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 photo-induced transformation of monuron (3-(4-chlorophenyl)-1,1 dimethylurea) was investigated in an aqueous solution containing nitrates and nitrites at 310 nm and 365 nm, respectively. In both NO(3)(-) and NO(2)(-) conditions, the degradation of monuron followed pseudo-first order kinetics. The intermediate products were identified by GC-MS, and the nitration, hydroxylation and coupling reactions were determined. In addition, the oxidation of the N-terminus group, the substitution of chlorine by ˙OH and the nitration by ˙NO(2) radical onto the phenyl ring were observed. The photo-induced transformation of monuron was studied under variable conditions of pH, inducer concentration, substrate concentration, humic acids, oxygen content and salts used as hydroxyl radical scavengers. The photodegradation rates were strongly influenced by all the above parameters. The degradation of monuron was also studied in the dark and in the presence of NO(2)(-) as well as in an aqueous solution with the addition of hydrogen peroxide.  相似文献   

2.
A bacterium, CP1, identified as Pseudomonas putida strain, was investigated for its ability to grow on and degrade mono-chlorophenols and phenols as sole carbon sources in aerobic shaking batch culture. The organism degraded up to 1.56 mM 2- and 3-chlorophenol, 2.34 mM 4-chlorophenol and 8.5 mM phenol using an ortho-cleavage pathway. P. putida CP1, acclimated to degrade 2-chlorophenol, was capable of 3-chlorocatechol degradation, while P. putida, acclimated to 4-chlorophenol degradation, degraded 4-chlorocatechol. Growth of P. putida CP1 on higher concentrations of the mono-chlorophenols, ≥1.56 mM 4-chlorophenol and ≥0.78 mM 2- and 3-chlorophenol, resulted in decreases in cell biomass despite metabolism of the substrates, and the formation of large aggregates of cells in the culture medium. Increases in cell biomass with no clumping of the cells resulted from growth of P. putida CP1 on phenol or on lower concentrations of mono-chlorophenol. Bacterial adherence to hydrocarbons (BATH) assays showed cells grown on the higher concentrations of mono-chlorophenol to be more hydrophobic than those grown on phenol and lower concentrations of mono-chlorophenol. The results suggested that increased hydrophobicity and autoaggregation of P. putida CP1 were a response to toxicity of the added substrates. Journal of Industrial Microbiology & Biotechnology (2002) 28, 316–324 DOI: 10.1038/sj/jim/7000249 Received 27 June 2001/ Accepted in revised form 09 February 2002  相似文献   

3.
In this study the chlorophenol-degrading actinobacterium, Arthrobacter chlorophenolicus A6, was tested for its ability to grow on mixtures of phenolic compounds. During the experiments depletion of the compounds was monitored, as were cell growth and activity. Activity assays were based on bioluminescence output from a luciferase-tagged strain. When the cells were grown on a mixture of 4-chlorophenol, 4-nitrophenol and phenol, 4-chlorophenol degradation apparently was delayed until 4-nitrophenol was almost completely depleted. Phenol was degraded more slowly than the other compounds and not until 4-nitrophenol and 4-chlorophenol were depleted, despite this being the least toxic compound of the three. A similar order of degradation was observed in non-sterile soil slurries inoculated with A. chlorophenolicus. The kinetics of degradation of the substituted phenols suggest that the preferential order of their depletion could be due to their respective pKa values and that the dissociated phenolate ions are the substrates. A mutant strain (T99), with a disrupted hydroxyquinol dioxygenase gene in the previously described 4-chlorophenol degradation gene cluster, was also studied for its ability to grow on the different phenols. The mutant strain was able to grow on phenol, but not on either of the substituted phenols, suggesting a different catabolic pathway for the degradation of phenol by this microorganism.  相似文献   

