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.
SARMAH  KOOKANA  & ALSTON 《Weed Research》1999,39(2):83-94
The degradation of chlorsulfuron and triasulfuron was investigated in alkaline soils (pH 7.1–9.4) spiked at 40 μg a.i. kg–1 under laboratory conditions at 25 °C and a moisture content corresponding to 70% field capacity (–33 kPa), using high-performance liquid chromatography. Degradation data for the two herbicides did not follow first-order kinetics, and observed DT50 values in surface soils ranged from 19 to 42 days and from 3 to 24 days for chlorsulfuron and triasulfuron respectively. Disappearance of both chlorsulfuron and triasulfuron was faster in non-sterile than in sterile soil, demonstrating the importance of microbes in the breakdown process. The persistence of chlorsulfuron increased with increasing depth, which can be attributed to the decline in the microbial populations down the profile. The DT50 value for chlorsulfuron at 30–40 cm depth was nearly four times higher than that in the top-soil. The results obtained show that persistence of these herbicides in alkaline surface soils at 25 °C and at a moisture content of 70% field capacity is similar to those reported in other European and North American soils. The study shows that if these herbicides are contained in surface soil layers, the risk of residue carry-over under southern Australian conditions is small. However, the rate of their degradation in alkaline subsoils is very slow, and under conditions conducive to leaching their prolonged persistence in the soil profile is possible.  相似文献   

2.
Hollaway  Kookana  McQuinn  Moerkerk  Noy  & Smal 《Weed Research》1999,39(5):383-397
The ability of bioassay, enzyme-linked immunosorbent assay (ELISA) and high-performance liquid chromatography (hplc) methods to detect sulfonylurea herbicides in soil was evaluated as part of a project studying the leaching and persistence of these herbicides in the alkaline soils of south-eastern Australia. Soil samples with known concentrations between 0.1 and 10 μg a.i. kg−1 chlorsulfuron, metsulfuron-methyl or triasulfuron were prepared by an independent laboratory and supplied in coded bags to separate laboratories for testing. The accuracy of the results was analysed, and the merits of each method are discussed. Bioassay was suitable for measuring biologically active residues from 0.1 to 1.0 μg a.i. kg−1. ELISA accurately measured residues in the range of 0.1–10 μg a.i. kg−1, making it the most widely adaptable assay tested. It will be useful for measuring residues in sodic subsoils where bioassay plants grow poorly. There was good reproducibility between the bioassay and ELISA. The hplc technique used in this study was not as accurate as bioassay or ELISA at quantifying residues of 3.0–10 μg a.i. kg−1 and could not detect residues at or below 1.0 μg a.i. kg−1.  相似文献   

3.
Photolysis of chlorsulfuron and metsulfuron-methyl was studied in methanol under UV light. Their rates of primary photolysis followed first-order kinetics. The main photoproducts were identified as 2-methoxy-4-methyl-1,3,5-triazin-6-amine, 2-chloro-benzenesulfonamide and methyl 2-(aminosulfonyl)benzoate, which entailed the cleavage of the two N–C ureic bonds. Further photolysis of benzenesulfonamide derivatives involved oxidation of −NH2, cyclisation with loss of CH3OH, and scission of the C–S bond A trace of methyl o-mercaptobenzoate was also detected. The corresponding photolysis pathways of chlorsulfuron and metsulfuron-methyl were tentatively proposed. © 1999 Society of Chemical Industry  相似文献   

4.
The sensitivity of 22 major crops, pastures and weeds from the north-east grain region of Australia to atrazine and chlorsulfuron residues was determined in a glasshouse using a soil-free bioassay system. A logistic equation was fitted to the seedling fresh weights as a function of the logarithm of herbicide concentration by non-linear regression and used to calculate the doses for 10%, 30% and 50% inhibition of seedling growth (ID10, ID30 and ID50). The ID50 for atrazine ranged from 0.03 to 0.04 mg a.i. L–1 for Salvia reflexa Hornem. and barley to 1.47 mg a.i. L–1 for sorghum. The ID50 for chlorsulfuron ranged from 0.19 to 0.21 μg a.i. L–1 for lucerne and snail medic to 102 μg a.i. L–1 for wheat. Based on ID50 values measured, the predicted responses of each species to a range of concentrations of atrazine and chlorsulfuron were classified into four categories ranging from no damage to severe damage. These sensitivity data will assist in planning cropping sequences in soils previously treated with atrazine or chlorsulfuron.  相似文献   

