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 detailed kinetic study was carried out to investigate the porcine pancreatic lipase-catalysed esterification reactions of p-cresol–acetic acid and lactic acid–stearic acid. The kinetic data were in agreement with a Ping Pong Bi–Bi mechanism being followed by the enzyme, where inhibition is indicated in the presence of p-cresol and lactic acid in the respective reactions. Mathematical analyses of experimentally observed initial rates yielded various kinetic parameters, K m(p-cresol) = 0.1, K m(acetic acid) = 0.54, K m(lactic acid) = 0.059 M, K m(stearic acid) = 0.04 M, V max(p-cresol–acetic acid) = 13.2(h–1), V max(lactic acid–stearic acid) = 0.00163 M/h, K i(p-cresol) = 0.59 and K i(lactic acid) = 0.079 M. The K m and K i values of p-cresol and lactic acid observed in the respective reactions showed both the competitive nature of binding between the substrates p-cresol and acetic acid on the one hand and lactic acid and stearic acid on the other and the inhibitory nature of p-cresol and lactic acid.  相似文献   

2.
Sequential photo-and biodegradation of p-cresol was studied using a mercury lamp, as well as KrCl and XeCl excilamps. Preirradiation of p-cresol at a concentration of 10?4 M did not affect the rate of its subsequent biodegradation. An increase in the concentration of p-cresol to 10?3 M and in the duration preliminary UV irradiation inhibited subsequent biodegradation. Biodegradation of p-cresol was accompanied by the formation of a product with a fluorescence maximum at 365 nm (λex = 280 nm), and photodegradation yielded a compound fluorescing at 400 nm (λex = 330 nm). Sequential UV and biodegradation led to the appearance of bands in the fluorescence spectra that were ascribed to p-cresol and its photolysis products. It was shown that sequential use of biological and photochemical degradation results in degradation of not only the initial toxicant but also the metabolites formed during its biodegradation.  相似文献   

3.
Summary Sulfate-reducing bacterial enrichments were obtained from a shallow anoxic aquifer for their ability to metabolize eithero-, m-, orp-cresol. GC/MS and simultaneous adaptation experiments suggested that the anaerobic decomposition ofp-cresol proceeds by the initial oxidation of the aryl methyl group to formp-hydroxybenzoic acid. This intermediate was then converted to benzoic acid. Benzoic acid and a hydroxybenzaldehyde were also found in spent culture fluids from ano-cresol-degrading enrichment culture. This result, in addition to others, suggested thato-cresol may also be anaerobically degraded by the oxidation of the methyl substituent. An alternate pathway for anaerobicm-cresol decomposition might exist. Enrichment cultures obtained with eitherp- oro-cresol degraded both of these substrates but notm-cresol. In contrast, am-cresol enrichment culture did not metabolize theortho orpara isomers. Anaerobic biodegradation in all enrichment cultures was inhibited by molybdate and oxygen, and was dependent on the presence of sulfate as a terminal electron acceptor. The stoichiometry of sulfate-reduction and substrate depletion by the various enrichment cultures indicated that the parent cresol isomers were completely mineralized. This result was confirmed by the conversion of14C-labeledp-cresol to14CO2. These results help clarify the fate of alkylated aromatic chemicals in anoxic aquifers.  相似文献   

4.
p-Cresol methylhydroxylases (PCMH) from aerobic and facultatively anaerobic bacteria are soluble, periplasmic flavocytochromes that catalyze the first step in biological p-cresol degradation, the hydroxylation of the substrate with water. Recent results suggested that p-cresol degradation in the strictly anaerobic Geobacter metallireducens involves a tightly membrane-bound PCMH complex. In this work, the soluble components of this complex were purified and characterized. The data obtained suggest a molecular mass of 124 ± 15 kDa and a unique αα′β2 subunit composition, with α and α′ representing isoforms of the flavin adenine dinucleotide (FAD)-containing subunit and β representing a c-type cytochrome. Fluorescence and mass spectrometric analysis suggested that one FAD was covalently linked to Tyr394 of the α subunit. In contrast, the α′ subunit did not contain any FAD cofactor and is therefore considered to be catalytically inactive. The UV/visible spectrum was typical for a flavocytochrome with two heme c cofactors and one FAD cofactor. p-Cresol reduced the FAD but only one of the two heme cofactors. PCMH catalyzed both the hydroxylation of p-cresol to p-hydroxybenzyl alcohol and the subsequent oxidation of the latter to p-hydroxybenzaldehyde in the presence of artificial electron acceptors. The very low Km values (1.7 and 2.7 μM, respectively) suggest that the in vivo function of PCMH is to oxidize both p-cresol and p-hydroxybenzyl alcohol. The latter was a mixed inhibitor of p-cresol oxidation, with inhibition constants of a Kic (competitive inhibition) value of 18 ± 9 μM and a Kiu (uncompetitive inhibition) value of 235 ± 20 μM. A putative functional model for an unusual PCMH enzyme is presented.  相似文献   