4.
We investigated the cellular responses of the white-rot basidiomycete Phanerochaete chrysosporium against vanillin. Based upon a proteomic survey, it was demonstrated that two flavin-containing monooxygenases (PcFMO1 and PcFMO2) are translationally up-regulated in response to exogenous addition of vanillin. To elucidate their catalytic functions, we cloned cDNAs and heterologously expressed them in Escherichia coli. The recombinant PcFMO1 showed catalytic activities against monocyclic phenols such as phenol, hydroquinone, and 4-chlorophenol. In addition, the product from hydroquinone was identified as 1,2,4-trihydroxybenzene, an important intermediate in a metabolic pathway of aromatic compounds in which the aromatic ring of 1,2,4-trihydroxybenzene can be further cleaved by fungal dioxygenases for mineralization. Thus, the ortho-cleavage pathway of phenolic compounds would presumably be associated with PcFMO1.  相似文献   

5.
To investigate the enhancement of phenol on the biodegradation of 4-chlorophenol (4-cp), metabolic profiling approach was performed for the first time to analyze metabolite changes of Pseudomonas sp. cbp1-3 using single substrate (succinate, phenol, and 4-cp) and dual substrate (mixtures of phenol and 4-cp). Phosphoric acid, γ-aminobutyric acid, 4-cp, 4-chlorocatechol, and catechol were shown to change significantly. Results indicated that phenols, especially 4-cp, depressed cell growth by inhibiting its primary metabolic pathway. In addition, the addition of phenol into the 4-cp-containing medium had a global influence on cells including the accumulation of amino acids, amines, saturated fatty acids, and monoacylglycerols as well as the concentration changes of metabolite participating in phenols biodegradation, thus enhancing the degradation of 4-cp. This study provided novel insights into the biodegradation of mixed phenolic compounds and the method could be used to investigate the biodegradation of complicated multi-pollutants.  相似文献   

6.
The sulfonephthalein indicator, phenol red, exhibits an unusually slow rate of oxidation by laccase from Poliporus pinsitus, in spite of the fact that it is a phenol and therefore a natural substrate for this phenoloxidase enzyme. Nevertheless, after prolonged exposure to laccase (24 h) phenol red is oxidized by more than 90%. We found that phenol red, which can be oxidatively converted into a resonance-stabilized phenoxy radical, performs as a mediator in the laccase-catalyzed oxidation of a nonphenolic substrate (4-methoxybenzyl alcohol) and also of a hindered phenol (2,4,6-tri-tert-butylphenol). In particular, phenol red was found to be at least 10 times more efficient than 3-hydroxyanthranilate (a reported natural phenolic mediator of laccase) in the oxidation of 4-methoxybenzyl alcohol. Other phenols, which do not bear structural analogies to phenol red, underwent rapid degradation and did not perform as laccase mediators. On the other hand, several variously substituted sulfonephthaleins, of different pK2 values, mediated the laccase catalysis, the most efficient being dichlorophenol red, which has the lowest pK2 of the series. The mediating efficiency of phenol red and dichlorophenol red was found to be pH dependent, as was their oxidation Ep value (determined by cyclic voltammetry). We argue that the relative abundance of the phenoxy anion, which is easier to oxidize than the protonated phenol, may be one of the factors determining the efficiency of a phenolic mediator, together with its ability to form relatively stable oxidized intermediates that react with the desired substrate before being depleted in undesired routes.  相似文献   

7.
The excited triplet state of 1-nitronaphthalene ((3)1NN*) reacts with OH(-) with a second-order reaction rate constant of (1.66 ± 0.08)×10(7) M(-1) s(-1) (μ±σ). The reaction yields the ˙OH radical and the radical anion 1NN(-)˙. In aerated solution, the radical 1NN(-)˙ would react with O(2) to finally produce H(2)O(2) upon hydroperoxide/superoxide disproportionation. The photolysis of H(2)O(2) is another potential source of ˙OH, but such a pathway would be a minor one in circumneutral (pH 6.5) or in basic solution ([OH(-)] = 0.3-0.5 M). The oxidation of H(2)O by (3)1NN*, with rate constant 3.8 ± 0.3 M(-1) s(-1), could be the main ˙OH source at pH 6.5.  相似文献   

