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.
M K Sim 《Life sciences》1991,48(20):1985-1990
The activities of monoamine oxidase and phenolsulfotransferase in the hypothalamus and anterior pituitary gland of spontaneously hypertensive rats and the normotensive control (Wistar Kyoto rat) rats were investigated. The monoamine oxidase activity (determined using dopamine as substrate) in both these tissues was not significantly different between the normo- and hypertensive animals. Hypothalamic phenolsulfotransferase does not sulfate-conjugate dopamine at pH of 6.5 and pituitary phenolsulfotransferase does not sulfate-conjugate dopamine or 3,4-dihydroxyphenylacetic acid at the same pH. Hypothalamic phenolsulfotransferase activity determined using 3,4-dihydroxyphenylacetic acid as substrate was significantly higher in the spontaneously hypertensive than the Wistar Kyoto rats, while pituitary enzyme (determined using phenol as substrate) was the same in both strains of animals. We proposed that in the spontaneously hypertensive rats the higher level of hypothalamic phenolsulfotransferase could (by removing 3,4-dihydroxyphenylacetic acid as sulfated acid) increase the deamination of dopamine by monoamine oxidase. This could in turn result in the presence of high amount of sulfated 3,4-dihydroxyphenylacetic acid in the anterior pituitary gland reported in our earlier study, and be partly responsible for the reduced central dopaminergic activity found in the hypertensive rats.  相似文献   

2.
Pseudomonas putida F6 was found to metabolize p-hydroxyphenylacetic acid through 3,4-dihydroxyphenylacetic acid, 3,4-dihydroxymandelic acid, and 3,4-dihydroxybenzaldehyde. Cell extracts of P. putida F6 catalyze the NAD(P)H-independent hydroxylation of p-hydroxyphenylacetic acid to 3,4-dihydroxyphenylacetic acid which is further oxidized to 3,4-dihydroxymandelic acid. Oxidation and decarboxylation of the latter yields 3,4-dihydroxybenzaldehyde. A red-brown color accompanies all of the above enzyme activities and is probably due to the polymerization of quinone-like compounds. 3,4-Dihydroxybenzaldehyde is further metabolized through extradiol ring cleavage.  相似文献   

3.
When 3,4-dihydroxyphenylacetic acid (homoprotocatechuic acid) was added to Pseudomonase acidovorans growing at the expense of succinate, enzymes required for degrading homoprotocatechuate to pyruvate and succinate semialdehyde were strongly induced. These enzymes were effectively absent from cell extracts of the organism grown with 4-hydroxyphenylacetic acid, and this substrate was metabolized by the catabolic enzymes of the homogentisate pathway. Two separate ring-fission dioxygenases for 3,4,5-trihydroxybenzoic acid (gallic acid) were present in cell extracts of Pseudomonas putida when grown with syringic acid, and gallate was degraded by reactions associated with meta fission. One of the two gallate dioxygenases also attacked 3-O-methylgallic acid; the other, which did not, was induced when cells were exposed to gallate. This organism possessed ortho fission enzymes, including protocatechuate 3,4-dioxygenase (EC 1.13.11.3) and cis,cis-carboxymuconate-lactonizing enzyme (EC 5.5.1.2), after induction with 3,4-dihydroxybenzoic acid (protocatechuic acid). Gallate was a substrate for protocatechuate 3,4-dioxygenase, with a Vmax about 3% of that of protocatechuate and with an apparent Km slightly lower. Gallate was a powerful competitive inhibitor of protocatechuate oxidation.  相似文献   

4.
Catabolism of 3- and 4-hydroxyphenylacetic acid by Klebsiella pneumoniae   总被引:3,自引:0,他引:3  
Klebsiella pneumoniae catabolizes both 4-hydroxyphenylacetic acid and 3-hydroxyphenylacetic acid via meta-cleavage of 3,4-dihydroxyphenylacetic acid, ultimately yielding pyruvate and succinate. The organism can synthesize two hydroxylases catalysing 3,4-dihydroxyphenylacetic acid formation, which differ in substrate specificity, cofactor requirement, kinetics and regulation. Five enzymes sequentially involved in the catabolism of 3,4-dihydroxyphenylacetic acid are encoded on a 7 kbp fragment of the K. pneumoniae chromosome that has been isolated in a recombinant plasmid.  相似文献   