5.
Phenol hydroxylase gene engineered microorganism (PHIND) was used to synthesize catechols from benzene and toluene by successive hydroxylation reaction. HPLC-MS and 1H NMR analysis proved that the products of biotransformation were the corresponding catechols via the intermediate production of phenols. It was indicated that the main products of toluene oxidation were o-cresol and p-cresol. 3-Methylcatechol was the predominant product for m-cresol biotransformation. Formation rate of catechol (25 μM/min/g cell dry weight) was 1.43-fold higher than that of methylcatechols. It was suggested that phenol hydroxylase could be successfully used to transform both benzene and toluene to catechols by successive hydroxylation.  相似文献   

6.
The substrate benzaldehyde (but not propionaldehyde) could elute aldehyde dehydrogenase from a p-hydroxyacetophenone-affinity column, and inhibit the esterase activity (Ki=47 μM), indicating that this simple aromatic aldehyde binds to the free enzyme and possibly in the substrate-binding site. Thus, the kinetic mechanism for aldehyde dehydrogenase might be dependent upon which aldehyde is used in the reaction. Chloramphenicol which also elutes the enzyme from the affinity column, shows a discriminatory effect by inhibiting the ALDH1 oxidation of benzaldehyde and activating that of propionaldehyde while showing no effect when assayed with hexanal or cyclohexane–carboxaldehyde. Chloramphenicol is an uncompetitive inhibitor against NAD when benzaldehyde is the substrate. We propose that this drug might interact with both the benzaldehyde and NAD binding sites.  相似文献   

7.
The anaerobic bacterium Desulfobacterium cetonicum oxidized p-cresol completely to CO2 with sulfate as the electron acceptor. During growth, 4-hydroxybenzylsuccinate accumulated in the medium. This finding indicated that the methyl group of p-cresol is activated by addition to fumarate, analogous to anaerobic toluene, m-xylene, and m-cresol degradation. In cell extracts, the formation of 4-hydroxybenzylsuccinate from p-cresol and fumarate was detected at an initial rate of 0.57 nmol min−1 (mg of protein)−1. This activity was specific for extracts of p-cresol-grown cells. 4-Hydroxybenzylsuccinate was degraded further to 4-hydroxybenzoyl-coenzyme A (CoA), most likely via β-oxidation. 4-Hydroxybenzoyl-CoA was reductively dehydroxylated to benzoyl-CoA. There was no evidence of degradation of p-cresol via methyl group oxidation by p-cresol-methylhydroxylase in this bacterium.  相似文献   

8.
The metabolic capability of denitrifying sludge to oxidize ammonium and p-cresol was evaluated in batch cultures. Ammonium oxidation was studied in presence of nitrite and/or p-cresol by 55 h. At 50 mg/L NH4+-N and 76 mg/L NO2-N, the substrates were consumed at 100% and 95%, respectively, being N2 the product. At 50 mg/L NH4+-N and 133 mg/L NO2-N, the consumption efficiencies decreased to 96% and 70%, respectively. The increase in nitrite concentration affected the ammonium oxidation rate. Nonetheless, the N2 production rate did not change. In organotrophic denitrification, the p-cresol oxidation rate was slower than ammonium oxidation. In litho-organotrophic cultures, the p-cresol and ammonium oxidation rates were affected at 133 mg/L NO2-N. Nonetheless, at 76 mg/L NO2-N the denitrifying sludge oxidized ammonium and p-cresol, but at different rate. Finally, this is the first work reporting the simultaneous oxidation of ammonium and p-cresol with the production of N2 from denitrifying sludge.  相似文献   