8.
The Cu(II) and Zn(II) complexes of phenoxyl radical species [M(II)(L1*)(NO3)]+ (M=Cu or Zn, L1H: 2-methylthio-4-tert-butyl-6-[[bis[2-(2-pyridyl)ethyl]amino]methyl]phenol ) and [M(II)(L2*)(NO3)]+ (M=Cu or Zn, L2H: 2,4-di-tert-butyl-6-[[bis[2-(2-pyridyl)ethyl]amino]methyl]phenol) are prepared as model complexes of the active form of galactose oxidase (GAO). Hydrogen atom abstraction of 1,4-cyclohexadiene and tert-butyl substituted phenols by the GAO model complexes proceeds very efficiently to give benzene and the corresponding phenoxyl radical or its C-C coupling dimer as the oxidation products, respectively. Kinetic analyses on the oxidation reactions have shown that the hydrogen atom abstraction of the phenol substrates is significantly enhanced by the coordinative interaction of the OH group to the metal ion center of the complex, providing valuable insight into the enzymatic mechanism of the alcohol oxidation. Details of the substrate-activation process have been discussed based on the activation parameters (deltaH* and deltaS*) of the reactions.  相似文献   

9.
Enhanced absorption is observed in the (15)N NMR spectra of (15)NO(-)(3) during decomposition of peroxynitrite and the peroxynitrite-CO(2) adduct at pH 5.25, indicating the formation of (15)NO(-)(3) in radical pairs [(15)NO(*)(2), HO(*)] and [(15)NO(*)(2), CO(*-)(3)]. During the reaction of peroxynitrite and the peroxynitrite-CO(2) adduct with L-tyrosine, the (15)N NMR signal of the nitration product 3-nitrotyrosine exhibits emission showing a radical pathway of its formation. The nuclear polarization is built up in radical pairs [(15)NO(*)(2), tyr(*)] generated by free radical encounters of nitrogen dioxide and tyrosinyl radicals. The (15)N NMR signal of (15)NO(-)(2) formed during reaction of peroxynitrite with L-tyrosine appears in emission. It is concluded that tyrosinyl radicals are generated by reaction of nitrogen dioxide with L-tyrosine. In contrast to this, (15)NO(-)(2) does not show (15)N chemically induced dynamic nuclear polarization (CIDNP) during reaction of the peroxynitrite-CO(2) adduct with L-tyrosine, indicating a different reaction mechanism, which is assumed to be a hydrogen transfer between CO(*-)(3) and L-tyrosine. Emission is also observed in the (15)N NMR signals of 2-nitro-4-fluorophenol, 3-nitro-4-hydroxyphenylacetic acid, 2-nitrophenol, and 4-nitrophenol during reaction of 4-fluorophenol, 4-hydroxyphenylacetic acid, and phenol with peroxynitrite and the peroxynitrite-CO(2) adduct. 3-Nitro-4-hydroxyphenylacetic acid is also observed in emission during reaction of phenylacetic acid with peroxynitrite, but is not formed with the peroxynitrite-CO(2) adduct. The magnitude of the (15)N CIDNP effect during reaction of peroxynitrite with 4-fluorophenol and of the peroxynitrite-CO(2) adduct with 4-fluorophenol and phenol is determined. It excludes the occurrence of nonradical reactions. Only weak emission signals are observed during the reaction of peroxynitrite with phenol in (15)NO(-)(2), 2-nitrophenol, and 4-nitrophenol. 2-Nitrophenol is only formed in traces, and 4-nitrophenol is only formed in higher yields. The latter might be generated in part via a nonradical pathway.  相似文献   