5.
Abstract: Administration of l -DOPA (50 mg/kg) elicits a significant increase in extracellular dopamine in striata of rats treated with the catecholaminergic neurotoxin 6-hydroxydopamine but not in striata of intact rats. To assess the role of dopaminergic nerve terminals in determining the effects of exogenous l -DOPA on extracellular dopamine levels in striatum, we examined the relative contributions of monoamine oxidase A and monoamine oxidase B to the catabolism of dopamine synthesized from exogenous l -DOPA. Extracellular concentrations of dopamine and its catabolite, 3,4-dihydroxyphenylacetic acid, were monitored with in vivo dialysis in striata of intact rats and of rats with unilateral 6-hydroxydopamine lesions of striatal dopamine. Clorgyline (2 mg/kg), an inhibitor of monoamine oxidase A, significantly increased dopamine and decreased 3,4-dihydroxyphenylacetic acid in intact but not in dopamine-depleted striata. Inhibition of monoamine oxidase B with either l -deprenyl (1 mg/kg) or Ro 19-6327 (1 mg/kg) did not significantly affect dopamine or 3,4-dihydroxyphenylacetic acid in striata of intact or dopamine-depleted rats. In intact rats, administration of clorgyline in conjunction with l -DOPA produced a >20-fold increase in dopamine and prevented the l -DOPA-induced increase in 3,4-dihydroxyphenylacetic acid. Although both l -deprenyl and Ro 19-6327 administered in combination with l -DOPA elicited a small but significant increase in dopamine, levels of 3,4-dihydroxyphenylacetic acid were not affected. In rats pretreated with 6-hydroxydopamine, clorgyline had no significant effect on the increases in dopamine and 3,4-dihydroxyphenylacetic acid elicited by l -DOPA. Furthermore, neither l -deprenyl nor Ro 19-6327 affected l -DOPA-induced increases in dopamine and 3,4-dihydroxyphenylacetic acid in dopamine-depleted striata. The present findings indicate that deamination by monoamine oxidase A is the primary mechanism for catabolism of striatal dopamine, both under basal conditions and after administration of exogenous l -DOPA. Loss of dopaminergic terminals eliminates this action of monoamine oxidase A but does not enhance deamination by monoamine oxidase B. These data support a model in which exogenous l -DOPA elicits enhanced extracellular accumulation of dopamine in the dopamine-depleted striatum because some transmitter synthesis occurs at nondopaminergic sites and the dopamine terminals that normally take up and catabolize this pool of transmitter are absent.  相似文献   

6.
Degradation of quercetin and luteolin by Eubacterium ramulus.   总被引:1,自引:0,他引:1  
The degradation of the flavonol quercetin and the flavone luteolin by Eubacterium ramulus, a strict anaerobe of the human intestinal tract, was studied. Resting cells converted these flavonoids to 3,4-dihydroxyphenylacetic acid and 3-(3,4-dihydroxyphenyl)propionic acid, respectively. The conversion of quercetin was accompanied by the transient formation of two intermediates, one of which was identified as taxifolin based on its specific retention time and UV and mass spectra. The structure of the second intermediate, alphitonin, was additionally elucidated by (1)H and (13)C nuclear magnetic resonance analysis. In resting-cell experiments, taxifolin in turn was converted via alphitonin to 3,4-dihydroxyphenylacetic acid. Alphitonin, which was prepared by enzymatic conversion of taxifolin and subsequent purification, was also transformed to 3,4-dihydroxyphenylacetic acid. The coenzyme-independent isomerization of taxifolin to alphitonin was catalyzed by cell extract or a partially purified enzyme preparation of E. ramulus. The degradation of luteolin by resting cells of E. ramulus resulted in the formation of the intermediate eriodictyol, which was identified by high-performance liquid chromatography and mass spectrometry analysis. The observed intermediates of quercetin and luteolin conversion suggest that the degradation pathways in E. ramulus start with an analogous reduction step followed by different enzymatic reactions depending on the additional 3-hydroxyl group present in the flavonol structure.  相似文献   

7.
A species of Acinetobacter and two strains of Pseudomonas putida when grown with 4-hydroxyphenylacetic acid gave cell extracts that converted 3,4-dihydroxyphenylacetic acid (homoprotocatechuic acid) into carbon dioxide, pyruvate, and succinate. The sequence of enzyme-catalyzed steps was as follows: ring-fission by a 2,3-dioxygenase, nicotinamide adenine dinucleotide-dependent dehydrogenation, decarboxylation, hydration, aldol fission, and oxidation of succinic semialdehyde. Two new metabolites, 5-carboxymethyl-2-hydroxymuconic acid and 2-hydroxyhepta-2,4-diene-1,7-dioic acid, were isolated from reaction mixtures and a third, 4-hydroxy-2-ketopimelic acid, was shown to be cleaved by extracts to give pyruvate and succinic semialdehyde. Enzymes of this metabolic pathway were present in Acinetobacter grown with 4-hydroxyphenylacetic acid but were effectively absent when 3-hydroxyphenylacetic acid or phenylacetic acid served as sources of carbon.  相似文献   

