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.
Sphingomonas xenophaga Bayram, isolated from the activated sludge of a municipal wastewater treatment plant, was able to utilize 4-(1-ethyl-1,4-dimethylpentyl)phenol, one of the main isomers of technical nonylphenol mixtures, as a sole carbon and energy source. The isolate degraded 1 mg of 4-(1-ethyl-1,4-dimethylpentyl)phenol/ml in minimal medium within 1 week. Growth experiments with five nonylphenol isomers showed that the three isomers with quaternary benzylic carbon atoms [(1,1,2,4-tetramethylpentyl)phenol, 4-(1-ethyl-1,4-dimethylpentyl)phenol, and 4-(1,1-dimethylheptyl)phenol] served as growth substrates, whereas the isomers containing one or two hydrogen atoms in the benzylic position [4-(1-methyloctyl)phenol and 4-n-nonylphenol] did not. However, when the isomers were incubated as a mixture, all were degraded to a certain degree. Differential degradation was clearly evident, as isomers with more highly branched alkyl side chains were degraded much faster than the others. Furthermore, the C9 alcohols 2,3,5-trimethylhexan-2-ol, 3,6-dimethylheptan-3-ol, and 2-methyloctan-2-ol, derived from the three nonylphenol isomers with quaternary benzylic carbon atoms, were detected in the culture fluid by gas chromatography-mass spectrometry, but no analogous metabolites could be found originating from 4-(1-methyloctyl)phenol and 4-n-nonylphenol. We propose that 4-(1-methyloctyl)phenol and 4-n-nonylphenol were cometabolically transformed in the growth experiments with the mixture but that, unlike the other isomers, they did not participate in the reactions leading to the detachment of the alkyl moiety. This hypothesis was corroborated by the observed accumulation in the culture fluid of an as yet unidentified metabolite derived from 4-(1-methyloctyl)phenol.  相似文献   

2.
Several nonylphenol isomers with alpha-quaternary carbon atoms serve as growth substrates for Sphingomonas xenophaga Bayram, whereas isomers containing hydrogen atoms at the alpha-carbon do not. Three metabolites of 4-(1-methyloctyl)-phenol were isolated in mg quantities from cultures of strain Bayram supplemented with the growth substrate isomer 4-(1-ethyl-1,4-dimethyl-pentyl)-phenol. They were unequivocally identified as 4-hydroxy-4-(1-methyl-octyl)-cyclohexa-2,5-dienone, 4-hydroxy-4-(1-methyl-octyl)-cyclohex-2-enone, and 2-(1-methyl-octyl)-benzene-1,4-diol by high pressure liquid chromatography-mass spectrometry and nuclear magnetic resonance spectroscopy. Furthermore, two metabolites originating from 4-n-nonylphenol were identified as 4-hydroxy-4-nonyl-cyclohexa-2,5-dienone and 4-hydroxy-4-nonyl-cyclohex-2-enone by high pressure liquid chromatography-mass spectrometry. We conclude that nonylphenols were initially hydroxylated at the ipso-position forming 4-alkyl-4-hydroxy-cyclohexa-2,5-dienones. Dienones originating from growth substrate nonylphenol isomers underwent a rearrangement that involved a 1,2-C,O shift of the alkyl moiety as a cation to the oxygen atom of the geminal hydroxy group yielding 4-alkoxyphenols, from which the alkyl moieties can be easily detached as alcohols by known mechanisms. Dienones originating from nongrowth substrates did not undergo such a rearrangement because the missing alkyl substituents at the alpha-carbon atom prevented stabilization of the putative alpha-carbocation. Instead they accumulated and subsequently underwent side reactions, such as 1,2-C,C shifts and dihydrogenations. The ipso-hydroxylation and the proposed 1,2-C,O shift constitute key steps in a novel pathway that enables bacteria to detach alpha-branched alkyl moieties of alkylphenols for utilization of the aromatic part as a carbon and energy source.  相似文献   

