Hostname: page-component-8448b6f56d-wq2xx Total loading time: 0 Render date: 2024-04-23T14:22:03.969Z Has data issue: false hasContentIssue false

Rapid Assay Evaluation of Plant Response to Protoporphyrinogen Oxidase (Protox)-Inhibiting Herbicides

Published online by Cambridge University Press:  20 January 2017

Jeanne S. Falk
Affiliation:
Department of Agronomy, Kansas State University, 2004 Throckmorton Hall, Manhattan, KS 66506
Kassim Al-Khatib*
Affiliation:
Department of Agronomy, Kansas State University, 2004 Throckmorton Hall, Manhattan, KS 66506
Dallas E. Peterson
Affiliation:
Department of Agronomy, Kansas State University, 2004 Throckmorton Hall, Manhattan, KS 66506
*
Corresponding author's E-mail: khatib@ksu.edu

Abstract

Protoporphyrinogen oxidase (protox)-inhibiting herbicides damage cell membranes, resulting in electrolyte leakage. A whole-plant dose-response study and a rapid assay that measured electrolyte leakage was used to determine the response of wild mustard, soybean, and protox inhibitor–susceptible and protox inhibitor–resistant common waterhemp to increasing doses of three protox inhibitors: acifluorfen, fomesafen, and sulfentrazone. For the dose-response study, whole plants were treated with the three protox-inhibitor herbicides. Electroconductivity assay 1 consisted of cutting discs from leaf tissue and submerging them in an incubation medium containing concentrations of acifluorfen, fomesafen, or sulfentrazone. In electroconductivity assay 2, the entire leaf was treated with solutions containing acifluorfen, fomesafen, or sulfentrazone. The whole-plant dose-response study showed increasing visible injury with increasing herbicide rates for all species and all herbicides. The order of visible injury was wild mustard > susceptible common waterhemp > resistant common waterhemp > soybean. In assay 1, electrolyte leakage from leaf discs treated with acifluorfen or fomesafen increased with increasing herbicide concentrations, and was similar for all species. In contrast, electrolyte leakage from leaf discs treated with sulfentrazone did not increase with increasing herbicide concentrations for any species. In assay 2, only wild mustard leaf discs increased in electrolyte leakage with increasing herbicide rates of acifluorfen, fomesafen, and sulfentrazone and followed the regression curves established by the whole-plant dose-response study. However, assay 2 was not able to distinguish between susceptible wild mustard and tolerant soybean, or between susceptible and resistant waterhemp.