5.
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   

6.
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   

7.
MIA-QSAR evaluation of a series of sulfonylurea herbicides   总被引:1,自引:0,他引:1  
  相似文献   

8.
Vapor pressure values for agricultural chemicals are necessary for estimating volatilization and dissipation through transport in the vapor phase. The low vapor pressures of the sulfonylurea herbicides have presented significant challenges in vapor pressure determination. We have used the Knudsen gas effusion method at elevated temperatures and extrapolated to 25 °C. Along with the Knudsen method, computer calculations using the Grain equation were also used to estimate vapor pressure. The gas saturation method with quantitation by high performance liquid chromatography (HPLC) provided an upper limit that confirmed the low vapor pressures obtained using Knudsen gas effusion and computer calculations. We report the best available experimental results for the vapor pressures of sulfonylurea herbicides. © 2000 Society of Chemical Industry  相似文献   

9.
10.
Since their introduction in the early 1980s, there have been a number of field studies conducted to assess the effects of sulfonylurea herbicides on non-target plants (i.e. plants not labeled for use). In these studies a wide variety of plant response assessment techniques have been used to measure effects on non-target plants. This paper examines the relationship of short-term plant response measurements to plant productivity measurements such as yield or quality. Whether short-term plant response measurements have a practical degree of accuracy and precision appropriate for hazard assessment on non-target plants from sulfonylureas is discussed. A comprehensive review of published literature and unpublished field studies of the effects of sulfonylureas on the yield and quality of non-target plant species is reported. When this information is coupled with exposure factors and environmental fate characteristics, the risks to non-target plants from sulfonylureas are similar to those from other herbicides used at higher application rates. © 1998 SCI.  相似文献   

11.
The behaviour and fate of chlorsulfuron in aqueous and soil systems were examined in laboratory studies. Aqueous hydrolysis was pH-dependent and followed pseudo-first-order degradation kinetics at 25°C, with faster hydrolysis occurring at pH 5 (half-life 24 days) than at either pH 7 or 9 (half-lives >365 days). Degradation occurred primarily by cleavage of the sulfonylurea bridge to form the major metabolites chlorobenzenesulfonamide (2-chlorobenzenesulfonamide) and triazine amine (4-methoxy-6-methyl-1,3,5-triazin-2-amine). This route is a major degradation pathway in water and soil systems. Aqueous photolysis (corrected for hydrolysis) proceeded much more slowly (half-life 198 days) than aqueous hydrolysis and is not expected to contribute significantly to overall degradation. Hydrolysis in soil thin-layer plates exposed to light (half-life 80 days), however, progressed at a much faster rate than in dark controls (half life 130 days), which suggests that a mechanism other than direct photolysis may have been operative. An aerobic soil metabolism study (25°C) in a Keyport silt loam soil (pH 6·4, 2·8% OM) showed that degradation was rapid (half-life 20 days). Dissipation in an anaerobic sediment/water system (initial pH of water phase 6·7, final pH 7·4) progressed much more slowly (half-life >365 days) than in aerobic soil systems. Major degradation products in aerobic soil included the chlorobenzenesulfonamide and triazine amine as in the aqueous hydrolysis study. Neither of these degradation products exhibited phytotoxicity to a variety of crop and weed species in a glasshouse experiment, and both exhibited an acute toxicological profile similar to that of chlorsulfuron in a battery of standard tests. Demethylation of the 4-methoxy group on the triazine moiety and subsequent cleavage of the triazine ring is another pathway found in both aqueous solution and soils, though different bonds on the triazine amine appear to be cleaved in the two systems. Hydroxylation of the benzenesulfonamide moiety is a minor degradation pathway found in soils. Two soils amended with 0·1 and 1·0 mg kg-1 chlorsulfuron showed slight stimulation of nitrification. The 1·0 mg kg-1 concentration of chlorsulfuron resulted in minor stimulation and inhibition of 14C-cellulose and 14C-protein degradation, respectively, in the same soils. Batch equilibrium adsorption studies conducted on four soils showed that adsorption was low in this system (Koc 13–54). Soil thin-layer chromatography of chlorsulfuron (Rf=0·55–0·86) and its major degradation products demonstrated that the chlorobenzenesulfonamide (Rf=0·34–0·68) had slightly less mobility and that the triazine amine (Rf=0·035–0·40) was much less mobile than chlorsulfuron. In an aged column leaching study, subsamples of a Fallsington sandy loam (pHwater 5·6, OM 1·4%) or a Flanagan silt loam (pHwater 6·4, OM 4·0%) were treated with chlorsulfuron, aged moist for 30 days in a glasshouse and then placed upon a prewet column of the same soil type prior to initiation of leaching. This treatment resulted in the retention of much more total radioactivity (including degradation products) than by a prewet column, where initiation of leaching began immediately after chlorsulfuron application, without aging (primarily chlorsulfuron parent). © 1998 SCI  相似文献   