3.
Recently we showed that degradation of several nonylphenol isomers with α-quaternary carbon atoms is initiated by ipso-hydroxylation in Sphingobium xenophagum Bayram (F. L. P. Gabriel, A. Heidlberger, D. Rentsch, W. Giger, K. Guenther, and H.-P. E. Kohler, J. Biol. Chem. 280:15526-15533, 2005). Here, we demonstrate with 18O-labeling experiments that the ipso-hydroxy group was derived from molecular oxygen and that, in the major pathway for cleavage of the alkyl moiety, the resulting nonanol metabolite contained an oxygen atom originating from water and not from the ipso-hydroxy group, as was previously assumed. Our results clearly show that the alkyl cation derived from the α-quaternary nonylphenol 4-(1-ethyl-1,4-dimethyl-pentyl)-phenol through ipso-hydroxylation and subsequent dissociation of the 4-alkyl-4-hydroxy-cyclohexadienone intermediate preferentially combines with a molecule of water to yield the corresponding alcohol and hydroquinone. However, the metabolism of certain α,α-dimethyl-substituted nonylphenols appears to also involve a reaction of the cation with the ipso-hydroxy group to form the corresponding 4-alkoxyphenols. Growth, oxygen uptake, and 18O-labeling experiments clearly indicate that strain Bayram metabolized 4-t-butoxyphenol by ipso-hydroxylation to a hemiketal followed by spontaneous dissociation to the corresponding alcohol and p-quinone. Hydroquinone effected high oxygen uptake in assays with induced resting cells as well as in assays with cell extracts. This further corroborates the role of hydroquinone as the ring cleavage intermediate during degradation of 4-nonylphenols and 4-alkoxyphenols.  相似文献   

