<<

BIBLIOGRAPHY WILD , STERILIS, A. FATUA, A. BARBATA, WEB OF SCIENCES JANUARY 21, 2007 KLAUS AMMANN

(1970). "Wild Oats and Tri-Allate." Agriculture 77(3): 144-&.://A1970F665200011

(1972). "Wild Oats and Wro." Agriculture 79(8): 364-&.://A1972N353100008

(1981). "Wild Oats an Annual Weed but a Perennial Problem." Milling Feed and Fertiliser 164(7): 30-31.://A1981LY33500005

(2005). "Caryopsis size and genotype effects on wild ( L.) - tame oat (Avena sativa L.) competition." Canadian Journal of Science 85(1): 166-166.://000228426900040

(2005). "Influence of wild oat (Avena fatua) time of emergence and density on tame oat (Avena sativa)." Canadian Journal of Plant Science 85(1): 165-165.://000228426900037

Adkins, D. W., S. Tanpipat, et al. (1998). "Influence of environmental factors on glyphosate efficacy when applied to Avena fatua or Urochloa panicoides." Weed Research 38(2): 129-138.://000073843100006 Glasshouse experiments were conducted to determine the effects of soil moisture content, irradiance, temperature and relative humidity on the efficacy of glyphosate applied to six isogenic lines of Avena fatua L. (wild oat) and four near- isogenic lines of Urochloa panicoides Beauv. (liverseed grass). The variables examined were four soil moisture conditions (29%, 42%, 55% and 100% of field capacity), two levels of irradiance (400 and 800 mu mol m(-2) s(-1)), two levels of relative humidity (> 92% and 65%) and four temperature regimes (20/15, 25/20, 30/25 and 35/30 degrees C: day/night), representing the environmental conditions of winter or summer fallows in the north-east grain region of . The efficacy of 360 g acid equivalent ha(-1) glyphosate was greatest under well-watered (100% of field capacity), warm (30/25 degrees C for A. fatua and 35/30 degrees C for U. panicoides) and humid (> 92/90%) conditions, The efficacy was least under severe water stress (29% of field capacity), warm (30/25 degrees C for A. fatua and 35/30 degrees C for U. panicoides) and moderately humid (65/60%) conditions. Efficacy was not altered by the level of irradiance nor was it different between isogenic lines.

Adkins, S. W. and A. L. Adkins (1994). "Effect of Potassium-Nitrate and Ethephon on Fate of Wild Oat (Avena-Fatua) in Soil." Weed Science 42(3): 353-357.://A1994PD56000005 A study under glasshouse conditions determined the effect of germination stimulants, potassium nitrate and ethephon, on emergence of several pure lines of wild oat from soil. Effect of the two chemicals on stimulating emergence was variable and no synergism noted when used together. Examination of exhumed seed showed that these treatments reduced seed viability. By counting the number of dead and adding to those that emerged, total seed removed from the soil was obtained. All treatments removed seed from the soil seedbank, especially when seed was old, near the soil surface, was primary seed, or from low or intermediate dormancy lines. However, because of the insensitivity of a large proportion of the seed and the high rates required it is unlikely that either potassium nitrate or ethephon has practical use for field application.

Adkins, S. W., M. Loewen, et al. (1986). "Variation within Pure Lines of Wild Oats (Avena-Fatua) in Relation to Degree of Primary Dormancy." Weed Science 34(6): 859-864.://A1986E740900009

Adkins, S. W., M. Loewen, et al. (1987). "Variation within Pure Lines of Wild Oats (Avena-Fatua) in Relation to Temperature of Development." Weed Science 35(2): 169-172.://A1987G444500007

Adkins, S. W., J. M. Naylor, et al. (1984). "The Physiological-Basis of Seed Dormancy in Avena-Fatua .5. Action of Ethanol and Other Organic-Compounds." Physiologia Plantarum 62(1): 18-24.://A1984TM20100004

Adkins, S. W. and J. D. Ross (1981). "Studies in Wild Oat Seed Dormancy .1. the Role of Ethylene in Dormancy Breakage and Germination of Wild Oat (Avena-Fatua L)." Plant Physiology 67(2): 358-362.://A1981LD91300032

Adkins, S. W. and G. M. Simpson (1988). "The Physiological-Basis of Seed Dormancy in Avena-Fatua .9. Characterization of 2 Dormancy States." Physiologia Plantarum 73(1): 15-20.://A1988N771200003

Adkins, S. W., G. M. Simpson, et al. (1984). "The Physiological-Basis of Seed Dormancy in Avena-Fatua .3. Action of Nitrogenous Compounds." Physiologia Plantarum 60(2): 227-233.://A1984SG89400021

Adkins, S. W., G. M. Simpson, et al. (1984). "The Physiological-Basis of Seed Dormancy in Avena-Fatua .4. Alternative Respiration and Nitrogenous Compounds." Physiologia Plantarum 60(2): 234-238.://A1984SG89400022

Adkins, S. W., G. M. Simpson, et al. (1984). "The Physiological-Basis of Seed Dormancy in Avena-Fatua .6. Respiration and the Stimulation of Germination by Ethanol." Physiologia Plantarum 62(2): 148-152.://A1984TR08700005

Adkins, S. W., G. M. Simpson, et al. (1985). "The Physiological-Basis of Seed Dormancy in Avena-Fatua .7. Action of Organic- Acids and Ph." Physiologia Plantarum 65(3): 310-316.://A1985AVM9100016

Adkins, S. W., S. J. Symons, et al. (1988). "The Physiological-Basis of Seed Dormancy in Avena-Fatua .8. Action of Malonic-Acid." Physiologia Plantarum 72(3): 477-482.://A1988M764000006

2

Ahrens, W. H. and R. J. Ehr (1991). "Tridiphane Enhances Wild Oat (Avena-Fatua) Control by Atrazine-Cyanazine Mixtures." Weed Technology 5(4): 799-804.://A1991HF52400017 Wild oat, a competitive weed in com, can be controlled by atrazine applied postemergence but at rates causing carryover problems on high pH soils rotated to susceptible crops. Experiments at Fargo, ND and Barnesville, MN were established in corn using atrazine rates ranging from 0.14 to 0.84 kg ai ha-1 supplemented with cyanazine to provide 1.68 kg ai ha-1 total triazine . Atrazine at 0.84 kg ha-1 plus cyanazine at 0.84 kg ha-1 controlled 2- to 3- wild oat 40 to 55%, but control increased to 90 to 95% by adding tridiphane at 0.56 or 0.84 kg ai ha-1. In greenhouse experiments, tridiphane synergistically increased wild oat control by atrazine and cyanazine. Tridiphane applied 1 or 2 d before atrazine or before an atrazine-cyanazine mixture generally controlled wild oat similar to the tridiphane-triazine tank mix.

Aibar, J., M. J. Ochoa, et al. (1991). "Field Emergence of Avena-Fatua L and Avena-Sterilis Ssp Ludoviciana (Dur) Nym in Aragon, Spain." Weed Research 31(1): 29-32.://A1991EU79600004 Emergence of Avena fatua and A. sterilis ssp. ludoviciana infesting winter cereals during two years and at two sites in Aragon began after sowing in late October and continued for 23 weeks, with 75% of seedlings appearing in the first 9 weeks. The start of emergence was associated with a fall in minimum air temperature to below 9-degrees-C and a maximum of less than 20-degrees-C. Soil moisture was not limiting, and during winter flushes of seedlings tended to be associated with rises in mean temperature. In contrast with results from other latitudes, A. fatua emerged mainly in autumn at the same time as A. sterilis ssp. ludoviciana.

Akalehiywot, T. and J. D. Bewley (1977). "Effects of Dehydration-Rehydration Treatments on Protein-Synthesis in Oat Grains (Avena-Fatua Cv Harmon) During Germination." Plant Physiology 59(6): 33-33.://A1977DM92400185

Akey, W. C. and I. N. Morrison (1983). "Effect of Moisture Stress on Wild Oat (Avena-Fatua) Response to Diclofop." Weed Science 31(2): 247-253.://A1983QH14300020

Akey, W. C. and I. N. Morrison (1984). "Effects of Soil-Moisture on the Vegetative Growth of Wild Oat (Avena-Fatua)." Weed Science 32(5): 625-630.://A1984TL30200012

Allard, R. W., G. R. Babbel, et al. (1972). "Evidence for Coadaptation in Avena-Barbata." Proceedings of the National Academy of Sciences of the United States of America 69(10): 3043-&.://A1972N749600066

Allard, R. W., P. Garcia, et al. (1993). " of Multilocus Genetic-Structure in Avena-Hirtula and Avena-Barbata." Genetics 135(4): 1125-1139.://A1993MH93800019 , an autotetraploid grass, is much more widely adapted than Avena hirtula, its diploid ancestor. We have determined the 14-locus genotype of 754 diploid and 4751 tetraploid from 10 and 50 Spanish sites, respectively. Allelic diversity is much greater in the tetraploid (52 alleles) than in the diploid (38 alleles): the extra alleles of the tetraploid were present in nonsegregating heteroallelic quadriplexes. Seven loci were monomorphic for the same allele (genotypically 11) in all populations of the diploid: five of these loci were also monomorphic for the same allele (genotypically 1111) in all populations of the tetraploid whereas two loci each formed a heteroallelic quadriplex (1122) that was monomorphic or predominant in the tetraploid. Seven of the 14 loci formed one or more highly successful homoallelic and/or heteroallelic quadriplexes in the tetraploid. We attribute much of the greater heterosis and wider adaptedness of the tetraploid to favorable within-locus interactions and interlocus (epistatic) interactions among alleles of the loci that form heteroallelic quadriplexes. It is difficult to account for the observed patterns in which genotypes are distributed ecogeographically except in terms of natural selection favoring particular alleles and genotypes in the specific genetic structure of different local populations as well as the adaptive landscape of both the diploid and tetraploid.

Andrews, C. J. and G. M. Simpson (1969). "Dormancy Studies in Seed of Avena-Fatua .6. Germinability of Immature Caryopsis." Canadian Journal of Botany 47(12): 1841-&.://A1969F193500003

Andrews, T. S., I. N. Morrison, et al. (1998). "Monitoring the spread of ACCase inhibitor resistance among wild oat (Avena fatua) patches using AFLP analysis." Weed Science 46(2): 196-199.://000073371800009 The relative genetic similarity of 37 wild oat samples was determined using amplified fragment length (AFLP) analysis. These data were compared with the herbicide-resistance characteristics of each sample to determine the importance of mutation and gene flow in the spread of Acetyl-CoA Carboxylase (ACCase) inhibitor resistance. There was a strong association between the genotypic clustering of samples, their herbicide-resistance characteristics, and their field of origin. Up to eight separate patches in a field were genetically similar, confirming that gene flow by seed movement is important in the distribution of resistant (R) wild oat seed. A greater emphasis on field scouting and treatment of wild oat patches could reduce the spread of resistance within fields. Samples collected from different fields were also found to be genetically similar, indicating that R wild oat was spread between fields. Improved sanitation of tillage and harvesting equipment and the use of certified seed could limit such seed movement.

Armstrong, L. J. and S. W. Adkins (1998). "Variation in photoperiod response of different isogenic lines of wild oats (Avena fatua)." Weed Science 46(1): 39-47.://000072323200007 For wild oats, persistence characters such as seed production and seed dormancy may be affected by generic variation and several environmental factors during the development of the parental plant. However, the effect of varying photoperiods on such characters is unclear. Previous studies have concentrated on natural populations and have nor studied the genetic variability within the population. Consequently a study was conducted to examine how photoperiod (10, 12, 14, 16, and 18 h at 20 C) influenced the persistence of different members of a synthetic population of wild oats. This synthetic population consisted of several isogenic lines with differing degrees of seed dormancy that originated from one region in . All lines were photoperiod sensitive, quantitative long day planes, with an increase in time to maturity for all lines (about 78 to 213 d) when photoperiod was reduced from 18 to 10 h. The lines within this synthetic population (with the exception of one, M73) showed a similar degree of photoperiod sensitivity. This may be expected, as 3

the lines are from one region with the same photoperiod environment. It could be concluded that such a photoperiod sensitivity is a beneficial character that has allowed persistence of these lines in this region. When photoperiod was reduced from 18 to 10 h, plants produced fewer seeds (about 141 to 61 primary seeds per plant) with a higher degree of seed dormancy (about 88 to 54% germination in 10(-4) M gibberellic acid; GA(3)). The reduced seed production under the shortest photoperiod (10 h) was due to poor panicle exsertion, which resulted in poor development of basal florets. There was no consistent relationship between photoperiod and other plant characters such as tiller production, caryopsis weight, and water content. However, significant relationships were found between seed dormancy characters and other plant characters such as maturity time, caryopsis weight, and water content under certain photoperiod treatments.

Ashford, R. and J. Holroyd (1977). "Postemergence Effects of Different Distribution Patterns of Tri-Allate Granules on Avena-Fatua L and Plants." Weed Research 17(6): 415-422.://A1977EM35700011

Aston, M. J. and M. M. Jones (1976). "Study of Transpiration Surfaces of Avena-Sterilis L Var Algerian Using Monosilicic Acid as a Tracer for Water-Movement." Planta 130(2): 121-129.://A1976BU21700004

Atwood, W. M. (1914). "A physiological study of the germination of Avena fatua." Botanical Gazette 57: 0386-0414.://000202130500025

Auld, B. A. (1988). "Dynamics of Pasture Invasion by 3 Weeds, Avena-Fatua L, Carduus-Tenuiflorus Curt. and Onopordum- Acanthium L." Australian Journal of Agricultural Research 39(4): 589-596.://A1988Q306900007

Aung, T. and H. Thomas (1976). "Transfer of Mildew Resistance from Wild Oat Avena-Barbata into Cultivated Oat." Nature 260(5552): 603-604.://A1976BM86700032

Aung, T. and H. Thomas (1978). "Structure and Breeding-Behavior of a Translocation Involving the Transfer of Mildew Resistance from Avena-Barbata Pott into the Cultivated Oat." Euphytica 27(3): 731-739.://A1978GH66700011

Aung, T., H. Thomas, et al. (1977). "Transfer of Gene for Mildew Resistance from Avena-Barbata (4x) into Cultivated Oat Avena- Sativa by an Induced Translocation." Euphytica 26(3): 623-632.://A1977EQ02900011

Bachthal.G (1966). "Present Distribution of Wild Oat (Avena Fatua L.) in Federal German Republic." Weed Research 6(3): 193- &.://A19668258200001

Baerg, R. J., J. W. Gronwald, et al. (1996). "Antagonism of diclofop control of wild oat (Avena fatua) by tribenuron." Weed Science 44(3): 461-468.://A1996VE21000003 Tribenuron antagonized diclofop control of wild oat in greenhouse studies, Wild oat shoot fresh weight was reduced 79% when plants were treated with diclofop at 1.12 kg ha(-1), but only 46% when treated with diclofop at 1.12 kg ha(-1) plus tribenuron at 18 g ha(-1). Increasing diclofop rate increased control of wild oat but did not overcome the antagonism. Separating the applications of diclofop and tribenuron by as little as 12 s reduced the antagonism by 40%, indicating the importance of close proximity of the two on the leaf surface for the antagonistic response. The inhibitory effect of diclofop at its target site (acetyl-CoA carboxylase) was not reduced by tribenuron. Metabolism of C-14-diclofop by wild oat leaves was not altered quantitatively or qualitatively by tribenuron. Tribenuron had no effect on spray retention or absorption of C-14-diclofop by wild oat leaves, and only slightly decreased the total amount of radiolabel translocated out of the treated zone on the leaf. However, tribenuron decreased basipetal translocation of diclofop from the treated zone by approximately 20%. The ability of tribenuron to reduce basipetal translocation of diclofop to meristematic regions of wild oat may be a factor in antagonism.

Baker, G. H., M. Amato, et al. (2003). "Influences of Aporrectodea trapezoides and A-rosea (Lumbricidae) on the uptake of nitrogen and yield of oats (Avena fatua) and lupins (Lupinus angustifolius)." Pedobiologia 47(5-6): 857-862.://000188421700069 The earthworms, Aporrectodea rosea and A. trapezoides (Lumbricidae), occur commonly in soils used for grain production in south-eastern Australia, but their distributions are patchy and their abundance low. In a field experiment, inoculations of A. trapezoides, but not A. rosea, increased the growth of oats at early tillering and oat straw and grain yield at harvest. Neither earthworm species influenced the growth and yield of lupins. At early tillering of the oats and concurrent flowering of the lupins, the amounts of N-15 (from fertiliser applied the year before) taken up by both oats and lupins were greater in the presence of A. trapezoides, but not A. rosea. However, there were no differences in the % N in lupin and oat plants between the earthworm treatments. At harvest, there were no differences in the amounts of N-15 taken up by the lupin straw, nor were there in the % N in straw or grain for either lupins or oats between earthworm treatments, The N-15 contents of the oat straw and the lupin and oat grain were higher in the presence of both earthworm species when compared with controls. More than 90 % of the N in the lupins was fixed symbiotically. The proportion of fixed N was reduced slightly by A. trapezoides. Inorganic N was analysed at various depths in the soil profile at sowing and early tillering and flowering. An earthworm effect on soil inorganic N was recorded in only one instance - at early tillering / flowering, when soil inorganic N was slightly higher at 0.6-1 m depth in the presence of A. trapezoides. Management practices, which improve earthworm abundance, in particular that of A. trapezoides, in cropping soils in southern Australia may enhance productivity of some, but not all, crops.

Baker, T. H. (1985). "Editorial Wild-Oats - Carmack,Edward,Ward and Tennessee Politics - Majors,Wr." Journal of American History 72(1): 164-164.://A1985ARM2500061

Baldwin, J. H. (1979). "Chemical Control of Wild Oats and Blackgrass - Review Based on Results from Adas Agronomy Trials." Adas Quarterly Review(33): 69-101.://A1979HW85100001

4

Banting, J. D. (1966). "Factors Affecting Persistence of Avena Fatua." Canadian Journal of Plant Science 46(5): 469-&.://A19668285900002

Banting, J. D. (1966). "Studies on Persistence of Avena Fatua." Canadian Journal of Plant Science 46(2): 129-&.://A19667467600005

Barnard, J. (1990). "Okeeffe 'Wild Oats'." Tls-the Times Literary Supplement(4538): 321-321.://A1990CV77300026

Barranco, A., D. J. Morgan, et al. (1978). "Note on Nutritive-Value of Wild Oats." Animal Production 27(AUG): 129-132.://A1978FU09900018

Barroso, J., C. Fernandez-Quintanilla, et al. (2004). "Spatial stability of ssp ludoviciana populations under annual applications of low rates of imazamethabenz." Weed Research 44(3): 178-186.://000221355400004 Long-term experiments were conducted in two winter barley fields in central Spain to determine the spatial stability of Avena sterilis ssp. ludoviciana populations under annual applications of low rates of imazamethabenz herbicide. Weed density was sampled every year (over 5 years in the first field and over 3 years in the second) on the same grid locations prior to herbicide application. Although weed patches were stable in their location, weed density decreased in most of the years. In the first field, the populations decreased exponentially over the 5-year period. The rates of population decline were dependent on the initial density of the population, being higher for the central core of the patches and lower for the low-density areas. Under the conditions present in this experiment, it was possible to reduce heavy weed patches (up to 1200 seedlings m(-2)) down to relatively safe levels (18 seedlings m(-2)) in a period of 3 years using a density-specific control programme, applying low rates of herbicides when weed densities were below a given level (1000 seedlings m(-2)). However, under adverse environmental conditions, half rates of the herbicide failed to control the weed populations adequately. The stability of the location of patches of A. sterilis ssp. ludoviciana suggest that weed seedling distributions mapped in one year are good predictors of future seedling distributions. However, the actual densities established each year will depend on the control level achieved the previous year and the climatic conditions present during the establishment period.

Barroso, J., L. Navarrete, et al. (2006). "Dispersal of Avena fatua and Avena sterilis patches by natural dissemination, soil tillage and combine harvesters." Weed Research 46(2): 118-128.://000235843400002 The dispersal of Avena spp. (A. fatua and A. sterilis) by natural dissemination and by agricultural operations was studied in four experiments conducted in Spain and Britain. Natural dispersal was very limited, with a maximum dispersal distance of 1.5 m. Dispersal was higher in the geographic direction that was downwind than in any of the other three geographic directions. Although plant movement was very small under no-tillage, an annual patch displacement of 2-3 m in the tillage direction was observed under conventional soil tillage. Ploughing downhill resulted in much larger dispersal distances than ploughing uphill. In the crops studied, combine harvesters dispersed few Avena spp. seeds, because of the fact that the plants had shed most of their seeds (> 90%) before harvest. The percentage of seeds available to be dispersed by the combine was dependent on the harvest time. Although combine harvesting may not contribute much to short-distance dispersal, it may play an important role in long-distance dispersal. In our studies, isolated plants were located up to 30 m from the original sources. This small proportion may have a significant effect on the distribution of the weed within a field, acting as foci for new patches.

Barroso, J., D. Ruiz, et al. (2005). "Comparison of sampling methodologies for site-specific management of Avena sterilis." Weed Research 45(3): 165-174.://000228488400001 The ability to manage weed infestations in a spatially precise manner requires efficient and accurate methods of mapping weed distributions. A study was conducted to compare four different ground-based methods for collecting georeferenced information on infestations of Avena sterilis in winter wheat and barley. Sampling was performed at harvest by scoring panicle density, either from the ground or from a combine, by counting the number of panicle contacts with a stick moved horizontally over the crop canopy by an observer walking through the field, and by sampling A. sterilis seed rain on the ground. No significant differences were observed among the populations estimated by the four methods. A partial budget analysis of the in-season costs and benefits of spraying patches using these methods showed that visual scoring from the combine was the most appropriate method for the creation of weed management maps to be used for patch spraying in the following season. As a large variety of spatial patterns may be found in fields, the recommended sampling method might be field-specific and optimality should be verified for general use.

Barton, D. L., D. C. Thill, et al. (1992). "Integrated Wild Oat (Avena-Fatua) Management Affects Spring Barley (Hordeum-Vulgare) Yield and Economics." Weed Technology 6(1): 129-135.://A1992HM60800022 The effect of barley seeding rate and row spacing, and triallate, diclofop, and difenzoquat herbicide rate on barley grain yield and quality, and wild oat control were evaluated in field experiments near Bonners Ferry, Idaho, in 1989 and 1990. The purpose of the study was to develop integrated control strategies for wild oat in spring barley. Barley row spacing (9 and 18 cm) did not affect barley grain yield. Barley grain yield was greatest when barley was seeded at 134 or 201 kg ha-1 compared to 67 kg ha-1. Wild oat control increased as wild oat herbicide rate increased and barley grain yield was greatest when wild oat herbicides were applied. However, barley grain yield was similar when wild oat biomass was reduced by either 65 or 85% by applications of half and full herbicide rates, respectively. Net return was greatest when the half rate of herbicide was applied to 100 wild oat plants per m2 and was greatest when half or full herbicide rates were applied to 290 wild oat plants per m2. Net return increased when the seeding rate was increased to 134 or 201 kg ha-1 when no herbicide was applied and when 290 wild oat plants per m2 were present.

Batish, D. R., H. P. Singh, et al. (2002). "Allelopathic effects of parthenin against two weedy species, Avena fatua and Bidens pilosa." Environmental and Experimental Botany 47(2): 149-155.://000174134500006 Parthenin is a natural constituent of Parthenium hysterophorus with phytotoxic and allelopathic properties. Its effect on two weedy species viz. Avena fatua and Bidens pilosa was studied with a view to explore its herbicidal potential. Germination 5

of both the weeds was reduced with increasing concentration of parthenin and a dose-response relationship was observed. This provided information on LC50 and Inhibition threshold concentrations of parthenin that could be useful for future studies. Further, parthenin also inhibited the growth of both the weeds in terms of root and shoot length and seedling dry weight. Inhibition of root growth was greater than that of shoot growth. Similar observations were made when the test weeds were grown in soil amended with different concentrations of parthenin. In addition to growth, there was a reduction of chlorophyll content in the growing seedlings. It also caused water loss in the weedy species. The study, therefore, reveals that parthenin exerts an inhibitory effect on the growth and development of both weeds and can be further explored as a herbicide for future weed management strategies. (C) 2002 Elsevier Science B.V. All rights reserved.

Baum, B. R. (1968). "On Some Relationships between Avena Sativa and a Fatua (Gramineae) as Studied from Canadian Material." Canadian Journal of Botany 46(8): 1013-&.://A1968B646300009

Baum, B. R. (1969). "Role of Lodicule and Epiblast in Determining Natural Hybrids of Avena Sativa X Fatua in Cultivated Oats." Canadian Journal of Botany 47(1): 85-&.://A1969C601300013

Baum, B. R. and V. E. Hadland (1973). "Scanning Electron-Microscopic Study of Epicuticular Waxes of Glumes in Avena-Magna, a- Murphyi, and Avena-Sterilis." Canadian Journal of Botany-Revue Canadienne De Botanique 51(12): 2381-&.://A1973S240800014

Bechler, R. (1995). "Okeeffe 'Wild Oats'." Tls-the Times Literary Supplement(4825): 20-20.://A1995RW64400034

Beckie, H. J., L. M. Hall, et al. (2005). "Patch management of herbicide-resistant wild oat (Avena fatua)." Weed Technology 19(3): 697-705.://000232674700027 A study was conducted at a 64-ha site in western Canada to determine how preventing seed shed from herbicide-resistant wild oat affects patch expansion over a 6-yr period. Seed shed was prevented in two patches and allowed to occur in two patches (nontreated controls). Annual patch expansion was determined by seed bank sampling and mapping. Crop management practices were performed by the grower. Area of treated patches increased by 35% over the 6-yr period, whereas nontreated patches increased by 330%. Patch expansion was attributed mainly to natural seed dispersal (nontreated) or seed movement by equipment at time of seeding (nontreated and treated). Extensive seed shed from plants in nontreated patches before harvest or control of resistant plants by alternative herbicides minimized seed movement by the combine harvester. Although both treated and nontreated patches were relatively stable over time in this cropping system, preventing seed production and shed in herbicide-resistant wild oat patches can markedly slow the rate of patch expansion.

Beckie, H. J. and F. A. Holm (2002). "Response of wild oat (Avena fatua) to residual and non-residual herbicides in canola (Brassica napus) in western Canada." Canadian Journal of Plant Science 82(4): 797-802.://000179839400025 It has been stated that soil residual herbicides, by controlling successive flushes of weeds, increase effective kill (efficacy) over the growing season, and thus impose a higher selection pressure for resistance in weeds than non-residual herbicides. To investigate this issue, the responses of wild oat to increasing rates of residual and non-residual herbicides in canola and wild oat recruitment in the following year were examined in a field study conducted in Saskatchewan, Canada, from 1997 to 2000. The rate-response curves of the wild oat variables indicated that efficacy of the soil residual herbicides, ethalfluralin and triallate, and of the non-residual herbicide, glufosinate, was generally lower than that of imazamox/imazethapyr (residual), sethoxydim, and glyphosate (non-residual). Emergence of wild oat in spring wheat (Triticum aestivum L.) grown in the following year did not differ among herbicides applied in the preceding crop year, nor was there a significant herbicide by rate interaction. The results suggest that the soil residual activity of these herbicides does not strongly influence selection pressure, estimated by reduction in wild oat seed return in canola.

Beckie, H. J. and K. J. Kirkland (2003). "Implication of reduced herbicide rates on resistance enrichment in wild oat (Avena fatua)." Weed Technology 17(1): 138-148.://000181848500021 Model simulations predict that lowering herbicide efficacy by reducing the application rate would slow the rate of enrichment of herbicide-resistant individuals in a weed population, but the resulting increase in density of susceptible plants would reduce crop yield and increase the weed seed bank. A study was conducted at three sites in Saskatchewan, Canada, from 1997 to 2000 to examine the implication of reduced rates of acetyl-CoA carboxylase (ACCase) inhibitors in a diverse 4-yr crop rotation, in conjunction with variable crop seeding rates, on the enrichment of resistant wild oat in a mixed (resistant and susceptible) population. Main-plot treatments were crop (barley, canola, field pea, and spring wheat), subplot treatments were crop seeding rate (recommended and high), and sub-subplot treatments were ACCase inhibitor rate (0, 0.33, 0.67, and 1.0 times the recommended rate). Herbicide rate frequently interacted with seeding rate in affecting wild oat seedling density, seed return, the viable fraction of the weed seed bank, and crop seed yield. As simulation models predict, reduced herbicide efficacy decreased the proportion of resistant individuals in the population. The high crop seeding rate compensated for a one-third reduction in herbicide rate by limiting total wild oat seed return and by reducing the number of resistant seedlings recruited from the seed bank. The level of resistance in the seed bank can be reduced without increasing the total (resistant plus susceptible) seed bank population by manipulating agronomic practices to increase crop competitiveness against wild oat when ACCase inhibitor rates are reduced to a maximum of two-thirds of that recommended.

Beckie, H. J., A. G. Thomas, et al. (1999). "Nature, occurrence, and cost of herbicide-resistant wild oat (Avena fatua) in small-grain production areas." Weed Technology 13(3): 612-625.://000085867600030 Surveys were conducted across the northern Great Plains of Canada in 1996 and 1997 to determine the nature and occurrence of herbicide-resistant (HR) biotypes of wild oat (Avena fatua). The surveys indicated that resistance to acetyl- CoA carboxylase (ACCase) inhibitors (Group I) occurred most frequently relative to other herbicide groups. Group 1-HR wild oat occurred in over one-half of fields surveyed in each of the three prairie provinces. Of particular concern was the relatively high incidence of multiple-group resistance in wild oat in Saskatchewan and Manitoba, In Saskatchewan, 18% of 6

Group 1-HR populations were also resistant to acetolactate synthase inhibitors (imidazolinones), even though these herbicides were not frequently used, In Manitoba, 27% of fields surveyed had wild oat resistant to herbicides from more than one group. Four populations were resistant to all herbicides registered for use in wheat (Triticum aestivum), Depending on the nature of resistance in wild oat, alternative herbicides available for their central may substantially increase costs to the grower. The cost to growers of managing HR wild oat in Saskatchewan and Manitoba using alternative herbicides is estimated at over $4 million annually. For some HR biotypes, alter-native herbicides either are not available or all have the same site of action, which restricts crop or herbicide rotation options and threatens the future sustainability of small-grain annual cropping systems where these infestations occur.

Beckie, H. J., A. G. Thomas, et al. (2002). "Survey of herbicide-resistant wild oat (Avena fatua) in two townships in Saskatchewan." Canadian Journal of Plant Science 82(2): 463-471.://000175508800030 Beckie, H. J., Thomas, A. G. and Stevenson, F. C. 2002. Survey of herbicide-resistant wild oat (Avena fatua) in two townships in Saskatchewan. Can. J. Plant Sci. 82: 463-471. The nature and occurrence of herbicide resistance in wild oat in annual crops grown in the Grassland and Parkland regions of Saskatchewan were determined in a systematic survey of fields in two townships in 1997. The survey found that over one-half of fields in both townships had populations resistant to Group 1 [acetyl-CoA carboxylase (ACCase) inhibitors], Group 2 [acetolactate synthase (ALS) inhibitors], and/or Group 8 (e.g., triallate, difenzoquat) herbicides. Forty-three percent of fields in the Grassland township and 48% of fields in the Parkland township had Group 1-resistant (HR) wild oat; 30 and 17% of fields in the Grassland and Parkland township, respectively, had populations exhibiting Group 2 resistance, whereas about 15% of fields in both townships had Group 8- HR wild oat. Single- (Groups 1, 2, or 8) and multiple-group resistance (1, 2; 1, 8; 2, 8; 1, 2, 8) were exhibited in populations in fields in both townships. Frequency of occurrence of resistance was not generally affected by farm size. The nature of resistance in wild oat populations is more diverse, differences in distribution and abundance of HR wild oat biotypes between Grassland and Parkland regions are generally less apparent, and occurrence of resistance is more prevalent than documented previously.

Beer, S. C., J. Goffreda, et al. (1993). "Assessment of Genetic-Variation in Avena-Sterilis Using Morphological Traits, Isozymes, and Rflps." Crop Science 33(6): 1386-1393.://A1993MR71200051 Genetic patterns revealed by proximity coefficients may be affected by choice of traits, method of scoring, any subsequent transformations of scores, and choice of proximity coefficient. Our objective was to compare restriction fragment length polymorphism (RFLP), isozyme polymorphism, and variation in qualitative and quantitative morphological traits in a geographically stratified set of 177 accessions of a hexaploid wild oat, Avena sterilis L. Jaccard similarity coefficients (SJ) and Russell and Rao similarity coefficients (SRR) were calculated for all pairs of genotypes from restriction fragments and, separately, from isozymes. Standard taxonomic (DSZ), Mahalanobis, and Good man distances were calculated from 26 morphological trait scores. Clustering of mean DSZ values or mean SJ(RFLP) values between pairs of countries produced similar dendrograms while SRR(isozymes) values resulted in different subgroups. Rankings of within-country diversity were similar for different types of traits but were unrelated to geographic proximity of provenances of the accessions. Proximity coefficients based on the same type of trait (either RFLP, isozyme, or morphology) were highly correlated (Mantel statistic) (0.6-0.9),while correlations for proximities based on different types of traits were less than or equal to 0.35. Isozyme and RFLP-based proximities were poorly correlated, and both were poor predictors of morphological relationships with the highest correlation being -0.35 between SJ(RFLP) and DSZ. While broad patterns of variation revealed by different types of traits were similar in this sample of A. sterilis, differences in pairwise estimates of relationship were sufficiently great to question the exclusive use of one type of trait for sampling and management of plant germplasm collections.

Belles, D. S., D. C. Thill, et al. (2000). "PP-604 rate and Avena fatua density effects on seed production and viability in Hordeum vulgare." Weed Science 48(3): 378-384.://000087829200016 High Avena fatua control costs have caused some Hordeum vulgare growers to use reduced rates of herbicides without fully understanding the consequences. Field studies near Moscow and Genesee, ID, were conducted to determine the effect of A. fatua density and PP-604 rate on A. fatua seed production in H. vulgare and on H. vulgare yield. PP-604 treatments were 25, 50, 100, 150, and 200 (minimum labeled rate) g ha(-1), and five A. fatua densities ranged from 0 to 386 plants m(-2). Visual A. fatua control was greater than 85% with 100 g ha(-1) PP-604 at all locations. Data from 1998 were used to construct nonlinear exponential decay and parabolic models to describe the effect of reduced herbicide rates on viable A. fatua seed production and relative H. vulgare grain yield, respectively. At A. fatua densities of 42 to 138 plants m(-2), 46 to 71% of the minimum labeled rate of PP-604 reduced seed production 95%. However, an estimated 140 to 235 seeds m(-2) were produced at this level of control, which may not ensure a decline in the A. fatua population over the long- term. Hordeum vulgare grain yield was maximum when 70 to 85% of the minimum labeled rate was applied to A. fatua densities of 42 to 138 plants m(-2). A higher rate of PP-604 likely will be required to ensure a decline in A. fatua populations over the long-term than needed to obtain maximum H. vulgare grain yield in a single growing season.

Bergmannova, E. and L. Taimr (1980). "Effect of Benzoylprop-Ethyl on Uptake, Transport and Metabolism of P-32 in Avena-Fatua L and Triticum-Aestivum L." Weed Research 20(4): 243-248.://A1980KM34400009

Bergmannova, E. and L. Taimr (1980). "The Effect of Chlorfenprop-Methyl on P-32 Uptake, Transport and Metabolism in Avena- Fatua L and Hordeum-Vulgare-L." Weed Research 20(1): 1-8.://A1980JW34600001

Bergmannova, E. and L. Taimr (1985). "Effect of Chlorfenprop-Methyl, Flamprop-Isopropyl and Benzoylprop-Ethyl on (Co2)-C-14 Fixation in Avena-Fatua L, Triticum-Aestivum L and Hordeum-Vulgare-L." Weed Research 25(5): 347-353.://A1985ASF0900005

Beyschlag, W., P. W. Barnes, et al. (1988). "Enhanced Uv-B Irradiation Has No Effect on Photosynthetic Characteristics of Wheat (Triticum-Aestivum L) and Wild Oat (Avena-Fatua L) under Greenhouse and Field Conditions." Photosynthetica 22(4): 516-525.://A1988R915100006 7

Billett, D. and R. Ashford (1978). "Differences in Phytotoxic Response of Wild Oats (Avena-Fatua) to Triallate and Trifluralin." Weed Science 26(3): 273-276.://A1978EX61600016

Bingham, I. J. (1995). "A Comparison of the Dynamics of Root-Growth and Biomass Partitioning in Wild Oat (Avena-Fatua L) and Spring Wheat." Weed Research 35(1): 57-66.://A1995RA78900009 The dynamics of early root growth and dry matter partitioning were compared in spring wheat (Triticum aestivum L.) and wild oat (Avena fatua L.) grown in solution culture. Total root length was greater in wheat than wild oat throughout the experiment; a result of a greater number of seminal axes and greater production of lateral root length per axis. The final number of adventitious roots was greater in wheat than in wild oat, but their length was similar. Relative growth rates were also similar as was shoot:root dry weight ratio and rate of root respiration. However, wheat used the dry matter partitioned to its roots more efficiently, producing a greater specific root length (SRL, length per unit weight). Caution must be exercised when relating these results to plants growing and competing in the field, but three general points are raised. First, the initial number of seminal axes can have a profound effect on the rate of early root development; second, the adventitious root system of wild oat is not inherently more vigorous than that of wheat; and third, future studies should compare SRL of wheat and wild oat in the field. If differences similar to those in the present study are found they may contribute to the greater competitive ability of wheat.

Blackshaw, R. E., J. T. Odonovan, et al. (1996). "Response of triallate-resistant wild oat (Avena fatua) to alternative herbicides." Weed Technology 10(2): 258-262.://A1996UP85600003 Wild oat populations resistant to triallate have been identified in Alberta. Dose response experiments were conducted in the greenhouse to determine if triallate-resistant wild oat was controlled by other selective wild oat herbicides. Triallate- resistant wild oat populations were effectively controlled by atrazine, ethalfluralin, fenoxaprop-P, flamprop, imazamethabenz, and tralkoxydim. EPTC and cycloate, which are chemically related to triallate, differed in their efficacy on triallate-resistant wild oats. EPTC at the 0.25x field use rate was more efficacious on triallate-resistant than triallate- susceptible wild oat. In contrast, cycloate at the 0.25 to 0.5x field use rate was less efficacious on triallate-resistant than susceptible wild oats. At higher rates, both EPTC and cycloate killed triallate-resistant wild oat populations. Growers have several herbicide choices to selectively control triallate-resistant wild oat in prairie field crops but should plan to rotate herbicides among different chemical families and adopt integrated weed management practices to reduce the risk of these wild oat populations developing resistance to other wild oat herbicides.

Blair, A. M. (1978). "Some Studies on the Sites of Uptake of Chlortoluron, Isoproturon and Metoxuron by Wheat, Avena-Fatua and Alopecurus-Myosuroides." Weed Research 18(6): 381-387.://A1978GN82600010

Blanco-Moreno, J. M., L. Chamorro, et al. (2006). "Spatial and temporal patterns of rigidum-Avena sterilis mixed populations in a cereal field." Weed Research 46(3): 207-218.://000237354100003 Through a detailed case study of a two-species (Lolium rigidum and Avena sterilis) weed community at contrasting scales, this paper examined factors that affect weed distribution across space and time in a commercial wheat field in north-east Spain. A. sterilis showed relatively stable spatial distribution and spatial structure of its population over time at large scale, with well-defined patches, although weed density rose quickly. L. rigidum showed poorly defined patches that were not stable across time. Interaction between species could explain to some degree the spatial distribution at large scale: a negative relationship was detected between the spatial structures of both weed populations. At fine scale, both species showed a clear interaction effect from primary dispersal (more important in A. sterilis) and secondary dispersal from combine harvesting (more important in L. rigidum).

Bourgeois, L., N. C. Kenkel, et al. (1997). "Characterization of cross-resistance patterns in acetyl-CoA carboxylase inhibitor resistant wild oat (Avena fatua)." Weed Science 45(6): 750-755.://000071078400003 The purpose of this study was to determine cross-resistance patterns among wild oat lines resistant to acetyl-CoA carboxylase (ACCase) inhibitors and to determine which, if any, cross-resistant type was more common than another Discriminatory concentrations of two aryloxyphenoxy-propionates (APP) and three cyclohexanediones (CHD) were determined using a petri-dish bioassay. These concentrations were then applied to 82 resistant wild oat lines identified in previous studies. In addition, two resistant standards (UM1 and UM33) and a susceptible standard (UM5) were included in the experiments. Coleoptile lengths expressed as percentages of untreated controls were used to assess the level of resistance to each herbicide. Large variations were observed among wild oat lines and herbicides. However, cluster analysis summarized the relationship between the five herbicides (variables) and the wild oat lines into three main cross- resistance types. Type A included wild oat lines with high resistance to APP herbicides and no or low resistance to CHD herbicides. Types B and C included those with low to moderate resistant and high levels of resistance to all five herbicides, respectively. Type C was the most common cross-resistance type. Relationships among herbicides were determined using pairwise correlation and principal component analysis (PCA). All correlations were high between APP herbicides and between CHD herbicides but not between APP and CHD herbicides. The first two axes of the PCA accounted for 88.4% of the total variance, with the first axis correlated to the CHD herbicides and the second axis correlated to the APP herbicides. In the PCA, wild oat lines were segregated into the three types identified in the cluster analysis. Although CHD and APP herbicides bind at the same region on the ACCase, resistant wild oat lines respond differently to them.

Bowden, B. A. and G. Friesen (1967). "Competition of Wild Oats (Avena Fatua L) in Wheat and Flax." Weed Research 7(4): 349- &.://A1967A536200009

Bradow, J. M., W. J. Connick, et al. (1990). "Germination Stimulation in Wild Oats (Avena-Fatua L) by Synthetic Strigol Analogs and Gibberellic-Acid." Journal of Plant Growth Regulation 9(1): 35-41.://A1990CE01000006

Brewster, B. D., A. P. Appleby, et al. (1977). "Control of Italian Ryegrass and Wild Oats in Winter-Wheat with Hoe 23408." Agronomy Journal 69(6): 911-913.://A1977EG05500004 8

Brezeanu, A. G., D. G. Davis, et al. (1976). "Ultrastructural Effects and Translocation of Methyl-2-(4-(2,4-Dichlorophenoxy)- Phenoxy)Propanoate in Wheat (Triticum-Aestivum) and Wild Oat (Avena-Fatua)." Canadian Journal of Botany-Revue Canadienne De Botanique 54(17): 2038-2048.://A1976CG85000006

Briggle, L. W., R. T. Smith, et al. (1975). "Protein Concentration and Amino-Acid Composition of Avena-Sterilis L Groats." Crop Science 15(4): 547-549.://A1975AN59700029

Brodny, U., L. W. Briggle, et al. (1976). "Reaction of United-States Crown Rust Resistant Oat Selections and Israeli Avena Sterilis Selections to Puccinia-Coronata Var Avenae." Plant Disease Reporter 60(11): 902-906.://A1976CM91600002

Brodny, U., I. Wahl, et al. (1983). "Factors Affecting the Survival of Physiologic Races of -Avenae on Avena- Sterilis in Israel." Phytopathology 73(2): 363-363.://A1983QC77800076

Brodny, U., I. Wahl, et al. (1988). "Factors Conditioning Dominance of Race 276 of Puccinia-Coronata-Avenae on Avena-Sterilis Populations in Israel." Phytopathology 78(2): 135-139.://A1988M438200003

Bryant, J. and S. Leather (1992). "Indiscriminate Use of Selectable Markers - Sowing Wild Oats - Reply." Trends in Biotechnology 10(11): 382-382.://A1992JZ38100005

Bullen, E. R. (1967). "Wild Oats." Agriculture 74(2): 60-&.://A1967A052800002

Burdon, J. J. and W. J. Muller (1987). "Measuring the Cost of Resistance to Puccinia-Coronata Cda in Avena-Fatua L." Journal of Applied Ecology 24(1): 191-200.://A1987G970900015

Burdon, J. J., J. D. Oates, et al. (1983). "Interactions between Avena and Puccinia Species .1. the Wild Hosts - Avena-Barata Pott Ex Link Avena-Fatua L Avena-Ludoviciana Durieu." Journal of Applied Ecology 20(2): 571-584.://A1983RD43600015

Burke, J. (1979). "Wild Oats - Epstein,J." Harpers 259(1550): 78-79.://A1979HB27600019

Buschiazzo, D. E. and N. Peinemann (1979). "Competition for Nutrients between Wheat and Weeds (Avena-Fatua and Chenopodium-Album)." Agrochimica 23(3-4): 208-214.://A1979HZ01000005

Cabrera, M., C. Martinez-Cocera, et al. (2002). " paniceum (wild oats) pollen counts and aeroallergens in the ambient air of Madrid, Spain." International Archives of Allergy and Immunology 128(2): 123-129.://000176569500006 Background. Madrid has a short but intensive grass pollen season, in which 79% of the total grass pollen load is released from the middle of May to the middle of June. The main objectives of this study were to quantify Trisetum paniceum (wild oats) aeroallergen in the atmosphere in Madrid from February to December 1996 and to correlate the aeroallergen concentrations with grass pollen counts. Methods: Two different samplers were used to assess allergen exposure; a Burkard spore trap was used to collect pollen grains and a high-volume air sampler to collect airborne particles. A total of 182 air filters were collected and extracted in 1 ml of PBS and analysed by ELISA inhibition, using pooled sera from highly allergic individuals. Results: T. paniceum aeroallergens were detected not only during the grass pollen season, but also before and after. Wild oat allergens had two main peaks of 1 and 1.9 mug/m(3), occurring in late May and July, respectively. The time series analysis established the existence of lags between the two main variables pollen counts and aeroallergen activity. Analysis of the data by the Spearman rank test and linear regression showed a weak correlation between grass allergenic activity and grass pollen counts (Spearman's rho = 0.29). Data obtained from time series analysis demonstrated that grass allergenic activity correlated strongly with current and 5-week-old grass pollen grain counts (r(2) = 0.73). Conclusions: Wild oats allergenic activity was detected during the entire year and not only during the pollen season. This fact is an important aspect to be considered in the clinical follow-up and treatment of grass pollen- sensitised patients in Madrid. Copyright (C) 2002 S. Karger AG, Basel.

Cairns, A. L. P. and O. T. Devilliers (1980). "Effect of Aluminum Phosphide Fumigation on the Dormancy and Viability of Avena- Fatua Seed." South African Journal of Science 76(7): 323-323.://A1980KD21000015

Cairns, A. L. P. and O. T. Devilliers (1986). "Breaking Dormancy of Avena-Fatua L Seed by Treatment with Ammonia." Weed Research 26(3): 191-197.://A1986C444100007

Cairns, A. L. P. and O. T. Devilliers (1986). "Physiological-Basis of Dormancy-Breaking in Wild Oat (Avena-Fatua L) Seed by Ammonia." Weed Research 26(5): 365-374.://A1986E115000009

Cairns, A. L. P. and O. T. Devilliers (1989). "Effect of Sucrose Taken up by Developing Avena-Fatua L Panicles on the Dormancy and Alpha-Amylase Synthesis of the Seeds Produced." Weed Research 29(3): 151-156.://A1989AA75300001

Campbell, G. L., F. W. Sosulski, et al. (1987). "Nutritive-Value of Irradiated and Beta-Glucanase-Treated Wild Oat Groats (Avena- Fatua L) for Broiler-Chickens." Animal Feed Science and Technology 16(4): 243-252.://A1987G387400001

Cardon, Z. G. and R. B. Jckson (1995). "Root Acid-Phosphatase-Activity in -Hordeaceus and Avena-Barbata Remains Unchanged under Elevated [Co2]." Plant Physiology 108(2): 148-148.://A1995RE28900812

Carlson, H. L. and J. E. Hill (1985). "Wild Oat (Avena-Fatua) Competition with Spring Wheat - Plant-Density Effects." Weed Science 33(2): 176-181.://A1985ADN5900007

9

Carlson, H. L. and J. E. Hill (1986). "Wild Oat (Avena-Fatua) Competition with Spring Wheat - Effects of Nitrogen-Fertilization." Weed Science 34(1): 29-33.://A1986AYK3500007

Carsten, L. D., M. R. Johnston, et al. (2000). "A field trial of crown rust (Puccinia coronata f. sp avenae) as a biocontrol agent of wild oats on San Clemente Island." Biological Control 19(2): 175-181.://000089920000009 San Clemente Island has been subjected to large-scale disturbance from grazing and subsequent weed infestation, particularly by two wild oat species, Avena fatua and A. barbata. An endemic disease, crown rust (caused by Puccinia coronata f.sp. avenae), was selected as a potential control agent. A strain with a short latent period was selected from a common island isolate. Three replicate sets of plots were established on the island, each set containing a plot treated with crown rust, a control plot, and a plot treated with a fungicide. Disease was assessed in these plots and outside of the inoculated plots to a distance of 100 m. Reproductive output of wild oats from these plots was measured and compared for differences among treatments. Results indicated that disease was established from artificial inoculations in both years, 2 weeks earlier than the onset of disease from natural infections in 1997, but not in 1998. This result may have been due to unusual weather patterns in 1998 that caused an earlier onset of natural infections and precluded higher success of artificial inoculations early in the season. Analysis of reproductive output of wild oats from these plots showed that biomass and seed number were significantly lower (P < 0.05) in inoculated plots than in control and fungicide-treated plots. These results indicated that, with augmentation over several years, crown rust may be effective in reducing populations of wild oats on San Clemente Island. (C) 2000 Academic Press.

Carsten, L. D., B. Maxwell, et al. (2001). "Impact of crown rust (Puccinia coronata f. sp avenae) on competitive interactions between wild oats (Avena fatua) and (Nassella pulchra)." Biological Control 22(3): 207-218.://000172172400002 Crown rust of oats (Puccinia coronata f. sp. avenae) was investigated as a biocontrol agent for wild oats (Avena fatua) on San Clemente Island, . Successful restoration of the native habitats of this island will involve the reduction of wild oats and revegetation with native grasses such as stipa (Nassella pulchra). Determination of the outcome of interference between wild oats and stipa is important in the prediction of the success of the biocontrol agent. An addition series design was used to investigate these interactions, with densities of each species ranging from 0 to 2000 seeds per m(2). Eight replicates were established, four of which were randomly chosen and infected with the pathogen. After 75 days, plant height, dry weight, and seed production were measured. The data were fit to a hyperbola surface model by use of a nonlinear regression procedure. Results indicate that wild oat is the superior competitor at the seedling stage; however, early rust infection greatly reduces fitness of wild oats, causing an increased fitness for stipa. Simulations with a plant community model constructed from the results of the greenhouse experiment and information in the literature indicated that an equilibrium may be established between wild oat and stipa if high initial seeding rates of stipa are used in revegetation. (C) 2001 Academic Press.

Caseley, J. C. and D. Coupland (1980). "Effect of Simulated Rain on Retention, Distribution, Uptake, Movement and Activity of Difenzoquat Applied to Avena-Fatua." Annals of Applied Biology 96(1): 111-118.://A1980KL84200013

Cavan, G., P. Biss, et al. (1998). "Herbicide resistance and gene flow in wild-oats (Avena fatua and Avena sterilis ssp. ludoviciana)." Annals of Applied Biology 133(2): 207-217.://000078154400007 The process of resistance evolution to fenoxaprop-P-ethyl was investigated in the cereal weeds wild-oats (Avena fatua L. and Avena sterilis ssp. ludoviciana Malzew) at a number of locations in England, including one farm where distinct patches occur within fields. Genetic fingerprints produced using PCR-based techniques provided evidence for hybridisation between the species and that resistance had spread from one patch to others. The proportion of total variation due to differences between populations (G(st)) was estimated at 33-42%, and herbicide-resistant patches contained on average less genetic diversity than herbicide-sensitive counterparts: both findings were consistent with a high degree of self- pollination. It was however concluded that cross-pollination occurs both within and possibly between species, and that this can result in the spread of herbicide resistance.

