Hostname: page-component-8448b6f56d-c4f8m Total loading time: 0 Render date: 2024-04-23T23:36:29.501Z Has data issue: false hasContentIssue false

Herbicides as Probes in Plant Biology

Published online by Cambridge University Press:  20 January 2017

Franck E. Dayan*
Affiliation:
U.S. Department of Agriculture–Agricultural Research Service, Natural Products Utilization Research Unit, P.O. Box 8048, University, MS 38677
Stephen O. Duke
Affiliation:
U.S. Department of Agriculture–Agricultural Research Service, Natural Products Utilization Research Unit, P.O. Box 8048, University, MS 38677
Klaus Grossmann
Affiliation:
BASF Agricultural Center Limburgerhof, Speyerer Strasse, D-67117 Limburgerhof, Germany
*
Corresponding author's E-mail: fdayan@olemiss.edu
Rights & Permissions [Opens in a new window]

Abstract

Core share and HTML view are not available for this content. However, as you have access to this content, a full PDF is available via the ‘Save PDF’ action button.

Herbicides are small molecules that inhibit specific molecular target sites within plant biochemical pathways and/or physiological processes. Inhibition of these sites often has catastrophic consequences that are lethal to plants. The affinity of these compounds for their respective target sites makes them useful tools to study and dissect the intricacies of plant biochemical and physiological processes. For instance, elucidation of the photosynthetic electron transport chain was achieved in part by the use of herbicides, such as terbutryn and paraquat, which act on photosystem II and I, respectively, as physiological probes. Work stemming from the discovery of the binding site of PS II–inhibiting herbicides was ultimately awarded the Nobel Prize in 1988. Although not as prestigious as the seminal work on photosynthesis, our knowledge of many other plant processes expanded significantly through the ingenious use of inhibitors as molecular probes. Examples highlight the critical role played by herbicides in expanding our understanding of the fundamental aspects of the synthesis of porphyrins and the nonmevalonate pathway, the evolution of acetyl-coenzyme A carboxylase, cell wall physiology, the functions of microtubules and the cell cycle, the role of auxin and cyanide, the importance of subcellular protein targeting, and the development of selectable markers.

Type
Symposium
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution license (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
Copyright © Weed Science Society of America

References

Literature Cited

Abdallah, I., Fischer, A. J., Elmore, C. L., Saltveit, M. E., and Zaki, M. 2006. Mechanism of resistance to quinclorac in smooth crabgrass (Digitaria ischaemum). Pestic. Biochem. Physiol. 84:3848.Google Scholar
Abell, L. M. 1996. Biochemical approaches to herbicide discovery: advances in enzyme target identification and inhibitor design. Weed Sci. 44:734742.Google Scholar
Alban, C., Baldet, P., and Douce, R. 1994. Localization and characterization of two structurally different forms of acetyl-CoA carboxylase in young pea leaves, of which one is sensitive to aryloxyphenoxypropionate herbicides. Biochem. J. 300:557565.Google Scholar
Anthony, R. G. and Hussey, P. J. 1999. Dinitroaniline herbicide resistance and the microtubule cytoskeleton. Trends Plant Sci. 4:112116.Google Scholar