8.
A rapid and highly sensitive procedure for simultaneous determination of serotonin, 5-hydroxyindoleacetic acid, 3,4-dihydroxyphenylacetic acid and homovanillic acid is described. After precipitation of proteins with perchloric acid the samples are applied directly to a high performance liquid chromatograph, with electrochemical detection. As little as 20 pg of serotonin, 5-hydroxyindoleacetic acid, and 3,4-dihydroxyphenylacetic acid and 200 pg of homovanillic acid can be detected. One chromatographic run requires less than 10 min.  相似文献   

9.
A microorganism capable of degrading DL-mandelic acid was isolated from sewage sediment of enrichment culture and was identified as Pseudomonas convexa. It was found to metabolize mandelic acid by a new pathway involving 4-hydroxymandelic acid, 4-hydroxybenzaldehyde, 4-hydroxybenzoic acid, and 3,4-dihydroxybenzoic acid as aromatic intermediates. All the enzymes of the pathway were demonstrated in cell-free extracts. L-Mandelate-4-hydroxylase, a soluble enzyme, requires tetrahydropteridine, nicotinamide adenine dinucleotide phosphate, reduced form, and Fe2+ for its activity. The next enzyme, L-4-hydroxymandelate oxidase (decarboxylating), a particulate enzyme, requires flavine adenine dinucleotide and Mn2+ for its activity. A nicotinamide adenine dinucleotide-dependent, as well as a nicotinamide adenine dinucleotide phosphate-dependent, benzaldehyde dehydrogenase has been resolved and partially purified.  相似文献   

10.
A saturable, specific, high-affinity binding site for [3H]1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine was found in rat brain homogenates. The CNS regional distribution, the subcellular fractionation, and the displacement by pargyline, clorgyline, and deprenyl suggest that this binding site may correspond to monoamine oxidase. 1-Methyl-4-phenyl-1,2,3,6-tetrahydropyridine inhibited the oxidative deamination of dopamine, both in vivo and in vitro. Striatal levels of 3,4-dihydroxyphenylacetic acid were significantly reduced shortly after intravenous administration, and returned to normal values after a few hours. The in vitro formation of 3,4-dihydroxyphenylacetic acid from dopamine was inhibited by concentrations of 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine comparable to those of pargyline.  相似文献   

11.
A new gas chromatographic method, using only flame ionization detection which can determine nanogram quantities of homovanillic acid, 3,4-dihydroxyphenylacetic acid, 3-methoxy-4-hydroxyphenylethyleneglycol and 3,4-dihydroxyphenylethyleneglycol in the same reaction, is described. These compounds are treated with diazoethane and n-butylboronic acid. Homovanillic acid and 3,4-dihydroxyphenylacetic acid are converted to their ethyl esters while 3-methoxy-4-hydroxyphenylethyleneglycol and 3,4-dihydroxyphenylethyleneglycol from cyclic boronates and are thus assayed. This method is quantitative, highly specific and sensitive. It has been applied to the analysis of these compounds in urine.  相似文献   

12.
A capillary gas chromatographic/mass spectrometric method was developed for the determination of homovanillic acid and 3,4-dihydroxyphenylacetic acid in discrete areas of human brain known to contain only small amounts of dopamine metabolites. The electron impact mass spectra of the trimethylsilyl derivatives of homovanillic acid and 3,4-dihydroxyphenylacetic acid and their deuterated isotopic variants were used for their identification and quantification.  相似文献   

13.
The degradation of the flavonol quercetin and the flavone luteolin by Eubacterium ramulus, a strict anaerobe of the human intestinal tract, was studied. Resting cells converted these flavonoids to 3,4-dihydroxyphenylacetic acid and 3-(3,4-dihydroxyphenyl)propionic acid, respectively. The conversion of quercetin was accompanied by the transient formation of two intermediates, one of which was identified as taxifolin based on its specific retention time and UV and mass spectra. The structure of the second intermediate, alphitonin, was additionally elucidated by 1H and 13C nuclear magnetic resonance analysis. In resting-cell experiments, taxifolin in turn was converted via alphitonin to 3,4-dihydroxyphenylacetic acid. Alphitonin, which was prepared by enzymatic conversion of taxifolin and subsequent purification, was also transformed to 3,4-dihydroxyphenylacetic acid. The coenzyme-independent isomerization of taxifolin to alphitonin was catalyzed by cell extract or a partially purified enzyme preparation of E. ramulus. The degradation of luteolin by resting cells of E. ramulus resulted in the formation of the intermediate eriodictyol, which was identified by high-performance liquid chromatography and mass spectrometry analysis. The observed intermediates of quercetin and luteolin conversion suggest that the degradation pathways in E. ramulus start with an analogous reduction step followed by different enzymatic reactions depending on the additional 3-hydroxyl group present in the flavonol structure.  相似文献   

