首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The degradation of the wild oat herbicide flamprop-methyl [methyl DL -N-benzoyl-N-(3-chloro-4-fluorophenyl)alaninate] in four soils has been studied under laboratory conditions using 14C-1abelled samples. The flamprop-methyl underwent degradation more rapidly than its analogue flamprop-isopropyl. However, similar degradation products were formed, namely the corresponding carboxylic acid and 3-chloro-4-fluoroaniline. The latter compound occurred mainly as ‘bound’ forms although evidence was obtained of limited ring-opening to give [14C]carbon dioxide. The time for depletion of 50% of the applied herbicide was approximately 1-2 weeks in sandy loam, clay and medium loam soils and 2-3 weeks in a peat soil.  相似文献   

2.
The degradation of [14C] benzoyl prop ethyl (SUFFIX,a ethyl N-benzoyl-N-(3,4-dichlorophenyl)-2-aminopropionate) in four soils has been studied under laboratory conditions. The major degradation product of benzoylprop ethyl at up to 4 months after treatment was its corresponding carboxylic acid (II). On further storage this compound became firmly bound to soil before it underwent a slow debenzoylation process which led to the formation of a number of products including N-3,4-dichlorophenylalanine (IV), benzoic acid, 3,4-dichloroaniline (DCA), which was mainly present complexed with humic acids, and other polar products. Although these polar products were not identified, they were probably degradation products of DCA, since they were also formed when DCA was added to soil. No 3,3′,4,4′-tetrachloroazobenzene (TCAB) was detected in any of the soils at limits of detectability ranging from 0.01-0.001 parts/million. Since N-3,4-dichlorophenylalanine (IV) and 3,4-dichloroaniline were transient degradation products of benzoylprop ethyl, the metabolism in soil of radiolabelled samples of these compounds was also studied. In these laboratory experiments the persistence of the herbicide increased as the organic matter content of the soil increased and the time for depletion of half of the applied benzoylprop ethyl varied from 1 week in sandy loam and clay loam soils to 12 weeks in a peat soil.  相似文献   

3.
The degradation of the wild oat herbicide flamprop-methyl [MATAVEN, methyl (±)-N-benzoyl-N-(3-chloro-4-fluorophenyl)-2-aminopropionate] was studied in soils stored under anaerobic conditions. Comparative experiments were carried out in which soil was either covered with water or stored in an atmosphere of nitrogen. Under these anaerobic conditions, the major product was the carboxylic acid analogue (II) of flamprop-methyl, which was also a major degradation product formed in soil stored under aerobic conditions. However, the 2-, 3-, and 4-hydroxy-benzoyl analogues of II were also detected in soils stored under nitrogen or water and they were present in highest concentrations in the waterlogged soil. A further new product was also detected in waterlogged soil and it was shown to be N-benzoyl-N-(3-chloro-4-hydroxyphenyl)-2-aminopropionic acid. Although no hydroxylated derivatives of flamprop-methyl were detected in soils stored under aerobic conditions, it is possible that they were formed but underwent further degradation.  相似文献   

4.
Kim J  Liu KH  Kang SH  Koo SJ  Kim JH 《Pest management science》2003,59(11):1260-1264
LGC-42153, 2-fluoro-1-[3-(4,6-dimethoxypyrimidin-2-ylcarbamoylsulfamoyl)pyridin-2-yl]propyl methoxyacetate, is a new sulfonylurea herbicide for use in rice. Its breakdown and metabolism were studied in soil under flooded condition using radioactive tracers labelled at either the propyl group or the pyrimidine ring. The half-life of LGC-42153 was approximately 3.0 days. The mass balance over 120 days ranged from 94.0 to 104.2% of applied radiocarbon, and no significant amount of volatiles or [14C]carbon dioxide were observed. Solvent non-extractable radiocarbon reached 11 approximately 14% of applied radiocarbon at 120 days after treatment. The major metabolic reaction was the cleavage of the carboxyl ester bond to give 1-(4,6-dimethoxypyrimidin-2-yl)-3-[2-(1-hydroxy-2-fluoropropyl)pyridine-3-sulfonyl]urea, which underwent hydrolysis of the sulfonylurea bridge giving 2-(1-hydroxy-2-fluoro)propyl-3-pyridinesulfonamide and 4,6-dimethoxy-2-aminopyrimidine.  相似文献   

