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 0 毫秒
1.
Neutral sterile and lyophilized extracts of fresh soil (NAFS Extract) degraded the following substrates: p-cresol: dl-3(3.4-dihydroxyphenylalanine)(dl DOPA): d(+) catechin and p-phenylcnediamine respectively (19.7); (5.3); (5.7) and (4.4) nmoles O2 mg C?1 min?1 as specific activity (sp. act.). NAFS Extract was fractionated by G100 Sephadex column chromatography into two major peaks (KD ~ 0 and ~ 1) without an increase in sp. act. DEAE DE 52 cellulose chromatography separated NAFS Extract into three fractions. The first fraction was free from humic acid, relatively homogeneous on polyacrylamide gel electrophoresis and had sp. act.: dl DOPA (17.2): d(+) catechin (8.5): p-phenylenediamine (22.9). The o- and p-diphenol oxidases which accompanied this fraction were well separated as complexes on G100 Sephadex column and were liberated by dialysis against distilled water. Following isolation, we obtained an o-diphenol oxidase on dl DOPA (23.4) and a laccase activity on p-phenylenediamine (33.8) in the free state; these activities being associated with nucleoprotein. Fractions (II) and (III) appeared to be relatively homogeneous in the form of “humic acid—enzyme complexes”. Specific activity were high in fraction (III): dl DOPA (10.8): d(+) catechin (0.7); p-phenylenediamine (5.4). The diphenol oxidase activity extracted from soil (NAFS Extract) was treated by salmine and SP Sephadex C25 to remove humic matter. The EFS Extract obtained had the following sp. act.: dl DOPA (14.0); p-phenylenediamine (6.3). This EFS Extract was separated into three fractions by means of G100 Sephadex column chromatography. The Kn were (I) ~0; (II) ~ 0.52; (III) ~ 1.3 respectively. The first fraction showed an increase of sp. act. only with p-phenylenediamine (9.1) and in the following two fractions the sp. act. were not augmented. The first fraction was further fractionated by means of DEAE cellulose chromatography into four fractions: the-first and the second had no activity on dl DOPA and p-phenylenediamine. The third was an o-diphenol oxidase on dl DOPA (11.0) with traces of laccase: p-phenylenediamine (0.7). The fourth was pure laccase: p-phenylenediamine (18.1). These results suggested that electrostatic, covalent and van der Waals forces contributed to the formation of humic acid enzymes complexes, associated in the tetramer to monomer forms of diphenol oxidases.  相似文献   

2.
Study of diphenol oxidases extracted from beech litter. Lyophilized neutral sterile extract from the fermentation (F) layer of beech litter (NALF Extract) exhibited the chemical characteristics of humic acids. It possessed diphenol oxidase activities. The specific activities (sp. act.), given in parentheses, are expressed in nmoles O2 absorbed mg C?1 min?1: p-crcsol (19.5); catechol (0.6); dl-3(3,4-dihydroxyphenyl)alanine (5.7); d(+) catechine (4.8) and p-phenylenediamine (7.1). The NALF Extract was polydispersed by G100 Sephadex column chromatography. The firsi peak kd ~ 0.05 (fraction I), the intermediate band (fractions II + III) and the second peak kd ~ 1.02 and 1.38 (fractions IV + V). Diphenol oxidases were localized in fractions I, II and III. Electrophoretic studies have shown that the fractions I, II and III are heterogeneous. Chromatography on DEAE cellulose of fraction I permitted the separation of 30 per cent of the laccase activity in a form which is free from humic material.  相似文献   

3.
The transformation of naturally occurring phenols to humic polymers through oxidative coupling reactions may involve oxidoreductive enzymes and soil minerals as catalysts. There is limited information on the possible inhibitory or synergistic interactions between oxidoreductases and mineral catalysts as they participate in oxidative coupling of phenolic substrates. In this study, a ternary system was investigated, in which a fungal enzyme (Trametes villosa laccase), birnessite (δ-MnO2), and a naturally occurring phenolic compound (catechol) were reacted together to model soil processes. Binary systems (catechol/laccase and catechol/birnessite) were included for comparison. In the absence of the mineral, T. villosa laccase (950 katal ml−1) transformed 31% of catechol, whereas birnessite (1 mg ml−1) in the absence of the enzyme showed a 24% catechol transformation. The percentages of catechol transformation in the binary systems did not accumulate in the ternary system; instead, birnessite and laccase tested together transformed only 36% of catechol. This suggested that birnessite had an inhibitory effect on substrate transformation by laccase catalysis. Enzyme assays indicated that inhibition was a result of enzyme deactivation by humic-like polymers produced by birnessite, and by Mn2+ ions released from the mineral. These observations underscore the importance of considering enzyme-soil mineral-organic matter interactions in studies of humus formation and contaminant removal.  相似文献   