Chancell.Rj and N. C. B. Peters (1974). "Time of Onset of Competition between Wild Oats (Avena-Fatua L) and Spring Cereals." Weed Research 14(3): 197-202.://A1974T338900008

Chang, F. Y., Stephens.Gr, et al. (1974). "Control of Wild Oats in Oats with Barban Plus Antidote." Weed Science 22(6): 546- 548.://A1974V038400003

Chang, P. R. and F. W. Sosulski (1985). "Functional-Properties of Dry Milled Fractions from Wild Oats (Avena-Fatua L)." Journal of Food Science 50(4): 1143-&.://A1985ALW1600067

Chao, J. F., A. I. Hsiao, et al. (1993). "Effects of Imazamethabenz on the Main Shoot Growth and Tillering of Wild Oat (Avena-Fatua L)." Journal of Plant Growth Regulation 12(3): 141-147.://A1993MK98000006 Foliar application of imazamethabenz at sublethal doses of 100 and 200 g a.i./ha to wild oat plants at the two-leaf stage without tillers greatly inhibited the growth of the main shoot but increased tillering. The near cessation of sheath and the main stem elongation indicated that the major effect of imazamethabenz on the main shoot was inhibition of intercalary growth. Low doses of imazamethabenz treatment resulted in more leaves (including leaf primordia) in the main stem but did not affect mature first and second leaves. Sublethal doses of imazamethabenz only briefly inhibited tiller growth. A later increase in tillering in treated plants resulted from the stimulated resumed growth of tillers and the increased initiation of tiller buds. Such enhanced tillering mainly resulted from the release of apical dominance due to the inhibition or cessation of the main stem growth with imazamethabenz treatment. Both doses of imazamethabenz (100 and 200 g a.i./ha) significantly reduced the biomass of shoots and roots, but increased the ratio of roots/shoots dry weight.

Chao, J. F., A. I. Hsiao, et al. (1997). "Tillers do not influence phytotoxicity of imazamethabenz in wild oat (Avena fatua)." Journal of Plant Growth Regulation 16(3): 173-179.://A1997XT77800009 10

The response of wild oat to imazamethabenz varies with the growth stage, but the role of tillers in this regard is unclear. Removal of tillers at the three-leaf stage before spraying with imazamethabenz did not significantly affect the total shoot fresh weight measured 3 weeks later. The leaf area and dry weight of intact plants at the three-leaf stage were 17-21% greater than for plants with coleoptilar and first leaf main shoot tillers (TO and TI) removed. The greater leaf area may have increased herbicide interception per plant. Similar fresh weight reductions in main shoot, total tillers, and total shoots were found whether imazamethabenz was applied to the plant at the two-leaf without tillers or the three-leaf with two tillers stage. Imazamethabenz applied only to the main shoot reduced total shoot dry weight more than an equivalent amount of imazamethabenz applied only to tiller TI or applied over the whole shoot. Imazamethabenz had the least inhibitory effect on whole plant growth when applied only to T1. When C-14-herbicide was applied to the first main shoot leaf of plants at the three-leaf stage with two tillers, the C-14 translocated 38% to roots, 33% to the main shoot, and nearly 30% to all tillers. When C-14-herbicide was applied to the first leaf of T1 then the C-14 translocated 50% to T1, 25% to the main shoot, 20% to roots, and 5% to all other tillers. The translocation pattern and fresh weight values suggested that the presence of early tillers during herbicide application neither increased nor decreased imazamethabenz efficacy in wild oat.

Chao, J. F., A. I. Hsiao, et al. (1994). "Effect of Decapitation on Absorption, Translocation, and Phytotoxicity of Imazamethabenz in Wild Oat (Avena-Fatua L)." Journal of Plant Growth Regulation 13(3): 153-158.://A1994PN67000007 The release of apical dominance by the physical destruction in situ of the apical meristem and associated leaf primordia (decapitation) promoted the growth of tillers in non-herbicide-treated wild oat plants, as indicated by increased tiller lengths and fresh weights. At 96 h after [C-14] herbicide treatment following decapitation, the absorption of [C-14]imazamethabenz and total translocation of radioactivity were respectively increased by 28% and 49%. By 96 h after [C-14]imazamethabenz application, the radioactivity detected in the roots of decapitated plants was 45% higher than that in the roots of nondecapitated plants while the radioactivity in tillers of decapitated plants was 2.6-fold that in tillers of intact plants. Decapitation together with foliar spraying of imazamethabenz at 200 g ha-1 further reduced tiller fresh weight, greatly decreased the total tiller number, and thereafter significantly increased overall phytotoxicity by 32% as measured by total shoot fresh weight. The results of this study support the hypothesis that main shoot apical dominance limits translocation of applied imazamethabenz to lateral shoots, rendering tillers less susceptible to growth inhibition by the herbicide.

Chao, J. F., W. A. Quick, et al. (1994). "Effect of Imazamethabenz on Histology and Histochemistry of Polysaccharides in the Main Shoot of Wild Oat (Avena-Fatua)." Weed Science 42(3): 345-352.://A1994PD56000004 Imazamethabenz, an acetohydroxyacid synthase (AHAS)-inhibiting herbicide, was applied at a sublethal rate (2.00 g al ha- 1) to foliage of greenhouse-grown wild oat plants at the 2-leaf stage without tillers. Main shoot tissues were sampled 1 to 2 wk after imazamethabenz treatment and prepared for light microscopic examination. Histological observations showed that imazamethabenz treatment did not produce drastic changes on leaf structures.but did affect chloroplast integrity of fully expanded 2nd leaf and inhibited differentiation of the young 3rd leaf 1 wk after application. The internode length and the number of cells in internodes of the herbicide-treated main stem were greatly reduced, presumably due to the inhibition by imazamethabenz of intercalary meristem cell division in early stages of internode elongation. One week after imazamethabenz application, histochemical studies revealed an increased accumulation of starch granules in chloroplasts of the 2nd leaf but reduced starch levels in the main stem. This confirmed the hypothesis that AHAS inhibitors affected phloem transport of photosynthates.

Chao, J. F., W. A. Quick, et al. (1994). "Influence of Nutrient Supply and Plant-Growth Regulators on Phytotoxicity of Imazamethabenz in Wild Oat (Avena-Fatua L)." Journal of Plant Growth Regulation 13(4): 195-201.://A1994PY13900005 The influences of nutrient supply and plant growth regulators on the phytotoxicity of imazamethabenz in wild oat (Avena fatua L.) were evaluated in the greenhouse. Wild oat plants supplied with half-strength rather than one-eighth-strength Hoagland solution were more susceptible to imazamethabenz, showing greater growth reduction in main shoot and tillers. The improved herbicide efficacy at higher nutrient levels appeared related to increased herbicide interception by the greater leaf surface available. Leaves develop at either nutrient level did not differ significantly in epicuticular wax, so differential absorption appeared unlikely. Wild oat plants supplemented with nutrient, switching from low to high levels at the time of herbicide application, were as susceptible to imazamethabenz or even more so than plants growing with a constant high level of nutrition. The wild oat pure-line Montana 73, a strongly tillering line, was more susceptible to imazamethabenz than the limited-tillering line, Crop Science 40. Both 2,4-D and GA3 reduced imazamethabenz-induced tillering. Imazamethabenz efficacy was increased by GA3 but not by 2,4-D. These results support the hypothesis that lowering apical dominance of wild oat increases imazamethabenz activity in tillers, and that increased tillering following sublethal doses of imazamethabenz treatment is associated with the release of apical dominance.

Chen, F. S., J. M. Mactaggart, et al. (1982). "Chemical-Constituents in Wild Oat (Avena-Fatua) Hulls and Their Effects on Seed- Germination." Canadian Journal of Plant Science 62(1): 155-161.://A1982NC69900023

Chen, S. C. and R. M. Elofson (1978). "Phytotoxicity of Organic and Inorganic Iodides to Avena-Fatua." Journal of Agricultural and Food Chemistry 26(1): 287-289.://A1978EH15800076

Chen, S. S. C. and W. M. Park (1973). "Early Actions of Gibberellic-Acid on Embryo and on Endosperm of Avena-Fatua Seeds." Plant Physiology 52(2): 174-176.://A1973Q598400018

Chen, S. S. C. and J. E. Varner (1969). "Metabolism of 14c-Maltose in Avena Fatua Seeds During Germination." Plant Physiology 44(5): 770-&.://A1969D457600023

Chen, S. S. C. and J. E. Varner (1970). "Respiration and Protein Synthesis in Dormant and after-Ripened Seeds of Avena-Fatua." Plant Physiology 46(1): 108-&.://A1970H431400020

Chiko, A. W. (1975). "Natural Occurrence of Barley Stripe Mosaic-Virus in Wild Oats (Avena-Fatua)." Canadian Journal of Botany- Revue Canadienne De Botanique 53(4): 417-420.://A1975V999200010 11

Chiko, A. W. (1983). "Reciprocal Contact Transmission of Barley Stripe Mosaic-Virus between Wild Oats and Barley." Plant Disease 67(2): 207-208.://A1983PZ44400032

Chiko, A. W. (1984). "Increased Virulence of Barley Stripe Mosaic-Virus for Wild Oats - Evidence of Strain Selection by Host Passage." Phytopathology 74(5): 595-599.://A1984ST28500022

Chow, P. N. P. (1976). "Effects of Postemergence Herbicides on Growth, Photosynthesis, and Photosynthate Translocation in Wild Oats." Canadian Journal of Plant Science 56(2): 429-430.://A1976BV47000055

Chow, P. N. P. (1982). "Wild Oat (Avena-Fatua) Herbicide Studies .1. Physiological-Response of Wild Oat to 5 Postemergence Herbicides." Weed Science 30(1): 1-6.://A1982NB52400001

Chow, P. N. P. (1986). "Sequential Application of Soil-Incorporated and Postemergence Herbicides for Controlling Wild Oat (Avena- Fatua L) and Green Foxtail (Setaria-Viridis (L) Beauv) in Spring Wheat." Crop Protection 5(3): 209-213.://A1986C748500009

Chow, P. N. P. (1988). "Wild-Oat Herbicide Studies - Barban Herbicidal Activity for Wild-Oat (Avena-Fatua L) Control as Influenced by Adjuvants." Crop Protection 7(1): 3-8.://A1988L758900001

Chow, P. N. P. and D. G. Dorrell (1979). "Response of Wild Oat (Avena-Fatua), Flax (Linum-Usitatissimum), and Rapeseed (Brassica-Campestris and B Napus) to Diclofop-Methyl." Weed Science 27(2): 212-215.://A1979GR79500016

Chow, P. N. P. and R. D. Dryden (1975). "Control of Wild Oats in Wheat with Benzoylprop-Ethyl." Canadian Journal of Plant Science 55(2): 379-383.://A1975AC35900004

Chow, P. N. P. and D. E. Laberge (1978). "Wild Oat Herbicide Studies .2. Physiological and Chemical Changes in Barley and Wild Oats Treated with Diclofop-Methyl Herbicide in Relation to Plant Tolerance." Journal of Agricultural and Food Chemistry 26(5): 1134- 1137.://A1978FP38900033

Ciastoch, K., J. Klos, et al. (2004). "Karyotype structure of hexaploid wild oat (Avena fatua)." Zeitschrift Fur Pflanzenkrankheiten Und Pflanzenschutz-Journal of Plant Diseases and Protection: 191-196.://000225651400024 Allohexaploid wild oat (Avena fatua L., : 2n = 6x = 42) is one of the world's worst weeds. In contrast to other Avena hexaploids its genomic structure has not been well documented. The C-banded karyotype of A. fatua described here in this study is different from those reported in previous studies. It was not possible to recognise C-genome or any other ancestral genomes due to differences in staining intensity.

Clamot, G. and R. Rivoal (1984). "Genetic-Resistance to Cereal Cyst Nematode Heterodera-Avenae Woll in Wild Oat Avena-Sterilis I-376." Euphytica 33(1): 27-32.://A1984AES4500006

Clegg, M. T. and R. W. Allard (1972). "Patterns of Genetic Differentiation in Slender Wild Oat Species Avena-Barbata." Proceedings of the National Academy of Sciences of the United States of America 69(7): 1820-&.://A1972N026000040

Clegg, M. T. and R. W. Allard (1973). "Genetics of Electrophoretic Variants in Avena - .2. Esterase E1, E2, E4, E5, E6 and Anodal Peroxidase Apx4 Loci in a-Fatua." Journal of Heredity 64(1): 3-6.://A1973P402600001

Clegg, M. T. and R. W. Allard (1973). "Viability Versus Fecundity Selection in Slender Wild Oat, Avena Barbata L." Science 181(4100): 667-668.://A1973Q361800024

Cluster, P. D. and R. W. Allard (1995). "Evolution of Ribosomal DNA (Rdna) Genetic-Structure in Colonial Californian Populations of Avena-Barbata." Genetics 139(2): 941-954.://A1995QE14500032 DNA samples from 980 plants of Avena barbata from 48 ecologically diverse sites in California and were assayed to determine their genotype for two duplicated loci governing rDNA variants. More than 40 different rDNA genotypes were observed among which 5 made up 96% of our sample in environmentally homogeneous sites; predominant genotypes were less frequent and recombinant genotypes were more frequent in environmentally heterogeneous sites. The spatial distribution of each predominant rDNA genotype was nearly an exact overlay on both macro- and microgeographical scales of a distinctive habitat and also of the distribution of an eight-locus morphological-allozyme variant genotype. In all, seven different habitat-genotype combinations () were distinguishable on the basis of their morphological- allozyme-rDNA genotypes. None of these seven genotypes has been found in ancestral Spanish populations; thus the above predominant multilocus genotypes (ecotypes) of the colonial populations evidently evolved subsequent to the recent introduction (within 150-200 generations) of A. barbata to California. The precise associations of specific alleles and genotypes of the morphological allozyme and rDNA loci with different specifiable habitats leads us to the conclusion that natural selection favoring particular multilocus combinations of alleles in different habitats was the main guiding force in shaping the internal genetic structure of local populations as well as the overall adaptive landscape of A. barbata over California and Oregon.

Cobb, R. D. and L. G. Jones (1962). "Germinating Dormant Seeds of Avena Fatua (Wild Oat)." American Journal of Botany 49(6): 658-&.://A19622650A00032

Comeau, A. (1981). "Tolerance to Barley Yellow Dwarf in Avena-Sterilis." Phytopathology 71(2): 210-210.://A1981LN69500120

12

Conn, J. S. (1986). "An Evaluation of Herbicides for Control of Wild Oats in Barley - Efficacy, Phytotoxicity, and Barley Variety Susceptibility Studies." Alaska Agricultural and Forestry Experiment Station Bulletin(70): 1-19.://A1986A855800001

Coupland, D., W. A. Taylor, et al. (1978). "Effect of Site of Application on Performance of Glyphosate on Agrypyron-Repens and Barban, Benzoylprop-Ethyl and Difenzoquat on Avena-Fatua." Weed Research 18(3): 123-&.://A1978FK07700001

Courtney, A. D. (1982). "Herbicides for the Control of Wild Oat (Avena-Fatua)." Agriculture in Northern Ireland 57(1): 11-13.://A1982NR51300003

Cousens, R., C. J. Doyle, et al. (1986). "Modeling the Economics of Controlling Avena-Fatua in Winter-Wheat." Pesticide Science 17(1): 1-12.://A1986A262600001

Cousens, R., S. E. Weaver, et al. (1992). "Growth and Development of Avena-Fatua (Wild-Oat) in the Field." Annals of Applied Biology 120(2): 339-351.://A1992JF04400015 The growth and development of field-grown A vena fatua plants were studied for autumn and spring sowings in two consecutive years. The duration of various growth stages from sowing until anthesis was quantified in terms of thermal time (accumulated degree days) or photothermal time (degree days modified by photoperiod). Base temperatures and photoperiods for developmental phases were estimated as those which minimised the coefficient of variation among sowing dates. Relationships were derived between leaf emergence, canopy height, plant leaf area, and photothermal time. Stem extension and flowering occurred earlier in autumn-sown plants than spring-sown plants. Autumn-sown plants produced more leaves on the main stem, and had greater leaf area and above-ground biomass at anthesis than spring- sown plants.

Cousens, R. D., M. P. Johnson, et al. (1992). "Comparative Rates of Emergence and Leaf Appearance in Wild Oats (Avena- Fatua), Winter Barley (Hordeum-Sativum) and Winter-Wheat (Triticum-Aestivum)." Journal of Agricultural Science 118: 149-156.://A1992HP65900002 Winter barley cv. Igri, winter wheat cv. Avalon and spring wild oats (Avena fatua) were studied in monocultures in pots and in the field. The field experiments were located near Bristol and Bury St Edmunds in 1988/89. Pot sowings were monthly, whereas there was a single sowing date for each of the two field locations. Base temperatures for emergence in pots were 2.2, 1.3 and 2.3-degrees-C for barley, wheat and A. fatua respectively. Barley was consistently the fastest to emerge. Leaf number was strongly correlated with photothermal time from emergence, with barley producing leaves at the greatest rate. Base temperatures for leaf appearance were -6, -5 and -3-degrees-C for barley, wheat and A. fatua respectively. The field studies confirmed the ranking of the species based on the pot experiments. Both a model based on photothermal time and one based on rate of change of daylength at emergence gave good descriptions of the data. It is argued that correlations of rates of development with individual environmental variables are not sufficient to deduce the underlying mechanisms.

Cousens, R. D., S. E. Weaver, et al. (1991). "Dynamics of Competition between Wild Oats (Avena-Fatua L) and Winter Cereals." Weed Research 31(4): 203-210.://A1991FZ80000005 This study examined the effects of competition on the growth of Avena fatua, winter wheat and winter barley. Plants were sampled at frequent intervals from replacement series experiments at two contrasting sites in the U.K. A. fatua was much slower to establish than the two cereals, but thereafter exhibited a faster rate of growth. In monoculture, it took a considerable time for A. fatua to reach a size equal to that of the cereals, but by the end of the experiments it was the largest of the three species. The change-over from cereal dominance in mixtures to A. fatua dominance was rapid, and in three cases coincided with cereal flag leaf emergence. In the fourth case, it appeared to coincide with the start of canopy height extension. At one site the order of competitiveness at anthesis was A. fatua = barley > wheat, and at the other site the order was A. fatua > wheat > barley. In order to explain and predict differences between years and sites, more studies are required on morphological development in relation to abiotic variables.

Cowan, A. K., S. L. Turner, et al. (1995). "Effect of Water-Stress and Diclofop-Methyl on Photosynthesis, Carotenoid and Abscisic- Acid Content of Leaves of Avena-Byzantina and Avena-Fatua." South African Journal of Botany-Suid-Afrikaanse Tydskrif Vir Plantkunde 61(1): 29-34.://A1995QH89700005 The effect of combined water stress and diclofop-methyl treatment on photosynthesis and carotenoid and abscisic acid (ABA) content of leaves of A. byzantina and A. fatua was investigated. Sublethal doses of diclofop-methyl caused a transient decline in net assimilation rate and a decrease in beta-carotene and zeaxanthin in leaves of both species. The decline in carotenoid levels occurred concomitant with a substantial but transient increase in ABA. A similar but less dramatic trend was observed for water-stressed plants. Recovery of photosynthesis in seedlings exposed simultaneously to diclofop-methyl and water stress, was associated with an increase in beta-carotene and zeaxanthin content and a return to basal ABA levels in leaaves of A. byzantina. By comparison, substantial accumulation of zeaxanthin was observed in leaves of A. fatua following combined water stress and herbicide treatment, apparently at the expense of ABA. Similar findings were made regarding levels of zeaxanthin when diclofop-methyl was applied to already water-stressed plants of A. byzantina and A. fatua. It is proposed that herbicide- and/or water-stress-induced alternations in acetyl-coenzyme A carboxylase activity coupled with reduced demand for fatty acid synthesis, facilitate channelling of photosynthetically fixed carbon into isoprenoids and that alterations in the capacity for terpenoid synthesis forms part of the mechanism by which drought stress antagonizes the activity of aryloxyphenoxypropionic acid herbicides.

Cox, D. J. and K. J. Frey (1984). "Improving Cultivated Oats (Avena-Sativa L) with Alleles for Vegetative Growth Index from Avena- Sterilis L." Theoretical and Applied Genetics 68(3): 239-245.://A1984SY75800006

Cox, T. S. and K. J. Frey (1985). "Complementarity of Genes for High Groat-Protein Percentage from Avena-Sativa L and Avena- Sterilis L." Crop Science 25(1): 106-109.://A1985AAX8500027

13

Craig, I. L., B. E. Murray, et al. (1974). "Avena-Canariensis - Morphological and Electrophoretic Polymorphism and Relationship to a Avena-Magna Avena-Murphyi Complex and a Avena-Sterilis." Canadian Journal of Genetics and Cytology 16(3): 677-689.://A1974V398800021

Cranston, H. J., R. R. Johnson, et al. (1999). "Isolation and characterization of a cDNA encoding a sar-like monomeric GTP- binding protein in Avena fatua L." Plant Science 145(2): 75-81.://000082023000003 Differential display of mRNAs from embryos of Avena fatua L. caryopses was used to isolate an mRNA more abundant in nondormant than dormant caryopses during early imbibition. The DNA sequence of the corresponding 579 bp cDNA, termed Af SAR1, is 92% identical to an Arabidopsis cDNA encoding the monomeric GTP-binding protein sar1p. Predicted amino acid sequences of the four conserved CTP binding and hydrolysis domains in Af SAR1 are 100% identical to sar1p. Af SAR1 mRNA levels increased 6-fold or more than 10-fold in nondormant embryos during the first 48 h of imbibition in water or GA(3), respectively. However, mRNA levels increased only slightly and transiently in dormant embryos imbibed in water, mRNA abundance was highest in meristematic and actively growing tissues of A. fatua seedlings. Af SAR1 belongs to a small (two to four members) gene family as judged by Southern hybridizations. Increased abundance of this mRNA during early germination and in actively growing tissues indicates that the respective protein is associated with rapid cell elongation, cell division and growth. (C) 1999 Elsevier Science Ireland Ltd. All rights reserved.

Cranston, H. J., A. J. Kern, et al. (1996). "Wound-induced ethylene and germination of embryos excised from dormant Avena fatua L caryopses." International Journal of Plant Sciences 157(2): 153-158.://A1996UC34000001 Intact caryopses of dormant AN265 and M73 wild oat (Avena fatua L.) lines did not germinate when imbibed in water at 14 degrees C, but germinated after exposure to ethylene during imbibition. Embryos excised from dormant caryopses and imbibed in water germinated readily. However, inhibitors of ethylene synthesis ([aminooxy]acetic acid and 2- [aminoethoxyvinyl] glycine) and ethylene action (2,5-norbornadiene and silver thiosulfate) delayed or almost completely inhibited germination of excised embryos. Embryos removed from inhibitor treatments after 15 d and incubated in water germinated normally. Except for 2,5-norbornadiene, inhibitors did nor reduce germination of intact nondormant (afterripened) caryopses or embryos excised from nondormant caryopses. Reduced germination rates of embryos excised from dormant caryopses and incubated in 2,5-norbornadiene and 2-(aminoethoxyvinyl) glycine were reversed by applications of 0.05 mu L/L ethylene. The results indicate that wound-induced ethylene synthesis may be responsible for germination of embryos excised from dormant wild oat caryopses.

Cremers, H. C. (1993). "Transgenic Bull to Sow Wild Oats." New Scientist 137(1855): 8-8.://A1993KG56600012

Crowley, J., J. T. Odonovan, et al. (1978). "Phytotoxicity of Soil-Applied Dichlorfop Methyl and Its Effect on Uptake of Ca-45 in Wild Oats, Barley and Wheat." Canadian Journal of Plant Science 58(2): 395-399.://A1978FA30300014

Cudney, D. W., L. S. Jordan, et al. (1989). "Developmental Rates of Wild Oats (Avena-Fatua) and Wheat (Triticum-Aestivum)." Weed Science 37(4): 521-524.://A1989AR21200006

Cudney, D. W., L. S. Jordan, et al. (1991). "Effect of Wild Oat (Avena-Fatua) Infestations on Light Interception and Growth-Rate of Wheat (Triticum-Aestivum)." Weed Science 39(2): 175-179.://A1991FY47400008 Wild oat reduced light penetration and growth of dwarf hard red spring wheat in field experiments performed under nonlimiting nitrogen and moisture conditions. Wild oat grew taller than wheat and had a greater portion of its canopy above 60 cm at maturity. Light penetration in a mixed canopy was similar to that in a monoculture wheat canopy when wild oat was clipped to the height of the wheat. A mathematical model was developed which accurately predicted the reduction in the growth rate of wheat from wild oat interference. The model also predicted that interference from wild oat was due to reduced leaf area of wheat at early growth stages and low wild oat densities, and reduced light penetration to wheat leaves at later growth stages and higher densities of wild oat.

Cudney, D. W., L. S. Jordan, et al. (1989). "Competitive Interactions of Wheat (Triticum-Aestivum) and Wild Oats (Avena-Fatua) Grown at Different Densities." Weed Science 37(4): 538-543.://A1989AR21200009

Cuming, A. C. and D. J. Osborne (1978). "Membrane Turnover in Imbibed and Dormant Embryos of Wild Oat (Avena-Fatua L) .1. Protein Turnover and Membrane Replacement." Planta 139(3): 209-217.://A1978EV20000004

Cuming, A. C. and D. J. Osborne (1978). "Membrane Turnover in Imbibed Dormant Embryos of Wild Oat (Avena-Fatua L) .2. Phospholipid Turnover and Membrane Replacement." Planta 139(3): 219-226.://A1978EV20000005

Cumming, B. G. and J. R. Hay (1958). "Light and Dormancy in Wild Oats (Avena-Fatua L)." Nature 182(4635): 609-610.://A1958ZQ49700057

Curran, W. S., L. A. Morrow, et al. (1987). "Lentil (Lens, Culinaris) Yield as Influenced by Duration of Wild Oat (Avena, Fatua) Interference." Weed Science 35(5): 669-672.://A1987K223500013

Curtis, A. (1977). "'Wild Oats'." Drama(124): 41-43.://A1977EP39500004

Darmency, H. (1982). "A Study of Polymorphism of Prolamines in a Population of Avena-Fatua L." Weed Research 22(4): 237- 243.://A1982NZ77900007

Darmency, H. and C. Aujas (1986). "Polymorphism for Vernalization Requirement in a Population of Avena-Fatua." Canadian Journal of Botany-Revue Canadienne De Botanique 64(4): 730-733.://A1986C437000005

14

Darmency, H. and C. Aujas (1987). "Character Inheritance and Polymorphism in a Wild Oat (Avena-Fatua) Population." Canadian Journal of Botany-Revue Canadienne De Botanique 65(11): 2352-2356.://A1987K962800020

Darmency, H. and C. Aujas (1992). "Genetic Diversity for Competitive and Reproductive Ability in Wild Oats (Avena-Fatua)." Weed Science 40(2): 215-219.://A1992JC08300010 Three wild oats were grown in wheat stands sown at different dates in greenhouse and field trials. Wild oats growth and seed output, and their effects on wheat biomass were not different among phenotypes when wild oats emerged 2 wk after the wheat. In experiments in which wild oats were planted in germinated wheat, one was shorter, weighed less, and produced fewer seed than the other phenotypes. Another phenotype reduced wheat biomass more than the other phenotypes. Vernalization increased vegetative growth and reduced spikelet production of one phenotype, but had no effect on its competitiveness with wheat.

Darwent, A. L. (1980). "Effects of Soil-Temperature on the Phytotoxicity of Trifluralin to Wild Oats." Canadian Journal of Plant Science 60(3): 929-938.://A1980KJ42600022

Daugovish, O., D. C. Thill, et al. (2002). "Competition between wild oat (Avena fatua) and yellow mustard (Sinapis alba) or canola (Brassica napus)." Weed Science 50(5): 587-594.://000178083600007 Wild oat, a troublesome weed in small grain cereals, infests about 11 million ha of cropland in the United States Diversifying cereal production with alternative crops, such as yellow mustard and canola, provides flexible cropping systems, decreases production risks, and may allow more effective weed suppression A greenhouse study was conducted to assess the competitive ability of yellow mustard and canola with wild oat in 1999 and 2000, using replacement series interference experiments to relate the results to plant development stages Yellow mustard, regardless of its proportion in mixture, reduced aboveground biomass of wild oat 33 to 66%, leaf and tiller number 34 to 36%, and panicle production 58% compared with wild oat in monoculture Canola did not affect wild oat biomass in mixtures Yellow mustard per plant biomass in 2000 and inflorescence production in 1999 decreased 30 and 20% with increased density of yellow mustard in mixtures Yellow mustard biomass was not affected by the addition of wild oat to the mixture, indicating the greater importance of intraspecific competition between yellow mustard relative to interspecific competition with wild oat Canola per plant biomass was affected more by interspecific competition with wild oat than by intraspecific competition A second greenhouse experiment was conducted to compare plant height and biomass accumulation by the three species over 7 wk Yellow mustard had the greatest biomass accumulation and plant elongation rate, followed by canola and wild oat The greater competitive ability of yellow mustard with wild oat, compared with canola, is likely associated with the rapid growth and canopy elevation of yellow mustard.

Daugovish, O., D. C. Thill, et al. (2003). "Modeling competition between wild oat (Avena fatua L.) and yellow mustard or canola." Weed Science 51(1): 102-109.://000180489500016 Wild oat, a troublesome weed in cereals, infests about 11 million hectares of cropland in the United States. Diversifying cereal production with alternative crops, such as yellow mustard and canola, can provide flexibility in cropping systems, decrease production risks, and allow for effective weed suppression. The objective of the study was to quantify the competitive ability of yellow mustard and canola relative to wild oat in addition series field experiments, which were conducted in 1999 and 2000. near Genesee, ID. Biomass and seed production of wild oat were reduced 67 and 80%, respectively, in mixtures with yellow mustard, which was three to four times greater than the reduction in corresponding mixtures with canola. In addition, yellow, mustard reduced the biomass and seed production of wild oat equally regardless of wild oat density. In contrast, the competitive effect of canola on wild oat biomass decreased 5 to 10 times when wild oat density increased from 100 to 200 plants Yellow mustard at all densities and at both biomass harvests suppressed wild oat biomass and seed production similarly. But suppression of wild oat by canola increased as canola density increased, and canola plants were more competitive at the flowering stage than at the rosette stage. Wild oat had little or no effect on yellow mustard seed yield but reduced canola seed yield 37%, when averaged over canola densities. Additionally, the oil content of canola seed was reduced 0.4% for every 1% of wild oat seed in the harvested seed. Models developed in this study accurately predicted plant populations of yellow mustard and canola that provided optimal weed suppression and crop yield for different wild oat populations.

Day, A. D. and R. K. Thompson (1966). "Inheritance of Awns in a Cross between C Koch and Avena Fatua L." Crop Science 6(6): 608-&.://A19668741700034

DelArco, M. J. S., C. Torner, et al. (1995). "Seed dynamics in populations of Avena sterilis ssp ludoviciana." Weed Research 35(6): 477-487.://A1995TR75100006 Seed populations of Avena sterilis ssp. ludoviciana (Durieu) Nyman were monitored in a naturally occurring infestation throughout its life cycle. Considering the large weed population present (298 panicles m(-2)), total seed production was relatively low: 3838 seeds m(-2). Only 68% of these seeds were recovered from the soil surface and a further 3% were removed with wheat grain and straw during harvest operations. The numbers of seeds from the stubble between mid-July and mid-September were relatively low (10%). Ploughing the stubble in October buried most of the recently produced seed rain and resulted in a relatively uniform vertical distribution of the seedbank. Maximum seed persistence in the soil ranged from 27 to 43 months (depending on the experimental technique used to do the study). Seed decline followed an exponential pattern on a yearly basis, with the greatest decline taking place between October and April (57-90% in year 1 and 10-40% in year 2). Between May and September the buried seed populations remained practically constant. Seedbank depletion was primarily due to seedling production (25%) and 'lethal' germination (24%). Although the depth of burial had very little effect on seed survival, the mode of seed disappearance was closely related to their depth in the soil. Seed depletion through 'lethal' germination increased with increasing depth in the soil, whereas depletion through seedling emergence decreased with increasing depth.

15

Delavega, M. P., P. Garcia, et al. (1991). "Multilocus Genetic-Structure of Ancestral Spanish and Colonial Californian Populations of Avena-Barbata." Proceedings of the National Academy of Sciences of the United States of America 88(4): 1202-1206.://A1991EY61700025 We have applied a multivariate log-linear technique to the analysis of interlocus allelic associations among 14 allozyme loci in a sample of 4011 plants from 42 Spanish populations of Avena barbata. The loci fell into three natural groups of five, five, and four loci. The five loci of the first group are invariant, or nearly so, throughout the range of the species. The genetic organization of the loci of this set is defined by a single five-locus genotype; each allele of this predominant genotype is a "wild-type" allele that contributes favorably to adaptedness in all single-locus and multilocus configurations regardless of environment. Although allelic diversity is high in Spain for the nine loci of the second and third sets, log-linear analyses showed that these loci are tied together in Spanish populations through complex networks of overlapping lower- order interlocus interactions. The ancestral Spanish and colonial Californian gene pools are closely similar in allelic composition on a locus-by-locus basis; however, Spanish allelic configurations at two-locus and higher-order levels are usually different from and much less tightly organized than in Californian populations. We conclude that the major force involved in the evolution of the colonial populations was selection that led to reorganization, at the interlocus level, of the ancestral Spanish allelic ingredients into different multilocus genotypes adapted to Californian habitats.

Devine, M. D., S. A. Macisaac, et al. (1992). "Investigation of the Mechanism of Diclofop Resistance in 2 Biotypes of Avena-Fatua." Pesticide Biochemistry and Physiology 42(1): 88-96.://A1992GV14000010

Devine, M. D. and A. Rashid (1993). "Antagonism of Tralkoxydim Activity in Avena-Fatua by Metsulfuron Methyl." Weed Research 33(2): 97-104.://A1993KU68000001 In growth chamber experiments, tralkoxydim activity against Avena fatua was significantly reduced by addition of metsulfuron methyl to the spray solution. This was apparent particularly at low rates of tralkoxydim (e.g., 125 g a.i. ha-1) and at high rates of metsulfuron methyl (8 or 12 g a.i. ha-1). Metsulfuron methyl alone reduced A. fatua growth significantly. Chlorsulfuron, which did not reduce A. fatua growth at rates of up to 33 g a.i. ha-1, caused only a slight loss in tralkoxydim activity. When applied in combination, metsulfuron methyl reduced the uptake of foliar-applied C-14- tralkoxydim 6 h after application, but did not affect uptake or translocation of C-14-tralkoxydim at subsequent sampling times. In in vitro assays, metsulfuron methyl at 10 or 100 nm in the assay medium had no effect on acetyl-coenzyme A carboxylase (ACCase) activity in A. fatua, nor did it interfere with the inhibition of ACCase by tralkoxydim. In addition, treatment of A. fatua plants with metsulfuron methyl at 4, 8 or 12 g ha-1 did not reduce the level of extractable ACCase activity when the plants were harvested 24-96 h after spraying. Thus, no apparent physiological or biochemical basis for the antagonism of tralkoxydim was identified, The temporary growth inhibition induced by metsulfuron methyl may be of sufficient duration for most of the tralkoxydim to be metabolized to inactive products, thereby reducing tralkoxydim-induced injury.

Dew, D. A. (1981). "Effect of Time of Removal of Wild Oats on Yield of Barley and Rape." Canadian Journal of Plant Science 61(2): 492-492.://A1981LT55500054

Dew, D. A. and C. H. Keys (1976). "Index of Competition for Estimating Loss of Rape Due to Wild Oats." Canadian Journal of Plant Science 56(4): 1005-1006.://A1976CP13300045

Dexter, J. E., K. R. Preston, et al. (1984). "The Effect of Various Levels of Barley, Wild Oats and Domestic Oats on the Milling and Baking Performance of Hard Red Spring Wheat." Canadian Journal of Plant Science 64(2): 275-283.://A1984SP47900004

Dhima, K. V., I. G. Eleftherohorinos, et al. (2000). "Interference between Avena sterilis, Phalaris minor and five barley cultivars." Weed Research 40(6): 549-559.://000166524900006 Field experiments were carried out in Northern Greece from 1994 to 1997 to study interference between Avena sterilis L. or Phalaris minor Retz. and five autumn-sown barley cultivars. Weed:crop interference began in early April. Avena sterilis at 120 plants m showed greater interference against barley than P. minor at 400 plants m(-2). The greatest grain yield and ear number reduction due to interference by either weed was recorded for cvs Klipper and Plaisant. and the least for cv. Athinaida; with cvs Carina and Thermi intermediate. Yield reduction due to A. sterilis for cvs Athinaida, Carina, Thermi, Klipper and Plaisant was 8, 16, 27, 61 and 67%, respectively, while corresponding losses to P. minor were 1, 8, 14, 45 and 55%. These results clearly indicate that growth and consequently yield components of cv. Athinaida were unaffected by the presence of either weed species, while those of cv. Carina were affected by A. sterilis. but not by P. minor. However, dry weight and panicle number of both weed species were severely reduced by interference with cvs Carina, Athinaida and Thermi compared with cvs Klipper and Plaisant. The order of interference of the five barley cultivars tested against A. sterilis and P. minor was Athinaida > Carina > Thermi > Klipper > Plaisant.

Dinoor, A. (1970). "Sources of Oat Crown Rust Resistance in Hexaploid and Tetraploid Wild Oats in Israel." Canadian Journal of Botany 48(1): 153-&.://A1970F432900021

Dunan, C. M., F. D. Moore, et al. (1994). "A Plant Process-Economic Model for Wild Oats Management Decisions in Irrigated Barley." Agricultural Systems 45(4): 355-368.://A1994NQ46000001 Development of integrated weed management systems requires the integration of different weed control methods: chemical, mechanical, cultural, genetic, and biological. Mathematical models provide an adequate tool for attaining this objective. An explanatory economic model was built to analyze the ecophysiological aspects of wild oats (Avena fatua L.) and barley (Hordeum vulgare L.) competitive ability, and to determine the impact of cultural and genetic control on chemical control of wild oats in irrigated barley. The competitive ability of both species was highly dependent on light interception during early growth rather than photosynthesis. Integration of genetic, cultural, and chemical control, accomplished by simulation, showed a good possibility for reducing herbicide use by using more competitive barley cultivars and planting them at higher seeding rates. 16

Dunan, C. M. and R. L. Zimdahl (1991). "Competitive Ability of Wild Oats (Avena-Fatua) and Barley (Hordeum-Vulgare)." Weed Science 39(4): 558-563.://A1991HA61200007 Replacement series and growth analysis experiments under greenhouse and field conditions quantified and explained the competitive ability of wild oats and barley. Barley was a stronger competitor than wild oats under greenhouse and field conditions. The reciprocal yield approach showed that barley's intraspecific competition was 7.3 times greater than its interspecific competition with wild oats when calculated on a dry weight per plant basis. When leaf area per plant was the yield variable, barley's intraspecific competition was only 2.4 times greater than its interspecific competition. The difference was explained by wild oats' higher leaf area ratio. Barley had a greater leaf area, root and shoot biomass, absolute growth rate, and shoot-root ratio than wild oats, but wild oats' leaf area ratio was always higher. No differences were detected in relative growth rate and net assimilation rate.

Dusky, J. A., D. G. Davies, et al. (1982). "Metabolism of Diclofop-Methyl in Cell-Cultures of Avena-Sativa and Avena-Fatua." Physiologia Plantarum 54(4): 490-494.://A1982NM59100019

Dyer, W. E. (1993). "Dormancy-Associated Embryonic Messenger-Rnas and Proteins in Imbibing Avena-Fatua Caryopses." Physiologia Plantarum 88(2): 201-211.://A1993LK47600001 The mechanisms controlling seed dormancy maintenance and release are not understood. To characterize the molecular events accompanying dormancy release, two-dimensional gel electrophoresis was used to monitor changes in soluble proteins and in vitro translation products of embryonic mRNA populations during imbibition of dormant and nondormant (after-ripened) Avena fatua L. caryopses. No differences were observed between in vitro translation products of mRNA extracted from dry dormant and nondormant embryos. However, the expression patterns of several imbibition- and germination-associated mRNAs were temporally modulated during the first 24 h of imbibition. Two dormancy-associated mRNAs, represented by polypeptides D1 and D2, were differentially overexpressed in dormant embryos after 3 h of imbibition. mRNA levels for D1 and D2 were about 8- and 3-fold higher, respectively, in dormant embryos than in nondormant embryos after 3 h of imbibition. Overexpression of D1 continued through 12 h of imbibition, while expression of both mRNAs fell to low and equivalent amounts in dormant and nondormant embryos after 24 h. Similar dormancy- associated changes in two soluble proteins were observed during imbibition. The results demonstrate that steady-state levels of specific mRNAs and proteins change during early imbibition of dormant and nondormant A. fatua embryos and indicate that these changes may be associated with differential gene expression responsible for the maintenance of dormancy.

Dyer, W. E., A. J. Kern, et al. (1996). "A novel mechanism of herbicide resistance in Avena fatua conferred by two recessive genes." Plant Physiology 111(2): 497-497.://A1996UR53400537

Eagles, H. A., R. M. Haslemore, et al. (1978). "Nitrogen-Utilization in Libyan Strains of Avena-Sterilis L with High Groat Protein and High Straw Nitrogen-Content." Journal of Agricultural Research 21(1): 65-72.://A1978FN25500008

Eberlein, C. V., T. L. Miller, et al. (1988). "Influence of Thiameturon and Dpx-L5300 on Wild Oats (Avena-Fatua) Control with Barban, Diclofop, Ac-222,293, and Difenzoquat." Weed Science 36(6): 792-799.://A1988R192800017

Ellern, S. J., J. L. Harper, et al. (1970). "Comparative Study of Distribution of Roots of Avena-Fatua and a-Striosa in Mixed Stands Using a C-14 Labelling Technique." Journal of Ecology 58(3): 865-&.://A1970I699800017

El-Shatnawi, M. K. J., H. M. Saoub, et al. (2004). "Growth and chemical composition of wild oat (Avena fatua) under Mediterranean conditions." Grass and Forage Science 59(1): 100-103.://000220736400012 Wild oat (Avena fatua) is an annual cool-season species that grows in areas with a Mediterranean climate and has potential as a forage source in Jordan. A field experiment was conducted during the growing seasons of 1999-2000 and 2000-2001 under sub-humid Mediterranean conditions at Samta in the Ajloun Mountains, Jordan. Data on seasonal herbage mass, morphology and chemical composition of wild oat were collected at 60, 80, 100, 120 and 140 days after emergence. Plant height increased rapidly beyond 100 days after emergence. The increase in herbage mass of dry matter was gradual and peaked at 140 days after emergence. The lowest concentration of crude fibre was at 60 and 80 days after emergence, with a range of 201-263 g kg(-1) DM. Crude fibre concentrations (610-630 g kg(-1) DM) peaked at 140 days after emergence (maturity). In contrast to concentrations of crude fibre, concentrations of crude protein decreased gradually with age. The calcium and phosphorus concentrations were sufficient to meet the maintenance requirements of ewes.

Evans, R. M., D. C. Thill, et al. (1991). "Wild Oat (Avena-Fatua) and Spring Barley (Hordeum-Vulgare) Density Affect Spring Barley- Grain Yield." Weed Technology 5(1): 33-39.://A1991FK00400005 Addition series field experiments were conducted near Moscow, ID, in 1987 and 1988 to determine the relative aggressiveness of spring barley and wild oat and to determine the effect of barley and wild oat density and proportion on barley grain yield and wild oat seed rain. Regression analysis was used to describe the relationship of the aboveground biomass and grain yield to species density. Barley was more aggressive than wild oat. Barley biomass was affected most by intraspecific competition, while wild oat biomass was affected most by interspecific competition. Barley aggressiveness changed little throughout the growing season. Wild oat aggressiveness varied but was always less than barley aggressiveness. Increasing wild oat density had a negative, asymptotic-type effect on barley grain yield at all barley densities. However, the effect of wild oat was greatest at the lower density of barley. Increasing barley density decreased wild oat seed rain.

Everson, C. S. and C. M. Breen (1983). "The Influence of Season on the Suppression by Pteridium-Aquilinum Leaf Extracts of Root-Growth in Avena-Fatua - a Possible Influence on the Distribution of Philipia-Evansii." South African Journal of Botany 2(4): 297- 300.://A1983RP82200003 17

Ewing, A. L. and J. W. Menke (1983). "Response of Soft Chess (Bromus-Mollis) and Slender Oat (Avena-Barbata) to Simulated Drought Cycles." Journal of Range Management 36(4): 415-418.://A1983RM22300002

Fay, P. K. and R. S. Gorecki (1978). "Stimulating Germination of Dormant Wild Oat (Avena-Fatua) Seed with Sodium Azide." Weed Science 26(4): 323-326.://A1978FE29500003

Fender, S. (1980). "'Wild Oats' - Epstein,J." Tls-the Times Literary Supplement(4013): 202-202.://A1980JG19800011

Fennimore, S. A. and M. E. Foley (1998). "Genetic and physiological evidence for the role of gibberellic acid in the germination of dormant Avena fatua seeds." Journal of Experimental Botany 49(318): 89-94.://000071821200012 Genetic and physiological data indicate that gibberellic acid does not have a primary role in the regulation of seed dormancy in wild oat (Avena fatua L.). The gibberellic acid sensitivity threshold of dormant caryopses imbibed for 7 d was 1 mu M. Intact dormant seeds were after-ripened for 0, 2, 4, 8, 12, 16, and 20 weeks at 40 degrees C and imbibed in H2O, 100 nM or 1 mu M gibberellic acid. The length of the after-ripening interval was inversely related to the mean base gibberellic acid concentration (the concentration resulting in 50% germination). Thus, after-ripening, not gibberellic acid, is the principal factor that regulates the release of seed dormancy in wild oat. F-2 caryopses (dormant x non-dormant) classified by germination response to progressively higher gibberellic acid concentrations, were pooled according to their gibberellic acid requirement: low, medium and high. Germination responses of the F-3 progeny from the low, medium, and high gibberellic acid requirement pools were regressed on to the F-2 parent values, and a heritability for germination response to gibberellic acid, h(2)=0.24, was calculated. Random amplified polymorphic DNA analysis of DNA samples from F-2 pools requiring low and high gibberellic acid concentrations were screened with 200 decamer primers and no polymorphisms were found. The findings of this investigation demonstrate that gibberellic acid is not the primary regulator of seed dormancy in wild oat.

Fennimore, S. A., W. E. Nyquist, et al. (1999). "A genetic model and molecular markers for wild oat (Avena fatua L.) seed dormancy." Theoretical and Applied Genetics 99(3-4): 711-718.://000082570500041 Seed dormancy allows weed seeds to persist in agricultural soils. Wild oat (Avena fatua L.) is a major weed of cereal grains and expresses a range of seed dormancy phenotypes. Genetic analysis of wild oat dormancy has been complicated by the difficulty of phenotypic classification in segregating populations. Therefore, little is known about the nature of the genes that regulate dormancy in wild oat. The objectives of our studies were to develop methods to classify the germination responses of segregating wild oat populations and to find molecular markers linked to quantitative trait loci (QTL) that regulate seed dormancy in wild oat. RAPD markers OPX-06 and OPT-04 explained 12.6% and 6.8% respectively, of the F-2 phenotypic variance. OPF-17 was not significant in a simple regression model, but it was linked in repulsion to OPT-04. A three-locus model of seed dormancy in wild oat is presented based on the 41-day germination profiles of F-1, F-2, F-3, BC1P1F1, BC1P1F2, and BC1P2F1 generations, and the 113 day germination profile of 126 F-7 recombinant inbred lines. Loci G(1) and G(2) promote early germination, and the D locus promotes late germination. If at least one copy of the dominant G(1) or G(2) alleles are present regardless of the genotype at D locus, then the individual will be nondormant. If the genotype is g(1)g(1)g(2)g(2)D_, then the phenotype will be dormant.

Fennimore, S. A., W. E. Nyquist, et al. (1998). "Temperature response in wild oat (Avena fatua L.) generations segregating for seed dormancy." Heredity 81: 674-682.://000077862700010 Crosses between parents with high and low levels of seed dormancy ill wild oat were used to produce F-1, F-2 and backcross populations. Germination phenotypes were determined by imbibing all populations at 15 and 20 degrees C. Rapid germination of genetically more dormant generations was favoured at the lower temperature, i.e. a generation by germination temperature interaction was observed. Evidence that dominance shifted from early germination at 15 degrees C to late germination at 20 degrees C is presented. Epistatic gene action may have been detected at 20 degrees C but not at 15 degrees C. Cumulative germination percentages of F-1 caryopses imbibed at 10, 15, 20 and 25 degrees C revealed an inverse relationship between germination rate and temperature. The narrow-sense family heritability of the seed dormancy phenotype of F-7 recombinant inbred lines pcr se was h(f,F=1)(2) = 0.75 with exact confidence limits of 0.64 and 0.83, Six factors were estimated to be segregating between the dormant and nondormant parents. Genotype by germination temperature interactions may play an adaptive role that allows wild oat to persist in diverse ecosystems.

Fernandezquinantilla, C., J. L. G. Andujar, et al. (1990). "Characterization of the Germination and Emergence Response to Temperature and Soil-Moisture of Avena-Fatua and Avena-Sterilis." Weed Research 30(4): 289-295.://A1990DQ72200008

Fernandez-Quintanilla, C., E. S. Leguizamon, et al. (2006). "Integrating herbicide rate, barley variety and seeding rate for the control of sterile oat (Avena sterilis spp. ludoviciana) in central Spain." European Journal of Agronomy 25(3): 223-233.://000241347900006 Three field experiments, each repeated over three seasons, were conducted to determine the effects of herbicide rate, winter barley variety and seeding rate on the control of sterile oat (Avena sterilis spp. ludoviciana) and the yield of barley. The herbicide rate giving 50% reduction in the weed fresh weight (ED50) was generally higher with diclofop than with imazamethabenz or tralkoxydim (26.6, 16.0 and 16.4% of full rate, respectively). Under adequate rainfall, reducing herbicide rate 50% resulted in sterile oat seed outputs very similar to those obtained with the full rate. However, in the presence of water stress, seed production was significant with low herbicide rates, particularly when weed pressure was high. Barley yields decreased gradually with decreasing herbicide rate. Although practically no yield losses were obtained by reducing the rate 50% of the recommended rate, reducing the rate down to 12.5% resulted in yields that were similar to the untreated control. Yield responses were variable in the 3 years. Reducing tralkoxydim to 50% of the recommended rate resulted in 32% crop yield loss in 1997 and only 2.7% yield loss in the other 2 years. The high yield losses in 1997 were associated with high weed densities and water stress conditions. Crop variety interacted with tralkoxydim rate to determine sterile oat growth and reproduction. The ED50 of tralkoxydim was 8% for 'Albacete' (tall and late maturing) and 18

18% for 'Barbarrosa' (medium height, medium maturing). The effectiveness of low tralkoxydim rates in reducing sterile oat fresh weight and seed production did not improve with high seeding rates. The results support the view that reducing herbicide rates down to 50% of the label rate is a safe practice when sterile oat densities are low or moderate. However, reducing the rates even further may increase the risks of yield losses and increased weed populations in subsequent years. This risk can be lowered by growing competitive barley varieties. (c) 2006 Elsevier B.V. All rights reserved.