5.
The degradation of the wild-oat herbicide flamprop-isopropyl, [isopropyl (±)-N-benzoyl-N-(3-chloro-4-fluorophenyl)alaninate], in four soils has been examined under laboratory conditions with sampling times of up to 45 weeks after treatment. The major degradation product of [14C]flamprop-isopropyl in all soils at up to 10 weeks after treatment was the carboxylic acid (±)-N-benzoyl-N-(3-chloro-4-fluorophenyl)alanine. This compound in turn underwent degradation by loss of the benzoyl group and the propionic acid moiety, with evolution of [14C]carbon dioxide to form 3-chloro-4-fluoroaniline (CFA). The CFA was formed slowly in soil and occurred mainly as a bound form. There was evidence to show that the CFA was subsequently converted into other polar products. The time for depletion of 50% of the applied herbicide was approximately 10 weeks in sandy loam and medium loam soils, 11 weeks in a clay loam soil and 23 weeks in a peat soil.  相似文献   

6.
LGC-42153, 2-fluoro-1-[3-(4,6-dimethoxypyrimidin-2-ylcarbamoylsulfamoyl)pyridin-2-yl] propyl methoxyacetate, is a new sulfonylurea herbicide for use in rice. Its breakdown and metabolism was studied in soil under flooded conditions using two radioactive tracer compounds labelled at either the propyl group or the pyrimidine ring. The half-life of LGC-42153 was approximately 3.0 days. The mass balance over 120 days ranged from 94.0 to 104.2% of applied radiocarbon, and no significant amount of volatiles or [14C]carbon dioxide were observed. Solvent non-extractable radiocarbon reached about 11-14% of applied radiocarbon at 120 days after treatment. The major metabolic reaction was the cleavage of the carboxyl ester bond to give 1-(4,6-dimethoxypyrimidin-2-yl)-3-[2-(1-hydroxy-2-fluoropropyl)pyridine-3-sulfonyl]urea, which underwent hydrolysis of the sulfonylurea bridge giving 2-(1-hydroxy-2-fluoro)propyl-3-pyridinesulfonamide and 4,6-dimethoxy-2-aminopyrimidine.  相似文献   

7.
In addition to the molecular structure of a pesticide, environmental conditions may influence its persistence through their effect on the growth and activity of pesticide-degrading micro-organisms. As a result, transformation rates may decrease rapidly when a compound is leached into subsoil. Metamitron sorption isotherms were determined and incubation series were set up for a sandy loam soil, simulating single and combination effects that occur during transport of metamitron into subsoils. KOC values increased with increasing depth from 185 to 700 litre kg−1. A combination of conditions that are unfavourable for microbial activity, such as low temperature (5°C), low concentrations (0·5 mg kg−1) and a large sorbed fraction (KOC = 700) resulted in half-lives of over one year. Oxygen inhibition decreased the transformation rate of metamitron from 0·058 to 0·019 day−1. In order of significance, the transformation of metamitron appears to be a function of temperature, oxygen availability and sorption to organic carbon. Increasing doses did not change transformation rates significantly, although different transformation pathways were observed.  相似文献   

8.
The degradation of the herbicide diclofop-methyl, ( ± )-methyl 2-[4-(2,4-dichloro-phenoxy)phenoxy]propionate, was investigated in two agricultural soils under aerobic and anaerobic conditions. Using two differently labelled forms of [14C]-diclofop-methyl the qualitative as well as the quantitative formation of extractable metabolites was followed for 64 days. The mineralisation of the uniformly labelled aromatic rings was pursued by monitoring the 14CO2 generated for 25 weeks. As a first step of the degradation a very rapid hydrolysis of the ester bond was detected under all conditions. Diclofop, the corresponding substituted propionic acid formed, was extensively degraded under aerobic conditions, the final product being 14CO2. As an intermediate, a compound later identified by GLC/MS to be 4-(2,4-dichlorophenoxy)phenol, was found in the extracts. Furthermore, traces of six other unknown metabolites were detected. Under anaerobic conditions the degradation proceeded to a small extent. At most 3% of the applied radioactivity was accounted for by the degradation product 4-(2,4-dichlorophenoxy)phenol. No other metabolite, including 14CO2, was observed, implying lack of any further degradation.  相似文献   