4.
The aquatic hyphomycete Clavariopsis aquatica was used to quantify the effects of extracellular laccase and intracellular reactions on the isomer-specific biotransformation of technical nonylphenol (t-NP). In laccase-producing cultures, maximal removal rates of t-NP and the isomer 4-(1-ethyl-1,4-dimethylpentyl)phenol (NP112) were about 1.6- and 2.4-fold higher, respectively, than in laccase-lacking cultures. The selective suppression of either laccase or intracellular reactions resulted in essentially comparable maximal removal rates for both compounds. Evidence for an unspecific oxidation of t-NP isomers was consistently obtained from laccase-expressing fungal cultures when intracellular biotransformation was suppressed and from reaction mixtures containing isolated laccase. This observation contrasts with the selective degradation of t-NP isomers by bacteria and should prevent the enrichment of highly estrogenic isomers in remaining t-NP. In contrast with laccase reactions, intracellular fungal biotransformation caused a significant shift in the isomeric composition of remaining t-NP. As a result, certain t-NP constituents related to more estrogenic isomers were less efficiently degraded than others. In contrast to bacterial degradation via ipso-hydroxylation, the substitution pattern of the quaternary α-carbon of t-NP isomers does not seem to be very important for intracellular transformation in C. aquatica. As-yet-unknown intracellular enzymes are obviously induced by nonylphenols. Mass spectral data of the metabolites resulting from the intracellular oxidation of t-NP, NP112, and 4-(1-ethyl-1,3-dimethylpentyl)phenol indicate nonyl chain hydroxylation, further oxidation into keto or aldehyde compounds, and the subsequent formation of carboxylic acid derivatives. Further metabolites suggest nonyl chain desaturation and methylation of carboxylic acids. The phenolic moieties of the nonylphenols remained unchanged.Nonylphenol ethoxylates (NPEOs) represent a major group of industrial nonionic surfactants. Technical nonylphenol (t-NP), used for the production of NPEOs, is synthesized by Friedel-Crafts alkylation of phenol with a mixture of differently branched nonenes. It therefore comprises a great variety of mainly para-substituted isomers, with variously branched nonyl chains. About 50 to 80 t-NP isomers were estimated to occur in environmentally relevant matrices (19). The incomplete bioconversion of NPEOs in wastewater treatment plants results in the formation of the less biodegradable t-NP and is considered a major source of this contaminant in the aquatic environment (57). The recalcitrance of t-NP to biodegradation is partly due to the presence in more than 85% of the t-NP isomers of a quaternary α-carbon in the branched nonyl chain. Such structural characteristics are considered to limit biological nonyl chain oxidation (11, 53, 55). Nonylphenols are known to disrupt normal endocrine functions in vertebrates (57). Certain isomers contained in t-NP have been reported to possess a considerably higher estrogenic activity than the t-NP mixture (15). Due to increasing concerns with respect to their largely unknown environmental fate and potentially adverse environmental and human health effects, nonylphenols have been listed as priority hazardous substances in the EU water framework directive.In light of the concerns above, microbial reactions with the potential to reduce nonylphenol concentrations in the environment but also offering new possibilities for applications such as effluent treatment have received increasing attention (11). Among environmental microorganisms, both aquatic and terrestrial fungi, as well as bacteria, have been shown to degrade t-NP (11). Fungal attack on nonylphenols differs from bacterial nonylphenol degradation. In the case of intracellular nonylphenol biotransformation reactions catalyzed by fungi, only metabolites modified in the alkyl chain have been described (23, 52). Metabolites indicative of oxidation of the phenolic ring have not been described to date. Bacterial degradation pathways have only been documented in the genera Sphingomonas and Sphingobium. Bacterial mineralization of the aromatic moiety of t-NP isomers to CO2 and H2O is initiated via ring hydroxylation at the ipso (C-4) position of the phenolic ring, and nonanols are produced from the nonyl chains (10, 11, 15, 16). Bacteria have been shown to utilize branched-chain nonylphenols as growth substrates (11, 12, 17, 43). In contrast, only one report describes the growth of a fungus, the yeast Candida aquaetextoris, on nonylphenol (the isomer 4-n-NP containing a linear nonyl chain) (52). With respect to fungal attack on t-NP, cometabolism seems to be the dominating process (11). Recent literature data indicate that certain t-NP isomers with an estrogenic potency higher than those of the original t-NP mixture can be enriched in remaining t-NP. This results from the selective removal of individual isomers upon bacterial ipso-substitution degradation mechanisms (15). However, the effects of fungal biotransformation reactions on the isomeric profile of t-NP have not yet been quantified.Laccases are extracellular multicopper oxidases. These have most frequently been described in white-rot basidiomycetes, which unspecifically oxidize via one-electron abstraction certain lignin constituents, as well many xenobiotic compounds. Thereby, organic radicals are generated as the primary oxidation products (3). Among the several groups of fungi found in aquatic environments, aquatic hyphomycetes (AQH) are a phylogenetically diverse group of mitosporic fungi specifically adapted to their exclusively aquatic lifestyle. AQH have been shown to metabolize several organic environmental pollutants, including t-NP (23), polycyclic musk fragrances (31), pesticide metabolites (2), and synthetic dyes (22). Therefore, with respect to the fungal attack on organic pollutants found in aquatic ecosystems, AQH are of special importance. Laccase production by strictly aquatic fungi such as AQH has already been demonstrated and discussed in the context of lignocellulose decay in aquatic ecosystems (1). A role of this enzyme in the AQH-catalyzed breakdown of aquatic environmental pollutants has been recently suggested. Here, laccase isolated from the AQH Clavariopsis aquatica was shown to act on nonylphenol (23) and polycyclic musk fragrances (31). Laccase has also been implicated in nonylphenol degradation by white-rot fungi (44, 45). Isolated extracellular laccases from several aquatic and terrestrial fungi were shown to catalyze the formation of oligomeric coupling products from nonylphenols via organic radical intermediates (6, 11). However, the effects of laccase reactions on the isomeric patterns of t-NP have not been assessed to date.The aim of the present study was to quantify the influence of extracellular laccase catalysis and intracellular biotransformation on nonylphenol removal rates and on the isomeric composition of t-NP. For this, C. aquatica was used as a model organism. The derived data were compared to effects of bacteria on nonylphenol isomers reported by other authors (15), and environmental and biotechnological implications of fungal t-NP biotransformation were deduced. At the same time we addressed metabolite formation from t-NP and the two major t-NP isomers 4-(1-ethyl-1,3-dimethylpentyl)phenol (NP111) and 4-(1-ethyl-1,4-dimethylpentyl)phenol (NP112) (Fig. (Fig.1).1). This was done to substantiate the apparent differences between fungi and bacteria in the intracellular oxidation of t-NP (11, 15).Open in a separate windowFIG. 1.Chemical structures of the nonylphenol isomers NP111 and NP112.  相似文献   

