Hostname: page-component-76fb5796d-wq484 Total loading time: 0 Render date: 2024-04-26T19:11:34.182Z Has data issue: false hasContentIssue false

Pesticide metabolism in plants and microorganisms

Published online by Cambridge University Press:  20 January 2017

Laura L. Van Eerd
Affiliation:
Department of Environmental Biology, University of Guelph, Guelph, ON, Canada N1G 2W1
Robert E. Hoagland
Affiliation:
Southern Weed Science Research Unit, United States Department of Agriculture, Agricultural Research Service, Stoneville, MS 38776
Robert M. Zablotowicz
Affiliation:
Southern Weed Science Research Unit, United States Department of Agriculture, Agricultural Research Service, Stoneville, MS 38776

Abstract

Understanding pesticide metabolism in plants and microorganisms is necessary for pesticide development, for safe and efficient use, as well as for developing pesticide bioremediation strategies for contaminated soil and water. Pesticide biotransformation may occur via multistep processes known as metabolism or cometabolism. Cometabolism is the biotransformation of an organic compound that is not used as an energy source or as a constitutive element of the organism. Individual reactions of degradation–detoxification pathways include oxidation, reduction, hydrolysis, and conjugation. Metabolic pathway diversity depends on the chemical structure of the xenobiotic compound, the organism, environmental conditions, metabolic factors, and the regulating expression of these biochemical pathways. Knowledge of these enzymatic processes, especially concepts related to pesticide mechanism of action, resistance, selectivity, tolerance, and environmental fate, has advanced our understanding of pesticide science, and of plant and microbial biochemistry and physiology. There are some fundamental similarities and differences between plant and microbial pesticide metabolism. In this review, directed to researchers in weed science, we present concepts that were discussed at a symposium of the American Chemical Society (ACS) in 1999 and in the subsequent book Pesticide Biotransformation in Plants and Microorganism: Similarities and Divergences, edited by J. C. Hall, R. E. Hoagland, and R. M. Zablotowicz, and published by Oxford University Press, 2001.

Type
Review Article
Copyright
Copyright © Weed Science Society of America 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Literature Cited

Alexander, M. 1999. Biodegradation and Bioremediation. 2nd ed. San Diego, CA: Academic. 453 p.Google Scholar
Anderson, M. P. and Gronwald, J. W. 1991. Atrazine resistance in a velvetleaf (Abutilon theophrasti) biotype due to enhanced glutathione S-transferase activity. Plant Physiol. 96:104110.Google Scholar
Andreae, W. A. and Good, N. E. 1957. Studies on 3-indoleacetic acid metabolism. IV. Conjugation with aspartic acid and ammonia as processes in the metabolism of carboxylic acids. Plant Physiol. 32:566572.Google ScholarPubMed
Andreoni, V., Colombo, M., Gennari, M., Negre, M., and Ambosoli, R. 1994. Cometabolic degradation of acifluorfen by a mixed microbial culture. J. Environ. Sci. Health B 29:963987.CrossRefGoogle ScholarPubMed
Arjmand, M. and Sandermann, H. Jr. 1985. Mineralization of chloroaniline/lignin conjugates and of free chloroanilines by the white rot fungus Phanerochaete chrysosporium . J. Agric. Food Chem. 40:20012007.Google Scholar
Armengaud, J. and Timmis, K. N. 1997. The reductase RedA2 of the multicomponent dioxin dioxygenase system of Sphingomonas sp. RW1 is related to class-I cytochrome P450-type reductases. Eur. J. Biochem. 247:833842.Google Scholar
Armstrong, R. N. 1994. Glutathione S-transferases: structure and mechanism of an archetypical detoxication enzyme. Adv. Enzymol. Relat. Areas Mol. Biol. 69:144.Google ScholarPubMed
Avila, L. Z. and Frost, J. W. 1988. Monomeric metaphosphate formation during radical-based dephosphorylation. J. Am. Chem Soc. 110:79047906.CrossRefGoogle Scholar
Avila, L. Z. and Frost, J. W. 1989. Phosphonium ion fragmentations relevant to organophosphonate biodegradation. J. Am. Chem Soc. 111:89698970.CrossRefGoogle Scholar
Balba, M. H. and Saha, J. G. 1974. Degradation of matacil by the ascorbic acid oxidation system. Bull. Environ. Contam. Toxicol. 11:193200.CrossRefGoogle ScholarPubMed
Banks, M. K., Lee, E., and Schwab, A. 1999. Evaluation of dissipation mechanisms for benzo[a]pyrene in the rhizosphere of tall fescue. J. Environ. Qual. 28:294298.Google Scholar
Barik, S. and Munnecke, D. M. 1982. Enzymatic hydrolysis of concentrated diazinon in soil. Bull. Environ. Contam. Toxicol. 29:235239.CrossRefGoogle ScholarPubMed
Barkovskii, A. L. 2001. Microbial reductive dehalogenation in the rhizosphere. Pages 4056 In Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M., eds. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. ACS Symposium Series 777. Washington, DC: American Chemical Society.Google Scholar
Barkovskii, A. L. and Adriaenes, P. 1998. Impact of humic constituents on microbial dechlorination of polychlorinated dioxins. Environ. Toxicol. Chem. 17:10131020.CrossRefGoogle Scholar
Barkovskii, A. L., Bouillant, M. L., and Balandreau, J. 1994. Polyphenolic compounds respired by bacteria. Pages 2842 In Anderson, T. A. and Coats, J. R., eds. Bioremediation through Rhizosphere Technology. ACS Symposium Series 563. Washington, DC: American Chemical Society.CrossRefGoogle Scholar
Barkovskii, A. L., Bouillant, M. L., Jocteur-Monrozier, L., and Balandreau, J. 1995. Azospirillum strains use phenolic compounds as intermediates for electron transfer under oxygen-limiting conditions. Microbial Ecol. 29:99144.CrossRefGoogle ScholarPubMed
Barr, D. P. and Augst, S. D. 1994. Pollutant degradation by white rot fungi. Rev. Environ. Contam. Toxicol. 138:4972.Google ScholarPubMed
Barrentine, W. L., Edwards, C. J. Jr., and Hartwig, E. E. 1976. Screening soybeans for tolerance to metribuzin. Agron. J. 68:351353.Google Scholar
Barrentine, W. L., Hartwig, E. E., Edwards, C. J. Jr., and Kilen, T. C. 1982. Tolerance of three soybean (Glycine max) cultivars to metribuzin injury. Weed Sci. 30:344348.CrossRefGoogle Scholar
Barrett, M. 1998. Cloning and Heterologous Expression of Pesticide Metabolizing Cytochrome P450 Genes. Springfield, VA: National Technical Information Service, Fedrip Database.Google Scholar
Barrett, M. 2000. The role of cytochrome P450 enzymes in herbicide metabolism. Pages 2537 In Cobbs, A. H. and Kirkwood, R. C., eds. Herbicides and Their Mechanisms of Action. Sheffield, Great Britain: Sheffield Academic.Google Scholar
Barry, G., Kishore, G., Padgette, S. et al. 1992. Inhibitors of amino acid biosynthesis: strategies for imparting glyphosate tolerance to crop plants. Pages 139145 In Singh, B. K., Flores, H. E., and Shannon, J. C., eds. Biosynthesis and Molecular Regulation of Amino Acids in Plants. Madison, WI: American Society of Plant Physiology.Google Scholar
Bayley, C., Trolinder, N., Ray, C., Morgan, M., Quesenberry, J. E., and Ow, D. W. 1992. Engineering 2,4-D resistance into cotton. Theor. Appl. Genet. 83:645649.CrossRefGoogle Scholar
Benyon, K. I., Stoydin, G., and Wright, A. N. 1972a. The breakdown of the triazine herbicide cyanazine in soils and maize. Pestic. Sci. 3:293305.CrossRefGoogle Scholar
Benyon, K. I., Stoydin, G., and Wright, A. N. 1972b. The breakdown of the triazine herbicide cyanazine in wheat and potatoes grown under indoor conditions in treated soils. Pestic. Sci. 3:379387.CrossRefGoogle Scholar
Bickerstaff, G. F. 1997. Immobilization of enzymes and cells: Some practical considerations. Pages 110 In Bickerstaff, G. F., ed. Methods in Biotechnology 1. Totowa, NJ: Humana.Google Scholar
Bilang, J. and Sturm, A. 1995. Cloning and characterization of a glutathione S-transferase that can be photolabelled with 5-azido-indole-3-acetic acid. Plant Physiol. 109:253260.Google Scholar
Blaabjerg, V. and Finster, K. 1998. Sulphate reduction associated with roots and rhizomes of the marine macrophyte Zostera marina . Aquat. Microb. Ecol. 15:311314.CrossRefGoogle Scholar
Blake-Kalff, M.M.A., Randall, R. A., and Coleman, J.O.D. 1997. Compartmentation of detoxified xenobiotics in plant cells. Pages 245259 In Hatzios, K. K., ed. Regulation of Enzymatic Systems Detoxifying Xenobiotics in Plants. NATO ASI Series. Dordrecht, The Netherlands: Kluwer Academic.CrossRefGoogle Scholar
Bollag, J.-M. 1972. Biochemical transformation of pesticides by soil fungi. CRC Crit. Rev. Microbiol. 2:3558.CrossRefGoogle ScholarPubMed
Bolwell, G. P., Bozak, K., and Zimmerlin, A. 1994. Plant cytochrome P450. Phytochemistry 37:14911505.CrossRefGoogle ScholarPubMed
Boundy-Mills, K. L., de Souza, M. L., Mandelbaum, R. T., Wackett, L. P., and Sadowsky, M. J. 1997. The atzB gene of Pseudomonas sp. strain ADP encodes the second enzyme of a novel atrazine degradation pathway . Appl. Environ. Microbiol. 63:916923.CrossRefGoogle ScholarPubMed
Bowling, C. C. and Hudgins, H. R. 1966. The effect of insecticides on the selectivity of propanil in rice. Weeds 14:9495.Google Scholar
Bradshaw, W. H., Conrad, H. E., Corey, E. J., Gunsalus, I. C., and Lednicer, D. 1959. Microbial degradation of (+)-camphor. J. Am. Chem. Soc. 81:5507.Google Scholar
Breaux, E. J. 1987. Initial metabolism of acetochlor in tolerant and susceptible seedlings. Weed Sci. 35:463469.Google Scholar
Breaux, E. J., Patanella, J. E., and Sanders, E. F. 1987. Chloroacetanilide herbicide selectivity: analysis of glutathione and homoglutathione in tolerant, susceptible, and safened seedlings. J. Agric. Food Chem. 35:474478.CrossRefGoogle Scholar
Brenner, S. 1988. The molecular evolution of genes and proteins: a tale of two serines. Nature 334:528530.CrossRefGoogle ScholarPubMed