12.
Summary. Adsorption and degradation rates of triasulfuron in 8 different soils were negatively correlated with soil pH and were generally lower in subsoils than in soils from the plough layer. The half-life at 20°C varied from 33 days in a top soil at pH 5·8 to 120 days in a subsoil at pH 7·4. Adsorption distribution coefficients in these two soils were 0·55 and 0·19, respectively. Movement and persistence of residues of chlorsulfuron, triasulfuron and metsulfuron-methyl were compared in a field experiment prepared in spring 1987. Triasulfuron was less mobile in the soil than the other two compounds. Residues of all three herbicides were largely confined to the upper 40–50 cm soil 148 days after application. With an initial dose of 32 g ha−1, residues in the surface soil layers were sufficient to affect growth of lettuce and sugar-beet sown approximately one year after application. Laboratory adsorption and degradation data were used with appropriate weather data in a computer model of herbicide transport in soil. The model gave good predictions of total soil residues during the first five months following application, and also predicted successfully the maximum depth of penetration of the herbicides into the soil during this period. However, more herbicide was retained close to the soil surface than was predicted by the model. The model predicted extensive movement of the herbicides in the soil during winter but did not predict that residues sufficient to affect crop growth could be present in the upper 15–20 cm soil after one year.  相似文献   

13.
用紫外线照射非荧光特性的砜嘧磺隆和氯磺隆,通过生成具有荧光特性的衍生物,分别研究了其在不同介质中的荧光特性及其影响因子,建立了测定土壤中砜嘧磺隆和氯磺隆残留的光化学荧光分析法(PCF)。结果表明:在2×10-3 mol/L、一定酸碱度(砜嘧磺隆pH 7、氯磺隆pH 12)的十六烷基三甲基氯化铵(CTAC)胶体分散体系中,紫外照射150 s是PCF法测定砜嘧磺 隆和氯磺隆残留的最佳条件,在此条件下砜嘧磺隆和氯磺隆的检出限(LOD)分别为0.7和0.6 μg/kg, 相对标准偏差(RSD)分别为1.7%和2.1%;在黄松田水稻土、黄红壤性水稻土和青紫泥田水稻土3种不同性质的土壤中,砜嘧磺隆和氯磺隆同时测定的平均回收率分别为99.0%±1.0%和98.7%±4.1%、97.6%±1.7%和97.0%±4.7%、96.7 %±2.3%和95.4%±5.5%;所建立的PCF法可有效、快速测定土壤中同时残留的微量砜嘧磺隆和氯磺隆。  相似文献   

