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 46 毫秒
1.
Brevibacillus laterosporus MTCC 2298 showed 87% decolorization of Golden Yellow HER within 48 h under static condition at the concentration 50 mg l?1; however no significant change in the decolorization performance was observed under shaking condition. Decolorization performance was maximum (74%) at the pH 7.0 and 30 °C. TLC and HPLC analysis confirmed the biodegradation of Golden Yellow HER. Biodegradation pathway was proposed using GC–MS and FTIR spectral analysis. Mainly elected metabolites are the 2,5-Dichloro-4 (3-hydrazino-2-hydroxy cyclopentylamino-) dibenzene-sulfonic acid (peak 1, m/z = 526), 4-(3-hydrazino-2-hydroxy cyclopentylamino)-benzene-sulfonic acid (peak 2, m/z = 455), 4-(3-amino-2-hydroxy-cyclopentylamino)-benzene-sulfonic acid and 5-amino-cyclohex-3-ene-sulfonic acid (peak 3, m/z = 183). Phytotoxicity results suggested that degradation products of Golden Yellow HER are non-toxic to the common crops such as Sorghum vulgare and Phaseolus mungo. Also, degradation products are non-toxic to B. laterosporus as well as ecologically important bacteria like Pseudomonas aeruginosa and Azotobacter vinelandii.  相似文献   

2.
Laccase-mediated grafting of functional molecules presents an eco-friendly approach to functionalize lignocellulose materials. In this study functional molecules in the form of reactive phenolic amines, hydrophobicity enhancing fluorophenols and selected wood preservatives, were for the first time successfully coupled onto the lignin model compound dibenzodioxocin (Db) as demonstrated by HPLC-MS analysis. A 1:1-coupling was demonstrated for various combinations including Db and tyramine (m/z 620.5), Db and 3-O-methyldopamine (m/z 650.5), Db and 4-hydroxy-3-methoxybenzylamine (m/z 636.5), Db and 4-fluoro-2-methylphenol (m/z 609.5), and Db and 2-phenylphenol (m/z 653.5). Fungal laccases from Trametes hirsuta and T. villosa were more efficient in mediating the coupling of tyramine to dibenzodioxocin and beech (Fagus sylvatica) wood than a Bacillus sp. laccase with lower redox potential. This work presents for the first time a model for functionalizing of lignocellulose using the lignin model dibenzodioxocin.  相似文献   

3.
We examined the degradation of amaranth, a representative azo dye, by Bjerkandera adusta Dec 1. The degradation products were analyzed by high performance liquid chromatography (HPLC), visible absorbance, and electrospray ionization time-of-flight mass spectroscopy (ESI-TOF-MS). At the primary culture stage (3 days), the probable reaction intermediates were 1-aminonaphthalene-2,3,6-triol, 4-(hydroxyamino) naphthalene-1-ol, and 2-hydroxy-3-[2-(4-sulfophenyl) hydrazinyl] benzenesulfonic acid. After 10 days, the reaction products detected were 4-nitrophenol, phenol, 2-hydroxy-3-nitrobenzenesulfonic acid, 4-nitrobenzene sulfonic acid, and 3,4′-disulfonyl azo benzene, suggesting that no aromatic amines were created. Manganese-dependent peroxidase activity increased sharply after 3 days culture. Based on these results, we herein propose, for the first time, a degradation pathway for amaranth. Our results suggest that Dec 1 degrades amaranth via the combined activities of peroxidase and hydrolase and reductase action.  相似文献   