9.
The kinetic behavior of a stable nitrifying consortium exposed to 2-chlorophenol (2-CP), phenol, p-cresol and p-hydroxybenzaldehyde (p-OHB) was evaluated in batch assays. Phenolic compounds were evaluated either individually or in mixture. In individual assays, 2-CP inhibited stronger the nitrification, diminishing the ammonium consumption efficiency (16%) and the nitrate production rate (at 91%). Nonetheless, the consumption efficiencies for all phenolics were of 100%. On the other hand, in mixture, the inhibitory effect of 2-CP diminished significantly, since ammonium consumption efficiency and nitrate production rate were improved. Consumption efficiencies for most of the phenolic compounds were high. Furthermore, the kinetic of 2-CP oxidation was 2.4-fold-faster than the individual assays. Finally, the experimental results showed the potential of nitrifying consortium for removing 2-CP, phenol, p-cresol and p-OHB. This is the first work showing the simultaneous removal of these pollutants and also this information might be useful for treating wastewaters of chemical complexity.  相似文献   

10.
The present work describes the development of HPLC-mass spectrometric systems equipped with an electrospray interface for the quantitative analysis of bile acids. Good separation of free as well as glycine- and taurine-conjugated bile acids was achieved with a C18 reversed-phase column (3 μm particle size, 70 × 4.6 mm I.D.) employing methanol-15 mM ammonium acetate as the mobile phase for both isocratic and gradient mode, at a flow-rate of 0.3 ml/min. This system permits post-column splitting of the eluate for analysis by two different detectors: (1) electrospray-mass spectrometer with a flow-rate of 18 μl/min; and (2) a complementary evaporative light scattering mass detector. When bile salts were ionized in the electrospray interface operating in the negative-ion mode, only [M  H] molecular ions were generated; the detection limit was 15 pg injected for all bile acids studied. In the second system, a semi-micro pre-column splitting apparatus (Acurate, LC Packings) was utilized: with this device the flow-rate from the HPLC pump was reduced to 1.4 μl/min and bile acids were separated with a micro-bore C18 column (3 μm particle size, 150 × 0.30 I.D.), using the same mobile phase as above. With this latter system, a head-column enrichment technique can be used: the amount injected can be increased from 60 to 200 nl, permitting an improvement in the detection limit to 5 pg injected. Application of the HPLC-electrospray-mass spectrometric method to bile and serum bile acid analysis is described; preliminary data on the ability of the first system to determine the 13C/12C isotope ratio in 13C-labeled bile acid enriched serum is also critically discussed.  相似文献   

11.
In this study, the effect of the organic micropollutant and known inhibitor of nitrification, p-cresol, was investigated on the metabolism of the ammonia oxidizing bacteria (AOB) Nitrosomonas eutropha C91 using MS-based quantitative proteomics. Several studies have demonstrated that AOB are capable of biotransforming a wide variety of aromatic compounds making them suitable candidates for bioremediation, yet the underlying molecular mechanisms are poorly described. The effect of two different concentrations of the aromatic micropollutant p-cresol (1 and 10 mg L?1) on the metabolism of N. eutropha C91, relative to a p-cresol absent control, was investigated. Though the rate of nitrification in N. eutropha C91 appeared essentially unaffected at both concentrations of p-cresol relative to the control, the expressional pattern of the proteins of N. eutropha C91 changed significantly. The presence of p-cresol resulted in the repressed expression of several key proteins related to N-metabolism, seemingly impairing energy production in N. eutropha C91, contradicting the observed unaltered rates of nitrification. However, the expression of proteins of the TCA cycle and proteins related to xenobiotic degradation, including a p-cresol dehydrogenase, was found to be stimulated by the presence of p-cresol. This indicates that N. eutropha C91 is capable of degrading p-cresol and that it assimilates degradation intermediates into the TCA cycle. The results reveal a pathway for p-cresol degradation and subsequent entry point in the TCA cycle in N. eutropha C91. The obtained data indicate that mixotrophy, rather than cometabolism, is the major mechanism behind p-cresol degradation in N. eutropha C91.  相似文献   