Argueso, C. T., Hansen, M., and Kieber, J. J. 2007. Regulation of ethylene biosynthesis. J. Plant Growth Regul. 26:92105.Google Scholar
Arias, R. S., Dayan, F. E., Michel, A., and Howell, J. L. 2006. Characterization of a higher plant herbicide-resistant phytoene desaturase and its use as a selectable marker. Plant Biotechnol. J. 4:263273.Google Scholar
Armstrong, J. I. 2007. Chemical genetics: catalysing pathway exploration and new target discovery. J. Sci. Food Agric. 87:19851990.Google Scholar
Badescu, G. O. and Napier, R. M. 2006. Receptors for auxin: will it all end in TIRs? Trends Plant Sci. 11:217223.Google Scholar
Bajguz, A. and Asami, T. 2004. Effects of brazzinoazole, an inhibitor of brassinosteroid biosynthesis, on light- and dark-grown Chlorella vulgaria . Planta. 218:869877.Google Scholar
Bajguz, A. and Asami, T. 2005. Suppression of Wolffia arrhiza growth by brazzinoazole, an inhibitor of brassinosteroid biosynthesis and its restoration by endogenous 24-epibrassinolide. Phytochemistry. 66:17871796.Google Scholar
Becerril, J. M., Duke, M. V., Nandihalli, U. B., Matsumoto, H., and Duke, S. O. 1992. Light control of porphyrin accumulation in acifluorfen-methyl treated Lemna pausicostata . Physiol. Plant. 86:616.Google Scholar
Blackwell, H. E. and Zhao, Y. D. 2003. Chemical genetic approaches to plant biology. Plant Physiol. 133:448455.Google Scholar
Blumenthal, S. G., Hendrickson, Y. P., Abrol, Y. P., and Conn, E. E. 1968. Cyanide metabolism in higher plants, III: the biosynthesis of β-cyanoalanine. J. Biol. Chem. 243:53025307.Google Scholar
Braun, N., Wyrzykowska, J., Muller, P., David, K., Couch, D., Perrot-Rechenmann, C., and Fleming, A. J. 2008. Conditional repression of auxin binding protein-1 reveals that it coordinates cell division and cell expansion during postembryonic shoot development in Arabidopsis and tobacco. Plant Cell. 20:27462762.Google Scholar
Brettel, K. 1997. Electron transfer and arrangement of the redox cofactors in photosystem I. Biochim. Biophys. Acta. 1318:322373.Google Scholar
Chitnis, P. R. 1996. Photosystem I. Plant Physiol. 111:661669.Google Scholar
Christian, M., Hannah, W. B., Luthen, H., and Jones, A. M. 2008. Identification of auxins by a chemical genomics approach. J. Exp. Bot. 59:27572767.Google Scholar
Cole, D., Pallett, K., and Rodgers, M. 2000. Discovering new modes of action for herbicides and the impact of genomics. Pestic. Outlook. 11:223229.Google Scholar
Comai, L., Sen, L. C., and Stalker, D. M. 1983. An altered aroA gene product confers resistance to the herbicide glyphosate. Science. 22:370371.Google Scholar
Corradi, H. R., Corrigall, A. V., Boix, E., Mohan, C. G., Sturrock, E. D., Meissner, P. N., and Acharya, K. R. 2006. Crystal structure of protoporphyrinogen oxidase from Myxococcus xanthus and its complex with the inhibitor acifluorfen. J. Biol. Chem. 281:3862538633.Google Scholar
Croteau, R. B. 1992. Clomazone does not inhibit the conversion of isopentenyl pyrophosphate to geranyl, farnesyl, or geranylgeranyl pyrophosphate in vitro. Plant Physiol. 98:15151517.Google Scholar
Dayan, F. E., Daga, P. R., Duke, S. O., Lee, R. M., Tranel, P. J., and Doerksen, R. J. Biochemical and structural consequences of a glycine deletion in the helix of protoporphyrinogen oxidase. Biochim. Biophysic. Acta. DOI: 10.1016/j.bbapap.2010.04.004.Google Scholar