5.
The chemical syntheses of a number of 14α-alkyl substituted 15-oxygenated sterols have been pursued to permit evaluation of their activity in the inhibition of the biosynthesis of cholesterol and other biological effects. Described herein are the first chemical syntheses of 14α-ethyl-5α-cholest-7-en-3β-ol-15-one, bis-3β,15α-acetoxy-14α-ethyl-5α-cholest-7-ene, 3β-acetoxy-14α-ethyl-5α-cholest-7-en-15β-ol, 14α-ethyl-5α-cholest-7-en-3β,15β-diol, 14α-ethyl-5α-cholest-7-en-3β,15α-diol, 3β-hexadecanoyloxy-14α-ethyl-5α-cholest-7-en-15α-ol, 3β-hexadecanoyloxy-14α-ethyl-5α-cholest-7-en-15β-ol, bis-3β,15α-hexadecanoyloxy-14α-ethyl-5α-cholest-7-ene, 3β-hexadecanoyloxy-14α-ethyl-5α-cholest-7-en-15-one, 3α-benzoyloxy-14α-ethyl-5α-cholest-7-en-15-one, 14α-ethyl-5α-cholest-7-en-3α-ol-15-one, 14α-ethyl-5α-cholest-7-en-15-on-3β-yl pyridinium sulfate, 14α-ethyl-5α-cholest-7-en-15-on-3β-yl potassium sulfate (monohydrate), 14α-ethyl-5α-cholest-7-en-15-on-3α-yl pyridinium sulfate, 14α-ethyl-5α-cholest-7-en-15-on-3α-yl potassium sulfate (monohydrate), 3β-ethoxy-14α-ethyl-5α-cholest-7-en-15-one, 3β-acetoxy-14α-n-propyl-5α-cholest-7-en-15-one, 14α-n-propyl-5α-cholest-7-en-3β-ol-15-one, bis-3β, 15α-acetoxy-14α-n-propyl-5α-cholest-7-ene, 3β-acetoxy-14α-n-propyl-5α-cholest-7-en-15β-ol, 14α-n-propyl-5α-cholest-7-en-3β, 15α-diol, 14α-n-propyl-5α-cholest-7-en-3β, 15β-diol, 14α-n-butyl-5α-cholest-7-en-3β-ol-15-one, 3β-acetoxy-14-α-n-butyl-5α-cholest-7-en-15-one, bis-3β,15α-acetoxy-14α-n-butyl-5α-cholest-7-ene, 3β-acetoxy-14α-n-butyl-5α-cholest-7-en-15β-ol, 14α-n-butyl-5β-cholest-7-en-3β, 15β-diol, and 14α-n-butyl-5α-cholest-7-en-3β, 15α-diol.  相似文献   

6.
This study evaluates the potential of bio-augmentation to improve the degradation of recalcitrant nonylphenol during the wastewater treatment in membrane bioreactors (MBR). One MBR containing activated sludge was bio-augmented using multistep inoculation with freeze dried Sphingomonas sp. strain TTNP3, whereas a second control reactor contained activated sludge solely. The 14C-labeled-nonylphenol isomer (4-[1-ethyl-1,3-dimethylpentyl]phenol) was applied as a single pulse. Bio-augmentation resulted in an immediate increase of dissolved radioactivity in the effluent in comparison to the control reactor (13% and 2% of initially applied radioactivity after 1 day, respectively). After 5 days of operation, the retentate of the bio-augmented reactor contained only 7% of the initial radioactivity in contrast to 50% in the control reactor. The radioactivity associated to the mixed liquor suspended solids, i.e., the suspension of biomass and other solids on the retentate side of the membrane, was mainly found as non-extractable residues that were increasingly formed during prolonged reactor operation, especially for the control MBR. HPLC-LSC and GC-MSn analyses revealed that the bio-augmented reactor produced more polar hydroquinone as main degradation intermediate, whereas the control reactor effluent contained a complex mixture of apolar compounds with shortened oxidized alkyl chains. Thus, the apparent differences in the behavior of nonylphenol between the reactors were due to the catabolism of nonylphenol conferred by bio-augmentation with Sphingomonas sp. strain TTNP3. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

7.
Nonylphenol (NP) is an endocrine disruptor present as a pollutant in river sediment. Biodegradation of NP can reduce its toxicological risk. As sediments are mainly anaerobic, degradation of linear (4-n-NP) and branched nonylphenol (tNP) was studied under methanogenic, sulphate reducing and denitrifying conditions in NP polluted river sediment. Anaerobic bioconversion was observed only for linear NP under denitrifying conditions. The microbial population involved herein was further studied by enrichment and molecular characterization. The largest change in diversity was observed between the enrichments of the third and fourth generation, and further enrichment did not affect the diversity. This implies that different microorganisms are involved in the degradation of 4-n-NP in the sediment. The major degrading bacteria were most closely related to denitrifying hexadecane degraders and linear alkyl benzene sulphonate (LAS) degraders. The molecular structures of alkanes and LAS are similar to the linear chain of 4-n-NP, this might indicate that the biodegradation of linear NP under denitrifying conditions starts at the nonyl chain. Initiation of anaerobic NP degradation was further tested using phenol as a structure analogue. Phenol was chosen instead of an aliphatic analogue, because phenol is the common structure present in all NP isomers while the structure of the aliphatic chain differs per isomer. Phenol was degraded in all cases, but did not affect the linear NP degradation under denitrifying conditions and did not initiate the degradation of tNP and linear NP under the other tested conditions.  相似文献   