9.
10.
We investigated the sorption of five widely used sterol biosynthesis inhibitor fungicides (SBIs: flusilazole, propiconazole, epoxiconazole, fenpropimorph and prochloraz) on a loam soil to assess availability of the SBI residues that are usually left in soil after crop treatments. We focused particularly on the soil moisture content effect, which is poorly documented and is difficult to investigate under realistic conditions. SBI sorption was determined (using diuron as a reference) at two soil moisture contents (26.1% and 46.6% w/w) over a period of 3 weeks using a direct soil solution sampling method. After 24 h of contact, <1% of each applied fungicide was recovered in the soil solution. Despite their low availability in the liquid phase, long‐term sorption was observed for all the compounds, reducing concentrations in the soil solution and doubling the value of the partition coefficient. Significant effects of soil moisture on long‐term sorption were observed, depending on the properties of the chemicals and the sorption mechanisms. Wershaw's humus model (humic substances have a membrane‐like structure) was adapted to fit our observations. Low soil moisture content is assumed to modify the structure of humic substances and to generate hydrophobic surfaces, which favour sorption of hydrophobic fungicides (flusilazole, propiconazole and epoxiconazole). This effect is likely to decrease with the increase in the hydrophobic character of non‐ionic pesticides. It becomes adverse for the more hydrophilic compounds (diuron), which are more sorbed at high soil moisture content due to their higher affinity for hydrophilic regions of humus and to diffusion. Soil moisture effects are more complex when compounds are likely to be protonated in soil. Weakly basic compounds (prochloraz) may partition rapidly into the liquid‐like interior of humus at low soil moisture content but increased diffusion at high soil moisture content may cause additional sorption by ion exchange at colloid surfaces. Strongly basic compounds (fenpropimorph) may essentially adsorb due to ionic interactions with colloids, and their sorption is enhanced at high soil moisture content due to diffusion. Consequences for environmental fate and biological activity of pesticides are briefly discussed. © 2000 Society of Chemical Industry  相似文献   

11.
The main processes which affect herbicide action in soil are considered to be the extent to which the herbicide is adsorbed by soil particles, the rate of absorption of the herbicide solution by plant roots and the rate at which the herbicide is lost from the soil by decomposition, evaporation or leaching. An attempt is made to show how these factors are interrelated and the possible use of mathematical models to optimise the performance of herbicides is briefly discussed.  相似文献   

12.
The route and rate of degradation of florasulam, a low‐rate triazolopyrimidine sulfonanilide herbicide, was investigated in six soil types under aerobic conditions at 20 or 25 °C. Degradation products were isolated and identified by mass spectroscopy. Florasulam was rapidly degraded by microbial action with an average half‐life of 2.4 days (range 0.7 to 4.5 days). The first step in the degradation pathway involved conversion of the methoxy group on the triazolopyrimidine ring to a hydroxy group to form N‐(2,6‐difluorophenyl)‐8‐fluoro‐5‐hydroxy[1,2,4]triazolo[1,5‐c]pyrimidine‐2‐sulfonamide. This metabolite degraded, with a half‐life of 10 to 61 days, via partial breakdown of the triazolopyrimidine ring to form N‐(2,6‐difluorophenyl)‐5‐aminosulfonyl‐1H‐1,2,4‐triazole‐3‐carboxylic acid. This was followed by cleavage of the sulfonamide bridge to form 5‐(aminosulfonyl)‐1H‐1,2,4‐triazole‐3‐carboxylic acid. Other degradation processes involved decarboxylation of the carboxylic acid metabolites and mineralisation to form carbon dioxide and non‐extractable residues. © 2000 Society of Chemical Industry  相似文献   

13.
Empirical equations were used to calculate the moisture content of surface soil from measurements of rainfall and daily maximum and minimum air temperatures. Air temperatures were also used to calculate soil temperatures. There was good agreement between calculated and measured moisture contents and temperatures from Wellesbourne and from some sites in North America. The equations were incorporated into a simulation model for the prediction of herbicide persistence. Results from the model were essentially the same, whether calculated or measured soil moistures and temperatures were used in the calculations.  相似文献   