Brown, H. M. and Kearney, P. C. 1991. Plant biochemistry, environmental properties and global impact of the sulfonylurea herbicides. Pages 3249 In Baker, D. R., Fenyes, J. G., and Moberg, W. K., eds. Synthesis and Chemistry of Agrochemicals II. American Chemical Society (ACS) Symposium Series 443. Washington, DC: American Chemical Society.Google Scholar
Bryant, D. W., McCalla, D. R., Leeksma, M., and Laneuville, P. 1981. Type I nitroreductases of Escherichia coli . Can. J. Microbiol. 27:8186.CrossRefGoogle ScholarPubMed
Buckland, J. L., Collins, R. F., and Pullin, E. M. 1973. Metabolism of bromoxynil octanoate in growing wheat. Pestic. Sci. 4:149162.CrossRefGoogle Scholar
Bujacz, B., Wieczorek, P., Krzysko-Lupicka, T., Golab, Z., Lejczak, B., and Kafarski, P. 1995. Organophosphonate utilization by the wild-type strain of Penicillium notatum . Appl. Environ. Microbiol. 61:29052910.CrossRefGoogle ScholarPubMed
Cabanne, F., Huby, D., Gaillardon, P., Scalla, R., and Durst, F. 1987. Effect of the cytochrome P-450 inactivator 1-aminobenzotriazole on the metabolism of chlortoluron and isoproturon in wheat. Pestic. Biochem. Physiol. 28:371380.CrossRefGoogle Scholar
Cai, B., Vuilleumier, S., and Wackett, L. P. 1998. Purification and characterization of the mutant enzyme W117Y of the dichloromethane dehalogenase from Methylophilus sp. strain DM11. Ann. N. Y. Acad. Sci. 264:210213.CrossRefGoogle Scholar
Canivenc, M.-C., Cagnac, B., Cabanne, F., and Scalla, R. 1989. Induced changes of chlorotoluron metabolism in wheat cell suspension cultures. Plant Physiol. Biochem. 27:193201.Google Scholar
Carey, V. F. III, Duke, S. O., Hoagland, R. E., and Talbert, R. E. 1995a. Resistance mechanism of propanil-resistant barnyardgrass I. Absorption, translocation, and site of action studies. Pestic. Biochem. Physiol. 52:182189.Google Scholar
Carey, V. F. III, Hoagland, R. E., and Talbert, R. E. 1995b. Verification and distribution of propanil-resistant barnyardgrass (Echinochloa crus-galli) in Arkansas. Weed Technol. 9:366372.CrossRefGoogle Scholar
Carey, V. F. III, Hoagland, R. E., and Talbert, R. E. 1997. Resistance mechanism of propanil-resistant barnyardgrass. II. In-vivo metabolism of the propanil molecule. Pestic. Sci. 49:333338.Google Scholar
Carringer, R. D., Rieck, C. E., and Bush, L. P. 1978. Effect of R-25788 on EPTC metabolism in corn (Zea mays). Weed Sci. 26:167171.CrossRefGoogle Scholar
Cassidy, M. B., Lee, H., Trevors, J. T., and Zablotowicz, R. M. 1999. Chlorophenol and nitrophenol metabolism by Sphingomonas sp. UG30. J. Ind. Microbiol. Biotechnol. 23:232241.Google Scholar
Cerniglia, C. E. 1992. Biodegradation of polycyclic aromatic hydrocarbons. Biodegradation 3:351368.CrossRefGoogle Scholar
Chaudhry, G. R., Ali, A. N., and Wheeler, W. B. 1988. Isolation of a methyl parathion-degrading Pseudomonas sp. that possesses DNA homologous to the opd gene from a Flavobacterium sp. Appl. Environ. Microbiol. 54:288293.CrossRefGoogle Scholar
Cheah, U. B., Kirkwood, R. C., and Lum, K.- Y. 1998. Degradation of four commonly used pesticides in Malaysian agricultural soils. J. Agric. Food Chem. 46:12171223.Google Scholar
Chen, C. M., Ye, Q. Z., Zhu, Z., Wanner, B. L., and Walsh, C. T. 1990. Molecular biology of carbon-phosphorus bond cleavage. J. Biol. Chem. 265:44614471.CrossRefGoogle ScholarPubMed
Choi, K. D., Jeohn, G. H., Rhee, J. S., and Yoo, O. J. 1990. Cloning and nucleotide sequence of an esterase gene from Pseudomonas fluorescens and expression of the gene in Escherichia coli . Agric. Biol. Chem. 54:20392045.Google ScholarPubMed
Colbert, S. F., Schroth, M. N., Weinhold, A. R., and Hendson, M. 1993. Enhancement of population densities of Pseudomonas putida PpG7 in agricultural ecosystems by selective feeding with the carbon source salicylate. Appl. Environ. Microbiol. 59:20642070.CrossRefGoogle ScholarPubMed
Cole, D. J., Cummins, I., Hutton, P. J., Dixon, D. P., and Edwards, R. 1997. Glutathione transferases in crops and major weeds. Pages 139154 In Hatzios, K. K., ed. Regulation of Enzymatic Systems Detoxifying Xenobiotics in Plants. NATO ASI Series. Dordrecht, The Netherlands: Kluwer Academic.Google Scholar
Cook, A. M. 1987. Biodegradation of s-triazine xenobiotics. FEMS Microbiol. Rev. 46:93116.CrossRefGoogle Scholar
Cook, A. M. and Hutter, R. 1981. s-Triazines as nitrogen sources for bacteria degradation of herbicides. J. Agric. Food Chem. 29:11351143.CrossRefGoogle Scholar
Cordeiro, M. L., Pompliano, D. L., and Frost, J. W. 1986. Degradation and detoxification of organophosphonates: cleavage of the carbon to phosphorus bond. J. Am. Chem. Soc. 108:332334.Google Scholar
Criddle, C. S., McCarty, P. L., Eliott, M. C., and Barker, J. F. 1986. Reduction of hexachloroethane to tetrachloroethylene in groundwater. Contam. Hydrol. 1:133142.Google Scholar
Crowley, D. E., Alvey, S., and Gilbert, E. S. 1997. Rhizosphere ecology of xenobiotic-degrading microorganisms. Pages 2036 In Kruger, E. L., Anderson, T. A., and Coats, J. R., eds. Phytoremediation of Soil and Water Contaminants. ACS Symposium Series 777. Washington, DC: American Chemical Society.CrossRefGoogle Scholar
Crowley, D. E., Luepromechai, E., and Singer, A. 2001. Metabolism of xenobiotics in the rhizosphere. Pages 333352 In Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M., eds. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. ACS Symposium Series 777. Washington, DC: American Chemical Society.Google Scholar
Cullimore, D. R. and Kohout, M. 1974. Isolation of a bacterial degrader of the herbicide bromoxynil from a Saskatchewan soil. Can. J. Microbiol. 20:14491452.Google Scholar
Cygler, M., Grochulski, P., and Schrag, J. B. 1995. Structural determinants defining common stereoselectivity of lipases toward secondary alcohols. Can. J. Microbiol. 41 (Suppl. 1): 289296.CrossRefGoogle ScholarPubMed
da Silva, J. F. and Warren, G. F. 1976. Effect of stage of growth on metribuzin tolerance. Weed Sci. 24:612615.Google Scholar
Davies, J., Caseley, J. C., Jones, O., Barrett, M., and Polge, N. 1997. Mode of action of napthalic anhydride as a safener for the herbicide AC263222 in maize. Pestic. Sci. 52:2938.3.0.CO;2-J>CrossRefGoogle Scholar
Davies, J., Caseley, J. C., and Jones, O.T.G. 1993. Enhancement of AC 263222 metabolism by the herbicide safener naphthalic anhydride. Proc. Br. Crop Prot. Conf.—Weeds. pp. 195200.Google Scholar
De Kok, L. J., Maas, F. M., Godeke, J., Haaksma, A. B., and Kuiper, P.J.C. 1986. Glutathione, a tripeptide which may function as a temporary storage compound of excessive reduced sulphur in H2S fumigated spinach plants. Plant Soil. 91:349352.Google Scholar
de Souza, M. L., Sadowsky, M. J., and Wackett, L. P. 1996. Atrazine chlorohydrolase from Pseudomonas sp. strain ADP: gene sequence, enzyme purification and protein characterization. J. Bacteriol. 178:48944900.Google Scholar
de Souza, M. L., Seffernick, J., Martinez, B., Sadowsky, M. J., and Wackett, L. P. 1998a. The atrazine catabolism gene atzABC are widespread and highly conserved. J. Bacteriol. 180:19511954.Google Scholar
de Souza, M. L., Wackett, L. P., Boundy-Mills, K. L., Mandelbaum, R. T., and Sadowsky, M. J. 1995. Cloning, characterization, and expression of a gene region from Pseudomonas sp. strain ADP involved in the dechlorination of atrazine. J. Appl. Environ. Microbiol. 61:33733378.Google Scholar
de Souza, M. L., Wackett, L. P., and Sadowsky, M. J. 1998b. The atzABC genes encoding atrazine catabolism are located on a self-transmissible plasmid in Pseudomonas sp. strain ADP. J. Appl. Environ. Microbiol. 64:23232326.CrossRefGoogle ScholarPubMed
Dec, J. and Bollag, J.-M. 1994. Use of plant material for the decontamination of water polluted with phenols. Biotechnol. Bioeng. 44:11321139.Google Scholar
Dec, J. and Bollag, J.-M. 2000. Phenoloxidase-mediated interactions of phenols and anilines with humic materials. J. Environ. Qual. 29:665676.Google Scholar
Dec, J. and Bollag, J.-M. 2001. Use of enzymes in bioremediation. Pages 182193 In Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M., eds. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. ACS Symposium Series 777. Washington, DC: American Chemical Society.Google Scholar
Dixon, D. P., Cole, D. J., and Edwards, R. 1997. Characterisation of multiple glutathione transferases containing the GST I subunit with activities toward herbicide substrates in maize (Zea mays). Pestic. Sci. 50:7282.3.0.CO;2-Z>CrossRefGoogle Scholar
Dixon, D. P., Cole, D. J., and Edwards, R. 1998a. Purification, regulation and cloning of a glutathione transferase (GST) from maize resembling the auxin-inducible type-III GSTs. Plant Mol. Biol. 36:7587.CrossRefGoogle ScholarPubMed
Dixon, D. P., Cummins, I., Cole, D. J., and Edwards, R. 1998b. Glutathione-mediated detoxification systems in plants. Curr. Opin. Plant Biol. 1:258270.Google Scholar
Droog, F. 1997. Plant glutathione S-transferases, a tale of theta and tau. J. Plant Growth Regul. 16:95107.CrossRefGoogle Scholar
Dudler, R., Hertig, C., Rebmann, G., Bull, J., and Mauch, F. 1991. A pathogen-induced wheat gene encodes a protein homologous to glutathione-S-transferases. Mol. Plant-Microbe Interact. 4:1418.Google Scholar
Durst, F. and O’Keefe, D. P. 1995. Plant cytochromes P450: an overview. Drug Metab. Drug Interact. 12:171186.Google Scholar
Dyer, E. and Wright, G. C. 1959. Thermal degradation of alkyl N-phenylcarbamates. J. Am. Chem. Soc. 81:21382143.Google Scholar
Edwards, R. and Dixon, D. P. 1991. Glutathione S-cinnamoyl transferases in plants. Phytochemistry 30:7984.Google Scholar
Ekler, Z., Dutka, F., and Stephenson, G. R. 1993. Safener effects on acetochlor toxicity, uptake, metabolism and glutathione S-transferase activity in maize. Weed Res. 33:311318.CrossRefGoogle Scholar
Ekler, Z. and Stephenson, G. R. 1989. Physiological responses of maize and sorghum to four different safeners for metazachlor. Weed Res. 29:181191.Google Scholar
