Wild Oat (Avena fatua) and Spring Barley (Hordeum vulgare) Interference in a Greenhouse Experiment

Weed Science ◽  
1991 ◽  
Vol 39 (2) ◽  
pp. 149-153 ◽  
Author(s):  
Don W. Morishita ◽  
Donald C. Thill ◽  
John E. Hammel

Intraspecific and interspecific interference effects on growth, gas exchange, and water potential of wild oat and spring barley were measured under greenhouse conditions using a 1:1.06 barley to wild oat replacement series. Intraspecific barley interference affected barley growth more than interspecific wild oat interference. Interspecific wild oat interference with barley reduced wild oat growth more than intraspecific interference. Wild oat plant height surpassed barley plant height near barley anthesis. Growth and gas exchange of barley and wild oat responded the same to short-term water stress.

Weed Science ◽  
2013 ◽  
Vol 61 (1) ◽  
pp. 26-31 ◽  
Author(s):  
Ilias S. Travlos

Studies were conducted to determine the growth, fecundity, and competitive ability of an acetyl-CoA carboxylase (ACCase)–inhibitor resistant (R) sterile wild oat biotype compared with a susceptible (S) biotype. Seed germination studies indicated that there were no differences in seed germination and seedling vigor between R and S biotypes at any temperature regime. R and S biotypes were grown under noncompetitive and competitive arrangement in the greenhouse. Under noncompetitive greenhouse conditions, growth of the R biotype was similar to that of the S biotype on the basis of plant height, canopy area, and plant biomass. Seed production and weight of R and S plants were also at the same levels. Furthermore, relative competitiveness among the R and S sterile wild oat biotypes was investigated by means of replacement series experiments. The R and S biotypes were compared under seven mixture proportions (6 : 0, 5 : 1, 4 : 2, 3 : 3, 2 : 4, 1 : 5, and 0 : 6). No significant differences in competitive ability were observed between R and S biotypes on the basis of plant height, canopy area, or plant biomass. In most cases, relative crowding coefficient (RCC) values at 20, 60, and 100 d after transplanting (DAT) were close to one, indicating equal competitiveness between the R and S biotypes of wild oat used in this competitive study. However, in some cases, the RCC value was 1.31 for plant height, evident of a slight competitive advantage for the R biotype at 100 DAT. In general, ACCase-inhibitor R and S sterile wild oat biotypes were equally competitive, clearly without any growth penalty for R plants in either noncompetitive or competitive conditions.


Weed Science ◽  
1977 ◽  
Vol 25 (4) ◽  
pp. 355-359 ◽  
Author(s):  
Wayne A. Olson ◽  
John D. Nalewaja

The tolerance of wheat (Triticum aestivum L. ‘Waldron’) and wild oat (Avena fatua L.) to various rates of flufenprop-methyl {methyl-2-[benzoyl(3-chloro-4-fluorophenyl)amino]propanoate} applied weekly after wheat and wild oat emergence was determined under field conditions. Wild oat control increased at all growth stages as flufenprop-methyl rate increased. Wild oat control was greater than 80% with flufenprop-methyl at all rates when applied up to 6 weeks after wild oat emergence: (anthesis stage), but decreased when application was delayed further. Wheat was most susceptible to flufenprop-methyl during anthesis. Flufenprop-methyl at 0.56 kg/ha injured weed-free wheat only at the boot and anthesis stages. Injury intensity and the number of weeks that injury remained evident increased as flufenprop-methyl rate increased. Flufenprop-methyl injury to wheat was expressed as reduced plant height, grain yield, and kernels per spike and increased grain protein. Plant height reductions were attributed to reduced cell elongation. Grain yield reductions resulted from reduced kernels per spike.


Weed Science ◽  
1988 ◽  
Vol 36 (1) ◽  
pp. 43-48 ◽  
Author(s):  
Don W. Morishita ◽  
Donald C. Thill

Barley (Hordeum vulgareL. ‘Advance’) and wild oat (Avena fatuaL. # AVEFA) were grown in the field in monoculture and mixed culture (additively) to compare their seasonal growth and development. Barley and wild oat tiller and tiller head production were reduced by the interference (higher density) of the other species. Plant height of either species was not affected by interference of the other. Wild oat biomass was reduced more and at an earlier growth stage (two to three tillers) than was barley biomass (heading) in mixed culture. Barley and wild oat grown in monoculture had similar total plant nitrogen content throughput the growing season. Gas exchange and water potential of barley and wild oat in monoculture and mixed culture were similar. All measurements indicated that barley and wild oat grown in monoculture had growth and development patterns that were similar. In mixed culture, however, barley was more competitive with wild oat than wild oat was with barley. Wild oat reduced barley yield component quality and grain yield.