14.
Degradation of triasulfuron in non-autoclaved and autoclaved soil incubated at different temperatures and moisture contents was evaluated in the laboratory using a maize root growth bioassay. Disappearance of triasulfuron was faster in non-autoclaved than in autoclaved soil, indicating the importance of microorganisms in the breakdown process. Degradation of the herbicide was faster at 30°C than at 10°C, with half-lives of 11–13 days at 30°C and 30–79 days at 10°C. Degradation of the herbicide was influenced more by temperature than by variation in soil moisture. Disappearance of the herbicide was rapid in the non-autoclaved soil at 30°C during the initial 30 days of incubation, but low levels of residues persisted for up to 90 days. A second application of the herbicide, to soil in which an initial dose of triasulfuron had degraded, disappeared at the same rate as herbicide added to previously untreated soil, indicating that there was no enhancement of degradation with repeated application of herbicide. Dégradation du triasulfuron dans le sol en conditions de laboratoire La dégradation du triasulfuron dans des sols non autoclavés et autoclavés, incubés à des températures et à des teneurs en humidité différentes, a étéévaluée au laboratoire en utilisant un bio essai sur la croissance d'une racine de maïs. La disparition du triasulfuron a été plus rapide en sol non autoclavé qu'en sol autoclavé, soulignant l'importance des microorganismes dans le processus de dégradation. La dégradation de l'herbicide a été plus rapide à 30°C qu'à 10°C avec des demi-vies respectives de 11–13 jours et de 30–79 jours. La dégradation de l'herbicide a été plus influencée par la température que par les variations d'humidité du sol. La disparition de l'herbicide a été rapide dans le sol non autoclavéà 30°C pendant les 30 premiers jours d'incubation, mais de faibles résidus persistaient au delà de 90 jours. Une seconde application d'herbicide sur un sol dans lequel une dose initiate de triasulfuron avait été dégradée, a disparu de la même façon qu'une dose appliquée sur un sol non traitd, montrant qu'il n'y avait pas d'augmentation de la dégradation à la suite d'une répétition d'application. Abbau von Triasulfuron im Boden unter Laborbedingungen Der Abbau von Triasulfuron in nicht sterilisiertem und sterilisiertem Boden bei verschiedener Temperatur und Bodenfeuchte wurde mit einem Maiswurzel-Wachstumstest untersucht. Die Menge des Triasulfurons nahm im nicht-sterilisierten Boden schneller ab als im sterilisierten, was auf mikrobiellen Abbau hinweist. Das Herbizid wurde bei 30 °C mit einer Halbwertszeit von 11 bis 13 Tagen schneller abgebaut als bei 10 °C mit einer von 30 bis 79 Tagen. Der Abbau wurde durch die Temperatur stärker beeinflußt als durch Änderung der Bodenfeuchte. Das Herbizid unterlag in den ersten 30 Tagen bei 30 °C im nichtsterilisierten Boden einem schnellen Abbau, doch geringe Rückstände wurden bis zu 90 Tagen gefunden. Bei einer zweiten Applikation des Herbizids auf Boden, in dem schon eine erste Dosis von Triasulfuron abgebaut worden war, nahm der Wirkstoff im selben Maße wie zuvor ab, so daß bei wiederholter Anwendung nicht mit einem verstärkten Abbau gerechnet werden kann.  相似文献   

15.
The measurement of the sulfonyl urea herbicides, chlorsulfuron and metsulfuron-methyl as pentafluorobenzyl derivatives has been investigated and the kinetics of reaction optimised. The reaction product was identified by mass spectroscopy as N,N-bis(pentafluorobenzyl)-2-chlorobenzenesulfonamide. Amounts of herbicide derivative as low as 0·1 pg per injection can be detected. Suitable conditions for both packed and capillary chromatography are given. Application of the method to residue determination is discussed and the method is shown to be suitable for residues in soil and water but less so for plant material.  相似文献   

16.
A Cyperus difformis L accession from Chonnam province, Korea was tested for resistance to the sulfonylurea herbicide, imazosulfuron. The accession was confirmed to be resistant (R) and was cross-resistant to other sulfonylurea herbicides, bensulfuron-methyl, cyclosulfamuron and pyrazosulfuron-ethyl, the pyrimidinyl thiobenzoate herbicide, bispyribac-sodium, and the imidazolinone herbicide imazapyr, but not to imazaquin. Multiple resistance was tested using twelve herbicides with target sites other than acetolactate synthase (ALS). The R biotype could be controlled by other herbicides with different modes of action such as butachlor, carfentrazone-ethyl, clomeprop, dithiopyr, esprocarb, mefenacet, oxadiazon, pretilachlor, pyrazolate and thiobencarb, applied to soil at recommended rates. Several sulfonylurea herbicide-based mixtures can control both the R and S biotypes of C difformis, except sulfonylurea plus dimepiperate, molinate or pyriftalid, and pyrazolate plus butachlor. Although mixtures of sulfonylurea herbicides might be more effective, they should be avoided and used only in special cases. In terms of in vitro ALS activity, the R biotype was 1139-, 3583-, 1482-, 416-, 5- and 9-fold more resistant to bensulfuron-methyl, cyclosulfamuron, imazosulfuron, pyrazosulfuron-ethyl, bispyribac-sodium and imazapyr, respectively, than the S biotype. The in vivo ALS activity of the R biotype was also less affected by the sulfonylurea herbicides, imazosulfuron and pyrazosulfuron-ethyl, than the S biotype. Results of in vitro and in vivo ALS assays indicated that the resistance mechanism of C difformis to ALS inhibitor herbicides was primarily due to an alteration in the target enzyme, ALS. Greenhouse experiments showed delayed flowering and reduced seed production of the R biotype, which could possibly result in reduced fitness. This unusual observation needs to be confirmed in field situations.  相似文献   

