环糊精和腐殖酸对手性除草剂禾草灵的水生毒理和生物有效性影响研究
详细信息    本馆镜像全文|  推荐本文 |  |   获取CNKI官网全文
摘要
农业生产中手性农药的广泛使用,使其不可避免地进入到附近的环境系统中。手性农药中对映体在生物过程中存在较大的差异性,使之表现出不同的生态毒理,为此,手性农药的使用与环境安全成为一个新的关注中心。禾草灵是一种具有代表性的手性芳氧丙酸类除草剂,但仍以外消旋的形式上市。本文首先合成了禾草灵和禾草灵酸的对映体化合物,研究了它们在藻液中的对映体差异性降解和毒理。并初步探讨了产生对映体差异性的机制。此外,还研究了农药慢性诱导作用对藻细胞的生长、通透以及降解性能的影响。
     禾草灵快速地吸附在藻细胞表面上,催化水解生成毒性较低的禾草灵酸,禾草灵和禾草灵酸进一步降解生成毒性更强的酚类代谢产物,表明禾草灵的这种降解没有降低其环境风险。禾草灵及其转化产物的毒性差异与它们的理化性质有关。测定农药在环境中转化产物的生态毒理和环境归宿,有助于完整准确地评价农药化合物带来的原生和次生的环境风险。禾草灵和禾草灵酸的毒性以及在藻液中的降解表现出了对映体差异性。蛋白核小球藻选择性吸收农药的R型对映体,普通小球藻和斜生栅藻表现出相反的对映体选择性。这种对映体选择性与藻细胞的通透性能具有一定的相关性,与化合物的种类和受试生物的种类有关,主要是由藻细胞的主动吸收造成的。
     通过多种技术手段,证实了环糊精与禾草灵和禾草灵酸之间存在包合作用。环糊精的包合作用增加了禾草灵的溶解度和稳定性,加快了农药的溶解速率,减少农药在土壤中的吸附。环糊精对水藻的生长影响不大,降低了藻细胞的被动吸收能力,减弱了藻细胞表面的疏水性,降低了农药的降解速率。实验结果表明环糊精的存在,可以使农药化合物更有效地表现出毒理效应,提高农药的药效。并且,环糊精改变了手性农药化合物在藻液中降解的对映体差异性。环糊精减少了藻细胞对禾草灵和禾草灵酸的吸收,但没有抑制它们的生物活性,显示了环糊精有望作为一种助剂,加快它们在植株中的传导,提高药效,降低使用量,从而降低使用过程中可能产生的生态风险。
     腐殖酸是天然水体中可溶性有机质中的主要成分,它可能通过氢键作用和疏水作用吸附在藻细胞上,增加了藻细胞的通透性,减弱了藻细胞表面疏水性能,
Widespread application of chiral pesticides results in their occurrence in adjacent environmenat systems such as atmosphere, soil, and surface water. In general, the enantiomers of chiral pesticides are different in some biological processes and thus pose enantioselective ecotoxicity to nontarget biology. Consequently, use and environmental safety of chiral pesticides have been a new point of scientific researches. In the present study, the enantiomers of diclofop methyl (DM) and diclofop (DC) prepared by ourselves were selected to investigate their enantioselective degradation and ecotoxicology in algal cell cultures.DM is rapidly adsorbed to algal cells and is catalytically hydrolyzed to less toxic DC which shows phytotoxicicty. They further metabolize to more toxic phenol metabolites, indicating an increase of environmental burden on DM degradation. DM and its transformation products have different toxicity to algae, associated with their physical chemical properties. Assessing the ecotoxicology and environmental fate of pesticide transformation products is necessary to accurately evaluate environmental risk of pesticides. Toxicology and degradation of chiral herbicides DM and DC is enantioselective, dependent of algae and chemicals. Chlorella pyrenoidosa preferentially uptakes or degrades (R)-enantiomers of DM and DC, however, Chlorella vulgaris and Scenedesmus obliquus have reverse enantioselelctivity with (S)-enantiomer degrading more rapidly than the R enantiomer. The enantioslelctivity of three algae is correlated with their respective cell permeability and is mainly ascribed to cell facilitated transport.Phase-solubility experiments and other methods or technologies demonstrate inclusion complexes between cyclodextrins (CDs) and DM/DC. The inclusion complexation of partially methylated cyclodextrin (MCD) enhances water solubility and stability of DM, improves solution rate of DM, and decreases adsorption of DM to soils, indicating an increase in pesticide bioavailability. MCD does not inhibit algal growth, and decreases passive uptake of algae and cell surface hydrophobicity, but reduces degradation rate of DM and DC in algal cell suspension solutions. The results
