Hostname: page-component-848d4c4894-m9kch Total loading time: 0 Render date: 2024-05-09T01:05:08.326Z Has data issue: false hasContentIssue false

Analytical Techniques to Measure Sethoxydim and Breakdown Products

Published online by Cambridge University Press:  12 June 2017

Antony R. Shoaf
Affiliation:
Div. Toxicol., Dep. Pharmacol. and Interdisciplinary Toxicol., Univ. Arkansas for Medical Sciences, Little Rock, AR 72205
William C. Carlson
Affiliation:
South. Forestry Res. Dep., Weyerhaeuser Co., Hot Springs, AR 71902

Abstract

A method was developed for the quantitative determination of trace levels of the widely used herbicide sethoxydim {2-[1-(ethoxyimino)butyl]-5-[2-(ethylthio)propyl]-3-hydroxy-2-cyclohexen-1-one} and its metabolites in an aqueous solution using reversed-phase high-performance liquid chromatography (HPLC). Optimum extraction of sethoxydim was with dichloromethane and was only 15% efficient at pH 3. The limit of detection by HPLC for sethoxydim was 5 ng on column and <5 ppb in soil. At least five different compounds were detected in the commercial formulation, in EPA reference standards, and in commercial sethoxydim standards. Sethoxydim undergoes a rapid decomposition in the presence of water to form more polar products, which accounts for the low extraction efficiency. Decomposition was greatest under alkaline conditions. Acid pH and soil inhibited decomposition and gave greater recoveries of parent compound. At least one breakdown product cochromatographed with a known sulfone derivative. The procedures are directly applicable to soils, environmental waters, and plant and animal tissues.

Type
Weed Control and Herbicide Technology
Copyright
Copyright © 1986 by the 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

Literatur Cited

1. Asare-Boamah, N. K. and Fletcher, R. A. 1983. Physiological and cytological effects of BAS 9052 OH on corn (Zea mays) seedlings. Weed Sci. 31:4955.Google Scholar
2. BASF Poast Technical Information Bulletin. 1983. BASF Wyandotte Corp., Parsippany, NJ.Google Scholar
3. Krebs, K. G., Heusser, D., and Wimmer, H. 1969. Spray Reagents. Pages 854911 in Stahl, E., ed. Thin-Layer Chromatography. Springer-Verlag, Berlin.Google Scholar
4. NABU (NP-55) Common Name: Sethoxydim – A New Selective Herbicide, 1982. Japan Pestic. Information No. 41. Pages 1821. Nippon Soda Co., Ltd. Google Scholar
5. Peters, R. A. 1980. Postemergence control of crabgrass with BAS 9052 OH in new alfalfa and red clover seedlings. Proc. Northeast Weed Sci. Soc. 34:101.Google Scholar
6. Shiau, S. Y., Huff, R. A., and Felkner, I. C. 1981. Pesticide mutagenicity in bacillus subtilis and salmonella typhimurium detectors. J. Agric. Food Chem. 29:268271.Google Scholar
7. Swisher, B. A. and Corbin, F. T. 1982. Behavior of BAS 9052 OH in soybean (Glycine max) and johnsongrass (Sorghum halepense) plant and cell cultures. Weed Sci. 30:640650.Google Scholar
8. Wilson, H. P. and Hines, T. E. 1980. Postemergence control of annual grasses in soybeans. Proc. Northeast. Weed Sci. Soc. 34: 610.Google Scholar