12.
Production of Skatole and para-Cresol by a Rumen Lactobacillus sp.   总被引:7,自引:3,他引:4       下载免费PDF全文
The objective of this study was to examine the substrate specificity of several ruminal strains of a Lactobacillus sp. which previously was shown to produce skatole (3-methylindole) by the decarboxylation of indoleacetic acid. A total of 13 compounds were tested for decarboxylase activity. The Lactobacillus strains produced p-cresol (4-methylphenol) by the decarboxylation of p-hydroxyphenylacetic acid, but did not produce either o-cresol or m-cresol from the corresponding hydroxyphenylacetic acid isomers. These strains also decarboxylated 5-hydroxyindoleacetic acid to 5-hydroxyskatole and 3,4-dihydroxyphenylacetic acid to methylcatechol. Skatole and p-cresol were produced in a 0.5:1 ratio, when indoleacetic acid and p-hydroxyphenylacetic acid were combined in equimolar concentrations. Competition studies with indoleacetic acid and p-hydroxyphenylacetic acid suggested that two different decarboxylating enzymes are involved in the production of skatole and p-cresol by these strains. This is the first demonstration of both skatole production and p-cresol production by a single bacterium.  相似文献   

13.
A. Feutry  R. Letouze 《Phytochemistry》1984,23(8):1557-1559
Hydroxycinnamate: CoA ligase was extracted from stems of in vitro willow cultures and characterized. One peak of activity was obtained after column chromatography on Sephadex G 100 or DEAE Sephacel. p-Coumaric acid gave the highest Vmax among the cinnamates examined. The Kmvalues for p-coumaric, caffeic and ferulic acid were 31.0, 4.7 and 46 μM, respectively. The MW of the CoA ligase was 57 000 and the pH optimum was 7.0. The characteristics of the enzyme correspond to its physiological role in lignin biosynthesis.  相似文献   

14.
The dissimilatory Fe(III) reducer, GS-15, is the first microorganism known to couple the oxidation of aromatic compounds to the reduction of Fe(III) and the first example of a pure culture of any kind known to anaerobically oxidize an aromatic hydrocarbon, toluene. In this study, the metabolism of toluene, phenol, and p-cresol by GS-15 was investigated in more detail. GS-15 grew in an anaerobic medium with toluene as the sole electron donor and Fe(III) oxide as the electron acceptor. Growth coincided with Fe(III) reduction. [ring-14C]toluene was oxidized to 14CO2, and the stoichiometry of 14CO2 production and Fe(III) reduction indicated that GS-15 completely oxidized toluene to carbon dioxide with Fe(III) as the electron acceptor. Magnetite was the primary iron end product during toluene oxidation. Phenol and p-cresol were also completely oxidized to carbon dioxide with Fe(III) as the sole electron acceptor, and GS-15 could obtain energy to support growth by oxidizing either of these compounds as the sole electron donor. p-Hydroxybenzoate was a transitory extracellular intermediate of phenol and p-cresol metabolism but not of toluene metabolism. GS-15 oxidized potential aromatic intermediates in the oxidation of toluene (benzylalcohol and benzaldehyde) and p-cresol (p-hydroxybenzylalcohol and p-hydroxybenzaldehyde). The metabolism described here provides a model for how aromatic hydrocarbons and phenols may be oxidized with the reduction of Fe(III) in contaminated aquifers and petroleum-containing sediments.  相似文献   

15.
Glycosyl hydrolase family 1 β-glucosidases are important enzymes that serve many diverse functions in plants including defense, whereby hydrolyzing the defensive compounds such as hydroxynitrile glucosides. A hydroxynitrile glucoside cleaving β-glucosidase gene (Llbglu1) was isolated from Leucaena leucocephala, cloned into pET-28a (+) and expressed in E. coli BL21 (DE3) cells. The recombinant enzyme was purified by Ni–NTA affinity chromatography. The optimal temperature and pH for this β-glucosidase were found to be 45 °C and 4.8, respectively. The purified Llbglu1 enzyme hydrolyzed the synthetic glycosides, pNPGlucoside (pNPGlc) and pNPGalactoside (pNPGal). Also, the enzyme hydrolyzed amygdalin, a hydroxynitrile glycoside and a few of the tested flavonoid and isoflavonoid glucosides. The kinetic parameters K m and V max were found to be 38.59 μM and 0.8237 μM/mg/min for pNPGlc, whereas for pNPGal the values were observed as 1845 μM and 0.1037 μM/mg/min. In the present study, a three dimensional (3D) model of the Llbglu1 was built by MODELLER software to find out the substrate binding sites and the quality of the model was examined using the program PROCHEK. Docking studies indicated that conserved active site residues are Glu 199, Glu 413, His 153, Asn 198, Val 270, Asn 340, and Trp 462. Docking of rhodiocyanoside A with the modeled Llbglu1 resulted in a binding with free energy change (ΔG) of ?5.52 kcal/mol on which basis rhodiocyanoside A could be considered as a potential substrate.  相似文献   