4.
The adsorption of herbicides on soil colloids is a major factor determining their mobility, persistence, and activity in soils. Solvent extraction could be a viable option for removing sorbed contaminants in soils. This study evaluated the extractability of three herbicides: 2,4 dichlorophenoxy-acetic acid (2,4-D), 4-chloro-2-methylphenoxypropanoic acid (mecoprop acid or MCPP), and 3,6-dichloro-2-methoxybenzoic acid (dicamba). Three solvents (water, methanol, and iso-propanol) and three methods of extraction (column, batch, and soxhlet) were compared for their efficiencies in removing the herbicides from three soils (loamy sand, silt loam, and silty clay). Both linear and non-linear Freundlich isotherms were used to predict sorption intensity of herbicides on soils subjected to various extraction methods and conditions. High Kdand Kfr, and low N values were obtained for all herbicides in silty clay soil by batch extraction. Methanol was the best solvent removing approximately 97% of all added herbicides from the loamy sand either by column or soxhlet extraction method. Isopropanol ranked second by removing over 90% of all herbicides by soxhelet extraction from all three soils. However, water was ineffective in removing herbicides from any of the soils using any of the three extracting procedures used in this study. In general, the extent of herbicide removal depended on soil type, herbicide concentration, extraction procedure, solvent type and amount, and extraction time.  相似文献   

5.
Leaching of the broadleaf herbicides 2,4-D and dicamba from home lawns was monitored with ceramic extraction plates placed at a 0.2 m depth beneath undisturbed sod. The site was located on a Merrimac sandy loam. Four treatments, consisting of two rates of herbicide applications coupled with two irrigation regimes. were evaluated on 12 plots. The low herbicide rate consisted of 1.1 and 0.1 kg ha?1 yr?1 of 2,4-D and dicamba, respectively. The high rate used was 3.3 and 0.33 kg ha?1 yr?1 of 2,4-D and dicamba applied in three equal applications. Irrigation treatments were (1) minimal irrigation to avoid drought stress and percolation from the root zone and (2) overwatering at 37.5 mm week?1. Geometric mean concentra tions of 2,4-D ranged from 0.55 to 0.87 μg L?1 compared to 0.26 to 0.55 μg L?1 for dicamba. The low application-minimum irrigation treatment generated significantly higher concentrations than the other treatments for both herbicides. The low concentrations observed for both herbicides suggest that excellent degradation conditions exist in the root zone of turfgrass during the summer months when application occurs.  相似文献   

6.
The phenoxyalkanoic acid herbicides constitute a group of chemically related molecules that have been widely used for over 50 years. A range of bacteria have been selected from various locations for their ability to degrade these compounds. Previously reported strains able to utilise 2,4-dichlorophenoxyacetic acid (2,4-D) include, Ralstonia eutropha JMP134, Burkholderia sp. RASC and Variovorax paradoxus TV1 and Sphingomonas sp. AW5 able to utilise 2,4,5-trichlorophenoxyacetic acid (2,4,5-T). In addition a novel set of mecoprop-degrading strains including Alcaligenes denitrificans, Alcaligenes sp. CS1 and Ralstonia sp. CS2 are here described. It has been reported recently that TfdA enzymes, initially reported to have a role in 2,4-D catabolism are also involved in the first-step cleavage of related phenoxyalkanoate herbicides. However, a diversity of tfdA gene sequences have been reported. We relate the tfdA gene type to the metabolic ability of these strains. The tfdA-like genes were investigated by polymerase chain reaction amplification using a set of specific tfdA primers. Degradation ability was observed via phenol production from a range of unsubstituted and substituted phenoxyalkanoics including, 2,4-D, 2-methyl 4-chlorophenoxyacetic acid (MCPA), racemic mecoprop, (R)-mecoprop, 2-(2,4-dichlorophenoxy) propionic acid (racemic 2,4-DP), 2,4,5-T, 2,4-dichlorophenoxybutyric acid (2,4-DB), 4-chloro-2-methylphenoxybutyric acid (MCPB) and phenoxyacetate. Mecoprop-degrading strains showed partial tfdA sequences identical to the one described for V. paradoxus TV1 (a strain isolated on 2,4-D). However, substrate specificity was not identical as V. paradoxus exhibited greatest activity towards 2,4-D and MCPA only, whereas the mecoprop-degrading strains showed intense activity towards 2,4-D, MCPA, racemic mecoprop and (R)-mecoprop as substrates. However, Sphingomonas sp. AW5 which has been shown to carry a very different tfdA-like gene was the only strain to utilise the phenoxybutyric acid MCPB as a sole carbon source. In this study, we thus demonstrate that sequence diversity is not related to substrate specificity within the tfdA-like gene family. However, phylogenetically unrelated sequences may govern substrate specific activity.  相似文献   