Dayan, F. E. and Duke, S. O. 1997. Phytotoxicity of protoporphyrinogen oxidase inhibitors: Phenomenology, mode of action and mechanisms of resistance. Pages 1135. in Roe, R. M., Burton, J. D., and Kuhr, R. J. eds. Herbicide Activity: Toxicology, Biochemistry and Molecular Biology. Washington, DC: IOS.Google Scholar
Dayan, F. E. and Duke, S. O. 2003. Herbicides: Protoporphyrinogen oxidase inhibitors. Pages 850863. in Plimmer, J. R., Gammon, D. W., and Ragsdale, N. N. eds. Encyclopedia of Agrochemicals. New York: J. Wiley.Google Scholar
DeBolt, S., Gutierrez, R., Ehrhardt, D. W., Melo, C. V., Ross, L., Cutler, S. R., Somerville, C., and Bonetta, D. 2007a. Morlin, an inhibitor of cortical microtubule dynamics and cellulose synthase movement. Proc. Nat. Acad. Sci. U. S. A. 104:58545859.Google Scholar
DeBolt, S., Gutierrez, R., Ehrhardt, D. W., and Somerville, C. 2007b. Nonmotile cellulose synthase subunits repeatedly accumulate within localized regions at the plasma membrane in Arabidopsis hypocotyl cells following 2,6-dichlorobenzonitrile treatment. Plant Physiol. 145:334338.Google Scholar
Della-Cioppa, G., Bauer, S. C., Klein, B. K., Shah, D. M., Fraley, R. T., and Kishore, G. M. 1986. Translocation of the precursor of 5-enolpyruvylshikimate-3-phosphate synthase into chloroplasts of higher plants in vitro. Proc. Nat. Acad. Sci. U. S. A. 83:68736877.Google Scholar
Della-Cioppa, G., Bauer, S. C., Taylor, M. L., Rochester, D. E., Klein, B. K., Shah, D. M., Fraley, R. T., and Kishore, G. M. 1987. Targeting a herbicide-resistant enzyme from Escherichia coli to chloroplasts of higher plants. Bio/Technology. 5:579584.Google Scholar
Della-Cioppa, G. and Kishore, G. M. 1988. Import of a precursor protein into chloroplasts is inhibited by the herbicide glyphosate. EMBO J. 7:12991305.Google Scholar
Délye, C., Straub, C., Michel, S., and Le Corre, V. 2004. Nucleotide variability at the acetyl coenzyme A carboxylase gene and the signature of herbicide selection in the grass weed Alopecurus myosuroides (Huds.). Mol. Biol. Evol. 21:884892.Google Scholar
Dharmasiri, N., Dharmasiri, S., and Estelle, M. 2005. The F-box protein TIR1 is an auxin receptor. Nature. 435:441445.Google Scholar
Dostatni, R., Meyer, H. E., and Oettmeier, W. 1988. Mapping of 2 tyrosine residues involved in the quinone-(QB) binding-site of the D-1 reaction center polypeptide of photosystem-II. FEBS Lett. 239:207210.Google Scholar
Draber, W., Tietjen, K., Kluth, J. F., and Trebst, A. 1991. Herbicides in photosynthesis research. Angew. Chem. Int. Ed. Engl. 30:16211633.Google Scholar
Draber, W., Trebst, A., and Harth, E. 1970. On a new inhibitor of photosynthetic electron-transport in isolated chloroplasts. Z. Naturforsch. 25b:11571159.Google Scholar
Duke, M. V. and Duke, S. O. 1997. Analysis and manipulation of the chlorophyll pathway in higher plants. Pages 229241. in Daschek, W. V. ed. Methods in Plant Biochemistry and Molecular Biology. Boca Raton, FL: CRC.Google Scholar
Duke, S. O., Kenyon, W. H., and Paul, R. N. 1985. FMC 57020 effects on chloroplast development in pitted morningglory (Ipomoea lacunosa) cotyledons. Weed Sci. 33:786794.Google Scholar
Duke, S. O., Paul, R. N., Becerril, J. M., and Schmidt, J. H. 1991. Clomazone causes accumulation of sesquiterpenoids in cotton (Gossypium hirsutum L.). Weed Sci. 39:339346.Google Scholar
Eisenreich, W., Bacher, A., Arigoni, D., and Rohdich, F. 2004. Biosynthesis of isoprenoids via the non-mevalonate pathway. Cell. Mol. Life Sci. 61:14011426.Google Scholar