14.
Metabolites of dopamine and norepinephrine measured in the plasma have long been associated with symptomatic severity and response to treatment in schizophrenic, bipolar and other psychiatric patients. Plasma concentrations of catecholamine metabolites are genetically regulated. The genes encoding enzymes that are involved in the synthesis and degradation of these monoamines are candidate targets for this genetic regulation. We have studied the relationship between the Val158Met polymorphism in catechol O-methyltransferase gene, variable tandem repeat polymorphisms in the monoamine oxidase A gene promoter, and plasma concentrations of 3-methoxy-4-hydroxyphenylglycol, 3,4-dihydroxyphenylacetic acid and homovanillic acid in healthy control subjects as well as in untreated schizophrenic and bipolar patients. We found that the Val158Met substitution in catechol O-methyltransferase gene influences the plasma concentrations of homovanillic and 3,4-dihydroxyphenylacetic acids. Although higher concentrations of plasma homovanillic acid were found in the high-activity ValVal genotype, this mutation did not affect the plasma concentration of 3-methoxy-4-hydroxyphenylglycol. 3,4-dihydroxyphenylacetic acid concentrations were higher in the low-activity MetMet genotype. Interestingly, plasma values 3-methoxy-4-hydroxyphenylglycol were greater in schizophrenic patients and in bipolar patients than in healthy controls. Our results are compatible with the previously reported effect of the Val158Met polymorphism on catechol O-methyltransferase enzymatic activity. Thus, our results suggest that this polymorphism, alone or associated with other polymorphisms, could have an important role in the genetic control of monoamine concentration and its metabolites.  相似文献   

15.
Lactate-grown cultures of Acinetobacter sp. strain 3B-1 synthesize constitutively all enzymes except the 4-hydroxyphenylacetic acid-3-hydroxylase. All enzymes are further synthesized when strain 3B-1 is grown with 4-hydroxyphenylacetic acid. Induction studies with two mutant strains, one defective in the 3-hydroxylase, and the other defective in the dehydrogenase, indicate that 4-hydroxyphenylacetic acid induces the 3-hydroxylase only, and the second metabolite 3,4-dihydroxyphenylacetic acid appears to induce 3,4-dihydroxyphenylacetic acid-2,3-dioxygenase and subsequent enzymes. Thus, the enzymes of the 4-hydroxyphenylacetic acid meta-cleavage pathway are synthesized following at least two sequential inductive events.  相似文献   

16.
Microdialysis was used to assess extracellular dopamine in striatum, nucleus accumbens, and medial frontal cortex of unanesthetized rats both under resting conditions and in response to intermittent tail-shock stress. The dopamine metabolites 3,4-dihydroxyphenylacetic acid and homovanillic acid also were measured. The resting extracellular concentration of dopamine was estimated to be approximately 10 nM in striatum, 11 nM in nucleus accumbens, and 3 nM in medial frontal cortex. In contrast, the resting extracellular levels of 3,4-dihydroxyphenylacetic acid and homovanillic acid were in the low micromolar range. Intermittent tail-shock stress increased extracellular dopamine relative to baseline by 25% in striatum, 39% in nucleus accumbens, and 95% in medial frontal cortex. 3,4-Dihydroxyphenylacetic acid and homovanillic acid also were generally increased by stress, although there was a great deal of variability in these responses. These data provide direct in vivo evidence for the global activation of dopaminergic systems by stress and support the concept that there exist regional variations in the regulation of dopamine release.  相似文献   