7.
Rhizosphere enhanced biodegradation of organic pollutants has been reported frequently and a stimulatory role for specific components of rhizodeposits postulated. As rhizodeposit composition is a function of plant species and soil type, we compared the effect of Lolium perenne and Trifolium pratense grown in two different soils (a sandy silt loam: pH 4, 2.8% OC, no previous 2,4-D exposure and a silt loam: pH 6.5, 4.3% OC, previous 2,4-D exposure) on the mineralization of the herbicide 2,4-D (2,4-dichlorophenoxyacetic acid). We investigated the relationship of mineralization kinetics to dehydrogenase activity, most probable number of 2,4-D degraders (MPN2,4-D) and 2,4-D degrader composition (using sequence analysis of the gene encoding α-ketoglutarate/2,4-D dioxygenase (tfdA)). There were significant (P<0.01) plant-soil interaction effects on MPN2,4-D and 2,4-D mineralization kinetics (e.g. T. pratense rhizodeposits enhanced the maximum mineralization rate by 30% in the acid sandy silt loam soil, but not in the neutral silt loam soil). Differences in mineralization kinetics could not be ascribed to 2,4-D degrader composition as both soils had tfdA sequences which clustered with tfdAs representative of two distinct classes of 2,4-D degrader: canonical R. eutropha JMP134-like and oligotrophic α-proteobacterial-like. Other explanations for the differential rhizodeposit effect between soils and plants (e.g. nutrient competition effects) are discussed. Our findings stress that complexity of soil-plant-microbe interactions in the rhizosphere make the occurrence and extent of rhizosphere-enhanced xenobiotic degradation difficult to predict.  相似文献   

8.
Two concentrations of 2,4-dichlorophenoxyacetic acid (2,4-D) 1.7 kg ha?1 and 3.4 kg ha?1 were applied to oats (Avena sativa L. ‘Orbit’) grown in terrestrial microcosms in a sandy soil. Carbon dioxide evolution and non-symbiotic N2 fixation (C2H2 reduction) were measured weekly. On day 70 of the study, 2,4-D was applied a second time at the same application rates and soil CO2 evolution and N2 fixation were measured more frequently. Soil CO2 evolution over 24 h period was significantly decreased by 40 to 50% at both application rates of 2,4-D. This response lasted less than 7 days. Nitrogen fixation was unaffected by 2,4-D except for an unexplained decrease observed in the 1.7 kg ha?1 treatment 35 days after 2,4-D application. This effect was not observed on the following sampling date. The second application of 2,4-D failed to produce any significant change in soil CO2 evolution or N2 fixation. From these studies we conclude soil microbial populations either degraded or became acclimated to 2,4-D as a result of the initial application and that 2,4-D has no significant adverse effect on N2 fixation or soil CO2 evolution.  相似文献   

9.

Purpose

We review 2,4-dichlorophenoxyacetic acid (2,4-D) and other phenoxy herbicide sorption experiments.