8.
We isolated three Sphingobium fuliginis strains from Phragmites australis rhizosphere sediment that were capable of utilizing 4-tert-butylphenol as a sole carbon and energy source. These strains are the first 4-tert-butylphenol-utilizing bacteria. The strain designated TIK-1 completely degraded 1.0 mM 4-tert-butylphenol in basal salts medium within 12 h, with concomitant cell growth. We identified 4-tert-butylcatechol and 3,3-dimethyl-2-butanone as internal metabolites by gas chromatography-mass spectrometry. When 3-fluorocatechol was used as an inactivator of meta-cleavage enzymes, strain TIK-1 could not degrade 4-tert-butylcatechol and 3,3-dimethyl-2-butanone was not detected. We concluded that metabolism of 4-tert-butylphenol by strain TIK-1 is initiated by hydroxylation to 4-tert-butylcatechol, followed by a meta-cleavage pathway. Growth experiments with 20 other alkylphenols showed that 4-isopropylphenol, 4-sec-butylphenol, and 4-tert-pentylphenol, which have alkyl side chains of three to five carbon atoms with α-quaternary or α-tertiary carbons, supported cell growth but that 4-n-alkylphenols, 4-tert-octylphenol, technical nonylphenol, 2-alkylphenols, and 3-alkylphenols did not. The rate of growth on 4-tert-butylphenol was much higher than that of growth on the other alkylphenols. Degradation experiments with various alkylphenols showed that strain TIK-1 cells grown on 4-tert-butylphenol could degrade 4-alkylphenols with variously sized and branched side chains (ethyl, n-propyl, isopropyl, n-butyl, sec-butyl, tert-butyl, n-pentyl, tert-pentyl, n-hexyl, n-heptyl, n-octyl, tert-octyl, n-nonyl, and branched nonyl) via a meta-cleavage pathway but not 2- or 3-alkylphenols. Along with the degradation of these alkylphenols, we detected methyl alkyl ketones that retained the structure of the original alkyl side chains. Strain TIK-1 may be useful in the bioremediation of environments polluted by 4-tert-butylphenol and various other 4-alkylphenols.4-tert-Butylphenol is an alkylphenol with a tertiary branched side chain of four carbon atoms at the para position of phenol. It is an industrially important chemical and is abundantly and widely used for the production of phenolic, polycarbonate, and epoxy resins. Production of 4-tert-butylphenol in the European Union in 2001 was 25,251 tons (t) (9). In Japan, according to the National Institute of Technology and Evaluation (http://www.safe.nite.go.jp/english/sougou/view/ComprehensiveInfoDisplay_en.faces), production of 4-tert-butylphenol amounted to 27,761 t in 2007. 4-tert-Butylphenol is widely distributed in aquatic environments, including river waters (20), seawaters (17), river sediments (17), marine sediments (23), and effluent samples from sewage treatment plants and wastewater treatment plants (22). Furthermore, 4-tert-butylphenol interacts with estrogen receptors (29, 30, 34, 35, 39) and exhibits other toxic effects on aquatic organisms and humans (4, 15, 16, 25, 26, 42, 43). Therefore, it is necessary to study the biodegradation of 4-tert-butylphenol to understand its fate in the aquatic environment, to establish technologies to treat the waters polluted by it, and to remove it from contaminated environments.Studies of the biodegradation of alkylphenols have focused mainly on branched 4-nonylphenol. Several strains of sphingomonad bacteria, including Sphingomonas sp. strain TTNP3 (38), Sphingobium xenophagum Bayram (11), and Sphingomonas cloacae S-3T (10), have recently been isolated from activated sludge. These strains can degrade branched 4-nonylphenol and utilize it as a sole carbon source. In addition, several Pseudomonas strains that can degrade medium-chain 4-n-alkylphenols (e.g., 4-n-butylphenol) and utilize them as sole carbon sources have been isolated from activated sludge or contaminated soil; they include Pseudomonas veronii INA06 (1), Pseudomonas sp. strain KL28 (21), and Pseudomonas putida MT4 (36). Biodegradation of branched 4-nonylphenol and 4-n-butylphenol has been well studied, but little is known about the biodegradation of 4-tert-butylphenol, although this compound has a structure similar to those of branched 4-nonylphenol and 4-n-butylphenol. There is only one report on the biotransformation of 4-tert-butylphenol—by resting cells of S. xenophagum strain Bayram grown on technical nonylphenol—but this strain cannot grow on 4-tert-butylphenol (11, 14). To our knowledge, there are no reports of bacteria that utilize 4-tert-butylphenol as the sole carbon source, and the biochemical pathway of 4-tert-butylphenol utilization has not been described.Here we characterize three Sphingobium fuliginis strains—TIK-1, TIK-2, and TIK-3—isolated from rhizosphere sediment of the reed Phragmites australis. These strains could use 4-tert-butylphenol as a sole carbon source. On the basis of additional tests of strain TIK-1, we propose that it degrades 4-tert-butylphenol through 4-tert-butylcatechol along a meta-cleavage pathway. We also show that strain TIK-1 cells grown on 4-tert-butylphenol can degrade a wide range of 4-alkylphenols via a meta-cleavage pathway.  相似文献   