4.
Microcystin-LR (MC-LR) and microcystin-RR (MC-RR) are the two most common microcystins (MCs) present in fresh water posing a direct threat to public health because of their hepatotoxicity. A novel MC-degrading bacterium designated MC-LTH1 capable of degrading MC-LR and -RR was isolated, and the degradation rates and mechanisms of MC-LR and -RR for this bacterium were investigated. The bacterium was identified as Bordetella sp. and shown to possess a homologous mlrA gene responsible for degrading MCs. To the best of our knowledge, this is the first report of mlrA gene detection in Bordetella species. MC-LR and -RR were completely degraded separately at rates of 0.31 mg/(L h) and 0.17 mg/(L h). However, the degradation rates of MC-LR and -RR decreased surprisingly to 0.27 mg/(L h) and 0.12 mg/(L h), respectively, when both of them were simultaneously present. Degradation products were identified by high performance liquid chromatography coupled with time-of-flight mass spectrometry. Adda (m/z 332.2215, C20H29NO3) commonly known as a final product of MC degradation by isolated bacteria was detected as an intermediate in this study. Linearized MC-LR (m/z 1013.5638, C49H76N10O13), linearized MC-RR (m/z 1056.4970, C49H77N13O13), and tetrapeptide (m/z 615.3394, C32H46N4O8) were also detected as intermediates. These results indicate that the bacterial strain MC-LTH1 is quite efficient for the detoxification of MC-LR and MC-RR, and possesses significant bioremediation potential.  相似文献   

5.
The white-rot basidiomycete Phanerochaete chrysosporium metabolized 3-(4′-ethoxy-3′-methoxyphenyl)-2-(4″-methoxyphenyl)propionic acid (V) in low-nitrogen, stationary cultures, conditions under which ligninolytic activity is expressed. The ability of several fungal mutant strains to degrade V reflected their ability to degrade [14C]lignin to 14CO2. 1-(4′-Ethoxy-3′-methoxyphenyl)-2-(4″-methoxyphenyl)-2- hydroxyethane (VII), anisyl alcohol, and 4-ethoxy-3-methoxybenzyl alcohol were isolated as metabolic products, indicating an initial oxidative decarboxylation of V, followed by α, β cleavage of the intermediate (VII). Exogenously added VII was rapidly converted to anisyl alcohol and 4-ethoxy-3-methoxybenzyl alcohol. When the degradation of V was carried out under 18O2, 18O was incorporated into the β position of the diarylethane product (VII), indicating that the reaction is oxygenative.  相似文献   

6.
The electrospray mass spectrum (ESI-MS) of cis-[Ru(NO)Cl(bpy)2]Cl2 (bpy=2,2-bipyridine), obtained from 50% CH3OH/50% H2O as the mobile solvent, exhibited ruthenium-containing ions derived from a {[RuII(NO+)Cl(bpy)2]2+, Cl}+ ion pair (m/z=514) and [RuII(NO+)Cl(bpy)2]2+ (m/z=239.5). [RuIIICl(bpy)2]2+, from the loss of NO from the 239.5 ion, is detected at m/z=224.5. Only the m/z 514 ion pair is detected when 100% CH3OH mobile solvent is used, but the presence of even small amounts of water prompted the additional detection of the m/z 239.5 and m/z 224.5 ions under tandem MS-MS conditions. Ruthenium-chloro-containing ions appear as a characteristic collection of eight main, and four lesser, intense ions created from combinations 104Ru, 102Ru, 101Ru, 99Ru, 98Ru, 96Ru, 35Cl and 37Cl isotopes with minor contributions from 13C, etc. For convenience of discussion, only the most abundant m/z species are mentioned herein as representative of all the isotopically distributed ions.Four fragmentation channels are detectable from the m/z=514 chloride ion pair: (1) the loss of HCl (main channel; ca. 50% of fragmentation events), (2) the loss of NO (ca. 12% ), (3) the loss of bpy (minor pathway), and (4) the loss of Cl atom (ca. 38% ).Loss of NO from ion m/z 514 yields ion m/z 484, which is the precursor of ions m/z 448 (by loss of HCl), m/z 328 (by loss of bpy) and m/z 292 (by loss of HCl and bpy). Loss of HCl from ion m/z 514 generates ion m/z 478, [RuII(NO+)Cl(bpyH)(bpy-H)]+, deprotonated at the ortho C-H of one bpy ligand. In MS-MS experiments, the m/z 478 ion was established to undergo loss of NO, producing ion m/z 448, rejoining further fragmentation process for ion m/z 448 at this point. Loss of neutral bipyridine from m/z 514 in low yield produces ion m/z 358, which undergoes further loss of NO to form [RuCl2(bpy)]+ ion (m/z=328). MS-MS “neutral loss of 30” spectra confirmed the NO loss events as part of the fragmentation sequence for all four pathways.A fourth species of m/z=479 from the “514” ion is obtained by an internal electron transfer from Cl of the ion pair, and loss of the resultant neutral Cl atom. The product [RuII(NO·)Cl(bpy)2]+ “479” fragment undergoes facile loss of NO to generate [RuIICl(bpy)2]+ (m/z=449). Ion m/z 449 gives rise to ions m/z 413 (loss of HCl) and m/z 257(loss of HCl and bpy). MS-MS experiments confirm the neutral loss of Cl from the m/z 514 ion, and the formation of the m/z 449 ion via m/z 479 and m/z 514 parents. This pathway was not observed in a prior study for the related complex, [Ru(NO)Cl(dpaH)(dpa)]+ (dpaH=2,2-dipyridylamine), which does not have an external Cl in an ion pair.  相似文献   