Methods

A database with 469 soil–water distribution coefficients K d (in liters per kilogram) was compiled: 271 coefficients are for the phenoxy herbicide 2,4-D, 9 for 4-(2,4-dichlorophenoxy)butyric acid, 18 for 2-(2,4-dichlorophenoxy)propanoic acid, 109 for 2-methyl-4-chlorophenoxyacetic acid, 5 for 4-(4-chloro-2-methylphenoxy)butanoic acid, and 57 for 2-(4-chloro-2-methylphenoxy)propanoic acid. The following parameters characterizing the soils, solutions, or experimental procedures used in the studies were also compiled if available: solution CaCl2 concentration, pH, pre-equilibration time, temperature, soil organic carbon content (f oc), percent sand, silt and clay, oxalate extractable aluminum, oxalate extractable iron (Oxalate Fe), dithionite–citrate–bicarbonate extractable aluminum, dithionite–citrate–bicarbonate extractable iron (DCB Fe), point of zero negative charge, anion exchange capacity, cation exchange capacity, soil type, soil horizon or depth of sampling, and geographic location. K d data were also compiled characterizing phenoxy herbicide sorption to the following well-defined sorbent materials: quartz, calcite, α-alumina, kaolinite, ferrihydrite, goethite, lepidocrocite, soil humic acid, Fluka humic acid, and Pahokee peat.

Results

The data review suggests that sorption of 2,4-D can be rationalized based on the soil parameters pH, f oc, Oxalate Fe, and DCB Fe in combination with sorption coefficients measured independently for humic acids and ferrihydrite, and goethite.

Conclusions

Soil organic matter and iron oxides appear to be the most relevant sorbents for phenoxy herbicides. Unfortunately, few authors report Oxalate Fe and DCB Fe data.  相似文献   

10.
The herbicides 2,4-diclorophenoxiacetic and 4-chloro-2-methylphenoxyacetic acids (2,4-D and MCPA) are widely used in agricultural practices worldwide. Not only are these practices responsible of surface waters contamination, but also agrochemical industries through the discharge of their liquid effluents. In this investigation, the ability of a 2,4-D degrading Delftia sp. strain to degrade the related compound MCPA and a mixture of both herbicides was assessed in batch reactors. The strain was also employed to remove and detoxify both herbicides from a synthetic effluent in a continuous reactor. Batch experiments were conducted in a 2-L aerobic microfermentor, at 28 °C. Continuous experiments were carried out in an aerobic downflow fixed-bed reactor. Bacterial growth was evaluated by the plate count method. Degradation of the compounds was evaluated by UV spectrophotometry, gas chromatography (GC), and chemical oxygen demand (COD). Toxicity was assessed before and after the continuous process by using Lactuca sativa seeds as test organisms. Delftia sp. was able to degrade 100 mg L?1 of MCPA in 52 h. When the biodegradation assay was carried out with a mixture of 100 mg L?1 of each herbicide, the process was accomplished in 56 h. In the continuous reactor, the strain showed high efficiency in the simultaneous removal of 100 mg L?1 of each herbicide. Removals of 99.7, 99.5, and 95.0% were achieved for 2,4-D, MCPA, and COD, respectively. Samples from the influent of the continuous reactor showed high toxicity levels for Lactuca sativa seeds, while toxicity was not detected after the continuous process.  相似文献   

11.
The effects of 20 herbicides on denitrification of nitrate in three soils were studied by determining the effects of 10 and 50μgg?1 soil of each herbicide on the amounts of nitrate lost and the amounts of nitrite, N2O and N2 produced when soil samples were incubated anaerobically after treatment with nitrate. The herbicides used were butylate, EPTC, chlorpropham, propham, diuron, linuron, monuron, siduron, alachlor, trifluralin, 2,4-D amine, 2,4-D ester, atrazine, cyanazine, metribuzin, simazine, dalapon, chloramben, dicamba and dinoseb.None of the herbicides studied significantly affected denitrification of nitrate when applied at the rate of 10 μg g?1 soil, but dinoseb increased the ratio of N2 to N2O in the gaseous products of denitrification when applied at this rate. Butylate, EPTC, diuron, simazine and dalapon had no significant effect on denitrification when applied at the rate of 50μgg?1 soil, whereas metribuzin and dinoseb enhanced denitrification when applied at this rate. The influence of the other herbicides on denitrification when applied at the rate of 50μgg?1soil depended on the soil, but all enhanced or inhibited denitrification in at least one soil.  相似文献   