Eisenreich, W., Rohdich, F., and Bacher, A. 2001. Deoxyxylulose phosphate pathway to terpenoids. Trends Plant Sci. 6:7884.Google Scholar
Eisenreich, W., Schwarz, M., Cartayrade, A., Arigoni, D., Zenk, M. H., and Bacher, A. 1998. The deoxyxylulose phosphate pathway of terpenoid biosynthesis in plants and microorganisms. Chem. Biol. 5:R221R233.Google Scholar
Ferhatoglu, Y. and Barrett, M. 2006. Studies of clomazone mode of action. Pestic. Biochem. Physiol. 85:714.Google Scholar
Fisher, D. D. and Cyr, R. J. 1998. Extending the microtubule/microfibril paradigm: cellulose synthesis is required for normal cortical microtubule alignment in elongating cells. Plant Physiol. 116:10431051.Google Scholar
Gershenzon, J. and Croteau, R. 1991. Terpenoids. Pages 165209. in Rosenthal, G. A. and Berenbaum, M. R. eds. Herbivores: Their Interactions with Secondary Plant Metabolites. New York: Academic.Google Scholar
Grossmann, K. 1996. A role for cyanide, derived from ethylene biosynthesis, in the development of stress symptoms. Physiol. Plant. 97:772775.Google Scholar
Grossmann, K. 1998. Quinclorac belongs to a new class of highly selective auxin herbicides. Weed Sci. 46:707716.Google Scholar
Grossmann, K. 2000a. The mode of action of auxin herbicides: a new ending to a long, drawn out story. Trends Plant Sci. 5:506508.Google Scholar
Grossmann, K. 2000b. The mode of action of quinclorac: a case study of a new auxin-type herbicide. Pages 181214. in Cobb, A. H. and Kirkwood, R. C. eds. Herbicides and Their Mechanisms of Action. Sheffield, UK: Sheffield Academic.Google Scholar
Grossmann, K. 2003a. Mediation of herbicide effects by hormone interactions. J. Plant Growth Regul. 22:109122.Google Scholar
Grossmann, K. 2003b. News from old compounds: the mode of action of auxin herbicides. Pages 131142. in Voss, G. and Ramos, G. eds. Chemistry of Crop Protection. Progress and Prospects in Science and Regulation. Weinheim, Germany: Wiley-VCH.Google Scholar
Grossmann, K. and Kwiatkowski, J. 1993. Selective induction of ethylene and cyanide biosynthesis appears to be involved in the selectivity of the herbicide quinclorac between rice and barnyardgrass. J. Plant Physiol. 142:457466.Google Scholar
Grossmann, K. and Kwiatkowski, J. 1995. Evidence for a causative role of cyanide, derived from ethylene biosynthesis, in the herbicidal mode of action of quinclorac in barnyard grass. Pestic. Biochem. Physiol. 51:150160.Google Scholar
Grossmann, K. and Kwiatkowski, J. 2000. The mechanism of quinclorac selectivity in grasses. Pestic. Biochem. Physiol. 66:8391.Google Scholar
Grossmann, K. and Scheltrup, F. 1997. Selective induction of 1-aminocyclopropane-1-carboxylic acid (ACC) synthase activity is involved in the selectivity of the auxin herbicide quinclorac between barnyard grass and rice. Pestic. Biochem. Physiol. 58:145153.Google Scholar
Guilfoyle, T. 2007. Sticking with auxin. Nature. 446:621622.Google Scholar
Hagen, G. and Guilfoyle, T. 2002. Auxin-responsive gene expression: genes, promoters and regulatory factors. Plant Mol. Biol. 49:373385.Google Scholar
Hall, R. D. 2006. Plant metabolomics: from holistic hope, to hype, to hot topic. New Phytol. 169:453468.Google Scholar
Hansen, H. and Grossmann, K. 2000. Auxin-induced ethylene triggers abscisic acid biosynthesis and growth inhibition. Plant Physiol. 124:14371448.Google Scholar