7.
The extracellular and intracellular metabolites formed upon exposure of activated sludge microorganisms to a sublethal concentration of N-ethylmaleimide were monitored by liquid chromatography with ion trap mass spectrometry. The metabolite N-ethylsuccinimido-S-glutathione (m/z 433) was converted rapidly to N-(2-oxoethyl)-2,2-(propionylamino)propanamide (m/z 187) and N-ethylmaleamic acid (m/z 144).  相似文献   

8.
A sensitive and accurate ultra-performance liquid chromatography electrospray ionization tandem mass spectrometry (UPLC–ESI-MS/MS) method was developed and validated for the simultaneous determination of vitexin-4″-O-glucoside (VGL), vitexin-2″-O-rhamnoside (VRH), rutin (RUT) and vitexin (VIT) in rat plasma after intravenous administration of hawthorn leaves flavonoids (HLF). Following protein precipitation by methanol, the analytes were separated on an ACQUITY UPLC BEH C18 column packed with 1.7 μm particles by gradient elution using a mobile phase composed of acetonitrile and water (containing 0.1% formic acid) at a flow rate of 0.20 mL/min. The analytes and diphenhydramine (internal standard, IS) were detected in the multiple reaction monitoring (MRM) mode by means of an electrospray ionization (ESI) interface (m/z 292.96 for vitexin-4″-O-glucoside, m/z 293.10 for vitexin-2″-O-rhamnoside, m/z 299.92 for rutin, m/z 310.94 for vitexin and m/z 166.96 for IS). The calibration curve was linear over the range 10–40,000 ng/mL for vitexin-4″-O-glucoside, 10–50,000 ng/mL for vitexin-2″-O-rhamnoside, 8–1000 ng/mL for rutin and 16–2000 ng/mL for vitexin. The intra- and inter-run precisions (relative standard deviation, RSD) of these analytes were all within 15% and the accuracy (the relative error, RE) ranged from −10% to 10%. The stability experiment indicated that the four analytes in rat plasma samples and plasma extracts under anticipated conditions were stable. The developed method was applied for the first time to pharmacokinetic studies of the four bioactive compounds of hawthorn leaves flavonoids following a single intravenous administration of 20 mg/kg in rats.  相似文献   

9.
Pyrolysis-gas chromatography-mass spectrometry (Py-GC–MS) and off-line thermochemolysis with tetramethylammonium hydroxide followed by GC–MS were used in the molecular characterisation of lignin in spruce wood decayed by Phanerochaete chrysosporium and Trametes versicolor. Mono-methoxyphenols were the main pyrolysis products from the undegraded lignin. Py-GC–MS provided qualitative evidence that 2-methoxy-4-(prop-2-enal)phenol and trans-2-methoxy-4-(1-hydroxy-prop-2-enyl)phenol content decreased whereas 1,2-dihydroxybenzene increased in intensity relative to other products upon fungal decay. Comparison of methylated phenols from thermochemolysis revealed that ratio of methyl 3,4-dimethoxybenzoate to 3,4-dimethoxybenzaldehyde increased from 0.69 in control spruce to 2.3 after decay by P. chrysosporium and 3.7 following growth of T. versicolor. The results indicate that white-rot fungi cleave alkyl side chains of β-O-4 linked mono-methoxyphenylpropane structures between the αβ carbon atoms to give lignin residues enriched in carboxylic acids as well as demethylating methoxy groups attached to aromatic nuclei to give dihydroxybenzene products. Py-GC–MS and thermochemolysis are complementary methods for tracking demethylation of aromatic nuclei and oxidation of alkyl side chains caused by white-rot fungi.  相似文献   