Engelhardt, G., Wallnofer, P. R., and Plapp, R. 1973. Purification and properties of an aryl acylamidase of Bacillus sphaericus, catalyzing the hydrolysis of various phenylamide herbicides and fungicides. Appl. Microbiol. 26:709718.Google Scholar
Erickson, L. E., Davis, L. C., and Narayanan, M. 1995. Bioenergetics and bioremediation of contaminated soil. Thermochim. Acta 250:353358.Google Scholar
Erickson, L. E. and Lee, K. H. 1989. Degradation of atrazine and related s-triazines. Crit. Rev. Environ. Control 19:114.Google Scholar
Ezra, G. and Gressel, J. 1982. Rapid effects of a thiocarbamate herbicide and its dichloroacetamide protectant on macromolecular syntheses and glutathione levels in maize cell cultures. Pestic. Biochem. Physiol. 17:4858.Google Scholar
Ezra, G., Krochmal, E., and Gressel, J. 1982. Competition between a thiocarbamate herbicide and herbicide protectants at the level of uptake into maize cells in culture. Pestic. Biochem. Physiol. 18:107112.Google Scholar
Fan, T. W., Lane, A. N., Pedler, J., Crowley, D., and Higashi, R. M. 1997. Comprehensive analysis of organic ligands in whole root exudates using nuclear magnetic resonance and gas chromatography-mass spectrometry. Anal. Biochem. 251:5768.CrossRefGoogle ScholarPubMed
Farago, S., Kreuz, K., and Brunold, C. 1993. Decreased glutathione levels enhance the susceptibility of maize seedlings to metolachlor. Pestic. Biochem. Physiol. 47:199205.Google Scholar
Feng, P.C.C. 1991. Soil transformation of acetochlor via glutathione conjugation. Pestic. Biochem. Physiol. 40:136142.Google Scholar
Feng, P.C.C. and Ruff, T. G. 2001. A review of strategies to engineer plant tolerance to the pyridine herbicides. Pages 129144 In Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M., eds. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. ACS Symposium Series 777. Washington, DC: American Chemical Society.Google Scholar
Feng, P.C.C. and Wratten, S. J. 1989. In vitro transformation of chloroacetanilide herbicides by rat liver enzymes: a comparative study of metolachlor and alachlor. J. Agric. Food Chem. 37:10881093.Google Scholar
Ferro, A., Sims, M.R.C., and Bugbee, B. 1994. Hycrest crested wheatgrass accelerates the degradation of pentachlorophenol in soil. J. Environ. Qual. 23:272279.Google Scholar
Feung, C., Hamilton, R. H., and Mumma, R. O. 1975. Metabolism of 2,4-dichlorophenoxyacetic acid. VII. Comparison of metabolities from five species of plant tissue cultures. J. Agric. Food Chem. 23:373376.Google Scholar
Feung, C., Hamilton, R. H., and Witham, F. H. 1971. Metabolism of 2,4-dichlorophenoxyacetic acid by soybean cotyledon callus tissue cultures. J. Agric. Food Chem. 19:475479.Google Scholar
Feung, C., Mumma, R. O., and Hamilton, R. H. 1974. Metabolism of 2,4-dichlorophenoxyacetic acid. VI. Biological properties of amino acid conjugates. J. Agric. Food Chem. 22:307309.Google Scholar
Field, J. A. and Thurmann, E. M. 1996. Glutathione conjugation and contaminant transformation. Environ. Sci. Technol. 30:14131418.CrossRefGoogle Scholar
Flanders, C., Dec, J., and Bollag, J.- M. 1999. Horseradish-mediated binding of 2,4-dichlorophenol to soil. Bioremed. J. 3:315321.CrossRefGoogle Scholar
Forlani, G., Mangiagalli, A., Nielsen, E., and Suardi, C. M. 1999. Degradation of the phosphonate herbicide glyphosate in soil: evidence for a possible involvement of unculturable microorganisms. Soil Biol. Biochem. 31:991997.CrossRefGoogle Scholar
Frear, D. S. 1976. Pesticide conjugates-glycosides. Pages 3554 In Kaufman, D. D., Still, G. G., Paulson, G. D., and Bandal, S. K., eds. Bound and Conjugated Pesticide Residues. ACS Symposium Series 29. Washington, DC: American Chemical Society.CrossRefGoogle Scholar
Frear, D. S., Mansager, E. R., and Swanson, H. R. 1989. Picloram metabolism in leafy spurge: isolation and identification of glucose and gentiobiose conjugates. J. Agric. Food Chem. 37:14081412.CrossRefGoogle Scholar
Frear, D. S., Mansager, E. R., Swanson, H. R., and Tanaka, F. S. 1983. Metribuzin metabolism in tomato: isolation and identification of N-glucoside conjugates herbicide, residues, tolerance. Pestic. Biochem. Physiol. 19:270281.CrossRefGoogle Scholar
Frear, D. S. and Still, G. G. 1968. The metabolism of 3,4-dichloropropionanilide in plants. Partial purification and properties of an aryl acylamidase from rice. Phytochemistry 7:913920.Google Scholar
Frear, D. S., Swanson, H. R., and Mansager, E. R. 1985. Alternate pathways of metribuzin metabolism in soybean: formation of N-glucoside and homoglutathione conjugates. Pestic. Biochem. Physiol. 23:5665.CrossRefGoogle Scholar
Frear, D. S., Swanson, H. R., Mansager, E. R., and Wien, R. G. 1978. Chloramben metabolism in plants: isolation and identification of glucose ester. J. Agric. Food Chem. 26:13471351.Google Scholar
Frear, D. S., Swanson, H. R., and Tanaka, F. S. 1993. Metabolism of flumetsulam (DE498) in wheat, corn and barley. Pestic. Biochem. Physiol. 45:178192.Google Scholar
Freedman, L. D. and Doak, G. O. 1957. The preparation and properties of phosphonic acids. Chem. Rev. 57:479523.Google Scholar
Frenzel, P., Bosse, U., and Janssen, P. H. 1999. Rice roots and methanogenesis in a paddy soil: ferric iron as an alternative electron acceptor in the rooted soil. Soil Biol. Biochem. 31:421430.Google Scholar
Frey, M., Chomet, P., Glaswischnig, E. et al. 1997. Analysis of a chemical plant defense mechanism in grasses. Science 277:696699.Google Scholar
Frost, J. W., Loo, S., Cordeiro, M. L., and Li, D. 1987. Radical-based dephosphorylation and organophosphonate biodegradation. J. Am. Chem. Soc. 109:21662171.Google Scholar
Frova, C., Sari Gorla, M., Pe, M. E., Greenland, A., Jepson, I., and Rossini, L. 1997. Role of the different GST isozymes of maize in herbicide tolerance: genetics and biochemical analysis. Pages 171181 In Hatzios, K. K., ed. Regulation of Enzymatic Systems Detoxifying Xenobiotics in Plants. NATO ASI Series. Dordrecht, The Netherlands: Kluwer Academic.Google Scholar
Fuenmayor, S. L., Wild, M., Boyes, A. L., and Williams, P. A. 1998. A gene cluster encoding steps in conversion of naphthalene to gentisate in Pseudomonas sp. strain U2. J. Bacteriol. 180:25222530.Google Scholar
Fuerst, E. P. and Gronwald, J. W. 1986. Induction of rapid metabolism of metolachlor in sorghum (Sorghum bicolor) shoots by CGA-92194 and other antidotes. Weed Sci. 34:354361.Google Scholar
Gaillard, C., Dufaud, A., Tommasini, R., Kreuz, K., Amhreim, N., and Martinoia, E. 1994. A herbicide antidote (safener) induces the activity of both the herbicide detoxifying enzyme and of vacuolar transporter for the detoxified herbicide. FEBS Lett. 352:219221.Google Scholar
Gaillardon, P., Cabanne, F., Scalla, R., and Durst, F. 1985. Effect of mixed function oxidase inhibitors on the toxicity of chlortoluron and isoproturon to wheat. Weed Res. 25:397402.Google Scholar
Gaynor, J. D. 1992. Microbial hydrolysis of diclofop-methyl in soil. Soil Biol. Biochem. 24:2932.Google Scholar
Gennari, M., Negre, M., Ambosoli, R., Andreoni, V., Vincenti, M., and Acquati, A. 1994. Anaerobic degradation of acifluorfen by different enrichment cultures. J. Agric. Food Chem. 42:12321236.Google Scholar
Gennari, M., Vincenti, M., Negre, M., and Ambosoli, R. 1995. Microbial metabolism of fenoxaprop-ethyl. Pestic. Sci. 44:299303.Google Scholar
Ghisalba, O., Kuenzi, M., Ramos Tombo, G. M., and Schar, H.- P. 1987. Microbial degradation and utilization of selected organophosphorus compounds—strategies and applications. Chimia 41:206215.Google Scholar
Gray, J. A., Balke, N. E., and Stoltenberg, D. E. 1996. Increased glutathione conjugation of atrazine confers resistance in a Wisconsin velvetleaf (Abutilon theophrasti) biotype. Pestic. Biochem. Physiol. 55:157167.Google Scholar
Griffith, G. D., Cole, J. R., Quensen, J. F. III, and Tiedje, J. M. 1992. Specific deuteration of dichlorobenzoate during reductive dehalogenation by Desulfomonile tiedjei in D2O. Appl. Environ. Microbiol. 58:409411.Google Scholar
Groenewegen, P.E.J., Breeuwer, P., van Helvoort, J.M.L.M., Langenhoff, A.A.M., de Vries, F. P., and de Bont, J.A. M. 1992. Novel degradative pathway of 4-nitrobenzoate in Comamonas acidovorans NBA-10. J. Gen. Microbiol. 138:15991605.CrossRefGoogle ScholarPubMed
Groenewegen, P.E.J. and de Bont, J.A.M. 1992. Degradation of 4-nitrobenzoate via 4-hydroxylaminobenzoate and 3,4-dihydroxybenzoate in Comamonas acidovorans NBA-10. Arch. Microbiol. 158:381386.Google Scholar
Gronwald, J. W., Fuerst, E. P., Eberlein, C. V., and Egli, M. A. 1987. Effect of herbicide antidotes on glutathione content and glutathione S-transferase activity of sorghum shoots. Pestic. Biochem. Physiol. 29:6676.CrossRefGoogle Scholar
Haby, P. A. and Crowley, D. E. 1996. Biodegradation of 3-chlorobenzoate as affected by rhizodeposition and selected carbon substrates. J. Environ. Qual. 25:304310.Google Scholar
Hagen, G., Uhrhammer, N., and Guilfoyle, T. J. 1988. Regulation of expression of an auxin-induced soybean sequence by cadmium. J. Biol. Chem. 263:64426446.CrossRefGoogle ScholarPubMed
Häggblom, M. M. 1990. Mechanisms of bacterial degradation and transformation of chlorinated monoaromatic compounds. J. Basic Microbiol. 2:115141.Google Scholar
Häggblom, M. M., Janke, D., and Salkinoja-Salonen, M. S. 1989. Hydroxylation and dechlorination of tetrachlorohydroquinone by Rhodoccus sp. strain CP-2 cell extracts. Appl. Environ. Microbiol. 55:516519.Google Scholar
Haider, K., Mosier, A., and Heinemeyer, O. 1987. The effect of growing plants on denitrification at high soil nitrate concentrations. Soil. Sci. Soc. Am. J. 51:97102.Google Scholar
Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M. 2001a. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. Washington, DC: American Chemical Society. 432 p.Google Scholar