Hemmerlin, A., Tritsch, D., Hartmann, M., Pacaud, K., Hoeffler, J-F., Van Dorsselaer, A., Rohmer, M., and Bach, T. J. 2006. A cytosolic Arabidopsis D-xylulose kinase catalyses the phosphorylation of 1-deoxy-D-xyulose in to a precursor of the plastidal isoprenoid pathway. Plant Physiol. 142:441457.Google Scholar
Hess, F. D. 2000. Light-dependent herbicides: an overview. Weed Sci. 48:160170.Google Scholar
Holliday, M. J. and Keen, N. T. 1982. The role of phytoalexins in the resistance of soybean leaves to bacteria: effect of glyphosate on glyceollin accumulation. Phytopathology. 72:14701474.Google Scholar
Kannanga, C. G. and Stumpf, P. K. 1972. Fat metabolism in higher plants, 54: procaryotic type acetyl CoA carboxylase in spinach chloroplasts. Arch. Biochem. Biophys. 152:8391.Google Scholar
Kelley, K. B. and Riechers, D. E. 2007. Recent developments in auxin biology and new opportunities for auxinic herbicide research. Pestic. Biochem. Physiol. 89:111.Google Scholar
Kepinski, S. and Leyser, O. 2005. The Arabidopsis TIR1 protein is an auxin receptor. Nature. 435:446451.Google Scholar
Kishore, G. M., Padgette, S. R., and Fraley, R. T. 1992. History of herbicide-tolerant crops, methods of development and current state of the art—emphasis on glyphosate tolerance. Weed Technol. 6:626634.Google Scholar
Koch, M., Breithaupt, C., Kiefersauer, R., Freigang, J., Huber, R., and Messerschmidt, A. 2004. Crystal structure of protoporphyrinogen IX oxidase: a key enzyme in haem and chlorophyll biosynthesis. EMBO J. 23:17201728.Google Scholar
Konishi, T. and Sasaki, Y. 1994. Compartmentalization of 2 forms of acetyl-CoA carboxylase in plants and the origin of their tolerance toward herbicides. Proc. Nat. Acad. Sci. U. S. A. 91:35983601.Google Scholar
Konishi, T., Shinohara, K., Yamada, K., and Sasaki, Y. 1996. Acetyl-CoA carboxylase in higher plants: most plants other than gramineae have both the prokaryotic and the eukaryotic forms of this enzyme. Plant Cell Physiol. 37:117122.Google Scholar
Kraft, M., Kuglitsch, R., Kwiatkowski, J., Frank, M., and Grossmann, K. 2007. Indole-3-acetic acid and auxin herbicides up-regulate 9-cis-epoxycarotenoid dioxygenase gene expression and abscisic acid accumulation in cleavers (Galium aparine): interaction with ethylene. J. Exp. Bot. 58:14971503.Google Scholar
Krijt, J., Psenak, O., Vokurka, M., Chlumska, A., and Fakan, F. 2003. Frantisek Experimental hepatic uroporphyria induced by the diphenyl-ether herbicide fomesafen in male DBA/2 mice. Toxicol. Appl. Pharmacol. 189:2838.Google Scholar
Kuzuyama, T., Shimizu, T., Takahashi, S., and Seto, H. 1998. Fosmidomycin, a specific inhibitor of 1-deoxy-D-xylulose 5-phosphate reductoisomerase in the nonmevalonate pathway for terpenoid biosynthesis. Tetrahedron Lett. 39:79137916.Google Scholar
Lancaster, C. R. D. and Michel, H. 1999. Refined crystal structures of reaction centres from Rhodopseudomonas viridis in complexes with the herbicide atrazine and two chiral atrazine derivatives also lead to a new model of the bound carotenoid. J. Mol. Biol. 286:883898.Google Scholar
Lange, B. M., Ketchum, R. E. B., and Croteau, R. B. 2001. Isoprenoid biosynthesis: metabolic profiling of peppermint oil gland secretory cells and application to herbicide target analysis. Plant Physiol. 127:305314.Google Scholar
Latunde-Dada, A. O. and Lucas, J. A. 1985. Involvement of the phytoalexin medicarpin in the differential response of callus lines of lucerne (Medicago sativa) to infection by Verticillium albo atrum . Physiol. Plant Pathol. 26:3142.Google Scholar