1997 ◽  
Vol 11 (2) ◽  
pp. 283-289 ◽  
Author(s):  
Robert N. Stougaard ◽  
Bruce D. Maxwell ◽  
Jerry D. Harris

Field experiments were conducted during 1992 and 1993 at Kalispell and Moccasin, MT, to determine the influence of application timing on the efficacy of reduced rate postemergence applications of imazamethabenz and diclofop in spring barley. Herbicides were applied at their respective 1 × and ½ × use rates at either 1, 2, or 3 weeks after crop emergence (WAE). While excellent wild oat control was sometimes achieved with reduced rates, there was no consistent relationship between wild oat growth stage and the level of control at either site regardless of the herbicide or rate applied. This response suggests that efficacy is governed not only by wild oat growth stage, but also by weed demographics and environmental considerations. Barley yield and adjusted gross return values were highest at Kalispell when imazamethabenz treatments were applied at 1 WAE, regardless of the level of wild oat control. Adjusted gross return values were similar for the 1 × and ½ × imazamethabenz treatments. Yields and adjusted gross returns with diclofop treatments were more related to the level of wild oat control at Kalispell, with the 1 × diclofop treatments providing the greatest yields and adjusted gross return values. The level of wild oat control at Moccasin had minimal effect on barley yield and adjusted gross returns, with both values being comparable to the nontreated check.


Weed Science ◽  
1998 ◽  
Vol 46 (3) ◽  
pp. 336-343 ◽  
Author(s):  
Michael J. Wille ◽  
Donald C. Thill ◽  
William J. Price

The efficacy of current wild oat herbicides and their high cost have resulted in use rates that are less than those recommended. While acceptable weed control may be attained at less cost, this practice does not consider the potential for increased wild oat seed production. The objective of this experiment was to determine the interaction of wild oat density and reduced imazamethabenz rates on wild oat seed production in spring barley. As wild oat densities increased from 8 to 1,100 plants m−2, wild oat seed production increased from 180 to 9,950 seed m−2without herbicide, and from 0 to 2,810 seed m−2using 0.53 kg ai ha−1imazamethabenz. This general pattern was modeled using a cumulative logistic function. Estimates from this model indicated that < 1 wild oat seed m−2was produced at population densities of ≤ 20 plants at any imazamethabenz rate. Imazamethabenz rates of 0.26 kg ha−1or greater at wild oat densities of less than approximately 190 plants m−2did not result in wild oat seed production above the initial population density. As wild oat density increased, however, imazamethabenz rates below 0.40 kg ha−1resulted in substantially greater wild oat seed production compared to the recommended rate.


Weed Science ◽  
1988 ◽  
Vol 36 (1) ◽  
pp. 37-42 ◽  
Author(s):  
Don W. Morishita ◽  
Donald C. Thill

Field experiments were conducted in 1983 and 1984 to measure the interference of wild oat (Avena fatuaL. # AVEFA) removed at various stages of development (two to three leaves, two to three tillers, two nodes, and heading), plus treatments with wild oat not removed, and wild oat-free stands on the growth and yield of spring barley (Hordeum vulgareL. ‘Advance’). The final plant density of barley and wild oat was 160 and 170 plants/m2, respectively. Based on aboveground barley biomass and yield, the critical duration of wild oat interference began about the two-node stage and continued until maturity of the barley. Wild oat interference reduced barley biomass, the number of tiller heads/plant, tiller heads/unit area, and tiller grain yield, but not the number or grain yield of the main-stem heads. Wild oat did not affect the soil matric potential or the barley total plant and soil nitrogen contents. However, wild oat did reduce total water and turgor potential in barley at the boot stage of growth, which may have affected tiller head formation.


1999 ◽  
Vol 79 (2) ◽  
pp. 303-312 ◽  
Author(s):  
J. T. O'Donovan ◽  
J. C. Newman ◽  
R. E. Blackshaw ◽  
K. N. Harker ◽  
D. A. Derksen ◽  
...  

Understanding the relative competitiveness and seed germination patterns of herbicide-resistant weeds has implications for integrated weed management. Replacement series experiments were conducted in the greenhouse to compare interspecific competition among two triallate/difenzoquat susceptible (S) and 10 resistant (R) wild oat (Avena fatua L.) populations. Each series included monocultures of each of the populations and three mixtures at relative S:R proportions of 3:1, 1:1 and 1:3. Shoot dry weight tended to be greater in the R than S populations, but results were not always statistically significant at the 5% level. Leaf area was more variable, but in most cases did not differ between R and S populations. With a few exceptions, relative crowding coefficients for shoot dry weight and leaf area were similar for S and R populations indicating little or no differences in competitiveness. In field experiments where two S and five R populations were grown in competition with wheat (Triticum aestivum L.), two of the R populations produced significantly (P < 0.05) greater shoot dry weight and seed yield than the S populations. Otherwise populations did not differ significantly. In seed germination studies, the S populations consistently displayed lower cumulative germination than the R populations. The higher seed germination associated with the R populations suggests that producers should be able to manage these populations effectively with a combination of alternative herbicides and cultural practices. Key words: Avena fatua, herbicide resistance, triallate, difenzoquat, relative competitiveness, seed germination


2013 ◽  
Vol 57 (4) ◽  
pp. 773-777 ◽  
Author(s):  
T. Inoue ◽  
Y. Yamauchi ◽  
A. H. Eltayeb ◽  
H. Samejima ◽  
A. G. T. Babiker ◽  
...  

Sign in / Sign up

Export Citation Format

Share Document