    show that less pesticide have the same biological acitivity when MCD exists. Besides, MCD changes enantioselective degradation of chiral herbicides in alga cultures. Thus, MCD probably increases herbicidal acitivity of DM and DC as additives, reduces application rates of pesticides, and decrease ecological risk of pesticides caused by overdose.Humic acid (HA) is a major component of dissolved organic matter in natural aquatic systems. HA binds DM and DC, and reduces the free pesticide compounds. In the study, however, HA, adsorbed to algal cells, increases cell permeability of three algae, decreases hydrophobicity of algal cell surface, and changes growth character of three algae in different degrees. Data for degradation of DM and DC and algal biomass in Chlorella vulgaris cultures shows that HA increases degradation or uptake of these pesticides and enhance their bioavailability. These suggest that cell permeability induced by HA plays more important role in their bioavailability than the binding interaction between HA and pesticides. At the same time, HA change degradation of DC and its two enantiomers, dependent of HA concentration, algae, and types and concentration of pestcicide.DC at low dose or concentration stimulates the growth of algae. On the algae exposed to 0.10 mg L"1 of DCs enantiomers for two months, however, the induction by DC results in some changes of algal growth and cell permeability. Induction by two enantiomers of DC stimulates the growth of Chlorella pyrenoidosa, inhibits the growth of Chlorella vulgaris, and have no observable effect on Scenedesmus obliquus. In general, induction increases the uptake of the same substrate by algal cells, for example, algal cells induced by R-DC uptake more R-DC than CK. Moreover, induction by two enantiomers reduces or even reverses the enantioselective degradation of chiral pesticides in alga cultures.
引文
1. Ali I., Aboul-Enein H. Y., Ghanem A. 2005. Enantioselective toxicity and carcinogenesis. Curr. Pharmaceu. Anal., 1,109-125.
    2. Ali I., Gupta V. K., Aboul-Enein H. Y. 2003. Chirality: a challenge for the environmental scientists. Curr. Sci., 84,152-156.
    3. Ashby J., Brady A., Elcombe C. R., Elliot B. M., Ishamel J., Odum J., Tugwood J. D., Kettle S., Purchase I. F. H.1994. An examination of the correlation between hepatic peroxisome proliferation and hepato carcinogenesis. Hum.Exp. Toxicol., 13 (Suppl. 2), S7-S18.
    4. Bentley P., Calder I., Elcombe C. R., Grasso P., Stringer D., Wiegand H. J. 1993. Hepatic peroxisome proliferationin rodents and its significance for humans. Fed. Chem. Toxicol., 31, 857-907.
    5. Bettoni G., Loiodice F., Tortorella V., Conte-Camarino D., Mambrini M., Ferrannini E., Bryant S. H. 1987. Stereospecificity of the chloride ion channel. The action of chiral clofibric acid analogues. J Med. Chem., 30, 1267-1270.
    6. Bewick D. W. 1986. Stereochemistry of fluzifop-butyl transformation in soil. Pestic. Sci., 1986, 17, 349-356.
    7. Beynen A. C., Geelen M. J. H. 1982. Short term inhibition of fatty acid biosynthesis in isolated hepatocytes by mono-aromatic compounds. Toxicol, 24, 184-197.
    8. Bucheli T. D., Muller S. R., Heberle S., Schwarzenbach R. P. 1998a. Occurrence and behavior of pesticides in rainwater, roof runoff, and artificial stormwater infiltration. Environ. Sci. Technol., 32, 3457-3464.