Lee, H. J., Duke, M. V., and Duke, S. O. 1993. Cellular localization of protoporphyrinogen-oxidizing activities of etiolated barley (Hordeum vulgare L.) leaves: relationship to mechanism of action of protoporphyrinogen oxidase inhibiting herbicides. Plant Physiol. 102:881889.Google Scholar
Lee, H. J. and Duke, S. O. 1994. Protoporphyrinogen oxidizing activities involved in the mode of action of peroxidizing herbicides. J. Agric. Food Chem. 42:26102618.Google Scholar
Lehnen, L. P., Sherman, T. D., Becerril, J. M., and Duke, S. O. 1990. Tissue and cellular localization of acifluorfen-induced porphyrins in cucumber cotyledons. Pestic. Biochem. Physiol. 37:239248.Google Scholar
Liang, W-S. 2003. Drought stress increases both cyanogenesis and β-cyanoalanine synthase activity in tobacco. Plant Sci. 165:11091115.Google Scholar
Liu, L., Punja, Z. K., and Rahe, J. E. 1997. Altered root exudation and suppression of induced lignification as mechanisms of predisposition by glyphosate of bean roots (Phaseolus vulgaris L.) to colonization by Pythium spp. Physiol. Mol. Plant Pathol. 51:111127.Google Scholar
Lloyd, C. and Chan, J. 2008. The parallel lives of microtubules and cellulose microfibrils. Curr. Opin. Plant Biol. 11:641646.Google Scholar
McMahon, J. M., Smith, J., and Arteca, R. N. 2000. Molecular control of ethylene production by cyanide in Arabidopsis thaliana . Physiol. Plant. 109:180187.Google Scholar
Michel, A., Arias, R. S., Scheffler, B. E., Duke, S. O., Netherland, M., and Dayan, F. E. 2004. Somatic mutation-mediated evolution of herbicide resistance in the nonindigenous invasive plant hydrilla (Hydrilla verticillata). Mol. Ecol. 13:32293237.Google Scholar
Michel, H. and Deisenhofer, J. 1988. Relevance of the photosynthetic reaction center from purple bacteria to the structure of photosystem II. Biochemistry. 27:17.Google Scholar
Miki, B. and McHugh, S. 2004. Selectable marker genes in transgenic plants: applications, alternatives and biosafety. J. Biotechnol. 107:193232.Google Scholar
Miller, J. M. and Conn, E. E. 1980. Metabolism of hydrogen cyanide by higher plants. Plant Physiol. 65:11991202.Google Scholar
Morejohn, L. C. and Fosket, D. E. 1991. The biochemistry of compounds with anti-microtubule activity in plant-cells. Pharmacol. Ther. 51:217230.Google Scholar
Mueller, C., Schwender, J., Zeidler, J., and Lichtenthaler, H. K. 2000. Properties and inhibition of the first two enzymes of the non-mevalonate pathway of isoprenoid bioynthesis. Biochem. Soc. Trans. 28:792793.Google Scholar
Mutwil, M., Debolt, S., and Persson, S. 2008. Cellulose synthesis: a complex complex. Curr. Opin. Plant Biol. 11:252257.Google Scholar
Nazakawa, M. and Matsui, M. 2003. Selection of hygromycin-resistant Arabidopsis seedlings. Biotechniques. 34:2830.Google Scholar
Oettmeier, W., Masson, K., Hohfeld, J., Meyer, H. E., Pfister, K., and Fischer, H. P. 1989. [I125]azido-ioxynil labels val249 of the photosystem-II D-1 reaction center protein. Z. Naturforsch. 44c:444449.Google Scholar
Ohad, N. and Hirschberg, J. 1992. Mutations in the D1 subunit of photosystem II distinguish between quinone and herbicide binding sites. Plant Cell. 4:273282.Google Scholar
Ohlrogge, J. and Browse, J. 1995. Lipid biosynthesis. The Plant Cell. 7:957970.Google Scholar
