Hostname: page-component-7479d7b7d-k7p5g Total loading time: 0 Render date: 2024-07-10T20:41:05.043Z Has data issue: false hasContentIssue false

Interspecific Differences in Weed Susceptibility to Steam Injury

Published online by Cambridge University Press:  20 January 2017

Ramon G. Leon*
Affiliation:
Universidad EARTH, Apartado 4442-1000, San Jose, Costa Rica
Dylan T. Ferreira
Affiliation:
Horticulture and Crop Science Department, California Polytechnic State University, San Luis Obispo, CA 93401
*
Corresponding author's E-mail: rleon@earth.ac.cr.

Abstract

Thermal weed control methods have been incorporated into weed control programs in organic and conventional production systems. Flaming is commonly used, but steaming has been proposed to increase efficiency of heat transfer to weeds and reduce the risk of fire. The objective of this research was to measure injury to leaves of plant species that differ in leaf morphology and to measure injury to plants at different stages of plant development. The study was conducted in a glasshouse and plants were exposed to steaming at 400 C for 0.36 s—equivalent to a steaming speed of 2 km/h. Overall, leaf thickness was the best morphological characteristic to predict injury ( = 0.51), with greater thickness resulting in less injury. For broadleaf species only, species with wider leaves were injured more than species with narrower leaves ( = 0.64). Injury was greatest when plants had fewer than six true leaves and when their shoots were less than 10 cm long. There was a wide range of injury across species, and the grass species bermudagrass and perennial ryegrass were injured (68 to 81%) more than other species such as common purslane and English daisy (23 to 34%). Biomass of all species tested was reduced by approximately 40%, indicating that leaf injury was not the sole effect of steaming on plant growth. These results indicated that considering both visual estimates of injury and morphological characteristics is important to properly assess thermal weed control effectiveness.

Type
Weed Management—Techniques
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

Ascard, J. 1994. Dose–response models for flame weeding in relation to plant size and density. Weed Res 34:377385.CrossRefGoogle Scholar
Ascard, J. 1995. Effects of flame weeding on weed species at different developmental stages. Weed Res 35:397411.Google Scholar
Ascard, J. 1997. Flame weeding: effects of fuel pressure and tandem burners. Weed Res 37:7786.CrossRefGoogle Scholar
Ascard, J. 1998a. Comparison of flaming and infrared radiation techniques for thermal weed control. Weed Res 38:6976.Google Scholar
Ascard, J. 1998b. Flame weeding: effects of burner angle on weed control and temperature patterns. Acta Agric. Scand. Sect. B Soil Plant Sci 48:248254.Google Scholar
Bond, W. and Grundy, A. C. 2001. Non-chemical weed management in organic farming systems. Weed Res 41:383405.Google Scholar
Daniell, J. W., Chappell, W. E., and Couch, H. B. 1969. Effect of sublethal and lethal temperatures on plant cells. Plant Physiol 44:16841689.CrossRefGoogle ScholarPubMed
Hansen, P. K., Kristoffersen, P., and Kristensen, K. 2004. Strategies for non-chemical weed control on public paved areas in Denmark. Pest Manage. Sci 60:600604.Google Scholar
Hansson, D. and Ascard, J. 2002. Influence of developmental stage and time of assessment on hot water weed control. Weed Res 42:307316.Google Scholar
Hatcher, P. E. and Melander, B. 2003. Combining physical, cultural and biological methods: prospects for integrated non-chemical weed management strategies. Weed Res 43:303322.Google Scholar
Melander, B., Rasmussen, I. A., and Barberi, P. 2005. Integrating physical and cultural methods of weed control-examples from European research. Weed Sci 53:369381.CrossRefGoogle Scholar
Norris, R. F. and Kogan, M. 2000. Interactions between weeds, arthropod pests, and their natural enemies in managed ecosystems. Weed Sci 48:94158.Google Scholar
Rahkonen, J. and Jokela, H. 2003. Infrared radiometry for measuring plant leaf temperature during thermal weed control treatment. Biosys. Eng 86:257266.Google Scholar
Rask, A. M. and Kristoffersen, R. 2007. A review of non-chemical weed control on hard surfaces. Weed Res 47:370380.Google Scholar
Sartorato, I., Zanin, G., Baldoin, C., and de Zanche, C. 2006. Observations on the potential of microwaves for weed control. Weed Res 46:19.Google Scholar
SAS 1998. SAS/STAT Useŕs Guide 7.0. Cary, NC: Statistical Analysis Systems Institute. 1028.Google Scholar
Sirvydas, A., Lazauskas, P., Stepanas, A., Nadzeikiene, J., and Kerpauskas, P. 2006. Plant temperature variation in the thermal weed control process. J. Plant Dis. Prot 20:355361.Google Scholar
Timmons, F. L. 2005. A history of weed control in the United States and Canada. Weed Sci 53:748761.Google Scholar