Fernandezquintanilla, C., L. Navarrete, et al. (1986). "Seedling Recruitment and Age-Specific Survivorship and Reproduction in Populations of Avena-Sterilis L Ssp Ludoviciana (Durieu) Nyman." Journal of Applied Ecology 23(3): 945-955.://A1986F652100017

Fernandezquintanilla, C., L. Navarrette, et al. (1987). "Influence of Herbicide Treatments on the Population-Dynamics of Avena- Sterilis Ssp Ludoviciana (Durieu) Nyman in Winter-Wheat Crops." Weed Research 27(5): 375-383.://A1987K196500009

Field, R. J. and J. C. Caseley (1987). "Abscisic-Acid as a Protectant of Avena-Fatua L against Diclofop-Methyl Activity." Weed Research 27(4): 237-244.://A1987J300800002

Fischer, G. W. and C. E. Claassen (1944). "Studies of stem rust (Puccinia graminis) from Poaampla, Avena fatua, and Agropyron spicatum in the Pullman, , region." Phytopathology 34(3): 301-314.://000200839700003

Fleischm.G (1969). "Resistance of Genes Isolated from Avena Sterilis to Isolates of Oat Crown Rust Prevalent in Canada in 1968." Canadian Journal of Botany 47(4): 623-&.://A1969D315800021

Fleischm.G (1970). "Effectiveness of Resistance Genes from Wild Oats, Avena-Sterilis, against Oat Crown Rust, Puccinia-Coronata F-Sp Avenae, in Canada in 1969." Canadian Journal of Botany 48(12): 2117-&.://A1970I408300006

Fleischm.G and R. J. Baker (1971). "Oat Crown Rust Race Differentiation - Replacement of Standard Differential Varieties with a New Set of Single Resistance Gene Lines Derived from Avena-Sterilis." Canadian Journal of Botany 49(8): 1433-&.://A1971K273200025

Fleischm.G and R. I. McKenzie (1967). "Seedling Resistance to Oat Crown Rust in Avena Sterilis." Phytopathology 57(8): 811- &.://A19679776700065

Fleischm.G and R. I. McKenzie (1968). "Inheritance of Crown Rust Resistance in Avena Sterilis." Canadian Journal of Genetics and Cytology 10(3): 762-&.://A1968C096800030

Fleischm.G, R. I. McKenzie, et al. (1971). "Inheritance of Crown Rust Resistance Genes in Avena-Sterilis Collections from Israel, Portugal, and Tunisia." Canadian Journal of Genetics and Cytology 13(2): 251-&.://A1971K160100013

Fleischm.G, R. I. McKenzie, et al. (1971). "Inheritance of Crown Rust Resistance in Avena-Sterilis L from Israel." Crop Science 11(3): 451-&.://A1971J590000043

Florell, V. H. (1929). "The synthetic formation of Avena sterilis." Journal of Heredity 20: 227-227.://000200600200037

Foley, M. E. (1986). "Effect of Fusicoccin and Gibberellic-Acid on Primary Dormant Avena-Fatua Caryopses." Physiologia Plantarum 67(4): 690-694.://A1986D910000030

Foley, M. E. (1987). "The Effect of Wounding on Primary Dormancy in Wild Oat (Avena-Fatua) Caryopses." Weed Science 35(2): 180-184.://A1987G444500010

Foley, M. E. (1992). "Effect of Soluble Sugars and Gibberellic-Acid in Breaking Dormancy of Excised Wild Oat (Avena-Fatua) Embryos." Weed Science 40(2): 208-214.://A1992JC08300009 Dormant line M73 wild oat caryopses were utilized to develop a system for the culture of excised embryos, to evaluate whether embryo dormancy exists, and to investigate the physiological basis for breaking dormancy. Dormant embryos cultured on N6 medium solidified with 0.25% Bacto agar displayed 70% germination in 2 d compared with approximately 20% for the other gelling agents. The non-plant-based gelling agent gellum at a concentration of 0.25% was selected for further experiments on breaking dormancy after it was determined that concentrations greater-than-or-equal-to 0.5% decreased the rate of germination. Amending N6 medium with concentrations of 0.1 to 10-mu-M gibberellic acid (GA) increased the rate and extent of germination. Embryos treated with 0 to 0.01-mu-M GA required 6 d to attain 90% germination. Germination of dormant embryos on N6 medium without GA suggested that either true embryo dormancy did not exist in M73 or some constituent of the N6 medium promoted breaking of dormancy. Subsequent experiments indicated that the 88 mM sucrose was the constituent in the N6 medium responsible for breaking dormancy. Concentrations of sucrose from 40 to 200 mM were effective in breaking dormancy. Ten-mu-M GA increased the rate and extent of germination of embryos cultured with 88 to 200 mM sucrose. At 88 mM, fructose, maltose, glucose, and sucrose all broke embryo dormancy. Fructose was the most active soluble sugar for breaking embryo dormancy, promoting nearly 100% germination in 4 d. As with sucrose, there was an interaction between GA and the soluble sugars in breaking dormancy. Ten-mu-M GA with 88 mM fructose provided nearly 100% germination in 1 d. Amylose, but not amylopectin or pullulan, may substitute for soluble sugars. However, with 10-mu-M GA amylose, amylopectin and pullulan were equally effective in breaking dormancy. Breaking dormancy of embryos on N6 medium was independent of temperatures from 12 to 24 C in the presence of GA, but in its absence the optimum was 12 C. Application of GA to dormant caryopses significantly increased and decreased the level of glucose and sucrose, respectively, in the embryo. Gibberellic acid had a similar effect on glucose and sucrose in the endosperm tissue, except the differences were not significant at all times after treatment. The change in carbohydrate metabolism, especially in embryo tissue, may be important when considered in 19

context with the observation that soluble sugars and GA act independently in breaking dormancy in excised M73 embryos. Breaking wild oat embryo dormancy with GA may be mainly a substitution for sugar requirement.

Foley, M. E. (1994). "Temperature and Water Status of Seed Affect Afterripening in Wild Oat (Avena-Fatua)." Weed Science 42(2): 200-204.://A1994NR43600009 Dormant wild oat seed require afterripening under warm-dry conditions for conversion to a nondormant state capable of germination. Research was conducted to determine the relationship between temperature and seed moisture levels on afterripening of dormant wild oat line M73 seed, and to evaluate the status of water binding in the dormant seed. Conversion of dominant wild oat seed to a nondormant state at 20 to 40 C occurs primarily in the range of 5 to 20% seed moisture. There is an inverse relationship between temperature and seed moisture content for after-ripening as measured by seed germination. As the afterripening temperature increases, the seed moisture content must decrease for maximum afterripening (germination) to occur. Moisture isotherms and derived enthalpy curves indicate three regions of water binding which reflect decreased binding of water to seed components as the moisture content in the dormant seed increases. Maximum afterripening, in the second region of water binding, corresponds to seed moisture contents of 7 to 22 %. In this region water is weakly associated with macromolecular surfaces and begins to have solvent properties. Because afterripening occurs mainly in the second region of water binding it is likely that individual enzymatic and nonenzymatic reactions, rather than metabolic processes, mediate the conversion of wild oat seed to the nondormant state.

Fox, S. L., P. D. Brown, et al. (1997). "Inheritance of crown rust resistance in four accessions of Avena sterilis L." Crop Science 37(2): 342-345.://A1997WU90500005 Genetic resistance in common oat, Avena sativa L., provides an effective means of controlling crown rust, caused by Puccinia coronata Cda. f. sp. avenae Eriks. Four Avena sterilis L. accessions (IB 1487, IB 2402, IB 2465, and IB 3432) were selected for a genetic study based on low disease reaction to 15 isolates of crown rust and ease of hybridization to common oat. Crosses were made to determine the mode of inheritance of the resistance in each accession and to detect the linkage relationship of the resistance genes with several known crown-rust resistance genes. The accessions were crossed to the susceptible common oat line Rodney 0, intercrossed among themselves, and crossed to six Lines, each with a single gene for crown rust resistance: Pc38, Pc56, Pc58, Pc61, Pc63, and Pc68. In a greenhouse, the F-2 populations were evaluated for seedling reaction after inoculation of the first leaf with crown rust isolate CR 13. Each accession had a single incompletely dominant or dominant gene conferring resistance, and genes were designated as Gene A in IB 1487, Gene B in IB 2402, Gene C in IB 2465, and Gene D in IB 3432. Gene A is allelic or closely linked to Pc56. Genes B and C are allelic or closely linked to Pc68. Having not been described previously, Gene D may be a useful addition to existing germplasm.

Fredeen, A. L. and C. B. Field (1995). "Contrasting Leaf and Ecosystem Co2 and H2o Exchange in Avena-Fatua Monoculture - Growth at Ambient and Elevated Co2." Photosynthesis Research 43(3): 263-271.://A1995RH51300009 Elevated CO2 (ambient + 35 Pa) increased shoot dry mass production in Avena fatua by similar to 68% at maturity. This increase in shoot biomass was paralleled by an 81% increase in average net CO2 uptake (A) per unit of leaf area and a 65% increase in average A at the 'ecosystem' level per unit of ground area. Elevated CO2 also increased 'ecosystem' A per unit of biomass. However, the products of total leaf area and light-saturated leaf A divided by the ground surface area over time appeared to lie on a single response curve for both CO2 treatments. The approximate slope of the response suggests that the integrated light saturated capacity for leaf photosynthesis is similar to 10-fold greater than the 'ecosystem' rate. 'Ecosystem' respiration (night) per unit of ground area, which includes soil and plant respiration, ranged from -20 (at day 19) to -18 (at day 40) mu mol m(-2) s(-1) for both elevated and ambient CO2 Avena. 'Ecosystem' below- ground respiration at the time of seedling emergence was similar to -10 mu mol m(-2) s(-1), while that occuring after shoot removal at the termination of the experiment ranged from -5 to -6 mu mol m(-2) s(-1). Hence, no significant differences between elevated and ambient CO2 treatments were found in any respiration measure on a ground area basis, though 'ecosystem' respiration on a shoot biomass basis was clearly reduced by elevated CO2. Significant differences existed between leaf and 'ecosystem' water flux. In general, leaf transpiration (E) decreased over the course of the experiment, possibly in response to leaf aging, while 'ecosystem' rates of evapotranspiration (ET) remained constant, probably because falling leaf rates were offset by an increasing total leaf biomass. Transpiration was lower in plants grown at elevated CO2, though variation was high because of variability in leaf age and ambient light conditions and differences were not significant. In contrast, 'ecosystem' evapotranspiration (ET) was significantly decreased by elevated CO2 on 5 out of 8 measurement dates. Photosynthetic water use efficiencies (A/E at the leaf level, A/ET at the 'ecosystem' level) were increased by elevated CO2. Increases were due to both increased A at leaf and 'ecosystem' level and decreased leaf E and 'ecosystem' ET.

Frederick, E. C., W. J. Aunan, et al. (1964). "High Moisture + Dry Wild Oats for Fattening Steers." Journal of Animal Science 23(3): 875-&.://A19643675B00094

Frey, K. J., T. McCarty, et al. (1975). "Straw-Protein Percentages in Avena-Sterilis L." Crop Science 15(5): 716-718.://A1975AY18200030

Friesen, G. (1967). "Efficiency of Barban as Influenced by Growth Stages of Wild Oats and Spring Wheat." Weeds 15(2): 160-&.://A19679983800019

Friesen, G. H. (1987). "Control of Wild Oats in Field Corn with Flamprop Methyl." Canadian Journal of Plant Science 67(1): 271- 274.://A1987G182000039

Friesen, H. A. and K. E. Bowren (1973). "Factors Affecting Control of Wild Oats in Rapeseed with Trifluralin." Canadian Journal of Plant Science 53(1): 199-205.://A1973O780300036

20

Friesen, H. A. and O. B. Litwin (1975). "Selective Control of Wild Oats in Barley with Ac-84777." Canadian Journal of Plant Science 55(4): 927-934.://A1975AY75400008

Friesen, H. A., P. A. Osullivan, et al. (1976). "Hoe 23408, a New Selective Herbicide for Wild Oats and Green Foxtail in Wheat and Barley." Canadian Journal of Plant Science 56(3): 567-578.://A1976CE92500021

Friesen, L. F., T. L. Jones, et al. (2000). "Identification of Avena fatua populations resistant to imazamethabenz, flamprop, and fenoxaprop-P." Weed Science 48(5): 532-540.://000089783100002 Three Avena fatua (wild oat) populations resistant to imazamethabenz, flamprop, and fenoxaprop-P were identified from the northwest agricultural region of Manitoba, Canada. These populations were identified after producer reports of failure of imazamethabenz to provide satisfactory control in the field. Although these A. fatua populations had previously been exposed to other herbicides, primarily ACCase inhibitors, imazamethabenz had never before been applied. In growth room experiments, resistant (R) plants were 7.2 and 8.7 times more resistant to imazamethabenz and flamprop, respectively than susceptible (S) plants, as measured by the ratio of dosages required to inhibit shoot dry matter accumulation by 50% (GR(50) R/S) The three populations did not differ significantly (P < 0.05) in levers of resistance to imazamethabenz. Similarly, the populations did not differ in levels of resistance to flamprop. The populations differed in their response to fenoxaprop-P; levels of resistance for two populations were 2.0-fold, while the remaining population was 2.9-fold. An experiment conducted in 1995 in one of the infested fields confirmed multiple herbicide resistance, with A. fatua panicle numbers in August being 36, 128, and 44% of untreated controls at recommended dosages of imazamethabenz, flamprop, and fenoxaprop-P, respectively. Three additional populations of A. fatua with multiple herbicide resistance from other areas of Manitoba were identified in a 1996 field experiment. For the six A. fatua populations in the 1996 experiment with multiple herbicide resistance, panicle numbers expressed as a percentage of the untreated controls varied from 44 to 77% for imazamethabenz, 57 to 83% for flamprop, and 43 to 88% for fenoxaprop-P (commercially recommended dosage of each herbicide). Multiple herbicide resistance in A. fatua is not rare; screening of A. fatua seed samples from across Manitoba and Saskatchewan has identified a number of additional R populations. The evolution of herbicide resistance in the absence of direct selection is a very serious development. as producers with multiple herbicide resistance in A. fatua are left with a very limited number of herbicide options for selective control in crops commonly grown in western Canada.

Froment, M. A. and N. D. Cooper (1996). "Evaluation of post-emergence herbicides for the control of wild oats (Avena fatua) in winter barley." Annals of Applied Biology 128: 24-25.://A1996WF77600012

Fuerst, E. P., M. K. Upadhyaya, et al. (1983). "A Study of the Relationship between Seed Dormancy and Pentose-Phosphate Pathway Activity in Avena-Fatua." Canadian Journal of Botany-Revue Canadienne De Botanique 61(3): 667-670.://A1983QH54600005

Gallandt, E. R., E. P. Fuerst, et al. (2004). "Effect of tillage, fungicide seed treatment, and soil fumigation on seed bank dynamics of wild oat (Avena fatua)." Weed Science 52(4): 597-604.://000222859200017 No-tillage offers potential for improved soil quality, reduced erosion, and equal or increased crop yields. We hypothesized that, compared with conservation tillage (CT), no-tillage (NT) offers conditions more conducive to microbial decay of weed seed. In NT systems seed remain at or near the soil surface where crop residues, moisture, and lack of disturbance create an environment with greater soil microbial diversity. In late fall of 1998 and 1999, dormant seed of wild oat, either individually glued to plastic toothpicks or mixed with soil and placed in mesh bags, were buried (mean seed depth of 2.5 cm) in replicated field plots managed by NT or CT since 1982. Treatments including fungicide seed treatment (thiram + metalaxyl + captan) and soil fumigation (propylene oxide) provided estimates of the contribution of microorganisms to observed mortality. Seed were retrieved in May and August, 1999 and 2000. Contrary to our original hypothesis, the proportion of dead seed was generally similar in NT and CT systems. Lack of tillage system by seed or soil treatments affecting the proportion of dead or decayed seed suggests that the contribution of microorganisms to seed fate is similar in these tillage environments. However, the proportion of dormant seed was consistently lower in the NT compared with CT treatments; there was a corresponding increase in the proportion of germinated seed. Overall, more than half of the wild oat seed bank losses could be directly attributed to germination whereas losses due to decay were relatively minor by comparison. Despite favorable distribution of seed and improved quality of the surface-strata of soil in NT systems, this study fails to provide evidence that enhanced microbial decay will contribute to a "weed-suppressive" capacity in such cropping systems.

Garber, R. J. (1922). "Origin of false wild oats." Journal of Heredity 13: 40-48.://000200599800007

Garber, R. J. and K. S. Quisenberry (1923). "Delayed germination and the origin of false wild oats." Journal of Heredity 14: 267- 274.://000200599900052

Garcia, P., M. I. Morris, et al. (1991). "Genetic Diversity and Adaptedness in Tetraploid Avena-Barbata and Its Diploid Ancestors Avena Hirtula and Avena-Wiestii." Proceedings of the National Academy of Sciences of the United States of America 88(4): 1207- 1211.://A1991EY61700026 Avena barbata, a tetraploid grass, is much more widely adapted and successful in forming dense stands than its diploid ancestors. The success of such polyploids has often been attributed to heterosis associated with ability to breed true for a highly heterozygous state in which allelic differences between the parents are fixed in the polyploid by chromosome doubling. We have examined the relationship between genetic diversity and adaptedness for 14 allozyme loci in A. barbata and its diploid ancestors in samples collected from diverse habitats in Israel and Spain. The relationship varied from locus to locus: superior adaptedness was associated with genetic uniformity for five loci, in part with genetic uniformity and in part with genetic diversity (monomorphism for a single heteroallelic quadriplex) for one locus, and with allelic diversity in the form of heteroallelic quadriplexes combined with genotypic diversity in the form of complex polymorphisms among different homoallelic and/or heteroallelic quadriplexes for the eight remaining loci. These results indicate that allelic diversity fixed in nonsegregating form through chromosome doubling was an important factor in the evolution of 21

adaptedness in A. barbata. However, it is unlikely that heterosis associated with heterozygosity contributed significantly to superior adaptedness in either the diploids or the tetraploid because virtually all loci (almost-equal-to 99%) were homozygous in the Avena diploids and tetraploid.

Garcia, P., F. J. Vences, et al. (1989). "Allelic and Genotypic Composition of Ancestral Spanish and Colonial Californian Gene Pools of Avena-Barbata - Evolutionary Implications." Genetics 122(3): 687-694.://A1989AE30400022

Gardner, K. M. and R. G. Latta (2006). "Identifying loci under selection across contrasting environments in Avena barbata using quantitative trait locus mapping." Molecular Ecology 15(5): 1321-1333.://000236584900010 We constructed recombinant inbred lines of a cross between naturally occurring ecotypes of Avena barbata (Pott ex Link), Poaceae, associated with contrasting moisture environments. These lines were assessed for fitness in common garden reciprocal transplant experiments in two contrasting field sites in each of two years, as well as a novel, benign greenhouse environment. An AFLP (amplified fragment length polymorphism) linkage map of 129 markers spanned 644 cM in 19 linkage groups, which is smaller, with more linkage groups, than expected. Therefore parts of the A. barbata genome remain unmapped, possibly because they lack variation between the ecotypes. Nevertheless, we identified QTL (quantitative trait loci) under selection in both native environments and in the greenhouse. Across years at the same site, the same loci remain under selection, for the same alleles. Across sites, an overlapping set of loci are under selection with either (i) the same alleles favoured at both sites or (ii) loci under selection at one site and neutral at the other. QTL under selection in the greenhouse were generally unlinked to those under selection in the field because selection acted on a different trait. We found little evidence that selection favours alternate alleles in alternate environments, which would be necessary if genotype by environment interaction were to maintain genetic variation in A. barbata. Additive effect QTL were best able to explain the genetic variation among recombinant inbred lines for the greenhouse environment where heritability was highest, and past selection had not eliminated variation.

Goffreda, J. C., W. B. Burnquist, et al. (1992). "Application of Molecular Markers to Assess Genetic-Relationships among Accessions of Wild Oat, Avena-Sterilis." Theoretical and Applied Genetics 85(2-3): 146-151.://A1992JY22200004 The Avena sterilis collection in the National Small Grains Collection (NSGC) is an invaluable source of genetic variation to be exploited by oat breeding programs. Prior knowledge of the structure and distribution of genetic variation within the A. sterilis collection would be useful to efficiently screen the collection for valuable traits. To determine genetic structure within a subset of the collection, restriction fragment length polymorphisms were analyzed in a stratified sample of 173 accessions originating in eight countries of Africa and South-west Asia. Of the 48 probes used for this study 43 detected polymorphism among accessions. The average number of RFLP patterns per probe ranged from 2.9 among Ethiopian accessions to 3.7 among those from . Genetic variation, as measured by genetic distances and polymorphic indexes, was highest in Iran and lowest in Ethiopia. The probability of drawing a genotype from Iran or Iraq that is not present in the more western regions was high, indicating large genetic divergence of the Iran-Iraq accessions from the other regional collections surveyed. Cluster analysis of genetic distances and probabilities of unique genotypes clearly differentiated the eastern region (Iran and Iraq) from the western region (Algeria, Ethiopia, Israel, Lebanon, Morocco, and Syria). The western region could be further subdivided into two clusters, an African cluster (Algeria, Ethiopia, and Morocco) and a southwestern Asia cluster (Israel, Lebanon, and Syria). Genetic distances were generally related to but not proportional to geographical distances.

Gonzalezandujar, J. L. and C. Fernandezquintanilla (1991). "Modeling the Population-Dynamics of Avena-Sterilis under Dry-Land Cereal Cropping Systems." Journal of Applied Ecology 28(1): 16-27.://A1991FN20800002 (1) A mathematical model for simulating the population dynamics of Avena sterilis ssp. ludoviciana (Dur.) Nyman has been constructed using previously reported data. The model considers the age structure of the population of seedlings as well as the effects of density on plant survivorship and reproduction. (2) The model is used to describe the behaviour of the population in the absence of control practices and to predict the effects of various control strategies. In the absence of control, and under continuous winter cereal cropping, the population grows hyperbolically, reaching equilibrium at a density of 535 plants m-2. Annual application of herbicides with < 85% control results in moderate reductions in the equilibrium level. To obtain a negative growth of the population it is necessary to apply herbicides annually with a control level of > 90%. Fallowing the land for 1 in every 2-3 years gave a practical method of containing the populations of A. sterilis. However, to eradicate this weed it was necessary to combine crop rotation with application of herbicides. (3) The effects of changing the values of the parameters on the output of the model were generally minor. The two processes most sensitive to parameter variation were dispersal and mortality of seeds after reproduction and the fecundity of the first cohort of plants. The contribution of late emerging plants to the overall dynamics of the population was rather small and could be disregarded. (4) The model was validated by comparing simulation results with those from long-term field studies. Model predictions closely matched experimental results from herbicide trials, but gave only a crude description of the population dynamics under various crop rotations.

Gonzalezandujar, J. L. and C. Fernandezquintanilla (1993). "Strategies for the Control of Avena-Sterilis in Winter-Wheat Production Systems in Central Spain." Crop Protection 12(8): 617-623.://A1993MH91400009 A bioeconomic model is described and used to investigate the agronomic and economic consequences of using a range of management strategies for the control of winter wild oats (Avena sterilis L.) in cereal cropping systems representative of central Spain. The results of simulations indicated that growing winter wheat continuously with the annual application of herbicides may be the optimum strategy, resulting in acceptable wild oat populations and maximum economic benefits. However, the practice of wheat monoculture was only a valid option as long as herbicides were applied annually: spraying herbicides in alternate years failed to control wild oats adequately and resulted in major economic losses. The rotation of wheat with a fallow year, with no herbicides applied in either of the two years, may be a satisfactory low-cost alternative when wild oat infestation levels are low, but it is not valid when infestation levels are high. The strategy that combines the use of a fallow year with herbicide application in the wheat year resulted in optimum wild oat control and moderate profitability under all conditions. However, the net returns obtained were substantially lower than in the continuous wheat plus herbicide strategy. The sensitivity of the model to variation in various key parameters was tested: wheat yield level 22

and fixed costs were the two parameters that had the largest effect on model output. In general, the effect of changing parameter values was more pronounced in continuous wheat systems than in wheat-fallow rotations

Gonzalezandujar, J. L. and J. N. Perry (1995). "Models for the Herbicidal Control of the Seed Bank of Avena-Sterilis - the Effects of Spatial and Temporal Heterogeneity and of Dispersal." Journal of Applied Ecology 32(3): 578-587.://A1995RQ01500012 1. A metapopulation neighbourhood model of the seed bank of an annual plant, that included the effects of heterogeneity in space and time, of stochastic local extinction and of dispersal, was modified using data reported previously, to examine control of the arable weed Avena sterilis. 2. In the absence of herbicide, for spatially homogeneous environments, few differences were found in the modelled mean predicted population for two levels of dispersal (strong and moderate), although the rate of spread and the variance of the number of seeds per cell were greater for the higher level of dispersal, For spatially heterogeneous environments, with strong dispersal, an increase of the spatial scale of patchiness increased the variance, whereas moderate dispersal had the opposite effect. The introduction of temporal heterogeneity did not affect the results greatly; nor did the inclusion of variation in the fecundity parameter. 3. With the introduction of a herbicide with spatially variable efficacy, the modelled metapopulation in all cases declined exponentially and became globally extinct in approximately 20 years; strongly dispersed populations with large-scale spatial heterogeneity were slightly more persistent. However, in all cases, it was usually possible to decrease the population within the model to acceptable levels (<10 seedlings m(-2)) within a period of between 3 and 5 years. Spatial variability was considerable and extreme patch persistence was occasionally observed.

Graham, S. B. (1977). "Reaping Wild Oats." American Bar Association Journal 63(AUG): 1046-1046.://A1977DQ94000005

Grahamyooll, A. (1987). "The Wild Oats They Sowed - Latin-American Exiles in Europe." Third World Quarterly 9(1): 246-253.://A1987F854800014

Grant, R. (1994). "Simulation of Competition between Barley and Wild Oats under Different Managements and Climates." Ecological Modelling 71(4): 269-287.://A1994MW91300005 The simulation of competition among different plant populations growing within a common ecosystem should be based upon the explicit simulation of the processes whereby these populations compete for irradiance, water and nutrients. In the mathematical ecosystem model presented here, each plant population is simulated independently within a common soil- atmosphere ecosystem. Exposure to irradiance is calculated from the vertical distribution of leaf area, calculated in turn from the elongation of each internode, sheath (if monocot) or petiole (if dicot), and leaf on each tiller or branch of each population within a common canopy. Access to water and nutrients is determined by the vertical distribution of root length and surface area, calculated from the elongation of primary and secondary root axes of each population through a common soil profile. The model reproduced losses of grain yield by barley (Hordeum vulgare L.) from 0 to 70% caused by different densities and emergence dates of wild oats (Avena fatua L.) that were recorded from several field trials in central Alberta. The sensitivity of simulated yield losses to wild oat competition under different climate and management was then compared to that reported in the literature. Examination of model results led to the hypothesis that sensitivity to competition among plant populations is determined by the availability of water, nutrients and other ecological resources at the site of study. Contrasting results of such competition in the literature may perhaps be explained by this hypothesis, but the absence of detailed data for soil and climate from experimental sites prevents rigorous testing.

Gressel, J. (1992). "Indiscriminate Use of Selectable Markers - Sowing Wild Oats." Trends in Biotechnology 10(11): 382-382.://A1992JZ38100004

Griffiths, D. J. and T. D. Johnston (1956). "Origin of the Common Wild Oat, Avena-Fatua L." Nature 178(4524): 99-100.://A1956ZQ46300037

Guilleme.R (1971). "Wild Oats in Vienne." Phytoma 23(232): 24-&.://A1971K698300003

Guma, I. R., M. P. de la Vega, et al. (2006). "Isozyme variation and genetic structure of populations of Avena barbata from ." Genetic Resources and Crop Evolution 53(3): 587-601.://000236615800016 Genetic diversity was analysed in 52 Argentinian populations of Avena barbata, a tetraploid grass introduced in America from Spain during the colonization period. Nine isozyme systems were studied and 14 loci identified, five of which were polymorphic. Cluster analysis based on Hedrick's index revealed a high similarity among populations. The total diversity (P-T) in the 52 populations was 0.144, the mean diversity (P-s) was 0.04, while between population diversity (D-ST) was 0.103. The resulting coefficient of differentiation (G(ST)) was 0.714, indicating that diversity among populations was an important contributor to the total variability. Genetic diversity was structured into multilocus associations; 122 different complexes were found among 3311 individuals, but only two complexes occurred at a high frequency. The distribution pattern of these frequent multilocus genotypes was associated with environmental factors, mainly rainfall and temperature. The comparison of these results with those of previous studies on A. barbata from Spain indicated that Spanish and Argentinian populations are closely similar in allelic composition on a locus-by-locus basis but different in multilocus genotypic composition. We concluded that selection was the main force involved in the reorganization of the Spanish genepool into novel multilocus associations adapted to specific habitats in Argentina.

Haizel, K. A. (1972). "Canopy Relationship of Pure and Mixed Populations of Barley (Hordeum-Vulgare L)White Mustard (Sinapis- Alba L) and Wild Oats (Avena-Fatua L)." Journal of Applied Ecology 9(2): 589-&.://A1972N479000022

Haizel, K. A. and J. L. Harper (1973). "Effects of Density and Timing of Removal on Interference between Barley, White Mustard and Wild Oats." Journal of Applied Ecology 10(1): 23-31.://A1973P627300003

23

Hakimelahi, A. and R. W. Allard (1983). "Distribution of Homeoalleles at 2 Loci in a Diploidized Tetraploid - Leucine Aminopeptidase Loci in Avena-Barbata." Journal of Heredity 74(5): 379-380.://A1983RH34800015

Halling, B. P. and R. Behrens (1983). "Effects of Difenzoquat on Photoreactions and Respiration in Wheat (Triticum-Aestivum) and Wild Oat (Avena-Fatua)." Weed Science 31(5): 693-699.://A1983RJ08100021

Hamadi, Z., Y. Steinberger, et al. (2000). "Decomposition of Avena sterilis litter under arid conditions." Journal of Arid Environments 46(3): 281-293.://000165718600006 The influence of abiotic conditions on Avena sterilis plant litter decomposition was studied at three sites along a topoclimatic gradient in the Judean Desert, Israel. Decomposition of A. sterilis plant litter followed a two-phase pattern: early and rapid mass loss during the winter season followed by a long period of low mass loss during the remaining seasons. Differences in decomposition rates were found between winter and summer, and between material on the soil surface and buried in the soil. There were significant differences in decomposition rates attributable to topoclimatic location, to which the short rainy-winter season contributed significantly. Lignin concentration was found to increase significantly during the rapid phase of decomposition, however a decrease in this value occurred during the remaining dry seasons. Litter nitrogen content oscillated during the study period such that in the buried litter a bi-phasic pattern similar to the litter decomposition in the above-ground litter was obtained. These results suggest that litter decomposition is not affected by climate but by litter lignin content, since an inverse relationship was found between lignin content and the decomposition rate. (C) 2000 Academic Press.

Hamrick, J. L. (1975). "Influence of Microhabitat Heterogeneity on Gene Frequency Distribution in Avena-Barbata." Genetics 80(3): S39-S39.://A1975AK01300107

Hamrick, J. L. and R. W. Allard (1972). "Microgeographical Variation in Allozyme Frequencies in Avena-Barbata." Proceedings of the National Academy of Sciences of the United States of America 69(8): 2100-&.://A1972N243300025

Hamrick, J. L. and R. W. Allard (1973). "Correlations between Quantitative Characters and Allozyme Genotypes in Avena- Barbata." Genetics 74(JUN): S105-S106.://A1973Q333000309

Hamrick, J. L. and R. W. Allard (1975). "Correlations between Quantitative Characters and Enzyme Genotypes in Avena-Barbata." Evolution 29(3): 438-442.://A1975BC35000005

Hamrick, J. L. and L. R. Holden (1979). "Influence of Microhabitat Heterogeneity on Gene-Frequency Distribution and Gametic Phase Disequilibrium in Avena-Barbata." Evolution 33(2): 521-533.://A1979HE57800001

Harder, D. E., J. Chong, et al. (1990). "Inheritance of Resistance to Puccinia-Coronata-Avenae and P-Graminis-Avenae in an Accession of Avena-Sterilis from Spain." Genome 33(2): 198-202.://A1990DD12200006

Harder, D. E. and J. W. Martens (1985). "Plant-Protection - Diseases and Insects Wild Oat Germplasm as a Source of Disease Resistance in Oats." Canadian Journal of Plant Science 65(3): 816-816.://A1985ANR6300060

Harder, D. E. and R. I. H. McKenzie (1984). "Complex Additive-Type of Resistance to Puccinia-Coronata in Avena-Sterilis." Canadian Journal of Plant Pathology-Revue Canadienne De Phytopathologie 6(2): 135-138.://A1984SY58900008

Harder, D. E., R. I. H. McKenzie, et al. (1980). "Inheritance of Crown Rust Resistance in 3 Accessions of Avena-Sterilis." Canadian Journal of Genetics and Cytology 22(1): 27-33.://A1980JP01000005

Harder, D. E., R. I. H. McKenzie, et al. (1984). "Inheritance of Adult-Plant Resistance to Crown Rust in an Accession of Avena- Sterilis." Phytopathology 74(3): 352-353.://A1984SH99500022

Harker, K. N., K. J. Kirkland, et al. (2003). "Early-harvest barley (Hordeum vulgare) silage reduces wild oat (Avena fatua) densities under zero tillage." Weed Technology 17(1): 102-110.://000181848500016 Effective, long-term wild oat management requires an integrated approach that uses management techniques beyond simple herbicide application. A 5-yr (1996 to 2000) zero tillage study was conducted to assess the influence of barley harvest timing on wild oat densities in subsequent years at Lacombe, AB, Canada and Melfort, SK, Canada. Harvest timings included barley harvested 1 wk after heading (early), approximately 14 to 16 d later at the soft dough stage (normal), and at maturity (grain). In the absence of herbicides, wild oat densities decreased in silage plots harvested early and increased in grain plots. Reductions were more distinct at Lacombe where barley phenological differences and whole plant moisture contents between early and normal silage harvests were greater than at Melfort. Half rates of wild oat herbicides (ICIA 0604 and imazamethabenz) did not augment reductions in wild oat densities after early silage harvest, but did improve wild oat management after normal silage harvest, and in grain production. At Lacombe, early silage harvest reduced wild oat densities more than did herbicides in grain production. Similar trends were apparent at Melfort but were not statistically significant. Early barley silage harvests may be an effective integrated weed management tool for wild oat.

Harlan, H. V. (1929). "The weedishness of wild oats - A reluctant and backbreaking study in adaptation." Journal of Heredity 20: 515-518.://000200600200081

Harrold, R. L., D. L. Craig, et al. (1980). "Nutritive-Value of Green or Yellow Foxtail, Wild Oats, Wild Buckwheat or Redroot Pigweed Seed as Determined with the Rat." Journal of Animal Science 51(1): 127-131.://A1980KC66300017

Hart, J. W. and A. M. M. Berrie (1966). "Germination of Avena Fatua under Different Gaseous Environments." Physiologia Plantarum 19(4): 1020-&.://A19668752300018 24

Hart, J. W. and A. M. M. Berrie (1968). "Relationship between Endogenous Levels of Malic Acid and Dormancy in Grain of Avena Fatua L." Phytochemistry 7(8): 1257-&.://A1968B447400006

Haynes, B., R. T. Koide, et al. (1991). "Phosphorus Uptake and Utilization in Wild and Cultivated Oats (Avena Spp)." Journal of Plant Nutrition 14(10): 1105-1118.://A1991GN70900009 In a solution culture experiment, the growth and nutrient uptake of wild oat plants (Avena fatua L.) was compared with that of cultivated oat plants (Avena sativa L. cv. Pennlo) at both low (5-mu-M P) and high (100-mu-M P) levels of phosphate. Solution phosphate depletions and plant growth variables were measured through daily solution sampling and destructive tissue harvests. Wild oat produced more root mass and accumulated P more rapidly than the cultivar in both low and high P levels. Wild oat plants also produced significantly more dry matter for a given accumulation of P than the cultivated oat plants in both low and high P solutions. We conclude that wild oat is likely to be competitively superior to the Pennlo cultivar under conditions of moderate P limitation due to its greater allocation of dry matter to roots and higher P use efficiency.

Heap, I. M., B. G. Murray, et al. (1993). "Resistance to Aryloxyphenoxypropionate and Cyclohexanedione Herbicides in Wild Oat (Avena-Fatua)." Weed Science 41(2): 232-238.://A1993LQ24300012 Resistance to aryloxyphenoxypropionate and cyclohexanedione herbicides was identified in four wild oat populations from western Canada. Populations UM1, UM2, and UM3 originated from northwestern Manitoba and UM33 from south-central Saskatchewan. Field histories indicated that these populations were exposed to repeated applications of diclofop-methyl and sethoxydim over the previous 10 yr. The populations differed in their levels and patterns of cross-resistance to these and five other acetyl-CoA carboxylase inhibitors (ACCase inhibitors). UM1, UM2, and UM3 were resistant to diclofop- methyl, fenoxaprop-p-ethyl, and sethoxydim. In contrast, UM33 was resistant to the aryloxyphenoxy propionate herbicides but not to sethoxydim. The dose of sethoxydim required to reduce growth of UM1 by 50% was 150 times greater than for a susceptible population (UM5) or UM33 based on shoot dry matter reductions 21 d after treatment. This population differed from UM2 and UM3 that had R/S ratios of less than 10. In the field UM1 also exhibited a very high level of resistance to sethoxydim. In contrast to susceptible plants that were killed at the recommended dosage, shoot dry matter of resistant plants treated at eight times the recommended dosage was reduced by only 27%. In growth chamber experiments none of the four populations was cross-resistant to herbicides from five different chemical families.

Henson, J. F. and L. S. Jordan (1982). "Wild Oat (Avena-Fatua) Competition with Wheat (Triticum-Aestivum and Triticum- Turgidum-Durum) for Nitrate." Weed Science 30(3): 297-300.://A1982NN70600016

Herman, D. J., K. K. Johnson, et al. (2006). "Root influence on nitrogen mineralization and nitrification in Avena barbata rhizosphere soil." Soil Science Society of America Journal 70(5): 1504-1511.://000240666800009 Micro-N-15 pool dilution was used to quantify rates of gross N mineralization, consumption, and nitrification in bulk soil and in soil within 2 mm of root sections of Avena barbata (slender wild oats), an annual grass common to California oak woodland-savannas. Rates of gross N mineralization in rhizosphere soil (9.2 mg N kg(-1)d(-1)) were about ten times higher than in bulk soil (1.0 mg N kg(-1)d(-1)). Total bacterial numbers in soil adjacent to roots were slightly higher than in bulk soil; protozoa biomass was not measurably different. Changes in bacterial numbers or standing stocks of bacterial N could not account for rates of N mineralization. Nitrification potential values were similar in bulk and rhizosphere soil, yet gross rates of nitrification were highly dependent on location along the root. Gross nitrification rates in soil near the root tip were the same as those in bulk soil, while rapid uptake of NH4 by older sections of root (8-16 cm from the tip), appeared to limit nitrification rates. Only small differences in microbial community structure between bulk and rhizosphere soil were detected by terminal restriction fragment length polymorphism (TRFLP) analysis. While the small increases in bacterial numbers and changes in community composition may in-part explain the increased rates of N mineralization, other microbial-root interactions are likely involved in accelerating the flux of N from organic sources to the plant-available NH4 pool. The high rates of N mineralization observed in soil immediately adjacent to roots should facilitate plant access to N. Most of the stocks and fluxes determined in these studies exhibited distinct spatial patterns along the plant root that may have significantly impacted N-availability to the plant.

Hetherington, S. D. and B. A. Auld (2001). "Host range of Drechslera avenacea, a fungus with potential for use as a biological control agent of Avena fatua." Australasian Plant Pathology 30(3): 205-210.://000170940400003 A range of plants was tested for their susceptibility to Drechslera avenacea (M.A. Curtis ex Cooke) Shoem., the possible active ingredient in a bioherbicide for use against the weed Avena fatua. Test plants were chosen on the basis of phylogenetic proximity to the weed, recorded susceptibility to closely related pathogens or physical proximity in agricultural systems to wheat, the crop in which the weed is to be controlled. The fungus caused mortality of A. fatua and Avena byzantina cv. Yarren. Sporulation occurred on A. fatua and, on one occasion (of three challenges), on Brassica napus cv. Oscar. Severe disease symptoms were recorded on 13 species of grass and mild infection on a further seven species. We observed a much broader host range than previously recorded. The reasons for, and implications of, this are discussed and the suitability of D. avenacea as a bioherbicide assessed.

Hetherington, S. D., H. E. Smith, et al. (2002). "Effects of some environmental conditions on the effectiveness of Drechslera avenacea (Curtis ex Cooke) Shoem.: a potential bioherbicidal organism for Avena fatua L." Biological Control 24(2): 103-109.://000176148000001 Drechslera avenacea is a potential bioherbicide for Avena fatua, wild oat, control in dryland wheat crops in southern Australia. Maximum disease severity (DS) (1.1 lesions per mm(2) of leaf tissue) was recorded following application of 1 x 10(5) spores per ml and exposure of weeds to a 12- to 16-h dew period at 20-25 degreesC. Weed seedlings between 3 and 5 weeks of age were the most susceptible to the disease. Occurrence of a low humidity (< 50% R.H.) for up to 8-h before the initiation of a dew period following inoculation did not reduce subsequent DS. Repeated 8-h (suboptimal) dew periods increased DS when compared with a single 8-h dew period (0.26 lesions per mm(2) compared to 0.39 lesions per mm(2), respectively). This was not the case for repeated 6-h dew periods. Following inoculation, temperature had a direct 25

effect on disease development rate. Plants recovered following inoculation and inoculated and uninoculated plants were not significantly different in terms of necrotic leaves 28 days after inoculation. Dew period limitations and the ability of D. avenacea to indirectly create a competitive advantage for wheat seedlings over wild oat are discussed in relation to the feasibility of this bioherbicide agent. (C) 2002 Elsevier Science (USA). All rights reserved.

Heun, M., J. P. Murphy, et al. (1994). "A Comparison of Rapd and Isozyme Analyses for Determining the Genetic-Relationships among Avena-Sterilis L Accessions." Theoretical and Applied Genetics 87(6): 689-696.://A1994MU94000009 Isozyme analysis is a valuable tool for determining genetic relationships among breeding lines and populations. The recently developed DNA technologies which can assay a greater proportion of the plant genome are providing a plentiful array of additional genomic markers. The objective of this research was to compare random amplified polymorphic DNA (RAPD) versus isozyme-based estimation of relationships among 24 accessions of a hexaploid wild oat, Avena sterilis L. The accessions were evaluated for variation in 23 enzyme systems and by 21 10-mer primers. A total of 77 polymorphic isozyme bands and 115 polymorphic RAPD bands were observed. Two matrices of genetic distances were estimated based on band presence/absence. These matrices were subsequently utilized in cluster analysis and principal coordinate analysis. Both isozymes and RAPDs were proficient at distinguishing between the 24 accessions. The correspondence between the elements of both distance matrices was moderate (r = 0.36**). Nevertheless, the overall representation of relationships among accessions by cluster analysis and ordination was in considerable agreement. The two techniques contrasted most notably in pair-by-pair comparisons of relationships. RAPD analysis resulted in a more definitive separation of clusters of accessions. The most significant impact of the DNA-based markers probably will be the more accurate determination of relationships between accessions that are too close to be accurately differentiated by isozymes.

Hill, B. D. and E. H. Stobbe (1978). "Effect of Light and Nutrient Levels on C-14 Benzoylprop Ethyl Metabolism and Growth- Inhibition in Wild Oat (Avena-Fatua L)." Weed Research 18(4): 223-229.://A1978FM87600006

Hill, B. D., B. G. Todd, et al. (1980). "Effect of 2,4-D on the Hydrolysis of Diclofop-Methyl in Wild Oat (Avena-Fatua)." Weed Science 28(6): 725-729.://A1980KT70200026

Hilton, J. R. (1984). "The Influence of Light and Potassium-Nitrate on the Dormancy and Germination of Avena-Fatua L (Wild Oat) Seed and Its Ecological Significance." New Phytologist 96(1): 31-34.://A1984SB46600005

Hilton, J. R. (1985). "The Influence of Light and Potassium-Nitrate on the Dormancy and Germination of Avena-Fatua L (Wild Oat) Seed Stored Buried under Natural Conditions." Journal of Experimental Botany 36(167): 974-979.://A1985AJZ4400013

Hilton, J. R. and C. J. Bitterli (1983). "The Influence of Light on the Germination of Avena-Fatua L (Wild Oat) Seed and Its Ecological Significance." New Phytologist 95(2): 325-333.://A1983RM70800018

Holm, F. A., K. J. Kirkland, et al. (2000). "Defining optimum herbicide rates and timing for wild oat (Avena fatua) control in spring wheat (Triticum aestivum)." Weed Technology 14(1): 167-175.://000086220300025 Knowledge of optimal combinations of graminicide rate and stage of application could improve the effectiveness and net benefit of commonly used graminicides. A study was conducted at two locations in Saskatchewan, Canada, from 1994 to 1997. Factorial combinations of five graminicides (CGA 184927, fenoxaprop-p-ethyl, ICIA 0604, imazamethabenz, and flamprop-methyl), three graminicide rates (full, two-thirds, and one-third recommended label rate), and three leaf stages of wild oat (Avena fatua; two-, four-, and six-leaf) were compared to determine their effect on wild oat fresh weight, wheat (Triticum aestivum) seed yield, and net return. Wild oat fresh weight increased and wheat seed yield decreased to a greater extent at Saskatoon (median wild oat fresh weight of 56 g/m(2)) than at Scott (median wild oat fresh weight of 85 g/m(2)) when graminicide rate was reduced from the recommended label rate. Net return consistently decreased at both locations and among all graminicides when application rate was reduced from two-thirds to one-third of the recommended label rate. Imazamethabenz applied at progressively later growth stages caused greater wild oat fresh weight at both locations and reduced wheat yield and net return. Applying other graminicides at the earliest (two-leaf) stage of wild oat generally resulted in more or similar levels of wild oat fresh weight compared with delayed applications, especially at Saskatoon. With the exception of imazamethabenz, crop yield and net return were unaffected by leaf stage at both locations. The optimal graminicide rate is mostly dependent on the level of wild oat infestation, and the best time to control wild oat is dependent mostly on the particular graminicide.

Holroyd, J. and A. G. Strickland (1978). "Roguing Wild Oats." Weed Research 18(3): 175-180.://A1978FK07700008

Hooley, R. (1982). "Protoplasts Isolated from Aleurone Layers of Wild Oat (Avena-Fatua L) Exhibit the Classic Response to Gibberellic-Acid." Planta 154(1): 29-40.://A1982NJ89600005

Hooley, R. (1984). "Gibberellic-Acid Controls Specific Acid-Phosphatase Isozymes in Aleurone Cells and Protoplasts of Avena- Fatua L." Planta 161(4): 355-360.://A1984TA96700010

Hooley, R. (1984). "Gibberellic-Acid Controls the Secretion of Acid-Phosphatase in Aleurone Layers and Isolated Aleurone Protoplasts of Avena-Fatua." Journal of Experimental Botany 35(155): 822-828.://A1984SX29100007

Hooley, R. (1992). "The Responsiveness of Avena-Fatua Aleurone Protoplasts to Gibberellic-Acid." Plant Growth Regulation 11(1): 85-89.://A1992HF24500014 The sensitivity to gibberellic acid (GA3) of aleurone protoplasts isolated from a single harvest of an inbred line of Avena fatua seed that had been after-ripened over anhydrous CaCl2 at 25 +/- 2-degrees-C and 4 +/- 2-degrees-C for three years was assessed. Protoplasts isolated from aleurones of seed stored at 25-degrees-C produced substantially more alpha- amylase in response to 10(-7) M GA3 than those isolated from aleurones of seed stored at 4-degrees-C. The apparent difference in responsiveness does not appear to be due to a change in the duration of the lag phase between addition of 26

GA3 and the production of alpha-amylase. The dose response of aleurone protoplasts to GA3, measured as alpha- amylase production, is complex and appears to have three phases. Protoplasts from seed stored at both temperatures respond appreciably to 10(-14) M GA3. With increasing concentrations of GA3, up to 10(-9) M, alpha-amylase production increases similarly in protoplasts from both lots of seed, reaching a level approximately 2.7-3.8 times greater than when no GA3 is applied. GA3-induced alpha-amylase production increases markedly as the concentration is raised from 10(-9) M to 10(-6) M, and the response then appears to be saturated. Over this part of the response curve protoplasts from the two seed lots differ markedly in their responsiveness to GA3. Those from seed stored at 25-degrees-C produce considerably more alpha-amylase, > 130-fold higher than the minus GA3 control, than those from seed stored at 4-degrees-C, < 35-fold higher than the minus GA3 Control. This apparent difference in the responsiveness of aleurone protoplasts to GA3 could be correlated with the loss of embryo dormancy in seed stored at 25-degrees-C. Seed stored at 4-degrees-C retained the dormancy characteristics present immediately after harvesting.

Hooley, R., M. H. Beale, et al. (1991). "Gibberellin Perception at the Plasma-Membrane of Avena-Fatua Aleurone Protoplasts." Planta 183(2): 274-280.://A1991ET35700019 A functional assay for gibberellin (GA) receptors is described based on the induction of alpha-amylase gene expression in isolated aleurone protoplasts of Avena fatua L. by GA4 immobilised to Sepharose beads. A 17-thiol derivative of GA4, shown to be biologically active with aleurone protoplasts, has been coupled to epoxy-activated Sepharose 6B. This GA4- 17-Sepharose induces high levels of alpha-amylase when incubated with isolated aleurone protoplasts, while cells of the intact aleurone layer do not respond appreciably to the immobilised GA4. In order to eliminate the possibility that GA4 may be released from the Sepharose when incubated with protoplasts, aleurone layers and isolated aleurone protoplasts have been co-incubated, and their responses to GA4, GA4-17-Sepharose and control Sepharose estimated by determining the relative amounts of alpha-amylase mRNA induced in each tissue. Evidence from these experiments is consistent with the view that GA4-17-Sepharose induces alpha-amylase gene expression in aleurone protoplasts by interacting with the protoplast surface. This indicates that GA receptors may be located at, or near, the external face of the aleurone plasma membrane.