Type
Research 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

Bailey, W. A., Wilson, H. P., and Hines, T. E. 2002. Response of potato (Solanum tuberosum) and selected weeds to sulfentrazone. Weed Technol. 16:651658.Google Scholar
Becerril, J. M. and Duke, S. O. 1989. Protoporphyrin IX content correlates with activity of photobleaching herbicides. Plant Physiol. 90:11751181.CrossRefGoogle ScholarPubMed
Blum, A. and Ebercon, A. 1981. Cell membrane stability as a measure of drought and heat tolerance in wheat. Crop Sci. 21:4347.Google Scholar
Choi, J. S., Lee, H. J., Hwang, I. T., Pyon, J. Y., and Cho, K. Y. 1999. Differential susceptibilities of wheat and barley to diphenyl ether herbicide oxyfluorfen. Pest. Biochem. Physiol. 65:6272.Google Scholar
Dayan, F. E., Green, H. M., Weete, J. D., and Hancock, H. G. 1996a. Postemergence activity of sulfentrazone: effects of surfactants and leaf surfaces. Weed Sci. 44:797803.Google Scholar
Dayan, F. E., Weete, J. D., Duke, S. O., and Hancock, H. G. 1997. Soybean (Glycine max) cultivar differences in response to sulfentrazone. Weed Sci. 45:634641.Google Scholar
Dayan, F. E., Weete, J. D., and Hancock, H. G. 1996b. Physiological basis for differential sensitivity to sulfentrazone by sicklepod (Senna obtusifolia) and coffee senna (Cassia occidentalis). Weed Sci. 44:1217.Google Scholar
Duke, S. O. and Kenyon, W. H. 1993. Peroxidizing activity determined by cellular leakage. in Boger, P. and Sandman, G., eds. Target Assays for Modern Herbicides and Related Phytotoxic Compounds. Boca Raton, FL: Lewis. Pp. 6166.Google Scholar
Duke, S. O., Lydon, J., and Paul, R. N. 1989. Oxadiazon activity is similar to that of p-nitro-diphenyl ether herbicides. Weed Sci. 37:152160.CrossRefGoogle Scholar
Fadayomi, O. and Warren, G. F. 1976. The light requirement for herbicidal activity of diphenyl ethers. Weed Sci. 24:598600.Google Scholar
Fadayomi, O. and Warren, G. F. 1977. Uptake and translocation of nitrofen and oxyfluorfen. Weed Sci. 25:111114.Google Scholar
Falk, J. S., Shoup, D. E., Al-Khatib, K., and Peterson, D. E. 2004. Survey of common waterhemp response to protox- and ALS-inhibiting herbicides in northeast Kansas. Proc. West. Soc. Weed Sci. 57:73.Google Scholar
Jacobs, J. M. and Jacobs, N. J. 1993. Porphyrin accumulation and export by isolated barley (Hordeum vulgare) plastids. Plant Physiol. 101:11811187.Google Scholar
Jacobs, J. M., Jacobs, N. J., Sherman, T. D., and Duke, S. O. 1991. Effect of diphenyl herbicides on oxidation of protoporphrinogen to protoporphyrin in organellar and plasma membrane enriched fractions of barley. Plant Physiol. 97:197203.Google Scholar
Johnson, W. O., Kollman, G. E., Swithenbank, C., and Yih, R. Y. 1978. RH-6201 (Blazer): a new broad spectrum herbicide for postemergence use in soybeans. J. Agric. Food Chem. 26:285286.Google Scholar
Kenyon, W. H., Duke, S. O., and Vaughn, K. C. 1985. Sequential of effects of acifluorfen on physiological and ultrastructural parameters in cucumber cotyledon discs. Pestic. Biochem. Physiol. 24:240250.Google Scholar
Knowles, N. R. and Knowles, L. O. 1989. Correlations between electrolyte leakage and degree of saturation of polar lipids from aged potato (Solanum tuberosum L.) tuber tissue. Ann. Bot. 63:331338.Google Scholar
Koo, S. J., Neal, J. C., and Di Tomaso, J. M. 1994. Quinclorac-induced electrolyte leakage in seedling grasses. Weed Sci. 42:17.Google Scholar
Krausz, R. F., Kapusta, G., and Matthews, J. L. 1998. Sulfentrazone for weed control in soybeans (Glycine max). Weed Technol. 12:684689.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
Li, J., Smeda, R. J., Nelson, K. A., and Dayan, F. E. 2004. Physiological basis for resistance to diphenyl ether herbicides in common waterhemp (Amaranthus rudis). Weed Sci. 52:333338.CrossRefGoogle Scholar
Li, Z., Walker, R. H., Wehtje, G., and Hancock, H. G. 2000. Using electrolyte leakage to detect soybean (Glycine max) cultivars sensitive to sulfentrazone. Weed Technol. 14:699704.Google Scholar
Liu, X. and Huang, B. 2000. Heat stress injury in relation to membrane peroxidation in creeping bentgrass. 40:503510.Google Scholar
Matringe, M., Camadro, J. M., Labette, P., and Scalla, R. 1989. Protoporphyrinogen oxidase as a molecular target for diphenyl ether herbicides. Biochem. J. 260:231235.Google Scholar
Matsumoto, H. and Duke, S. O. 1990. Acifluorfen-methyl effects on prophyrin synthesis in Lemna pausicostata Hegelm. J. Agric. Food Chem. 38:20662071.Google Scholar
Matsunaka, S. 1969. Acceptor of light energy in photoactivation of diphenylether herbicides. J. Agric. Food. Chem. 17:171175.Google Scholar
Nunes, M. E. S. and Smith, G. R. 2003. Electrolyte leakage assay capable of quantifying freezing resistance in rose clover. Crop Sci. 43:13491357.Google Scholar
Patzoldt, W. L., Hager, A. G., and Tranel, P. J. 2002. An Illinois waterhemp biotype with resistance to PPO, ALS, and PSII inhibitors. Proc. North Cent. Weed Sci. Soc. 57:161.Google Scholar
Peterson, D. E., Regehr, D. L., Thompson, C. R., and Al-Khatib, K. 2001. Herbicide Mode of Action. Publication C-715. Manhattan, KS: Kansas Cooperative Extension Service. Pp. 1415.Google Scholar
Ritter, R. L. and Coble, H. D. 1981a. Penetration, translocation, and metabolism of acifluorfen in soybean (Glycine max), common ragweed (Ambrosia artemisiifolia), and common cocklebur (Xanthium pensylvanicum). Weed Sci. 29:474480.Google Scholar
Ritter, R. L. and Coble, H. D. 1981b. Influence of temperature and relative humidity on the activity of acifluorfen. Weed Sci. 29:480485.Google Scholar
Sherman, T. D., Becerril, J. M., Matsumoto, H., Duke, M. V., Jacobs, J. M., Jacobs, N. J., and Duke, S. O. 1991. Physiological basis for differential sensitivities of plant species to protoporphyrinogen oxidase-inhibiting herbicides. Plant Physiol. 97:280287.Google Scholar
Shoup, D. E., Al-Khatib, K., and Peterson, D. E. 2003. Common waterhemp (Amaranthus rudis) resistance to protoporphyrinogen oxidase-inhibiting herbicides. Weed Sci. 51:145150.Google Scholar
Sweat, J. K., Horak, M. J., Peterson, D. E., Lloyd, R. W., and Boyer, J. E. 1998. Herbicide efficacy on four Amaranthus species in soybean (Glycine max). Weed Sci. 12:315321.Google Scholar
Unland, R. D., Al-Khatib, K., and Peterson, D. E. 1999. Interactions between imazamox and diphenylethers. Weed Sci. 47:462466.Google Scholar
Vanstone, D. E. and Stobbe, E. H. 1977. Electrolytic conductivity—a rapid measure of herbicide injury. Weed Sci. 25:352354.CrossRefGoogle Scholar
Vencill, W. K. 2002. Herbicide Handbook. 8th ed. Lawrence, KS. Weed Science Society of America. 477 p.Google Scholar
Wilson, D. E., Nissen, S. J., and Thompson, A. 2002. Potato (Solanum tuberosum) variety and weed response to sulfentrazone and flumioxazin. Weed Technol. 16:567574.Google Scholar
Witkowski, D. A. and Halling, B. P. 1989. Inhibition of plant protoporphyrinogen oxidase by herbicide acifluorfen-methyl. Plant Physiol. 90:12391242.Google Scholar
Yanase, D., Andoh, A., and Yasudomi, N. 1990. A new simple bioassay to evaluate photosynthetic electron-transport inhibition utilizing paraquat phytotoxicity. Pestic. Biochem. Physiol. 38:9298.Google Scholar