9.
《Phytochemistry》1987,26(5):1513-1515
Five new lignans, machilin A[(2S,3R)-2,3-dimethyl-1,4-dipiperonyl-butane], machilin B [(2S,3S)-2,3-dihydro-7-methoxy-3-methyl-2-piperon threo-2-(2-methoxy-4-trans-propenylphenoxy)-1-(4-hydroxy-3-methoxyphenyl)propan-1-ol], machilin E (erythro-1-acetoxy-2-(2-methoxy-4-trans-(3-hydroxy-1-propenyl)phenoxy]-l-piperonylpropane) were isolated from the bark of Machilus thunbergii and their structures were characterized.  相似文献   

10.
The biotransformation of [14C]4-n-nonylphenol (5 mg l–1; 10 mg l–1) by Agrostemma githago cell suspensions was studied using a batch two-liquid-phase system (medium/n-hexadecane 200:1, v/v). The highly lipophilic 4-n-nonylphenol was applied via n-hexadecane phase. After 7 d of incubation, more than 85% of applied 4-n-nonylphenol was absorbed by the cells, and 40% was transformed to 10 side-chain monohydroxylated metabolites (two with additional double bond at side-chain). The primary metabolites were analyzed by GC-EIMS. In the cells, the monohydroxylated products and residual 4-n-nonylphenol were present as glycosides. The method proved to be suitable for the production of primary metabolites of 4-n-nonylphenol on a larger scale for identification purposes and for metabolic profiling of the compound.  相似文献   

11.
Recently we showed that degradation of several nonylphenol isomers with alpha-quaternary carbon atoms is initiated by ipso-hydroxylation in Sphingobium xenophagum Bayram (F. L. P. Gabriel, A. Heidlberger, D. Rentsch, W. Giger, K. Guenther, and H.-P. E. Kohler, J. Biol. Chem. 280:15526-15533, 2005). Here, we demonstrate with 18O-labeling experiments that the ipso-hydroxy group was derived from molecular oxygen and that, in the major pathway for cleavage of the alkyl moiety, the resulting nonanol metabolite contained an oxygen atom originating from water and not from the ipso-hydroxy group, as was previously assumed. Our results clearly show that the alkyl cation derived from the alpha-quaternary nonylphenol 4-(1-ethyl-1,4-dimethyl-pentyl)-phenol through ipso-hydroxylation and subsequent dissociation of the 4-alkyl-4-hydroxy-cyclohexadienone intermediate preferentially combines with a molecule of water to yield the corresponding alcohol and hydroquinone. However, the metabolism of certain alpha,alpha-dimethyl-substituted nonylphenols appears to also involve a reaction of the cation with the ipso-hydroxy group to form the corresponding 4-alkoxyphenols. Growth, oxygen uptake, and 18O-labeling experiments clearly indicate that strain Bayram metabolized 4-t-butoxyphenol by ipso-hydroxylation to a hemiketal followed by spontaneous dissociation to the corresponding alcohol and p-quinone. Hydroquinone effected high oxygen uptake in assays with induced resting cells as well as in assays with cell extracts. This further corroborates the role of hydroquinone as the ring cleavage intermediate during degradation of 4-nonylphenols and 4-alkoxyphenols.  相似文献   