    9. Bucheli T. D., Muller S. R., Voegelin A., Schwarzenbach R. P. 1998b. Bituminous roof sealing membranes as major sources of the herbicide (R,S)-mecoprop in roof runoff waters: potential contamination of groundwater and surface waters. Environ. Sci. Technol., 32, 3465-3471.
    10. Burton J. D., Gronwald J. W., Keith R. A., Somers D. A., Gengenbach B. G., Wyse D. A. 1991. Kinetics of inhibition of acetyl coenzyme A carboxylase by sethoxydim and haloxyfop. Pestic. Biochem. Physiol., 39, 100-109.
    11. Buser H. R., Muller M. D. 1998. Occurrence and transformation reactions of chiral and achiral phenoxyalkanoic acid herbicides in lakes and rivers in Switzerland. Enviorn. Sci. Technol., 32, 626-633.
    12. Carr J. E., Davies L. G, Cobb A. H., Pallett K. E., 1986. Uptake, translocation and metabolism of fluazifop-butyl in Setraria viridis. Ann. Appl. Biol., 108,115-123.
    13. Carson M. 1987. Ribbon Models of Macromolecules. J. Mol. Graphics 5,103-106.
    14. Cartwright D. 1989. The synthesis, stability and biological acitivity of the enantiomers of pyridylxoyphenoxypropionates, in Proceedings of 1989 British Crop Protection Conference on Weeds, BCPC Publications, Surrey, pp. 707-716.
    15. Cohen, AJ., Grasso, P., 1981. Review of the hepatic response to hypolipidaemic drags in rodents and assessment of its toxicological significance to man. Food Chem. Toxicol. 19,585-605.
    16. de Fuentes I. E., Viseras C. A., Ubiali D., Terreni M, Alcantara A. R. 2001. Different phyllosilicates as supports for lipase immobilization. Journal of Molecular Catalysis B: Enzymatic, 11, 657-663.
    17. DeLorenzo M. E., Scott G. I., Ross P. E. 2001. Toxicity of pesticides to aquatic microorganisms: A review. Environ. Toxicol. Chem., 20, 84-98.
    18. Dicks J. W., Slater J. W., Bewick D. W. 1985. PP005-the R-enantiomer of fluazifop butyl, in Proceedings of 1985 British Crop Protection Conference on Weeds, BCPC Publications, Surrey, pp. 271-280.
    19. Donaldson D., Kiely T., Grube A. 2002. 1998 and 1999 market estimates. In: EPA (ed) Pesticides industry sales and usage.
    20. Easson L. H., Stedman E. 1933. Studies on the relationship between chemical constitution and physiological action. V. Molecular dissymmetry and physiological activity. Biochem. J.,27,1257-1266.
    21. Edwards T, Cole D. J. 1996. Glutathione transferases in wheat (Triticum) species with activity toward fenoxaprop-ethyl and other herbicides. Pestic. Biochem. Physiol., 54, 96-104.
    22. Eliel E. L., Wilen S. H. Stereochemistry of Organic Compounds;Wiley & Sons: New York,1993;p 208.
    23. Eliyahu D., Applebaum S., Rafaeli A. 2003. Moth sex-pheromone biosynthesis is inhibitied by the herbicide diclofop. Pestic. Biochem. Physiol., 77,75-81.
    24. Fang Z. H., Wen Y. Z., Liu W. P. 2005. Influence of kaolinite on chiral hydrolysis of methyl dichlorprop enantiomers. J. Zhejiang Univ. Sci., 6B, 1028-1032.
    25. Felding G 1995. Leaching of phenoxyalkanoic acid herbicides from farmland. Sci. Total Environ., 168,11-18.
    26. Fernandez-Lorente G, Terreni M., Mateo C, Bastida A., Fernandez-Lafuente R., Dalmases P., Huguet J., Guisan J. M. 2001. Modulation of lipase properties in macroaqueous systems by controlled enzyme immobilization: enantioselective hydrolysis of a chiral ester by immobilized Pseudomonas lipase. Enzyme and Microbial Technology, 28, 389-396.