Hooley, R., M. H. Beale, et al. (1992). "Gibberellin Perception and the Avena-Fatua Aleurone - Do Our Molecular Keys Fit the Correct Locks." Biochemical Society Transactions 20(1): 85-89.://A1992HG45300017

Hooley, R., S. J. Smith, et al. (1993). "In-Vivo Photoaffinity-Labeling of Gibberellin-Binding Proteins in Avena-Fatua Aleurone." Australian Journal of Plant Physiology 20(4-5): 573-584.://A1993MC82100014 It is generally accepted that specific recognition between plant hormones and proteins involved in their biosynthesis, metabolism, transport and perception are of profound importance in the hormonal regulation of plant growth and development. The identification and detailed characterisation of hormone-binding proteins which perform these functions is an important component of research that aims to understand plant hormone action. In this report the development of photoaffinity reagents for gibberellin (GA)-binding proteins is reviewed, and their use as probes with which to identify and characterise GA-binding proteins in Avena fatua aleurone is described. In vivo GA-photoaffinity labelling of aleurone layers, using the new photoaffinity probe GA(4)-17-sulfoxyethyl-p-azido-[I-125] iodosalicylate, leads to the covalent attachment of this reagent to numerous aleurone polypeptides. Biologically active and inactive GAs used as competitors during in vivo GA-photoaffinity labelling help discriminate between specific and non-specific binding. The biologically active GA(4) and GA(4)-17-yl-1'-(1'-thia)propan-3'-ol-4-azido-5-iodosalicylate compete for photoaffinity labelling of a 60 kDa aleurone polypeptide, while the inactive GA(8) does not. These GA-photoaffinity labelling characteristics suggest that the 60 kDa aleurone polypeptide may interact specifically with active GAs. This is the first report to identify a specific GA- binding protein in aleurone. We suggest that this specific interaction observed in vivo, under conditions where the aleurone cells are responding to the GA-photoaffinity probe by expressing alpha-amylase genes, may be of significance in the perception and action of GA in this tissue.

Hoppe, H. D. and W. Pohler (1989). "Hybrids between Avena-Barbata and Avena-Macrostachya." Cereal Research Communications 17(2): 129-134.://A1989AR76300005

Horvath, D. P., R. Schaffer, et al. (2003). "Identification of genes induced in emerging tillers of wild oat (Avena fatua) using Arabidopsis microarrays." Weed Science 51(4): 503-508.://000184390400004 Arabidopsis complementary DNA (cDNA) microarrays were hybridized with labeled cDNA from mature leaves and emerging tillers of wild oat to determine if they could identify gene expression profiles in distantly related species. More than 23% of the > 11,000 cDNAs on the array hybridized to the wild oat probe. Transcription patterns detected by hybridization to the arrays are indicators for physiological processes in the tissues tested. Coordinated expression patterns for these genes in Arabidopsis indicate common signals involved in their regulation. The results demonstrate that probing cDNA-based arrays from well-characterized species can provide valuable insight into the signal transduction processes regulating growth and development of poorly characterized species.

Hou, J. Q., E. J. Kendall, et al. (1997). "Water uptake and distribution in non-dormant and dormant wild oat (Avena fatua L) caryopses." Journal of Experimental Botany 48(308): 683-692.://A1997XA41700006 The influence of seed coat modification and light quality on water uptake and distribution in caryopses of dormant and non- dormant lines of wild oat (Avena fatua L.) was determined using NMR microimaging. Non-dormant seeds absorbed water more rapidly than dormant seeds during imbibition on distilled water. This effect was detected first in the embryo-scutellar region (8 h) and later in the proximal endosperm (12 h). Cutting the testa and pericarp close to the embryo or scarification with KOH promoted rapid embryo/scutellum hydration and germination. Cutting at the middle part of the caryopsis did not enhance embryo hydration nor did it greatly improve germination. The sensitivity of water distribution to the phytochrome germination effect was examined. Significant differences in imbibitional water uptake by embryo-scutellum tissue were detected by 18 h following red-light (germination promoter) compared with far-red (germination inhibitor) treatment. The results indicated that both the rate and the sequence of embryo/scutellum hydration were important in initiating 27

germination in dormant seeds. A refinement of the model that describes water imbibition in wild oat seeds during the early stages of germination is discussed.

Hou, J. Q. and G. M. Simpson (1990). "Phytochrome Action and Water Status in Seed-Germination of Wild Oats (Avena-Fatua)." Canadian Journal of Botany-Revue Canadienne De Botanique 68(8): 1722-1727.://A1990DZ76000014

Hou, J. Q. and G. M. Simpson (1991). "Effects of Prolonged Light on Germination of 6 Lines of Wild Oat (Avena-Fatua)." Canadian Journal of Botany-Revue Canadienne De Botanique 69(7): 1414-1417.://A1991GE79600004 Effects of prolonged light irradiation on seed germination of wild oat (Avena fatua L.) were studied in three nondormant and three dormant genetic lines. Light responses were observed in each of the lines tested. The expression of the light reaction is related to the genetic variability and dormancy states of the seeds. The light reaction can be observed in the dormant lines by removing the dormancy blocks in the seeds either through afterripening or by treatment with gibberellic acid or azide. Prolonged far-red, blue, and white light are inhibitory to germination. Prolonged red light had neutral, or inhibitory, effects compared with the corresponding dark germination. Germination responses to light depend on both the condition of phytochrome established by the light environment and the state of dormancy in wild oat seeds.

Hou, J. Q. and G. M. Simpson (1992). "After-Ripening and Phytochrome Action in Seeds of Dormant Lines of Wild Oat (Avena- Fatua)." Physiologia Plantarum 86(3): 427-432.://A1992KA47900012 The effects of a short exposure to red, far-red or alternate red/far-red light on the germination of seeds after-ripened for different periods of time were studied in dormant lines of wild oat (Avena fatua L.). Three stages were distinguishable in the after-ripening period in the response of germination to light. Seeds stayed dormant and showed no response to light during stage I. Phytochrome-mediated germination was observed in seeds during stage II. The phytochrome action disappeared during stage III, i.e. seeds fully germinated following treatments of all light qualities. When the seeds were imbibed in polyethylene glycol solutions, dark germination was reduced and phytochrome again had an effect, which suggested the involvement of phytochrome in water uptake of the seed.

Hou, J. Q. and G. M. Simpson (1993). "Germination Response to Phytochrome Depends on Specific Dormancy States in Wild Oat (Avena-Fatua)." Canadian Journal of Botany-Revue Canadienne De Botanique 71(11): 1528-1532.://A1993MQ64700018 Effects of brief red and far-red light on germination of seeds from dormant lines of wild oat (Avena fatua L.) were studied in combination with mechanical injury to the seed coat, application of gibberellin A,, or changes in relative humidity during afterripening. Aberrant germination responses to phytochrome action were observed in the mechanically injured seeds in some of the lines, i.e., brief red light inhibited or delayed germination induced by injury, and immediately following far-red light cancelled the negative effects. Phytochrome action influenced germination of the gibberellin-treated seeds in a normal fashion, although effects of the gibberellic acid and brief red light on germination were not additive. Brief red light inhibited germination of seeds afterripened in zero relative humidity; the same light promoted germination of those in 30 and 60% relative humidity. Germination response to phytochrome in wild oat depends on specific seed dormancy states, illustrated by genetic origins, dormancy-breaking methods and afterripening conditions.

Hou, J. Q. and G. M. Simpson (1994). "Effects of Immersing Dry Seeds in Alkaline-Solutions on Seed Dormancy and Water-Uptake in Wild Oat (Avena-Fatua)." Canadian Journal of Plant Science 74(1): 19-24.://A1994NB31300004 Effects of immersing dry seeds in KOH and NaOH solutions on seed dormancy and water uptake were studied in three dormant lines of wild oat (Avena fatua L.). KOH was more effective than NaOH in breaking dormancy. Maximum dormancy-breaking effect of 5.3 N KOH could be achieved with a 10- or 15-min treatment. Increase in treatment time did not necessarily increase germination; rather, it caused damage to the seeds. For 10-min treatment, 5.3 and 7.6 N KOH solutions were more effective than 3 and 9.8 N. Genetic lines responded differently to the KOH treatment. Initial rate and amount of water uptake by KOH-treated seed were significantly higher than by the untreated. It is believed that breaking dormancy by the alkaline treatment is related to removing the barrier to water uptake formed by the seed coat.

Hsiao, A. I. (1979). "Effect of Sodium-Hypochlorite and Gibberellic-Acid on Seed Dormancy and Germination of Wild Oats (Avena- Fatua)." Canadian Journal of Botany-Revue Canadienne De Botanique 57(16): 1729-1734.://A1979HP22900008

Hsiao, A. I. (1979). "Factors Affecting the Sodium-Hypochlorite Seed Viability Test for Wild Oats (Avena-Fatua L)." Plant Physiology 63(5): 68-68.://A1979GX47600379

Hsiao, A. I., M. E. Macgregor, et al. (1979). "Use of Sodium-Hypochlorite in Testing the Seed Viability of Wild Oats." Canadian Journal of Plant Science 59(4): 1047-1052.://A1979HY43200019

Hsiao, A. I. and G. I. McIntyre (1988). "Induction of Vivipary in Avena-Fatua." Physiologia Plantarum 73(1): 128-133.://A1988N771200020

Hsiao, A. I., G. I. McIntyre, et al. (1983). "Seed Dormancy in Avena-Fatua .1. Induction of Germination by Mechanical Injury." Botanical Gazette 144(2): 217-222.://A1983RG06200008

Hsiao, A. I. and W. A. Quick (1984). "Actions of Sodium-Hypochlorite and Hydrogen-Peroxide on Seed Dormancy and Germination of Wild Oats, Avena-Fatua L." Weed Research 24(6): 411-419.://A1984TU08200005

Hsiao, A. I. and W. A. Quick (1985). "Wild Oats (Avena-Fatua-L) Seed Dormancy as Influenced by Sodium-Hypochlorite, Moist Storage and Gibberellin-A3." Weed Research 25(4): 281-288.://A1985ANJ0300007

Hsiao, A. I. H. and G. M. Simpson (1971). "Dormancy Studies in Seed of Avena-Fatua .7. Effects of Light and Variation in Water Regime on Germination." Canadian Journal of Botany 49(8): 1347-&.://A1971K273200013

28

Hunter, D. E., J. P. Murphy, et al. (1995). "Isozyme Variation in Avena-Sterilis L Collected in Turkey." Crop Science 35(5): 1477- 1482.://A1995RR62100035 The National Small Grains (NSG) collection of Avena sterilis L. is an important germplasm resource for cultivated oat (A. sativa L. and A. byzantina C. Koch) improvement. Complete evaluation and utilization of the almost 7000 accessions will be protracted. In the interim, new accessions will he acquired. Previous studies showed that Turkish accessions of A. sterilis exhibited a broad array of isozyme phenotypes found in the wild collection and in over 400 cultivated accessions. The objectives of this research were (i) to determine if recent Turkish acquisitions of A. sterilis (Forsberg-Simons collections of 1986) have contributed additional isozyme variation to that already in storage and (ii) to illustrate relationships among Turkish A. sterilis and cultivated germplasm. Zymogram phenotypes of 341 A. sterilis and 405 cultivated samples were recorded for 13 enzyme systems. The 115-plant sample from the Forsberg-Simons collection included six phenotypes not previously observed in Turkish accessions in the NSG collection. Cluster analysis suggested that approximately 40% of the Forsberg-Simons acquisitions differed from much of the existing NSG Turkish accessions. Both local and regional diversity occurred within A. sterilis germplasm. Isozyme patterns in cultivated germplasm were most similar to collections of A. sterilis with a wide geographical distribution.

Huskins, C. L. (1927). "On the genetics and cytology of fatuoid or false wild oats." Journal of Genetics 18(3): 315-U15.://000200616200002

Huskins, C. L. (1932). "Fatuoids or false wild oats." Nature 130: 132-133.://000188238100079

Hutchinson, E. S. (1984). "Seed Size and Quantitative Characters in Avena-Barbata." Heredity 52(FEB): 25-33.://A1984SF88000004

Hutchinson, E. S., A. Hakimelahi, et al. (1983). "The Genetics of the Diploidized Tetraploid Avena-Barbata - Acid-Phosphatase, Esterase, Leucine Aminopeptidase, Peroxidase, and 6-Phosphogluconate Dehydrogenase Loci." Journal of Heredity 74(5): 325- 330.://A1983RH34800003

Hutchinson, E. S., S. C. Price, et al. (1983). "An Experimental-Verification of Segregation Theory in a Diploidized Tetraploid - Esterase Loci in Avena-Barbata." Journal of Heredity 74(5): 381-383.://A1983RH34800016

Imam, A. G. and R. W. Allard (1965). "Population Studies in Predominantly Self-Pollinated Species .6. Genetic Variability between and within Natural Populations of Wild Oats from Differing Habitats in California." Genetics 51(1): 49-&.://A19656051100005

Ivens, G. W. (1978). "Weeds of Note .1. Wild Oats." New Zealand Journal of Agriculture 136(8): 20-21.://A1978FL89400007

Iwig, M. M. and H. W. Ohm (1976). "Genetic-Control of Protein from Avena-Sterilis L." Crop Science 16(6): 749-752.://A1976CT06500001

Jain, J. C., W. A. Quick, et al. (1982). "Studies of Acid-Soluble Phosphorus-Compounds in Genetically Pure Lines of Avena-Fatua with Different Dormancy Characteristics." Canadian Journal of Botany-Revue Canadienne De Botanique 60(10): 2099-2104.://A1982PP88500026

Jain, J. C., W. A. Quick, et al. (1983). "Atp Synthesis During Water Imbibition in Caryopses of Genetically Dormant and Non- Dormant Lines of Wild Oat (Avena-Fatua L)." Journal of Experimental Botany 34(141): 381-387.://A1983QL71800002

Jain, R. and W. H. Vandenborn (1989). "Morphological and Histological Effects of 3 Grass Selective Herbicides on Developing Wild Oat (Avena-Fatua) Stems." Weed Science 37(4): 575-584.://A1989AR21200015

Jain, S. K. and D. R. Marshall (1967). "Population Studies in Predominantly Self-Pollinating Species .X. Variation in Natural Populations of Avena Fatua and a Barbata." American Naturalist 101(917): 19-&.://A19678988600002

Jain, S. K. and D. R. Marshall (1970). "Within-Family Selection in Avena-Fatua and a-Barbata." Theoretical and Applied Genetics 40(2): 73-&.://A1970G702600006

Jain, S. K. and K. N. Rai (1974). "Population Biology of Avena .4. Polymorphism in Small Populations of Avena-Fatua." Theoretical and Applied Genetics 44(1): 7-11.://A1974R655400002

Jain, S. K. and K. N. Rai (1977). "Natural-Selection During Germination in Wild Oat (Avena-Fatua) and California Burclover (Medicago-Polymorpha Var-Vulgaris) Populations." Weed Science 25(6): 495-498.://A1977ED90400007

Jain, S. K., K. N. Rai, et al. (1981). "Population Biology of Avena .11. Variation in Peripheral Isolates of Avena-Barbata." Genetica 56(3): 213-215.://A1981MS90900008

Jana, S., S. N. Acharya, et al. (1979). "Dormancy Studies in Seed of Avena-Fatua .10. Inheritance of Germination Behavior." Canadian Journal of Botany-Revue Canadienne De Botanique 57(15): 1663-1667.://A1979HN23300009

Jana, S. and J. M. Naylor (1980). "Dormancy Studies in Seed of Avena-Fatua .11. Heritability for Seed Dormancy." Canadian Journal of Botany-Revue Canadienne De Botanique 58(1): 91-93.://A1980JH18100011

29

Jana, S. and J. M. Naylor (1982). "Adaptation for Herbicide Tolerance in Populations of Avena-Fatua." Canadian Journal of Botany- Revue Canadienne De Botanique 60(9): 1611-1617.://A1982PK90400006

Jana, S. and K. M. Thai (1987). "Patterns of Changes of Dormant Genotypes in Avena-Fatua Populations under Different Agricultural Conditions." Canadian Journal of Botany-Revue Canadienne De Botanique 65(8): 1741-1745.://A1987J816600025

Jana, S., M. K. Upadhyaya, et al. (1988). "Genetic-Basis of Dormancy and Differential Response to Sodium-Azide in Avena-Fatua Seeds." Canadian Journal of Botany-Revue Canadienne De Botanique 66(4): 635-641.://A1988N187500004

Jeffcoat, B. and W. N. Harries (1975). "Selectivity and Mode of Action of Flamprop-Isopropyl, Isopropyl (+/-)-2-[N-(3-Chloro-4- Fluorophenyl)Benzamido]Propionate, in Control of Avena-Fatua in Barley." Pesticide Science 6(3): 283-296.://A1975AV50000009

Jensen, K. I. N. and J. C. Caseley (1990). "Antagonistic Effects of 2,4-D Amine and Bentazone on Control of Avena-Fatua with Tralkoxydim." Weed Research 30(6): 389-395.://A1990EJ99700002

Johansen-Morris, A. D. and R. G. Latta (2006). "Fitness consequences of hybridization between ecotypes of Avena barbata: Hybrid breakdown, hybrid vigor, and transgressive segregation." Evolution 60(8): 1585-1595.://000240530300006 Hybridization is an important factor in the evolution of plants; however, many of the studies that have examined hybrid fitness have been concerned with the study of early generation hybrids. We examined the early- and late-generation fitness consequences of hybridization between two ecotypes of the selfing annual Avena barbata in a greenhouse environment as well as in two natural environments. Fitness of early generation (F-2) hybrids reflects both the action of dominance effects (hybrid vigor) and recombination (hybrid breakdown) and was not significantly different from that of the midparent in any environment. Fitness of later generation (F-6) recombinant inbred lines (RILS) derived from the cross reflect both the loss of early generation heterozygosity as well as disruption of any coadapted gene complexes present in the parents. In all environments, F-6 RILs were on average significantly less fit than the (equally homozygous) midparent, indicating hybrid breakdown through the disruption of epistatic interactions. However, the inbred F-6 were also less fit than the heterozygous F-2, indicating that hybrid vigor also occurs in A. barbata, and counteracts hybrid breakdown in early generation hybrids. Also, although the F-6 generation mean is lower than the midparent mean, there are individual genotypes within the F-6 generation that are capable of outperforming the parental ecotypes in the greenhouse. Fewer hybrid genotypes are capable of outperforming the parental ecotypes in the field. Overall, these experiments demonstrate how a single hybridization event can result in a number of outcomes including hybrid vigor, hybrid breakdown, and transgressive segregation, which interact to determine long-term hybrid fitness.

Johnson, C. D. (1998). "'Transcendental Wild Oats' or the cost of an idea (Louisa May Alcott fiction)." American Transcendental Quarterly 12(1): 45-65.://000072972100003

Johnson, L. D. R. and P. G. Rothman (1986). "Resistance to Stem Rust in Avena-Fatua." Phytopathology 76(10): 1147-1147.://A1986F034600712

Johnson, R. R., M. E. Chaverra, et al. (1996). "Degradation of wild oat (Avena fatua L) avenin and protein Z mRNAs during seed development, after ripening, and germination." Plant Physiology 111(2): 115-115.://A1996UR53400159

Johnson, R. R., H. J. Cranston, et al. (1995). "Characterization of Cdna Clones for Differentially Expressed Genes in Embryos of Dormant and Nondormant Avena-Fatua L Caryopses." Plant Molecular Biology 28(1): 113-122.://A1995RF03600011 The molecular regulation of seed dormancy was investigated using differential display to visualize and isolate cDNAs representing differentially expressed genes during early imbibition of dormant and nondormant Avena fatua L. embryos. Of about 3000 cDNA bands examined, 5 cDNAs hybridized with mRNAs exhibiting dormancy-associated expression patterns during the first 48 h of imbibition, while many more nondormancy-associated cDNAs were observed. Dormancy-associated clone AFD1 hybridized with a 1.5 kb mRNA barely detectable in dry dormant and nondormant embryos that became more abundant in dormant embryos after 24 h of imbibition. Clone AFD2 hybridized with two mRNAs, a 1.3 kb message constitutively expressed in dormant and nondormant embryos and a 0.9 kb message present at higher levels in dormant embryos after 3 h of imbibition. Nondormancy-associated clones AFN1, AFN2 and AFN3 hybridized with 1.5 kb, 1.7 kb and 1.1 kb mRNAs, respectively, that were more abundant in nondormant embryos during imbibition. Expression patterns of some mRNAs in dormant embryos induced to germinate by GA(3) treatment were different than water controls, but were not identic al to those observed in nondormant embryos. DNA sequence analysis revealed 76% sequence identity between clone AFN3 and a Citrus sinensis glutathione peroxidase-like cDNA, while significant sequence similarities with known genes were not found for other clones. Southern hybridization analyses showed that all clones represent low (1 to 4) copy number genes.

Johnston, M. R., L. D. Carsten, et al. (2000). "Epidemic development and virulence in 1995-1998 of Puccinia coronata, a potential biocontrol agent of wild oats on San Clemente Island." Biological Control 17(3): 250-257.://000085743500006 The biology of islands differs from that of large land masses in having less complex ecosystems. Introduced exotic weeds are often a major threat to fragile island ecosystems because of their expansion into habitats previously occupied by endemic species. San Clemente Island, 120 km off the California coastline, is an example of this process in which numerous exotic weed species have been introduced and some are endangering the native flora. Crown rust of oats caused by Puccinia coronata f.sp, avenae was investigated as a potential biocontrol agent against two wild oat species, Avena barbata and Avena fatua, introduced on San Clemente Island. Epidemiology and virulence of this rust were studied. The island was surveyed from 1995 to 1998 for occurrence of P. coronata on wild oats. Wild oats were found sprouting in the northern part of the island shortly after autumn rainfall and subsequently covered the main grasslands of the island. The rust also appeared first on the northern part of the island and progressively spread south. However, disease severities 30

in the south were considerably lower than those in the north. Diverse virulence types, although related to Californian and Mexican forms, were detected among the isolates. The potential use of P. coronata as an augmentative biocontrol agent for wild oat species on San Clemente Island is discussed. (C) 2000 Academic Press.

Jones, E. T. (1932). "Fatuoids or false wild oats." Nature 129: 617-617.://000188272600323

Jones, H. D., S. Kurup, et al. (2000). "Identification and analysis of proteins that interact with the Avena fatua homologue of the maize transcription factor VIVIPAROUS 1." Plant Journal 21(2): 133-142.://000085607300002 The Avena fatua (wild oat) homologue of VIVIPAROUS 1 (AfVP1) has been implicated in controlling the maintenance of embryo dormancy in mature imbibed seeds, but the detailed mechanisms by which this transcription factor family activates embryo maturation pathways and simultaneously represses germination are not known. A two-hybrid screen in yeast identified three proteins that interacted specifically with AfVP1 (AfVP1 interacting proteins; AfVIPs). AfVIPs 2 and 3 interacted with the C-terminus of AfVP1, which contains the B2 + B3 domains, previously shown to bind DNA, whereas AfVIP1 interacted with the isolated B3 domain. Using purified proteins in in vitro experiments, all three AfVIPs were shown also to interact with the Arabidopsis homologue ABSCISIC ACID INSENSITIVE 3 (ABI3). The three AfVIPs were expressed in both dormant and non-dormant embryos, but the abundance of AfVIP1 and 3 transcripts was greater in germinated than dormant seeds, whereas transcripts of AfVIP2 (and AfVP1) were more highly expressed in dormant embryos. The AfVIP3 protein has homology to a human cell-crisis gene with a predicted role in the cell cycle; AfVIP2 contains a ring-type zinc finger motif. These homologies, together with analysis of expression studies, suggest that these proteins may play specific roles in AfVP1-mediated regulation of the dormancy to germination transition in A. fatua seeds.

Jones, H. D., N. C. B. Peters, et al. (1997). "Genotype and environment interact to central dormancy and differential expression of the VIVIPAROUS 1 homologue in embryos of Avena fatua." Plant Journal 12(4): 911-920.://A1997YF45400017 Embryo dormancy is a reversible developmental state during which germination is repressed. In this study, inbred lines of Avena fatua were used to analyse the influence of genotype and environment on the dormant phenotype, and on expression of the homologue of the maize transcription factor VIVIPAROUS 1 (afVP 1). The cDNA for afVP 1 was cloned from mature embryos. Analysis of the predicted protein sequence revealed a high degree of similarity to other VP 1/ABI 3- related transcription factors, in particular in four regions previously shown to be highly conserved, including the BR2 region that has been shown to interact with several classes of sequence-specific DNA binding proteins. The potential of imbibed mature embryos for dormancy was analysed and shown to be determined primarily by genotype and secondarily by previous environmental experience of the mature seed acting on embryo genotype. Under all conditions studied, expression of afVP 1 and the A. fatua homologue of Pm (shown in maize to be regulated by VP 1 during embryo maturation) were positively correlated with the dormant phenotype, whereas expression of A. fatua AMY-related RNAs was negatively correlated with dormancy (in barley AMY 6-4 has been shown to be repressed by VP 1). Expression of afVP 1 RNA was also shown in the dry seed to be positively correlated with the length of time required for seeds of the inbred lines to after-ripen. These results suggest new functions for the VP 1 transcription factor family in the control of dormancy-related processes in embryo cells of mature seeds, and the up-regulation of afVP 1 and afEm RNAs in the dormant state suggests that they are regulated by a switching mechanism in the mature seed that shows some aspects of reversibility.

Jones, R. and R. Medd (1997). "Economic analysis of integrated management of wild oats involving fallow, herbicide and crop rotational options." Australian Journal of Experimental Agriculture 37(6): 683-691.://A1997YB43500010 The primary objective of this study was to estimate the economic benefits associated with an integrated weed management approach for wild oats (Avena fatua and A. ludoviciana) in northern New South Wales involving chemical and non-chemical controls. The paper presents a framework for assessing the population dynamics of wild oats and the economics of a range of control options over 15 years. Wild oats is a weed primarily of winter crops which, as a consequence of its persistence and its impact upon yields, leads to significant economic losses in the grain growing regions of Australia. In this study, a dynamic programming model is developed to examine the impact of a range of management strategies for the control of wild oats in wheat. The strategies evaluated include conventional herbicide control to reduce weed densities, the use of selective herbicides to reduce seed set of the weed, and summer crop and winter fallow rotational options which provide a break in the cereal cycle and allow accelerated control of wild oat populations. The hypothesis for the study, that strategies which involve measures that directly reduce seed production and minimise wild oats seed bank populations will yield the greatest economic benefit, is acceptable based on the findings of the study. The work also shows that a dynamic programming model provides a means of determining the optimal combination of strategies over time for various initial values of the seed bank. The methodology is considered to have general application as a framework for evaluating the economics of weed control problems in annual cropping systems.

Jones, R. B. (1986). "Editorial Wild-Oats - Carmack,Edward,Ward and Tennessee Politics - Majors,Wr." Appalachian Journal 13(3): 309-311.://A1986C817200013

Joseph, O. O., S. L. A. Hobbs, et al. (1990). "Diclofop Resistance in Wild Oat (Avena-Fatua)." Weed Science 38(6): 475-479.://A1990EN88400005 The differences in tolerance, morphology, and physiological response of diclofop-resistant and -susceptible wild oat biotypes collected from fields in Saskatchewan, Canada, were investigated under growth room and field conditions. Under herbicide-free conditions the resistant biotype had more upright leaves with about 12% less leaf area and 50% less leaf width than the susceptible biotype. A marked difference in the level of tolerance to diclofop was observed. Photosynthesis was initially significantly reduced in both biotypes after treatment with diclofop at the rate of 0.70 kg ai ha-1, but the resistant biotype was able to recover. Injury to the susceptible biotype was reduced by coating the seeds with 1,8- naphthalic anhydride. Differential foliar retention was not an important factor in selectivity of diclofop among the two biotypes.

31

Kafiz, B., J. P. Caussanel, et al. (1989). "Interaction between Diclofop-Methyl and 2,4-D in Wild Oat (Avena-Fatua L) and Cultivated Oat (Avena-Sativa L), and Fate of Diclofop-Methyl in Cultivated Oat." Weed Research 29(4): 299-305.://A1989AH38800009

Kahler, A. L., R. W. Allard, et al. (1980). "Associations between Isoenzyme Phenotypes and Environment in the Slender Wild Oat (Avena-Barbata) in Israel." Theoretical and Applied Genetics 56(1-2): 31-47.://A1980HZ96000006

Kahler, A. L., R. W. Allard, et al. (1976). "Isozyme Phenotype Environment Associations in Natural-Populations of Avena-Barbata in Israel." Genetics 83(3): S37-S38.://A1976BZ03400107

Kahler, A. L., K. M. Shumaker, et al. (1979). "Hidden Alleles at Enzyme Loci in 3 Plant-Species (Avena-Barbata, Hordeum-Vulgare and Zea-Mays)." Genetics 91(4): S57-S58.://A1979HF94700138

Karlowsky, J. D., A. L. Brule-Babel, et al. (2006). "Inheritance of multiple herbicide resistance in wild oat (Avena fatua L.)." Canadian Journal of Plant Science 86(1): 317-329.://000240051300078 To gain some insight into the surprisingly frequent occurrence of multiple herbicide resistant wild oat in western Canada, the inheritance of multiple herbicide resistance was studied in two wild oat (Avena fatua L.) populations, UMWO12-01 and UMWO12-03, from Manitoba, Canada. Both populations are resistant to each of three distinct herbicides, imazamethabenz-methyl, flamprop-methyl, and fenoxaprop-p-ethyl (hereafter referred to as imazamethabenz, flamprop, and fenoxaprop-P, respectively). Crosses were made between each resistant (R) population and a susceptible (S) wild oat population (UM5) (R/S crosses), and between the two resistant populations (R/R crosses). Subsets of parental, F-2 plants, and F-2-derived F-3 (F-2:3) families were treated separately with each of the three herbicides and classified as R or S for individual plants, and homozygous R, segregating, or homozygous S for F-2:3 families. F-2 plants and F-2,F-3 families from R/S crosses segregated in 3R:1S and I homozygous R:2 segregating:1 homozygous S ratios, respectively. These ratios indicate that a single dominant or semi-dominant nuclear gene controls resistance to each of these herbicides in each population. F-2 plants and F-2:3 families from R/R crosses segregated for resistance/susceptibility when treated with either imazamethabenz or flamprop. Therefore, the genes for resistance to these two herbicides are different in each R population. Individual F-2:3 family response demonstrated that the genes were not independent of each other, indicating possible linkage between the genes for resistance to each herbicide. Genetic linkage could explain how the wild oat populations developed multiple resistance in the absence of selection by two of the herbicides, imazamethabenz and flamprop.

Keng, J. P. (1991). "Effect of Afterripening on the Status of Protein and Nonstructural Carbohydrate in Dormant Avena-Fatua Caryopses." Journal of the Agricultural Association of China(156): 24-32.://A1991JA81200004 Darmant Avena fatua L. caryopses were afterripened for periods ranging from 0 to 24 months. The rate and final percentage germination of these caryopses increased in proportion to the duration of afterripening. SDS-polyacrylamide gel electrophoresis revealed that a 60 and 30 kD soluble endosperm protein appeared and disappeared, respectively, during afterripening. There were no apparent qualitative changes during the period of afterripening in soluble and SDS- extractable proteins in the embryo, and in SDS-extractable protein in the endosperm. There were no changes in the quantity of embryo and endosperm protein throughout the period of afterripening.The level of starch decreased by 6-mu- mole per caryopsis and sucrose increased by 0.039-mu-mole per mg in the endosperm after a 14 month period of afterripeing. These results indicate that the termination of dormancy in A. fatua caryopses is related to changes both in protein and in carbohydrate profiles during the process of afterripening.

Keng, J. P. and M. E. Foley (1987). "Effect of Gibberellin on Protein and Nonstructural Carbohydrate in Dormant Avena-Fatua Caryopses." Plant Science 51(1): 37-41.://A1987J893800006

Kepczynski, J., B. Bialecka, et al. (2006). "Regulation of Avena fatua seed germination by smoke solutions, gibberellin A(3) and ethylene." Plant Growth Regulation 49(1): 9-16.://000237796400002 Dormant, intact Avena fatua L. (wild oat) seeds germinate poorly at 20 degrees C. Removing the hulls slightly increased germination. Treatment with smoke solutions increased the germination of both intact seeds and caryopses. Exogenous GA(3), alone or in the presence of smoke solution, increased the germination of caryopses, while ACC shows a tendency to increase germination of caryopses only when applied in combination with smoke solution. Results suggest that GA(3) and ethylene, but not smoke solutions, are involved in the regulation of alpha-amylase activity during germination. However, the participation of smoke solutions in the control of ACC oxidase activity cannot be excluded.

Kern, A. J., C. T. Colliver, et al. (1996). "Characterization of wild oat (Avena fatua L) populations and an inbred line with multiple herbicide resistance." Weed Science 44(4): 847-852.://A1996WA91300014 Repeated use of the pre-emergence herbicide triallate has selected for wild oat populations that are resistant (R)(3) to field use rates. Field collections and an inbred R line were shown in greenhouse and petri dish dose response experiments to be 6- to 20-fold more tolerant to triallate than susceptible (S)(3) lines, R populations and the inbred line were also resistant (8-fold) to the related thiocarbamate herbicide diallate, as well as to the chemically unrelated postemergence herbicide difenzoquat (60-fold), C-14-triallate uptake and translocation patterns were similar between R and S lines for the first 24 h after application. However, translocation of radioactivity was more rapid in S lines than R lines from 24 through 60 h after application. C-14-difenzoquat uptake was the same in R and S lines 12 h after application, but was 10 to 20% higher in R lines than S lines by 24 through 96 h after application. Similarly, translocation of radioactivity after C-14-difenzoquat application was 7 to 14% greater in R than S lines after 12 h, although translocated radioactivity amounts were not significantly different between R and S lines. The relatively minor differences in triallate and difenzoquat uptake and translocation patterns between R and S lines are most likely not of sufficient magnitude to explain the observed resistance levels.

32

Kern, A. J. and W. E. Dyer (1998). "Compartmental analysis of herbicide efflux in susceptible and difenzoquat-resistant Avena fatua L. suspension cells." Pesticide Biochemistry and Physiology 61(1): 27-37.://000075282300004 The cellular localization of difenzoquat was investigated using suspension cell cultures derived from inbred wild oat (Avena fatua L.) lines that were susceptible (S) or resistant (R) to the herbicide. Compartmental analysis of S cell cultures resolved [3-C-14]difenzoquat efflux from three cellular compartments which contained 75, 19, and 6%, respectively, of the total absorbed difenzoquat. Greater than 98% of the efflux from S cells occurred within 420 min. In contrast, compartmental analysis of R cell cultures differentiated only two cellular components, containing 73 and 27% of the absorbed difenzoquat. Further, difenzoquat efflux from R cell cultures occurred more slowly, requiring >2000 min for 98% efflux to be achieved. Washing cell cultures preloaded with [3-C-14]difenzoquat in methanol:chloroform indicated that difenzoquat was preferentially bound in cell wall material of R cell cultures. In whole plants, a greater proportion of total leaf-absorbed [3-C- 14]difenzoquat was present in an insoluble residue in R plants than in S as determined by tissue homogenization and liquid scintillation counting. By 96 h after treatment with [3-C-14]difenzoquat, >90% of the absorbed radiolabel was extractable from S plants, whereas <10% was extractable from R tissues. The results indicate that tight binding of difenzoquat in R cell walls may be responsible for difenzoquat resistance in wild oats. (C) 1998 Academic Press.

Kern, A. J. and W. E. Dyer (1998). "Physiological, biochemical, and genetic studies of triallate resistance in wild oat (Avena fatua L.)." Abstracts of Papers of the American Chemical Society 215: U29-U30.://000072414400092

Kern, A. J., L. L. Jackson, et al. (1997). "Fatty acid and wax biosynthesis in susceptible and triallate-resistant Avena fatua L." Pesticide Science 51(1): 21-26.://A1997XV80300003 The recent characterization of triallate-resistant lines of wild oat (Avena fatua L.) deficient in triallate sulfoxidation provides an experimental system to investigate and differentiate the effects of triallate and triallate sulfoxide on wax and lipid biosynthesis. Greenhouse applications of triallate dramatically reduced epicuticular wax deposition in susceptible (S) but not resistant (R) wild oats. Triallate treatment had no effect on in-vivo concentrations of C-12 to C-26 fatty acids and fatty alcohols in R plants, while elongated fatty acid fractions (C > 18) were significantly reduced in S plants. In contrast, treatment with triallate sulfoxide reduced in-vivo concentrations of elongated fatty acids equally in R and S, supporting the hypothesis that triallate sulfoxide is more inhibitory than triallate towards fatty acid elongases. Although de-novo synthesis of short-chain fatty acids was not affected by triallate or triallate sulfoxide in R or S plants, synthesis of elongated fatty acid fractions was dramatically reduced in S plants by triallate. Fatty acid biosynthesis in R and S plants was equally sensitive to triallate sulfoxide. The results support the idea that in-vivo triallate sulfoxidation is necessary for herbicidal activity, and confirm that reduced rates of triallate sulfoxidation confer resistance in R wild oats.

Kern, A. J., T. M. Myers, et al. (2002). "Two recessive gene inheritance for triallate resistance in Avena fatua L." Journal of Heredity 93(1): 48-50.://000175627300008 Extensive use of the preemergence herbicide triallate over the last three decades has selected for resistant (R) Avena fatua L. populations in several areas of the United States and Canada. R plants are also cross-resistant to the unrelated pyrazolium herbicide difenzoquat. We made reciprocal crosses between inbred R and susceptible (S) lines to determine the genetic basis of triallate resistance. Seeds from parental lines and F-2 populations were treated with soil applications of 0.275, 0.55, or 1.1 kg/ha triallate in the greenhouse and plant heights recorded after 37 days. Surviving F-2 plants were selfed and the resulting F-3 families were screened with 1.1 kg/ha triallate. In the F-2 populations, assortment of S and R phenotypes fit a 15:1 segregation ratio, suggesting that resistance was controlled by the two independently segregating recessive genes TRR1 and TRR2. None of the 912 F-3 progeny from 51 R F-2 individuals was susceptible to triallate treatment, further supporting a two-gene mode of inheritance. There was a possible maternal effect on susceptibility at the highest triallate rate tested.

Kern, A. J., D. M. Peterson, et al. (1996). "Triallate resistance in Avena fatua L is due to reduced herbicide activation." Pesticide Biochemistry and Physiology 56(3): 163-173.://A1996WW63500001 Extensive use of triallate, a preemergence herbicide used for wild oat (Avena fatua L.) control in cereal crops, has selected for resistant (R) wild oat populations. Triallate is thought to be activated via metabolic sulfoxidation to create the more potent triallate sulfoxide. Treatment of R and susceptible (S) wild oat lines with [1-C-14]triallate showed that triallate is metabolized to the same primary endproduct, 2,3,3-trichloropropene sulfonic acid, in both types. However, the rate of triallate metabolism was more than 12-fold slower in R than in S plants. Dose-response studies indicated that although R plants were 6- to 20-fold more resistant than S plants to triallate treatment, both types were equally sensitive to in vitro synthesized triallate sulfoxide. In addition, [1-C-14]triallate sulfoxide was metabolized to the same endproducts and at the same rate in R and S plants. The data indicate that resistance is conferred by a reduced rate of triallate sulfoxidation and represent the first documented case of herbicide resistance in plants conferred by reduced metabolism. (C) 1996 Academic Press.

Kibite, S., K. N. Harker, et al. (1995). "Inheritance of Resistance to Diclofop-Methyl and Fenoxaprop-P-Ethyl in 2 Avena-Sativa X Avena-Fatua Populations." Canadian Journal of Plant Science 75(1): 81-85.://A1995QN02900013 Since the late 1980s, several reports of herbicide-resistant wild oat have raised concern about potential weed control problems in western Canada. This study was conducted to determine the mode of inheritance, number of genes and type of gene action governing herbicide resistance in two Avena sativa x A. fatua crosses. The herbicide-susceptible A. sativa cultivars, Random and Derby, were crossed with a resistant A. fatua genotype, GP-HR-01. Parents, F(2)s and F-2:3 families were tested for their reactions to two post-emergent wild oat herbicides, diclofopmethyl and fenoxaprop-p-ethyl, in the greenhouse. inheritance of resistance to diclofop-methyl and fenoxaprop-p-ethyl was dominant and monogenic in the Derby/GP-HR-01 cross, but was controlled by two dominant complementary genes in the Random/GPHR-01 cross. Resistance to both herbicides appeared to be controlled by the same genes or groups of tightly linked genes.

Kiehn, F. A., R. I. H. McKenzie, et al. (1976). "Inheritance of Resistance to Puccinia-Coronata-Avenae and Its Association with Seed Characteristics in 4 Accessions of Avena-Sterilis." Canadian Journal of Genetics and Cytology 18(4): 717-726.://A1976CZ94200014 33

Kim, H. B. (1974). "Inheritance of Resistance to Puccinia-Coronata Var Avenae in 6 Selections of Avena-Sterilis." Euphytica 23(1): 174-180.://A1974S350200025

Kim, H. B., I. M. Atkins, et al. (1968). "Inheritance of Anther Size and Spikelet Characteristics - Their Interrelationship in Crossing Cultivars Bronco Avena Sativa L and a Sterilis L Ssp Macrocarpa (Moench) Briq." Crop Science 8(1): 101-&.://A1968A775100030

Kirkland, K. J. (1993). "Spring Wheat (Triticum-Aestivum) Growth and Yield as Influenced by Duration of Wild Oat (Avena-Fatua) Competition." Weed Technology 7(4): 890-893.://A1993MW67000017 The effect of duration of wild oat competition on spring wheat yield and growth was determined in time-of-removal experiments conducted over a three year study period in Saskatchewan, Canada. Failure to remove wild oat reduced wheat yield 28 and 39% at wild oat populations of 64 and 188 plants per m(2), respectively. Wheat yield was not reduced by wild oat densities of 64 or 118 plants per m(2) until the six- and seven-leaf stage of wild oat, respectively. Removing wild oat at 64 plants per m(2) before the seven-leaf stage and 118 plants per m(2) before the five-leaf stage did not increase wheat culm or fresh weight production.

Kirkland, K. J. (1994). "Efficacy of Fall Incorporated and Nonincorporated Granular Triallate on Wild Oat (Avena-Fatua) and Wheat (Triticum-Aestivum)." Weed Technology 8(3): 607-611.://A1994PJ72000033 The influence of incorporation vs no incorporation on the efficacy of granular triallate applied in mid-October at 1400 and 1700 g ai/ha to control wild oat was evaluated in spring wheat in west central Saskatchewan over a 7-yr period. All fall- applied triallate reduced wild oat panicles and fresh weight, and increased yield compared to untreated checks. With applications in standing wheat stubble there was no difference in wild oat control from incorporation versus no incorporation. All triallate treatments reduced wild oat panicles and fresh weight by over 95%, and resulted in wheat yield increases ranging from 29 to 67%. In tilled fallow, incorporated granules provided better wild oat control than when there was no incorporation. Wheat yield increases ranged from 50 to 85% for triallate treatments with yield for incorporated triallate approximately 15% greater than non-incorporated. The rate of triallate did not affect the level of wild oat control achieved with either incorporation method. In separate tolerance studies triallate incorporation method did not affect spring wheat emergence or subsequent development. Nomenclature: Triallate, S-(2,3,3-trichloro-2-propenyl) bis(1-methylethyl) carbamothioate; wild oat, Avena fatua L. #3 AVEFA; spring wheat, Triticum aestivum L. 'Neepawa.'

Kirkland, K. J., K. N. Harker, et al. (1989). "Influence of Metribuzin and Cyanazine on the Phytotoxicity of Graminicides on Wild Oats and Barley." Canadian Journal of Plant Science 69(1): 195-203.://A1989T501500021

Kirkland, K. J. and J. H. Hunter (1991). "Competitiveness of Canada Prairie Spring with Wild Oat (Avena-Fatua L)." Canadian Journal of Plant Science 71(4): 1089-1092.://A1991GY53900014 Three spring wheat (Triticum aestivum L.) genotypes (Neepawa, a hard red spring, and HY320 and HY355 both Canada Prairie Spring wheats) were subjected to four levels of wild oat density at two locations over 4 yr to assess the effects of wild oat competition on biomass production, culm formation and yield. Wheat biomass and culm production were reduced at all wild oat density levels in each of the three cultivars. There were no significant differences among cultivars. In general, yields of all cultivars were reduced as wild oat density levels increased. Actual yield reductions at comparable wild oat densities tended to be greater at Regina than at Scott. There was a significant cultivar-by-density interaction for yield with yield reductions in HY320 > HY355 > Neepawa.

Kirkland, K. J., E. N. Johnson, et al. (2001). "Control of wild oat (Avena fatua) in wheat with MKH 6562." Weed Technology 15(1): 48-55.://000175240600009 MKH 6562 is a new acetolactate synthase inhibitor herbicide that would provide an alternative to control wild oat. Three experiments were conducted at Scott, SK, Canada, from 1996 to 1999 to evaluate MKH 6562 rates (20 and 30 g ai/ha). carrier volumes (30, 50, and 100 L/ha), time of applications (two- to three-leaf and three- to four-leaf stages of wild oat), and broadleaf weed herbicide tank mixtures. Reduced wild oat control (lower visual rating of percent control and higher fresh weight) often occurred when MKH 6562 was applied at a rate of 20 vs. 30 g/ha, with carrier volumes of 30 vs. 50 and 100 L/ha, and at the three- to four-leaf vs. two- to three-leaf stages of wild oat. Wild oat control generally was lower for MKH 6562 tank mixed with dicamba + mecoprop + MCPA 1:1:4.4 and bromoxynil + MCPA 1:1 compared with MKH 6562 applied alone or tank mixed with other broadleaf herbicides. MKH 6562 tank mixed with MCPA, and to a lesser extent 2,4- D, resulted in decreased wild oat control when applied at the three- to four-leaf stage of wild oat, but not at the two- to three-leaf stage. Wheat grain yield usually was not affected by MKH 6562 rate and carrier volume. Yield was 7% lower when MKH 6562 was tank mixed with dicamba + mecoprop + MCPA or fluroxypyr + 2,4-D and was 8% lower when MKH 6562 was applied in a mixture with 2,4-D formulations or bromoxynil + MCPA at the three- to four-leaf stage compared with the two- to three-leaf stage. Similar yields were achieved when MKH 6562 was applied alone at both leaf stages of wild oat. Wild oat control and wheat yield with MKH 6562 were as good as or better than with ICIA 0604 and imazamethabenz and were as good as or poorer than with CGA 184927. MKH 6562 provides adequate control of wild oat in wheat when applied early with the recommended carrier volume. Satisfactory control may be achieved with reduced rates if wild oat infestations are light.

Kirkland, K. J. and P. A. Osullivan (1984). "Control of Wild Oats in Wheat with Barban, Diclofop Methyl, Flamprop Methyl and Difenzoquat." Canadian Journal of Plant Science 64(4): 1019-1021.://A1984TY90500030

Kiviharju, E. and E. Pehu (1998). "The effect of cold and heat pretreatments on anther culture response of Avena sativa and A- sterilis." Plant Cell Tissue and Organ Culture 54(2): 97-104.://000079293100005 The effect of stress pretreatments on embryo induction in anther cultures of selected genotypes of Avena sativa and A. sterilis was tested. A heat pretreatment of isolated anthers at +32 degrees C for 5 days was best for the A. sativa line WW 18019 and for A. sterilis line CAV 2648. Genotype dependency may exist since in 'Stout' heat pretreatment did not 34

increase embryo production. For A. sterilis 13 green and three albino regenerants were produced, of which five plants (haploids) survived transfer to the greenhouse. For A. sativa, 30 various differentiation media/treatment combinations were used in an attempt to regenerate plants from embryos, with no success. Seven day cold treatment of cut tillers increased slightly the response level in 'Stout' and was routinely used in subsequent experiments. Maltose proved to be better then sucrose as a carbon source for the genotypes tested. Fourteen percent maltose promoted the highest induction in A. sterilis, but the quality of embryos was improved in the presence of 10% maltose for both species. Sub-optimal carbohydrate levels did not enhance embryo induction in oats.

Kiviharju, E., M. Puolimatka, et al. (1997). "Regeneration of anther-derived plants of Avena sterilis." Plant Cell Tissue and Organ Culture 48(2): 147-152.://A1997XJ61100012 Anther culture response of accessions of several hexaploid Avena species was studied with respect to requirement for auxin. The highest callus induction frequency was achieved with A. sativa L. cv. Stout (17.3%). The three most responsive genotypes, two of A. sativa and one of A. nuda L., had two-fold higher anther culture responses on growth regulator-free medium than on a medium containing 2,4-D. The A. sterilis L. accession CAV 2648 was the only genotype which consistently produced white, embryogenic structures (on solid MS+ 10% sucrose) and did so irrespective of the presence of 2,4-D. When transferred onto medium with lower sucrose concentration and an auxin transport inhibitor (TIBA), three green (two diploid [2n=6x=42] and one haploid [1n=3x=21]) and two albino plantlets were regenerated. The diploid regenerants set seed in the greenhouse. This first report of plants recovered from anther culture in the wild oat A. sterilis may provide an avenue to understand better and possibly overcome problems associated with androgenesis in cultivated oat.

Kiviharju, E., M. Puolimatka, et al. (1998). "The effect of genotype on anther culture response of cultivated and wild oats." Agricultural and Food Science in 7(3): 409-422.://000077353500008 Anther culture ability was tested for 44 oat (Avena saliva L.), six naked oat (A. sativa L., naked type) and 15 wild oat (Avena sterilis L.) genotypes, in addition to progeny of five intraspecific crosses of A. sativa and two interspecific crosses of A. sativa x A. sterilis. Anther culture response was affected considerably by genotype. Thirty one oat genotypes responded by callus growth on induction medium and seven of them produced embryo structures, two of the lines consistently. All naked oat genotypes produced embryo structures. Embryo production rates for the wild oat lines were comparable with those for the naked oat genotypes, and higher than for oat: 13 of the 15 genotypes tested produced embryo structures. plant regeneration was possible only from wild oat. The regeneration ability was inherited in the progeny of the A, sativa x A. sterilis cross cv. Puhti x CAV 2648. The response of anthers of oat genotypes was inhibited by auxin on the induction medium, while naked oat, wild oat and A. sativa x A. sterilis crosses responded better on a medium containing 2,4-dichlorophenoxy-acetic acid.

Kiviharju, E. M. and A. A. Tauriainen (1999). "2,4-dichlorophenoxyacetic acid and kinetin in anther culture of cultivated and wild oats and their interspecific crosses: plant regeneration from A-sativa L." Plant Cell Reports 18(7-8): 582-588.://000079282200011 The effect of 2,4-dichlorophenoxyacetic acid (2,4-D) and kinetin was studied in anther culture of oat Avena sativa L., wild oat A. sterilis L, and progeny of crosses between them. A high 2,4-D concentration (5-6 mg l(-1)) increased embryo production in genotypes of both species and promoted plant regeneration in anther cultures of A. sterilis and A. sativaxA. sterilis progeny, while kinetin caused severe browning. However, a low concentration: of kinetin was essential for initiation of regenerable embryos from anther culture of A. sativa cv. Kolbu: one green and one albino plant were produced. Tn addition, medium containing W-14 salts gave higher regenerant recovery compared with medium containing Murashge and Skoog salts, when cross progeny were tested.

Koide, R., M. Li, et al. (1988). "Role of Mycorrhizal Infection in the Growth and Reproduction of Wild Vs Cultivated Plants .1. Wild Vs Cultivated Oats." Oecologia 77(4): 537-543.://A1988R443600015

Koide, R. T. and X. H. Lu (1992). "Mycorrhizal Infection of Wild Oats - Maternal Effects on Offspring Growth and Reproduction." Oecologia 90(2): 218-226.://A1992HU05200009 The objective of this study was to determine whether infection of Avena fatua L. plants by the mycorrhizal fungus Glomus intraradices Schenck & Smith could influence the vigor of the offspring generation. Two experiments demonstrated that mycorrhizal infection of the maternal generation had slight but persistent positive effects on offspring leaf expansion in the early stages of growth. In two other experiments, mycorrhizal infection of mother plants had several long lasting effects on their offspring. Offspring produced by mycorrhizal mother plants had greater leaf areas, shoot and root nutrient contents and root:shoot ratios compared to those produced by non-mycorrhizal mother plants. Moreover, mycorrhizal infection of mother plants significantly reduced the weight of individual seeds produced by offspring plants while it increased the P concentrations of the seeds and the number of seeds per spikelet produced by offspring plants. The effects of mycorrhizal infections of maternal plants on the vigor and performance of offspring plants were associated with higher seed phosphorus contents but generally lighter seeds. The results suggest that mycorrhizal infection may influence plant fitness by increasing offspring vigor and offspring reproductive success in addition to previously reported increases in maternal fecundity.