10.
Imidacloprid (IMI), a potent insecticide, belongs to the neonicotinoid family and is of great concern due to the fact that its persistence in the soil is a threat to both plants and vertebrates. The present study was aimed at the isolation and characterization of a bacterial strain from paddy field soil at Punjab (India), which has a history of 9–10 years of imidacloprid contamination. Among the various isolates, a soil bacterium was selected and identified by 16S rRNA gene sequencing as Enterobacter sp. strain ATA1. It grew well in pH ranging from 6.0 to 7.0 at 37°C, and it was found to be a competent bacterium for the degradation of IMI. The presence of glucose in minimal salt medium (MMG; 0.1% w/v) as compared with any other co-substrate provokes the dissipation of IMI as a co-metabolite. Initially, incubation of IMI for 72 h in the MMG resulted in 30–40% degradation; thereafter, no significant change in its amount was found until 15 days of incubation, which explains the disappearance of any viable cells in the medium. Among the various identified metabolites, imidacloprid urea (m/z = 212) and imidacloprid guanidine (m/z = 211) were found to be the end products of IMI degradation, whereas others remained unidentified (m/z = 99 and m/z = 119).  相似文献   

11.
Accumulation of phosphatidylcholine hydroperoxide (PCOOH), a primary oxidation product of phosphatidylcholine, in blood plasma has been observed in various pathological conditions, including atherosclerosis. In this study, we investigated the use of liquid chromatography–tandem mass spectrometry (LC–MS/MS) to develop a method for accurate quantification of PCOOH (1-palmitoyl-2-hydroperoxyoctadecadienoyl-sn-glycero-3-phosphocholine, 16:0/HpODE PC), focusing on isomers such as 16:0/13-HpODE PC and 16:0/9-HpODE PC. Sodiated PCOOH ([M+Na]+, m/z 812) provided not only a known product ion (m/z 147) but also characteristic product ions (m/z 541 for 16:0/13-HpODE PC and m/z 388 for 16:0/9-HpODE PC). Thus, three multiple reaction monitorings (MRMs) could be performed. MRM (812/147) enabled determination of 16:0/HpODE PC, and MRM (812/541) and MRM (812/388) allowed specific measurement of 16:0/13-HpODE PC and 16:0/9-HpODE PC, respectively. By using this method, we could determine plasma PCOOH concentrations in healthy subjects and patients with angiographically significant stenosis. In healthy subject and patient plasma, the concentration of 16:0/HpODE PC was close to the sum of the concentrations of 16:0/13-HpODE PC and 16:0/9-HpODE PC. This finding shows that radical and/or enzymatic oxidation, rather than singlet oxygen oxidation, is recognized to cause peroxidation of PC. The newly developed LC–MS/MS method appears to be a powerful tool for developing a better understanding of in vivo lipid peroxidation and its involvement in human diseases.  相似文献   

12.
Oxidative cleavage of poly(cis-1,4-isoprene) by rubber oxygenase RoxA purified from Xanthomonas sp. was investigated in the presence of different combinations of 16O2, 18O2, H216O, and H218O. 12-Oxo-4,8-dimethyl-trideca-4,8-diene-1-al (ODTD; m/z 236) was the main cleavage product in the absence of 18O-compounds. Incorporation of one 18O atom in ODTD was found if the cleavage reaction was performed in the presence of 18O2 and H216O. Incubation of poly(cis-1,4-isoprene) (with RoxA) or of isolated unlabeled ODTD (without RoxA) with H218O in the presence of 16O2 indicated that the carbonyl oxygen atoms of ODTD significantly exchanged with oxygen atoms derived from water. The isotope exchange was avoided by simultaneous enzymatic reduction of both carbonyl functions of ODTD to the corresponding dialcohol (12-hydroxy-4,8-dimethyl-trideca-4,8-diene-1-ol (HDTD; m/z 240) during RoxA-mediated in vitro cleavage of poly(cis-1,4-isoprene). In the presence of 18O2, H216O, and alcohol dehydrogenase/NADH, incorporation of two atoms of 18O into the reduced metabolite HDTD was found (m/z 244), revealing that RoxA cleaves rubber by a dioxygenase mechanism. Based on the labeling results and the presence of two hemes in RoxA, a model of the enzymatic cleavage mechanism of poly(cis-1,4-isoprene) is proposed.  相似文献   

