties of the chicken liver phosphoenolpyruvate carboxykinase isozymes were rather similar, despite the marked difference in amino acid compositions.  相似文献   

5.
The biosynthesis of staurosporine, rebeccamycin, and goadsporin, which are produced by actinomycetes and contain characteristic heterocyclic rings, was characterized by genetic methods. Staurosporine and rebeccamycin contain an indolocarbazole ring synthesized from two molecules of tryptophan, with indolepyruvic acid imine and chromopyrrolic acid as biosynthetic intermediates. A tetrameric hemoprotein synthesizes chromopyrrolic acid, and cytochrome P450 peroxidase catalyzes the intramolecular C–C coupling and decarboxylation of chromopyrrolic acid to form the indolocarbazole core. Goadsporin is a thiopeptide containing thiazole and oxazole heterocyclic rings. The structural gene godA is ribosomally translated to a goadsporin precursor peptide, and oxazole, methyloxazole, and thiazole rings are derived from serine, threonine, and cystein through post-translational modifications. On the basis of these knowledges, a wide variety of indolocarbazole and goadsporin analogs through the rational gene recombination and disruption of these biosynthetic genes were successfully produced.  相似文献   

6.
L M Abell  M H O'Leary 《Biochemistry》1988,27(16):5933-5939
The decarboxylation of histidine by the pyruvate-dependent histidine decarboxylase of Lactobacillus 30a shows a carbon isotope effect of k12/k13 = 1.0334 +/- 0.0005 and a nitrogen isotope effect k14/k15 = 0.9799 +/- 0.0006 at pH 4.8, 37 degrees C. The carbon isotope effect is slightly increased by deuteriation of the substrate and slightly decreased in D2O. The observed nitrogen isotope effect indicates that the imine nitrogen in the substrate-Schiff base intermediate complex is ordinarily protonated, and the pH dependence of the carbon isotope effect indicates that both protonated and unprotonated forms of this intermediate are capable of undergoing decarboxylation. As with the pyridoxal 5'-phosphate dependent enzyme, Schiff base formation and decarboxylation are jointly rate-limiting, with the intermediate histidine-pyruvate Schiff base showing a decarboxylation/Schiff base hydrolysis ratio of 0.5-1.0 at pH 4.8. The decarboxylation transition state is more reactant-like for the pyruvate-dependent enzyme than for the pyridoxal 5'-phosphate dependent enzyme. These studies find no particular energetic or catalytic advantage to the use of pyridoxal 5'-phosphate over covalently bound pyruvate in catalysis of the decarboxylation of histidine.  相似文献   

7.
Although in mammals, acetone formation from acetoacetic acid is normally regarded as a non-enzymatic (spontaneous) process, the existence of an acetoacetic acid decarboxylase activity was postulated recently. The results imply that this enzymatic activity can be relatively important at the physiological concentrations of ketone bodies found in the rat and that acetone acts as a competitive inhibitor of this enzyme.  相似文献   

8.
Indolepyruvate decarboxylase, a key enzyme for indole-3-acetic acid biosynthesis, was found in extracts of Enterobacter cloacae. The enzyme catalyzes the decarboxylation of indole-3-pyruvic acid to yield indole-3-acetaldehyde and carbon dioxide. The enzyme was purified to apparent homogeneity from Escherichia coli cells harboring the genetic locus for this enzyme obtained from E. cloacae. The results of gel filtration experiments showed that indolepyruvate decarboxylase is a tetramer with an M(r) of 240,000. In the absence of thiamine pyrophosphate and Mg2+, the active tetramers dissociate into inactive monomers and dimers. However, the addition of thiamine pyrophosphate and Mg2+ to the inactive monomers and dimers results in the formation of active tetramers. These results indicate that the thiamine pyrophosphate-Mg2+ complex functions in the formation of the tetramer, which is the enzymatically active holoenzyme. The enzyme exhibited decarboxylase activity with indole-3-pyruvic acid and pyruvic acid as substrates, but no decarboxylase activity was apparent with L-tryptophan, indole-3-lactic acid, beta-phenylpyruvic acid, oxalic acid, oxaloacetic acid, and acetoacetic acid. The Km values for indole-3-pyruvic acid and pyruvic acid were 15 microM and 2.5 mM, respectively. These results indicate that indole-3-acetic acid biosynthesis in E. cloacae is mediated by indolepyruvate decarboxylase, which has a high specificity and affinity for indole-3-pyruvic acid.  相似文献   