Korber, H. D. (1982). "Wild Oats (-Spica-Venti) as Cause of Colic in Horses." Deutsche Tierarztliche Wochenschrift 89(7): 285-288.://A1982PC00500007

Koscelny, J. A. and T. F. Peeper (1997). "Herbicides for winter-hardy wild oat (Avena fatua) control in winter wheat (Triticum aestivum)." Weed Technology 11(1): 35-38.://A1997WX37900007 Diclofop at 840 g ai/ha, fenoxaprop at 90 g ai/ha, and imazamethabenz at 530 g ai/ha fall-applied controlled wild oat 96, 99, and 95% and increased wheat grain yields 26, 29, and 24%, respectively. These herbicides controlled wild oat over a wider range of growth stages than current labels indicate, The same treatments applied in March were less effective for wild oat control and did not increase wheat yield. 35

Kovacs, M. I. P. and G. M. Simpson (1976). "Dormancy and Enzyme Levels in Seeds of Wild Oats." Phytochemistry 15(4): 455- 458.://A1976BM93300002

Kropac, Z., T. Havranek, et al. (1986). "Effect of Duration and Depth of Burial on Seed Survival of Avena-Fatua in Arable Soil." Folia Geobotanica & Phytotaxonomica 21(3): 249-262.://A1986E077400003

Kudsk, P. and S. K. Mathiassen (1994). "Effects of Broadleaf Herbicides on Imazamethabenz-Methyl Performance on Wild Oat (Avena-Fatua L)." Weed Research 34(4): 251-263.://A1994NZ25100002 The influence of broadleaf herbicides on imazamethabenz-methyl performance on wild oat (Avena fatua L.) plants grown outdoors in pots was examined. The foliar activity of imazamethabenz-methyl was reduced when applied in mixture with salt formulations of MCPA and mecoprop, salt and ester formulations of bromoxynil and bentazone. In contrast, chlorsulfuron had no influence and ester formulations of MCPA and mecoprop either had no effect or promoted imazamethabenz-methyl performance. A comparison of the activity of imazamethabenz-methyl in mixture with the ester formulation of MCPA and the corresponding formulation blank revealed that the antagonistic effect of MCPA active ingredient in the salt and ester formulations were similar. However, because of an enhancing effect of the formulation constituents in the ester formulation antagonism was masked in mixture with the latter. Similarly, addition of a mineral oil adjuvant only masked but did not eliminate the antagonistic interaction between the two herbicides. Volume rate did not affect imazamethabenz-methyl activity nor the antagonistic effect of the salt formulation of MCPA. The ester formulation of MCPA, however, promoted imazamethabenz-methyl performance significantly more at the low than at the high volume rate. Sequential application reduced antagonism of the salt formulation of MCPA on imazamethabenz-methyl. Selective placement of droplets revealed that antagonism occurred only if imazamethabenz-methyl and the salt formulation of MCPA were applied in the same droplet but not as separate droplets. Neither foliar- nor root-applied MCPA reduced the performance of root-applied imazamethabenz-methyl. The results indicate that antagonism of MCPA was caused by a reduced uptake. The implications of the results of the present study in relation to the performance of tank mixtures of imazamethabenz-methyl and broadleaf herbicides under field conditions are discussed.

Kumaratilake, A. R., D. F. Lorraine-Colwill, et al. (2002). "A comparative study of glufosinate efficacy in rigid ryegrass (Lolium rigidum) and sterile oat (Avena sterilis)." Weed Science 50(5): 560-566.://000178083600003 Glufosinate efficacy was examined in two major grass weed species, rigid ryegrass and sterile oat Dose-response pot experiments under controlled environmental conditions showed that sterile oat was more successfully controlled by glufosinate than was rigid ryegrass Glutamine synthetase was extracted from both species and assayed in vitro Glufosinate readily inhibited glutamine synthetase activity in both species, indicating no differential sensitivity to the target enzyme Thin-layer chromatography analysis of glufosinate showed no significant metabolism of glufosinate in either species Absorption and translocation studies with C-14-glufosinate showed that the radiolabel was rapidly absorbed into the leaves of both species However, translocation of radiolabeled glufosinate from the treated leaf to the meristematic regions was significantly greater in sterile oat, whereas translocation to the tip of the leaf was significantly greater in rigid ryegrass This indicates that there is a difference in glufosinate distribution between the two species It is likely that this difference in the distribution of glufosinate results in sterile oat being more easily controlled by glufosinate than is rigid ryegrass.

Ladizinsky, G. (1995). "Domestication Via Hybridization of the Wild Tetraploid Oats Avena-Magna and a-Murphyi." Theoretical and Applied Genetics 91(4): 639-646.://A1995RW05700014 The wild tetraploid (2n = 28) oat species Avena magna and A. murphyi have been domesticated by having been transferred from the common oat, A sativa (2n = 42), the characteristics of non-shedding spikelets glabrous and yellow lemma, and reduced awn formation. Domestication has been achieved by crossing the common oat with either of the tetraploid species and then backcrossing the pentaploid hybrids with pollen of the tetraploid wild parent. Among the BC plants obtained only a few produced some seeds. Fertile tetraploids exhibiting the domesticated syndrome have been selected for in the F-2 generation. Although morphologically they were almost indistinguishable from the common oat, they were tetraploids. Wild x domesticated A. magna hybrids were vigorous and fertile. They retained their spikelets at maturity, lemma color and pubescence were intermediate between the parental lines, and awns were formed only on the lower floret of the spikelet. Each of these characteristics segregated in a 3:1 fashion, indicating single gene control, as in the common oat. These four characteristics formed a linkage group in one F-2 family and two linkage groups in the other two families. The usefulness of the domesticated tetraploids for oat research and production has been discussed. Taxonomically, the domesticated tetraploids were ranked as subspecies: A. magna ssp. domestica, and A. murphyi ssp. rigida.

Ladizinsky, G. and R. Fainstein (1977). "Domestication of Protein-Rich Tetraploid Wild Oats Avena-Magna and Avena-Murphyi." Euphytica 26(1): 221-223.://A1977CY78400027

Lake, J. R. and W. A. Taylor (1974). "Effect of Form of a Deposit on Activity of Barban Applied to Avena-Fatua L." Weed Research 14(1): 13-18.://A1974S397600002

Lamon, L. C. (1985). "Editorial Wild Oats - Carmack,Edward,Ward and Tennessee Politics - Majors,Wr." Journal of Southern History 51(3): 462-463.://A1985APD4600039

Landenmark, O. (1978). "Mechanical Separation of Wild Oat (Avena-Fatua) Seeds in Cereal Samples with Bardex Q1." Seed Science and Technology 6(2): 543-553.://A1978FU71200018

Landes, A. and J. R. Porter (1990). "Development of the Inflorescence in Wild Oats." Annals of Botany 66(1): 41-50.://A1990DQ97200006

36

Landry, B., A. Comeau, et al. (1984). "Genetic-Analysis of Resistance to Barley Yellow Dwarf Virus in Hybrids between Avena- Sativa Lamar and Virus-Resistant Lines of Avena-Sterilis." Crop Science 24(2): 337-340.://A1984SJ72400029

Latta, R. G., J. L. MacKenzie, et al. (2004). "Divergence and variation of quantitative traits between allozyme genotypes of Avena barbata from contrasting habitats." Journal of Ecology 92(1): 57-71.://000187891600006 1 Avena barbata occurs as two multilocus allozyme genotypes associated with moist (mesic) and dry (xeric) habitats in California. We examined the divergence of quantitative traits between these genotypes, and heritable trait variation in the progeny of a cross. 2 A replacement series showed that the mesic genotype was competitively superior to the xeric. Early germination could alter this competitive outcome, but there was very little difference between the genotypes in germination time. 3 The mesic genotype exhibited larger seeds, and seedlings, as well as greater fecundity. However, there was no difference in relative growth rate between the genotypes. Thus the early size advantage of the mesic genotype seems to be due primarily to larger seed size. 4 Seedlings of the xeric genotype expressed a greater root mass ratio (RMR), and allocated a higher fraction of their root mass deeper in the soil when grown in artificial soil columns than did the mesic genotype. 5 We crossed the two genotypes and allowed the F1 to propagate by self-fertilization to yield F3 families. There was significant among-family (i.e. genetic) variation in RMR, root mass allocation by depth, and seed size. F3 families did not vary significantly in rooting depth per unit shoot mass. 6 In a competition experiment similar to that above, there was significant variation among F3 families in fecundity, but not dry mass, and family by competitor interactions were not significant. 7 There was little correlation among traits across F3 family means, indicating no common genetic basis to the traits. This suggests that novel combinations of characters can be created through crossing between genetically diverged populations. In addition, there was a significant family by block (environment) interaction for RMR, indicating that the degree of plasticity in this character has heritable variation that can respond to natural selection.

Lawrence, P. K. and K. J. Frey (1975). "Backcross Variability for Grain Yield in Oat Species Crosses (Avena-Sativa L X Avena- Sterilis L)." Euphytica 24(1): 77-85.://A1975W050000011

Lawrence, T. L. (1973). "Utilization by Growing Pig of Barley Adulterated with Seeds of Wild Oat (Avena-Fatua)." Veterinary Record 93(26): 666-667.://A1973R631300005

Legere, A., H. J. Beckie, et al. (2000). "Survey of management practices affecting the occurrence of wild oat (Avena fatua) resistance to acetyl-CoA carboxylase inhibitors." Weed Technology 14(2): 366-376.://000166693500018 A survey conducted across agricultural ecoregions of Saskatchewan in 1996 revealed that wild oat (Avena fatua) populations resistant to acetyl-CoA carboxylase (ACCase) inhibitors were present in approximately 10% of Saskatchewan fields (2.4 million ha). In the Aspen Parkland and Boreal Transition ecoregions, this increased to 17%. The objective of this study was to determine if agronomic practices promoted or delayed resistance and to assess producer awareness of herbicide resistance. Weed resistance and management questionnaire data from the 1996 resistance survey and management questionnaire data from the 1995 Saskatchewan weed survey were submitted to multiway frequency analysis. The frequency of occurrence of herbicide-resistant wild oat was related directly to ACCase inhibitor use. Resistance to cyclohexanedione (CHD) herbicides was not related to CHD use but to frequency of ACCase inhibitor use (i.e., CHD + aryloxyphenoxypropanoate [AOPP]), suggesting that the pressure imposed by AOPPs contributed to the selection of CHD resistance in wild oat. ACCase inhibitor use was more extensive in the Aspen Parkland and Boreal Transition ecoregions than in the Mixed and Moist Mixed Grassland ecoregions, Crop rotations were not conducive to rotation of herbicides with different sites of action, Frequency of ACCase inhibitor use increased with frequency of annual crops, in spite of the inclusion of cereal and dicot crops in the rotation. Producers utilizing conservation tillage practices in the Grassland ecoregions used proportionally more ACCase inhibitors than those using conventional tillage practices. This increase in ACCase use in conservation tillage systems did not result in an increased incidence of wild oat populations resistant to ACCase inhibitors. Producers reporting troublesome wild oat populations tended to have proportionally more ACCase resistant wild oat. Producers who reported practicing weed sanitation were less likely to have resistant wild oat than those who were less careful. Increased awareness and implementation of management practices that will reduce the dependency on ACCase herbicides are required to better enable producers to prevent, delay, or manage herbicide- resistant wild oat populations.

Leguizamon, E. S., C. Fernandez-Quintanilla, et al. (2005). "Using thermal and hydrothermal time to model seedling emergence of Avena sterilis ssp ludoviciana in Spain." Weed Research 45(2): 149-156.://000227351300009 Historical records on the field emergence of seedlings of Avena sterilis in winter cereal crops were used to develop thermal/hydrothermal models to predict the emergence of this weed as a function of meteorological conditions. When water was not a limiting factor, a Weibull function provided a good description of the relationship between thermal time and seedling emergence. The variability in the rate of seedling emergence observed under these conditions was associated with variation in the differential between measured and base temperatures. When water was limiting in the soil, the rate of seedling emergence was reduced, a different function was required to describe the process. The use of a hydrothermal model enabled the use of a single function to describe the patterns of seedling emergence of the nine site-years considered in the study. Using this model we can conclude that 70% seedling emergence will be reached, under very diverse conditions, within 300 hydrodegree days. The hydrothermal model proposed was validated with independent seedling emergence data, supporting the idea that this model is robust enough to be used as a predictive tool for seedling emergence of A. sterilis in a variety of conditions.

Leist, N. (1981). "Investigations into Hybrids between Cultivated Oats (Avena Sativa, a Byzantina) and Wild Oats (a Fatua, a Sterilis) as Well as into Fatuoids." Seed Science and Technology 9(3): 781-805.://A1981MZ60700012

Li, B. L. and M. E. Foley (1994). "Differential Polypeptide Patterns in Imbibed Dormant and after-Ripened Avena-Fatua Embryos." Journal of Experimental Botany 45(271): 275-279.://A1994MY77600015 To test the hypothesis that differential gene expression is responsible for maintaining seed dormancy in wild oat (Avena fatua L.), we compared the polypeptide patterns of dormant and after-ripened (non-dormant) embryos of the inbred line 37

M73. Polypeptides labelled in vivo or translated in vitro were separated by two-dimensional electrophoresis followed by fluorography. Many of the newly synthesized proteins were more abundant in dormant embryos, while only a few were more prevalent in after-ripened embryos. The results demonstrate that steady-state levels of specific proteins and/or mRNAs differ between dormant and after-ripened embryos during early imbibition, and these differences may be associated with maintaining and/or breaking seed dormancy.

Li, B. L. and M. E. Foley (1995). "Cloning and characterization of differentially expressed genes in imbibed dormant and afterripened Avena fatua embryos." Plant Molecular Biology 29(4): 823-831.://A1995TL29900018 To analyze the patterns of gene expression associated with seed dormancy in wild oat (Avena fatua), we have isolated cDNA clones corresponding to genes that are differentially expressed in dormant and afterripened line M73 embryos. Gene transcripts of these clones were maintained in embryos of imbibed dormant caryopses, but declined rapidly in afterripened embryos after imbibition. GA, treatment of dormant caryopses, which breaks dormancy, could lower the transcript levels in dormant embryos. When the germination of afterripened caryopses was inhibited by high temperature (35 degrees C), the decline in abundance of the transcripts in afterripened embryos was arrested. These genes were expressed to various degrees in water-stressed, but not in unstressed, 7-day-old seedlings. The expression of the genes was also ABA-inducible in afterripened embryos. The expression patterns in non-dormant line SH430 wild oat were similar to those of afterripened M73. DNA sequence analyses indicated that some of the cDNA clones encode LEA (late embryogenesis-abundant) proteins and aldose reductase. The significance of the expression of these genes in maintaining seed dormancy or longevity is discussed.

Litav, M. (1965). "Effects of Soil Type and Competition on Occurrence of Avena Sterilis L in Judean Hills (Israel)." Israel Journal of Botany 14(2): 74-&.://A19657208800003

Liu, S. H., A. I. Hsiao, et al. (1995). "Effect of Sodium Bisulfate on the Phytotoxicity, Retention, Foliar Uptake, and Translocation of Imazamethabenz on Wild Oats (Avena-Fatua L)." Weed Science 43(1): 40-46.://A1995QK71000008 Studies determined the effect of sodium bisulfate (NaHSO4) on the phytotoxicity, retention, uptake, and translocation of the suspension concentrate formulation of imazamethabenz in wild oats, NaHSO4 completely solubilized this herbicide formulation when added in an equimolar concentration and did not affect herbicidal activity at NaHSO4 concentrations below 28 mM when used in a carrier volume of 100 L ha(-1), NaHSO4 improved phytotoxicity at a carrier volume of 33 L ha(-1), NaHSO4 at 28 mM increased the efficacy of imazamethabenz applied as individual drops on growth chamber- grown wild oats, The herbicide retention was not changed but foliar absorption and acropetal translocation were increased, while basipetal translocation was decreased, The increase in phytotoxicity of imazamethabenz with NaHSO4 was related to absorption of imazamethabenz by the target plants.

Liu, S. H., W. A. Quick, et al. (1994). "Effect of Mcpa on the Phytotoxicity of Imazamethabenz-Methyl Applied to Wild Oats (Avena- Fatua L)." Weed Research 34(6): 425-431.://A1994PZ48200005 Greenhouse studies were conducted to determine the effect of both ester and amine formulations of MCPA on the phytotoxicity of imazamethabenz applied to wild oats (Avena fatua). The MCPA ester antagonized activity of the liquid concentrate (LC) formulation of imazamethabenz but not the suspension concentrate (SC) formulation of imazamethabenz when the combination was applied to wild oats at two- to three-leaf stage without tiller. The MCPA amine antagonized the efficacy of both formulations of imazamethabenz on wild oats at the two- to three-leaf stage. When the herbicides were applied at the three- to four-leaf stage with one or two tillers, the antagonism was found only with MCPA amine and imazamethabenz-LC. In general, antagonism is most likely to occur at an early stage of wild oats. MCPA amine was more antagonistic than MCPA ester on the phytotoxicity of imazamethabenz. Imazamethabenz-SC can be tank mixed with MCPA ester but should not be tank mixed with MCPA amine. Imazamethabenz-LC should not be tank mixed with either ester or amine of MCPA because of antagonism, especially when herbicides are applied at early stage. When MCPA was applied at intervals of from 2 to 24 h following applications of imazamethabenz-LC, there was no antagonistic effect of MCPA and the same level of phytotoxicity was produced as with imazamethabenz-LC alone. In the reverse sequence of MCPA followed by imazamethabenz-LC, the greatest antagonism of phytotoxicity of imazamethabenz-LC occurred when 2 h separated the two applications. This effect was lessened when the elapsed time increased.

Lockhart, S. J. and K. A. Howatt (2004). "Split applications of herbicides at reduced rates can effectively control wild oat (Avena fatua) in wheat." Weed Technology 18(2): 369-374.://000222194700026 Split application of herbicides for wild oat control may minimize wild oat competition with wheat while reducing the number of wild oat seeds returned to the soil. Field experiments were conducted in 2000 and 2001 to evaluate the effects of CGA- 184927, fenoxaprop-P, flucarbazone, and ICIA 0604 at labeled and reduced rates on wild oat control, wild oat seed rain, and wheat yield. Each herbicide was applied once at 25, 33, and 100% of the labeled rate at the two-leaf stage of wild oat or split applied at 50 and 66% of the full rate as two equal applications. Excellent full-season wild oat control was obtained with CGA-184927, flucarbazone, and ICIA 0604 applied twice at reduced rates. ICIA 0604 or CGA-184927 split applied at 25 and 33% rates (totaled 50 and 66% of the full rate) provided wild oat control equal to one application of labeled rates. Wild oat seed rain was similar among all herbicide treatments, except plots treated with fenoxaprop-P once at 25 and 33% rates where seed rain was higher and equal to 47% of untreated plots. Wheat yields and net returns were highest and similar after treatment with CGA-184927 or ICIA 0604 applied either once at the labeled rate or split applied at 25 or 33% rates.

Loskutov, I. G. (2001). "Influence of vernalization and photoperiod to the vegetation period of wild species of oats (Avena spp.)." Euphytica 117(2): 125-131.://000165378600005 This paper presents the results of a six year field study of wild Avena species and their response to vernalization and photoperiod. The accessions of twenty one wild and weedy species were tested under 12-hr and 18-hr daylength and cold temperatures (for 40 days at +2 C-degrees) treatments and without it (as a control). The results demonstrate that for the majority of species evaluated, cold temperature requirements had a greater influence on heading date and the duration of the vegetative period than daylength. Genotypes with neutral, weak and strong reactions to all treatments were found and 38

spring and winter types were selected. The results further demonstrated that daylength-insensitive forms occurred in the south Mediterranean region and adjacent southern territories.

Lovegrove, A. W. (1985). "The Effect of Glyphosate Applied through a Rope Wick Applicator on the Viability of Avena-Fatua Seeds." Annals of Applied Biology 106: 102-103.://A1985A226700052

Lu, X. and T. Koide (1991). "Avena-Fatua L Seed and Seedling Nutrient Dynamics as Influenced by Mycorrhizal Infection of the Maternal Generation." Plant Cell and Environment 14(9): 931-939.://A1991GY05500006 The objective of this study was to determine how mycorrhizal infection of one generation of plants influences the nutrient dynamics of seeds and seedlings comprising the subsequent generation. We showed that, for Avena fatua L., seeds produced by mycorrhizal (M) plants consistently contained significantly more phosphorus (particularly the phytate P and residual P fractions) than seeds produced by non-mycorrhizal (NM) plants. We also followed the development of spikelets produced by M and NM plants. The rates of increase in spikelet dry weight and nitrogen content were largely unaffected by mycorrhizal infection. However, the rate of P accumulation into spikelets was significantly increased by mycorrhizal infection. Greater endosperm P reserves in seeds produced by M plants were associated with greater rates of P accumulation in resultant seedlings. Moreover, offspring plants (all NM) produced by M mother plants had significantly higher root and rhizosphere phosphatase, ATPase and phytase activities than offspring plants produced by NM mother plants. This persistent maternal effect has never before been described. Our results suggest that mycorrhizal infection of one generation of plants may have substantial positive effects on the offspring generation, and thus, may influence plant population dynamics.

Luby, J. J. and D. D. Stuthman (1983). "Evaluation of Avena-Sativa-L/a Fatua-L Progenies for Agronomic and Grain Quality Characters." Crop Science 23(6): 1047-1052.://A1983SA02800007

Luby, J. J., D. D. Stuthman, et al. (1985). "Micronuclei Frequency and Character Coherence in Avena-Sativa L Avena-Fatua L Crosses." Theoretical and Applied Genetics 69(4): 367-373.://A1985AJR3600005

Lutz, A. W. and R. W. Feeny (1973). "New Class of Herbicides for Selective Control of Wild Oats." Abstracts of Papers of the American Chemical Society(AUG26): 65-65.://A1973Q295700966

Lyrene, P. M. and H. L. Shands (1974). "Groat Protein Percentage in Avena-Sativa L Fatuoids and in a Fatuoid X a-Sterilis L Cross." Crop Science 14(5): 765-767.://A1974U548700045

Lyrene, P. M. and H. L. Shands (1975). "Associations among Traits in Progenies from Avena-Sativa L X a Sterilis L Crosses." Crop Science 15(3): 361-363.://A1975AF85500024

Lyrene, P. M. and H. L. Shands (1975). "Groat Protein Percentage in Avena-Sativa X a Sterilis Crosses in Early Generation." Crop Science 15(3): 398-400.://A1975AF85500034

Lyrene, P. M. and H. L. Shands (1975). "Heading Dates in 6 Avena-Sativa L X a Sterilis L Crosses." Crop Science 15(3): 359- 360.://A1975AF85500023

Macias, F. A., D. Marin, et al. (2006). "Structure-activity relationship (SAR) studies of benzoxazinones, their degradation products, and analogues. Phytotoxicity on problematic weeds Avena fatua L. and Lolium rigidum Gaud." Journal of Agricultural and Food Chemistry 54(4): 1040-1048.://000235627800009 Avena fatua L. (wild oat) and Lolium rigidum Gaud. (rigid ryegrass) are highly problematic weeds affecting a wide variety of cereal crops worldwide. The fact that both of these weeds have developed resistance to several herbicide groups made them optimal candidates as target organisms for ongoing research about the potential application of allelochemicals and analogue compounds as natural herbicide models. Benzoxazinones, a family of natural allelochemicals present in corn, wheat, and rye, including 2,4-dihydroxy-(2H)-1,4-benzoxazin-3(4H)-one and 2,4-dihydroxy-7-methoxy-(2H)-1,4- benzoxazin-3(4H)-one, together with some degradation products, found in crop soils as well as in other systems, and some synthetic analogues of them were tested on wild oat and rigid ryegrass seeds; the results were statistically treated, and some structure-activity relationships, useful in further development of natural herbicide models, were elucidated. The most active compounds were the synthetic benzoxazinone 2-acetoxy-(2H)-1,4-benzoxazin-3(4H)-one and the degradation product 2-aminophenoxazin-3-one, with highly significant inhibition on the development of both weeds. The ecological role of these compounds is discussed by considering both degradability and phytotoxicity. The bioactivity of aminophenoxazines has been correlated by their aqueous solubility-lipophilicity predicted by means of computational methods.

Mackay, P. (1978). "Wild Oats." Theatre Crafts 12(3): 10-10.://A1978EQ18200013

Mahall, B. E., V. T. Parker, et al. (1981). "Growth and Photosynthetic Irradiance Responses of Avena-Fatua L and Bromus- Diandrus Roth and Their Ecological Significance in Californian Savannas." Photosynthetica 15(1): 5-15.://A1981LP60000002

Mahercha.N and J. M. Naylor (1972). "Onset of Synthesis of Rna and Protein in Mature Aleurone Tissue of Wild Oats (Avena- Fatua)." Canadian Journal of Botany 50(2): 305-&.://A1972L803800008

Mahercha.Nj and J. M. Naylor (1971). "Variability in DNA Content and Nuclear Morphology of Aleurone Cells of Avena-Fatua (Wild Oats)." Canadian Journal of Genetics and Cytology 13(3): 578-&.://A1971L149500023

39

Maherchandani, N. (1975). "Effects of Gamma-Radiation on Dormant Seed of Avena-Fatua L." Radiation Botany 15(4): 439- 443.://A1975BE34400017

Maherchandani, N. (1976). "Stimulation of Germination by Gamma-Radiations of Dormant Seeds of Avena-Fatua L." Current Science 45(17): 629-630.://A1976CC92600019

Malcherek, K., J. Breuer, et al. (1998). "Metabolism of 4-nitrophenol in aseptically cultivated plants of the species wheat (Triticum aestivum L.), soybean (Glycine max L.), wild oat (Avena fatua L.) and corn cockle (Agrostemma githago L.)." Journal of Plant Physiology 153(1-2): 192-199.://000075669800028 The metabolism and behaviour of the root applied xenobiotic [U-C-14] 4-nitrophenol was studied in intact, aseptically grown plants of wheat ( Triticum aestivum L.), soybean (Glycine max L.), wild oat (Avena fatua L.), and corn cockle (Agrostemma githago L.). After 7 days of exposition, different port-ions of the applied radioactivity were found absorbed into the plants: wild oat, 11.4%; corn cockle, 16.9%; wheat, 18.2%; soybean, 47.8%. Complementary long-term experiments performed only with wild oat and corn cockle showed chat the uptake continued, and after 28 days of exposition, C-14 portions found in the plants were 24.7% and 19.9%, respectively. In all species studied, the main percentages of absorbed radioactivity were confined to che roots (>75% after 7 days of incubation). After 28 days, however, considerable portions of C-14 were translocated to the aerial parts of wild oat (43.6% of absorbed C-14) and corn cockle (57.9%). 4-Nitrophenol was almost quantitatively transformed by all plant species examined to polar soluble conjugates, besides minor amounts of non-extractable residues (wild oat: 13.2% after 1 day, corn cockle: 2.0% after 28 days). The primary conjugate was 4-nitrophenyl-beta-D-glucoside, which was identified by TLC and HPLC. In the two monocotyledonous species, this product was partially further conjugated to 4-nitrophenyl-beta-D-gentiobioside, and an unidentified 4-nitrophenyl-(1 --> 4)-diglucoside. Additionally the 6'-O-malonylated beta-glucoside was formed in wheat. In both dicotyledonous species, the beta-glucoside was in part found to be 6'-O-esterified with malonic acid.

Maneechote, C., J. A. M. Holtum, et al. (1994). "Resistant Acetyl-Coa Carboxylase Is a Mechanism of Herbicide Resistance in a Biotype of Avena-Sterilis Ssp Ludoviciana." Plant and Cell Physiology 35(4): 627-635.://A1994NR70500012 A biotype of Avena sterilis ssp. ludoviciana is highly resistant to a range of herbicides which inhibit a key enzyme in fatty acid synthesis, acetyl-CoA carboxylase (ACCase). Possible mechanisms of herbicide resistance were investigated in this biotype. Acetyl-CoA carboxylase from the resistant biotype is less sensitive to inhibition by herbicides to which resistance is expressed. I,, values for herbicide inhibition of ACCase were 52 to 6 times greater in the resistant biotype than in the susceptible biotype. This was the only major difference found between the resistant and susceptible biotypes. The amount of ACCase in the meristems of the resistant and susceptible is similar during ontogeny and no difference was found in distribution of ACCase between the two biotypes. Uptake, translocation and metabolism of [C-14]diclofop-methyl were not different between the two biotypes. In vivo, ACCase activity in the meristems of the susceptible biotype was greatly inhibited by herbicide application whereas only 25% inhibition occurred in the resistant biotype. Depolarisation of plasma membrane potential by 50 mu M diclofop acid was observed in both biotypes and neither biotype showed recovery of the membrane potential following removal of the herbicide. Hence, a modified form of ACCase appears to be the major determinant of resistance in this resistant wild oat biotype.

Maneechote, C., C. Preston, et al. (1997). "A diclofop-methyl-resistant Avena sterilis biotype with a herbicide-resistant acetyl- coenzyme A carboxylase and enhanced metabolism of diclofop-methyl." Pesticide Science 49(2): 105-114.://A1997WH56900001 An Avena sterilis biotype was found to be highly resistant to aryloxyphenoxypropionate (APP) herbicides, especially diclofop-methyl. At the enzyme level, this biotype contained a modified acetyl-coenzyme A carboxylase (ACCase) with six- fold resistance to diclofop acid. Absorption and translocation of [C-14]diclofop-methyl applied to the leaf axil of the two-leaf stage plants were similar in both susceptible and resistant biotypes. However, the rate of metabolism of [C-14]diclofop was increased 1.5-fold in this resistant biotype compared to the susceptible. Experiments with tetcyclacis, a cytochrome P450 monooxygenase inhibitor, indicated that inhibition of this enhanced diclofop metabolism increased diclofop-methyl phytotoxicity in this biotype. Studies with ten individual families of the resistant biotype indicated that both mechanisms of resistance, an altered target site and enhanced metabolism, are present in each individual of the population. Hence, it is likely that these two mechanisms of resistance both contribute to resistance in this biotype.

Manisterski, J. (1983). "Slow-Rusting Type of Resistance to Stem Rust Disease in Avena-Sterilis - Importance, Stability and Mechanisms Controlling the Phenomenon." Phytoparasitica 11(3-4): 217-217.://A1983RX43700043

Mansooji, A. M., J. A. Holtum, et al. (1992). "Resistance to Aryloxyphenoxypropionate Herbicides in 2 Wild Oat Species (Avena- Fatua and Avena-Sterilis Ssp Ludoviciana)." Weed Science 40(4): 599-605.://A1992KK93300016 Resistance to the methyl ester of diclofop, an aryloxyphenoxypropionate graminicide, was shown for a wild oat (Avena fatua) population from Western Australia, and marked resistance to a range of aryloxyphenoxypropionate and cyclohexanedione graminicides was detected in a winter wild oat (Avena sterilis ssp. ludoviciana) population from South Australia. The A. sterilis biotype exhibited high levels of resistance to the aryloxyphenoxypropionate herbicides diclofop, fluazifop, haloxyfop, fenoxaprop, quizalofop, propaquizafop, and quinfurop and low levels of resistance to the cyclohexanedione herbicides sethoxydim, tralkoxydim, and cycloxydim. Ratios of LD50 values for responses of resistant and susceptible A. sterilis to the aryloxyphenoxypropionate herbicides were between 20 for propaquizafop and > 1,000 for fluazifop, and were between 2.5 and 3 for the cyclohexanedione herbicides. The LD50 value for diclofop for the A. fatua biotype was 442 g ai ha-1 which was 2.7-fold that of a susceptible control. Thirty-three percent of the plants survived at the registered rate of application.

Mariot, M. P., M. Sereno, et al. (1999). "Inheritance of plant height and main panicle length for the crosses between Avena sativa L and Avena sterilis L." Pesquisa Agropecuaria Brasileira 34(1): 77-82.://000078623200011 The wild species Avena sterilis L. has been widely used for crossing with cultivated oat. The high affinity between them enables an increase in genetic variability and transference of important agronomic genes. However, undesirable 40

characters such as great plant height and length of the panicle can be transferred to progenies. These characters were analysed to evaluate the genetic variability and inheritance in crossing between Avena sativa L. and Avena sterilis L. Estimates of phenotypical and genetical variances were high while the environmental variance was low, providing a high broad sense heritability. The additive gene effect was the most important for explaining the genetic variation for both characters.

Marshall, D. R. and R. W. Allard (1969). "Genetics of Electrophoretic Variants in Avena .I. Esterase E4 E9 E10 Phosphatase P5 and Anodal Peroxidase Apx5 Loci in a Barbata." Journal of Heredity 60(1): 17-&.://A1969D086300004

Marshall, D. R. and R. W. Allard (1970). "Isozyme Polymorphisms in Natural Populations of Avena-Fatua and a-Barbata." Heredity 25: 373-&.://A1970H402200004

Marshall, D. R. and R. W. Allard (1970). "Maintenance of Isozyme Polymorphisms in Natural Populations of Avena-Barbata." Genetics 66(2): 393-&.://A1970I346300014

Marshall, D. R. and S. K. Jain (1967). "Cohabitation and Relative Abundance of 2 Species of Wild Oats." Ecology 48(4): 656-&.://A1967A010300019

Marshall, D. R. and S. K. Jain (1968). "Phenotypic Placticity of Avena Fatua and a Barbata." American Naturalist 102(927): 457- &.://A1968C396600006

Marshall, D. R. and S. K. Jain (1969). "Genetic Polymorphism in Natural Populations of Avena Fatua and a Barbata." Nature 221(5177): 276-&.://A1969C452500049

Marshall, D. R. and S. K. Jain (1969). "Interference in Pure and Mixed Populations of Avena Fatua and a Barbata." Journal of Ecology 57(1): 251-&.://A1969D234800016

Marshall, D. R. and S. K. Jain (1970). "Speed Predation and Dormancy in Population Dynamics of Avena-Fatua and a-Barbata." Ecology 51(5): 886-&.://A1970H939400015

Martin, M. and R. J. Field (1987). "Competition between Vegetative Plants of Wild Oat (Avena-Fatua L) and Wheat (Triticum- Aestivum L)." Weed Research 27(2): 119-124.://A1987G569400006

Martin, R. J., B. R. Cullis, et al. (1987). "Prediction of Wheat Yield Loss Due to Competition by Wild Oats (Avena Spp)." Australian Journal of Agricultural Research 38(3): 487-499.://A1987J717700004

Martin, R. J. and W. L. Felton (1993). "Effect of Crop-Rotation, Tillage Practice, and Herbicides on the Population-Dynamics of Wild Oats in Wheat." Australian Journal of Experimental Agriculture 33(2): 159-165.://A1993LF98800006 The effects of crop rotation, tillage practice, and herbicide use on the population dynamics of wild oats (Avena fatua and A. sterilis ssp. ludoviciana) were studied in a field experiment in northern New South Wales. In the third and fourth years of a continuous wheat rotation, cultivated fallow using tines increased wild oat density and reduced grain yield compared with a no-tillage fallow. Tillage did not affect the vertical distribution of wild oat seeds in the soil, and about 80% of wild oat seeds were in the top 5 cm of soil in both tillage treatments. The seed reservoir at the end of the experiment was smaller under a no-tillage fallow regime. The half-life of wild oat seeds in the soil was about 6 months, and rotation of wheat with sorghum was the most effective means of reducing the wild oat seed reservoir. Rotational strategies for weed control are also likely to be effective in delaying or minimising the development of herbicide resistance, particularly where the average seed bank life time, as shown for wild oats in this study, is short. Annual use of either tri-allate or flamprop-methyl in 4 successive wheat crops did not prevent a massive build-up of wild oat seed. The poor performance of herbicides was partly attributed to below-average rainfall in autumn and early winter in 1985 and 1986. However, wild oats are well adapted to continuous cropping with wheat, where recruitment of 3-6% of the soil seed reservoir maintained the population despite the use of selective herbicides. The results of this experiment indicate that a continuous wheat rotation using herbicides to control wild oats is likely to be much less effective in reducing the wild oat seed reservoir.

Martin, R. J. and W. L. Felton (1994). "Effect of Crop-Rotation, Tillage Practice, and Herbicides on the Population-Dynamics of Wild Oats in Wheat - Reply." Australian Journal of Experimental Agriculture 34(5): 639-639.://A1994PW72400011

Martin, R. J., W. L. Felton, et al. (1989). "A Comparison of Tri-Allate Formulations for Control of Wild Oats in Wheat in Northern New-South-Wales." Australian Journal of Experimental Agriculture 29(2): 215-221.://A1989U760000011

Martin, W. W. (1978). "Dichlorfop-Methyl Controls Wild Oats." Agricultural Research 27(1): 16-16.://A1978FJ42400014

Martinoli, G. and L. Bagnoli (1962). "Effetti Delle Radiazioni Ionizzanti Su Cellule Meristematiche Apicali Di Avena Sterilis (2n = 42)." Caryologia 15(2): 327-&.://A19628471A00010

McBeath, D. K., D. A. Dew, et al. (1970). "Competition between Barley and Wild Oats as Affected by Nitrogen, Barban and Time of Seeding." Canadian Journal of Plant Science 50(5): 541-&.://A1970H446100005

McIntyre, G. I., A. J. Cessna, et al. (1996). "Seed dormancy in Avena fatua: Interacting effects of nitrate, water and seed coat injury." Physiologia Plantarum 97(2): 291-302.://A1996UT36400012 In experiments conducted under controlled conditions, KNO3 (50 or 100 mM) promoted germination of a dormant strain (AN 474) of Avena fatua when either one or two holes were pierced in the lower (adaxial) surface of the caryopsis in contact with the nitrate solution. Germination was increased by increasing either the KNO3 concentration or the number of 41

holes in the seed coat. The germination response induced by the application of water to a hole pierced in the upper surface of the caryopsis was increased by pre-treatment of the intact caryopsis with KNO3. Treatment with either 50 or 100 mM KNO3 caused a transient reduction in embryo water content of intact caryopses, but increased the nitrate and amino-N content of pierced caryopses prior to germination. Supplying a 100 mM solution of KNO3 to pierced caryopses reduced the total water potential and osmotic potential of the embryo, and increased its pressure potential by the same amount as an equimolar solution of KCl; however, while both treatments promoted germination, the KNO3 induced more rapid germination than the KCl. Both treatments also increased the K+ content of the embryo, the KNO3 again having the greater effect. These results are consistent with the hypothesis, based on our previous investigations, that KNO3 promotes germination of dormant caryopses by accumulating in the embryo where it acts osmotically to increase water uptake. It is also postulated that, in contrast to KCl, KNO3 may combine an osmotic effect on water uptake with a nutritional effect on protein synthesis.

McIntyre, G. I. and A. I. Hsiao (1985). "Seed Dormancy in Avena-Fatua .2. Evidence of Embryo Water-Content as a Limiting Factor." Botanical Gazette 146(3): 347-352.://A1985AVU2100011

McKenzie, R. I. and G. Fleischmann (1964). "Inheritance of Crown Rust Resistance in Selections from 2 Israeli Collections of Avena Sterilis." Canadian Journal of Genetics and Cytology 6(2): 232-&.://A19648041A00020

McKenzie, R. I., J. W. Martens, et al. (1970). "Inheritance of Oat Stem Rust Resistance in a Tunisian Strain of Avena-Sterilis." Canadian Journal of Genetics and Cytology 12(3): 501-&.://A1970H893800013

McKersie, B. D. and D. T. Tomes (1980). "Effects of Dehydration Treatments on Germination, Seedling Vigor, and Cytoplasmic Leakage in Wild Oats and Birdsfoot-Trefoil." Canadian Journal of Botany-Revue Canadienne De Botanique 58(4): 471-476.://A1980JK10500012

McMullen, M. S., R. L. Phillips, et al. (1982). "Meiotic Irregularities in Avena-Sativa L/a Sterilis L Hybrids and Breeding Implications." Crop Science 22(4): 890-897.://A1982PG45700045

Medd, R. W. (1979). "Difenzoquat and Seed Viability of Wild Oats." Pans 25(1): 91-92.://A1979GN74500015

Medd, R. W. (1994). "Effect of Crop-Rotation, Tillage Practice, and Herbicides on the Population-Dynamics of Wild Oats in Wheat - Comment." Australian Journal of Experimental Agriculture 34(5): 637-639.://A1994PW72400010

Mengistu, L. W., C. G. Messersmith, et al. (2005). "Genetic diversity of herbicide-resistant and -susceptible Avena fatua populations in North Dakota and Minnesota." Weed Research 45(6): 413-423.://000233171200002 Genetic diversity within and among 20 herbicide-resistant (HR) and 16 herbicide-susceptible (HS) Avena fatua multi-field populations was determined using 82 polymorphic loci resulting from two intersimple sequence repeat (ISSR) primers and one long-primer random amplified polymorphic DNA (LP-RAPD) primer. Collections from the Red River Valley of North Dakota and Minnesota, sampled in 1964 and 2000, represented A. fatua populations before and after intensive exposure to herbicides. A 1995 collection from south-west North Dakota represented A. fatua exposed to low herbicide selection. Despite differences in years of herbicide exposure among collections, both HR and HS populations from every collection maintained nearly similar levels of ISSR and RAPD diversity. Genetic differentiation among populations (G(ST)) varied from 11% to 13% among HR populations and from 9% to 16% among HS populations, indicating that 84-91% of total variation remained within HS or within HR populations. Minimal difference in gene diversity between HR and HS is consistent with multiple origins of resistance, where HR A. fatua most likely evolved from diverse founding individuals.

Merritt, C. R. (1982). "The Influence of Form of Deposit on the Phytotoxicity of Difenzoquat Applied as Individual Drops to Avena- Fatua." Annals of Applied Biology 101(3): 517-&.://A1982PS71900013

Mesbah, A., S. D. Miller, et al. (1995). "Wild Mustard (Brassica-Kaber) and Wild Oat (Avena-Fatua) Interference in Sugar-Beets (Beta-Vulgaris L)." Weed Technology 9(1): 49-52.://A1995QQ42600008 Two furrow irrigated field experiments were conducted for two years at the Research and Extension Center, Powell, WY to determine the influence of various mixed densities and durations of wild oat and wild mustard interference in sugarbeet. Sugarbeet root yields were reduced by competition from all examined densities of wild oat and wild mustard, alone and in combination. Root yield reduction was less than additive with mixed densities of wild oat and wild mustard. Root yields decreased as the duration of interference after sugarbeet emergence from a mixed density of wild oat and wild mustard increased. Sucrose content of sugarbeet was not altered by competition. Based on regression analysis, the minimum time that a mixed density of 0.8 wild mustard and 1 wild oat/m of row can interfere with sugarbeet before causing an economic root yield loss was approximately 1.6 weeks after sugarbeet emergence.

Metzger, J. D. (1983). "Role of Endogenous Plant-Growth Regulators in Seed Dormancy of Avena-Fatua .2. Gibberellins." Plant Physiology 73(3): 791-795.://A1983RR85800051

Metzger, J. D. and D. K. Sebesta (1982). "Role of Endogenous Growth-Regulators in Seed Dormancy of Avena-Fatua .1. Short Chain Fatty-Acids." Plant Physiology 70(5): 1480-1485.://A1982PR21900044

Mickelson, J. A. and W. E. Grey (2006). "Effect of soil water content on wild oat (Avena fatua) seed mortality and seedling emergence." Weed Science 54(2): 255-262.://000236728600009 Field experiments were established in fall 1999 and 2000 near Huntley, MT to determine the effects of soil water content on wild oat seed mortality and seedling emergence. Four supplemental irrigation treatments were implemented from June through September to establish plots with varying soil water content. Wild oat seed mortality during the summer increased linearly as soil water content increased. For seed banks established in 1999 (1999SB), seed mortality increased, on 42

average, from 36 to 55% in 2000, and 15 to 55% in 2001 as soil water content increased from 6 to 24%. For seed banks established in 2000 (2000SB), seed mortality increased, on average, from 38 to 88% in 2001 and 53 to 79% in 2002 as soil water content increased from 6 to 24%. Increasing soil water content likely increased the activity of microorganisms chat cause mortality in wild oat seeds. The increasing seed mortality rates (due to increasing soil water content) resulted in greater annual declines of wild oat seed banks and 2-yr cumulative decline rates. Total season emergence percentage was not affected by irrigation treatment. Results show that weed seed bank decline is more rapid in moist than in dry soils and suggest that management practices that increase or conserve soil moisture will also increase the rate of wild oat seed bank decline.

Mickelson, J. A. and R. N. Stougaard (2003). "Assessment of soil sampling methods to estimate wild oat (Avena fatua) seed bank populations." Weed Science 51(2): 226-230.://000181737400014 Accurate and precise estimation of weed seed bank populations is critical to studying to (1) examine the spatial distribution of wild oat seed banks on a small scale (1 m(2) plots), (2) compare wild oat seed bank density sample means and precision between two soil samplers, and (3) predict the sample area needed to quantify a range of wild oat seed bank densities at several levels of precision. Seed bank sample means obtained with a large sampler (10- by 10-cm box) were greater than means obtained with a small sampler (4.4- or 3.8-cm-diam cylinder) for 15 of 18 seed banks. There was no clear advantage in precision (SE/mean) when sampling seed banks using a large number of small soil samples rather than using a small number of large soil samples. Furthermore, at very low seed bank densities, using a small number of large samples gave better precision. Precision improved as sample number increased for each seed bank at each site-year. High-density seed banks tended to have better precision than low-density seed banks at any given sample number. Seed banks had an aggregated spatial distribution when sampled with either soil sampler. As the desired precision level decreases (becomes more precise), the predicted sample area required increases greatly. A seed bank containing 6,000 seeds m(-2) has a predicted sample area of 0.5, 0.7, 1.3, and 2.9% of the total area to obtain precision levels of 0.5, 0.4, 0.3, and 0.2, respectively.

Miller, R. D. and R. W. Allard (1976). "Additional Patterns of Genetic Differentiation in Avena-Barbata in California." Genetics 83(3): S50-S50.://A1976BZ03400142

Miller, S. D. and J. D. Nalewaja (1980). "Wild Oat (Avena-Fatua) Control with Fall-Applied and Spring-Applied Triallate." Weed Science 28(4): 416-418.://A1980KA96500010

Miller, S. D. and J. D. Nalewaja (1990). "Influence of Burial Depth on Wild Oats (Avena-Fatua) Seed Longevity." Weed Technology 4(3): 514-517.://A1990ED29500012

Miller, S. D., J. D. Nalewaja, et al. (1985). "Wild Oats Seed Longevity and Production." North Dakota Farm Research 43(1): 15- 18.://A1985APU3000004

Miller, S. D., J. D. Nalewaja, et al. (1979). "Wild Oat (Avena-Fatua) Control with Flufenprop-Methyl." Weed Science 27(1): 91- 95.://A1979GJ48700022

Miller, S. D., J. D. Nalewaja, et al. (1978). "Barban-Aqueous Nitrogen Combinations for Wild Oat (Avena-Fatua) Control." Weed Science 26(4): 344-348.://A1978FE29500007

Miller, S. D., J. D. Nalewaja, et al. (1978). "Difenzoquat for Wild Oat (Avena-Fatua) Control." Weed Science 26(6): 571-576.://A1978GB53000013

Mohamed, M. A., T. M. Mohamed, et al. (2000). "Distribution of lipases in the Gramineae. Partial purification and characterization of esterase from Avena fatua." Bioresource Technology 73(3): 227-234.://000086383600005 The activity levels of esterase, lipid acylhydrolase and lipase were quantitatively screened in 23 species and cultivars of Gramineae. Their activity levels expressed as units g(-1) seeds, were found to range from 10 to 123 for esterase, 0.28 to 7.67 for lipid acylhydrolase and 13.1 to 93.9 for lipase. Avena fatua, one of the grass species, exhibited the highest levels of esterase and lipase and could be potentially a good starting material for preparation of lipases. A. fatua esterase has been partially purified and characterized. Four isoenzymes, EI, EII, EIII and EIV, were separated by ion exchange chromatography. Esterases EII and EIII had Km values of 0.52 and 0.38 mM and a pH optimum at 9.0 with half maximal activities at pHs 8.5, 10 and 8, 10.5, respectively. Esterases EII and EIII had optimum activities at temperatures of 75 degrees C and 65 degrees C with activation energies of 3.3 and 4.3 kcal mol(-1), respectively. The enzymes were thermally stable as esterases EII and EIII retaining 39% and 23% of their activities at 90 degrees C, respectively. Esterases EII and EIII were stimulated by Ba2+ and Ca2+ but were inhibited by Mn2+ and Zn2+. A. fatua esterases exhibited optimum storage stability and were stable at high temperatures and alkaline pH. They possessed high affinity toward substrate and were resistant to inhibition by most divalent cations that were examined. These are important properties when considering the industrial application of these enzymes. (C) 2000 Elsevier Science Ltd. All rights reserved.

Molberg, E. S., E. V. McCurdy, et al. (1964). "Placement of Di-Allate + Tri-Allate for Control of Wild Oats in Wheat." Canadian Journal of Plant Science 44(4): 351-&.://A19648141A00006

Morikawa, T. (1989). "Genetic-Analysis on Dwarfness of Wild Oats, Avena-Fatua." Japanese Journal of Genetics 64(5): 363- 371.://A1989CB38500004

Morishita, D. W. and D. C. Thill (1988). "Factors of Wild Oat (Avena-Fatua) Interference on Spring Barley (Hordeum-Vulgare) Growth and Yield." Weed Science 36(1): 37-42.://A1988M231300008

43

Morishita, D. W. and D. C. Thill (1988). "Wild Oat (Avena-Fatua) and Spring Barley (Hordeum-Vulgare) Growth and Development in Monoculture and Mixed Culture." Weed Science 36(1): 43-48.://A1988M231300009

Morishita, D. W., D. C. Thill, et al. (1991). "Wild Oat (Avena-Fatua) and Spring Barley (Hordeum-Vulgare) Interference in a Greenhouse Experiment." Weed Science 39(2): 149-153.://A1991FY47400003 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.