10.
We have reported that low levels of peroxynitrite (PN) can cause inactivation of the heme-thiolate protein prostacyclin (PGI2)-synthase by nitration of a tyrosine residue. To prove that iron catalysis is involved we studied the interaction of PN with microperoxidase and P450nor, a heme-thiolate protein of known structure. Spectral and kinetic analyses allow to conclude on a ferryl nitrogen dioxide complex as an intermediate which decomposes in the presence of an excess of PN under formation of dioxygen, nitrite, and nitrate. This occurs in a catalytic cycle which was more efficient with P450nor than with microperoxidase. If phenol was added to the reaction mixtures of PN and the ferric complexes the ratio of hydroxylated to nitrated phenols decreased compared to the metal-free system. Phenol competed with the formation of dioxygen indicating that the ferryl intermediate was involved in both pathways. One therefore can postulate that the ferryl complex reacts with phenol to give the phenoxyradical which is nitrated in the presence of nitrogen dioxide but does not give hydroxylated products as with metal-free PN. Alternately, the ferryl nitrogen dioxide complex can oxidize a second PN molecule to the radical, *OONO, which can decompose to dioxygen and NO. The latter forms N2O3, with the remaining *NO2 radical. A third pathway consists in the isomerization to nitrate which also is catalyzed by the heme proteins since the ratio of nitrite/nitrate does not change significantly during the catalytic reaction with excess of PN. Our data explain the mechanism of nitration of PGI2-synthase, suggest a role of P450nor as a PN scavenger, and favor heme-thiolate complexes for trapping PN.  相似文献   

11.
The free radical intermediates of phenol derivatives, produced by the methemoglobin-hydrogen peroxide system at pH 5 and 7, are detected by electron paramagnetic resonance equipped with a continuous-flow apparatus. All the radicals from phenols are the phenoxy radicals, as identified by analyzing the observed hyperfine structures of the spectra with the aid of SCF-LCAO MO calculations. Comparing with the reaction of Fenton's reagent, it is concluded that free OH radical, even if it exists, is not liberated into the solution in the methemoglobin-hydrogen peroxide system.  相似文献   

12.
Wang Y  Jiang X  Liu J  Zhao M  Kang G  Wu J  Peng L  Peng S 《Molecular bioSystems》2011,7(5):1678-1683
Nitronyl nitroxides are capable of preventing cells, tissues, and organs from radical-induced damage through scavenging NO˙, ˙O(2)(-) and ˙OH. In order to explore the conversions of nitronyl nitroxides in biological systems with and without NO˙, HPLC-MS aided PC12 cell systems were developed, and the conversions of 2-(3'-nitrophenyl)-4,4,5,5-tetramethylimidazoline-1-oxyl -3-oxide (3-nitro-PTIO), 1-oxyl-2-(3'-nitrophenyl)-4,4,5,5-tetramethylimidazoline (3-nitro-PTI), and 1-hydroxyl-2-(3'-nitrophenyl)-4,4,5,5-tetramethylimidazoline (3-nitro-PTIH) were quantitatively monitored. In these systems 3-nitro-PTIO and 3-nitro-PTI were time-dependently converted to 3-nitro-PTIH, while no conversion of 3-nitro-PTIH was detected. Free radical NO˙ donors (sodium nitroprusside, SNP) accelerated the conversions, but had no effect upon the conversion product. In the in vitro and in vivo assays the 3-nitro-PTIH treated cells and mice exhibited no toxic response.  相似文献   

13.
The α-class carbonic anhydrases (CAs, EC 4.2.1.1) from the bacterial pathogens Neisseria gonorrhoeae (NgCAα) and Vibrio cholerae (VchCAα) were investigated for their inhibition by a panel of phenols and phenolic acids. Mono-, di- and tri-substituted phenols incorporating additional hydroxyl/hydroxymethyl, amino, acetamido, carboxyl, halogeno and carboxyethenyl moieties were included in the study. The best NgCAα inhibitrs were phenol, 3-aminophenol, 4-hydroxy-benzylalcohol, 3-amino-4-chlorophenol and paracetamol, with KI values of 0.6–1.7 µM. The most effective VchCAα inhibitrs were phenol, 3-amino-4-chlorophenol and 4-hydroxy-benzyl-alcohol, with KI values of 0.7–1.2 µM. Small changes in the phenol scaffold led to drastic effects on the bacterial CA inhibitory activity. This class of underinvestigated bacterial CA inhibitors may thus lead to effective compounds for fighting drug resistant bacteria.  相似文献   