    27. Garrison A. W. 2006. Probing the enantioselectivity of chiral pesticides. Enviorn. Sci. Technol., 40,16-23.
    28. Garrison A. W., Schmitt P., Martens D., Kettrup A. 1996. Enantiomeric selectivity in the environmental degradation of dichlorprop as determined by high-performance capillary electrophoresis. Environ. Sci. Technol., 30, 2249-2455.
    29. Gerwick B. C, Jackson L. A., Handly J. Gray N. R., Russell J. W. 1988. Preemergence and postemergence activities of the R and S enantiomers of haloxyfop. Weed Sci., 36, 453-456.
    30. Gintautas P. A., Daniel S. R., Macalady D. L. 1992. Phenoxyalkanoic acid herbicides in municipal landfill leachates. Environ. Sci. Technol., 26, 517-521.
    31. Gorbach S. G., Kuenzler K., Asshauer J. 1977. On the metabolism of HOE 23408 OH in wheat. J. Agric. Food Chem., 25, 507-511.
    32. Granzer E., Nahm H. 1973. HCG 004, a new highly potent hypolipidemic drug. Arzneim Forsch, 23, 1353-1354
    33. Grasso P., Hinton R. H. 1991. Evidence for and possible mechanisms of non-genotoxic carcinogenesis in rodent liver. Mut. Res., 248, 271-290.
    34. Hansen W. H., Quaife M. L., Habermann R. T., Fitzhug, O. G. 1971. Chronic toxicity of 2,4-dichlorophenoylacetic acid in rats and dogs. Toxicol. Appl. Pharmacol., 20, 122-129.
    35. Hayes T. B., Collins A., Lee M., Mendoza M., Noriega N., Stuart A. A., Vonk A. 2002. Hermaphroditic, demasculinized frogs after exposure to the herbicide atrazine at low ecologically relevant doses. Proc. Natl. Acad. Sci. USA., 99, 5476-5480.
    36. Hegg E. L., Que L. Jr. 1997. The 2-His-1-carboxylate facial triad—an emerging structural motif in mononuclear non-heme iron(Ⅱ) enzymes. Eur J Biochem, 250, 625-629
    37. Hemmen C., Konradt M.1991. Gesunde Pfanz, 143, 189.
    38. Hendley P., Dicks J. W., Monace T. J., Slyfield S. M., Tummon O. J., Barrett J. C. 1985. Translocation and metabolism of pyridinyloxyphenoxypropanoate herbicides in rhizomatous quackgrass (Agropyron repens). Weed Sci., 33, 11-24.
    39. Heron G., Christensen T. H. 1992. Degradation of the herbicide mecoprop in an aerobic aquifer determined by laboratory batch studies. Chemosphere, 24, 547-557.
    40. Hofer M., Hoggett J. G. 1981. Transport across biological membranes. Pitman Publishing Ltd., Boston, Mass.
    41. Hofer M., Kotyk A. 1968. Tight coupling of monosaccharide transport and metabolism in Rhodotorula gracilis. Folia Microbiol (Praha)., 13, 197-204.
    42. Huff H. P., Bottner B., Ebert E., Langeloddeke P. 1989. HOE 046360-the optical active isomer of fenoxaprop-ethyl for broad spectrum grass-weed control in dicotyledonous crops, in Proceedings of 1989 British Crop Protection Conference on Weeds, BCPC Publications, Surrey,, pp. 717-722.
    43. Jacobson A., Shimabukuro R. H., McMichael C. 1985. Response of wheat and oat seedling to root applied diclofop-methyl and 2,4-dichlorophenoxyacetic acid. Pestic. Biochem. Physiol., 24, 61-67.
    44. Jeffcoat B., Harris W. N. 1975. Selectivity and mode of action of flampropisopropyl (±)-2-[N-(3-chloro-4-flurophenyl)benzamido-]propionate in the control of Avena fatua in barley. Pestic. Sci., 6, 283-296.
    45. Kohler H. P. E., Angst W., Giger W. 1997. Environmental fate of chiral pollutants-the necessary of considering sterochemistry. Chimia, 51,947-950.