13.
Anaerobic degradation of the aromatic hydrocarbon ethylbenzene was studied with sulfate as the electron acceptor. Enrichment cultures prepared with marine sediment samples from different locations showed ethylbenzene-dependent reduction of sulfate to sulfide and always contained a characteristic cell type that formed gas vesicles towards the end of growth. A pure culture of this cell type, strain EbS7, was isolated from sediment from Guaymas Basin (Gulf of California). Complete mineralization of ethylbenzene coupled to sulfate reduction was demonstrated in growth experiments with strain EbS7. Sequence analysis of the 16S rRNA gene revealed a close relationship between strain EbS7 and the previously described marine sulfate-reducing strains NaphS2 and mXyS1 (similarity values, 97.6 and 96.2%, respectively), which grow anaerobically with naphthalene and m-xylene, respectively. However, strain EbS7 did not oxidize naphthalene, m-xylene, or toluene. Other compounds utilized by strain EbS7 were phenylacetate, 3-phenylpropionate, formate, n-hexanoate, lactate, and pyruvate. 1-Phenylethanol and acetophenone, the characteristic intermediates in anaerobic ethylbenzene degradation by denitrifying bacteria, neither served as growth substrates nor were detectable as metabolites by gas chromatography-mass spectrometry in ethylbenzene-grown cultures of strain EbS7. Rather, (1-phenylethyl)succinate and 4-phenylpentanoate were detected as specific metabolites in such cultures. Formation of these intermediates can be explained by a reaction sequence involving addition of the benzyl carbon atom of ethylbenzene to fumarate, carbon skeleton rearrangement of the succinate moiety (as a thioester), and loss of one carboxyl group. Such reactions are analogous to those suggested for anaerobic n-alkane degradation and thus differ from the initial reactions in anaerobic ethylbenzene degradation by denitrifying bacteria which employ dehydrogenations.  相似文献   

14.
The sulfonated azo dye Reactive Orange 16 is the commonly used representative of reactive dyes, but is hard to be degraded by some conventional treatment methods. In order to develop more efficient and more cost-effective treatment methods for degrading this recalcitrant dye, the capability of the white-rot fungus Ganoderma sp. En3 isolated by our laboratory to decolorize and detoxify Reactive Orange 16 was investigated in this study. Ganoderma sp. En3 had a strong ability to decolorize high concentrations of Reactive Orange 16 and simulated textile wastewater containing Reactive Orange 16 in submerged cultures. Decolorization of Reactive Orange 16 and its simulated dye effluents by this fungus resulted in the decrease of phytotoxicity. Ganoderma sp. En3 had strong adaptability and tolerance to high concentrations of Reactive Orange 16. Compared with some previous research, Ganoderma sp. En3 was superior to some other fungal strains reported previously in the rate and extent of decolorizing Reactive Orange 16. It was also found that the real textile wastewater could be efficiently decolorized by Ganoderma sp. En3 in submerged cultures. The crude enzyme produced by Ganoderma sp. En3 could also efficiently decolorize Reactive Orange 16 and simulated textile wastewater under in vitro conditions.  相似文献   