14.
K Wimalasena  K R Alliston 《Biochemistry》1999,38(45):14916-14926
Previous studies have shown that the dopamine beta-monooxygenase (DbetaM; E.C. 1.14.17.1)/1-(2-aminoethyl)-1,4-cyclohexadiene (CHDEA) reaction partitions between side chain and ring H-abstraction to produce the side-chain-hydroxylated product, 2-amino-1-(1, 4-cyclohexadienyl)ethanol, and the aromatized product, phenylethylamine, and that the two pathways do not crossover. [Wimalasena, K., and May, S. W. (1989) J. Am. Chem. Soc. 111, 2729-2731; Wimalasena, K., and Alliston, K. R. (1995) J. Am. Chem. Soc. 117, 1220-1224]. We now report that the ring H-abstraction pathway of the reaction further partitions to produce the ring hydroxylated product, CHDEA-6OH, and the aromatized product, PEA, at the carbon-oxygen bond formation step. The ring hydroxylation is shown to be stereospecific, exclusively producing the (S) product. The absolute stereospecificity of the ring and side-chain hydroxylations of the DbetaM/CHDEA reaction suggests that the side-chain pro-R hydrogen of the enzyme-bound substrate is close to perpendicular to the aromatic ring of the phenylethylamine substrate or cyclohexadiene ring of CHDEA. The relative activation energy parameters suggest that the partitioning of the ring H abstraction pathway between aromatized and ring hydroxylated products is due to the partitioning of the high-energy intermediates, the cyclohexadienyl radical and the Cu(II)-O(*) species, between carbon-oxygen bond formation and direct electron transfer. The relatively high activation enthalpic favorability and entropic unfavorability for the carbon-oxygen bond formation strongly suggest that the critical balancing of these two opposing forces is mandatory for the desired product formation.  相似文献   

15.
Peroxynitrite (ONOO(-)), a reactive nitrogen species, is capable of nitrating tyrosine residue of proteins. Here we show in vitro evidence that plant phenolic compounds can also be nitrated by an ONOO(-)-independent mechanism. In the presence of NaNO(2), H(2)O(2), and horseradish peroxidase (HRP), monophenolic p-coumaric acid (p-CA, 4-hydroxycinnamic acid) was nitrated to form 4-hydroxy-3-nitrocinnamic acid. The reaction was completely inhibited by KCN, an inhibitor for HRP. The antioxidant ascorbate suppressed p-CA nitration and its suppression time depended strongly on ascorbate concentration. We conclude that nitrogen dioxide radical (NO(2)(radical)), but not ONOO(-), produced by a guaiacol peroxidase is the intermediate for phytophenolic nitration.  相似文献   

16.
A solid–liquid two-phase partitioning bioreactor (TPPB) in which the non-aqueous phase consisted of polymer (HYTREL) beads was used to degrade a model mixture of phenols [phenol, o-cresol, and 4-chlorophenol (4CP)] by a microbial consortium. In one set of experiments, high concentrations (850 mg l−1 of each of the three substrates) were reduced to sub-inhibitory levels within 45 min by the addition of the polymer beads, followed by inoculation and rapid (8 h) consumption of the total phenolics loading. In a second set of experiments, the beneficial effect of using polymer beads to launch a fermentation inhibited by high substrate concentrations was demonstrated by adding 1,300 and 2,000 mg l−1 total substrates (equal concentrations of each phenolic) to a pre-inoculated bioreactor. At these levels, no cell growth and no degradation were observed; however, after adding polymer beads to the systems, the ensuing reduced substrate concentrations permitted complete destruction of the target molecules, demonstrating the essential role played by the polymer sequestering phase when applied to systems facing inhibitory substrate concentrations. In addition to establishing alternative modes of TPPB operation, the present work has demonstrated the differential partitioning of phenols in a mixture between the aqueous and polymeric phases. The polymeric phase was also observed to absorb a degradation intermediate (arising from the incomplete biodegradation of 4CP), which opens the possibility of using solid–liquid TPPBs during biosynthetic transformation to sequester metabolic byproducts.  相似文献   