    46. Kohler H. P. E., Nickel K., Zipper C. 2000. Effect of chirality on the microbial degradation and the environmental fate of chiral pollutants. In: Schink B (ed) Advances in microbial ecology. Kluwer, New York, pp 201-231
    47. Kuranyi-Covacs I., Rudali G., Imbert J. 1976. Bioassay of 2,4,5-trichloro-phenoxyacetic acid for carcinogenicity in mice. Br. J. Cancer 33,626-633.
    48. Kurihara N., Miyamoto J. 1998. Chirality in agrochemicals. John Wiley & Sons Ltd., pp. 186.
    49. Lewis D. L., Garrison A. W., Wommack K. E., Whittemore A., Steudler P., Melille J. 1999. Influence of environmental changes on degradation of chiral pollutants in soils. Nature, 401, 898-901.
    50. Liu W. P., Gan J., Shneak D., Jury W. A. 2005. Enantioselectivity in environmental safety of current chiral insecticides. Proc. Natl. Acad. Sci. USA, 102, 701-706.
    51. Loos M. A. In Degradation of Herbicides (Kearney E C.;Kaufmann D. D. eds), Marcel Dekker, Inc., New York, 1969, p. 1.
    52. Ludwig P., Gunkel W., Huhnerfuss H. 1992a. Chromatographic separation of the enantiomers of marine pollutants. Part 5: enantioselective degradation of phenoxycarboxylic acid herbicides by marine microorganisms. Chemosphere. 24. 1423-1429.
    53. Lyngkide J., Christensen T. H. 1992. Fate of organic contaminants in the redox zone of a landfill leachate pollution plume. J. Contam. Hydrol., 10, 291-307.
    54. McBlain W. A. 1987. The Levo Enantiomer of o, p'-DDT Inhibits the Binding of 17 Beta-estradiol to the Estrogen Receptor. Life Sci, 40, 215-221.
    55. McFadden J. J., Frear D. S., Mansanger E. R. 1989. Aryl hydroxylation of diclofop by a cytochrome P-450 dependent monooxygenase from wheat. Pestic. Biochem. Physiol., 34, 92-100.
    56. Mesecar A. D., Koshland D. E. Jr 2000. A new model for protein stereospecificity. Nature, 403, 614-615.
    57. Morvan Y., Panis P. E., Strathmann S. 1991. Fluoroglycofen-ethyl plus dichlorprop-P a new herbicide for post-emergence broadleaved weed control in cereals. Mededelingen Van De Faculteit Landbouwwetenschappen Rijksuniversiteit Gent 56(3 Part A), 711-717.
    58. Muller M. D., Buser H. R. 1997. Conversion reactions of various phenoxyalkanoic acid herbicides in soil. 1. Enantiomerization and enantioselective degradation of the chiral 2-phenoxypropionic acid herbicides. Environ. Sci. Technol., 31, 1953-1959.
    59. Muller T. A., Byrde S., Fleischmann T., van der Meer J.R., Kohler H. P. E. 2004. Genetic analysis chiral phenoxyalkanoic acid herbicides degradation in Sphingomonas herbicidovorans MH. Appl. Environ. Microbiol., 70, 6066-6075.
    60. Muller T. A., Kohler H. P. E. 2004. Chirality of pollutants-effects on metabolism and fate. Appl. Microbiol. Biotechnol., 64, 300-316.
    61. Mustaers J. H. G. M., Kooreaman H. J. 1991. Preparation of optically pure 2-aryl and 2-aryloxy-propionates by selective enzymic hydrolysis. Rec. Tray. Chim. Patsbas, 110, 185-188.
    62. Nestler H. J., Bieringer H. 1980. Synthese und herbizide Wirksamkeit der D- and L--enantiomeren des 2-[4-(2.4-Dichlorphenoxy)-phenoxy]-propionsaure-methyl esters. Z Naturforsch, 85b, 366-371.
    63. Nicholls A., Sharp K. A., Honig B.1991. Protein folding and association: insights from the interfacial and thermodynamic properties of hydrocarbons. Proteins, 11,281-296.