17.
A metabolic pathway for L-tyrosine catabolism involves 3,4-dihydroxyphenylacetic acid (homoprotocatechuic acid) as substrate for fission of the benzene nucleus. Cell extracts of an organism tentatively identified as a Micrococcus possessed the enzymes required for degrading homoprotocatechuate to succinate and pyruvate, and stoichiometry was established for several of these reactions. When the required coenzymes were added, cell extracts degraded L-tyrosine to the ring-fission product of homoprotocatechuate 2,3-dioxygenase and also converted 4-hydroxyphenylpyruvic acid into 4-hydroxyphenylacetic acid. This compound, in turn, gave stoichiometric amounts of the ring-fission product of homoprotocatechuate by the action of a nicotinamide adenine dinucleotide phosphate-dependent 3-hydroxylase coupled with homoprotocatechuate 2,3-dioxygenase. Evidence is presented that this route for L-tyrosine catabolism is taken by five other gram-positive strains, including Micrococcus lysodeikticus and a species of Bacillus. Five other gram-positive bacteria from other genera employed the alternative homogentisate pathway.  相似文献   

18.
Outer scales of yellow onion bulbs turn brown during maturing. The brown outer scales contain an antifungal component, 3,4-dihydroxybenzoic acid. An aim of the present study is to elucidate the mechanism of formation of the benzoic acid. In a browning scale, the scale was divided into three areas; fleshy, drying and dried brown areas. Levels of quercetin glucosides in dried brown areas were less than 10% of the glucosides in fleshy and drying areas, whereas levels of quercetin were high in dried brown areas. This result suggests that quercetin was formed by deglucosidation of quercetin glucosides on the border between drying and dried brown areas. Peroxidase (POX) activity of dried brown areas was about 10% of those of fleshy and drying areas. Quercetin was oxidized by autooxidation, and cell-free extracts of drying areas and POX isolated from onion scales enhanced the oxidation even in the absence of externally added hydrogen peroxide. The enhancement of quercetin oxidation was suppressed by catalase. No tyrosinase-like activity was detected in the cell-free extracts and the POX preparation. These results suggest that, during the enhanced oxidation of quercetin, hydrogen peroxide is formed. 3,4-Dihydroxybenzoic acid and 2,4,6-trihydroxyphenylglyoxylic acid, which were the oxidation products of quercetin, were found in dried brown area. These results suggest that an antifungal agent 3,4-dihydroxybenzoic acid is formed by POX-dependent oxidation of quercetin on browning of onion scales.  相似文献   

19.
Eight structurally similar dihydroxy and trihydroxyphenolic acids (protocatechuic acid, 3,4-dihydroxyphenylacetic acid, hydrocaffeic acid, caffeic acid, gallic acid, 3,4,5-trihydroxyphenylacetic acid, 3-(3,4,5-trihydroxyphenyl)propanoic acid and 3-(3,4,5-trihydroxyphenyl)propenoic acid) were examined for their total antioxidant capacity (TAC). Furthermore, their ability to scavenge peroxyl radicals, generated by AAPH in liposomes, was determined. The antioxidant/pro-oxidant activity of the compounds was screened using the 2'-deoxyguanosine assay. All compounds behave as radical scavengers, with 3,4,5-trihydroxyphenylacetic acid being the most potent. Nevertheless, in the lipid peroxidation assay an inverse ranking order was observed, 3,4-dihydroxyphenylacetic acid being the most effective compound. All the dihydroxylated compounds showed a pro-oxidant behaviour leading to an increase of 50% in 8-OH-dG induction. From the structure-antioxidant activity relationship studies performed it may be concluded that the number of phenolic groups and the type of the alkyl spacer between the carboxylic acid and the aromatic ring strongly influence the antioxidant activity.  相似文献   

20.
Eight structurally similar dihydroxy and trihydroxyphenolic acids (protocatechuic acid, 3,4-dihydroxyphenylacetic acid, hydrocaffeic acid, caffeic acid, gallic acid, 3,4,5-trihydroxyphenylacetic acid, 3-(3,4,5-trihydroxyphenyl)propanoic acid and 3-(3,4,5-trihydroxyphenyl)propenoic acid) were examined for their total antioxidant capacity (TAC). Furthermore, their ability to scavenge peroxyl radicals, generated by AAPH in liposomes, was determined. The antioxidant/pro-oxidant activity of the compounds was screened using the 2′-deoxyguanosine assay. All compounds behave as radical scavengers, with 3,4,5-trihydroxyphenylacetic acid being the most potent. Nevertheless, in the lipid peroxidation assay an inverse ranking order was observed, 3,4-dihydroxyphenylacetic acid being the most effective compound. All the dihydroxylated compounds showed a pro-oxidant behaviour leading to an increase of 50% in 8-OH-dG induction. From the structure–antioxidant activity relationship studies performed it may be concluded that the number of phenolic groups and the type of the alkyl spacer between the carboxylic acid and the aromatic ring strongly influence the antioxidant activity.  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