16.
A new method of assay for the A6-desaturation of linoleic acid was developed. This method, which uses HPLC for separation of the fatty acid substrate and product, exhibited a lower coefficient of variation (0.3%) than the reported TLC method (3.5%) [l], and avoided the step of methylation of the saponified fatty acid substrate and product. Using this new method of assay, the kinetics of the Δ6-desaturase in a multi-enzyme system were analysed. A number of factors that could have striking effects on desaturase kinetics were investigated, including the effect of (i) endogenous microsomal linoleic acid on total substrate concentration, and (ii) the pre-reaction catalysed by acyl-CoA synthetase and competing reactions catalysed by lysophospholipid acyltransferase and acyl-CoA hydrolase. Endogenous free linoleate in the hepatic microsomes was found to be 2.9 ± l.0 AM (0.5 mg microsomal protein/ml), which was comparable to added substrate concentrations (l.8 to 7.9 μM). The kinetics of the Δ6-desaturase were dissected from the kinetics of the above mentioned pre-reaction and competing reactions through a combination of experimental approaches and computer modeling. From computer modeling, a Km and Vmax of l.5 μM and 0.063 nmol/min were calculated for the Δ6-desaturase, compared to Km and Vmax of 10.7 μM and 0.08 nmol/min calculated directly from data uncorrected for endogenous substrate. It was concluded that lysophospholipid acyltransferase, acyl-CoA synthetase and endogenous linoleic acid significantly affect the kinetic measurements of hepatic microsomal Δ6-desaturase. These results have implications for kinetic analyses of all desaturases in microsomal systems.  相似文献   

17.
There is limited knowledge of interspecies interactions in biofilm communities. In this study, Pseudomonas sp. strain GJ1, a 2-chloroethanol (2-CE)-degrading organism, and Pseudomonas putida DMP1, a p-cresol-degrading organism, produced distinct biofilms in response to model mixed waste streams composed of 2-CE and various p-cresol concentrations. The two organisms maintained a commensal relationship, with DMP1 mitigating the inhibitory effects of p-cresol on GJ1. A triple-labeling technique compatible with confocal microscopy was used to investigate the influence of toxicant concentrations on biofilm morphology, species distribution, and exopolysaccharide production. Single-species biofilms of GJ1 shifted from loosely associated cell clusters connected by exopolysaccharide to densely packed structures as the p-cresol concentrations increased, and biofilm formation was severely inhibited at high p-cresol concentrations. In contrast, GJ1 was abundant when associated with DMP1 in a dual-species biofilm at all p-cresol concentrations, although at high p-cresol concentrations it was present only in regions of the biofilm where it was surrounded by DMP1. Evidence in support of a commensal relationship between DMP1 and GJ1 was obtained by comparing GJ1-DMP1 biofilms with dual-species biofilms containing GJ1 and Escherichia coli ATCC 33456, an adhesive strain that does not mineralize p-cresol. Additionally, the data indicated that only tower-like cell structures in the GJ1-DMP1 biofilm produced exopolysaccharide, in contrast to the uniform distribution of EPS in the single-species GJ1 biofilm.  相似文献   