Morrison, I. N. and L. Dushnicky (1982). "Structure of the Covering Layers of the Wild Oat (Avena-Fatua) Caryopsis." Weed Science 30(4): 352-359.://A1982NY77300006

Morrison, I. N., B. D. Hill, et al. (1979). "Histological Studies on the Effects of Benzoylprop Ethyl and Flamprop Methyl on Growth and Development of Wild Oats." Weed Research 19(6): 385-393.://A1979JM05100009

Morrison, I. N., M. G. Owino, et al. (1981). "Effects of Diclofop on Growth, Mitotic Index, and Structure of Wheat (Triticum- Aestivum) and Wild Oat (Avena-Fatua) Adventitious Roots." Weed Science 29(4): 426-432.://A1981MB15500014

Morrow, L. A. and D. R. Gealy (1983). "Growth-Characteristics of Wild Oat (Avena-Fatua) in the Pacific Northwest." Weed Science 31(2): 226-229.://A1983QH14300015

Mortensen, K. and A. I. Hsiao (1987). "Fungal Infestation of Seeds from 7 Populations of Wild Oats (Avena-Fatua L) with Different Dormancy and Viability Characteristics." Weed Research 27(4): 297-304.://A1987J300800009

Moyer, J. R. (1979). "Soil Organic-Matter, Moisture, and Temperature - Effect on Wild Oats Control with Trifluralin." Canadian Journal of Plant Science 59(3): 763-768.://A1979HJ32100024

Moyer, J. R. and R. D. Dryden (1977). "Effects of Combined Applications of Triallate or Trifluralin with Solution Nitrogen on Wheat, Wild Oats and Green Foxtail." Canadian Journal of Plant Science 57(2): 479-484.://A1977DE07400023

Moyer, J. R. and R. D. Dryden (1979). "Wild Oats, Green Foxtail, and Broad-Leaved Weeds - Control and Effect on Corn Yield at Brandon, Manitoba." Canadian Journal of Plant Science 59(2): 383-389.://A1979GW71100015

Moyer, J. R., R. D. Dryden, et al. (1979). "Effect of Barban and Flamprop Methyl with Solution Nitrogen on Wheat, Wild Oats and Green Foxtail." Canadian Journal of Plant Science 59(2): 351-356.://A1979GW71100011

Murphy, J. P. and T. D. Phillips (1993). "Isozyme Variation in Cultivated Oat and Its Progenitor Species, Avena-Sterilis L." Crop Science 33(6): 1366-1372.://A1993MR71200048 Effective identification of wild accessions with potential to enhance variation for complex, low heritability traits is a prerequisite to broader utilization of conserved genetic resources. In two previous studies, 23 enzyme systems were assayed in 405 oat cultivars (Avena sativa L. and A. byzantina C. Koch) and in 1005 accessions of the progenitor species. A. sterilis L. The objectives of the present report were to (i) compare isozymic variation in cultivated oat with a broad geographical sample of accessions of the progenitor species and (ii) propose a strategy to assist in the efficient sampling of progenitor germplasm by North American oat breeders. Avena sterilis displayed a greater level of isozymic diversity compared to cultivated germplasm based upon number and frequencies of variants. Three sampling strategies are discussed whereby a representative core of A. sterilis accessions could be selected from the progenitor germplasm pool. A combined strategy is outlined that incorporates elements of all three, with selection of accessions from (i) the center of isozymic diversity (Turkey), (ii) six clusters of A. sterilis accessions identified by multivariate analysis of genetic distances between accessions without regard to provenance data, and (iii) those accessions with variants present at intermediate to high frequencies in A. sterilis from individual countries or clusters yet absent in cultivated germplasm. Selected A. sterilis accessions could be used in combining ability analyses with cultivated germplasm. Subsequent, more extensive, exploitation of the germplasm collection might be based on results from these exploratory evaluations of breeding potential.

Murray, B. G., A. L. BruleBabel, et al. (1996). "Two distinct alleles encode for acetyl-CoA carboxylase inhibitor resistance in wild oat (Avena fatua)." Weed Science 44(3): 476-481.://A1996VE21000005 The objectives of this study were to determine the inheritance of aryloxyphenoxypropionate (APP) resistance in the wild oat population UM33 and to determine the genetic relationship between resistance in UM33 and another population, UM1, which has a different cross-resistance pattern. Reciprocal crosses were made between UM33 and a susceptible population UM5, and between UM33 and UM1. Initial screenings of F-1 and F-2 populations derived from crosses between UM33 and UM5 were conducted over a range of fenoxaprop-P rates to determine a discriminatory dosage. F-2 populations and F-2-derived F-3 families were then screened at this dosage (1200 g ai ha(-1)) to determine segregation patterns. Results from reciprocal UM33 x UM5 F-1 dose-response experiments, and F-2 and F-2-derived F-3 segregation experiments indicated that UM33 resistance to fenoxaprop-P was governed by a single, partially dominant nuclear gene system. To determine if resistance in UM1 and UM33 results from alterations at the same gene locus, 584 F-2 plants derived from reciprocal UM33 x UM1 crosses were screened with 150 g ha(-1) fenoxaprop-P. This dosage was sufficient to kill susceptible plants (UM5), but was not sufficient to kill plants with a resistance allele from either parent. None of the treated F-2 plants exhibited injury or death, indicating that UM1 and UM33 resistance genes did not segregate independently. From this it was concluded that resistance in both populations is encoded at the same gene locus. 44

Murray, B. G., L. F. Friesen, et al. (1996). "A seed bioassay to identify acetyl-CoA carboxylase inhibitor resistant wild oat (Avena fatua) populations." Weed Technology 10(1): 85-89.://A1996UF10300013 A seed bioassay was developed and tested for the rapid identification of aryloxyphenoxypropionate (APP) and cyclohexanedione (CHD) resistance in wild oat. Two susceptible (S) genotypes, UM5 and Dumont, were treated with fenoxaprop-P and sethoxydim over a range of dosages on filter paper and agar. The former is a wild oat line and the latter a tame oat cultivar. Within 5 d, shoot and root development of both genotypes were completely inhibited by 10 mu M fenoxaprop-P and 5 mu M sethoxydim. These dosages were then tested to determine if they were suitable for distinguishing between resistant (R) and susceptible (S) plants, Agar medium was preferred over filter paper because of the ease of preparation and maintenance. Four known R wild oat populations were included in the tests. Those with high levels of resistance produced significantly longer coleoptiles and roots than S genotypes, but those with moderate or low levels of resistance could not be separated statistically from S biotypes based on quantitative measurements. However, after exposing the germinating, treated seeds to light for 24 to 48 h, all the R populations produced green coleoptiles and initiated a first leaf, unlike the S genotypes which did not turn green or produce any new growth. This procedure proved useful in discriminating between R and S genotypes and in ranking populations in terms of relative levels of resistance.

Murray, B. G., I. N. Morrison, et al. (1995). "Inheritance of Acetyl-Coa Carboxylase Inhibitor Resistance in Wild Oat (Avena-Fatua)." Weed Science 43(2): 233-238.://A1995RB70200013 Resistance to fenoxaprop-P and other aryloxyphenoxypropionate and cyclohexanedione herbicides in the wild oat population, UM1, is controlled by a single, partially dominant, nuclear gene, In arriving at this conclusion, parents, F-1 hybrids, and F-2 plants derived from reciprocal crosses between UM1 and a susceptible wild oat line, UM5, were treated with fenoxaprop-P over a wide range of dosages, Based on these experiments, a dosage of 400 g al ha(-1) fenoxaprop-P was selected to discriminate between three response types, At this dosage, susceptible plants were killed and resistant plants were unaffected, whereas plants characterized as intermediate in response were injured but recovered, Treated F-2 plants segregated in a 1:2:1 (R, I, S) ratio, indicative of single nuclear gene inheritance, This was confirmed by selfing F-2 plants and screening several F-3 families, Families derived from intermediate F-2 plants segregated for the three characteristic response types, whereas those derived from resistant F-2 plants were uniformly resistant, Chi-square analysis indicated the F-2 Segregation ratios fit those expected for a single partially dominant nuclear gene system, In addition, F-2 populations from both crosses were screened with a mixture of fenoxaprop-P and sethoxydim, The dosages of both herbicides (150 g al ha(-1) fenoxaprop-P and 100 g ha(-1) sethoxydim) were sufficient to control only susceptible plants. Treated F-2 populations segregated in a 3:1 (R:S) pattern, thereby confirming that resistance to the two chemically unrelated herbicides results from the same gene alteration.

Myers, S. P., M. E. Foley, et al. (1997). "Developmental differences between germinating after-ripened and dormant excised Avena fatua L embryos." Annals of Botany 79(1): 19-23.://A1997WE46100004 Based on physiological and molecular differences associated with the germination of after-ripened and dormant caryopses and excised embryos, it has been hypothesized that various methods of after-ripening are the only treatments that facilitate the transition of dormant wild oat embryos to a non-dormant state. To further investigate this hypothesis, analytical methods were used to evaluate physical and temporal changes associated with germination and subsequent growth of after-ripened and dormant excised embryos (AR-embryos and D-embryos, respectively) induced to germinate with fructose (Fru) and/or gibberellic acid (GA). While chemical treatments of Fru, GA, and Fru + GA have little effect on the germination and short-term growth of AR-embryos, they do induce germination of D-embryos. Growth following germination of D-embryos varied according to treatment with the combination of Fru + GA inducing the greatest growth over the duration of the experiment. Even considering differences in the time to complete germination, growth of D- embryos was not comparable with that of AR-embryos. This provides physical evidence that chemical treatments induce germination without fulfilling the requirements for normal after-ripening-enhanced germination/growth, and indicates that fructose and/or gibberellic acid do not remove the dormancy-block or rate limiting step in the same manner as after- ripening. (C) 1997 Annals of Botany Company

Nandula, V. K. and C. G. Messersmith (2000). "Mechanism of wild oat (Avena fatua L.) resistance to imazamethabenz-methyl." Pesticide Biochemistry and Physiology 68(3): 148-155.://000166132900003 Imazamethabenz-methyl, an acetolactate synthase inhibitor, is used to control wild oat and blackgrass in wheat. Its selectivity is due to differential rates of metabolism to the biologically active imazamethabenz acid from the parent ester in wheat and susceptible species. Imazamethabenz-methyl is a mixture of meta- and para-isomers (2:3), with the meta- isomer being more phytotoxic to wild oat than the para-isomer. Several studies were conducted to characterize the physiological mechanism of resistance in a wild oat biotype resistant to imazamethabenz-methyl. Dose responses in greenhouse experiments indicated an ED50 of 0.5 kp ai/ha for the susceptible wild oat biotype and 4.0 kg/ha for the resistant biotype. Acetolactate synthase enzyme extracts from the susceptible wild oat, resistant wild oat, and wheat were equally sensitive to imazamethabenz acid, with I-50 values of 2.3 x 10(-7) M, 2.5 x 10(-7) M, and 3.3 x 10(-7) M, respectively. The meta-isomer was absorbed better than the para-isomer in both resistant and susceptible biotypes, and there were only minor differences in absorption patterns of the same isomer between biotypes. Increased translocation of [C-14] imazamethabenz was also observed in the susceptible biotype compared to the resistant biotype. Finally, more meta-isomer was metabolized to the acid form to a greater extent in the susceptible than in the resistant biotype. Apparently, the primary mechanism of resistance to imazamethabenz-methyl in wild oat is due to reduced metabolism of imazamethabenz-methyl to the biologically active imazamethabenz acid and is not due to an altered target site. (C) 2000 Academic Press.

Nandula, V. K. and C. G. Messersmith (2001). "Resistance to BAY MKH 6562 in wild oat (Avena fatua)." Weed Technology 15(2): 343-347.://000175240800025 BAY MKH 6562 [flucarbazone- sodium (proposed)], an acetolactate synthase (ALS)-inhibiting herbicide of the sulfonylaminocarbonyltriazolinone family, provides postemergence wild oat control in wheat. Whole-plant dose responses and in vitro ALS sensitivity assays were used to evaluate the magnitude and nature of cross-resistance to BAY MKH 6562 45

in a wild oat accession (ARI) with metabolism-based resistance to imazamethabenz, an ALS inhibitor of the imidazolinone family. An imazamethabenz- susceptible wild oat accession (AHS2), five BAY MKE 6562-resistant wild oat accessions, AN104, AN205, AN307, AN406, and ASBII, and wheat were also evaluated. AHS2 and ARI dose responses to BAY MKH 6562 indicated a resistant/susceptible (R/S) herbicide dose required to cause 50% growth reduction (GR(50)) ratio of 200. Inhibition of ALS from the AHS2 and ARI wild oat by BAY MKH 6562 was similar, with a concentration of herbicide required to cause 50% inhibition of enzymatic activity (I-50) of 0.007 mumoles, indicating that cross-resistance was not due to an altered ALS enzyme. The GR(50) for BAY MKH 6562 for the AN104, AN205, AN307, AN406, and ASB I l wild oat accessions was 0.23, 0.07, 0.23, 0.22, and 0.12 kg ai/ha, respectively, and the R/S ratio to the GR(50) value for the AHS2 accession was 230, 70, 230, 220, and 120, respectively. Studies on ALS sensitivity to BAY MKH 6562 indicated that the I-50 for the AN 104, AN205, AN307, AN406, and ASB I I wild oat accessions was 5.2, 0.003, 0.008, 9.8, and 0.007 mumoles, respectively, and the R/S ratio to the I-50 value for the AHS2 accession was 759, 0.5, 1, 1,444, and 1, respectively. Of the five wild oat accessions resistant to BAY MKH 6562, accessions AN 104 and AN406 had high R/S I-50 ratios indicative of an altered target site and accessions AN205, AN307, and ARI had low R/S I-50 ratios indicative of resistance based on metabolic degradation. Hard red spring wheat (2371) was 800-fold tolerant to BAY MKH 6562 and inhibition of ALS from wheat by BAY MKH 6562 was similar to that of ALS from the susceptible accession AHS2.

Nandula, V. K. and C. G. Messersmith (2002). "Imazamethabenz-resistant wild oat (Avena fatua L.) is resistant to diclofop-methyl." Pesticide Biochemistry and Physiology 74(2): 53-61.://000181368600001 An imazamethabenz-resistant wild oat accession AR1 was found to be resistant to diclofop-methyl. Experiments were conducted to determine the response of this accession to other acetyl-CoA-carboxylase-inhibiting herbicides and the physiological mechanism of its resistance to diclofop-methyl. Diclofop-methyl dose responses in greenhouse experiments indicated an ED50 of 1.04 kg ha(-1) for the AHS2 susceptible accession and 43.8 kg ha(-1) for the AR1 resistant accession with a AR1/AHS2 ED50 ratio of 42. The diclofopmethyl- resistant AR1 accession was susceptible, i.e., not cross resistant, to both aryloxyphenoxypropionate herbicides, fenoxaprop-P and clodinafop, and cyclohexanedione herbicides, sethoxydim, clethodim and tralkoxydim. Diclofop similarly inhibited the acetyl-CoA carboxylase from the AR1 (I-50 = 31 muM) and AHS2 (I-50 = 35 muM) accessions. Absorption of [C-14]diclofop-methyl was similar for the two accessions for at least the first 12 h after treatment (HAT), but was greater by the AHS2 than the AR1 accession at 24, 72, and 168 HAT. Metabolism of [C-14]diclofop-methyl to diclofop was similar in AR1 and AHS2 wild oat at 24 and 72 HAT. Therefore, the mechanism of resistance to diclofop-methyl in the AR1 wild oat accession is not due to an altered ACCase enzyme nor due to differential metabolic activation of diclofop-methyl to diclofop. Difference in the pattern of absorption of [C- 14]diclofop-methyl between AR1 and AHS2 accessions may have a minor role. The role of other possible mechanisms of resistance is discussed. (C) 2002 Elsevier Science (USA). All rights reserved.

Navarrete, L. and C. F. Quintanilla (1996). "The influence of crop rotation and soil tillage on seed population dynamics of Avena sterilis ssp ludoviciana." Weed Research 36(2): 123-131.://A1996UH10900003 The dynamics of seed populations of Avena sterilis ssp. ludoviciana (Durieu) Nyman in plots maintained under different crop rotations and tillage systems was studied over a 5-year period. The seed reserves buried in the soil were practically depleted during this period by using any of the three cropping systems evaluated. No significant differences were found between using a continuous barley rotation with annual application of herbicides and using a barley:fallow rotation with herbicides applied only as needed. The decline in seed populations was not affected by the tillage system used during the fallow period (ploughing or no tillage). The tillage practices used in the various treatments, combined with the variable seed output from each treatment, resulted in different vertical distribution of the seeds in the soil profile. Shallow cultivation resulted in an accumulation of seeds in the upper soil layers; in contrast, a large proportion of the seeds were buried and maintained below 15 cm in the ploughed treatments. Although the seed distribution patterns produced by the various tillage systems had an effect on seedling recruitment, the largest changes in this parameter were associated with the different cropping sequences. Four times more seedlings were recruited in years under barley cropping than in those under fallow.

Naylor, J. M. (1966). "Dormancy Studies in Seed of Avena Fatua .5. on Response of Aleurone Cells to Gibberellic Acid." Canadian Journal of Botany 44(1): 19-&.://A19667243200003

Naylor, J. M., S. N. Acharya, et al. (1979). "Genetic-Basis of Dormancy in Seed of Avena-Fatua L." Genetics 91(4): S88-S88.://A1979HF94700208

Naylor, J. M. and P. Fedec (1978). "Dormancy Studies in Seed of Avena-Fatua .8. Genetic Diversity Affecting Response to Temperature." Canadian Journal of Botany-Revue Canadienne De Botanique 56(18): 2224-2229.://A1978FT17200006

Naylor, J. M. and S. Jana (1975). "Studies on Genetic Adaptation in Wild Oats." Canadian Journal of Genetics and Cytology 17(3): 464-464.://A1975AZ75600058

Naylor, J. M. and S. Jana (1976). "Genetic Adaptation for Seed Dormancy in Avena-Fatua." Canadian Journal of Botany-Revue Canadienne De Botanique 54(3-4): 306-312.://A1976BJ52400012

Naylor, J. M. and G. M. Simpson (1961). "Bioassay of Gibberellic Acid Using Excised Embryos of Avena Fatua L." Nature 192(480): 679-&.://A19618749B00625

Nichols, M. B., S. P. Meyers, et al. (1995). "Physical Characterization of Germination in Dormant and Afterripened Wild Oat (Avena- Fatua) Caryopses." Plant Physiology 108(2): 97-97.://A1995RE28900511

Nishiyama, I. (1981). "Trisomic Analysis of Genome-B in Avena-Barbata Pott (Aabb)." Japanese Journal of Genetics 56(2): 185- 192.://A1981LN20700007

46

Odonovan, J. T. (1988). "Wild Oat (Avena-Fatua) Infestations and Economic Returns as Influenced by Frequency of Control." Weed Technology 2(4): 495-498.://A1988R109400019

O'Donovan, J. T., M. N. Baig, et al. (2000). "Persistence of herbicide resistance in wild oat (Avena fatua L.) populations after cessation of triallate or difenzoquat use." Canadian Journal of Plant Science 80(2): 451-453.://000087505500032 In experimental plots, the proportion of resistant (R) to susceptible (S) wild oat (Avena fatua L.) increased between 1992 and 1996 and was highest in continuous wheat (2.4:1) and lowest in continuous barley (1.2:1). Conversely, the relative frequency of R plants decreased in most farmers' fields between 1990 and 1997 and varied from 12 to 100%. Frequency of resistance was positively correlated with the number of triallate or difenzoquat applications since 1988.

O'Donovan, J. T., R. E. Blackshaw, et al. (2006). "Wheat seeding rate influences herbicide performance in wild oat (Avena fatua L.)." Agronomy Journal 98(3): 815-822.://000237867600048 Field experiments were conducted at three locations in Alberta for 3 yr to determine if spring wheat (Triticum aestimin L.) seeding rate (75 and 150 kg ha(-1)) influenced the effects of recommended and reduced herbicide rates on wild oat (Avena fatua L.) shoot biomass, wild oat seed in the soil seed bank, and wheat yield and net economic return. Wild oat biomass and seed in the soil seed bank decreased nonlinearly at both seeding rates as herbicide rates increased. The herbicides were more effective in reducing wild oat shoot biomass and seed in the soil seed bank when wheat was seeded at the higher rate. The lowest wheat yields and net economic returns occurred when no herbicides were applied and both variables increased nonlinearly with increasing herbicide rate. In most cases, wheat yield and net economic return were greater at the higher seeding rate. On average, wheat yield improved by 19% and net economic return by 16% when wheat was seeded at the higher rate. The results indicate that seeding wheat at relatively high rates can contribute positively to herbicide performance and result in better wild oat management and higher wheat yields and economic returns. In some cases, there was little difference between applying the herbicides at 75 or 100% of the recommended rate but reducing rates below 75% almost always resulted in higher wild oat shoot biomass and seed, and reduced yields and net economic returns, even at the higher wheat seeding rate.

Odonovan, J. T., E. A. Destremy, et al. (1985). "Influence of the Relative-Time of Emergence of Wild Oat (Avena-Fatua) on Yield Loss of Barley (Hordeum-Vulgare) and Wheat (Triticum-Aestivum)." Weed Science 33(4): 498-503.://A1985AMN8900015

O'Donovan, J. T., K. N. Harker, et al. (2003). "Effects of variable tralkoxydim rates on wild oat (Avena fatua) seed production, wheat (Triticum aestivum) yield, and economic return." Weed Technology 17(1): 149-156.://000181848500022 Field experiments were conducted at Lacombe, Lethbridge, and Vegreville, Alberta, Canada and Kalispell, MO, over several years to determine the influence of recommended (minimum label) and lower-than-recommended tralkoxydim rates on wild oat seed production, spring wheat yield, and economic return. Wild oat seed production as a function of tralkoxydim rate varied considerably among locations and years. For example, at the recommended rate, wild oat seed production varied from none at both Lethbridge and Vegreville in 1994 to over 800 seeds/m(2) at Vegreville in 1995. At 50% of recommended rate, seed production varied from none at Lethbridge in 1994 to over 3,000 seeds/m(2) at Vegreville in 1995. In most cases, wheat yield response to tralkoxydim rate was curvilinear. Yields generally increased exponentially as rates increased up to about 40 or 50% of the recommended rate, but then plateaued as rates were increased further. In some cases economic returns tended to plateau or decrease at rates higher than this, but reductions in economic returns at the recommended herbicide rates were, in most cases, relatively slight. In contrast, at Lethbridge in 1993 and 1995 and at Kalispell in 1994 and 1996, yield and economic returns generally increased as herbicide rate increased, and there was an economic disadvantage to reducing the tralkoxydim rate below that recommended. In view of the variable effects on wild oat seed production, and the questionable economic benefit, our study suggests that reducing the rate of tralkoxydim below that recommended may not be without risk. Further studies are necessary to determine the long-term implications of returning relatively large amounts of wild oat seed to the soil seedbank at reduced herbicide rates.

O'Donovan, J. T., K. N. Harker, et al. (2003). "Influence of variable rates of imazamethabenz and difenzoquat on wild oat (Avena fatua) seed production, and wheat (Triticum aestivum) yield and profitability." Canadian Journal of Plant Science 83(4): 977-985.://000188252400045 Field experiments to investigate the effects of variable imazamethabenz rates on wild oat seed production and wheat yield and profitability were conducted at Lacombe, Lethbridge and Vegreville, Alberta, and Kalispell, Montana, over several years. Similar studies with difenzoquat were conducted at Lacombe and Lethbridge. In most cases, reducing the herbicide rates below those recommended resulted in increases in wild oat seed production, but the potential for returning relatively large amounts of wild oat seed to the soil seedbank depended on the extent of the rate reduction. For example, averaged over locations and years, reducing the rate of imazamethabenz to 75% of the recommended rate resulted in wild oat seed production increasing by 25% compared with an increase of over 100% when the rate was reduced to 50%. Wheat yields and economic returns as functions of rate also varied for both herbicides. It was more economical, in most cases, to apply imazamethabenz at 50 or 75% of the recommended rate compared with the full rate. However, an economic loss occurred in four and three of the 11 location-years when the imazamethabenz rate was reduced to 50 and 75%, respectively, and losses were more severe at the 50% rate. Compared with imazamethabenz, reducing the rate of difenzoquat tended to be more risky in terms of increased wild oat seed production and reduced net economic return.

O'Donovan, J. T., K. N. Harker, et al. (2000). "Wild oat (Avena fatua) interference in barley (Hordeum vulgare) is influenced by barley variety and seeding rate." Weed Technology 14(3): 624-629.://000166693600026 Field experiments were conducted at Vegreville and Lacombe, AB, to determine the influence of barley (Hordeum vulgare) variety and seeding rate on interference of wild oat (Avena fatua) with barley. Barley variety and seeding rate affected barley density, height at maturity, and seed yield, as well as wild oat shoot dry weight and seed yield in most experiments, but there was no variety by seeding rate interaction. As expected, the semidwarf varieties Falcon and CDC Earl were the shortest. Barley seedling emergence and subsequent plant densities varied among varieties, locations, and years. The hull-less varieties Falcon and CDC Dawn had the poorest emergence in most cases, whereas AC Lacombe and Seebe had the highest emergence. Wild oat shoot dry matter and seed production was highest in the Falcon, CDC Dawn, and 47

CDC Earl plots, suggesting that these were the least competitive with wild oat. Barley yield loss from wild oat interference also tended to be highest in these varieties. Poor emergence of Falcon and CDC Dawn and the shorter stature of Falcon and CDC Earl likely contributed to their relatively poor competitiveness with wild oat. Increasing the seeding rate improved the competitiveness of all varieties, as evidenced by reduced wild oat shoot dry matter and seed production and, in some cases, improved barley yields.

Odonovan, J. T., A. Rashid, et al. (1996). "A seedling bioassay for assessing the response of wild oat (Avena fatua) populations to triallate." Weed Technology 10(4): 931-935.://A1996WD96900039 Germinated seeds of wild oat populations that were susceptible (S) or resistant (R) to triallate at the recommended soil- applied rate (1.7 kg/ha) were treated with six triallate concentrations on filter paper in petri dishes. Measurement of shoot length 8 d after treatment provided an accurate indication of differences among populations, and was more reliable than determining shoot fresh weight. ED(50) values (herbicide concentrations that reduced shoot length by 50% relative to untreated controls), derived from nonlinear regression analysis, indicated four and five levels of response to triallate among eight S and seven R populations, respectively, The ED(50) values varied from 0.11 to 11 ppm a.i. triallate for the most susceptible to the most resistant populations, respectively. Routine testing of wild oat samples suspected of resistance, at triallate concentrations of 0.5 or 1 ppm in the petri dish bioassay, effectively identified populations that had become resistant to the recommended soil-applied rate.

Odonovan, J. T., M. P. Sharma, et al. (1994). "Wild Oat (Avena-Fatua) Populations Resistant to Triallate Are Also Resistant to Difenzoquat." Weed Science 42(2): 195-199.://A1994NR43600008 In response to farmer complaints of poor triallate performance, wild oat seed was collected from 34 fields in Alberta in the fall of 1990. Screening trials in the greenhouse indicated that 15 of the populations were highly resistant to triallate applied at the equivalent of the recommended field rate (1.7 kg ha-1), whereas the other 19 populations were adequately controlled. All triallate-resistant populations were also highly resistant to difenzoquat applied at 1.7 kg ha-1 (equivalent to twice the recommended field rate). The effect of increasing rates of both herbicides on dry weight of five of the resistant and two of the susceptible populations was determined in greenhouse experiments. Triallate applied up to 3.4 kg ha-1 had little or no effect on the resistant populations, whereas the susceptible populations were controlled at 1.7 kg ha-1. At rates of 6.8 kg ha-1 or higher, there were differences among the resistant populations and among individuals within the populations in the response to triallate. Response of the resistant populations to increasing difenzoquat rates was variable between experiments, but in all cases the effect of difenzoquat on wild oat dry weight was considerably less in triallate- resistant than triallate-susceptible populations. Effects of increasing rates of triallate and difenzoquat on resistant and susceptible wild oat populations growing with barley in field experiments were generally similar to the responses in the greenhouse.

Ohm, H. W. and Patterso.Fl (1973). "6-Parent Diallel Cross Analysis for Protein in Avena-Sterilis L." Crop Science 13(1): 27-30.://A1973P163900009

Ohm, H. W. and D. M. Peterson (1975). "Protein Composition in Developing Groats of an Avena-Sativa-L Cultivar and an Avena- Sativa X Avena-Sterilis L Selection." Crop Science 15(6): 855-858.://A1975BC16300032

Oleary, N. F., J. T. Odonovan, et al. (1980). "Effect of Diclofop Methyl and 2,4-D Alone and in Combination on Leaf Cell-Membrane Permeability of Wild Oats and Barley." Canadian Journal of Plant Science 60(2): 773-775.://A1980JW35500061

Olson, W. A. and J. D. Nalewaja (1981). "Antagonistic Effects of Mcpa on Wild Oat (Avena-Fatua) Control with Diclofop." Weed Science 29(5): 566-571.://A1981MK02000010

Osullivan, P. A. (1980). "Control of Wild Oats and Tartary Buckwheat with Mixtures of Metribuzin and Various Postemergence Wild Oat Herbicides." Canadian Journal of Plant Science 60(4): 1255-1261.://A1980LB50600023

Osullivan, P. A. (1981). "Control of Avena-Fatua and Fagopyrum-Tataricum with Tank Mixtures of Linuron or Linuron + Mcpa and Sequential Applications of Linuron, and Post-Emergence Avena-Fatua Herbicides." Weed Research 21(5): 211-217.://A1981MP75900003

Osullivan, P. A. (1981). "Control of Wild Oats, Green Foxtail and Tartary Buckwheat with Mixtures of Propanil or Propanil-Mcpa and Postemergence Wild Oat Herbicides." Canadian Journal of Plant Science 61(2): 383-390.://A1981LT55500026

Osullivan, P. A. (1983). "Influence of Picloram Alone or Plus 2,4-D on Control of Wild Oats (Avena-Fatua) with 4 Postemergence Herbicides." Weed Science 31(6): 889-891.://A1983RR71400026

Osullivan, P. A., P. N. P. Chow, et al. (1982). "Control of Green Foxtail in Cereals with Ac 206,784, Alone and in Mixtures with Triallate for Wild Oats." Canadian Journal of Plant Science 62(4): 995-1001.://A1982PY41500023

Osullivan, P. A. and K. J. Kirkland (1984). "Chlorsulfuron Reduced Control of Wild Oat (Avena-Fatua) with Diclofop, Difenzoquat, and Flamprop." Weed Science 32(3): 285-289.://A1984SR03300001

Osullivan, P. A. and K. J. Kirkland (1984). "Control of Avena-Fatua L and Cirsium-Arvense (L) Scop with Mixtures of 3,6- Dichloropicolinic Acid and 4 Herbicides for Control of Avena-Fatua." Weed Research 24(1): 23-28.://A1984SA71000004

Page, E. R., R. S. Gallagher, et al. (2006). "Modeling site-specific wild oat (Avena fatua) emergence across a variable landscape." Weed Science 54(5): 838-846.://000240538100005 The spatial and temporal pattern of wild oat emergence in eastern Washington is affected by the steep, rolling hills that dominate this landscape. The objective of this study was to assess the impact of landscape position and crop residue on 48

the emergence phenology of wild oat. Emergence of a natural wild oat infestation was characterized over two growing seasons (2003 and 2004), at two wheat residue levels (0 and 500 g m(-2)), and at five landscape positions differing in slope, aspect, and elevation in a no-till winter wheat field. Wild oat emerged 1 to 2 wk earlier at south-facing landscape positions than at north-facing landscape positions. Crop residue delayed wild oat emergence by 7 to 13 d relative to bare soil at south-facing positions in 2003 and had a reduced effect on emergence at north-facing landscape positions. Therefore, preserving surface residues tended to synchronize emergence across the landscape and may facilitate better timing of weed control where residue is present. Emergence of wild oat was modeled as a function of thermal time adjusted by water potential using a Weibull function. Temperature explained more variation in the model than water potential. This model explained much of the variability in wild oat emergence among landscape positions over these 2 yr and may be useful as a tool to predict the timing of wild oat emergence. Results also indicate that site-specific modeling is a plausible approach to improving prediction of weed seedling emergence.

Pan, W. P. and E. G. Hammond (1983). "Stereospecific Analysis of Triglycerides of Glycine-Max, Glycine-Soya, Avena-Sativa and Avena-Sterilis Strains." Lipids 18(12): 882-888.://A1983RX45000006

Pandey, S. and R. W. Medd (1990). "Integration of Seed and Plant Kill Tactics for Control of Wild Oats - an Economic-Evaluation." Agricultural Systems 34(1): 65-76.://A1990DQ97700005

Pandey, S. and R. W. Medd (1991). "A Stochastic Dynamic-Programming Framework for Weed-Control Decision-Making - an Application to Avena-Fatua L." Agricultural Economics 6(2): 115-128.://A1991GV94300002 This paper develops a stochastic multi-period decision model to analyse a continuous wheat cropping system infested by wild oats (Avena fatua L.), in southern Australia. The multi-period solutions is obtained by employing a dynamic programming model in conjunction with a bioeconomic simulation model. An empirically estimated dose response function is used to derive the optimal herbicide rate. Uncertainties due to environmental effects on the performance of herbicide and crop yields are modelled and optimal decision rules derived. The results indicate that substantial economic gains can be realised if herbicide dose decisions are taken by considering future profit effects of current decisions, as opposed to the more common approach of only considering the current-period effect.

Pannell, D. J. and G. S. Gill (1994). "Mixtures of Wild Oats (Avena-Fatua) and Ryegrass (Lolium-Rigidum) in Wheat - Competition and Optimal Economic-Control." Crop Protection 13(5): 371-375.://A1994NX67900008 The joint control of wild oats (A vena fatua) and ryegrass (Lolium rigidum) in mixture in a wheat crop using a single postemergence herbicide, diclofop-methyl, was examined. The two weeds were found to have different competitive strengths against wheat and different degrees of sensitivity to the herbicide. Response surfaces to different rates of herbicide and different weed densities were estimated statistically and these functions were used to examine optimal control practices for different mixtures of the weeds.

Paterson, J. G. (1977). "Interaction between Herbicides, Time of Application and Genotype of Wild Oats (Avena-Fatua L)." Australian Journal of Agricultural Research 28(4): 671-680.://A1977DR47300012

Paterson, J. G., W. J. R. Boyd, et al. (1976). "Effect of Temperature and Depth of Burial on Persistence of Seed of Avena-Fatua L in Western-Australia." Journal of Applied Ecology 13(3): 841-847.://A1976CP89100015

Paterson, J. G., W. J. R. Boyd, et al. (1976). "Vernalization and Photoperiod Requirement of Naturalized Avena-Fatua and Avena- Barbata Pott Ex Link in Western Australia." Journal of Applied Ecology 13(1): 265-272.://A1976BP46100020

Paterson, J. G., N. A. Goodchild, et al. (1976). "Effect of Storage Temperature, Storage Duration and Germination Temperature on Dormancy of Seed of Avena-Fatua L and Avena-Barbata Pott Ex Link." Australian Journal of Agricultural Research 27(3): 373- 379.://A1976BU67300005

Pearce, F. (1987). "A Time for Sowing Wild Oats." New Scientist 116(1591): 10-10.://A1987L315300015

Perez, F. J. (1990). "Allelopathic Effect of Hydroxamic Acids from Cereals on Avena-Sativa and Avena-Fatua." Phytochemistry 29(3): 773-776.://A1990CU56000014

Perez, F. J. and J. Ormenonunez (1991). "Root Exudates of Wild Oats - Allelopathic Effect on Spring Wheat." Phytochemistry 30(7): 2199-2202.://A1991FT56300020 Root exudates from the undisturbed root system of wild oats Avena fatua were collected by a modification of the Tang and Young method. Exudates inhibited root and coleoptile growth of spring wheat seedlings (Triticum aestivum). Scopoletin, coumarin, p-hydroxybenzoic and vanillic acid were tentatively identified from the root exudates by HPLC.

Peters, N. C. B. (1982). "The Dormancy of Wild Oat Seed (Avena-Fatua L) from Plants Grown under Various Temperature and Soil- Moisture Conditions." Weed Research 22(4): 205-212.://A1982NZ77900004

Peters, N. C. B. (1982). "Production and Dormancy of Wild Oat (Avena-Fatua) Seed from Plants Grown under Soil Waterstress." Annals of Applied Biology 100(1): 189-196.://A1982MZ20200020

Peters, N. C. B. (1984). "Time of Onset of Competition and Effects of Various Fractions of an Avena-Fatua L Population on Spring Barley." Weed Research 24(5): 305-315.://A1984TL61900001

Peters, N. C. B. (1985). "Competitive Effects of Avena-Fatua L Plants Derived from Seeds of Different Weights." Weed Research 25(1): 67-77.://A1985AAX1900010

49

Peters, N. C. B. (1986). "Factors Affecting Seedling Emergence of Different Strains of Avena-Fatua L." Weed Research 26(1): 29- 38.://A1986AYN5600004

Peters, N. C. B. (1991). "Seed Dormancy and Seedling Emergence Studies in Avena-Fatua L." Weed Research 31(2): 107-116.://A1991FC63400006 Studies were made of seedling emergence of three phenotypes of Avena fatua L. with different coloured lemmas (fA, fB, fC) originally collected from one site. Each phenotype was grown under the same conditions in 1975 and the resulting seed buried 25 mm deep in soil immediately after collection. The soil was either left uncultivated (all phenotypes) or cultivated monthly (phenotype fA only). Seedling emergence was assessed weekly and the number of remaining viable seeds was determined at the end of three years. Seeds of different lines of A. fatua (fA phenotypes) obtained from separate locations were grown under the same conditions in 1983, buried and seedling emergence monitored for 8.5 months. Without cultivation, overall emergence (mean of three phenotypes) in successive autumns and springs was 9 %, 18 %, 9 %, 33 %, < 1 %, 14 % and < 1 %. A further 3 % of viable seeds were recovered at the end of the experiment. Periodicity of emergence was the same for all phenotypes. Actual numbers of seedlings emerging in each of the periods varied between phenotypes. Total emergence of seed from the inner zone of panicles was significantly less than that from the outer zone, although there was little difference between the two zones in each of the natural emergence periods. Cultivation increased emergence, particularly from secondary seed in the first spring after burial, but did not changes its periodicity. By the second spring seed numbers had declined, and seedling counts were similar from cultivated and non- cultivated soil. No viable seed remained in the cultivated soil after three years. Emergence from the two lines of fA was very different. Seedling emergence occurred after hot dry conditions, or in warm periods immediately after periods of chilling, particularly those below 4-degrees-C.

Peters, N. C. B. and B. J. Wilson (1983). "Some Studies on the Competition between Avena-Fatua-L and Spring Barley .2. Variation of Avena-Fatua Emergence and Development and Its Influence on Crop Yield." Weed Research 23(5): 305-311.://A1983RH84200009

Pfaelzer, J. (1989). "The Sentimental Promise and the Utopian Myth, Davis,Rebecca,Harding the 'Harmonists' and Alcott,Louisa,May 'Transcendental Wild Oats'." American Transcendental Quarterly 3(1): 85-99.://A1989U115100006

Pfleeger, T. G. and C. C. Mundt (1998). "Wheat leaf rust severity as affected by plant density and species proportion in simple communities of wheat and wild oats." Phytopathology 88(7): 708-714.://000074422800015 While it is generally accepted that dense stands of plants exacerbate epidemics caused by foliar pathogens, there is little experimental evidence to support this view. We grew model plant communities consisting of wheat and wild oats at different densities and proportions and exposed these communities to Puccinia recondita to induce wheat leaf rust. Wild oats was included because it is a common competitor of wheat and may act as a barrier to the dispersal of P. recondita spores among wheat plants. Disease severity was estimated as percentage of wheat flag leaves covered by rust lesions. Seeding density rarely had a significant influence on rust severity, probably because of compensation due to increased tillering at low seeding densities. In contrast, increasing the proportion of wheat in mixtures with wild oats consistently increased wheat leaf rust severity. Regression parameters describing wheat leaf rust severity as a function of wheat seeding density did not differ significantly between pure wheat stands and wheat-wild oat mixtures and, thus, failed to support an effect of wild oats on wheat leaf rust other than through its competitive impact on wheat tiller density.

Pfleeger, T. G., C. C. Mundt, et al. (1999). "Effects of wheat leaf rust on interactions between wheat and wild oats planted at various densities and proportions." Canadian Journal of Botany-Revue Canadienne De Botanique 77(11): 1669-1683.://000085329100015 The importance of competition as a major influence on the composition and structure of plant communities has recently been questioned, because other types of interactions can cause significant compositional changes. The goal of this research was to broaden our understanding of disease as a process structuring plant communities under a variety of competitive scenarios. Two cultivars of spring wheat (Triticum aestivum L., cv. Twin and cv. Penawawa) and wild oats (Avena fatua L.) were planted at three densities and at five proportions. One-half of the experimental material was inoculated with uredospores of Puccinia recondita. Increasing the proportion of wheat or oats in mixtures led to significant increases in the amount of aboveground biomass and total seed weight for that species. The seed weight and aboveground biomass per culm or per planted seed decreased for wheat and wild oats as the proportion of wild oats increased in mixtures, indicating a competitive advantage for wild oats when grown with wheat. Wild oats generally did not respond significantly to the effects of leaf rust on wheat, while wheat performance declined. Lowered wheat performance in inoculated stands was the main reason for lower relative biomass ratios of wheat to wild oats. Puccinia recondita infections occurred late in the life cycle of wheat, thereby decreasing the potential impact on wild oats' adults through competitive interactions.

Phillips, T. D., J. P. Murphy, et al. (1993). "Isozyme Variation in Germplasm Accessions of the Wild Oat Avena-Sterilis L." Theoretical and Applied Genetics 86(1): 54-64.://A1993KV86200007 Optimal exploitation of crop genetic resources requires a knowledge of the range and structure of the variation present in the gene pool of interest. Avena sterilis L., the cultivated oat progenitor, contains a store of genetic diversity that is readily accessible to the oat breeder. The objectives of the present paper were: (1) to evaluate isozyme polymorphisms in a sample of A. sterilis accessions from the U.S. National Small Grains Collection, (2) to analyze the distribution of isozyme diversity across the geographic range of the accessions, (3) to classify the accessions into groups based on isozyme variation, and (4) to suggest strategies for efficient sampling of this germplasm collection. One thousand and five accessions from 23 countries and 679 collection sites were screened for variation using 23 enzyme systems. Due to limited information about the genetic relationship among individual members of families of isozymes in hexaploid oat species, data were recorded solely for band presence. The frequencies of bands in accessions from the various countries were used to calculate the probability of genotypic identity (I(x.y)), the probability of a unique genotype (U(x.y)), and an adjusted polymorphic index (H(x)). Accessions from Turkey and Lebanon had the largest polymorphic index values, 50

Turkish and Moroccan accessions displayed the greatest numbers of bands. Accessions from Iran, Turkey, Iraq, and Lebanon had the largest mean probabilities of containing unique genotypes. Based on isozyme data, Turkey appeared to represent the center of diversity in this germplasm collection. Band frequencies calculated among countries were used in a principal component analysis. Accessions from Israel and Morocco clustered together; accessions from Iran, Iraq, Turkey, and Ethiopia formed another group; and Algerian accessions formed an outlying group. Several isozyme bands had a regional distribution. These results suggested that choosing accessions from countries based on their groupings in the principal component analysis should secure a greater range of diversity than sampling from the collection at random. Cluster analyses based on Jaccard's distances calculated for all pairwise combinations of the 1005 accessions revealed six broad genetic groups of accessions. Groups 1 and 6 contained accessions from many countries and encompassed half of all accessions. Groups 2 and 4 were heavily populated by accessions from Israel and Morocco. Groups 3 and 5 were composed almost exclusively of accessions from Iran, Iraq, and Turkey. By selecting representative accessions from these six groups, oat breeders could most effectively sample the range of genetic variation in this A. sterilis collection.

Philp, J. (1932). "Fatuoid or false wild oats." Nature 129: 796-796.://000188272600423

Philpotts, H. (1975). "Control of Wild Oats in Wheat by Winter Fallowing and Summer Cropping." Weed Research 15(4): 221- 225.://A1975AH54600004

Pillmoor, J. B. (1985). "Influence of Temperature on the Activity of Ac-222,293 against Avena-Fatua L and Alopecurus-Myosuroides Huds." Weed Research 25(6): 433-442.://A1985AWM9600006

Pillmoor, J. B. and J. C. Caseley (1984). "The Influence of Growth Stage and Foliage or Soil Application on the Activity of Ac 222,293 against Alopecurus-Myosuroides and Avena-Fatua." Annals of Applied Biology 105(3): 517-&.://A1984AFQ5200013

Pillmoor, J. B. and J. C. Caseley (1987). "The Biochemical and Physiological-Effects and Mode of Action of Ac-222,293 against Alopecurus-Myosuroides Huds and Avena-Fatua L." Pesticide Biochemistry and Physiology 27(3): 340-349.://A1987G442900010

Pittman, U. J. (1970). "Magnetotropic Responses in Roots of Wild Oats." Canadian Journal of Plant Science 50(3): 350-&.://A1970G558000020

Polanco, C. and M. P. Delavega (1995). "Length Polymorphism in the Ribosomal DNA Intergenic Spacer of Rye and Slender Wild Oats." Journal of Heredity 86(5): 402-407.://A1995RW53300013 Variation in ribosomal DNA spacer length was analyzed in nine populations of three species of grasses by restriction fragment length polymorphism analysis using non radioactive labeling. Three populations of rye. Secale cereale, four of the tetraploid slender wild oat, Avena barbata, and two of its diploid ancestor A. hirtula were analyzed. Extensive spacer length polymer phism, attributed to the differences in number of spacer subrepeats within rDNA arrays, existed in the three species. The fragment lengths generated by EcoRI in the two oat species ranged from 2.2 to 3.2 kb, while the fragments generated by BamHI in rye ranged from 4.7 to 6.2 kb. The phenotypic variability observed agrees with the expectations from the different mating types and number of rDNA loci of these three species. In the two oats species, the most frequent spacer length variants were shared for both oat species and were present in all the populations; likewise, frequent spacer length variants were present in the three rye populations. The results are compared with previous data on rDNA polymorphism in rye and in slender oats populations and on genetic variability for isozyme loci in the same populations of both oats and rye.

Polley, H. W., H. B. Johnson, et al. (1992). "Growth and Gas-Exchange of Oats (Avena-Sativa) and Wild Mustard (Brassica-Kaber) at Subambient Co2 Concentrations." International Journal of Plant Sciences 153(3): 453-461.://A1992JU77800021 A repeated sequence of monocultures and mixtures of oats (A vena sativa L.) and wild mustard (Brassica kaber (DC.) Wheeler) was grown along a daytime gradient of CO2 concentrations ([CO2]) from near 330 to a minimum of 150 mumol mol-1. The objectives were to determine effects of subambient [CO2] on leaf gas exchange, biomass production, and competitive interactions of these C3 species. A decrease in stomatal conductance did not prevent a nearly linear increase in leaf internal [CO2] and net assimilation of oat leaves as [CO2] increased. Net assimilation of oats and wild mustard increased from 5.0 and 2.5 mumol m-2 s-1 at 150 mumol mol-1, respectively, to 16.1 and 15.9 mumol m-2 s-1 at 330 mumol mol-1 CO2, respectively, when measured at 1,200-1,500 mumol m-2 s-1 incident light. Aboveground biomass per plant of wild mustard and oats increased 106% and 198%, respectively, and leaf area rose more than two- and threefold, respectively, from 154 to 331 mumol mol-1 CO2. The CO2-induced increase in aboveground biomass of plants of each species did not vary among monocultures and mixtures. Responses of oats and wild mustard to higher subambient [CO2] were large relative to reported responses of C3 species to comparable increases above the current atmospheric [CO2]. This suggests that past changes in atmospheric [CO2], including the 27% rise since the beginning of the nineteenth century, may have profoundly altered the productivity of C3 plants.

Pomilio, A. B., S. R. Leicach, et al. (2000). "Constituents of the root exudate of Avena fatua grown under far-infrared-enriched light." Phytochemical Analysis 11(5): 304-308.://000089517700005 The constituents of the root exudate of plantlets of wild oats (Avena fatua) grown in the greenhouse under far-IR-enriched radiation were analysed for the first time by HPTLC and capillary GC, and further separated by medium-pressure liquid chromatography (MPLC; monitored by TLC) and characterised by GC-EIMS and IH-NMR. This combination of methodologies permitted the study of root exudates under controlled conditions and is appropriate for examining a variety of ecophysiological/environmental effects in a short time period. In this instance, the response compounds of the irradiated plantlets were identified as linear and branched alcohols, linear and branched alkanes, a bicyclic monoterpene, mono- and bicyclic sesquiterpenes, and free and esterified fatty acids. An ester, isopropyl myristate, was the main component present, although the unusual occurrence of branched compounds, sesquiterpenoids and related compounds of isoprenoid 51

metabolism is noteworthy. The effect of far-IR-enriched light on the production of these compounds is discussed. Copyright (C) 2000 John Wiley & Sons, Ltd.

Ponce, R. G. (1987). "Competition for N and P between Wheat and Wild Oats (Avena-Sterlis L) According to the Proximity of Their Time of Emergence." Plant and Soil 102(1): 133-136.://A1987J629700023

Ponce, R. G. (1988). "Competition between Avena-Sterilis Ssp Macrocarpa Mo and Cultivars of Wheat." Weed Research 28(5): 303- 307.://A1988Q247700001

Ponce, R. G. and A. Lamela (1987). "Improving the Efficiency of Nitrogen-Fertilizer in Wheat Treated with Herbicide for the Control of Wild Oats." Journal of Plant Nutrition 10(9-16): 1771-1778.://A1987L885600093

Ponce, R. G. and M. L. Salas (1989). "Effects of Different Nitrogen-Compounds and Temperatures on the Germination of Avena- Sterilis Spp Macrocarpa Mo." Biologia Plantarum 31(4): 261-268.://A1989AV54700004

Ponce, R. G. and I. Santin (2001). "Competitive ability of wheat cultivars with wild oats depending on nitrogen fertilization." Agronomie 21(2): 119-125.://000167651700001 In a field experiment, wheat (Triticum aestivum L.) grew with the infesting weed wild oat (Avena sterilis ssp. sterilis L.) in semi-arid conditions. Both species were affected by the drought conditions, although drought caused more damage to the growth and seed production of A. sterilis than of the wheat. Both species benefited from nitrogen fertilization, but much more so in the year when water was not a limiting factor. In comparison with the semi-dwarf cultivar (Anza), the tall wheat cultivar (Pane 247) is competitively superior to A. sterilis. This competitive superiority was related more to the height of the wheat plant and supposed competition for light than to tiller production. The competition affected the production of A. sterilis panicles and spikelets. In addition, the height and competitive ability of the tall wheat cultivar increased more than that of the semi-dwarf cultivar with increasing nitrogen doses. The choice of the wheat cultivars and the use of N doses depending on the climatic conditions is discussed with regard to the control possibilities for A. sterilis.

Prakash, A., K. C. Sharma, et al. (1986). "Control of Phalaris-Minor and Wild Oats in Wheat by Cultural and Chemical Methods." Indian Journal of Agronomy 31(4): 411-413.://A1986K410100029

Price, S. C., J. E. Hill, et al. (1988). "The Morphological and Physiological-Response of Slender Oat (Avena-Barbata) to the Herbicides Barban and Difenzoquat." Weed Science 36(1): 60-69.://A1988M231300012

Quail, P. H. and O. G. Carter (1968). "Survival and Seasonal Germination of Seeds of Avena Fatua and a Ludoviciana." Australian Journal of Agricultural Research 19(5): 721-&.://A1968B992000002

Quail, P. H. and O. G. Carter (1969). "Dormancy in Seeds of Avena Ludoviciana and a Fatua." Australian Journal of Agricultural Research 20(1): 1-&.://A1969C902600001

Quick, W. A. and A. I. Hsiao (1984). "Changes in Inorganic-Phosphate and Seed Germinability During Afterripening of Wild Oats." Canadian Journal of Botany-Revue Canadienne De Botanique 62(11): 2469-2471.://A1984TX11400039

Quick, W. A., A. I. Hsiao, et al. (1997). "Dormancy implications of phosphorus levels in developing caryopses of wild oats (Avena fatua L)." Journal of Plant Growth Regulation 16(1): 27-34.://A1997WP88100005 The element phosphorus made up 0.5% of the dry weight of dehulled Avena fatua caryopses 7 days after anthesis (DAA), half of it inorganic (P-i). Caryopses detached and pierced 7 DAA germinated in vitro with a rapid drop in P-i levels. By 15- 20 DAA caryopsis dry weight had increased three- to fourfold, but phosphorus made up less than 0.04% of the dry weight of this enlarged caryopsis. Caryopses at this stage germinated readily without piercing if incubated in vitro. A further decrease in P-i accompanied by a marked increase in phytate phosphorus began about 15 DAA and continued during later seed maturation. By 20 DAA, when embryos were relatively mature and endosperm cell division had ceased, a decrease in caryopsis water content (as a percentage of dry weight) began, and seed dormancy became apparent. As starch and phytate reserves accumulated, P-i and water levels of the caryopsis diminished. Higher levels of endogenous P-i coincided with the anabolic events of initial seed formation and, to a lesser extent, with anabolic events of seed germination. Decreasing P-i levels coincided with accumulation of nutrient reserves, lowering of water content, and the initiation of dormancy. The data suggest that (1) enzymes associated with the formation and development of the embryo may be activated by the high P-i levels present during initial seed differentiation; (2) embryo quiescence and dormancy are facilitated by the drop of P-i levels which accompanies the accumulation of starch and phytate reserves; and (3) the increase in P-i which accompanies seed afterripening aids in the termination of dormancy and the resumption of germination.