9.
Control of pH in aminoethylation reactions is critical for maintaining high selectivity towards cysteine modification. Measurement of aminoethylation rate constants by liquid chromatography mass spectrometry demonstrates reaction selectivity of cysteine>amino-terminus>histidine. Lysine and methionine were not reactive at the conditions used. For thiol modification, the acid/base property of the gamma-thialysine residue measured by NMR results in a 1.15 decrease in pK(a) (relative to a lysine residue). NMR confirms ethylene imine is the reactive intermediate for alkylation of peptide nucleophiles with bromoethylamine. Conversion of bromoethylamine into ethylene imine prior to exposure to the target thiol, provides a reagent that promotes selectivity by allowing precise control of reaction pH. Reaction selectivity plots of relative aminoethylation rates for cysteine, histidine, and N-terminus imine demonstrate increasing alkaline conditions favors thiol modification. When applied to protein modification, the conversion of bromoethylamine into ethylene imine and buffering at alkaline pH will allow optimal cysteine residue aminoethylation.  相似文献   

10.
The rates of the acid-catalyzed decarboxylation and amide hydrolysis of α-ketoglutaramic acid, the keto analog of glutamine, were investigated and the products of the reactions were characterized. In strong acid at 100°C, amide hydrolysis and decarboxylation occur with about equal facility, yielding α-ketoglutaric acid and 5-hydroxy-2-pyrrolidone, respectively. 5-Hydroxy-2-pyrrolidone undergoes further amide hydrolysis so that the products of complete acid hydrolysis of α-ketoglutaramic acid are ammonia (100%), carbon dioxide (50%), and approximately equal yields (50%) of α-ketoglutaric acid and succinic semialdehyde (β-formylpropionic acid). At increasing pH values, the relative rate of decarboxylation to amide hydrolysis of α-ketoglutaramic acid increases, such that, at pH values of 2 or greater, decarboxylation occurs almost exclusively. The decarboxylation product 5-hydroxy-2-pyrrolidone, was characterized chromatographically and by its infrared and pmr spectra; the compound may be regarded as the cyclized form of succinamic semialdehyde. A mechanism for the competing amide hydrolysis and decarboxylation reactions is proposed, and the potential biological significance of the decarboxylation pathway is discussed.  相似文献   

11.
Trichosporon cutaneum degraded 4-hydroxyphenylacetic acid to acetoacetic and malic acids. 3.4-Dihydroxyphenylacetic acid, an intermediate in the reaction sequence, underwent hydroxylation before the benzene ring was cleaved.  相似文献   

12.
L M Abell  M H O'Leary 《Biochemistry》1988,27(16):5927-5933
The pyridoxal 5'-phosphate dependent histidine decarboxylase from Morganella morganii shows a nitrogen isotope effect k14/k15 = 0.9770 +/- 0.0021, a carbon isotope effect k12/k13 = 1.0308 +/- 0.0006, and a carbon isotope effect for L-[alpha-2H]histidine of 1.0333 +/- 0.0001 at pH 6.3, 37 degrees C. These results indicate that the overall decarboxylation rate is limited jointly by the rate of Schiff base interchange and by the rate of decarboxylation. Although the observed isotope effects are quite different from those for the analogous glutamate decarboxylase from Escherichia coli [Abell, L. M., & O'Leary, M. H. (1988) Biochemistry 27, 3325], the intrinsic isotope effects for the two enzymes are essentially the same. The difference in observed isotope effects occurs because of a roughly twofold difference in the partitioning of the pyridoxal 5'-phosphate-substrate Schiff base between decarboxylation and Schiff base interchange. The observed nitrogen isotope effect requires that the imine nitrogen in this Schiff base is protonated. Comparison of carbon isotope effects for deuteriated and undeuteriated substrates reveals that the deuterium isotope effect on the decarboxylation step is about 1.20; thus, in the transition state for the decarboxylation step, the carbon-carbon bond is about two-thirds broken.  相似文献   