12.
l-threo-2,3-Hexodiulosono-1,4-lactone 2-(3-chlorophenylhydrazone) and 4- (2-acetoxyethylidene)-4-hydroxy-2,3-dioxobutano-1,4-lactone 2-(3-chlorophenylhydrazone) were prepared. The two geometric isomers of the corresponding bis(hydrazone) underwent an intramolecular rearrangement to 1-(3-chlorophenyl)- 3-(l-threo-glycerol-1-yl)-4,5-pyrazoledione 4-(3-chlorophenylhydrazone), which gave a tri-O-acetyl derivative upon acetylation and the anticipated formyl derivative upon periodate oxidation. Oxidation of the bis(hydrazone) with cupric chloride afforded the bicyclic compound 3,6-anhydro-3-C-(3-chlorophenylazo)-l- xylo-2-hexulosono-1,4-lactone 2-(3-chlorophenylhydrazone), whose acetylation afforded the mono-O-acetyl derivative.  相似文献   

13.
The use of the microorganism Sporotrichum sulfurescens (ATCC 7159) to oxygenate organic molecules has been extended to several dialkylbenzenes. Oxygenation of 1,4-di-t-butylbenzene (1) gave 4-t-butyl(1-hydroxy-2-methyl)isopropylbenzene (2) and 1,4-di-(1-hydroxy-2-methyl)isopropylbenzene (3); of 1,4-diisopropylbenzene (4) gave (R,R)-1,4-di-(1-hydroxy)isopropylbenzene (5); of 1,3-diisopropylbenzene (6) gave 1,3-di-(2-hydroxy)isopropylbenzene (7), 3-(1-hydroxy)isopropyl-(2-hydroxy)isopropylbenzene (8), and 1,3-di-(1-hydroxy)isopropylbenzene (9); and of p-isobutylisopropylbenzene (20) gave 1-(p-2-hydroxyisopropylphenyl)-2-methylpropan-2-ol (15) and 1-(p-1-hydroxyisopropylphenyl)-2-methylpropan-2-ol (16). Monohydroxydialkylbenzenes also served as useful substrates in this reaction as suggested by the fact that 2 is an intermediate in the formation of 3 from 1. Oxygenation of 1-(p-isopropylphenyl)-2-methylpropan-2-ol (14), conveniently prepared from 2-(p-isopropylphenyl)propene (12) via oxygenative isomerization with thallium trinitrate to 13 followed by addition of methyl magnesium bromide, gave 15 and 16. Oxygenation of 2-(p-isobutylphenyl)propan-2-ol (18) gave 15, 2-(p-isobutylphenyl)-propan-1,2-diol (21), and 1-(p-2-hydroxyisopropylphenyl)-2-methylpropan-3-ol (22). Compound 16, obtained from substrate 14, was converted to (2R)-2-[4-(2-hydroxy-2-methylpropyl)phenyl]propionic acid (11), the enantiomer of a metabolite of the antiinflammatory agent, 2-(4-i-butyl)phenylpropionic acid (10).  相似文献   

14.
The degradation of bisphenol A and nonylphenol involves the unusual rearrangement of stable carbon-carbon bonds. Some nonylphenol isomers and bisphenol A possess a quaternary α-carbon atom as a common structural feature. The degradation of nonylphenol in Sphingomonas sp. strain TTNP3 occurs via a type II ipso substitution with the presence of a quaternary α-carbon as a prerequisite. We report here a new degradation pathway of bisphenol A. Consequent to the hydroxylation at position C-4, according to a type II ipso substitution mechanism, the C-C bond between the phenolic moiety and the isopropyl group of bisphenol A is broken. Besides the formation of hydroquinone and 4-(2-hydroxypropan-2-yl)phenol as the main metabolites, further compounds resulting from molecular rearrangements consistent with a carbocationic intermediate were identified. Assays with resting cells or cell extracts of Sphingomonas sp. strain TTNP3 under an 18O2 atmosphere were performed. One atom of 18O2 was present in hydroquinone, resulting from the monooxygenation of bisphenol A and nonylphenol. The monooxygenase activity was dependent on both NADPH and flavin adenine dinucleotide. Various cytochrome P450 inhibitors had identical inhibition effects on the conversion of both xenobiotics. Using a mutant of Sphingomonas sp. strain TTNP3, which is defective for growth on nonylphenol, we demonstrated that the reaction is catalyzed by the same enzymatic system. In conclusion, the degradation of bisphenol A and nonylphenol is initiated by the same monooxygenase, which may also lead to ipso substitution in other xenobiotics containing phenol with a quaternary α-carbon.  相似文献   

15.
Both isomers of 1,2:4,6-di-O-benzylidene-α-d-glucopyranose (and their 3-O-acetyl and 3-O-benzyl derivatives) have been prepared and their 1H- and 13C-n.m.r. spectra assigned. The mode of hydrogenolysis of the dioxolane ring in these isomers by the LiAlH4AlCl3 reagent is determined by the configuration at the acetal carbon and is independent of the electronic character of the two oxygen atoms.  相似文献   