    64. Nickel K., Suter M. J. F., Kohler H. P. E. 1997. Involvement of two ketoglutarate-dependent dioxygenases in enantioselective degradation of (R)- and (S)-mecoprop by Sphingomonas herbicidovorans MH. J. Bacteriol., 179, 6674-6679.
    65. Ogston A. G. 1948. Interpretation of experiments on metabolic processes, using isotopic tracer elements. Nature, 162, 963-963.
    66. Palomo J. M., Fernandez-Lorente G., Mateo C., Ortiz C., Fernandez-Lafuente R., Guisan J. M. 2002. Modulation of the enantioselectivity of lipases via controlled immobilization and medium engineering: hydrolytic resolution of mandelic acid esters. Enzyme and Microbial Technology, 31, 775-783.
    67. Palut D., Kostka G, Wiadrowska B., Bankowski R. 2002. Effect of diclofop on the activity of some drug-metabolizing enzymes in the liver of male Wistar rats. Rocz. Panstw. Zakl. Hig.,53,1-9.
    68. Palut D., Ludwicki J. K., Kostka G, Kopec-Szlezak J. 2001. Studies of early hepatocellular proliferation and peroxisomal proliferation in Wistar rats treated with herbicide diclofop. Toxicol., 158,119-126.
    69. Palut, D., 1997. Peroxisome proliferation and hepatocarcinogenesis. Rocz. PZH 48,1-11.
    70. Peterson H. G., Boutin C, Martin P. A., Freemark K. E., Ruecker N. J., Moody M. J. 1994. Aquatic phyto-toxicity of 23 pesticides applied at expected environmental concentrations. Aquatic Toxicology, 28, 275-292.
    71. Petit E, Goff L. G, Cravede J. P., Valotaire Y., Pakdel F. 1997. Two complementary bioassays for screening the estrogenic potency of xenobiotics: Recombinant yeast for trout estrogen and trout hepatocyte cultures. J. Mol. Endocrinol., 19, 321-335.
    72. Ralph S. A., D'Ombrain M. C, McFadden G. I. 2001. The apicoplast as an antimalarial drug target. Drug Res Updates, 4,145-151.
    73. Ratterman D. M., Balke N. E. 1989. Diclofop-methyl increases the proton permeability ofisolated oat-root tonoplasts. Plant Physiol., 91, 756-765.
    74. Reddy J. K., Chu R., 1996. Peroxisome proliferators induced pleiotropic responses: pursuit of a phenomenoa Ann. NY Acad. Sci. 804, 176-201.
    75. Reddy J. K., Lawani N. D. 1983. Carcinogenesis by hepatic peroxisome proliferators: Evaluation of the risk of hypolipidemic drugs and industrial plasticizers to humans.CRC Crit.Rev. Toxicol., 12,1-58.
    76. Rendina A. R., Felts J., Beaudoin J. D., Craig-Kinnard A. D., Look L. L., Paraskos S. L., Hagenah J. A. 1988. Kinetic characterisation, stereoselectivity, and species selectivity of the inhibition of plant acetyl-CoA carboxylase by the aryloxyphenoxypropionic acid grass herbicides. Arch Biochem. Biophys., 265, 219-225.
    77. Roberts R. A., Nebert D. W., Hickman J. A., Richburg J. H., Goldworthy T. L. 1997. Perturbation of mitosis/apoptosis balance, a fundamental mechanism in toxicology. Fundament. Appl. Toxicol., 38, 107-115.
    78. Romero E., Matallo M. B., Pena A., Sanchez-Rasero F., Schmitt-Kopplin P., Dios G. 2001. Dissipation of racemic mecoprop and dichlorprop and their pure R-enantiomers in three calcareous soils with and without peat addition. Environ. Pollut., 111, 209-215.