14.
The sorption kinetics of photosystem II herbicides in the unicellular microalgae Ankistrodesmus braunii are independent of the herbicide concentration. While diuron-type herbicides attain an equilibrium state of sorption after at most 5 min, it takes up to 60 min for phenol-type herbicides. The kinetics of the inhibition of photosynthetic electron transport display a good correlation to the kinetics of sorption. For both types of herbicides, a phase of prevalent partitioning into the lipid phase of the cell membranes preceeds binding to thylakoid membranes. With phenol-type herbicides this phase is much more extended than with diuron-type compounds. As a consequence, there is no displacement of previously bound herbicides of this type from algal cells by addition of an excess of other photosystem II herbicides. The sorption of phenol-type herbicides to microalgae is under strong influence of environmental factors, such as light, temperature, pH, or oxygen concentration. Thus, besides the lipophilicity of these biocides, the response of the target cells to environmental factors is at least as important for herbicide sorption in the cells. From these results it is concluded that a prediction of the biocide accumulation in aquatic microorganisms can only be partly deduced from the properties of the pesticide molecule. The target cells and their complex responses to their environment strongly relativize correlations of chemical properties, e.g., lipophilicity and biological activity.  相似文献   

15.
Photosynthesis is the single most important source of O2 and organic chemical energy necessary to support all non-autotrophic life forms. Plants compartmentalize this elaborate biochemical process within chloroplasts in order to safely harness the power of solar energy and convert it into usable chemical units. Stresses (biotic or abiotic) that challenge the integrity of the plant cell are likely to affect photosynthesis and alter chlorophyll fluorescence. A simple three-step assay was developed to test selected herbicides representative of the known herbicide mechanisms of action and a number of natural phytotoxins to determine their effect on photosynthesis as measured by chlorophyll fluorescence. The most active compounds were those interacting directly with photosynthesis (inhibitors of photosystem I and II), those inhibiting carotenoid synthesis, and those with mechanisms of action generating reactive oxygen species and lipid peroxidation (uncouplers and inhibitors of protoporphyrinogen oxidase). Other active compounds targeted lipids (very-long-chain fatty acid synthase and removal of cuticular waxes). Therefore, induced chlorophyll fluorescence is a good biomarker to help identify certain herbicide modes of action and their dependence on light for bioactivity.  相似文献   

16.
In not experiments with Agropyron repens, Stellaria media and dwarf bean (Phaseolus vulgaris), several acids which are complexing agents enhanced the effects of glyphosate and dichlorprop. Both herbicides were activated by orthophosphoric citric, tartaric, lactic, oxalic and glycollic acids. Other acids which are not complexing agents had little or no effect. Sodium or ammonium salts often acted like the parent acids. Glyphosate activity was increased by EDTA while dichlorprop effects were enhanced by FDTA, NTA and polyphosphates. It is suggested that activation was due to interactions with calcium or other metallic ions, which would otherwise immobilize the herbicides. In field experiments, orthophosphoric or oxalic acids increased the effects of glyphosate on Agrophron bud viability. Ammonium sulphate had a similar effect. However, in contrast to glasshouse experiments, most three-way mixtures of an acid, ammonium sulphate glyphosate wer antagonistic. Les agents complexants en tant qu adjuvants pour les herbicides Dans des experiences en pots avec Agropyron repens, Stellariamedia et des haricots (Phaseolus vulgaris), plusieurs acides qui sont des agents complexants out augmenté l'activitié du glyphosate et du dichlorprop. Ces deux herbicides ont été activés par les acides orthophosphorique, citrique, tartrique, lactique, oxalique et glycollique. D'autres acides, qui ne sont pas des agents complexants, n'ont eu que peu ou pasd'effet. Les sels de sodium ou d'ammonium présentent souvent la méme activité que les acides don't ils dérivent. L'activite du glyphosate a été accrue par l'EDTA, cependant que les effets du dichloprop ont augmenté avec l'EDTA, le NTA et les polyphosphates. II est suggéré que cette activation était due aux interactions avec le calcium ou d'autres ions métalliques qui pourraient, par ailleurs, immobiliser les herbicides. Dans des expériences au champ, les acides orthophosphorique ou oxalique ont augmenté les effets du glyphosate sur la viabilité des bourgeons d'Agropyron. Le sulfate d'ammonium a eu en effet similaire. Toutefois, en opposition avec les expériences en serre, la plupart des mélanges ternaries d'un acide, de sulfate d'ammonium et de glyphosate se sont révélés antagonists. Komplxbildner als Additive für Herbizide In Gefässversuchen mit Agropyron repens, Stellaria media und Phasealus vulgaris wurde die Wirkung von Glyphosat und Diehlorprop durch komplexbildende Säuren verstärkt. Beide Herbizide wurden a tiviert durch: ortho-Phosphor. Zitronen, Wein, Milch-, Oxal- und Glykolsäure. Andere Säuren, die micht komplexbildend sind, wirkten nur schwach oder gar nicht. Natrium-oder Ammoniumsalze wirkten häufig wie die entsprechenden Säuren. Die Aktivität von Glyphosat wurde durch EDTA verstärkt, während die Wirkung von Dichlorprop durch EDTA, NTA und durch Polyphosphate verbessert wurde. Es wird vermutet, dass die bessere Wirkung in Interaklionen mit Kalzium oder anderern Metallionen zu suchen ist, die sonst die Herbizide immobilisieren. In Feldversuchen wurde durch ortho-Phosphorsäure und Oxalsäure die Wirkung von Glyphosat auf die Lebensfähigkeit von Agropyron-Knospen erhöht. Ammoniumsulfat wirkte ähnlich Im Gegensatz zu Versuchen im Gewächshaus wirkien aber die meisten Dreikomponenten Mischungen von Säure. Ammoniumsulfat und Glyphosat, antagonistisch  相似文献   