O'Looney, N. and Fry, S. C. 2005. Oxaziclomefone, a new herbicide, inhibits wall expansion in maize cell-cultures without affecting polysaccharide biosynthesis, xyloglucan transglycosylation, peroxidase action or apoplastic ascorbate oxidation. Ann. Bot. (Lond.) 96:10971107.Google Scholar
Paredez, A. R., Somerville, C. R., and Ehrhardt, D. W. 2006. Visualization of cellulose synthase demonstrates functional association with microtubules. Science. 312:14911495.Google Scholar
Pfister, K., Steinback, K. E., Gardner, G., and Arntzen, C. J. 1981. Photoaffinity-labeling of an herbicide receptor protein in chloroplast membranes. Proc. Nat. Acad. Sci. U. S. A. 78:981985.Google Scholar
Planchais, S., Glab, N., Inzé, D., and Bergounioux, C. 2000. Chemical inhibitors: a tool for plant cell cycle studies. FEBS Lett. 476:7883.Google Scholar
Popper, Z. A. 2008. Evolution and diversity of green plant cell walls. Curr. Opin. Plant Biol. 11:286292.Google Scholar
Porter, J. W. and Spurgeon, S. L. 1981. Biosynthesis of Isoprenoid Compounds. New York: Wiley.Google Scholar
Rajangam, A. S., Kumar, M., Aspeborg, H., et al. 2008. MAP20, a microtubule-associated protein in the secondary cell walls of hybrid aspen, is a target of the cellulose synthesis inhibitor 2,6-dichlorobenzonitrile. Plant Physiol. 148:12831294.Google Scholar
Rebeiz, C. A., Amindari, S., Reddy, K. N., Nandihalli, U. B., Moubarak, M. B., and Velu, J. A. 1994. δ-Aminolevulinic acid based herbicides and tetrapyrrole biosynthesis modulators. Am. Chem. Soc. Symp. Ser. 559:4864.Google Scholar
Reinhardt, D., Kende, H., and Boller, T. 1994. Subcellular localization of 1-aminocyclopropane-1-carboxylate oxidase in tomato cells. Planta. 195:142146.Google Scholar
Rohmer, M. 1999. The discovery of a mevalonate-independent pathway for isoprenoid biosynthesis in bacteria, algae and higher plants. Nat. Prod. Rep. 16:565574.Google Scholar
Ross, J. J., O'Neill, D. P., Wolbang, C. M., Symons, G. M., and Reid, J. B. 2001. Auxin-gibberellin interactions and their role in plant growth. J. Plant Growth Regul. 20:346353.Google Scholar
Sabba, R. P. and Vaughn, K. C. 1999. Herbicides that inhibit cellulose biosynthesis. Weed Sci. 47:757763.Google Scholar
Scheible, W. R., Eshed, R., Richmond, T., Delmer, D., and Somerville, C. 2001. Modifications of cellulose synthase confer resistance to isoxaben and thiazolidinone herbicides in Arabidopsis Ixr1 mutants. Proc. Nat. Acad. Sci. U. S. A. 98:1007910084.Google Scholar
Scheltrup, F. and Grossmann, K. 1995. Abscisic acid is a causative factor in the mode of action of the auxinic herbicide quinmerac in cleaver (Galium aparine L.). J. Plant Physiol. 147:118126.Google Scholar
Schwartz, S. H., Qin, X., and Zeevaart, J. A. D. 2003. Elucidation of the indirect pathway of abscisic acid biosynthesis by mutants, genes, and enzymes. Plant Physiol. 131:15911601.Google Scholar
Schwarz, M. K. 1994. TerpenBiosynthese in Ginkgo biloba: Eine uberraschende Geschichte [Terpene Biosynthesis in Ginkgo biloba: A Surprising Story]. . Zurich, Switzerland: Eidgenössische Technische Hochschule. [In Swiss].Google Scholar
Sharon, A., Amsellem, Z., and Gressel, J. 1992. Glyphosate suppression of an elicited response. Increased susceptibility of Cassia obtusifolia to a mycoherbicide. Plant Physiol. 98:654659.Google Scholar
Siegien, I. and Bogatek, R. 2006. Cyanide action in plants—from toxic to regulatory. Acta Physiol. Plant. 28:483497.Google Scholar