    79. Scheidleder A., Grath J., Winkler G, Stark U., Koreimann C, Gmeiner C, Nixon S., Casillas J., Gravesen P., Leonard J., Elvira M. 1999. Groundwater quality and quantity in Europe. European Environment Agency, Copenhagen
    80. Schneiderheinze J. M., Armstrong D. W., Berthod A. 1999. Plant and soil enantioselective biodegradation of racemic phenoxyalkanoic herbicides. Chirality, 11, 330-337.
    81. Shaner D. 2003. Herbicide safety relative to common targets in plants and mammals. Pest Manag. Sci., 60,17-24.
    82. Shimabukuro M. A., Shimabukuro R. H., Walsh W. C. 1982. The antagonism of IAA-induced hydrogen ion extrusion and coleoptile growth by diclofop-methyl. Physiol. Plant., 56,
     444-452.
    83. Shimabukuro R. H. 1990. Selectivity and mode of action of the post emergence herbicide diclofop-methyl. Plant Growth Reg. Soc. Am., 18, 37-54.
    84. Shimabukuro R. H., Hoffer B. L. 1989. Metabolism of diclofopmethyl in annual ryegrass, Lolium rigidum, and its relationship to recovery of membrane potential. Weed Sci. SocAm. Abstr., 29, 78
    85. Shimabukuro R. H., Hoffer B. L. 1995. Enantiomers of diclofop-methyl and their role in herbicide mechanism of action. Pestic. Biochem. Physiol., 51, 68-82.
    86. Shimabukuro R. H., Walsh W. C, Jacobson A. 1987. Aryl-O-glucoside of diclofop: a detoxification product in wheat shoots and wild oat suspension culture. J. Agric. Food Chem.,35,393-397.
    87. Spindler F., Fruh T. Chiral Acylanilides and Chiral Triazole-Related Fungicides. In Kurihara, N., Miyamoto, J., Eds;Chirality in Agrochemicals;Wiley & Sons: Chichester, UK, 1998, pp143-145.
    88. Stuart D. A., McCall C. M. 1992. Induction of Somatic Embryogenesis Using Side Chain and Ring Modified Forms of Phenoxy Acid Growth Regulators. Plant Physiol., 99,111-118.
    89. Tal A., Romano M. L., Stephenson G R., Schwan A. L., Hall J. C. 1993. Glutathione conjugation: a detoxification pathway for fenoxaprop-ethyl in barley, crabgrass, oat and wheat. Pestic. Biochem. Physiol., 46,190-199.
    90. Tomlin C. 1994. The pesticide manual: a world compendium, 10th edn, BCPC Publications,Bershire.
    91. Twardowska I. 2004. Ecotoxicology, environmental safety, and sustainable development- challenges of the third millennium. Ecotoxicol. Environ. Saf., 58, 3-6.
    92. Vencill W. K. Herbicide handbook, 8th edn, Weed Science Society of America, Lawrence, KS, USA, 493 pp (2002).
    93. Westendorf A., Benndorf D., Muller R. H., Babel W. 2002. The two enantio-specific dichlorprop/a-ketoglutarate-dioxygenases from Delftia acidovorans MC1-protein and sequence date of RdpA and SdpA. Microbiol. Res., 157,1-6.
    94. Westendorf A., Muller R. H., Babel W. 2003. Purification and characterisation of the enantiospecific dioxygenases from Delftia acidovorans MCI initiating the degradation of phenoxypropionate and phenoxyacetate herbicides. Acta Biotechnol., 23, 3-17,
    95. Williams A. 1996. Opportunities for chiral agrochemicals. Pestic. Sci., 46, 3-9.
    96. Williams G, M., Harrison I., Carlick C. A., Crowley O. 2003. Changes in enantiomeric fraction as evidence of natural attenuation of mecoprop in a limestone aquifer. J. Contam. Hydrol., 64, 253-267.
    97. Wilson R. J. M. 2002. Progress with parasite plastids. J. Mol. Biol., 319, 257-274.
    98. Wink O., Luley U. 1988. Enantioselective transformation of the herbicides diclofop-methyl and fenoxaprop-ethyl in soil. Pestic. Sci., 22, 31-40.