Hall, J. C., Wickenden, J. S., and Yau, K.Y.F. 2001b. Biochemical conjugation of pesticides in plants and microogranisms: an overview of similarities and divergences. Pages 89118 In Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M., eds. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. Washington, DC: American Chemical Society.Google Scholar
Hallas, L. E., Hahn, E. M., and Korndorfer, C. 1988. Characterization of microbial traits associated with glyphosate biodegradation in industrial activated sludge. J. Ind. Microbiol. 3:377385.Google Scholar
Hammond, P. M., Price, C. P., and Scaven, M. D. 1983. Purification and properties of aryl acylamidase from Pseudomonas fluorescens ATCCC 39004. Eur. J. Biochem. 132:651655.Google Scholar
Han, S. and Hatzios, K. K. 1991. Physiological interactions between the herbicide pretilachlor and the safener fenclorim on rice. Pestic. Biochem. Physiol. 39:270280.Google Scholar
Hardcastle, W. S. 1974. Differences in the tolerance of metribuzin by varieties of soybeans. Weed Res. 14:181184.CrossRefGoogle Scholar
Hardcastle, W. S. 1979. Soybean (Glycine max) cultivar response to metribuzin in solution culture. Weed Sci. 27:278279.Google Scholar
Hartwig, E. E., Barrentine, W. L., and Edwards, J. Jr. 1980. Registration of Tracy-M soybeans. J. Crop Sci. 20:825.Google Scholar
Hatzios, K. K. 1991. Biotransformations of herbicides in higher plants. Pages 141185 In Grover, R. and Cessna, A. J., eds. Environmental Chemistry of Herbicides. Boca Raton, FL: CRC Press.Google Scholar
Hatzios, K. K. 1997. Regulation of xenobiotic degrading enzymes with herbicide safeners. Pages 275288 In Hatzios, K. K., ed. Regulation of Enzymatic Systems Detoxifying Xenobiotics in Plants. NATO ASI Series. Dordrecht, The Netherlands: Kluwer Academic Publishers.Google Scholar
Hatzios, K. K. 2001. Functions and regulation of plant glutathione-S-transferases. Pages 218239 In Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M., eds. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. Washington, DC: American Chemical Society.Google Scholar
Hatzios, K. K. and Hoagland, R. E. 1989. Development of herbicide safeners: industrial and university perspectives. Pages 338 In Hatzios, K. K. and Hoagland, R. E., eds. Crop Safeners for Herbicides: Development, Uses, and Mechanisms of Action. San Diego, CA: Academic Press.Google Scholar
Hausladen, A. and Alscher, R. G. 1993. Glutathione. Pages 130 In Alscher, R. G. and Hess, J. L., eds. Antioxidants in Higher Plants. Boca Raton, FL: CRC Press.Google Scholar
Hershey, H. P. and Stoner, T. D. 1991. Isolation and characterization of cDNA clones for RNA species induced by substituted benzenesulfonamides in corn. Plant Mol. Biol. 17:679690.Google Scholar
Hirase, K. and Matsunaka, S. 1991. Purification and properties of propanil hydrolase in Pseudomonas picketti . Pestic. Biochem. Physiol. 39:302308.CrossRefGoogle Scholar
Hoagland, R. E. 1975. Hydrolysis of 3’,4'-dichloropropionanilide by an aryl acylamidase from Taraxacum officinale . Phytochemistry 14:383386.Google Scholar
Hoagland, R. E. 1978. Isolation and some properties of an aryl acylamidase from red rice, Oryza sativa L., that metabolizes 3’,4'-dichloropropionanilide. Plant Cell Physiol. 19:10191027.Google Scholar
Hoagland, R. E. and Graf, G. 1972. An aryl acylamidase from tulip which hydrolyzes 3',4'-dichloropropionanilide. Phytochemistry 11:521527.Google Scholar
Hoagland, R. E. and Zablotowicz, R. M. 1995. Rhizobacteria with exceptionally high aryl acylamidase activity. Pestic. Biochem. Physiol. 52:190200.Google Scholar
Hoagland, R. E. and Zablotowicz, R. M. 1998. Biotransformations of fenoxaprop-ethyl by fluorescent Pseudomonas strains. J. Agric. Food Chem. 45:47594765.Google Scholar
Hoagland, R. E. and Zablotowicz, R. M. 2001. The role of plant and microbial hydrolytic enzymes in pesticide metabolism. Pages 5888 In Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M., eds. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. ACS Symposium Series 777. Washington, DC: American Chemical Society.Google Scholar
Hoagland, R. E., Zablotowicz, R. M., and Locke, M. A. 1997. An integrated phytoremediation strategy for chloroacetamides in soil. Pages 3853 In Kruger, E. L., Anderson, T. A., and Coats, J. R., eds. Phytoremediation of Soil and Water Contaminants. ACS Symposium Series 664. Washington, DC: American Chemical Society.Google Scholar
Hodgson, R. H., Frear, D. S., Swanson, H. R., and Regan, L. A. 1973. Alteration of diphenamid metabolism in tomato by ozone. Weed Sci. 21:542549.CrossRefGoogle Scholar
Hoffman, O. L. 1962. Chemical seed treatments as herbicidal antidotes. Weed Sci. 10:322323.Google Scholar
Hückelhoven, R., Schuphan, I., Thiede, B., and Schmidt, B. 1997. Biotransformation of pyrene by cell cultures of soybean (Glycine max L.), wheat (Triticum aestivum L.), jimsonweed (Datura stramonium L.), and purple foxglove (Digitalis purpurea L.). J. Agric. Food Chem. 45:263269.Google Scholar
Hustler, A. and Marschner, H. 1994. The influence of root exudates on the uptake of PCDD/PCDE by plants. Organohalogen Compounds 20:3134.Google Scholar
Incledon, B. J. and Hall, J. C. 1997. Enzymatic de-esterification of xenobiotics in plants. Pages 6782 In Hatzios, K. K., ed. Regulation of Enzymatic Systems Detoxifying Xenobiotics in Plants. NATO ASI Series. Dordrecht, The Netherlands: Kluwer Academic Publishers.Google Scholar
Iwan, J. 1976. Miscellaneous conjugations—acylation and alkylation of xenobiotics in physiologically active systems. Pages 132152 In Kaufman, D. D., Still, G. G., Paulson, G. D., and Bandal, S. K., eds. Bound and Conjugated Pesticide Residues. ACS Symposium Series 29. Washington, DC: American Chemical Society.Google Scholar
Jepson, I., Andrews, C. J., Roussel, V., Skipsey, M., and Towson, J. K. 1999. Transgenic approaches to understanding glutathione S-transferases. Abstr. Weed Sci. Soc. Am. 39:355.Google Scholar
Jepson, I., Lay, V. J., Holt, D. C., Bright, S.W.J., and Greenland, A. 1994. Cloning and characterization of maize herbicide safener-induced cDNAs encoding subunits of glutathione S-transferase isoforms I, II, and IV. J. Plant Mol. Biol. 26:18551866.Google Scholar
Jones, A. M. 1994. Auxin-binding proteins. Annu. Rev. Plant Physiol. Plant Mol. Biol. 45:393420.Google Scholar
Jordahl, J. L., Foster, L., Schnoor, J. L., and Alvarez, P. J. 1997. Effect of hybrid poplar trees on microbial populations important to hazardous waste bioremediation. J. Environ. Toxicol. Chem. 16:13181321.CrossRefGoogle Scholar
Joshi, D. K. and Gold, M. H. 1993. Degradation of 2,4,5-trichlorophenol by the lignin-degrading basidiomycete Phanerochaete chrysosporium . Appl. Environ. Microbiol. 59:17791785.Google Scholar
Kadiyala, V. and Spain, J. C. 1998. A two-component monooxygenase catalyzes both the hydroxylation of p-nitrophenol and the oxidative release of nitrite from 4-nitrocatechol in Bacillus sphaericus JS905. Appl. Environ. Microbiol. 64:24792484.Google Scholar
Kafarski, P., Lejczak, B., and Forlani, G. 2001. Biodegradation of pesticides containing carbon-to-phosphorous bond. Pages 145163 In Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M., eds. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. ACS Symposium Series 777. Washington, DC: American Chemical Society.Google Scholar
Katayama, A. and Matsumura, F. 1993. Degradation of organochlorine pesticides, particularly endosulfan, by Trichoderma harzianum . Environ. Toxicol. Chem. 12:10591065.Google Scholar
Kaufman, D. D., Blake, J., and Miller, D. E. 1971. Methylcarbamates affect acylanilide herbicide residues in soil. J. Agric. Food Chem. 19:204206.Google Scholar
Kearney, P. C. and Kaufman, D. D. 1965. Enzyme from soil bacterium hydroylzes phenylcarbamate herbicides. Science 147:740741.Google Scholar
Ketchersid, M. L., Vietor, D. M., and Merkle, M. G. 1982. CGA-43089 effects on metolachlor uptake and membrane permeability in grain sorghum (Sorghum bicolor). J. Growth Regul. 1:285294.Google Scholar
Kim, S. K., Makino, K., Amemura, M., Shinagawa, H., and Nakata, A. 1993. Molecular analysis of the phoH gene, belonging to the phosphate regulon in Escherichia coli . J. Bacteriol. 175:13161324.Google Scholar
Kinouchi, T. and Ohnishi, Y. 1983. Purification and characterization of 1-nitropyrene nitroreductases from Bacteroides fragilis . Appl. Environ. Microbiol. 46:596604.CrossRefGoogle ScholarPubMed
Klambt, D. H. 1961. Stoffwechselprodukte der naphthyl-1-essigsaure und 2,4-dichlorphenoxyessigsaure und der vergleich mit jenen der indol-3-essigsaure und benzoesaure. Planta 57:339353.Google Scholar
Köcher, H., Kellner, H. M., Lötzsch, K., Dorn, E., and Wink, O. 1982. Mode of action and metabolic fate of the herbicide fenoxaprop-ethyl, HOE 33171. Br. Crop Prot. Conf. Weeds 1:341347.Google Scholar
Komoβa, D., Gennity, I., and Sandermann, H. Jr. 1992. Plant metabolism of herbicides with C-P bonds: glyphosate. Pestic. Biochem. Physiol. 43:8594.Google Scholar
Komoβa, D. and Sandermann, H. Jr. 1992. Plant metabolism of herbicides with C-P bonds: phosphinothricin. Pestic. Biochem. Physiol. 43:95102.Google Scholar
Korpraditskul, R., Katayama, A., and Kuwatsuka, S. 1993. Degradation of atrazine by soil bacteria in the stationary phase. J. Pestic. Sci. 18:293298.Google Scholar
Kreslavski, V. D., Vasilyeva, G. K., Comfort, S. D., Drijber, R. A., and Shea, P. J. 1999. Accelerated transformation and binding of 2,4,6-trinitrotoluene in rhizosphere soil. Bioremed. J. 3:5969.Google Scholar
Kreuz, K., Gaudin, J., Stingelin, J., and Ebert, E. Z. 1991. Metabolism of the aryloxyphenoxypropanoate herbicide, CGA 184927, in wheat, barley and maize: differential effects of the safener CGA 185072. Z. Naturforsch. 46c:901905.Google Scholar
Kreuz, K., Tommasini, R., and Martinoia, E. 1996. Old enzymes for a new job: herbicide detoxification in plants. Plant Physiol. 111:349353.Google Scholar
Krzysko-Lupicka, T. and Orlik, A. 1997. The use of glyphosate as the sole source of phosphorus or carbon for the selection of soil-borne fungal strains capable to degrade this herbicide. Chemosphere 34:26012605.Google Scholar