12.
The ability of 22 strains of Rhizobium to degrade catechol, protocatechuic acid, p-hydroxybenzoic acid and salicylic acid all at 1 mm concentration was examined. In the presence of 4.8 mm Na-glutamate, all rhizobia tested degraded catechol (99–100%), p-hvdroxybenzoic acid (79–99%), protocatechuic acid (81 97%) and salicylic acid (20–83%).The concentration of Na glutamate in the medium affected the degradation of the phenolic compounds at 1 mm concentration. Increased glutamate favoured degradation of p-hydroxybenzoic and salicylic acids but had little effect on catechol. Degradation of protocatechuic acid was inhibited by increased glutamate concentration.Rhizobium phaseoli 405 grown with 8.0 mM Na-glutamate, directly cleaved catechol and protocatechuic acid. p-Hydroxybenzoic acid was converted to protocatechuic acid before ring cleavage. Salicylic acid was converted to gentisic acid before further oxidation. O2 uptake experiments showed that R. phaseoli 405 grown with p-hydroxybenzoic acid was adapted to this compound and protocatechuic acid. A lag of 30 min was required for catechol and salicylic acid.  相似文献   

13.
The efficiency of Trametes versicolor laccase in the transformation of phenols (caffeic acid, catechol, hydroxytyrosol, methylcatechol, protocatechuic acid, syringic acid, m-tyrosol, 3-hydroxybenzoic acid, 3-hydroxyphenylacetic acid, 2,6-dihydroxybenzoic acid, 4-hydroxybenzaldehyde) usually present in waste water, such as that derived from an olive oil factory, was investigated. According to their response to 24 h laccase action the 11 phenolic compounds were classified in three groups: reactive (88-100% transformation), intermediate reactive (transformation lower than 50%), and recalcitrant (not transformed at all). The enzyme was able to transform the 11 substrates even when they were present in a mixture and also toward a phenolic extract from a Moroccan olive oil mill waste water (OMW) sample. The disappearance of protocatechuic, 3-hydroxyphenylacetic, and 2,6-dihydroxybenzoic acids, and 4-hydroxybenzaldehyde was enhanced whereas that of caffeic acid and m-tyrosol was depressed when the phenols were present in the mixture. A reduction of enzyme activity occurred in single and/or complex phenolic mixtures after enzymatic oxidation. No correspondence between phenol transformation and disappearance of enzymatic activity was, however, observed. The overall results suggest that laccases are effective in the transformation of simple and complex phenolic mixtures.  相似文献   

14.
Research on organoclays as sorbents of pesticides has shown the usefulness of these materials as pesticide supports to prolong the efficacy of soil-applied pesticides and to reduce the large transport losses that usually affect pesticides applied in an immediately available form. Nevertheless, little information exists on the availability of organoclay-formulated pesticides for bacterial degradation. In this work, laboratory experiments were conducted to determine the adsorption-desorption behavior of two hexadecyltrimethylammonium-treated Arizona montmorillonites (SA-HDTMA50 and SA-HDTMA100) for the herbicide 2,4-dichlorophenoxyacetic acid (2,4-D), and to evaluate the ability of these organoclays to slow the release of the herbicide and to reduce herbicide leaching losses as compared to the free (technical) compound. The kinetics of mineralization of free and formulated 2,4-D by adapted bacteria was also determined. Organoclay-based formulations of 2,4-D displayed slow release properties in water and reduced herbicide leaching through soil columns, while maintained a herbicidal efficacy similar to that of the free (technical) 2,4-D. The total amount of 14C-2,4-D mineralized at the end of the biodegradation experiment (t=130 h) ranged between 30% and 46% of the formulated herbicide, which represented 53-81% of the amount of free 2,4-D mineralized in the same conditions. The release, leaching, and mineralization patterns of the formulated herbicide were found to depend both on the affinity of the organoclay for the herbicide and on the degree of interaction promoted during the preparation of the herbicide-organoclay complex. This suggests the possibility to select diverse preparations to achieve the desired release, leaching and biodegradation behavior.  相似文献   

15.
Abstract

Short‐term response of microbial respiration after treatment with different doses of the herbicides metsulfuron methyl (MET), 2,4‐D, and glyphosate (GLY) was studied in microcosms of soils collected in three agricultural sites of the Southern Pampas region, Buenos Aires, Argentina. The influence of diammonium phosphate [(NH4)2PO4] on carbon dioxide (CO2) evolution, when applied with the highest doses of the herbicides, was also investigated. MET had no effect on microbial respiration of an acidic soil of San Román (pH 6.06), even at the highest rate. However, MET inhibited microbial respiration in soils of Bordenave (pH 7.44), at a rate of 0.1 mg kg?1 soil. Low application rates of GLY and 2,4‐D produced only transitory effects on CO2 evolution, whereas the addition of high doses of these herbicides stimulated microbial activity. On the other hand, the addition of fertilizer to soil treated with a high dose of GLY temporarily inhibited CO2 release.  相似文献   