13.
Feng WY  Austin TJ  Chew F  Gronert S  Wu W 《Biochemistry》2000,39(7):1778-1783
The mechanism of orotidine 5'-monophosphate decarboxylase (OMP decarboxylase, ODCase) was studied using the decarboxylation of orotic acid analogues as a model system. The rate of decarboxylation of 1,3-dimethylorotic acid and its analogues as well as the stability of their corresponding carbanion intermediates was determined. The results have shown that the stability of the carbanion intermediate is not a critical factor in the rate of decarboxylation. On the other hand, the reaction rate is largely dependent on the equilibrium constant for the formation of a zwitterion. Based on these results, we have proposed a new mechanism in which ODCase catalyzes the decarboxylation of OMP by binding the substrate in a zwitterionic form and providing a destabilizing environment for the carboxylate group of OMP.  相似文献   

14.
Homoisocitrate dehydrogenase (HIcDH, 3-carboxy-2-hydroxyadipate dehydrogenase) catalyzes the fourth reaction of the alpha-aminoadipate pathway for lysine biosynthesis, the conversion of homoisocitrate to alpha-ketoadipate using NAD as an oxidizing agent. A chemical mechanism for HIcDH is proposed on the basis of the pH dependence of kinetic parameters, dissociation constants for competitive inhibitors, and isotope effects. According to the pH-rate profiles, two enzyme groups act as acid-base catalysts in the reaction. A group with a p K a of approximately 6.5-7 acts as a general base accepting a proton as the beta-hydroxy acid is oxidized to the beta-keto acid, and this residue participates in all three of the chemical steps, acting to shuttle a proton between the C2 hydroxyl and itself. The second group acts as a general acid with a p K a of 9.5 and likely catalyzes the tautomerization step by donating a proton to the enol to give the final product. The general acid is observed in only the V pH-rate profile with homoisocitrate as a substrate, but not with isocitrate as a substrate, because the oxidative decarboxylation portion of the isocitrate reaction is limiting overall. With isocitrate as the substrate, the observed primary deuterium and (13)C isotope effects indicate that hydride transfer and decarboxylation steps contribute to rate limitation, and that the decarboxylation step is the more rate-limiting of the two. The multiple-substrate deuterium/ (13)C isotope effects suggest a stepwise mechanism with hydride transfer preceding decarboxylation. With homoisocitrate as the substrate, no primary deuterium isotope effect was observed, and a small (13)C kinetic isotope effect (1.0057) indicates that the decarboxylation step contributes only slightly to rate limitation. Thus, the chemical steps do not contribute significantly to rate limitation with the native substrate. On the basis of data from solvent deuterium kinetic isotope effects, viscosity effects, and multiple-solvent deuterium/ (13)C kinetic isotope effects, the proton transfer step(s) is slow and likely reflects a conformational change prior to catalysis.  相似文献   

15.
Liver and kidney microsomes have been found to catalyze a rapid decarboxylation of retinoic acid in vitro. The reaction requires NADPH and Fe(2+), and is further stimulated by the presence of pyrophosphate. Thiamine pyrophosphate contained sufficient iron as an impurity to provide strong enhancement of the reaction in the absence of added iron. The decarboxylation could also be shown to occur nonenzymatically in the presence of ascorbate, Fe(2+), and boiled microsomes, but there was little autoxidation resulting in decarboxylation. The reaction was strongly inhibited by chelating agents, N,N'-diphenyl-p-phenylene diamine, phenazine methosulfate, and ferricyanide, and resembled lipid peroxidation in both its cofactor requirements and response to inhibitors. The product of the reaction appeared to lack only the C-15 of the original retinoic acid molecule. It was not retained by diethylaminoethyl cellulose, was more polar than retinoic acid upon silicic acid chromatography, had a lower UV absorption maximum (295 m micro ) than the starting product, and seemed to have an aldehyde group at C-14. The physiological significance of the decarboxylation remains to be assessed, but its rapidity makes it important to in vitro work on retinoic acid.  相似文献   