Kullman, S. W. and Matsumura, F. 1996. Metabolic pathways utilized by Phanerochaete chrysosporium for degradation of the cyclodiene pesticide endosulfan. Appl. Environ. Microbiol. 62:593600.Google Scholar
Kusaba, M., Takahashi, Y., and Nagata, T. 1996. A multiple-stimuli-responsive as-1-related element of parA gene confers responsiveness to cadmium but not to copper. Plant Physiol. 111:11611167.Google Scholar
Lamar, R. T. and Dietrich, D. M. 1990. In situ depletion of pentachlorophenol from contaminated soil by Phanerochaete spp . Appl. Environ. Microbiol. 56:30933100.CrossRefGoogle ScholarPubMed
Lamoureux, G. L. and Frear, D. S. 1979. Pesticide metabolism in higher plants: in vitro enzyme studies. Pages 77128 In Paulson, G. D., Frear, D. S., and Marks, E. P., eds. Xenobiotic Metabolism: In Vitro Methods. ACS Symposium Series 97. Washington, DC: American Chemical Soceity.Google Scholar
Lamoureux, G. L. and Rusness, D. G. 1980. Pentachloronitrobenzene metabolism in peanut. 1. Mass spectral characterization of seven glutathione-related conjugates produced in vivo or in vitro . J. Agric. Food Chem. 28:10571070.Google Scholar
Lamoureux, G. L. and Rusness, D. G. 1986. Xenobiotic conjugation in higher plants. Pages 62105 In Paulson, G. D., Caldwell, J., Hutson, D. H., and Menn, J. J., eds. Xenobiotic Conjugation Chemistry. ACS Symposium Series 97. Washington, DC: American Chemical Society.Google Scholar
Lamoureux, G. L. and Rusness, D. G. 1989. The role of glutathione-S-transferases in pesticide metabolism, selectivity, and mode of action in plants and insects. Pages 153196 In Dolphin, D., Polson, R., and Avramovic, O., eds. Coenezymes and Cofactors. Glutathione: Chemical, Biochemical and Medical Aspects. Volume 3. New York: Wiley.Google Scholar
Lamoureux, G. L. and Rusness, D. G. 1991. The effect of BAS 145138 safener on chlorimuron ethyl metabolism and toxicity in corn. Z. Naturforsch. 46c:882886.Google Scholar
Lamoureux, G. L. and Rusness, D. G. 1993. Glutathione in the metabolism and detoxification of xenobiotics in plants. Pages 221237 In De Kok, L. J., Stulen, I., Rennenberg, H., Brunold, C., and Rauser, W., eds. Sulfur Nutrition and Assimilation in Higher Plants. The Hague, The Netherlands: SPB Academic Publishing.Google Scholar
Lamoureux, G. L., Shimabukuro, R. H., and Frear, D. S. 1991. Glutathione and glucoside conjugation in herbicide selectivity. Pages 227261 In Casely, J. C., Cussans, G. W., and Atkin, R. K., eds. Herbicide Resistance in Weeds and Crops. Oxford: Butterworth-Heinemann.Google Scholar
Lamoureux, G. L., Simoneaux, B., and Larson, J. 1998. The metabolism of atrazine and related 2-chloro-4,6-bis(alkylamino)-s-triazines in plants. Pages 6081 In Ballantine, L. G., McFarland, J. E., and Hackett, D. S., eds. Triazine Herbicides: Risk Assessment. ACS Symposium Series 68. Washington, DC: American Chemical Society.CrossRefGoogle Scholar
Lay, M. M. and Casida, J. E. 1976. Dichloroacetamide antidotes enhance thiocarbamate sulfoxide detoxification by elevating corn root glutathione content and glutathione S-transferase activity. Pestic. Biochem. Physiol. 6:442456.Google Scholar
Leah, J. M., Caseley, J. C., Riches, C. R., and Valverde, B. 1994. Association between elevated activity of aryl acylamidase and propanil resistance in jungle-rice Echinochloa colona . Pestic. Sci. 42:281289.Google Scholar
Leavitt, J.R.C. and Penner, D. 1979. In vitro conjugation of glutathione and other thiols with acetanilide herbicides and EPTC sulfoxide and the action of the herbicide antidote R-25688. J. Agric. Food Chem. 27:533536.Google Scholar
Leisinger, T., Bader, R., Hermann, R., Schmid-Appert, M., and Vuilleumier, S. 1994. Microbes, enzymes and genes involved in dichloromethane utilization. Biodegradation 5:237248.Google Scholar
Lenke, H. and Knackmuss, H. J. 1992. Initial hydrogenation during catabolism of picric acid by Rhodococcus erythropolis HL 24-2. Appl. Environ. Microbiol. 58:29332937.Google Scholar
Lenke, H., Pieper, D. H., Bruhn, C., and Knackmuss, H. J. 1992. Degradation of 2,4-dinitrophenol by two Rhodococcus erythropolis strains, HL 24-1 and HL 24-2. Appl. Environ. Microbiol. 58:29282932.Google Scholar
Leung, K. T., Cassidy, M. B., Shaw, K. W., Lee, H., Trevors, J. T., Lohmeier-Vogel, E. L., and Vogel, H. J. 1997. Pentachlorophenol biodegradation by Pseudomonas spp. UG25 and UG30. World J. Microbiol. Biotech. 13:305313.Google Scholar
Li, Z.-S., Zhao, Y., and Rhea, P. A. 1995a. Magnesium adenosine 5'-triphosphate-energized transport of glutathione S-conjugates by plant vacuolar membrane vesicles. Plant Physiol. 107:12571268.Google Scholar
Li, Z.-S., Zhen, R.-G., and Rea, P. A. 1995b. 1-Chloro-2,4-dinitrobenzeneelicited increase in vacuolar glutathione-S-conjugate transport activity. Plant Physiol. 109:177185.Google Scholar
Liu, S. Y., Schocken, M. J., and Rosazza, J.P.N. 1996. Microbial transformations of clomazone. J. Agric. Food Chem. 44:313319.Google Scholar
Llewellyn, D. and Last, D. 1996. Genetic engineering of crops for tolerance to 2,4-D. Pages 159174 In Duke, S. O., ed. Herbicide Resistant Crops. Boca Raton, FL: CRC Press.Google Scholar
Locke, M. A., Gaston, L. A., and Zablotowicz, R. M. 1997. Acifluorfen sorption and sorption kinetics in soil. J. Agric. Food Chem. 45:286293.Google Scholar
Lovley, D. T., Coates, J. D., Blunt-Harris, E. L., Phillips, E.J.P., and Woodward, J. C. 1996. Humic substances as electron acceptors for microbial respiration. Nature 382:445448.Google Scholar
Low, P. S. and Merida, J. R. 1996. The oxidative burst in plant defense: function and signal transduction. Physiol. Plant. 96:533542.Google Scholar
Lusby, W. R., Oliver, J. E., and Kearney, P. C. 1980. Metabolism of 2,6-dinitro-4-(trifluoromethyl)benzenamine by a Streptomyces isolated from soil. J. Agric. Food Chem. 28:641644.Google Scholar
Macnicol, P. K. 1987. Homoglutathione and glutathione synthetases of legume seedlings: partial purification and substrate specificity. Plant Sci. 53:229235.Google Scholar
Malik, J., Barry, G., and Kishore, G. M. 1989. The herbicide glyphosate. Biofactors 2:1725.Google Scholar
Mandelbaum, R. T., Allan, D. L., and Wackett, L. P. 1995. Isolation and characterization of a Pseudomonas sp. that mineralizes the s-triazine herbicide atrazine. Appl. Environ. Microbiol. 61:14511457.CrossRefGoogle ScholarPubMed
Mangeot, B. L., Slife, F. E., and Rieck, C. E. 1979. Differential metabolism of metribuzin herbicide by two soybean (Glycine max) cultivars. Weed Sci. 27:267269.Google Scholar
Marrs, K. A. 1996. The functions and regulation of glutathione S-transferases in plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 47:127158.Google Scholar
Marrs, K. A., Alfenito, M. R., Lloyd, A. M., and Walbot, V. 1995. A glutathione S-transferase involved in vacuolar transfer encoded by the maize gene Bronze-2 . Nature 375:397400.Google Scholar
Martinez, B., Tompkins, J., Wackett, L. P., Wing, R., and Sadowsky, M. J. 2001. Complete nucleotide sequence and organization of the atrazine catabolic plasmid pADP-1 from Pseudomonas sp. strain ADP. J. Bacteriol. 183:56845697.Google Scholar
Martinoia, E., Grill, E., Tommasini, R., Kreuz, K., and Amhreim, N. 1993. ATP-dependent glutathione S-conjugate ‘export’ pump in the vacuolar membrane of plants. Nature 364:247249.Google Scholar
Matsunaka, S. 1968. Propanil hydrolysis: inhibition in rice plants by insecticides. Science 160:13601361.Google Scholar
McBride, K. E., Kenny, J. W., and Stalker, D. M. 1986. Metabolism of the herbicide bromoxynil by Klebsiella pneumoniae subsp. ozaenae . Appl. Environ. Microbiol. 52:325330.Google Scholar
McCarthy, D. L., Navarrete, S., Willett, W. S., Babbitt, P. C., and Copley, S. D. 1996. Exploration of the relationship between tetrachlorohydroquinone dehalogenase and the glutathione S-transferase superfamily. Biochemistry 35:14 63414 642.Google Scholar
McDaniel, C. S., Harper, L. L., and Wild, J. R. 1988. Cloning and sequencing of a plasmid-borne gene (opd) encoding a phosphotriesterase. J. Bacteriol. 170:23062311.Google Scholar
McGonigle, B., Lau, S.-M. C., and O’Keefe, D. P. 1997. Endogenous reactions and substrate specificity of herbicide metabolizing enzymes. Pages 918 In Hatzios, K. K., ed. Regulation of Enzymatic Systems Detoxifying Xenobiotics in Plants. NATO ASI Series. Dordrecht, The Netherlands: Kluwer Academic Publishers.Google Scholar
Menendez, J. and De Prado, R. 1997. Detoxification of chlorotoluron in a chlorotoluron-resistant biotype of Alopecurus myosuroides . Comparison between cell cultures and whole plants. Physiol. Plant. 99:97104.Google Scholar
Metcalf, W. W. and Wanner, B. L. 1991. Involvement of the Escherichia coliphn (psiD) gene cluster I assimilation of phosphorus in the form of phosphonates, phosphite, Pi esters, and Pi . J. Bacteriol. 173:587600.Google Scholar
Milcamps, A. and deBruijn, F. J. 1999. Identification of a novel nutrientdeprivation-induced Sinorhizobium meliloti gene (hmgA) involved in the degradation of tyrosine. Microbiology 145:935947.Google Scholar
Mine, A., Miyakado, M., and Matsunaka, S. 1975. The mechanism of bentazon selectivity. Pestic. Biochem. Physiol. 5:566574.Google Scholar
Mochida, K., Nakamura, T., Li, W. X., and Ozoe, Y. 1993. Purification of extracellular aryl acylamidase from a coryneform bacterium, strain A-1. J. Pestic. Sci. 18:211216.Google Scholar
Mougin, C., Cabanne, F., Canivenc, M.-C., and Scalla, R. 1990. Hydroxylation and N-demethylation of chlorotoluron by wheat microsomal enzymes. Plant Sci. 66:195203.Google Scholar
Mougin, C. P., Corio-Costet, M.-F., and Werck-Reichhart, D. 2001. Plant and fungal cytochrome P-450s: their role in pesticide transformation. Pages 166182 In Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M., eds. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. ACS Symposium Series 777. Washington, DC: American Chemical Society.Google Scholar