16.
The ability of Pleurotus eryngii, Pleurotus ostreatus, Pleurotus pulmonarius and Pleurotus sajor-caju to degrade the aromatic pollutants 2,4-dichorophenol (2,4-DCP) and benzo(a)pyrene [B(a)P] in liquid culture and microcosm (using wheat straw as growth substrate and sea sand as a xenobiotic carrier) was investigated by HPLC and 14CO2 release from labeled pollutants. We found that 100 μM 2,4-DCP was very quickly transformed by the four fungi, disappearing 24 h after its addition to the liquid cultures. However, a 2-week incubation period was required to transform 100 μM B(a)P up to 75% by P. eryngii and P. pulmonarius. Whereas the fungi were able to begin degradation of the two pollutants with high transformation rates, their complete degradation (mineralization) rates were very low. Mineralization of B(a)P in liquid cultures was only observed with P. eryngii and P. pulmonarius, although the four Pleurotus species studied were able to mineralize this compound in solid state fermentation (SSF). The ligninolytic enzymes laccase and versatile peroxidase (VP), together with aryl-alcohol oxidase (AAO) providing extracellular H2O2, were found in liquid cultures. Except AAO, these enzymes were also detected in SSF experiments. In order to investigate the role of ligninolytic enzymes in the process, their action on both pollutants (50 μM) was studied in vitro in the absence and presence of redox mediators. As observed with the fungal cultures, 2,4-DCP was oxidized faster than B(a)P by both laccase (60% transformation after 6 h) and VP (100% transformation after 1 h). Moreover, laccase oxidation was strongly increased (up to 90% transformation after 3 h), by the presence of the mediators 2,2′-azino-bis-(3-ethylbenzothiazoline-6-sulphonic acid) (ABTS) or 1-hydroxybenzotriazole (HBT). In the case of B(a)P, the presence of ABTS or HBT was strictly required for oxidation by laccase (25% transformation after 8 h). Degradation of B(a)P was also observed in reactions with VP (40% transformation after 6 h). The results obtained suggest that Pleurotus species can be used in applications focused to the degradation of aromatic pollutants using wheat straw as a growth substrate, and provide the first evidence on the direct transformation of recalcitrant aromatic pollutants by VP.  相似文献   

17.
Summary Wheat seedlings, treated with the auxine 2,4-dichlor-phenoxy acetic acid (2,4-D) during germination developed only a residual root system. Root elongation was extremely restricted and root tips were deformed to thick club-shaped tumours. When 2,4-D was added in a later stage of plant growth the plants developed additional nodule-like knots along primary roots. Root and shoot dry-matter production was slightly repressed in all 2,4-D treatments and N translocation from roots to shoots was repressed as well. When transferred to an auxine-free growth medium, the 2,4-D-affected roots were not capable of complete recovery. In plants inoculated gnotobiotically with Azospirillum brasilense, either with the wild type or with the NH 4 + -excreting mutant strain C3, a 2,4-D addition increased rhizosphere acetylene-reduction activity at pO2 1.5 kPa. The O2 sensitivity of root-associated nitrogenase activity tended to be reduced. The number of root-colonizing bacteria, at approximately 108 colony-forming units (cfu) per g dry root, was similar in the 2,4-D treatments and untreated controls. Plant treatment with high concentrations of the chemical isomer 3,5-dichlor-phenoxy acetic acid (3,5-D) did not have comparable effects, either on plant development or on rhizosphere-associated nitrogenase activity. Root-tumour tissue inhabited by A. brasilense showed purple staining when subjected to a tetrazolium chloride solution, which may indicate intensive local nitrogenase activity in this tissue. Exposed to an 15N2-enriched atmosphere, plants treated with 2,4-D and with A. brasilense incorporated significantly higher amounts of 15N than untreated controls. In all cases the highest values of 15N enrichment were found following inoculation with the NH 4 + -excreting mutant strain C3. Present address: F. A. Janssens Laboratory of Genetics, Catholic University of Leuven, Willem de Croylan 42, B-3001 Heverlee, Belgium  相似文献   