Quick, W. A., A. I. Hsiao, et al. (1983). "Endogenous Inorganic-Phosphate in Relation to Seed Dormancy and Germination of Wild Oats." Plant Science Letters 28(2): 129-135.://A1983PZ00900002

Qureshi, F. A. and W. H. Vandenborn (1979). "Interaction of Diclofop-Methyl and Mcpa on Wild Oats (Avena-Fatua)." Weed Science 27(2): 202-205.://A1979GR79500014

Radics, L., J. Alkamper, et al. (1985). "Studies of Nitrogen Nutrient Competition between Summer Barley and Avena-Fatua L and Sinapsis-Arvensis L, Resp, as Well as between Maize and Echinochloa-Crus-Galli Lpb and Amaranthus-Retroflexus L, Resp in the Period Preceding Weed-Control." Novenytermeles 34(5): 399-408.://A1985AYG8500008

Rai, K. N. (1974). "Mutation Insitu and Origin of Localized Polymorphism in Avena-Barbata." Evolution 28(2): 330-332.://A1974T392500017 52

Rai, K. N. (1985). "Regional Patterns of Polymorphisms in Natural-Populations of Avena-Barbata." Canadian Journal of Genetics and Cytology 27(6): 639-643.://A1985AZA0400003

Rai, K. N. and S. K. Jain (1982). "Population Biology of Avena .9. Gene Flow and Neighborhood Size in Relation to Microgeographic Variation in Avena-Barbata." Oecologia 53(3): 399-405.://A1982NU18500023

Rains, D. W. (1971). "Lead Accumulation by Wild Oats (Avena-Fatua) in a Contaminated Area." Nature 233(5316): 210-&.://A1971K295500043

Rajhathy, T. (1969). "A Re-Examination of Chromosomes of Tetraploid Avena-Sterilis from Sardinia." Canadian Journal of Genetics and Cytology 11(4): 1001-&.://A1969F433200027

Raju, M. V. S. (1984). "Studies on the Inflorescence of Wild Oats (Avena-Fatua) - Morphology and Anatomy of the Awn in Relation to Its Movement." Canadian Journal of Botany-Revue Canadienne De Botanique 62(11): 2237-2247.://A1984TX11400007

Raju, M. V. S. (1985). "Aleurone Cells in Wild Oats (Avena-Fatua-L)." American Journal of Botany 72(6): 827-828.://A1985ALY7500102

Raju, M. V. S. (1986). "Behavior of Aleurone Cells of Wild Oats (Avena-Fatua L) - a Light Microscopic Study." Flora 178(5): 351- 362.://A1986E278600007

Raju, M. V. S. and R. J. Barton (1984). "On Dislodging Caryopses of Wild Oats." Botanical Magazine-Tokyo 97(1045): 127-130.://A1984SN77200010

Raju, M. V. S., A. I. Hsiao, et al. (1986). "Seed Dormancy in Avena-Fatua .3. the Effect of Mechanical Injury on the Growth and Development of the Root and Scutellum." Botanical Gazette 147(4): 443-452.://A1986G182600010

Raju, M. V. S., A. I. Hsiao, et al. (1988). "Seed Dormancy in Avena-Fatua .4. Further Observations on the Effect of Mechanical Injury on Water-Uptake and Germination in Different Pure Lines." Botanical Gazette 149(4): 419-426.://A1988U423900009

Raju, M. V. S., G. J. Jones, et al. (1985). "Floret Anthesis and Pollination in Wild Oats (Avena-Fatua)." Canadian Journal of Botany- Revue Canadienne De Botanique 63(12): 2187-2195.://A1985A088400018

Raju, M. V. S. and S. N. Ramaswamy (1983). "Studies on the Inflorescence of Wild Oats (Avena-Fatua)." Canadian Journal of Botany-Revue Canadienne De Botanique 61(1): 74-78.://A1983QE09400006

Raju, M. V. S. and T. A. Steeves (1998). "Growth, anatomy and morphology of the mesocotyl and the growth of appendages of the wild oat (Avena fatua L.) seedling." Journal of Plant Research 111(1101): 73-85.://000074284100009 Morphological and histological studies were made on the mesocotyl and the emergence of seedlings of a nondormant strain (CS40) of wild oats (Avena fatua L.). The elongation of the mesocotyl was primarily responsible for the emergence of seedlings from deeper levels of soil. The mesocotyl of the seedling is here interpreted as the hypocotyl. The functionally suctorial scutellum together with coleoptile constitutes the first cotyledon and the first true-leaf is regarded as the second cotyledon. The development of tillers from scutellar and first-leaf buds depends on the depth at which level the seeds (caryopses) germinated and the seedlings emerged above the soil surface. The first-leaf axillary buds, regardless of depths, develop into dominant tillers. The scutellar buds, especially at greater depths, remain inhibited. At shallower levels, however, they develop into tillers. The scutellar buds, at deeper levels, behave as reserve ramets which feature adds to the success of the species as a weed in the agricultural prairies.

Raju, M. V. S. and A. Walther (1988). "Heterogeneity and Behavior of Aleurone Cells in the Caryopsis of Wild Oats (Avena-Fatua)." Flora 180(5-6): 417-427.://A1988P393500006

Raju, M. V. S., A. Walther, et al. (1988). "Growth and Development of Embryo Parts During the Germination of Caryopses of the Wild Oat (Avena-Fatua L)." Botanical Magazine-Tokyo 101(1061): 9-23.://A1988N032900002

Ramsdale, B. K. and C. G. Messersmith (2002). "Low-rate split-applied herbicide treatments for wild oat (Avena fatua) control in wheat (Triticum aestivum)." Weed Technology 16(1): 149-155.://000176901100023 Field experiments were conducted in 1999 and 2000 to evaluate reduced-rate split-applied treatments of imazamethabenz and ICIA 0604 for wild oat control in hard red spring wheat. Single herbicide treatments were applied at standard timing to wild oat at the two- to four-leaf stage. The first split-applied treatments were made when most of the wild oat had emerged but prior to the two-leaf stage. The second split-applied treatments were made after the surviving wild oat had visibly recovered and prior to the two-leaf stage of late-emerging wild oat. Split applications of imazamethabenz or ICIA 0604 allowed the herbicide rate to be reduced by half while maintaining or improving wild oat control. Both imazamethabenz and ICIA 0604 were most effective with methylated vegetable oil adjuvants. Wheat yield and net returns were greatest for half- rate split applications of imazamethabenz and ICIA 0604 in both years, likely because of less early-season wild oat competition and reduced wheat injury caused by ICIA 0604.

Ramsey, R. J. L., G. R. Stephenson, et al. (2002). "Effect of relative humidity on the uptake, translocation, and efficacy of glufosinate ammonium in wild oat (Avena fatua)." Pesticide Biochemistry and Physiology 73(1): 1-8.://000178019900001 Relative humidity (RH) can greatly affect the uptake and efficacy of glufosinate ammonium (D,L-homoalanine-4yl-(methyl)- phosphinate) in wild oat (Avena fatua L.). Exposure to high (>95%) RH as opposed to low (40%) RH increased glufosinate 53

ammonium efficacy and it was high RH within 12 h of spraying that was most crucial in affecting the efficacy. Subsequent dose-response experiments at 40% RH indicated that exposure of wild oat to high RH for as little as 30 min before and after spraying significantly increased the efficacy when compared to plants grown continuously at 40% RH. Furthermore, when [C-14]glufosinate ammonium was applied as a spray to wild oat plants grown at 40% RH, there was a significant increase in uptake into wild oat plants exposed to high RH for 30 min before and after treatment compared to those left continuously at low RH. Conversely, when the same experiment was conducted with [C-14]glufosinate ammonium applied as ten 1-mul droplets, exposure to high RH for 30 min before and after treatment did not increase the herbicide uptake. Based on these results, we hypothesize that droplet size and hence drying time is the major factor in effecting changes in glufosinate ammonium uptake in response to exposure to high versus low RH. Applying [C-14]glufosinate ammonium as large droplets resulted in an overestimation of uptake and masked the inhibitory effect of low RH on glufosinate ammonium as determined by dose-response experiments. Conversely, spray application of [C-14]glufosinate ammonium produced data that correlated with those from dose-response experiments. (C) 2002 Elsevier Science (USA). All rights reserved.

Ramsey, R. J. L., G. R. Stephenson, et al. (2006). "Effect of humectants on the uptake and efficacy glufosinate in wild oat (Avena fatua) plants and cuticles under dry conditions." Weed Science 54(2): 205-211.://000236728600003 A series of dose-response experiments were performed at low humidity to determine if glufosinate efficacy Could be increased by lengthening the drying time through the addition of humectants. Of several humectants evaluated, only 5% glycerol or 5% triethylene glycol when applied with glufosinate produced dry weight reductions and mortality similar to exposure to high humidity. C-14-glufosinate movement through isolated wild oat Cuticles was greater at high humidity, poorest at low humidity, but intermediate at low humidity in the presence of 5% glycerol in the spray solution. I-lie increases in uptake observed at high humidity and with 5% glycerol at low humidity were characterized by greater initial uptake that continued much longer than that observed at low humidity without humectant.

Rao, D. V. and M. V. S. Raju (1985). "Radial Elongation of the Epidermal-Cells of Scutellum During Caryopsis Germination of Wild Oats (Avena-Fatua)." Canadian Journal of Botany-Revue Canadienne De Botanique 63(10): 1789-1793.://A1985ASX9100019

Rashid, A., C. I. Johnson, et al. (1997). "Effects of triallate and difenzoquat on fatty acid composition in young shoots of susceptible and resistant Avena fatua populations." Pesticide Biochemistry and Physiology 57(1): 79-85.://A1997XL63800008 The composition of fatty acid moieties was similar in young shoot tissues of selected wild oat (Avena fatua L.) populations which were characterized to be either susceptible or resistant to the herbicide triallate. Gas chromatographic analysis of fatty acid methyl eaters showed that these tissues contained fatty acids of chain length C14 to C24, but the major components were C18 (C18:2, C18:3 +C18:1) and C16 chain lengths. All the other fatty acid chains were present as very minor components. Application of triallate at a concentration equivalent to the recommended field rate in either a soil- or it filter paper-based system for seedling growth caused a significant (P < 0.05) reduction in the level of all major fatty acids in the susceptible populations, but did not affect their levels in the resistant populations. In the susceptible populations, triallate also caused a significant increase in the level of C15 fatty acid. Application of difenzoquat (10 ppm), a herbicide to which triallate-resistant populations have shown cross-resistance, to young seedlings did not cause any significant change in the fatty acid composition of either susceptible or resistant populations. This would suggest that in these wild oat seedlings, fatty acid biosynthesis is not sensitive to difenzoquat; therefore, it may not be involved in cross-resistance to this herbicide. (C) 1997 Academic Press.

Rashid, A., J. T. O'Donovan, et al. (1998). "A possible involvement of gibberellin in the mechanism of Avena fatua resistance to triallate and cross-resistance to difenzoquat." Weed Research 38(6): 461-466.://000078334300008 Avena fatua L. populations from numerous fields in Alberta have simultaneously developed resistance to recommended rates of two dissimilar herbicides, triallate and difenzoquat. We used exogenous applications of gibberellic acid (GA(3)) to investigate the possibility that endogenous gibberellins were involved in the A. fatua's resistance mechanism. For control plants, without applied GA(3), shoots of the most resistant (R) populations emerged more rapidly and elongated faster than shoots of the susceptible (S) populations. Increases in shoot elongation in response to exogenous GA(3) were significantly lower in R populations compared with S populations. This suggests that R populations may have elevated endogenous gibberellin levels, relative to S populations. Additionally, inhibition of S population shoot elongation and shoot anatomical abnormalities caused by relatively low concentrations of triallate and difenzoquat could be prevented by exogenous application of GA(3). These results suggest that there may be a phytohormonal involvement in the mechanism of triallate/difenzoquat resistance in A. fatua. That is, higher endogenous gibberellin levels in R populations may result in meristematic growth that is rapid enough to preclude phytotoxic levels of these herbicides from reaching the shoot meristem, which is the likely site of action.

Rashid, A., J. T. Odonovan, et al. (1997). "Response of triallate-resistant and -susceptible wild oat (Avena fatua) populations to difenzoquat and EPTC in a seedling bioassay." Weed Technology 11(3): 527-531.://A1997XY34100019 A seedling bioassay was used to determine the response of triallate-resistant (R) and -susceptible (S) wild oat populations to difenzoquat and EPTC. The bioassay, based on seedling shoot length at 10 d after treatment, provided a reliable and rapid means of determining if wild oat populations were resistant to difenzoquat. Using a bioassay concentration of 15 ppm difenzoquat, it was possible to identify populations that were resistant to the recommended foliar-applied rate (0.85 kg ai/ha). Expected herbicide dosages that reduced shoot length by 50% (ED50) derived from nonlinear regression analysis indicated three and two levels of response to difenzoquat among eight S and seven R populations, respectively, indicating within population variability in their response to difenzoquat. Of the populations tested, none was resistant to EPTC. On the contrary, some R populations had lower ED50 values than did S populations, suggesting an increased sensitivity to EPTC.

Rathmann, D. P. and S. D. Miller (1981). "Wild Oat (Avena-Fatua) Competition in Soybean (Glycine-Max)." Weed Science 29(4): 410-414.://A1981MB15500011

54

Regnier, E. E. and K. B. Bakelana (1995). "Crop Planting Pattern Effects on Early Growth and Canopy Shape of Cultivated and Wild Oats (Avena-Fatua)." Weed Science 43(1): 88-94.://A1995QK71000015 Field studies were conducted to determine the effects of cultivated oats planting pattern on early canopy shape and growth of cultivated oats and wild oats, in part to test the assumption of radial plant canopy expansion on which previous theoretical models of crop-weed interference models have been based. Cultivated oats density was kept constant as the pattern rectangularity was varied, and single wild oats plants were centered within each pattern. Individual plant canopies, photographed from above 31 days after emergence (DAE), were radial for wild oats in all crop planting patterns and for cultivated oats planted in triangular and square planting patterns. Canopy radius perpendicular to the crop row axis in rectangular patterns was similar to canopy radius along the same cardinal axis in equidistant patterns, but was reduced along the crop row axis, resulting in a rectangular canopy shape and decreased canopy area in a rectangular compared to equidistant patterns. Cultivated oats dry weight and leaf area at crop flowering (64 DAE) also decreased with increasing rectangularity of crop planting pattern. Reductions in cultivated oats growth in rectangular patterns were associated with earlier intraspecific interference and delayed crop canopy closure in rectangular compared to equidistant patterns. Wild oats leaf area and tiller number 64 DAE decreased with more equidistant crop planting patterns, consistent with reduced canopy area 31 DAE and earlier crop canopy closure in equidistant patterns. The data suggest that individual oats canopy expansion during early growth is essentially radial and also support previous theoretical predictions of crop planting pattern effects on weed suppression.

Reich, J. M. and M. A. Brinkman (1984). "Inheritance of Groat Protein Percentage in Avena-Sativa L X Avena-Fatua L Crosses." Euphytica 33(3): 907-913.://A1984AEX0700032

Renault, S., A. Shukla, et al. (1997). "Plasma membrane lipid composition and herbicide effects on lipoxygenase activity do not contribute to differential membrane responses in herbicide-resistant and -susceptible wild oat (Avena fatua L) biotypes." Journal of Agricultural and Food Chemistry 45(8): 3269-3275.://A1997XR85600082 Plasma membrane lipid composition of herbicide-resistant (R) and -susceptible (S) wild oat biotypes was analyzed to determine the basis for the differential effect of diclofop on the transmembrane electrogenic potential between the two biotypes and reduced herbicide uptake into protoplasts of the R biotype. In addition, Lipoxygenase (LOX) activity was examined in herbicide-treated and untreated R and S plants to determine its involvement in herbicide action and resistance. Overall, no significant differences in lipid composition were found between the two biotypes. Glycolipids represented 41 and 36%, phospholipids 29 and 37%, and free sterols 30 and 27% of the total plasma membrane lipid in the R and S biotypes, respectively No differences in LOX activity were observed between the herbicide-treated and untreated wild oat biotypes. It was concluded that differences in membrane transport of diclofop and its effect on plasma membrane potential in the R and S biotypes are not related to differences in membrane lipid composition or to differential effects of herbicides on LOX activity in the two biotypes.

Reynolds, G. J. and R. Hooley (1992). "Cdna Cloning of a Tetraubiquitin Gene, and Expression of Ubiquitin-Containing Transcripts, in Aleurone Layers of Avena-Fatua." Plant Molecular Biology 20(4): 753-758.://A1992KB31200022 A lambdagt11 cDNA library, constructed from poly(A)+ mRNA isolated from A vena fatua aleurone layers incubated with 1 muM gibberellin A1 (GA1) for 4 days, was screened with an anti-idiotypic antiserum raised against the GA-specific monoclonal antibody MAC 182. One positive clone was isolated, sequenced and shown to encode a tetraubiquitin based on the deduced amino acid sequence. This polyubiquitin cDNA exhibited a high degree of homology to a cloned wheat hexaubiquitin in its 3'-non-coding region. Analysis of total RNA isolated from A. fatua aleurone layers, treated without or with a range of concentrations of GA1 from 10(-11) to 10(-6) M, by northern blotting using the cDNA probe revealed 8 different ubiquitin-containing transcript classes all of which are constitutively expressed in aleurone and are regulated by GA1.

Rezai, A. and K. J. Frey (1988). "Variation in Relation to Geographical-Distribution of Wild Oats - Seed Traits." Euphytica 39(2): 113-118.://A1988T716700004

Rines, H. W. and R. P. Halstead (1988). "Agronomic Evaluation of Oat Cultivars with Substituted Avena-Fatua and Avena-Sterilis Cytoplasms." Crop Science 28(5): 805-809.://A1988Q150800017

Rines, H. W., D. D. Stuthman, et al. (1980). "Collection and Evaluation of Avena-Fatua for Use in Oat Improvement." Crop Science 20(1): 63-68.://A1980JK33900015

Rolston, M. P. (1981). "Wild Oats in New-Zealand - a Review." New Zealand Journal of Experimental Agriculture 9(1): 115-121.://A1981LV11400021

Rooney, J. M. (1991). "Influence of Growth Form of Avena-Fatua L on the Growth and Yield of Triticum-Aestivum L." Annals of Applied Biology 118(2): 411-416.://A1991FQ43100016 In a field experiment, all components of winter wheat grain yield, except grain number per ear, were reduced by wild-oat competition, and more so by a prostrate form of wild-oat than by an upright form. The prostrate form, 'Churston', shaded the crop at anthesis. Although there were no significant differences in photosynthetic activity in the flag leaves of wheat at saturating light intensities during the later stages of grain filling, it is likely that differences in growth form, height, weight and area of these two forms of wild-oat, competing principally for light, accounted for their different effects on wheat grain yield.

Rooney, J. M., P. Brain, et al. (1989). "The Influence of Temperature on Leaf Production and Vegetative Growth of Avena-Fatua." Annals of Botany 64(4): 469-479.://A1989AX82400011

Ross, D. M. and R. C. Van Acker (2005). "Effect of nitrogen fertilizer and landscape position on wild oat (Avena fatua) interference in spring wheat." Weed Science 53(6): 869-876.://000233835300016 55

Site-specific fertilizer application is of primary interest among those investigating site-specific crop management. Site- specific soil moisture levels may be associated with the relative success of certain weed species, and weed competition can be affected by fertilizer application. A field study was conducted to investigate how landscape position-based site- specific nitrogen fertilizer application would influence wild oat interference in spring wheat. Wild oat was allowed to grow in spring wheat at foot and knoll landscape positions in the presence or absence of 80 to 90 kg ha(-1) of spring preseed broadcast nitrogen fertilizer. In all three site-years of the study, landscape position did not affect wild oat competitiveness in wheat or relative wild oat biomass, and there were no significant landscape position by nitrogen fertilizer interactions for these variables. The lack of landscape position effect may be attributed to the lack of substantive differences in soil characteristics between landscape positions at the sites used in this study and to normal to above-average seasonal precipitation levels in all three site-years. Wild oat competitiveness in wheat was significantly greater in the presence of nitrogen fertilizer for all three site-years. The results of this study suggest that, when there are not great differences in typical soil characteristics between landscape positions and precipitation levels are normal or above normal, landscape- based site-specific nitrogen fertilizer application does not affect wild oat competitiveness in spring wheat, but preseed spring broadcast nitrogen fertilizer does make wild oat more competitive in spring wheat.

Rubins, J. (1979). "Wild Oats - Epstein,J." Saturday Review 6(13): 43-44.://A1979GX65000020

Ruiz, D., C. Escribano, et al. (2006). "Assessing the opportunity for site-specific management of Avena sterilis in winter barley fields in Spain." Weed Research 46(5): 379-387.://000240404700004 The abundance of Avena sterilis in dryland barley fields was studied in four Spanish provinces. During two growing seasons, differential geopositioning system (DGPS)-geo-referenced A. sterilis infestations were obtained in 31 fields. The majority of the infestations were concentrated in few large but irregularly shaped patches, with a higher number of smaller and more regular patches accounting for a small proportion of the infestation. A multitude of very small and irregular patches completed the inventory. The implications of this spatial structure were studied. Site-specific adjusted-dose herbicide application offered 61-74% potential herbicide savings. However, given the low levels of infestation and the low economic returns obtained in most of the provinces, the most profitable strategy was generally no herbicide application. Site-specific treatments were advantageous only in high-returns systems. Because few large patches provided the majority of the infestation, zone-specific treatments would be advisable, until such time that weed detection and site-specific application technologies become more efficient.

Ruiz, D., C. Escribano, et al. (2006). "Identifying associations among sterile oat (Avena sterilis) infestation level, landscape characteristics, and crop yields." Weed Science 54(6): 1113-1121.://000242599800021 Information on weed spatial distribution could improve weed management decisions. Herbicide use could be reduced if applied only to field zones with an infestation level higher than a specific economic threshold. In this study, we surveyed 31 winter barley fields in different regions of Spain to describe the spatial distribution of sterile oat and to analyze the relationship between sterile oat infestation level and landscape and crop yield attributes. Elevation and slope angle, crop yield and slope aspect were the main factors in order of importance in explaining the distribution of sterile oat. In general, greater infestation levels were observed in flat lowland and concave landscapes, with a low crop yield, and on northern exposures (when slope = 2%). We could define field zones with a higher risk (> 6 times probability) of having sterile oat problems. High-probability zones, defined by topographic attributes alone, occupied 24% of the total area and contained 46% of the high infestation levels, whereas zones defined by topographic and crop yield attributes constituted 14% of the total area and contained 31% of the infestation.

Rummens, F. H. A., D. C. Rummensditters, et al. (1975). "Effects of Diallate and Its Isomers on Growth of Wild Oats." Weed Science 23(1): 11-14.://A1975V459600004

Rushe, D. (1978). "'Wild Oats' + Play by Okeefe,John." Eire-Ireland 13(1): 124-127.://A1978EZ19000010

Sabri, N. and D. D. Clarke (1996). "The relative tolerances of wild and cultivated oats to infection by Erysiphe graminis f sp avenae .1. The effects of infection on vegetative growth and yield." Physiological and Molecular Plant Pathology 49(6): 405-421.://A1996WU95500005 The effects of infection by the powdery mildew fungus Erysiphe graminis f.sp avenae were studied in one line of wild oat (Avena fatua), and two cultivars, Lustre and Peniarth, of cultivated oat (A. sativa) to determine if the wild oat was more tolerant of infection than the two cultivated oats. Seven weeks after inoculation, when the plants were 10-weeks-old with fully expanded flag leaves, the fungus had colonized approx. 40% of the leaf surfaces of wild oat and cv. Lustre but only about 30% of the leaf surfaces of cv. Peniarth. The lower leaves of cv. Peniarth were at least as susceptible, if not more so, than those of the other two lines but the upper leaves, including the flag leaf, were much more resistant. Although cv. Peniarth supported the production of about half the number of mildew conidia as the wild oat and cv. Lustre its total dry weight and grain yield were reduced to the greatest extent. The wild oat was clearly much more tolerant of mildew infection than cv. Peniarth and slightly more tolerant than cv. Lustre. The greater tolerances of the wild oat and cv. Lustre compared to cv. Peniarth appeared to be due to the lower sensitivities of their metabolism to the activities of the mildew fungus. (C) 1996 Academic Press Limited.

Sabri, N., P. J. Dominy, et al. (1997). "The relative tolerances of wild and cultivated oats to infection by Erysiphe graminis f.sp. avenae .2. The effects of infection on photosynthesis and respiration." Physiological and Molecular Plant Pathology 50(5): 321- 335.://A1997XY65800003 The effects of infection by the powdery mildew fungus Erysiphe graminis f.sp. avenae on photosynthesis and respiration in the third leaf of one Line of wild oat (Avena fatua) and two cultivars of cultivated oat (Avena sativa), Lustre and Peniarth, were studied. Gross and net photosynthesis and chlorophyll levels were reduced by infection in all three lines, but to the greatest extent in cv. Peniarth, and to the least extent in the wild oat. Infection increased respiration in the two cultivated oats, particularly in cv. Peniarth, but it had no significant effect on respiration in the wild oat. However, the highest level of 56

mildew biomass developed on the wild oat, and so the study clearly demonstrated that the wild oat possessed more tolerance of infection than either of the two cultivated oats, particularly cv. Peniarth. (C) 1997 Academic Press Limited.

Saini, H. S., P. K. Bassi, et al. (1986). "Interactions among Ethephon, Nitrate, and after-Ripening in the Release of Dormancy of Wild Oat (Avena-Fatua) Seed." Weed Science 34(1): 43-47.://A1986AYK3500010

Satorre, E. H. and R. W. Snaydon (1992). "A Comparison of Root and Shoot Competition between Spring Cereals and Avena- Fatua L." Weed Research 32(1): 45-55.://A1992HC09600006 Two cultivars of each of three spring cereals (wheat, barley and oats) were grown with Avena fatua in a box experiment, where the effects of root and shoot competition were separated using soil and aerial partitions. Measures of resource complementarity (RYT) and of the relative severity of competition indicated that competition for soil resources, particularly nitrogen, was more severe than competition for aerial resources, i.e. light. When both root and shoot competition occurred, oats and barley were generally more competitive than wheat, but there were significant differences in competitive ability between cultivars of all three cereals. Although competition between cereals and A. fatua occurred predominantly below- ground, the various cultivars only differed slightly in root competitive ability against weeds. Nitrogen fertilizer did not significantly change the ranking of competitive abilities of the cultivars. Although shoot competition had less effect than root competition, the various cereals differed markedly in shoot competitive ability. The two oat cultivars, and the barley cultivar Egmont, had higher shoot competitive abilities than the two wheat cultivars or the barley cultivar Goldmarker. There was little evidence that differences in shoot competitive ability were due to differences in plant height.

Sawhney, R. (1989). "Temperature Control of Dormancy and Germination in Embryos Isolated from Seeds of Dormant and Nondormant Lines of Wild Oats (Avena-Fatua)." Canadian Journal of Botany-Revue Canadienne De Botanique 67(1): 128-134.://A1989T509800019

Sawhney, R., A. I. Hsiao, et al. (1984). "Temperature Control of Germination and Its Possible Role in the Survival of a Non-Dormant Population of Avena-Fatua." Physiologia Plantarum 61(3): 331-336.://A1984TD13800005

Sawhney, R., A. I. Hsiao, et al. (1986). "The Influence of Diffused Light and Temperature on Seed-Germination of 3 Genetically Nondormant Lines of Wild Oats (Avena-Fatua) and Its Adaptive Significance." Canadian Journal of Botany-Revue Canadienne De Botanique 64(9): 1910-1915.://A1986E155300012

Sawhney, R. and J. M. Naylor (1979). "Dormancy Studies in Seed of Avena-Fatua .9. Demonstration of Genetic-Variability Affecting the Response to Temperature During Seed Development." Canadian Journal of Botany-Revue Canadienne De Botanique 57(1): 59- 63.://A1979GJ22500012

Sawhney, R. and J. M. Naylor (1980). "Dormancy Studies in Seed of Avena-Fatua .12. Influence of Temperature on Germination Behavior of Non-Dormant Families." Canadian Journal of Botany-Revue Canadienne De Botanique 58(5): 578-581.://A1980JM93500013

Sawhney, R. and J. M. Naylor (1982). "Dormancy Studies in Seed of Avena-Fatua .13. Influence of Drought Stress During Seed Development on Duration of Seed Dormancy." Canadian Journal of Botany-Revue Canadienne De Botanique 60(6): 1016-1020.://A1982NZ50500031

Sawhney, R., W. A. Quick, et al. (1985). "The Effect of Temperature During Parental Vegetative Growth on Seed-Germination of Wild Oats (Avena-Fatua L)." Annals of Botany 55(1): 25-28.://A1985AJY9000002

Schumacher, W. J., D. C. Thill, et al. (1983). "Allelopathic Potential of Wild Oat (Avena-Fatua) on Spring Wheat (Triticum- Aestivum) Growth." Journal of Chemical Ecology 9(8): 1235-1245.://A1983RN50400023

Scursoni, J. A. and R. B. Arnold (2002). "Effect of nitrogen fertilization timing on the demographic processes of wild oat (Avena fatua) in barley (Hordeum vulgare)." Weed Science 50(5): 616-621.://000178083600011 Wild oat is the most important weed in Argentine barley and one of the worst weeds in wheat and barley worldwide During 1997 and 1998, field experiments were performed to determine the response of wild oat to N fertilizer at planting and at early tillering at 55 kg ha(-1) with and without the use of diclofop-methyl Seedling emergence and survival rate, individual fecundity, and preharvest seed dispersal rate of wild oat were assessed The effect on the individual growth of wild oat was also studied N fertilizer increased both wild oat seedling survival rate and fecundity, especially when it was applied at early tillering However, the effect of N fertilization was negligible with diclofop-methyl treatment In addition, the application of N fertilizer reduced the seed dispersal rate immediately before crop harvest Quantification of the effects of each agronomic practice, such as N fertilization, is useful to develop simulation models that predict the population dynamics When N was applied at sowing, the annual growth rate of the wild oat population was almost 25% lower than when applied at early tillering.

Scursoni, J. A. and E. H. Satorre (2005). "Barley (Hordeum vulgare) and wild oat (Avena fatua) competition is affected by crop and weed density." Weed Technology 19(4): 790-795.://000233962200003 Wild oat is the most serious grass weed in Argentine barley crops and its control has concentrated on herbicide strategies. Increasing crop density could be an effective strategy to reduce the effect of wild oat on barley yield. However, limited research has been conducted to evaluate the effect of crop density on the competitive balance between barley and spontaneous populations of wild oat. A field experiment was conducted in 1992, 1993, and 1999, to study the effect of spontaneous populations of wild oat on barley sown at densities of 160, 220, and 280 plants/ml. Wild oat density averaged 84 plants/m(2). Wild oat biomass increased linearly with weed density in all treatments but was reduced by increasing barley seeding rates. Barley biomass and yield were not affected by wild oat at high crop sowing densities, but for the low and medium barley densities, yield loss was almost 25% when 70 wild oat plants/m(2) were established. Barley yield loss 57

was mostly related to competition from the early emerged wild oat. The relationship between yield losses and wild oat density was equally significant when the whole population or only early emerged individuals of the weed were considered. Nomenclature: Wild oat, Avena fatua L. #(3) AVEFA; barley, Hordeum vulgare L. Additional index words: Wild oat interference in barley, integrated wild oat management, seeding rates of barley.

Sebesta, J. and F. Kuhn (1990). "Avena-Fatua L Subsp Fatua V Glabrata Peterm Subv Pseudo-Basifixa Thell as a Source of Crown Rust Resistance Genes." Euphytica 50(1): 51-55.://A1990EE08300007

Seefeldt, S. S., E. P. Fuerst, et al. (1996). "Mechanisms of resistance to diclofop of two wild oat (Avena fatua) biotypes from the Willamette Valley of Oregon." Weed Science 44(4): 776-781.://A1996WA91300003 Laboratory experiments were conducted to determine the mechanism of resistance to diclofop in two wild oat biotypes (designated 'B' and 'C' biotypes) from the Willamette Valley of Oregon. Resistance could not be attributed to differential absorption, translocation, or metabolism of diclofop, Resistance was not correlated with membrane plasmalemma repolarization following diclofop acid treatment. Compared to a susceptible ('S') wild oat biotype, acetyl CoA carboxylase from the B and C biotypes showed a 10.3 and 4.5 fold increase in the level of resistance, respectively, to diclofop acid. Cross-resistance to fenoxaprop acid was 5.5 and 7.3 times higher in the B and C biotypes, respectively than the S biotype. Correlation between resistance at the whole plant level and at the ACCase level was good for diclofop and fenoxaprop in the B biotype. For the C biotype, this correlation was not as good. Possible reasons for the discrepancy are given.

Seefeldt, S. S., D. R. Gealy, et al. (1994). "Cross-Resistance of Several Diclofop-Resistant Wild Oat (Avena-Fatua) Biotypes from the Willamette Valley of Oregon." Weed Science 42(3): 430-437.://A1994PD56000017 The first occurrences of wild oat resistance to diclofop in the Willamette Valley of Oregon were reported in 1990. Among eight resistant biotypes, GR50 values for diclofop were 3 to 64 times greater than the GR50 for a susceptible wild oat biotype. GR50 Values for other aryloxyphenoxypropionate herbicides varied from 1 to over 100 times greater than a susceptible biotype. Only one resistant biotype was resistant to cyclohexanedione herbicides, and this was only a three- fold increase in GR50. Except for one biotype that had a low level of resistance to pronamide, none of the wild oat biotypes were cross-resistant to any other commonly used wild oat herbicide. Levels of resistance and cross-resistance did not follow a consistent pattern among biotypes in this study, suggesting more than one resistance trait. There were significant differences in the light use efficiency, height, dry weight, leaf area, and extent and timing of tillering and flowering of four wild oat biotypes studied. These physiological and morphological differences suggest that these resistant biotypes were selected independently. The diversity of resistance patterns and the coevolution of resistance at several locations will add to the difficulty of resistance management.

Seefeldt, S. S., D. L. Hoffman, et al. (1998). "Inheritance of diclofop resistance in wild oat (Avena fatua L.) biotypes from the Willamette Valley of Oregon." Weed Science 46(2): 170-175.://000073371800005 Inheritance of resistance to diclofop was studied in three wild oat biotypes (designated B, C, and H) from the Willamette Valley of Oregon. Cultivated oat (cultivar 'Monida') was crossed, including reciprocals, to three wild oat biotypes. Leaves of each F-1 plant were spotted with diclofop as a nondestructive test for resistance or susceptibility. All F-1 hybrids were resistant, indicating that resistance is dominant and is under nuclear control. The F-2 plants where Monida was the maternal parent were screened with diclofop, and F-2 plants of the Monida/C cross were screened with fenoxaprop because the parent C biotype was resistant to fenoxaprop. At lower doses, a 3:1 (R:S) segregation ratio in F-2 was observed and at higher doses a 1:3 (R:S) segregation ratio was often observed. The F-2:3 families segregated in a 1:2:1 (all resistant: segregating resistant and susceptible all susceptible) ratio when treated with a 1.1-kg ae ha(-1) dose of diclofop. This confirms that resistance to diclofop in the B, C, and H biotypes is primarily under monogenic control, with resistance being dominant to susceptibility at lower herbicide doses. At increased doses, susceptibility becomes dominant. Knowledge of the inheritance of resistance may help in the development of containment measures to prevent the spread of herbicide-resistance genes.

Sereno, M., M. H. Bodanese-Zanettini, et al. (1998). "Genetic variability in Avena sativa, Avena sterilis and in their hybrids through the agronomic evaluation." Pesquisa Agropecuaria Brasileira 33(10): 1601-1607.://000079527000007 This study aimed to compare the genetic variability of the cultivated oat group and of the wild-type of this cereal. Several agronomic traits like plant stature, number of primary and secondary anthoecium on the main panicle, vegetative cycle (head date), number of kernel only on the main panicle, 1000 kernel weight and tiller number were studied in the field in different groups of cultivated oat genotypes (UFRGS 7, UFRGS 8 and UPF 7) and wild-types (I-303, I-320, I-325, I-377, I- 378, I-428 and I-ARG). Statistical differences were detected on all of the traits tested, except for the tillers number. Higher interspecific than intraspecific genetic variability was also found. The action of natural selection on all of the Avena sterilis L. group and the selection pressure applied by man (artificial selection) on the Avena sativa L. group could be the main point of difference between the two species. The I-325, I-377 and I-378 introductions showed similar vegetative cycle to the cultivated group, and increased the possibility of interspecific hybridization. The I-ARG wild type showed to have intermediate agronomic traits in relation to cultivated and wild oat types.

Serenotavares, M., M. H. Bodanesezanettini, et al. (1995). "Hybridization among Brazilian Oat (Avena-Sativa L) Cultivars and the Wild Oat Avena-Sterilis L." Pesquisa Agropecuaria Brasileira 30(3): 383-388.://A1995RM73300012 Hybrids among Avena sativa (L.) Brazilian cultivars and A. sterilis L. accessions were obtained without in vitro embryo culture. Parental samples and their hybrids were investigated by morphological characters of spikelets, meiotic chromosome behaviour and esterase isoenzymes analysis. The F-1 hybrids were intermediate to the parental lines for the morphological spikelet characters. The high esterase zymotype affinity made difficult the diagnostic of the hybrid condition. The hybrids presented more meiotic irregularities than the parental lines. However, predominant bivalent formation shows that the corresponding chromosomes in A, sativa and A. sterilis are able to pair effectively and should result in the recombination of characters of the two species.

Sexsmith, J. J. (1967). "Varietal Differences in Seed Dormancy of Wild Oats." Weeds 15(3): 252-&.://A19679983900018 58

Sharma, M. P., D. K. McBeath, et al. (1976). "Studies on Biology of Wild Oats .1. Dormancy, Germination and Emergence." Canadian Journal of Plant Science 56(3): 611-618.://A1976CE92500025

Sharma, M. P., D. K. McBeath, et al. (1977). "Studies on Biology of Wild Oats .2. Growth." Canadian Journal of Plant Science 57(3): 811-817.://A1977DR41000028

Sharma, M. P., F. A. Qureshi, et al. (1982). "The Basis for Synergism between Barban and Flamprop on Wild Oat (Avena-Fatua)." Weed Science 30(2): 147-152.://A1982NH00700006

Sharma, M. P. and W. H. Vandenborn (1978). "Biology of Canadian Weeds .27. Avena-Fatua L." Canadian Journal of Plant Science 58(1): 141-157.://A1978EL19600020

Sharma, M. P., W. H. Vandenborn, et al. (1977). "Effects of Post-Emergence Wild Oat Herbicides on Transpiration of Wild Oats." Canadian Journal of Plant Science 57(1): 127-132.://A1977CZ47000019

Sharma, M. P., W. H. Vandenborn, et al. (1978). "Spray Retention, Foliar Penetration, Translocation and Selectivity of Asulam in Wild Oats and Flax." Weed Research 18(3): 169-&.://A1978FK07700007

Shepeleva, E. M. (1939). "Karyosystematic study on cultivated and wild oats." Comptes Rendus De L Academie Des Sciences De L Urss 25: 228-231.://000201892400064

Shirtliffe, S. J. and M. H. Entz (2005). "Chaff collection reduces seed dispersal of wild oat (Avena fatua) by a combine harvester." Weed Science 53(4): 465-470.://000230999200007 Combine harvesters have the potential to disperse weed seeds great distances. Reducing this dispersal may be important in an integrated weed management system. The objectives of this study were to determine the distance that wild oat seeds are dispersed by a combine harvester and the effect of chaff collection on combine harvester seed dispersal. This was measured by sampling wild oat seeds at varying distances behind a combine equipped with a removable chaff collection system after it passed through a wild oat patch. Chaff collection consistently reduced the amount and distance that wild oat seeds were dispersed. This occurred because more than 74% of the total wild oat seed that were ejected from the combine were in the chaff. Because most of the chaff falls in a row directly behind the combine, chaff collection only affected dispersal in this area. In 1996, chaff collection reduced wild oat seed dispersal past the wild oat patch to less than 10 seeds m(-2) at 45 m, whereas without chaff collection, there was greater than 10 seeds m(-2) up to 145 m. At distances beyond 145 m, chaff collection had no significant effect on seed dispersal. Chaff collection may be an important tool in an integrated weed management program because it may slow weed invasions and reduce the expansion of weed patches.

Shirtliffe, S. J., M. H. Entz, et al. (2000). "Avena fatua development and seed shatter as related to thermal time." Weed Science 48(5): 555-560.://000089783100005 Avena fatua seeds remaining on the plant at harvest and taken into the combine harvester may be dispersed over large areas. The objective of this study was to characterize the development of A. fatua in comparison to spring Triticum aestivum. As part of this objective, the rate of seed shed in A. fatua relative to development of T. aestivum was determined. Avena fatua and T. aestivum had similar phyllochron intervals within locations but differed between locations. Plant development as measured by the Zadoks plant development scale was consistent within plant species between locations. Seed shed in A. fatua was also consistent between locations. Most of the seed shed occurred within 2 wk, and the cumulative seed shed followed a sigmoidal pattern. The seed shed occurred as T. aestivum was ripening, and the percentage of seed shed appears to be related to the water content of the T. aestivum spike. Because of this relationship, the proportion of seed remaining on A. fatua at harvest could be managed by changing the timing of crop harvest.

Shorter, R., P. Gibson, et al. (1978). "Outcrossing Rates in Oat Species Crosses (Avena-Sativa L X Avena-Sterilis L)." Crop Science 18(5): 877-878.://A1978FW50200051

Shu, H. L., A. I. Hsiao, et al. (1992). "Effects of Sodium Bisulfate, Acidic Buffers and Ammonium-Sulfate on Imazamethabenz Phytotoxicity to Wild Oats." Crop Protection 11(4): 335-340.://A1992JD86600006 Greenhouse studies determined the effects of sodium bisulphate (NaB), acidic buffers and ammonium sulphate on the phytotoxicity of imazamethabenz, (+/-)-2-[4.5-dihydro-4-methyl-4-(1-methylethyl)-5-oxo-1H-imidazol-2-yl]- 4-(and 5)- methylbenzoic acid (3:2), applied to wild oat plants at the 2- to 3-leaf stage. At an equivalent dosage of imazamethabenz the liquid concentrate (LC) formulation had higher herbicidal activity and much lower pH than did the suspension concentrate (SC) formulation. NaB added to the SC formulation lowered solution pH by >3 units to approximately 2.2 and enhanced herbicidal activity at 200-300 g ha-1 active ingredient (a.i.), but not at 400 g a.i. ha-1. NaB did not improve the activity of the LC formulation. Increasing the spray solution pH from 5.6 to approximately 7.3 with 1 M KOH antagonized phytotoxicity of both LC and SC formulations of imazamethabenz alone, and imazamethabenz-SC plus NaB. Citric acid phosphate buffers at pH 3 6, or HCl-KCl buffers at pH 1 and 2, either had no effect or antagonized the activity of the SC formulation of imazamethabenz. However. the same HCl-KCl buffers plus 1% (v/v) ammonium sulphate enhanced herbicidal activity at 250 g a.i. ha-1, while 1% ammonium sulphate, alone, enhanced herbicidal activity at both 200 and 250 g ha-1. NaB provided both bisulphate ion and low pH concurrently. Acidic buffers did not improve the activity of the SC formulation except in the presence of bisulphate or sulphate ions. It is suggested that at low pH the herbicide becomes cationic, forms the bisulphate or sulphate salt of the methyl ester of imazamethabenz, and enters true solution. In this form it readily penetrates the cuticle. High solution pH acts against this process.

Shuma, J. M., W. A. Quick, et al. (1995). "Germination of Seeds from Plants of Avena-Fatua L Treated with Glyphosate." Weed Research 35(4): 249-255.://A1995RW46600005 59

Glyphosate was applied at four rates under greenhouse conditions to Avena fatua L. plants at four stages of seed development. Application at anthesis completely prevented the formation of viable seeds. Application five days after anthesis (DAA) of the terminal floret of the panicle significantly reduced seed production at ail herbi-cide rates used, and at 1.76 kg a.i, ha(-1) no viable seeds were produced. When applied 10 DAA, only the highest rate of glyphosate resulted in substantial reduction in number of primary seeds, but seed viability suffered at all herbicide levels. Glyphosate applied 15 DAA still produced a significant decrease in primary and secondary seed production and biomass. Both the viability and the germination rate of seeds from treated plants were significantly affected. When the herbicide was applied to plants 5 DAA, no viable seeds were produced by plants surviving the highest rate, and all rates significantly reduced germination. Glyphosate applied 10 DAA significantly suppressed germination, with 1.76 kg a.i. ha(-1) being the most effective rate. When applied to plants 15 DAA, only the highest rate of glyphosate significantly affected the overall germination of both primary and secondary seeds, but the normal imposition of dormancy was partially blocked in seeds from plants treated with 0.44 and 0.88 kg a.i. ha(-1). These findings are relevant to chemical summerfallow and crop desiccation practices.

Shuma, J. M. and M. V. S. Raju (1993). "A Histological Study of the Effect of Glyphosate on Seed Development in the Wild Oat (Avena-Fatua L)." Weed Research 33(1): 43-51.://A1993KK39200006 Glyphosate was applied to panicles of wild oats (Avena fatua L.) before, at and after anthesis. The florets in spikelets exposed to the herbicide before, at and immediately after anthesis produced seeds with abnormal embryos and endosperm. Histological and embryological studies showed that in all cases the chemical did not stop the normal process of double fertilization in the florets. The zygote divided, but did not proceed beyond the proembryo stage. The development of both the proembryo and the endosperm was abnormal. The proembryos and the early stages of embryos showed no normal provascular or vascular tissue differentiation. Externally also there was no proper development of embryo parts, such as the scutellum, epiblast, shoot and root apices. The embryonic tissues showed considerable obliteration and shrinkage and the nuclei in cells appeared small and shrunken. The cells, especially in the meristem regions, appeared lacunose. The embryo and the endosperm in the caryopses treated with glyphosate 9 days after anthesis or later, were morphologically and structurally very similar to those of the controls of comparable age. It was concluded that the greater the delay of glyphosate application after anthesis, the less was the effect on seed development.

Shumaker, K. M., R. W. Allard, et al. (1982). "Cryptic Variability at Enzyme Loci in 3 Plant-Species, Avena-Barbata, Hordeum- Vulgare, and Zea-Mays." Journal of Heredity 73(2): 86-90.://A1982NK45900001

Simmonds, J. A. (1971). "Increased Participation of Pentose Phosphate Pathway in Response to Afterripening and Gibberellic Acid Treatment in Caryopses of Avena-Fatua." Canadian Journal of Botany 49(10): 1833-&.://A1971K653000013

Simmonds, J. A. and G. M. Simpson (1972). "Regulation of Krebs Cycle and Pentose Phosphate Pathway Activities in Control of Dormancy of Avena-Fatua." Canadian Journal of Botany 50(5): 1041-&.://A1972M654200016

Simons, M. D. (1965). "Seedling Resistance to Puccinia Coronata Avenae Race 264 Found in Avena Sterilis." Phytopathology 55(6): 700-&.://A19656545500024

Simons, M. D. (1971). "Crown Rust Tolerance of Avena-Sativa-Type Oats Derived from Wild Avena-Sterilis." Phytopathology 61(8): 911-&.://A1971K035400180

Simons, M. D. (1972). "Crown Rust Tolerance of Avena-Sativa-Type Oats Derived from Wild Avena-Sterilis." Phytopathology 62(12): 1444-1446.://A1972O554200018

Simons, M. D. (1979). "Influence of Genes for Resistance to Puccinia-Coronata from Avena-Sterilis on Yield and Rust Reaction of Cultivated Oats." Phytopathology 69(5): 450-452.://A1979HM95300006

Simons, M. D. (1981). "Transfer of Crown Rust Field-Resistance from Avena-Sterilis to Cultivated Oats by Backcrossing." Phytopathology 71(8): 904-904.://A1981MD74100476

Simons, M. D. (1985). "Transfer of Field-Resistance to Puccinia-Coronata from Avena-Sterilis to Cultivated Oats by Backcrossing." Phytopathology 75(3): 314-317.://A1985AEK2400013

Simons, M. D. and L. W. Briggle (1984). "Screening for Tolerance to Puccinia-Coronata in Progenies of Visually Susceptible Strains of Avena-Fatua." Phytopathology 74(10): 1271-1271.://A1984TN52600055

Simons, M. D., L. D. Robertson, et al. (1985). "Association of Host Cytoplasm with Reaction to Puccinia-Coronata in Progeny of Crosses between Wild and Cultivated Oats." Plant Disease 69(11): 969-971.://A1985ATV7700018

Simpson, G. M. (1965). "Dormancy Studies in Seed of Avena Fatua .4. Role of Gibberellin in Embryo Dormancy." Canadian Journal of Botany 43(7): 793-&.://A19656667400003

Simpson, G. M. (1966). "Suppression by (2-Chloroethyl)Trimethylammoniumchloride of Synthesis of a Gibberellin-Like Substance by Embryos of Avena Fatua." Canadian Journal of Botany 44(1): 115-&.://A19667243200018

Smith, A. M. and P. N. P. Chow (1990). "The Influence of Agral-90 Surfactant on the Activity of Imazamethabenz in Wild Oats (Avena-Fatua L)." Weed Research 30(5): 355-362.://A1990ED13600005

Smith, S. J. and R. Hooley (2002). "An increase in the speed of response, and sensitivity, of Avena fatua aleurone layers and protoplasts to gibberellin." Journal of Plant Physiology 159(4): 355-360.://000175707000004 60

In wild oat (Avena fatua) aleurone protoplasts and layers, a-amylase induction by gibberellin is very slow compared with other members of the Gramineae. Wild oat aleurone protoplasts for example do not secrete a-amylase until approximately 72 hours after treatment with gibberellin while in aleurone layers or protoplasts from wheat or barley this response takes between 8 and 24 hours. In this study, we demonstrate that the duration of the lag phase in A. fatua aleurone protoplasts can be reduced substantially by incubating the protoplasts in the absence of hormone for up to 4 days before treating with gibberellin. RNA gel blot analysis revealed substantial induction of alpha-amylass mRNA within 14 hours of gibberellin treatment in aged aleurone protoplasts. A reduction in the duration of the lag phase is also observed in intact wild oat aleurone layers that have been incubated in the absence of hormone for 4 days prior to gibberellin treatment. Interestingly, this effect is only observed in aleurones from which the endosperm has been removed. Dose response analysis revealed that the concentration of gibberellin required to induce a maximum response was approximately one order of magnitude lower in the 4 day-old tissue compared with freshly isolated material. These observations suggest that the long duration of the lag phase of freshly isolated aleurone layers and protoplasts of wild oat is due in part to gibberellin independent events.