17.
Summary The biodegradation of the three isomeric monochlorophenols 2-(2CP), 3- (3CP) and 4-chlorophenol (4CP) and phenol by the constructed strain Alcaligenes sp. A7-2 was investigated. Mineralization took place in the order: phenol >4CP >2CP >3CP, whereas 3CP was mineralized only co-metabolically. In substrate mixtures with phenol, degradation of 4CP was decelerated but degradation of 2CP was accelerated. Free cells in batch culture showed biphasic growth with an equimolar mixture of 2CP and 4CP as substrates, perhaps due to diauxie. Degradation patterns obtained with free cells in batch culture were confirmed with immobilized cells in continuous culture. Immobilized cells of Alcaligenes sp. A7-2 built up a biofilm on the lava that was used as filling material in the packed-bed reactors. The continuous cultures remained stable despite increasing input rates of chlorophenol and phenol mixtures up to 1.16 mMo1.1–1.h–1 for several weeks. Correspondence to: H.-J. Rehm  相似文献   

18.
Nitration of protein tyrosine residues to 3-nitrotyrosine (NO2Tyr) serves as both a marker and mediator of pathogenic reactions of nitric oxide (*NO), with peroxynitrite (ONOO-) and leukocyte peroxidase-derived nitrogen dioxide (*NO2) being proximal mediators of nitration reactions in vivo. Cytochrome c is a respiratory and apoptotic signaling heme protein localized exofacially on the inner mitochondrial membrane. We report herein a novel function for cytochrome c as a catalyst for nitrite (NO2-) and hydrogen peroxide (H2O2)-mediated nitration reactions. Cytochrome c catalyzes both self- and adjacent-molecule (hydroxyphenylacetic acid, Mn-superoxide dismutase) nitration via heme-dependent mechanisms involving tyrosyl radical and *NO2 production, as for phagocyte peroxidases. Although low molecular weight phenolic nitration yields were similar for cytochrome c and the proteolytic fragment of cytochrome c microperoxidase-11 (MPx-11), greater extents of protein nitration occurred when MPx-11 served as catalyst. Partial proteolysis of cytochrome c increased both the peroxidase and nitrating activities of cytochrome c. Extensive tyrosine nitration of Mn-superoxide dismutase occurred when exposed to either cytochrome c or MPx-11 in the presence of H2O2 and NO2-, with no apparent decrease in catalytic activity. These results reveal a post-translational tyrosine modification mechanism that is mediated by an abundant hemoprotein present in both mitochondrial and cytosolic compartments. The data also infer that the distribution of specific proteins capable of serving as potent catalysts of nitration can lend both spatial and molecular specificity to biomolecule nitration reactions.  相似文献   

19.
The type and the position of the substituent in the phenolic ring, the bond dissociation energy and the exogenously supplied carbon source as well as the inductive and resonance effect phenomena of the substituents adjust the biodegradability of the phenolic compounds. The comparative biodegradation study of mono-nitrophenols (electron acceptors) and mono-methylphenols (electron donors) revealed that it is a completely photoregulated process. The closer the donor group (OH(-)) of the phenolic ring is to the acceptor group (NO(2)(-)), the higher the biodegradation values are (2-nitrophenol>3-nitrophenol>4-nitrophenol); the further the donor group (OH(-)) of the phenolic compound is from the second donor group (CH(3)(+)), the higher the biodegradation values are (2-methylphenol<3-methylphenol<4-methylphenol). However, there are compounds without a specific role of acceptor or donor such as mono-iodophenols, where a type of counteraction between the inductive and resonance effect determines the behavior of the substituent. This fact combined with the presence of the hydroxyl group in the phenolic ring gave the observed stabilization in the biodegradation results of mono-iodophenols (2-iodophenol approximately 3-iodophenol approximately 4-iodophenol).  相似文献   

20.
The removal of phenol (Ph) and 2-chlorophenol (2-CPh) from aqueous solution by native and heat inactivated fungus Funaliatrogii pellets were investigated. The effects of contact time, solid/liquid ratio, optimum pH and temperature on the phenols removal capacity by the pellets were established. The removal efficiency of phenols increased significantly with increasing biomass dose. The optimum pH was detected to be 8.0. The second-order equations are described and evaluated on the basis of a comparative estimation of the corresponding coefficients. The phenol removal equilibrium isotherm was modeled by the Langmuir equations. The enthalpy change values were obtained between −7.62 and −10.64 kJ/mol. This indicated that the uptake of phenols either on native or heat inactivated fungal pellets was based on a physical adsorption process.  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