Mougin, C., Pericaud, C., Dubroca, J., and Asther, M. 1997. Enhanced mineralization of lindane in soils supplemented with the white rot basidiomycete Phanerochaete chrysosporium . Soil Biol. Biochem. 29:13211324.Google Scholar
Mougin, C., Pericaud, C., Malosse, C., Laugero, C., and Asther, M. 1996. Biotransformation of the insecticide lindane by the white rot basidiomycete Phanerochaete chrysosporium . Pestic. Sci. 47:5159.Google Scholar
Mulbry, W. W. and Karns, J. S. 1989. Parathion hydrolase specified by the Flavobacterium opd gene relationship between the gene and protein. J. Bacteriol. 171:67406746.Google Scholar
Munnecke, D. M. 1976. Enzymatic hydrolysis of organophosphate insecticides, a possible disposal method. Appl. Environ. Microbiol. 32:713.Google Scholar
Nagata, Y., Miyauchi, K., and Takagi, M. 1999. Complete analysis of genes and enzymes for γ-hexachlorocyclohexane degradation in Sphingomonas paucimobilis UT26. J. Ind. Microbiol. Biotechnol. 23:380390.Google Scholar
Nannipieri, P. and Bollag, J.-M. 1991. Use of enzymes to detoxify pesticide-contaminated soils and waters. J. Environ. Qual. 20:510517.Google Scholar
Nardi, S., Reniero, F., and Concheri, G. 1997. Soil organic matter mobilization by root exudates of three maize hybrids. Chemosphere 35:22372244.Google Scholar
Neuefeind, T., Huber, R., Dasenbrock, H., Prade, L., and Bieseler, B. 1997a. Crystal structure of herbicide-detoxifying maize glutathione S-transferase-I in complex with lactoylglutathione: evidence for an inducedfit mechanism. J. Mol. Biol. 274:446453.Google Scholar
Neuefeind, T., Huber, R., Knablein, J., Prade, L., Mann, K., and Bieseler, B. 1997b. Cloning, sequencing, crystallization and x-ray structure of glutathione S-transferase-III from Zea mays var. mutin: a leading enzyme in detoxification of maize herbicides. J. Mol. Biol. 274:577587.Google Scholar
Nichols, T. D., Wolf, D. C., Roders, H. B., Beyrouty, C. A., and Renolds, C. M. 1997. Rhizosphere microbial populations in contaminated soils. Water Air Soil Pollut. 95:165178.Google Scholar
Nishida, M., Harada, S., Noguchi, S., Satow, Y., Inoue, H., and Takahashi, K. 1998. Three-dimensional structure of Escherichia coli glutathione S-transferase complexed with glutathione sulfonate: catalytic roles of Cys10 and His106. J. Mol. Biol. 281:135147.Google Scholar
Nishino, S. F. and Spain, J. C. 1993. Degradation of nitrobenzene by a Pseudomonas pseudocaligenes . Appl. Environ. Microbiol. 59:25202525.Google Scholar
Nomura, N. S. and Hilton, H. W. 1977. The absorption and degradation of glyphosate in five Hawaiian sugarcane soils. Weed Res. 17:113121.Google Scholar
Norsworthy, J. K., Talbert, R. E., and Hoagland, R. E. 1999. Chlorophyll fluorescence evaluation of agrochemical interactions with propanil on propanil-resistant barnyardgrass (Echinochloa crus-galli). Weed Sci. 47:1319.Google Scholar
Obojska, A., Lejczak, B., and Kubrak, M. 1999. Degradation of phosphonates by streptomycete isolates. Appl. Microbiol. Biotechnol. 51:872876.Google Scholar
Oyamada, M. and Kuwatsuka, S. 1989. Reduction mechanism of the nitro group of chlornitrofen, a diphenyl ether herbicide, in flooded soils. J. Pestic. Sci. 14:321327.Google Scholar
Padgette, S. R., Re, D. B., Barry, G. F., Eichholtz, D. E., Delannay, X., Fuchs, R. L., Kishore, G. M., and Fraley, R. T. 1996. New weed control opportunities: development of soybeans with a Roundup Ready gene. Pages 5384 In Duke, S. O., ed. Herbicide Resistant Crops. Boca Raton, FL: CRC Press.Google Scholar
Pallet, K. E., Little, J. P., Sheekey, M., and Veerasekaran, P. 1998. Mode of action of isoxaflutole I. Physiological effects of metabolism and selectivity. Pestic. Biochem. Physiol. 62:113124.Google Scholar
Pimental, D. and Levitan, L. 1986. Pesticides: amounts applied and amounts reaching pests. Biosciences 36:8691.Google Scholar
Plaisance, K. L. and Gronwald, J. W. 1999. Enhanced catalytic constant for glutathione S-transferase (atrazine) activity in an atrazine-resistant Abutilon theophrasti biotype. Pestic. Biochem. Physiol. 63:3449.Google Scholar
Pothuluri, J. V., Hinson, J. A., and Cerniglia, C. E. 1991. Propanil: toxicological characteristics, metabolism, and biodegradation potential in soil. J. Environ. Qual. 20:330347.Google Scholar
Prade, L., Huber, R., and Bieseler, B. 1998. Structures of herbicides in complex with their detoxifying enzyme glutathione S-transferase—explanations for the selectivity of the enzyme in plants. Structure 6:14451452.Google Scholar
Preston, C., Tardif, F. J., Christopher, J. T., and Powles, S. B. 1996. Multiple resistance to dissimilar herbicide chemistries in a biotype of Lolium rigidum due to enhanced activity of several herbicide degrading enzymes. Pestic. Biochem. Physiol. 54:123134.Google Scholar
Probst, G. W. and Tepe, J. B. 1969. Trifluralin and related compounds. Pages 255282 In Kearney, P. C. and Kaufman, D. D., eds. Degradation of Herbicides. New York: Marcel Dekker.Google Scholar
Quinn, J. P., Peden, J.M.M., and Dick, R. E. 1989. Carbon-phosphorus bond cleavage by gram-positive and gram-negative soil bacteria. Appl. Microbiol. Biotechnol. 31:283287.Google Scholar
Ramsey, R.J.L., Mena, F. L., and Stephenson, G. R. 2001. Effects of chemical safeners on herbicide action and metabolism in plants. Pages 310332 In Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M., ed. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. ACS Symposium Series 777. Washington, DC: American Chemical Society.Google Scholar
Reddy, G.V.B., Joshi, D. K., and Gold, M. H. 1993. Degradation of chlorophenoxyacetic acids by the lignin-degrading fungus Dichomitus squalens . Microbiology 143:23532360.Google Scholar
Reilley, K. A., Banks, M. K., and Schwab, A. P. 1996. Dissipation of polycyclic aromatic hydrocarbons in the rhizosphere. J. Environ. Qual. 25:212219.Google Scholar
Reinemer, P., Prade, L., Hof, P. et al. 1996. Three-dimensional structure of glutathione S-transferase from Arabidopsis thaliana at 2.2 resolution: structural characterization of herbicide-conjugating plant glutathione S-transferases and a novel active site architecture. J. Mol. Biol. 255:289309.Google Scholar
Rennenberg, H. 1982. Glutathione metabolism and possible biological roles in higher plants. Phytochemistry 21:27712781.Google Scholar
Rennenberg, H. and Brunold, C. 1994. Significance of glutathione metabolism in plants under stress. Prog. Bot. 55:142156.Google Scholar
Rheinwald, J. G., Chakrabarty, A. M., and Gunsalus, I. C. 1973. A transmissible plasmid controlling camphor oxidation in Pseudomonas putida . Proc. Natl. Acad. Sci. USA 70:885889.Google Scholar
Riley, P. S. and Behal, F. J. 1971. Amino acid-naphthylamide hydrolysis by Pseudomonas aeruginosa arylamidase. J. Bacteriol. 108:809816.Google Scholar
Robineau, T., Batard, Y., Nedelkina, S., Cabello-Hurtado, F., LeRet, M., Sorokine, O., Didierjean, L., and Werck-Reichhart, D. 1998. The chemically inducible plant cytochrome P450 CYP76B1 actively metabolizes phenylureas and other xenobiotics. Plant Physiol. 118:10491056.Google Scholar
Roden, E. E. and Wetzel, R. G. 1996. Organic carbon oxidation and suppression of methane production by microbial Fe(III) oxide reduction in vegetated and unvegetated freshwater wetland sediments. Limnol. Oceanogr. 41:17331748.Google Scholar
Romano, M. L., Stephenson, G. R., Tal, A., and Hall, J. C. 1993. The effect of monooxygenase and glutathione S-transferase inhibitors on the metabolism of diclofop-methyl and fenoxaprop-ethyl in barley and wheat. Pestic. Biochem. Physiol. 46:181189.Google Scholar
Rossjohn, J., Polekhina, G., Feil, S. C., Allocati, N., Masulli, M., De Illio, C., and Parker, M. W. 1998. A mixed disulfide bond in bacterial glutathione transferase: functional and evolutionary implications. Structure 6:721734.Google Scholar
Rozman, K. K. and Klaassen, C. D. 1996. Absorption, distribution, and excretion of toxicants. Pages 177183 In Casarett, L. and Doull, J., eds. Toxicology: The Basic Science of Poisons. New York: McGraw-Hill Health Professions Division.Google Scholar
Ruepple, M. L., Brighthwell, B. B., Schaefer, J., and Marvel, J. T. 1977. Metabolism and degradation of glyphosate in soil and water. J. Agric. Food Chem. 25:517528.Google Scholar
Rushmore, T. H. and Pickett, C. B. 1993. Glutathione S-transferases, structure, regulation, and therapeutic implications. J. Biol. Chem. 268:11 47511 478.Google Scholar
Rusness, D. G. and Lamoureux, G. L. 1980. Pentachloronitrobenzene fungicide metabolism in peanut. 2. Characterization of chloroform-soluble metabolites produced in vivo. J. Agric. Food Chem. 28:10701077.Google Scholar
Sadowsky, M. J., Tong, Z., de Souza, M., and Wackett, L. P. 1998. AtzC is a new member of the amidohydrolase protein superfamily and is homologous to other atrazine-metabolizing enzymes. J. Bacteriol. 180:152158.Google Scholar
Sadowsky, M. J. and Wackett, L. P. 2001. Genetics of atrazine and s-triazine degradation by Psedomonas sp. strain ADP and other bacteria. Pages 268282 In Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M., eds. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. ACS Symposium Series 777. Washington, DC: American Chemical Society.Google Scholar
Sanchez-Fernandez, R., Ardiles-Diaz, W., Van Montagu, M., Inze, D., and May, M. 1998. Cloning and expression analyses of AtMRP4, a novel MRP-like gene from Arabidopsis thaliana . J. Mol. Genet. 258:655662.Google Scholar
Sandermann, H. Jr., Arjmand, M., Gennity, I., Winkler, R., Stuble, C. B., and Aschbacher, P. W. 1990. Animal bioavailability of defined xenobiotic lignin metabolites. J. Agric. Food Chem. 38:18771880.Google Scholar