16.
4-Ethylphenol methylenehydroxylase from Pseudomonas putida JD1 acts by dehydrogenation of its substrate to give a quinone methide, which is then hydrated to an alcohol. It was shown to be active with a range of 4-alkylphenols as substrates. 4-n-Propylphenol, 4-n-butylphenol, chavicol, and 4-hydroxydiphenylmethane were hydroxylated on the methylene group next to the benzene ring and produced the corresponding chiral alcohol as the major product. The alcohols 1-(4′-hydroxyphenyl)propanol and 1-(4′-hydroxyphenyl)-2-propen-1-ol, produced by the biotransformation of 4-n-propylphenol and chavicol, respectively, were shown to be R(+) enantiomers. 5-Indanol, 6-hydroxytetralin, 4-isopropylphenol, and cyclohexylphenol, with cyclic or branched alkyl groups, gave the corresponding vinyl compounds as their major products.  相似文献   

17.
《Carbohydrate research》1998,310(4):229-238
Eight positional isomers of 61,6m-di-O-α-d-mannopyranosyl-cyclomaltooctaose (γCD) (m=2–5) and 6-O-α-(n-O-α-d-mannopyranosyl)-d-mannopyranosyl-γCD (n=2, 3, 4, and 6) in a mixture of products from γCD and d-mannose by condensation reaction of α-mannosidase from jack bean were isolated by HPLC. The structures of four isomers of 6-O-α-(n-O-α-d-mannopyranosyl)-d-mannopyranosyl-γCD were elucidated by NMR spectroscopy. On the other hand, four positional isomers of 61,6m-di-O-α-d-mannopyranosyl-γCD were determined by LC–MS analysis of degree of polymerization of the branched oligosaccharides produced by enzymatic degradation with bacterial saccharifying α-amylase (BSA), and combination of BSA and glucoamylase. Similarly cyclomaltodextrin glucanotransferase also digested these isomers.  相似文献   

18.
In the seeds of Cucurbita pepo three closely related 24-ethyl-Δ7-steryl glucosides were identified by hydrolytic studies and spectral analysis as spinasteryl-β-D-glucopyranoside, the new 3-O-(β-D-glucopyranosyl)-24β-ethyl-5α-cholesta-7,25(27)-dien-3β-ol and the corresponding Δ22E,25,(27)-trienol. Except for its occurrence in cucumber seeds the latter is so far unknown as a natural product.  相似文献   

19.
《Phytochemistry》1983,22(10):2259-2262
Four new spirostanosides were isolated from the methanol extract of Asparagus sprengeri roots and characterized as 3-O-[β-D-xylo-(1 → 4)-β-D-gluco]-(25R)-spirost-5-en-3β-ol; 3-O-[α-L-rhamno-(1 → 6)-β-D-gluco]-(25R)-spirost-5-en-3β-ol; 3-O-{[α-L-rhamno-(1 → 2)] [β-D-xylo-(1 → 4)]-β-D-gluco}-(25R)-spirost-5-en-3β-ol and 3- O-{[α-L-rhamno-(1 → 2)] [α-L-rhamno-(1 → 6)]-β-D-gluco}-(25R)-spirost-5-en-3β-ol respectively.  相似文献   

20.
Three spirostanol and two furostanol glycosides were isolated from a methanol extract of the roots of Asparagus curillus and characterized as 3-O-[α-l-arabinopyranosyl (1→4)- β-d-glucopyranosyl]-(25S)-5β-spirostan-3β-ol, 3-O-[{α-l-rhamnopyranosyl (1→2)} {α-l-arabinopyranosyl (1→4)}-β-d-glucopyranosyl]-(25S)-5β-spirostan- 3β-ol, 3-O-[{β-d-glucopyranosyl (1→2)} {α-l-arabinopyranosyl (1→4)}-β- d-glucopyranosyl]-(25S)-5β-spirostan-3β-ol, 3-O-[{β-d-glucopyranosyl (1→2)} {α-l-arabinopyranosyl (1→4)}-β-d-glucopyranosyl]-26-O-[β-d-glucopyranosyl]- 22α-methoxy-(25S)-5β-furostan-3β, 26-diol and 3-O-[{β-d-glucopyranosyl (1→2)} {α-l-arabinopyranosyl (1→4)}-β-d-glucopyranosyl]-26-O-[β-d-glucopyranosyl]- (25S)-5β-furostan-3β, 22α, 26-triol respectively.  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