    99. Worthing C. R., Hance R. J. 1991. The pesticide manual—a world compendium, 9th edn. The British Crop Council, Farnham
    100. Wright J. P., Shimabukuro R. H. 1987. Effects of diclofop and diclofop-methyl on the membrane potentials of wheat and oat coleoptiles. Plant Physiol., 85, 188-193.
    101. Zhang H. L., Tweel B., Tong L. 2004. Molecular basis for the inhibition of the carboxytransferase domain of acetyl-coenzyme-A carboxylase by haloxyfop and diclofop.
     Proc. Natl. Acad. Sci. USA, 101, 5910-5915.
    102. Zhang H. L., Yang Z. R., Shen Y., Tong L. 2003. Crystal structure of the carboxytransferase domain of acetyl-coenzyme-A carboxylase. Science, 299, 2064-2067.
    103. Zimdahl R. L. 2002. My view. Weed Sci., 50, 687.
    104. Zipper C. 1998. Microbial degradation and environmental fate of chiral phenoxyalkanoic acid herbicides. Ph.D. thesis, Swiss Federal Institute of Technology, Zurich.
    105. Zipper C., Bolliger C., Fleischmann T., Suter M. J. F., Angst W., Muller M. D., Kohler H. P. E. 1999. Fate of the herbicides mecoprop, as determined by laboratory batch studies and enantiomerspecific analysis. Biodegradation, 10, 271-278.
    106. Zipper C., Bunk M., Zehnder A. J. B., Kohler H. P. E. 1998a. Enantioselective uptake and degradation of the chiral herbicide Sphingomonas herbicidovorans MH. J Bacteriol., 180, 3368-3374.
    107. Zipper C., Nickel K., Angst W., Kohler H. P. E. 1996. Complete microbial degradation of both enantiomers of the chiral herbicide mecoprop [(RS)-2- (4-chloro-2-methylphenoxy) propionic acid] in an enantioselective manner by Sphingomonas herbicidovorans sp. nov. Appl. Environ. Microbiol., 62, 4318-4322.
    108. Zipper C., Suter M. J. F., Haderlein S. B., Gruhl M., Kohler H. P. E. 1998b. Changes in the enantiomeric ratio of (R)- to (S)-mecoprop indicate in situ biodegradation of this chiral herbicide in a polluted aquifer. Environ. Sci. Technol., 32, 2070-2076.
    109.方兆华,2005a.脂肪酶与手性苯氧丙酸类除草剂的对映选择性相互作用及其影响因素.PH D. Thesis,浙江大学,杭州.
    110.方兆华,刘维屏,周珊珊,文岳中.2005b.硅藻土吸附增强的(RS)-2,4-DP对映体选择性酶促水解.土壤学报 42,301-305.
    111.高贵,韩四平,王智,翁梁,王柏婧,冯雁,曹淑桂.2002.硅藻土固定化脂肪酶.吉林大学学报(理学版),40,324-326.
    112.蒋木庚,杨红,杨春龙,王鸣华.2001.21世纪手性农药发展展望.世界农药,23,14-41.
    113.刘军民,许建和,刘幽燕,许学书,袁勤生.1998.混合溶剂系统中固定化脂肪酶对酮基布洛芬的催化酯化反应.高等学校化学学报,19,1959-1963.
    114.马云,徐超,陈胜文,文岳中,刘维屏.2005.淤泥优势菌对手性农药2,4-滴丙酸甲酯的对映体选择性降解.环境科学,26,152-155.
    115.任康太,李永红,杨华铮.1999.芳氧苯氧丙酸酯和环己二酮类除草剂的作用机制.农药,38,1-4.
    116.王鸣华,章维华,蒋木庚.野燕麦除草剂的研究进展.农药,2003,42(8):6-7,5.
    117.文辉,边庆花,王敏.2004.光学纯2-氯丙酸及其酯的合成与在手性农药中的应用.农药,43,363-366.

© 2004-2018 中国地质图书馆版权所有 京ICP备05064691号 京公网安备11010802017129号

地址:北京市海淀区学院路29号 邮编:100083

电话:办公室:(+86 10)66554848;文献借阅、咨询服务、科技查新:66554700