Sterling, T. M. and Hall, J. C. 1997. Mechanism of action of natural auxins and the auxinic herbicides. Pages 111141. in Roe, R. M., Burton, J. D., and Kuhr, R. J. eds. Herbicide Activity: Toxicology, Biochemistry and Molecular Biology. Amsterdam: IOS.Google Scholar
Sundar, I. K. and Sakthivel, N. 2008. Advances in selectable marker genes for plant transformation. J. Plant Physiol. 165:16981716.Google Scholar
Tan, X., Calderon-Villalobos, L. I. A., Sharon, M., Zheng, C., Robinson, C. V., Estelle, M., and Zheng, N. 2007. Mechanism of auxin perception by the TIR1 ubiquitin ligase. Nature. 446:640645.Google Scholar
Taylor, I. B., Sonneveld, T., Bugg, T. D. H., and Thompson, A. J. 2005. Regulation and manipulation of the biosynthesis of abscisic acid, including the supply of xanthophyll precursors. J. Plant Growth Regul. 24:253273.Google Scholar
Tittle, F. L., Goudey, J. S., and Spencer, M. S. 1990. Effect of 2,4-dichlorophenoxyacetic acid on endogenous cyanide, β-cyanoalanine synthase activity, and ethylene evolution in seedlings of soybean and barley. Plant Physiol. 94:11431148.Google Scholar
Trebst, A. 2007. Inhibitors in the functional dissection of the photosynthetic electron transport system. Photosynth. Res. 92:217224.Google Scholar
Trebst, A., Wietoska, H., Draber, W., and Knops, H. J. 1978. Inhibition of photosynthetic electron flow in chloroplasts by the dinitrophenylether of bromo-nitrothymol or iodo-nitrothymol. Z. Naturforsch. 33c:919923.Google Scholar
Vanneste, S. and Friml, J. 2009. Auxin: a trigger for change in plant development. Cell. 136:10051016.Google Scholar
Vaughn, K. C. and Lehnen, L. P. J. 1991. Mitotic disrupter herbicides. Weed Sci. 39:450457.Google Scholar
Ververidis, P. and Dilley, D. R. 1995. Catalytic and noncatalytic inactivation of ACC oxidase. Pages 183189. in. Proceedings of the 22nd Annual Meeting of the Plant Growth Regulation Society of America. LaGrange, GA: PGRSA.Google Scholar
Walsh, T. A. 2007. The emerging field of chemical genetics: potential applications for pesticide discovery. Pest Manag. Sci. 63:11651171.Google Scholar
Walsh, T. A., Neal, R., Merlo, A. O., Honma, M., Hicks, G. R., Wolff, K., Matsumura, W., and Davies, J. P. 2006. Mutations in an auxin receptor homolog AFB5 and in SGT1b confer resistance to synthetic picolinate auxins and not to 2,4-dichlorophenoxyacetic acid or indole-3-acetic acid in Arabidopsis . Plant Physiol. 142:542552.Google Scholar
Witt, K. 1973. Further evidence of X-320 as a primary acceptor of photosystem II in photosynthesis. FEBS Lett. 38:116118.Google Scholar
Woodward, A. W. and Bartel, B. 2005. Auxin: regulation, action, and interaction. Ann. Bot. (Lond.) 95:707735.Google Scholar
Wurtele, E. S., Nikolau, B. J., and Conn, E. E. 1985. Subcellular and developmental distribution of ß-cyanoalanine synthase in barley leaves. Plant Physiol. 78:285290.Google Scholar
Yang, S. F. and Hoffman, N. E. 1984. Ethylene biosynthesis and its regulation in higher plants. Ann. Rev. Plant Physiol. 35:155189.Google Scholar
Yip, W-K. and Yang, S. F. 1998. Ethylene biosynthesis in relation to cyanide metabolism. Bot. Bull. Acad. Sin. (Taipei) 39:17.Google Scholar
Zeidler, J., Schwender, J., Müller, C., Weisner, J., Weidemeyer, C., Beck, E., Jomaa, H., and Lichtenthaler, H. K. 1998. Inhibition of the non-mevalonate 1-deoxy-D-xylulose-5-phosphate pathway of plant isoprenoid biosynthesis by fosmidomycin. Z. Naturforsch. 54c:980986.Google Scholar