Sandermann, H. Jr., Haas, M., Messner, B., Pflumacher, S., Schroder, P., and Wetzel, A. 1997. The role of glucosyl and malonyl conjugation in herbicide selectivity. Pages 211231 In Hatzios, K. K., ed. Regulation of Enzymatic Systems Detoxifying Xenobiotics in Plants. NATO ASI Series. Dordrecht, The Netherlands: Kluwer Academic Publishers.Google Scholar
Sandermann, H. Jr., Hertkorn, N., May, R. G., and Lange, B. M. 2001. Bound pesticidal residues in crop plants: chemistry, bioavailability, and toxicology. Pages 119128 In Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M., eds. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. ACS Symposium Series 777. Washington, DC: American Chemical Society.Google Scholar
Sandermann, H. Jr., Scheel, D., and Trenck, T.v.d. 1983. Metabolism of environmental chemicals by plants—copolymerization into lignin. J. Appl. Polym. Sci.: Appl. Polym. Symp. 37:407420.Google Scholar
Sarkar, J. M., Malcolm, R. L., and Bollag, J.- M. 1988. Enzymatic coupling of 2,4-dichlorophenol to stream fulvic acid in the presence of oxidoreductases. Soil Sci. Soc. Am. J. 52:688694.Google Scholar
Schmidt, B. 2001. Metabolic profiling using plant cell suspension cultures. Pages 4056 In Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M., eds. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. ACS Symposium Series 777. Washington, DC: American Chemical Society.Google Scholar
Schocken, M. J. 2001. In vitro methods in metabolism and environmental fate studies. Pages 3039 In Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M., eds. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. ACS Symposium Series 777. Washington, DC: American Chemical Society.Google Scholar
Schocken, M. J., Mao, J., and Schabacker, D. J. 1997. Microbial transformations of the fungicide cyprodinil (CGA-219417). J. Agric. Food Chem. 45:36473651.Google Scholar
Schröder, P. 1997. Fate of glutathione S-conjugation in plants: degradation of gluathione moiety. Pages 233244 In Hatzios, K. K., ed. Regulation of Enzymatic Systems Detoxifying Xenobiotics in Plants. NATO ASI Series. Dordrecht, The Netherlands: Kluwer Academic Publishers.Google Scholar
Schwab, A. P., Banks, M. K., and Arunachalam, M. 1995. Biodegradation of polycyclic aromatic hydrocarbons in rhizosphere soil. Pages 2329 In Hinchee, R. E., Hoeppel, R. E., and Anderson, D. B., eds. Bioremediation of Recalcitrant Organics. Columbus, OH: Battelle Memorial Institute.Google Scholar
Shaner, D. L. and Mallipudi, M. 1991. Mechanisms of selectivity of the imidazolinone herbicides. Pages 91102 In Shaner, D. L. and O’Connor, S. L., eds. The Imidazolinone Herbicides. Boca Raton, FL: CRC Press.Google Scholar
Shaner, D. L. and Tecle, B. 2001. Designing herbicide tolerance based on metabolic alteration: the challenges and the future. Pages 353374 In Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M., eds. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. ACS Symposium Series 777. Washington, DC: American Chemical Society.Google Scholar
Sharma, Y. K. and Davis, K. R. 1994. Ozone-induced expression of stressrelated genes in Arabidopsis thaliana . Plant Physiol. 105:10891096.Google Scholar
Shields, M. S., Reagin, M. J., Gerger, R. R., Campbell, R., and Somerville, R. 1995. TOM, a new aromatic degradative plasmid from Burkholderia (Pseudomonas) cepacia G4. Appl. Environ. Microbiol. 61:13521356.Google Scholar
Shimabukuro, R. H. 1985. Detoxification of herbicides. Pages 215240 In Duke, S. O., ed. Weed Physiology. Volume 2. Boca Raton, FL: CRC Press.Google Scholar
Shiota, N., Inui, H., and Okhawa, H. 1996. Metabolism of the herbicide chlortoluron in transgenic tobacco plants expressing the fused enzyme between rat cytochrome P4501A1 and yeast NADPH-cytochrome P450 oxidoreductase. Pestic. Biochem. Physiol. 54:190198.CrossRefGoogle Scholar
Siminszky, B., Corbin, F. T., Ward, E. R., Fleischmann, T. J., and Dewey, R. E. 1999. Expression of a soybean cytochrome P450 monooxygenase cDNA in yeast and tobacco enhances the metabolism of phenylurea herbicides. Proc. Natl. Acad. Sci. USA 96:17501755.Google Scholar
Skidmore, M. W., Paulson, G. D., Kuiper, H. A., Ohlin, B., and Reynolds, S. 1998. Bound xenobiotic residues in food commodities of plant and animal origin. Pure Appl. Chem. 70:14231447.Google Scholar
Smith, A. E. and Aubin, A. J. 1990. Degradation studies with 14C-fenoxaprop in Prairie soils. Can. J. Soil Sci. 70:343350.Google Scholar
Smith, A. E., Phatak, S. C., and Emmatty, D. A. 1989. Metribuzin metabolism by tomato cultivars with low, medium, and high levels of tolerance to metribuzin. Pestic. Biochem. Physiol. 35:284290.Google Scholar
Sobera, M., Wieczorek, P., Lejczak, B., and Kafarski, P. 1997. Organophosphonate utilization by the wild-type strain Chladosporum resinae . Toxicol. Environ. Chem. 61:229235.Google Scholar
Sommer, A. and Böger, P. 1999. Characterization of recombinant corn glutathione S-transferase isoforms I, II, III, and IV. Pestic. Biochem. Physiol. 63:127138.Google Scholar
Sommer, A. and Böger, P. 2001. Enzymological studies on recombinant isoforms of glutathione s-transferase from corn. Pages 253267 In Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M., eds. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. ACS Symposium Series 777. Washington, DC: American Chemical Society.Google Scholar
Sprankle, P., Meggitt, W. F., and Penner, D. 1975. Absorption, mobility, and microbial degradation of glyphosate in the soil. Weed Sci. 23:229234.Google Scholar
Stalker, D. M., Kiser, J. A., Baldwin, G., Coulombe, B., and Houck, C. M. 1996. Cotton weed control using the BXN system. Pages 93105 In Duke, S. O., ed. Herbicide Resistant Crops. Boca Raton, FL: CRC Press.Google Scholar
Stalker, D. M., Malyj, L. B., and McBride, K. E. 1988a. Purification and properties of a nitrilase specific for the herbicide bromoxynil and corresponding nucleotide sequence analysis of the bxn gene. J. Biol. Chem. 263:63106314.Google Scholar
Stalker, D. M. and McBride, K. E. 1987. Cloning and expression in Escherichia coli of a Klebsiella ozaenae plasmid-borne gene encoding a nitrilase specific for the herbicide bromoxynil. J. Bacteriol. 169:955960.Google Scholar
Stalker, D. M., McBride, K. E., and Malyj, L. B. 1988b. Herbicide resistance in transgenic plants expressing a bacterial detoxification gene. Science 242:419423.Google Scholar
Steiert, J. G. and Crawford, R. L. 1986. Catabolism of pentachlorophenol by a Flavobacterium sp. Biochem. Biophys. Res. Commun. 141:825830.Google Scholar
Steinkamp, R. and Rennenberg, H. 1985. Degradation of glutathione in plant cells: evidence against the participation of a γ-glutamyltranspeptidase. Z. Naturforsch. 40c:2933.Google Scholar
Stephenson, G. R., Bunce, N. J., Makowski, R. I., Bergsma, M. D., and Curry, J. C. 1979. Structure-activity relationships for thiocarbamate herbicides in corn. J. Agric. Food Chem. 27:543547.Google Scholar
Stephenson, G. R., Bunce, N. J., Makowski, R. I., and Curry, J. C. 1978. Structure-activity relationships for S-ethyl N,N-dipropylthiocarbamate (EPTC) antidotes in corn. J. Agric. Food Chem. 26:137140.Google Scholar
Stephenson, G. R., Tal, A., Vincent, N. A., and Hall, J. C. 1993. Interactions of fenoxaprop-ethyl with fenchlorazole-ethyl in annual grasses. Weed Technol. 7:163168.Google Scholar
Still, G. G. and Mansager, E. R. 1972. Aryl hydroxylation of isopropyl-3-chlorocarbanilate by soybean plants. Phytochemistry 11:515520.Google Scholar
Still, G. G. and Mansager, E. R. 1973. Soybean shoot metabolism of isopropyl-3-chlorocarbanilate: ortho and para aryl hydroxylation. Pestic. Biochem. Physiol. 3:8795.Google Scholar
Stomp, A. M., Han, K. H., Wilbert, S., and Gordon, M. P. 1993. Genetic improvement of tree species for remediation of hazardous wastes. In Vitro Cell. Dev. Biol. Plant 29:227232.Google Scholar
Streber, W. R., Timmis, K. N., and Zenk, M. H. 1987. Analysis, cloning, and high-level expression of 2,4-dichlorophenoxyacetate monooxygenase gene tfdA of Alcaligenes eutrophus JMP134. J. Bacteriol. 169:29502955.Google Scholar
Suflita, J. M., Loll, M. J., Snipes, W. C., and Bollag, J.- M. 1981. Electron spin resonance study of free radicals generated by a soil extract. Soil Sci. 131:145150.Google Scholar
Suzuki, T. J. 1983. Methylation and hydroxylation of pentachlorophenol by Mycobacterium sp. isolated from soil. J. Pestic. Sci. 8:419428.Google Scholar
Tal, J. A., Hall, J. C., and Stephenson, G. R. 1995. Non-enzymatic conjugation of fenoxaprop-ethyl with glutathione and cysteine in several grass species. Weed Res. 35:133139.Google Scholar
Tal, J. A., Romano, M. L., Stephenson, G. R., Schwan, A. L., and Hall, J. C. 1993. Glutathione conjugation: a detoxification pathway for fenoxaprop-ethyl in barley, crabgrass, oat and wheat. Pestic. Biochem. Physiol. 46:190199.Google Scholar
Tate, R. L. and Alexander, M. 1974. Formation of dimethylamine and diethylamine in soil treated with pesticides. Soil Sci. 118:317321.Google Scholar
Taylor, J. L., Fritzemeier, K.-H., Häuser, I., Kombrink, E., Rohwer, F., Schröder, M., Strittmatter, G., and Hahlbrock, K. 1990. Structural analysis and activation by fungal infection of a gene encoding a pathogenesisrelated protein in potato. Mol. Plant-Microbe Interact. 3:7277.Google Scholar
Tebbe, C. C. and Reber, H. H. 1988. Utilization of the herbicide phosphinothricin as a nitrogen source by soil bacteria. Appl. Microbiol. Biotechnol. 29:103105.Google Scholar
Tecle, B., Da Cunha, A., and Shaner, D. L. 1993. Differential routes of metabolism of imidazolinones: basis for soybean (Glycine max) selectivity. Pestic. Biochem. Physiol. 46:120130.Google Scholar
Tenhaken, R., Levine, A., Brisson, L. F., Dixon, R. A., and Lamb, C. 1995. Function of the oxidative burst in hypersensitive disease resistance. Proc. Natl. Acad. Sci. USA 92:41584163.Google Scholar
Ternan, N. G., McGrath, J. W., McMullan, G., and Quinn, J. P. 1998. Review: organophosphonates: occurrence, synthesis and biodegradation by microorganisms. World J. Microbiol. Biotechnol. 14:635647.Google Scholar