16.
Catalytic antibody 27C1 bears binding sites for both a substrate- and a functionalized small nonprotein component in the active site. We investigated the possibility of exploiting imine and enamine intermediates using a primary amine molecule into the active site of antibody 27C1. The antibody catalyzed β-keto acid decarboxylation with a rate enhancement (kcat/Km/kuncat) of 140,000, as well as highly regioselective cross-aldol reactions of ketones and p-nitrobenzaldehyde. These studies provide new strategies for the generation of catalytic antibodies possessing binding sites for functionalized components.  相似文献   

17.
The apparent binding constants and the effect of semisynthetic derivatives of carminomycin and rubomycin (anthracycline antibiotics) on DNA fusion were studied. The following semisynthetic derivatives were used. 13-dihydrocarminomycin, 14-hydroxycarminomycin, 13-(4-methylpiperazinyl) imine carminomycin, 13-benzoylhydrazone carminomycin (carminazone), 13-tret-butoxycarbonyl hydrazone rubomycin, 13-(4-methylpiperazinyl) imine rubomycin, 14-(1-hydroxyl-2,2,6,6-tetramethylpiperidyl-4)-acetoxyrubomycin (spin-labeled rubomycin). The above derivatives slightly differed from the initial antibiotics by their affinity to DNA. The binding constants of methylpiperazinyl imines was 2-3 times higher than those of the respective antibiotics.  相似文献   

18.
The mechanism of inactivation of rodent ornithine decarboxylase by alpha-difluoromethylornithine (DFMO) was studied using the inhibitor labelled with 14C in both the 1 and the 5 positions. [1-14C]DFMO was a substrate and was decarboxylated by the enzyme yielding 14CO2. A radioactive metabolite derived from [5-14C]DFMO was bound to the enzyme, and the extent of binding paralleled the irreversible inactivation of ornithine decarboxylase. The partition ratio of decarboxylation to binding was approx. 3.3. These results provide support for the postulated mechanism of action of DFMO [Metcalf, Bey, Danzin, Jung, Casera & Vevert (1978) J. Am. Chem. Soc. 100, 2551-2553], in which enzymic decarboxylation of the inhibitor leads to the generation of a conjugated imine, which then alkylates a nucleophilic residue on the enzyme.  相似文献   

19.
The preparation of 24-nor-5β-chol-22-enes from formyloxy-5β-cholanic acids by oxidative decarboxylation with lead tetraacetate is described. NMR data is presented with other physical constants for the norcholenes derived from cholic, chenodeoxycholic, ursodeoxycholic, hyodeoxycholic, and deoxycholic acids. The facile synthesis of these norcholenes demonstrates the applicability of the formyloxy protecting group to oxidative decarboxylations in the bile acid series.  相似文献   

20.
Purified branched-chain 2-oxo acid dehydrogenase (BCODH) and pyruvate dehydrogenase (PDH) had apparent Km values (microM) for 2-oxobutyrate of 26 and 114, with a relative Vmax. (% of Vmax. for 3-methyl-2-oxobutyrate and pyruvate) of 38 and 45% respectively. The phosphorylation state of both complexes in extracts of mitochondria from rat liver, kidney, heart and skeletal muscle was shown to influence oxidative decarboxylation of 2-oxobutyrate. Inhibitory antibodies to BCODH and an inhibitor of PDH (3-fluoropyruvate) were used with mitochondrial extracts to determine the relative contribution of both complexes to oxidative decarboxylation of 2-oxobutyrate. Calculated rates of 2-oxobutyrate decarboxylation in mitochondrial extracts, based on the kinetic constants given above and the activities of both complexes, were the same as the measured rates. Hydroxyapatite chromatography of extracts of mitochondria from rat liver revealed only two peaks of oxidative decarboxylation of 2-oxobutyrate, with one peak associated with PDH and the other with BCODH. Competition studies with various 2-oxo acids revealed a different inhibition pattern with mitochondrial extracts from liver compared with those from heart or skeletal muscle. We conclude that both intramitochondrial complexes are responsible for oxidative decarboxylation of 2-oxobutyrate. However, the BCODH is probably the more important complex, particularly in liver, on the basis of kinetic analyses, activity or phosphorylation state of both complexes, competition studies, and the apparent physiological concentration of pyruvate, 2-oxobutyrate and the branched-chain 2-oxo acids.  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