15.
We performed x-ray crystallographic analyses of the 6-aminohexanoate oligomer hydrolase (NylC) from Agromyces sp. at 2.0 Å-resolution. This enzyme is a member of the N-terminal nucleophile hydrolase superfamily that is responsible for the degradation of the nylon-6 industry byproduct. We observed four identical heterodimers (27 kDa + 9 kDa), which resulted from the autoprocessing of the precursor protein (36 kDa) and which constitute the doughnut-shaped quaternary structure. The catalytic residue of NylC was identified as the N-terminal Thr-267 of the 9-kDa subunit. Furthermore, each heterodimer is folded into a single domain, generating a stacked αββα core structure. Amino acid mutations at subunit interfaces of the tetramer were observed to drastically alter the thermostability of the protein. In particular, four mutations (D122G/H130Y/D36A/E263Q) of wild-type NylC from Arthrobacter sp. (plasmid pOAD2-encoding enzyme), with a heat denaturation temperature of Tm = 52 °C, enhanced the protein thermostability by 36 °C (Tm = 88 °C), whereas a single mutation (G111S or L137A) decreased the stability by ∼10 °C. We examined the enzymatic hydrolysis of nylon-6 by the thermostable NylC mutant. Argon cluster secondary ion mass spectrometry analyses of the reaction products revealed that the major peak of nylon-6 (m/z 10,000–25,000) shifted to a smaller range, producing a new peak corresponding to m/z 1500–3000 after the enzyme treatment at 60 °C. In addition, smaller fragments in the soluble fraction were successively hydrolyzed to dimers and monomers. Based on these data, we propose that NylC should be designated as nylon hydrolase (or nylonase). Three potential uses of NylC for industrial and environmental applications are also discussed.  相似文献   

16.
G.l.c.-mass spectrometry has been used to provide information on the O-trimethylsilyl derivatives of the products of alkaline degradation of 3,4-di- and 3,4,6-tri-O-methyl-D-glucose, and 3,4,6-tri-O-methyl-D-galactose. During reaction with sodium hydroxide-sodium borohydride mixtures, reduction occurs more rapidly than β-elimination and the only detectable products were the corresponding alditols and the epimeric 3-deoxyalditols. Extended reaction with sodium hydroxide alone, followed by treatment with sodium borohydride, gives mixtures of aldonic acids including the epimeric 3-deoxy-4-O-methylaldonic acids (metasaccharinic acids), 3-deoxyaldonic acids (with loss of the 4-O-methyl substituent), and 3,4-dideoxy-aldonic acids. Possible reaction-pathways are discussed.  相似文献   

17.
Reflecting the known biological activity of isoniazid-based hydrazones, seventeen hydrazones of 4-(trifluoromethyl)benzohydrazide as their bioisosters were synthesized from various benzaldehydes and aliphatic ketones. The compounds were screened for their in vitro activity against Mycobacterium tuberculosis, nontuberculous mycobacteria (M. avium, M. kansasii), bacterial and fungal strains. The most antimicrobial potent derivatives were also investigated for their cytostatic and cytotoxic properties against three cell lines. Camphor-based molecule, 4-(trifluoromethyl)-N′-(1,7,7-trimethylbicyclo[2.2.1]heptan-2-ylidene)benzohydrazide, exhibited the highest and selective inhibition of M. tuberculosis with the minimum inhibitory concentration (MIC) of 4?µM, while N′-(4-chlorobenzylidene)-4-(trifluoromethyl)benzohydrazide was found to be superior against M. kansasii (MIC?=?16?µM). N′-(5-Chloro-2-hydroxybenzylidene)-4-(trifluoromethyl)benzohydrazide showed the lowest MIC values for gram-positive bacteria including methicillin-resistant Staphylococcus aureus as well as against two fungal strains of Candida glabrata and Trichophyton mentagrophytes within the range of ≤0.49–3.9?µM. The convenient substitution of benzylidene moiety at the position 4 or the presence of 5-chloro-2-hydroxybenzylidene scaffold concomitantly with a sufficient lipophilicity are essential for the noticeable antimicrobial activity. This 5-chlorosalicylidene derivative avoided any cytotoxicity on two mammalian cell cultures (HepG2, BMMΦ) up to the concentration of 100?µM, but it affected the growth of MonoMac6 cells.  相似文献   