18.
A new high-performance liquid chromatographic method for simultaneous determination of primidone (PRM) and of its three major metabolites, phenobarbital (PB), p-hydroxyphenobarbital (p-HO-PB) and phenylethylmalonamide (PEMA), in rat urine, was developed. After acid hydrolysis, these compounds were extracted from urine by means of a Bond Elut Certify LRC column with good clean-up. The extracts were chromatographed on a C18 reversed-phase column using isocratic elution at 40°C, with UV detection at 227 nm. The limit of detection was 0.5 mg/ml for the four compounds. Good linearity (r2>0.99) was observed within the calibration ranges studied: 37.4–299.3 μg/ml for PRM, 26.4–211.2 μg/ml for PB, 12.5–100.2 μg/ml for p-HO-PB and 12.1–97.0 μg/ml for PEMA. Repeatability was in the range 3.1–6.8%. This method constitutes a useful tool for studies on the influence of various parameters on primidone metabolism.  相似文献   

19.
A bacterial dipeptidyl carboxypeptidase inhibitor was isolated from the culture broth of a bacterium identified as Bacillus subtilis. The inhibitor was purified 33-fold from the culture supernatant of B. subtillis 3-16-20 strain by Q-, and S-Sepharose fast flow, C18 column chromatography, ethanol treatment, and ODS column chromatography. The purified inhibitor has an amino acid sequence of glycyl-prolyl-phenylalanyl-prolylisoleucine. IC50 values of the inhibitor were 177 μM (rabbit lung ACE) and 35 μM (bacterial DCP).  相似文献   

20.
Wild-type toluene 4-monooxygenase (T4MO) of Pseudomonas mendocina KR1 oxidizes toluene to p-cresol (96%) and oxidizes benzene sequentially to phenol, to catechol, and to 1,2,3-trihydroxybenzene. In this study T4MO was found to oxidize o-cresol to 3-methylcatechol (91%) and methylhydroquinone (9%), to oxidize m-cresol and p-cresol to 4-methylcatechol (100%), and to oxidize o-methoxyphenol to 4-methoxyresorcinol (87%), 3-methoxycatechol (11%), and methoxyhydroquinone (2%). Apparent Vmax values of 6.6 ± 0.9 to 10.7 ± 0.1 nmol/min/ mg of protein were obtained for o-, m-, and p-cresol oxidation by wild-type T4MO, which are comparable to the toluene oxidation rate (15.1 ± 0.8 nmol/min/mg of protein). After these new reactions were discovered, saturation mutagenesis was performed near the diiron catalytic center at positions I100, G103, and A107 of the alpha subunit of the hydroxylase (TmoA) based on directed evolution of the related toluene o-monooxygenase of Burkholderia cepacia G4 (K. A. Canada, S. Iwashita, H. Shim, and T. K. Wood, J. Bacteriol. 184:344-349, 2002) and a previously reported T4MO G103L regiospecific mutant (K. H. Mitchell, J. M. Studts, and B. G. Fox, Biochemistry 41:3176-3188, 2002). By using o-cresol and o-methoxyphenol as model substrates, regiospecific mutants of T4MO were created; for example, TmoA variant G103A/A107S produced 3-methylcatechol (98%) from o-cresol twofold faster and produced 3-methoxycatechol (82%) from 1 mM o-methoxyphenol seven times faster than the wild-type T4MO (1.5 ± 0.2 versus 0.21 ± 0.01 nmol/min/mg of protein). Variant I100L produced 3-methoxycatechol from o-methoxyphenol four times faster than wild-type T4MO, and G103S/A107T produced methylhydroquinone (92%) from o-cresol fourfold faster than wild-type T4MO and there was 10 times more in terms of the percentage of the product. Variant G103S produced 40-fold more methoxyhydroquinone from o-methoxyphenol than the wild-type enzyme produced (80 versus 2%) and produced methylhydroquinone (80%) from o-cresol. Hence, the regiospecific oxidation of o-methoxyphenol and o-cresol was changed for significant synthesis of 3-methoxycatechol, methoxyhydroquinone, 3-methylcatechol, and methylhydroquinone. The enzyme variants also demonstrated altered monohydroxylation regiospecificity for toluene; for example, G103S/A107G formed 82% o-cresol, so saturation mutagenesis converted T4MO into an ortho-hydroxylating enzyme. Furthermore, G103S/A107T formed 100% p-cresol from toluene; hence, a better para-hydroxylating enzyme than wild-type T4MO was formed. Structure homology modeling suggested that hydrogen bonding interactions of the hydroxyl groups of altered residues S103, S107, and T107 influence the regiospecificity of the oxygenase reaction.  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