17.
The olive industry generates residues which can be applied as amendments to soils in their original form (olive cake) or after composting or vermicomposting processes. The addition, fresh or incubated, of these amendments to soil and of their different organic fractions was studied in relation to the sorption/desorption of three sulfonylurea herbicides, bensulfuron‐methyl, chlorsulfuron and prosulfuron. Herbicide sorption was low or very low, slightly promoted by the addition of the agricultural by‐products, especially olive cake, and mainly affected by pH of the soil solution, with the organic carbon content having no significant effect on herbicide retention. Desorption was only reduced when fresh olive cake was added. The incubation of soil and amendments for 3 months did not modify herbicide sorption, but made desorption reversible except for olive cake. The transformation of the organic matter of the amendments due to humification and maturity processes are likely to be responsible for this behaviour. Different organic fractions were removed to assess the influence of each fraction on sulfonylurea sorption. Only the removal of all studied organic fractions increased herbicide sorption, revealing the role of humin and mineral fractions in this process. Therefore, the use of organic amendments is not useful for reducing the risk of movement of ionisable molecules in soil.  相似文献   

18.
磺酰脲类除草剂是使用最广泛的除草剂之一。微生物降解是磺酰脲类除草剂在土壤中降解的重要方式。本文简要概述了土壤微生物对磺酰脲类除草剂的降解、降解机理及影响微生物降解磺酰脲类除草剂的因素。  相似文献   

19.
磺酰脲类除草剂的降解机制及代谢产物的研究进展   总被引:6,自引:0,他引:6  
磺酰脲类除草剂是一类高效,低毒和高选择性的除草剂,该类除草剂能有效防除阔叶杂草,其中有些品种对禾本科杂草也有一定的抑制作用。但同时因其用量低、对哺乳动物低毒以及独特的除草活性等特点而得到广泛应用。因此,了解磺酰脲类除草剂在土壤中的环境行为及归趋对于其科学合理使用、防止作物药害和保护农业生态环境具有非常重要的意义。根据笔者对磺酰脲类除草剂的深入研究,并总结归纳国内外的相关文献报道,对磺酰脲类除草剂的降解机制及其代谢产物的研究进行了综述,最后展望了磺酰脲类除草剂未来的发展趋势。  相似文献   

20.
The hydrolysis of triasulfuron, metsulfuron‐methyl and chlorsulfuron in aqueous buffer solutions and in soil suspensions at pH values ranging from 5.2 to 11.2 was investigated. Hydrolysis of all three compounds in both aqueous buffer and soil suspensions was highly pH‐sensitive. The rate of hydrolysis was much faster in the acidic pH range (5.2–6.2) than under neutral and moderately alkaline conditions (8.2–9.4), but it increased rapidly as the pH exceeded 10.2. All three compounds degraded faster at pH 5.2 than at pH 11.2. Hydrolysis rates of all three compounds could be described well with pseudo‐first‐order kinetics. There were no significant differences (P = 0.05) in the rate constants (k, day−1) of the three compounds in soil suspensions from those in buffer solutions within the pH ranges studied. A functional relationship based on the propensity of nonionic and anionic species of the herbicides to hydrolyse was used to describe the dependence of the ‘rate constant’ on pH. The hydrolysis involving attack by neutral water was at least 100‐fold faster when the sulfonylurea herbicides were undissociated (acidic conditions) than when they were present as the anion at near neutral pH. In aqueous buffer solution at pH > 11, a prominent degradation pathway involved O‐demethylation of metsulfuron‐methyl to yield a highly polar degradate, and hydrolytic opening of the triazine ring. It is concluded that these herbicides are not likely to degrade substantially through hydrolysis in most agricultural alkaline soils. © 2000 Society of Chemical Industry  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