18.
Microbial treatment of environmental pollutants including dyes with white rot fungi has received wide attention as a potential alternative for conventional methods in wastewater treatment. The degradation products from dyes and mechanism underlying fungal degradation of dyes is desirable to be understood. Capillary electrophoresis coupled with mass spectrometry (CE-MS) was used in this study to determine biodegradation products of 4-[(4-hydroxyphenyl)azo]-benzenesulfonic acid, sodium salt (4HABA) and Acid Orange 7 (C.I. 15510), produced by a white rot fungus, Pleurotus ostreatus. Two major degradation products, benzenesulfonic acid and 4-hydroxy-benzenesulfonic acid, from both sulfonated compounds, were identified and their kinetic profiles in biodegradation were followed by CE-MS. Another product, 1,2-naphthoquinone, from Acid Orange 7 was identified using HPLC. Formation of these products in fungal degradation is discussed.Revisions requested 8 October 2004; Revision received 12 November 2004  相似文献   

19.
Vanillic acid (4-hydroxy-3-methoxybenzoic acid) supported the anaerobic (nitrate respiration) but not the aerobic growth of Pseudomonas sp. strain PN-1. Cells grown anaerobically on vanillate oxidized vanillate, p-hydroxybenzoate, and protocatechuic acid (3,4-dihydroxybenzoic acid) with O2 or nitrate. Veratric acid (3,4-dimethoxybenzoic acid) but not isovanillic acid (3-hydroxy-4-methoxybenzoic acid) induced cells for the oxic and anoxic utilization of vanillate, and protocatechuate was detected as an intermediate of vanillate breakdown under either condition. Aerobic catabolism of protocatechuate proceeded via 4,5-meta cleavage, whereas anaerobically it was probably dehydroxylated to benzoic acid. Formaldehyde was identified as a product of aerobic demethylation, indicating a monooxygenase mechanism, but was not detected during anaerobic demethylation. The aerobic and anaerobic systems had similar but not identical substrate specificities. Both utilized m-anisic acid (3-methoxybenzoic acid) and veratrate but not o- or p-anisate and isovanillate. Syringic acid (4-hydroxy-3,5-dimethoxybenzoic acid), 3-O-methylgallic acid (3-methoxy-4,5-dihydroxybenzoic acid), and 3,5-dimethoxybenzoic acid were attacked under either condition, and formaldehyde was liberated from these substrates in the presence of O2. The anaerobic demethylating system but not the aerobic enzyme was also active upon guaiacol (2-methoxyphenol), ferulic acid (3-[4-hydroxy-3-methoxyphenyl]-2-propenoic acid), 3,4,5-trimethoxycinnamic acid (3-[3,4,5-trimethoxyphenyl]-2-propenoic acid), and 3,4,5-trimethoxybenzoic acid. The broad specificity of the anaerobic demethylation system suggests that it probably is significant in the degradation of lignoaromatic molecules in anaerobic environments.  相似文献   

20.
The purpose of this study was to develop a fungal bioremediation method that could be used for soils heavily contaminated with persistent organic compounds, such as polyaromatic hydrocarbons (PAHs). Sawmill soil, contaminated with PAHs, was mixed with composted green waste (1:1) and incubated with or without fungal inoculum. The treatments were performed at the laboratory and field scales. In the laboratory scale treatment (starting concentration 3500 mg kg−1, sum of 16 PAH) the high molecular weight PAHs were degraded significantly more in the fungal-inoculated microcosms than in the uninoculated ones. In the microcosms inoculated with Phanerochaete velutina, 96% of 4-ring PAHs and 39% of 5- and 6-ring PAHs were removed in three months. In the uninoculated microcosms, 55% of 4-ring PAHs and only 7% of 5- and 6-ring PAHs were degraded. However, during the field scale (2 t) experiment at lower starting concentration (1400 mg kg−1, sum of 16 PAH) the % degradation was similar in both the P. velutina-inoculated and the uninoculated treatments: 94% of the 16 PAHs were degraded in three months. In the field scale experiment the copy number of gram-positive bacteria PAH-ring hydroxylating dioxygenase genes was found to increase 1000 fold, indicating that bacterial PAH degradation also played an important role.  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