17.
氟乐灵的微生物生态效应   总被引:1,自引:0,他引:1  
研究了3种浓度氟乐灵对南疆棉田2种土壤(沙土、黏土)4种土壤酶(过氧化氢酶、转化酶、脲酶和磷酸酶)活性的影响。结果表明,同一时间同一浓度氟乐灵处理土壤除过氧化氢酶外,其余3种酶均表现为沙土高于黏土;过氧化氢酶酶活性变化趋势为“下降—升高—趋于平缓”。转化酶和脲酶的变化趋势类似,为“下降—升高—下降—趋于平缓”。磷酸酶变化趋势为“上升—剧烈下降—逐渐升高并趋于平缓”。即不同酶对氟乐灵的反应不同。3种浓度的氟乐灵处理两种土壤,其酶活性经过显著性分析无明显差异,即棉田用氟乐灵作土壤封闭处理对酶活性无影响。  相似文献   

18.
A. WALKER 《Weed Research》1987,27(2):143-152
The movement and persistence of residues of propyzamide, linuron, isoxaben and R-40244 were measured in a sandy loam soil in field experiments prepared in spring and autumn. None of the herbicides moved to depths greater than 12 cm in the soil during the winter period, following application in autumn, and none moved more than 6 cm in the soil, following application in spring. The general order of persistence of total soil residues was isoxaben > linuron = R-40244 > propyzamide. Appropriate constants to describe the moisture and temperature dependence of degradation were derived from laboratory incubation experiments and used with measurements of the strengths of adsorption of the different herbicides by the soil, in a computer model of herbicide movement. The model, in general, gave good predictions of total soil residues, but overestimated herbicide movement, particularly in winter. Measurements of herbicide desorption from the soil at intervals, during a laboratory incubation experiment, demonstrated an apparent increase in the strength of adsorption with time. When appropriate allowance was made for these changes in adsorption in the computer model, improved predictions of the vertical distribution of the herbicide residues were obtained.  相似文献   

19.
A stable mixed bacterial culture which degrades the herbicide linuron was isolated from soil by enrichment with linuron in a liquid mineral medium. Radio-respirometry studies showed that the culture mineralised linuron completely. No intermediate degradation products were detected in the medium. The culture was able to utilise linuron as a source of both nitrogen and carbon and was also able to degrade the related herbicides monolinuron and chlorbromuron and the possible intermediate degradation products of linuron: 3,4-dichlorophenyl-l-methylurea, 3,4-dichlorophenylurea and 3,4-dichloroaniline. The culture was unable to degrade the 1,1-dimethyl substituted ureas monuron, diuron or metoxuron. The culture contained Gram-negative aerobic rods, and Gram-positive aerobic non-spore-forming rods and cocco-bacilli. Of 124 isolates from the mixed culture, none degraded linuron in pure culture, indicating that a consortium of organisms is involved. Further investigation suggested that Pseudomonas spp. were important components of the population responsible for degradation.  相似文献   

20.
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