18.
Seventy-six rhizobial isolates belonging to four different genera were obtained from the root nodules of several legumes (Vicia sativa, Vicia faba, Medicago sativa, Melilotus sp., Glycine max and Lotus corniculatus). The action of five commonly used herbicides [2,4-dichlorophenoxyacetic acid (2,4-D), glyphosate (GF), dicamba, atrazine and metsulfuron-methyl] on the growth of rhizobial strains was assessed. Subsequently, GF and 2,4-D were tested in a minimum broth as C and energy sources for 20 tolerant strains. The ability of these strains to metabolize different carbon sources was studied in order to detect further differences among them. Tolerance of the bacteria to agrochemicals varied; 2,4-D and GF in solid medium inhibited and diminished growth, respectively, in slow-growing rhizobial strains. Among slow-growing strains we detected Bradyrhizobium sp. SJ140 that grew well in broth + GF as the sole C and energy source. No strain was found which could use 2,4-D as sole C source. The 20 strains studied exhibited different patterns of C sources utilization. Cluster analysis revealed three groups, corresponding to four genera of rhizobia: Rhizobium (group I), Sinorhizobium (group II) and Mesorhizobium–Bradyrhizobium (group III). On the basis of the results obtained on responses to herbicides and C sources utilization by the isolates investigated, it was possible to differentiate them at the level of strains. These results evidenced a considerable diversity in rhizobial populations that had not been previously described for Argentinean soils, and suggested a physiological potential to use natural and xenobiotic C sources.  相似文献   

19.
The biodegradation of organic compounds in soil is a key process that has major implications for different ecosystem services such as soil fertility, air and water quality, and climate regulation. Due to the complexity of soil, the distributions of organic compounds and microorganisms are heterogeneous on sub-cm scales, and biodegradation is therefore partly controlled by the respective localizations of organic substrates and degraders. If they are not co-localized, transfer processes become crucial for the accessibility and availability of the substrate to degraders. This spatial interaction is still poorly understood, leading to poor predictions of organic compound dynamics in soils. The objectives of this work were to better understand how the mm-scale distribution of a model pesticide, 2,4-dichlorophenoxyacetic acid (2,4-D), and its degraders drives the fate of 2,4-D at the cm soil core scale. We constructed cm-scale soil cores combining sterilized and “natural” soil aggregates in which we controlled the initial distributions of 2,4-D and soil microorganisms with the following spatial distributions: i) a homogeneous distribution of microorganisms and 2,4-D at the core-scale, ii) a co-localized distribution of microorganisms and 2,4-D in a single spot (360 mm3) and iii) a disjoint localization of microorganisms and 2,4-D in 2 soil spots (360 mm3) separated by 2 cm. Two sets of experiments were performed: one used radiolabeled 14C-2,4-D to study the fate of 2,4-D, and the other used 12C-2,4-D to follow the dynamics of degraders. Microcosms were incubated at 20 °C and at field capacity (−31.6 kPa). At the core scale, we followed 2,4-D mineralization over time. On three dates, soil cores with microorganisms and 2,4-D localized in soil spots, were cut out in slices and then in 360 mm3 soil cubes. The individual soil cubes were then independently analysed for extractable and non-extractable 14C and for degraders (quantitative PCR of tfdA genes). Knowing the initial position of each soil cube allowed us to establish 3D maps of 2,4-D residues and degraders in soil. The results indicated that microorganisms and pesticide localizations in soil are major driving factors of i) pesticide biodegradation, by regulating the accessibility of 2,4-D to degrading microorganisms (by diffusion); and ii) the formation of non-extractable residues (NER). These results also emphasized the dominant role of microorganisms in the formation and localization of biogenic NER at a mm-scale. To conclude, these results demonstrate the importance of considering micro-scale processes to better understand the fate of pesticides and more generally of soil organic substrates at upper scales in soil and suggest that such spatial heterogeneity should not be neglected when predicting the fate of organic compounds in soils.  相似文献   

20.
A phenol-utilizing strain of Rhodotorula glutinis was isolated from Rothamsted soil. Washed, phenol-grown R. glutinis cells oxidized 3- and 4-chlorophenols to 4-chlorocatechol, and 4-bromophenol presumably to 4-bromocatechol. Phenol-grown cells consumed O2 in the presence of 2-, 3- or 4-chlorophenol, 4-bromophenol, 2,4-dichloro- and 2,4-dibromophenol. Benzoate-grown cells showed O2-uptakes in the presence of 2-, 3- or 4-chlorobenzoate but the formation of catechol compounds was not definitely established.  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