Hostname: page-component-76fb5796d-zzh7m Total loading time: 0 Render date: 2024-04-26T17:00:11.327Z Has data issue: false hasContentIssue false

The Influence of Temperature, Moisture, and Prior EPTC Application on the Degradation of EPTC in Soils

Published online by Cambridge University Press:  12 June 2017

Tim Obrigawitch
Affiliation:
Dep. Agron., Univ. of Nebraska, Lincoln, NE 68583, Scottsbluff, NE 69361, Lincoln, NE 68583, and Clay Center, NE 68933
Robert G. Wilson
Affiliation:
Dep. Agron., Univ. of Nebraska, Lincoln, NE 68583, Scottsbluff, NE 69361, Lincoln, NE 68583, and Clay Center, NE 68933
Alex R. Martin
Affiliation:
Dep. Agron., Univ. of Nebraska, Lincoln, NE 68583, Scottsbluff, NE 69361, Lincoln, NE 68583, and Clay Center, NE 68933
Fred W. Roeth
Affiliation:
Dep. Agron., Univ. of Nebraska, Lincoln, NE 68583, Scottsbluff, NE 69361, Lincoln, NE 68583, and Clay Center, NE 68933

Abstract

Laboratory and field studies were conducted to compare the degradation rate of EPTC (S-ethyl dipropylthiocarbamate) in soils with a previous history of EPTC application vs. soils with no prior EPTC application. Laboratory experiments showed a rapid breakdown of 14C-carbonyl EPTC to 14CO2 in Kennebec silt loam (sil) and Tripp very fine sandy loam (vfsl) with prior exposure to EPTC. A single prior application of EPTC to Tripp vfsl was sufficient to increase the rate of 14CO2 evolution. The rate of EPTC degradation in Tripp vfsl with a history of EPTC exposure was dependent on soil moisture from below 3% and independent of moisture above 3%. Patterns of degradation at 5, 15, and 25 C in the Kennebec sil with and without prior history were described by exponential decay. EPTC was degraded more rapidly at 15 and 25 C in the Kennebec sil with a prior EPTC history. Under field conditions the breakdown of EPTC in a Kennebec sil with and without prior exposure to EPTC was similar to laboratory results.

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

1. Ashton, F. M. and Sheets, T. J. 1959. The relationship of soil adsorption of EPTC to oats injury in various soil types. Weeds 7:8890.Google Scholar
2. Audus, L. J. 1949. Biological detoxication of 2,4-D. Plant Soil 2:3135.Google Scholar
3. Audus, L. J. 1964. The effects of soil microorganisms on herbicides. Pages 165184 in Audus, L. J., ed. The Physiology and Biochemistry of Herbicides. Academic Press, New York.Google Scholar
4. Cohn, M. and Monod, J. 1953. Specific inhibition and induction of enzyme biosynthesis. Symp. III. Soc. Gen. Microbiol., Camb. Univ. Press. 132 pp.Google Scholar
5. Danielson, L. L. and Gentner, W. A. 1964. Influence of air movement on persistence of EPTC on soil. Weeds 12:9294.Google Scholar
6. Danielson, L. L., Gentner, W. A., and Jansen, L. L. 1961. Persistence of soil-incorporated EPTC and other carbamates. Weed Sci. 9:463476.Google Scholar
7. Fang, S. C., Theisen, P., and Freed, V. H. 1961. Effects of water evaporation, temperature and rates of application on the retention of EPTC in various soils. Weeds 9:569574.Google Scholar
8. Fryer, J. D. and Kirkland, K. 1970. Field experiments to investigate long-term effects of repeated applications of MCPA, triallate, simazine and linuron: report after six years. Weed Res. 10:133158.Google Scholar
9. Gray, R. A. and Weierich, A. J. 1965. Factors affecting vapor loss of EPTC from soils. Weeds 13:141147.CrossRefGoogle Scholar
10. Gray, R. A. and Weierich, A. J. 1968. Behavior and persistence of thiocarbamate herbicides in soils under different environmental conditions. Proc. 9th Br. Weed Control Conf. pp. 94101.Google Scholar
11. Hamaker, J. W. 1972. Decomposition: quantitative aspects. Pages 253340 in Goring, C.A.I. and Hamaker, J. W., eds. Organic Chemicals in the Soil Environment. Vol. I. Marcel Dekker Inc., New York.Google Scholar
12. Hyzak, D. L. and Zimdahl, R. L. 1974. Rate of degradation of metribuzin and two analogs in soil. Weed Sci. 22:7579.Google Scholar
13. Kaufman, D. D. 1967. Degradation of carbamate herbicides in soil. J. Agric. Food Chem. 15:582591.Google Scholar
14. Kaufman, D. D. and Kearney, P. C. 1970. Microbial degradation of triazine herbicides. Residue Rev. 32:235265.Google Scholar
15. Kirkland, K. and Fryer, J. D. 1972. Degradation of several herbicides in a soil previously treated with MCPA. Weed Res. 12:9095.Google Scholar
16. McRae, I. C. and Alexander, M. 1965. Microbial degradation of selected herbicides in soil. J. Agric. Food Chem. 13:7275.Google Scholar
17. Nalewaja, J. D., Behrens, R., and Schmid, A. R. 1964. Uptake, translocation and fate of EPTC in alfalfa. Weeds 12:269272.Google Scholar
18. Rahman, A., Atkinson, G. C., and Douglas, J. A. 1979. Eradicane causes problems. N.Z. J. Agric. 139(3):4749.Google Scholar
19. Sheets, T. J. 1959. Effects of soil type and time on the herbicidal activity of CDAA, CDEC and EPTC. Weeds 7:442448.Google Scholar
20. Torstensson, N.T.L., Stark, J., and Goransson, B. 1975. The effect of repeated applications of 2,4-D and MCPA on their breakdown in soil. Weed Res. 15:159164.Google Scholar
21. Walker, R. L. and Newman, A. S. 1956. Microbial decomposition of [(2,4-dichlorophenoxy)acetic acid]. Appl. Microbiol. 4:201206.Google Scholar