Thiessen, E. P. 1978. Barban Plus Naphthalic Anhydride for the Selective Control of Wild Oats in Oats. . University of Guelph, Guelph, ON, Canada.Google Scholar
Timmerman, K. P. 1989. Molecular characterization of corn glutathione S-transferase isozymes involved in herbicide detoxification. Plant Physiol. 77:323342.Google Scholar
Tong, Z., Board, P. G., and Anders, M. W. 1998a. Glutathione transferase zeta catalyses the oxygenation of the carcinogen dichloroacetic acid to glyoxylic acid. Biochem. J. 371:371374.Google Scholar
Tong, Z., Board, P. G., and Anders, M. W. 1998b. Glutathione transferase zeta-catalyzed biotransformation of dichloroacetic acid and other alpha-haloacids. Chem. Res. Toxicol. 11:13321338.CrossRefGoogle ScholarPubMed
Torstensson, N.T.L. and Aamisepp, A. 1977. Detoxification of glyphosate in soil. Weed Res. 17:209211.Google Scholar
Trower, M. K., Sariaslani, F. S., and Kitson, F. G. 1988. Xenobiotic oxidation by cytochrome P-450-enriched extracts of Streptomyces griseus . Biochem. Biophys. Res. Commun. 157:14171422.Google Scholar
Tsuchida, S. and Sato, K. 1992. Glutathione transferases and cancer. CRC Crit. Rev. Biochem. Mol. Biol. 27:337384.Google Scholar
Tweedy, B. G., Loeppy, C., and Ross, J. A. 1970. Metabolism of 3-(pbromophenyl)-1-methoxy-1-methylurea (metobromuron) by selected soil microorganisms. Science 168:482483.Google Scholar
Valli, K. and Gold, M. H. 1991. Degradation of 2,4-dichlorophenol by the lignin-degrading fungus Phanerochaete chrysosporium . J. Bacteriol. 173:345352.Google Scholar
van den Brink, H.J.M., van Gorcom, R.F.M., van den Hondel, C.A.M., and Punt, P. J. 1998. Cytochrome P450 enzyme systems in fungi. Fungal Genet. Biol. 23:117.Google Scholar
van den Tweel, W.J.J., van der Kok, J. B., and de Bont, J.A.M. 1987. Reductive dechlorination of 2,4-dichlorobenzoate to 4-chlorobenzoate and hydrolytic dehalogenation of 4-chloro-, 4-bromo, and 4-iodobenzoate by Alcaligenes denitrificans NTB-1. Appl. Environ. Microbiol. 53:810815.Google Scholar
van der Krol, D., Schuphan, I., Thiede, B., and Schmidt, B. 1995. Metabolism of [ring-2,6-14C]parathion in plant cell suspension cultures of carrot (Daucus carota), purple foxglove (Digitalis purpurea), soybean, thorn apple (Datura stramonium) and wheat (Triticum aestivum). Pestic. Sci. 45:143152.Google Scholar
van Hylckama, J.E.T., Kingma, J., Kruizinga, W., and Janssen, D. B. 1999. Purification of a glutathione S-transferase and a glutathione conjugate-specific dehydrogenase involved in isoprene metabolism in Rhodococcus sp. strain AD45. J. Bacteriol. 181:20942101.Google Scholar
van Hylckama, J.E.T., Kingma, J., van den Wijngaard, A. J., and Janssen, D. B. 1998. A glutathione S-transferase with activity towards cis- 1,2-dichloroepoxyethane is involved in isoprene utilization by Rhodococcus sp. strain AD45. Appl. Environ. Microbiol. 64:28002805.Google Scholar
Vuilleumier, S. 2001. Bacterial glutathione S-transferases and the detoxification of xenobiotics: dehalogenation through glutathione conjugation and beyond. Pages 240252 In Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M., eds. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. ACS Symposium Series 777. Washington, DC: American Chemical Society.Google Scholar
Vuilleumier, S., Gisi, D., Stumpp, M. T., and Leisinger, T. 1999. Bacterial dichloromethane dehalogenase: a particular brand of glutathione S-transferases. Clin. Chem. Enzymol. Commun. 8:367378.Google Scholar
Vuilleumier, S. and Leisinger, T. 1996. Protein engineering studies of dichloromethane dehalogenase/glutathione S-transferase from Methylophilus sp. strain DM11 Ser12 but not Tyr6 is required for enzyme activity. Eur. J. Biochem. 239:410417.Google Scholar
Vuilleumier, S., Sorribas, H., and Leisinger, T. 1997. Identification of a novel determinant of glutathione affinity in dichloromethane dehalogenases/glutathione S-transferases. Biochem. Biophys. Res. Commun. 238:452456.Google Scholar
Wackett, L. P., Wanner, B. L., Venditti, C. P., and Walsh, C. T. 1987. Involvement of the phosphate regulon and the psiD locus in carbon-phosphorus lyase activity of Escherichia coli K-12. J. Bacteriol. 169:17531756.Google Scholar
Walton, B. T., Hoylman, A. M., Perez, M. M., Anderson, T. A., Johnson, T. R., Guthrie, E. A., and Christman, R. F. 1994. Rhizosphere microbial communities as a plant defense against toxic substances in soils. Pages 8292 In Anderson, T. A. and Coats, J. R., eds. Bioremediation through Rhizosphere Technology. ACS Symposium Series 563. Washington, DC: American Chemical Society.Google Scholar
Wang, T. G. and Peverly, J. H. 1999. Iron oxidation states on root surfaces of a wetland plant (Phragmites australis). Soil Sci. Soc. Am. J. 63:247252.Google Scholar
Wanner, B. L. and Boline, J. A. 1990. Mapping and molecular cloning of the phn (psiD) locus for phosphonate utilization in Escherichia coli . J. Bacteriol. 172:11861196.Google Scholar
Wanner, B. L. and McSharry, R. 1982. Phosphate-controlled gene expression in Escherichia coli K12 using Mudl-directed lacZ fusion. J. Mol. Biol. 158:347363.Google Scholar
Wanner, B. L. and Metcalf, W. W. 1992. Molecular genetic studies of a 10.9-kb operon in Escherichia coli for phosphonate uptake and biodegradation. FEMS Microbiol. Lett. 100:133140.Google Scholar
Werwath, J., Arfmann, H.-A., Piepers, D. H., Timmis, K. H., and Wittich, R.- M. 1998. Biochemical and genetic characterization of a gentisate 1,2-dioxygenase from Sphingomonas sp. strain RW5. J. Bacteriol. 180:41714176.Google Scholar
Williams, P. A. and Sayers, J. R. 1994. The evolution of pathways for aromatic hydrocarbon oxidation in Pseudomonas . Biodegradation 5:195217.Google Scholar
Wu, J., Cramer, C. L., and Hatzios, K. K. 1999. Characterization of two cDNAs encoding glutathione S-transferases in rice and induction of their transcripts by the herbicide safener fenclorim. Physiol. Plant. 105:102108.Google Scholar
Yaacoby, T., Hall, J. C., and Stephenson, G. R. 1991. Influence of fenchlorazole-ethyl on the metabolism of fenoxaprop-ethyl in wheat barley and crabgrass. Pestic. Biochem. Physiol. 41:296304.Google Scholar
Yee, D. C., Maynard, J. A., and Wood, T. K. 1998. Rhizoremediation of trichloroethylene by a recombinant, root-colonizing Pseudomonas fluorescens strain expressing toluene ortho-monooxygenase constitutively. Appl. Environ. Microbiol. 64:112118.Google Scholar
Yenne, S. P. and Hatzios, K. K. 1990. Molecular comparisons of selected herbicides and their safeners by computer-aided molecular modeling. J. Agric. Food Chem. 38:19501956.Google Scholar
Yenne, S. P., Hatzios, K. K., and Meredith, S. A. 1990. Uptake, translocation, and metabolism of oxabetrinil and CGA-133205 in grain sorghum (Sorghum bicolor) and their influence on metolachlor metabolism. J. Agric. Food Chem. 38:19571961.Google Scholar
Yoshioka, H., Nagasawa, T., and Yamada, H. 1991. Purification and characterization of aryl acylamidase from Nocardia globerula . Eur. J. Biochem. 199:1724.Google Scholar
Zablotowicz, R. M., Hoagland, R. E., Lee, H., Alber, T., Trevors, J. T., Hall, J. C., and Locke, M. A. 2001. Transformation of nitroaromatic pesticides and related xenobiotics by microorganisms and plants. Pages 194216 In Hall, J. C., Hoagland, R. E., and Zablotowicz, R. M., eds. Pesticide Biotransformation in Plants and Microorganisms: Similarities and Divergences. ACS Symposium Series 777. Washington, DC: American Chemical Society.Google Scholar
Zablotowicz, R. M., Hoagland, R. E., and Locke, M. A. 1994. Glutathione S-transferase activity in rhizosphere bacteria and the potential for herbicide detoxification. Pages 184198 In Anderson, T. A. and Coats, J. R., eds. Bioremediation through Rhizosphere Technology. ACS Symposium Series 563. Washington, DC: American Chemical Society.Google Scholar
Zablotowicz, R. M., Hoagland, R. E., Locke, M. A., and Hickey, W. J. 1995. Glutathione-S-transferase activity and metabolism of glutathione conjugates by rhizosphere bacteria. Appl. Environ. Microbiol. 61:10541060.Google Scholar
Zablotowicz, R. M., Hoagland, R. E., Staddon, W. J., and Locke, M. A. 2000. Effects of pH on chemical stability and de-esterification of fenoxaprop-ethyl by purified enzymes, bacterial extracts, and soils. J. Agric. Food Chem. 48:47114716.Google Scholar
Zablotowicz, R. M., Leung, K. T., Alber, T., Cassidy, M. B., Trevors, J. T., Lee, H., Veldhuis, L., and Hall, J. C. 1999. Degradation of 2,4-dinitrophenol and selected nitroaromatic compounds by Sphingomonas sp. UG30. Can. J. Microbiol. 45:840848.Google Scholar
Zablotowicz, R. M., Locke, M. A., and Hoagland, R. E. 1997. Aromatic nitroreduction of acifluorfen in soils, rhizospheres, and pure cultures of rhizobacteria. Pages 3853 In Kruger, E. L., Anderson, T. A., and Coats, J. R., eds. Phytoremediation of Soil and Water Contaminants. ACS Symposium Series 664. Washington, DC: American Chemical Society.Google Scholar
Zajc, A., Neuefeind, T., Prade, L., Reinemer, P., Huber, R., and Bieseler, B. 1999. Herbicide detoxification by glutathione S-transferases as implicated from X-ray structures. Pestic. Sci. 55:248252.Google Scholar
Zama, P. and Hatzios, K. K. 1986. Effects of CGA-92194 on the chemical reactivity of metolachlor with glutathione and metabolism of metolachlor in grain sorghum (Sorghum bicolour). Weed Sci. 34:834841.Google Scholar
Zaranyika, M. F. and Nyandoro, M. G. 1993. Degradation of glyphosate in the aquatic environment: an enzymatic kinetic model that takes into account microbial degradation of both free and colloidal (or sediment) particle adsorbed glyphosate. J. Agric. Food Chem. 41:838842.Google Scholar
Zboinska, E., Maliszewska, I., Lejczak, B., and Kafarski, P. 1992. Degradation of organophosphonates by Penicillium citrinum . Lett. Appl. Microbiol. 15:269272.Google Scholar