Smith, S. J., R. P. Walker, et al. (1993). "Biological-Activity of Some Gibberellins and Gibberellin Derivatives in Aleurone Cells and Protoplasts of Avena-Fatua." Phytochemistry 33(1): 17-20.://A1993LB56100004 The induction of alpha-amylase in aleurone cells and protoplasts of Avena fatua by a number of gibberellins (GAs) and GA-derivatives has been determined. GA1, GA4 and GA3 are of similar high activity in aleurone protoplasts, as are the more lipophilic synthetic GA4 derivatives, 2,2-dimethylGA4 and 2,2-diethylGA4. 3-epiGA4 and GA4 methyl ester have moderate activity. 3-epiGA4 methyl ester has low activity. The 2beta-hydroxylated GA8 and GA51 have very low activity in both aleurone cells and protoplasts. GA51 does not inhibit GA4-induction of alpha-amylase in aleurone protoplasts.

Somersalo, S., P. Makela, et al. (1998). "Morpho-physiological traits characterizing environmental adaptation of Avena barbata." Euphytica 99(3): 213-220.://000072085100009 Seventeen morphological and physiological characteristics of three Avena barbata L. populations from Israel were measured in order to define possible combinations explaining adaptation of these populations to different precipitation, temperature and altitude regimes. Five genotypes from each A. barbata populations were collected from Ashqelon (31 degrees 63'N, low annual precipitation), En Hamifraz (32 degrees 46'N, high temperature), and Mount Carmel (32 degrees 73'N, high altitude), Israel. The behavior of the populations was followed by measuring the morphophysiological characteristics under well-watered and moderately drought stressed conditions. The experiment was conducted at the Department of Plant Production, University of Helsinki, Finland (60 degrees 13'N). The measured traits characterized macro-morphology, transpiration rate, photosynthesis and chloroplast features. The data were subjected to principal component and discriminant analyses and the characteristic combinations that most adequately accounted for the differences among A. barbata populations were established. Differences among the populations were related to adaptation to low water availability and high altitude characterized by special light conditions. The Mount Carmel population (high water availability, high light intensities and increased proportion of UV-light) was characterized by higher tillering, hairy leaf sheaths, high transpiration, high stomatal conductance, slow fluorescence quenching capacity, and less starch granules per chloroplast when compared with populations adapted to lower altitudes. The En Hamifraz population (high mean temperature) was characterized by a high CO2 exchange rate and both En Hamifraz and Ashqelon populations (both adapted to arid conditions) used water sparingly when moderately drought stressed.

Somody, C. N., P. K. Fay, et al. (1982). "The Wild Oats Pilot Project." North Dakota Farm Research 39(4): 25-30.://A1982ND21600005

Somody, C. N., J. D. Nalewaja, et al. (1984). "The Response of Wild Oat (Avena-Fatua) and Avena-Sterilis Accessions to Photoperiod and Temperature." Weed Science 32(2): 206-213.://A1984SH55800015

Somody, C. N., J. D. Nalewaja, et al. (1984). "Wild Oat (Avena-Fatua) and Avena-Sterilis Morphological-Characteristics and Response to Herbicides." Weed Science 32(3): 353-359.://A1984SR03300014

Somody, C. N., J. D. Nalewaja, et al. (1984). "Wild Oat (Avena-Fatua) Seed Environment and Germination." Weed Science 32(4): 502-507.://A1984TC58500016

Sosulski, F. (1983). "Wild Oats for Breakfast Tables." Milling 166(6): 48-48.://A1983QX51300010

Sosulski, F. W. (1983). "Potential of Wild Oats as a High Protein Food." Journal of the American Oil Chemists Society 60(4): 708- 708.://A1983QL66600104

Sosulski, F. W. and K. Elkowicz (1982). "Potential of Wild Oats as a High Protein Food." Canadian Institute of Food Science and Technology Journal-Journal De L Institut Canadien De Science Et Technologie Alimentaires 15(3): R21-R21.://A1982PE91000095

Spandl, E., B. R. Durgan, et al. (1997). "Wild oat (Avena fatua) control in spring wheat (Triticum aestivum) and barley (Hordeum vulgare) with reduced rates of postemergence herbicides." Weed Technology 11(3): 591-597.://A1997XY34100030 Rates and application timings of postemergence herbicides for wild oat control in spring wheat and barley were evaluated at Crookston, MN, from 1994 to 1996. Diclofop, imazamethabenz, and fenoxaprop plus MCPA plus thifensulfuron plus tribenuron were applied to one-to three-leaf wild oat; and difenzoquat, imazamethabenz, fenoxaprop plus MCPA plus thifensulfuron plus tribenuron, and fenoxaprop plus 2,4-D plus MCPA were applied to four-to five-leaf wild oat at 1/2 X, 3/4 X, and 1 X rates. Wild oat response to herbicide rate and timing was similar in wheat and barley. Wild oat control with 1/2 X rates generally was less than that with 3/4 X rates, which was lower than or similar to that with 1 X rates. Wild oat biomass was often reduced less with 1/2 X rates than 1 X rates. However, reducing herbicide rates generally did not 61

influence grain yields or net economic return. Grain yields and net economic return were generally greater in herbicide- treated plots than in the nontreated control.

Steidl, R. P., J. A. Webster, et al. (1979). "Cereal Leaf Beetle Coleoptera, Chrysomelidae Plant-Resistance - Antibiosis in an Avena-Sterilis Introduction." Environmental Entomology 8(3): 448-450.://A1979HA23800017

Stevens, J. B. and M. A. Brinkman (1986). "Performance of Avena-Sativa L Avena-Fatua L Backcross Lines." Euphytica 35(3): 785-792.://A1986F503600013

Stevenson, F. C., F. A. Holm, et al. (2000). "Optimizing wild oat (Avena fatua) control with ICIA 0604." Weed Technology 14(3): 608-616.://000166693600024 Wild oat (Avena fatua) control often is an integral management practice in cropping systems that include cereal crops. Experiments were conducted at two locations in Saskatchewan (Saskatoon and Scott), Canada, from 1994 to 1997 to determine the influence of ICIA 0604 rate (50, 100, 150, and 200 g ai/ha), water volume (30, 50, and 100 L/ha), spray mixture pH (unbuffered, close to pH 7.0; reduced, pH 4.0), late morning and evening application times, and sodium bicarbonate concentration of water source (Saskatoon water, negligible; Scott water, 695 mg/L) on wild oat fresh weight and wheat (Triticum aestivum) grain yield. Reducing ICIA 0604 rate below the recommended label rate (200 g/ha) increased wild oat fresh weight by 22% and decreased wheat grain yield by 7% when applied with 50 or 100 L/ha of water. Applications with 30 L/ha of water resulted in more wild oat growth (19%) and less wheat yield (6%), regardless of the ICIA 0604 rates. Spray mixture pH or time of application did not modify the effects of ICIA 0604 rate and water volume on wild oat fresh weight and wheat yield at Saskatoon. At Scott, the negative effects of ICIA 0604 rates lower than 200 g/ha applied with 50 or 100 L/ha of water were most apparent when applications were made in the morning, especially with an unbuffered spray mixture. ICIA 0604 applications made in the evening with 50 or 100 L/ha of water resulted in the lowest wild oat fresh weights and greatest wheat yields, regardless of the ICIA 0604 rate or spray mixture pH. Antagonism between sodium bicarbonate in the unbuffered water from Scott, as indicated by the spray mixture pH effect, and the time of application effect were important factors controlling treatment responses at Scott. Lower than recommended ICIA 0604 rates often maintained net returns, even though wheat yield responded negatively to reduced ICIA 0604 rates. Understanding the effects of water quality on wild oat control will allow producers to make prudent decisions regarding the optimal application parameters for ICIA 0604.

Stinson, R. H. and R. L. Peterson (1979). "Sowing Wild Oats." Canadian Journal of Botany-Revue Canadienne De Botanique 57(11): 1292-1295.://A1979HC48100015

Stoklosa, A., A. Janeczko, et al. (2006). "Isothermal calorimetry as a tool for estimating resistance of wild oat (Avena fatua L.) to aryloxyphenoxypropionate herbicides." Thermochimica Acta 441(2): 203-206.://000235417100017 The application of isothermal calorimetry for the early detection of the resistance of wild oat to fenoxaprop(1) and diclofop(2) was investigated. In the first test, three leaf tillers were sprayed with field doses of fenoxaprop or diclofop. For resistant biotypes, the rate of heat flow after 48 h was similar to that in control plants. In susceptible biotypes, fenoxaprop significantly reduced and diclofop significantly increased the rate of heat flow. In the second test, 3-day-old seedlings were put into calorimetric ampoules on filter paper moistured with herbicide solution (152% and 40% of the field dose for fenoxaprop and diclofop, respectively). Rate of heat flow was measured for 72 h, however, differences were already visible in the first hours of germination on each herbicide. Rate of heat flow for seedlings resistant to both herbicides was higher than for susceptible ones. The most evident differences between susceptible and resistant biotypes were noticed after 10- 20 h and 25-40 h (of the seedlings' growth) on fenoxaprop and diclofop, respectively, when a sharp increase of rate of heat flow was observed. In conclusion, calorimetry may be used as a rapid test for the detection of the resistance of wild oat biotypes to fenoxaprop and diclofop. (C) 2005 Elsevier B.V. All rights reserved.

Stougaard, R. N. (1997). "Adjuvant combinations with quizalofop for wild oat (Avena fatua) control in peppermint (Mentha piperita)." Weed Technology 11(1): 45-50.://A1997WX37900009 Field experiments were conducted at Kalispell, MT, to determine the optimum adjuvant combination for wild oat control in peppermint with quizalofop. Quizalofop was applied to four- and eight-leaf wild oat plants at 20 and 50 g ai/ha with either a nonionic surfactant (NIS) or methylated seed oil (MSG) alone or in combination with 28% urea ammonium nitrate (UAN) liquid fertilizer. Differences among adjuvants were most apparent when quizalofop was applied at the lowest rate. MSO was more effective than NIS for enhancing quizalofop activity. Quizalofop efficacy with both adjuvants increased when applied with UAN. Greater than 90% wild oat control was obtained with the lowest, rate when applied with MSO plus UAN to four-leaf wild oat plants. These results demonstrate the potential to improve the consistency of weed control as well as reduce postemergence herbicide rates when applied with the proper adjuvant combination.

Stougaard, R. N., B. D. Maxwell, et al. (1997). "Influence of application timing on the efficacy of reduced rate postemergence herbicides for wild oat (Avena fatua) control in spring barley (Hordeum vulgare)." Weed Technology 11(2): 283-289.://A1997XH42700012 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 X and 1/2 X 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 X and 1/2 X 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 X diclofop treatments providing the 62

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.

Stougaard, R. N. and Q. Xue (2005). "Quality versus quantity: spring wheat seed size and seeding rate effects on Avena fatua interference, economic returns and economic thresholds." Weed Research 45(5): 351-360.://000231707400004 A 3-year field experiment was conducted at Kalispell, Montana, USA, to investigate the effects of spring wheat seed size and seeding rate on wheat yield loss (YL), economic returns and economic thresholds (ETs), as a function of Avena fatua density. Crop competitive ability increased as wheat seeding rate and seed size increased, with the greatest differences among treatment factors being observed at low weed densities. Both treatment factors decreased spring wheat YL, increasing economic returns during all 3 years of the study despite the higher associated seed costs. Averaged over all other factors, adjusted gross returns (AGR) were 477 and 537$ ha(-1) for the low and high seeding rates, while values of 453, 521 and 547$ ha(-1) were obtained for the small, bulk and large seed size classes respectively. Weed-free yield potential varied yearly. As yield potential increased, A. fatua competitive effects were more evident and ETs decreased. Nonetheless, both treatment factors increased ETs in 2 of 3 years. These results demonstrate that the use of higher seeding rates and larger seed size classes both improve wheat competitive ability towards A. fatua while simultaneously increasing economic returns.

Stougaard, R. N. and Q. W. Xue (2004). "Spring wheat seed size and seeding rate effects on yield loss due to wild oat (Avena fatua) interference." Weed Science 52(1): 133-141.://000188218200018 The development of competitive cropping systems could minimize the negative effects of wild oat competition on cereal grain yield, and in the process, help augment herbicide use. A 3-yr field experiment was conducted at Kalispell, MT, to investigate the effects of spring wheat seed size and seeding rate on wheat spike production, biomass, and grain yield under a range of wild oat densities. Wheat plant density, spikes, biomass, and yield all increased as seed size and seeding rates increased. Averaged across all other factors, the use of higher seeding rates and larger seed sizes improved yields by 12 and 18%, respectively. Accordingly, grain yield was more highly correlated with seed size than with seeding rate effects. However, the combined use of both tactics resulted in a more competitive cropping system, improving grain yields by 30%. Seeding rate effects were related to spike production, whereas seed size effects were related to biomass production. As such, plants derived from large seed appear to have greater vigor and are able to acquire a larger share of plant growth factors relative to plants derived from small seed.

Suneson, C. A. and H. G. Marshall (1967). "Cold Resistance in Wild Oats." Crop Science 7(6): 667-&.://A1967A406100033

Surface, F. M. (1916). "On the inheritance of certain glume characters in the cross Avena fatua X a Sativa var Kherson." Proceedings of the National Academy of Sciences of the United States of America 2: 478-484.://000201969200124

Surface, F. M. (1916). "Studies on oat breeding. III. On the inheritance of certain glume characters in the cross Avena fatua x A- sativa Var. Khersow." Genetics 1(3): 252-286.://000201582000002

Symons, S. J., J. M. Naylor, et al. (1986). "Secondary Dormancy in Avena-Fatua - Induction and Characteristics in Genetically Pure Dormant Lines." Physiologia Plantarum 68(1): 27-33.://A1986E206500004

Symons, S. J., G. M. Simpson, et al. (1987). "Secondary Dormancy in Avena-Fatua - Effect of Temperature and after-Ripening." Physiologia Plantarum 70(3): 419-426.://A1987J316800007

Szekeres, F. (1982). "Effect of Light, Range of Temperature and Photoperiodic Induction on Germinating-Seeds and Caryopses of Avena-Fatua L." Acta Phytopathologica Academiae Scientiarum Hungaricae 17(3-4): 285-290.://A1982RF05900011

Sztejnberg, A. and I. Wahl (1976). "Mechanisms and Stability of Slow Stem Rusting Resistance in Avena-Sterilis." Phytopathology 66(1): 74-80.://A1976BD88200015

Tadmor, N. H. and G. Orshan (1965). "Competition between Avena Sterilis L and Stipa Tortilis Desf under Conditions of Adequate Water Supply." Israel Journal of Botany 13(2-4): 234-&.://A19656808900017

Takeda, K., T. B. Bailey, et al. (1985). "Changes in Mean, Variance, and Covariation among Agronomic Traits in Successive Backcross Generations of Interspecific Matings (Avena-Sativa L X Avena-Sterilis L) of Oats." Canadian Journal of Genetics and Cytology 27(4): 426-432.://A1985APF1700009

Takeda, K. and K. J. Frey (1976). "Contributions of Vegetative Growth-Rate and Harvest Index to Grain-Yield of Progenies from Avena-Sativa by Avena-Sterilis Crosses." Crop Science 16(6): 817-821.://A1976CT06500019

Takeda, K. and K. J. Frey (1977). "Growth-Rate Inheritance and Associations with Other Traits in Backcross Populations of Avena- Sativaxa-Sterilis." Euphytica 26(2): 309-317.://A1977DP29400009

Takeda, K. and K. J. Frey (1979). "Protein Yield and Its Relationship to Other Traits in Backcross Populations from an Avena-Sativa X Avena-Sterilis Cross." Crop Science 19(5): 623-628.://A1979HT43800017

Takeda, K. and K. J. Frey (1985). "Simultaneous Selection for Grain-Yield and Protein Percentage in Backcross Populations from Avena-Sterilis X a-Sativa Matings by Using the Independent Culling Levels Procedure." Theoretical and Applied Genetics 69(4): 375- 382.://A1985AJR3600006

63

Takeda, K. and K. J. Frey (1987). "Improving Grain-Yield in Backcross Populations from Avena-Sativa Xa Sterilis Matings by Using Independent Culling for Harvest Index and Vegetative Growth Index or Unit Straw Weight." Theoretical and Applied Genetics 74(5): 659-665.://A1987K146900019

Tal, M. (1977). "Abscisic-Acid and Germination in Avena-Sterilis L." Israel Journal of Botany 26(2): 100-103.://A1977DT56500006

Taubel, N. (1980). "Elimination of Competition from Wild Oats by Application of Diclofop-Methyl in Spring Wheat in a Pot Trial." Zeitschrift Fur Pflanzenkrankheiten Und Pflanzenschutz-Journal of Plant Diseases and Protection 87(2): 113-120.://A1980JK71800006

Taylor, C. and A. J. Macnab (2000). "Pediatric eye injury due to Avena fatua (Wild Oats)." Journal of Investigative Medicine 48(1): 1A-1A.://000086346600014

Taylor, C. and A. J. Macnab (2001). "Pediatric eye injury due to Avena fatua (wild oats)." Pediatric Emergency Care 17(5): 358- 360.://000171949100011 Objective: We report on florid and unusual ophthalmic physical signs in three children where the trauma was caused by seeds from Avena fatua, a grass common in western North America. Design: Case series and literature review. Setting: Three local emergency departments (ED) during the fall of 1998. Patients or Participants: Three children reporting to an ED with an acutely painful eye from which the foreign body was identified botanically as Avena fatua. Interventions: None. Main Outcome Measures: Symptoms, interventions, duration of problem. Results: Three male children (6, 10, 14 years) presented separately following incidents in which they had sustained direct eye injury. Each child immediately experienced severe pain and profuse watering of the eye. Severe localized edema of the conjunctiva and inflammation was evident with conjunctival vessel injection leading to bleeding, reminiscent of a chemical "burn." Initially, two children appeared to have an eyelash caught behind the lower lid. In both instances, the emergency physicians initially dismissed the possibility of there being a significant foreign body, but because of the severity of the pain, conjunctival vessel injection, and edema, they attempted to remove the "lash." Removal of the foreign body proved difficult in all three cases, requiring far greater traction than anticipated. Intact seedpods had become embedded in the subconjunctival space. Ophthalmic analgesia relieved the pain immediately, but in one child who was treated with topical antibiotic alone, significant pain was experienced for 18 hours, until steroid-antibiotic therapy was instituted. All injuries occurred in late summer when the grass propagates. Conclusions: The physical signs of scleral vasculitis and conjunctival edema can be mistaken for chemical injury or allergic chemosis, but where a foreign body resembling a hair or eyelash is visible, the presence of a seed-pod retained in the subconjunctival space must be considered, particularly if the patient reports exposure to wild grass. Application of local analgesia, foreign body removal, and steroid-antibiotic treatment is recommended.

Taylor, C. R. and O. R. Burt (1984). "Near-Optimal Management Strategies for Controlling Wild Oats in Spring Wheat." American Journal of Agricultural Economics 66(1): 50-60.://A1984SJ41100006

Taylor, H. F. and M. P. C. Loader (1984). "Research on the Control of Wild Oats and Broad-Leaved Weeds by Herbicide Mixtures." Outlook on Agriculture 13(2): 58-68.://A1984SW84700001

Taylor, J. S. and G. M. Simpson (1980). "Endogenous Hormones in after-Ripening Wild Oat (Avena-Fatua) Seed." Canadian Journal of Botany-Revue Canadienne De Botanique 58(9): 1016-1024.://A1980JV50900002

Taylor, M. J., P. Ayres, et al. (1982). "Effect of Surfactants and Oils on the Phytotoxicity of Difenzoquat to Avena-Fatua, Barley and Wheat." Annals of Applied Biology 100(2): 353-363.://A1982NH04700018

Thai, K. M., S. Jana, et al. (1989). "Cell-Membrane Permeability and Ultrastructural Effects of Difenzoquat on Wild Oats (Avena- Fatua)." Weed Science 37(1): 98-106.://A1989T277700018

Thai, K. M., S. Jana, et al. (1985). "Variability for Response to Herbicides in Wild Oat (Avena-Fatua) Populations." Weed Science 33(6): 829-835.://A1985AVB3900017

Thill, D. C., J. T. Odonovan, et al. (1994). "Integrated Weed Management Strategies for Delaying Herbicide Resistance in Wild Oats." Phytoprotection 75: 61-70.://A1994PZ57700008 Herbicide-resistant biotypes of wild oats (Avena fatua) infest most major cereal producing regions in the western United States and Canada. This paper reviews potential integrated weed management strategies that can be used to prevent or delay selection of herbicide-resistant wi Id oats plants. An integrated wild oats management strategy to delay or prevent the development of herbicide resistance should be based on preventing the movement of wild oats seed into the soil. Two ways to achieve this are by preventing the immigration of seed into the field from external sources, and by reducing or eliminating seed production by wild oats already in the field. It is becoming increasingly clear that reliance on continuous herbicide use as the sole means of weed control will fail to eliminate wild oats and other weed seed from the soil seedbank. On the contrary, evidence is mounting that this practice will select for biotypes that are resistant to the herbicides used, especially where herbicides of the same mode of action are used continuously. It is essential, therefore, that herbicides be considered as just one component of an overall integrated system together with cultural control and other management strategies, and that agronomic principles be considered when developing this system.

Thomas, H. and T. Aung (1976). "Transfer of Gene for Mildew Resistance from Avena-Barbata (2n=28) into Cultivated Oat a-Sativa (2n=42)." Heredity 37(DEC): 453-453.://A1976CP68700024

Thomas, H., J. M. Leggett, et al. (1975). "Addition of a Pair of Chromosomes of Wild Oat Avena-Barbata (2n=28) to Cultivated Oat a Avena-Sativa L (2n=42)." Euphytica 24(3): 717-724.://A1975AY30400020 64

Thomas, H., W. Powell, et al. (1980). "Interfering with Regular Meiotic Behavior in Avena-Sativa as a Method of Incorporating the Gene for Mildew Resistance from Avena-Barbata." Euphytica 29(3): 635-640.://A1980LD28100015

Thurston, J. M. (1951). "A Comparison of the Growths of Wild and of Cultivated Oats in Manganese-Deficient Soils." Annals of Applied Biology 38(1): 289-&.://A1951UZ20700018

Thurston, J. M. (1951). "Some Experiments and Field Observations on the Germination of Wild Oat (Avena-Fatua and a- Ludoviciana) Seeds in Soil and the Emergence of Seedlings." Annals of Applied Biology 38(4): 812-&.://A1951UZ21000006

Thurston, J. M. (1954). "A Survey of Wild Oats (Avena-Fatua and a-Ludoviciana) in England and Wales in 1951." Annals of Applied Biology 41(4): 619-636.://A1954UZ22200008

Thurston, J. M. (1966). "Survival of Seeds of Wild Oats (Avena Fatual and Avena Ludoviciana Dur) and Charlock (Sinapis Arvensis L)in Soil under Leys." Weed Research 6(1): 67-&.://A19667638700006

Thurston, J. M. (1979). "Symposium on Wild Oats, Zajecar, Yugoslavia, June 1979." Pans 25(4): 474-475.://A1979JB31700019

Tilsner, H. R. and M. K. Upadhyaya (1985). "Induction and Release of Secondary Seed Dormancy in Genetically Pure Lines of Avena-Fatua." Physiologia Plantarum 64(3): 377-382.://A1985ANH6600016

Tilsner, H. R. and M. K. Upadhyaya (1987). "Action of Respiratory Inhibitors on Seed-Germination and Oxygen-Uptake in Avena- Fatua L." Annals of Botany 59(5): 477-482.://A1987H374700001

Tilsner, H. R. and M. K. Upadhyaya (1989). "The Effect of Ph on the Action of Respiratory Inhibitors in Avena-Fatua Seeds." Annals of Botany 64(6): 707-711.://A1989CF79600012

Tinnin, R. O. and C. H. Muller (1971). "Allelopathic Potential of Avena-Fatua - Influence on Herb Distribution." Bulletin of the Torrey Botanical Club 98(5): 243-&.://A1971L198800001

Tinnin, R. O. and C. H. Muller (1972). "Allelopathic Influence of Avena-Fatua - Allelopathic Mechanism." Bulletin of the Torrey Botanical Club 99(6): 287-292.://A1972P837400003

Todd, B. G. and E. H. Stobbe (1977). "Selectivity of Dichlofop Methyl among Wheat, Barley, Wild Oat (Avena-Fatua) and Green Foxtail (Setaria-Viridis)." Weed Science 25(5): 382-385.://A1977DX81900002

Todd, B. G. and E. H. Stobbe (1980). "The Basis of the Antagonistic Effect of 2,4-D on Diclofop-Methyl Toxicity to Wild Oat (Avena- Fatua)." Weed Science 28(4): 371-377.://A1980KA96500003

Torner, C., J. L. G. Andujar, et al. (1991). "Wild Oat (Avena-Sterilis L) Competition with Winter Barley - Plant-Density Effects." Weed Research 31(5): 301-307.://A1991GJ04200007 The competitive interactions between Avena sterilis ssp. ludoviciana (Dur.) Nyman and winter barley have been studied, taking into consideration the densities of both species. As the density of A. sterilis increased, barley yield decreased exponentially. A 10% reduction in yield was found with wild oat densities ranging from 20-80 panicles m-2, and yield losses reached 50%, with densities of > 300 panicles m-2. Barley grain yield was reduced by wild oats through a reduction in the number of fertile tillers. Climatic conditions during the growing seasons affected the response of barley to wild oat competition. In general, barley yields were relatively unaffected by seeding rates, with similar responses observed in the presence and in the absence of wild oat infestations. However, the highest yield losses were obtained with the lowest seeding rate (100 kg ha-1). Furthermore, low barley densities allowed the wild oat plants to produce more seeds, increasing the potential infestation during the following season.

Turk, M. A. and A. M. Tawaha (2003). "Allelopathic effect of black mustard (Brassica nigra L.) on germination and growth of wild oat (Avena fatua L.)." Crop Protection 22(4): 673-677.://000182541500013 Black mustard (Brassica nigra L.) contains water-soluble substances that inhibited the germination and seedling growth of wild oat (Avena fatua L.). Determining where allelochemicals may be found in B. nigra in the greatest concentration would aid in trying to isolate the compound or compounds responsible for autotoxicity. This study investigated the allelopathic effects of various R nigra plant parts on A. fatua L. germination and seedling growth. Aqueous extracts of B. nigra leaf, stem, flower and root plant part were made to determine their effects on germination and dry weights of hypocotyl and radicle length of 8-d old A. fatua L. seedlings over a range of extract concentrations. Increasing the aqueous extract concentrations of separated R nigra L., plant parts significantly inhibited A. fatua L. germination, seedling length and weight. Radicle length was more sensitive to extract source than seed germination or hypocotyl length. Based on 8-d-old A. fatua L. plant radicle length growth, averaged across all extract concentrations, the degree of toxicity of different R nigra plant parts can be classified in order of decreasing inhibition as follows: leaf, mixture of all plant parts, flower, stem and root. Soil incorporation of fresh B. nigra roots only or both roots and shoots reduced A. fatua emergence, plant height, and dry weight per plant. (C) 2003 Elsevier Science Ltd. All rights reserved.

Tyler, A. (1979). "'Wild Oats' - Epstein,J." New York Times Book Review 84(24): 14-14.://A1979GX92500014

Upadhyaya, M. K. (1986). "Effects of Salicylhydroxamate on Respiration, Seed-Germination and Seedling Growth in Avena-Fatua." Physiologia Plantarum 67(1): 43-48.://A1986C675000007 65

Upadhyaya, M. K. (1987). "Inhibition of Gibberellic Acid-Induced Starch Mobilization by Salicylhydroxamic Acid in Avena-Fatua L Seed." Annals of Botany 59(3): 265-268.://A1987G458000002

Upadhyaya, M. K., A. I. Hsiao, et al. (1986). "Differential Response of Pure Lines of Avena-Fatua L to Substituted Phthalimides - Germination and Endosperm-Mobilization Studies." Annals of Botany 58(4): 455-463.://A1986E520900003

Upadhyaya, M. K., J. M. Naylor, et al. (1982). "Co-Adaptation of Seed Dormancy and Hormonal Dependence of Alpha-Amylase Production in Endosperm Segments of Avena-Fatua." Canadian Journal of Botany-Revue Canadienne De Botanique 60(7): 1142- 1147.://A1982PC17600015

Upadhyaya, M. K., J. M. Naylor, et al. (1982). "The Physiological-Basis of Seed Dormancy in Avena Fatua L .1. Action of the Respiratory Inhibitors Sodium-Azide and Salicylhydroxamic Acid." Physiologia Plantarum 54(4): 419-424.://A1982NM59100007

Upadhyaya, M. K., J. M. Naylor, et al. (1983). "The Physiological-Basis of Seed Dormancy in Avena-Fatua .2. on the Involvement of Alternative Respiration in the Stimulation of Germination by Sodium-Azide." Physiologia Plantarum 58(1): 119-123.://A1983QS64000021

Upadhyaya, M. K., G. M. Simpson, et al. (1981). "Levels of Glucose-6-Phosphate and 6-Phosphogluconate Dehydrogenases in the Embryos and Endosperms of Some Dormant and Non-Dormant Lines of Avena-Fatua During Germination." Canadian Journal of Botany-Revue Canadienne De Botanique 59(9): 1640-1646.://A1981MH75400011

Van Wychen, L. R., B. D. Maxwell, et al. (2004). "Wild oat (Avena fatua) habitat and water use in cereal grain cropping systems." Weed Science 52(3): 352-358.://000221464600006 The advent of site-specific weed management has generated research aimed at predicting weed spatial distributions from existing weed maps or correlations with soil properties and edaphic factors. Forecasting the spatial distribution of annual weeds requires knowledge of fecundity, dispersal, management, and suitable habitat distribution. We hypothesized that wild oat habitat was limited by field-scale heterogeneity in plant-available water. We eliminated seed number and dispersal limitations by seeding wild oat in areas with and without historical wild oat patches in three similarly managed spring wheat fields that differed in soil properties and wild oat infestations and were situated within a 160-km radius of Great Falls, MT Wild oat habitat was quantified by wild oat leaf area growth rate, mature shoot biomass, seeds produced per plant, biomass water use efficiency, and competitive ratio with spring wheat. Soil texture and plot elevation correlated with existing wild oat patch areas in individual fields, but no site properties consistently correlated with wild oat patch areas in all three fields. Soil water use (SWU) and almost all habitat-defining variables for wild oat were similar between historic patch and nonpatch areas. Wild oat grew and produced seed regardless of existing patch boundaries and field-scale heterogeneity in SWU. This research suggested that (1) wild oat habitat may be unlimited in cereal grain cropping systems of the Northern Great Plains and (2) soil properties are a poor predictor of weed distribution for a generalist such as wild oat.

Vangardingen, P. R., C. E. Jeffree, et al. (1989). "Variation in Stomatal Aperture in Leaves of Avena-Fatua L Observed by Low- Temperature Scanning Electron-Microscopy." Plant Cell and Environment 12(9): 887-897.://A1989CJ68800003

Vecchio, V., J. Gasquez, et al. (1982). "Morphological and Enzymatic Variability in a Mixed Population of Avena-Fatua L and Avena-Sterilis L." Weed Research 22(5): 263-269.://A1982PH70700004

Wahl, I. (1970). "Prevalence and Geographic Distribution of Resistance to Crown Rust in Avena Sterilis." Phytopathology 60(5): 746- &.://A1970G234000001

Walker, R. P., W. M. Waterworth, et al. (1994). "Gibberellin-Photoaffinity Labeling of Wild Oat (Avena-Fatua L) Aleurone Protoplasts." Plant Growth Regulation 15(3): 271-279.://A1994PW97500010 Aleurone protoplasts of wild oat (Avena fatua L.), and subcellular fractions isolated from them, were photoaffinity labeled using the synthetic gibberellin (GA) derivative GA(4)-17-yl-1'-(1'-thia)propan-3'-ol-4-azido-5-[I-125]iodosalicylate. Labeled polypeptides were identified by electrophoresis under denaturing conditions followed by autoradiagraphy. GA-photoaffinity labeling of both intact protoplasts and isolated subcellular fractions led to the covalent attachment of the reagent to many polypeptides. A 50 kD polypeptide in the soluble fraction of homogenates of aleurone protoplasts GA-photoaffinity labeled in vivo showed specific binding. The biologically active GA(1), GA(4) and GA(4)-17-yl-1'(1'-thia)propan-3'-ol-4- azidosalicylate completed for binding whereas the biologically inactive GA(8) and GA(34) did not. The GA-photoaffinity labeling characteristics of this polypeptide suggested that:it might interact specifically with biologically active GAs in vivo. Attempts to detect specific GA-binding in in vitro GA-photoaffinity labeling experiments met with only limited success perhaps indicating the labile nature of specific binding observed in vivo. The potential of GA-photoaffinity labeling for identifying GA-binding proteins in aleurone and other GA-responsive tissues is discussed.

Walker, R. P., W. M. Waterworth, et al. (1993). "Preparation and Polypeptide Composition of Plasma-Membrane and Other Subcellular-Fractions from Wild Oat (Avena-Fatua) Aleurone." Physiologia Plantarum 89(2): 388-398.://A1993MD17900022 Plasma membranes can be isolated from a variety of plant tissues by first preparing a post-mitochondrial membrane fraction enriched in plasma membranes, by differential centrifugation, and partitioning this on a dextran-polyethylene glycol two-phase system. With wild oat aleurone, however, we observed that differential centrifugation could not be used to produce a microsomal fraction enriched in plasma membrane. Approximately 70% of the plasma membrane in aleurone homogenates was pelleted by sequential centrifugation at 100 g x 10 min and 1 000 g x 10 min. The remainder sedimented at 112 000 g x 1 h. All the material that was pelletable by centrifugation was, therefore, subjected to dextran- 66

polyethylene glycol two-phase partitioning. The plasma membrane marker enzymes glucan synthase II (GSII, EC 2.4.1.34) and UDP-glucose:sterol glucosyltransferase (SGT, EC 2.4.1.) were enriched in the upper phase, whereas cytochrome c oxidase activity (EC 1.9.3.1), a mitochondrial marker enzyme, was depleted. The presence of endoplasmic reticulum (ER) and protein body membranes in the phase system was assessed by probing western blots, of SDS-PAGE separated proteins, with polyclonal antiserum either to binding protein (BiP, an ER marker) or to tonoplast intrinsic protein (TIP, a protein body membrane marker). BiP and TIP were present in the lower phase, but were not detected in the upper phase. In addition, the polypeptide patterns of material in the upper and lower phases were very different. These observations suggested that high purity aleurone plasma membrane had been isolated. Although the procedure for isolating plasma membranes was applicable to both aleurone protoplasts and layers, the polypeptide patterns of plasma membranes prepared from these sources were very different. The major protein components of wild oat aleurone were 7 S and 12 S storage globulins. These proteins were present in the lower phase, but not in the plasma membrane enriched upper phase, after aqueous two-phase partitioning. Differential centrifugation studies showed that it was necessary to homogenise aleurone in a buffer of pH 6.0 or less if a soluble protein fraction. essentially devoid of storage globulins, was to be obtained. The use of these fractionation techniques is discussed in relation to photoaffinity labelling of gibberellin (GA)-binding proteins in aleurone.

Wall, D. A. (1993). "Comparison of Green Foxtail (Setaria-Viridis) and Wild Oat (Avena-Fatua) Growth, Development, and Competitiveness under 3 Temperature Regimes." Weed Science 41(3): 369-378.://A1993MD60800008 In a replacement series study, barley was more competitive than green foxtail and wild oat at 28/22 and 22/16 C. Wild oat was more competitive with green foxtail at 22/16 C than at 28/22 C. Maximum green foxtail dry weight and leaf area was produced at 28/22 C. As temperature increased, maximum dry weight and leaf area occurred earlier during plant growth. Similarly maximum wild oat leaf area occurred earlier with increasing temperature, but the greatest leaf area was observed at 16/10 C. Leaf area ratio (LAR) and relative growth rate (RGR) of wild oat did not differ markedly between temperature regimes. Green foxtail LAR was higher while RGR was lower at 16/10 C than at 22/16 or 28/22 C. Under the 16/10 C regime green foxtail produced little mature seed. Greatest seed numbers were produced at 28/22 C. Wild oat produced mature seed under all temperature regimes with the greatest seed numbers produced at 22/16 C.

Walter, H., W. Koch, et al. (1980). "Effect of the Type of Application on the Penetration and Translocation of Diclofop-Methyl in Wild Oats (Avena-Fatua L)." Weed Research 20(6): 325-331.://A1980LN60400001

Waterson, H. A. and G. J. Davies (1973). "Distribution of Avena-Fatua L, Avena-Strigosa Schreb and Agropyron-Repens (L) Beauv in Barley Crops in West of Scotland." Weed Research 13(2): 192-199.://A1973Q295300007

Weaver, S. E., M. J. Kropff, et al. (1993). "A Simulation-Model of Avena-Fatua L (Wild-Oat) Growth and Development." Annals of Applied Biology 122(3): 537-554.://A1993LZ98100014 An eco-physiological simulation model of the growth and development of A vena fatua was parameterised and tested. The model simulates growth of A. fatua, in kg dry matter ha-1 day-1, from sowing to maturity as a function of irradiance, temperature and various species characteristics. Parameter values were derived from the literature and from field experiments, including both autumn and spring sowings of A. fatua over three years at two sites in southern England. With two exceptions, a single set of parameter values was sufficient to accurately simulate the emergence, growth, and development of both autumn and spring cohorts over all years and sites. The two exceptions were the result of differences between autumn and spring cohorts of A. fatua in the rate of early leaf area growth and in the relationship between specific leaf area and developmental stage.

Weaver, S. E., M. J. Kropff, et al. (1994). "A Simulation-Model of Competition between Winter-Wheat and Avena-Fatua for Light." Annals of Applied Biology 124(2): 315-331.://A1994NM34500010 An eco-physiological simulation model of competition between Avena fatua and winter wheat (Triticum aestivum cv. Avalon) for light was parameterised and tested. The model simulates growth of each species, in kg dry matter ha-1 day-1, from sowing to maturity as a function of irradiance. temperature and various species characteristics. Parameter values were derived from the literature and from field experiments with each species grown in monoculture. Model performance was tested against 50:50 mixtures of the two species grown at two sites in southern England. Sensitivity analyses were performed in which the canopy height of each species and the density and time of emergence of A. fatua were systematically varied while all other parameters were left unchanged. Accurate simulations of growth in mixtures depended upon an accurate description of the canopy height of each species throughout the growing season. Model predictions of winter wheat yield losses in relation to A. fatua density and time of emergence showed good agreement with previously published data.

Webster, G. R. B., C. F. Shaykewich, et al. (1978). "Availability of Herbicide Trifluralin for Control of Wild Oats as Influenced by Soil Characteristics in 4 Manitoba Soils." Canadian Journal of Soil Science 58(3): 397-404.://A1978FU35500013

Westdal, P. H. and Richards.Hp (1969). "Susceptibility of Cereals and Wild Oats to an Isolate of Aster Yellows Pathogen." Canadian Journal of Botany 47(5): 755-&.://A1969D415300019

Whalley, R. D. B. and J. M. Burfitt (1972). "Ecotypic Variation in Avena-Fatua L, a Avena-Sterilis L (a Avena-Ludoviciana), and a Avena-Barbata Pott in New South-Wales and Southern Queensland." Australian Journal of Agricultural Research 23(5): 799-&.://A1972O213900007

Whitting.Wj, J. Hillman, et al. (1970). "Light and Temperature Effects on Germination of Wild Oats." Heredity 25(NOV): 641-&.://A1970I858800012

Wiese, A. F. and R. S. Dunham (1954). "Fall Applications of Ipc and Cipc for Killing Wild Oats (Avena-Fatua) Prior to Sowing Oats." Agronomy Journal 46(8): 358-361.://A1954UC65300002 67

Wille, M. J., D. C. Thill, et al. (1998). "Wild oat (Avena fatua) seed production in spring barley (Hordeum vulgare) is affected by the interaction of wild oat density and herbicide rate." Weed Science 46(3): 336-343.://000074501300011 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 oar seed production increased from 180 to 9,950 seed m(-2) without herbicide, and from 0 to 2,810 seed m(-2) using 0.53 kg ai ha(-1) imazamethabenz. This general pattern was modeled using a cumulative logistic function. Estimates from this model indicated that < 1 wild oat seed m(-2) was produced at population densities of less than or equal to 20 plants at any imazamethabenz rate. Imazamethabenz rates of 0.26 kg ha(-1) or greater at wild oat densities of less than approximately 190 plants m(-2) did 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(-1) resulted in substantially greater wild oat seed production compared to the recommended rate.

Willenborg, C. J., W. E. May, et al. (2005). "Influence of wild oat (Avena fatua) relative time of emergence and density on cultivated oat yield, wild oat seed production, and wild oat contamination." Weed Science 53(3): 342-352.://000229456600010 Wild oat is a serious weed in cultivated oat because there are no herbicides to selectively control it. Considering the effect of time of emergence on weed-crop interference is critical for the development of accurate crop yield loss models and weed density thresholds. Therefore, field experiments were conducted at two locations in Saskatchewan, Canada, in 2002 and 2003 to determine the effect of wild oat density and time of emergence on cultivated oat yield and quality. Wild oat was planted at 50 growing degree day (GDD) intervals ranging from 100 GDD before to 100 GDD after crop planting. Wild oat density ranged from 0 to 320 plants m(-2). High densities of early emerging wild oat greatly reduced cultivated oat yield and increased wild oat contamination, with observed oat yield losses as great as 70% and wild oat contamination levels of 15%. Wild oat that emerged before cultivated oat caused considerably more yield and quality loss and had higher reproductive output than wild oat that emerged after cultivated oat. The yield loss caused by individual wild oat plants at low densities (parameter I) ranged from 0.40 to 0.49%. The effect of relative time of wild oat emergence (parameter C) always varied sigh in significantly between site-years. However, little variation in absolute values wit years was observed for cultivated oat yield loss, wild oat seed production, and wild oat contamination, suggesting that relative time of wild oat emergence influences all similarly. The results of this study emphasize both the need to control early emerging wild oat, as well as the importance of time of emergence in the prediction of crop conducting an emergence study based on yield loss. Furthermore, our approach thermal time is novel and demonstrates a robust, mechanistic method of estimating crop yield losses due to relative time of emergence.

Willenborg, C. J., B. G. Rossnagel, et al. (2005). "Effects of relative time of emergence and density of wild oat (Avena fatua L.) on oat quality." Canadian Journal of Plant Science 85(3): 561-567.://000232484400004 Selective control of wild oat with herbicides is not possible in oat (Avena sativa L.) crops and, consequently, the high quality that is imperative to market success is continually jeopardized. Field experiments were conducted to determine the effects of relative time of emergence and density of wild oat on oat kernel size and weight. Thousand-kernel weight and percentage plump kernels were generally reduced, and percentage thin kernels were generally increased with increasing wild oat densities and earlier wild oat emergence. However, reductions were generally small and occurred only at the highest wild oat densities, which are not commonly observed in commercial fields. These findings suggest that oat growers should not control wild oat based on perceived reductions in oat physical grain quality.

Williams, G. C. and J. M. Thurston (1964). "Effect of Temperature in Sack-Drier on Survival of Insects ( Oryzaephilus Surinamensis ) L ) ) Col Silvanidae ) ) + Weed Seeds ) Avena Fatua L + a Ludoviciana Dur )." Annals of Applied Biology 53(1): 29-&.://A19644410A00014

Wilson, B. J. (1978). "Long-Term Decline of a Population of Avena-Fatua L with Different Cultivations Associated with Spring Barley Cropping." Weed Research 18(1): 25-31.://A1978EW83100004

Wilson, B. J. (1979). "Effect of Controlling Alopecurus-Myosuroides Huds and Avena-Fatua L Individually and Together, in Mixed Infestations on the Yield of Wheat." Weed Research 19(3): 193-199.://A1979HD91700006

Wilson, B. J. (1979). "Post-Emergence Control of Wild Oats in Queensland with Difenzoquat, Flamprop-Methyl, Dichlofop Methyl and Barban." Australian Journal of Experimental Agriculture 19(96): 108-117.://A1979GN11100018

Wilson, B. J. (1981). "The Influence of Reduced Cultivations and Direct Drilling on the Long-Term Decline of a Population of Avena- Fatua L in Spring Barley." Weed Research 21(1): 23-28.://A1981LU14200004

Wilson, B. J. (1985). "Effect of Seed Age and Cultivation on Seedling Emergence and Seed Decline of Avena-Fatua L in Winter Barley." Weed Research 25(3): 213-219.://A1985AHY8500006

Wilson, B. J., R. Cousens, et al. (1990). "The Response of Spring Barley and Winter-Wheat to Avena-Fatua Population-Density." Annals of Applied Biology 116(3): 601-609.://A1990DV06700022

Wilson, B. J. and G. W. Cussans (1975). "Study of Population-Dynamics of Avena-Fatua L as Influenced by Straw Burning, Seed Shedding and Cultivations." Weed Research 15(4): 249-258.://A1975AH54600008

Wilson, B. J. and G. W. Cussans (1978). "Effects of Herbicides, Applied Alone and in Sequence, on Control of Wild-Oats (Avena- Fatua) and Broad-Leaved Weeds, and on Yield of Winter-Wheat." Annals of Applied Biology 89(3): 459-466.://A1978FK82300012 68

Wilson, B. J. and N. C. B. Peters (1982). "Some Studies of Competition between Avena-Fatua L and Spring Barley .1. the Influence of Avena-Fatua on Yield of Barley." Weed Research 22(3): 143-148.://A1982NV87900004

Wilson, B. J. and W. A. Taylor (1978). "Field Trials with Controlled Drop Application of Barban and Difenzoquat for Control of Wild Oats (Avena-Fatua L) in Spring Barley." Weed Research 18(4): 215-221.://A1978FM87600005

Wimschneider, W., G. Bachthaler, et al. (1990). "Competitive Effects of Avena-Fatua L on Wheat as a Basis for Effective Weed- Control." Weed Research 30(1): 43-52.://A1990CL78800006

Winfield, R. J. and J. J. B. Caldicott (1975). "Difenzoquat, "1,2-Dimethyl-3,5-Diphenylpyrazolium Ion, a Selective Herbicide for Control of Wild Oats (Avena Spp) in Wheat and Barley." Pesticide Science 6(3): 297-303.://A1975AV50000010

Wong, L. S. L., R. I. H. McKenzie, et al. (1983). "The Inheritance of Resistance to Puccinia-Coronata and of Floret Characters in Avena-Sterilis." Canadian Journal of Genetics and Cytology 25(4): 329-335.://A1983RE78100003

Wright, G. M. and N. L. Shillito (1974). "False Wild Oats in Mapua-70 Crop." New Zealand Journal of Agriculture 128(4): 45-45.://A1974S965600015

Wright, M. (1986). "The Acquisition of Gravisensitivity During the Development of Nodes of Avena-Fatua." Journal of Plant Growth Regulation 5(1): 37-47.://A1986C266500004

Wright, M. and P. Doherty (1985). "Auxin Levels in Single Half Nodes of Avena-Fatua Estimated Using High-Performance Liquid- Chromatography with Coulometric Detection." Journal of Plant Growth Regulation 4(2): 91-100.://A1985AMG2400004

Xie, H. S., A. I. Hsiao, et al. (1994). "Effect of Shading on Activity of Imazamethabenz and Fenoxaprop in Wild-Oat (Avena-Fatua)." Weed Science 42(1): 66-69.://A1994NJ44200011 Growth chamber and greenhouse studies were conducted to evaluate effects of long-term low-light intensity on wild oat control with imazamethabenz and fenoxaprop. Seventy percent shading imposed during the entire experimental period resulted in enhanced activities for both herbicides applied at early and later growth stages. Such shading also reduced wild oat regrowth following application of imazamethabenz and fenoxaprop. When applied to plants exposed to 70 to 90% prespraying shading, both herbicides had phytotoxicity similar to, or better, than plants grown under continuous shading. Postspraying shading has less effect on herbicidal activity than prespraying shading or prolonged shading, especially with imazamethabenz. Full-light treatment more adversely affected fenoxaprop activity than imazamethabenz activity.

Xie, H. S., A. I. Hsiao, et al. (1994). "Impact of Temperature on the Phytotoxicity of Imazamethabenz and Fenoxaprop to Wild Oat (Avena-Fatua)." Crop Protection 13(5): 370-380.://A1994NX67900009 Controlled environmental experiments were conducted to determine the effect of short- and long-term temperature stresses on the phytotoxicity of imazamethabenz and fenoxaprop to wild oat. Short-term (16 h) temperature stresses, imposed either before or after herbicide spraying, had no influence on imazamethabenz phytotoxicity. However, short-term temperature stresses affected the phytotoxicity of fenoxaprop, with pretreatment high temperature (35-degrees-C) reducing the phytotoxicity and post-treatment high temperatures (25-35-degrees-C) enhancing the phytotoxicity. Imazamethabenz phytotoxicity was similar at long-term temperatures between 20/15-degrees-C and 30/20-degrees-C. Constant low temperature (10/5-degrees-C) reduced imazamethabenz phytotoxicity, whereas such temperature had no effect if only imposed before or after spraying. Among three constant temperature regimes, fenoxaprop was more effective at 20/15-degrees-C, followed by 10/5-degrees-C. Constant high temperature (30/20-degrees-C) greatly decreased fenoxaprop phytotoxicity. The adverse effect of long-term low temperature on fenoxaprop phytotoxicity was due mainly to the influence of post-treatment low temperature, whereas long-term high temperature imposed either before or after spraying reduced the phytotoxicity of fenoxaprop.

Xie, H. S., A. I. Hsiao, et al. (1996). "Influence of temperature and light intensity on absorption, translocation, and phytotoxicity of fenoxaprop-ethyl and imazamethabenz-methyl in Avena fatua." Journal of Plant Growth Regulation 15(2): 57-62.://A1996VG47400002 The absorption and translocation of fenoxaprop-ethyl and imazamethabenz-methyl were investigated in wild oat (Avena fatua L.) plants grown under different temperature and light intensity conditions by using C-14 tracer techniques. The phytotoxicity of both herbicides, applied as individual droplets, was also determined under similar environments. The absorption of fenoxaprop-ethyl and imazamethabenz-methyl was increased by high temperature (30/20 degrees C) and to a lesser extent by 70% shading; low temperature (10/5 degrees C) had limited effect on the absorption. The basipetal translocation of fenoxaprop-ethyl was not affected by high temperature, and the increase in imazamethabenz-methyl translocation at high temperature was likely the result of the increased absorption. Low temperature decreased total translocation and translocation efficiency in both fenoxaprop-ethyl and imazamethabenz-methyl, Low light intensity tended to reduce the efficiency of basipetal translocation of both herbicides. Fenoxaprop-ethyl phytotoxicity was reduced by high temperature but not by low temperature. Temperature had little effect on imazamethabenz-methyl effectiveness. Under 70% shading, the phytotoxicity of both herbicides was enhanced.

Xie, H. S., A. I. Hsiao, et al. (1997). "Effect of environment on wild oat (Avena fatua) control with imazamethabenz or fenoxaprop tank-mixed with additives or MCPA." Journal of Plant Growth Regulation 16(2): 63-67.://A1997WY02800002 Greenhouse and growth chamber experiments were conducted to determine the effect of soil moisture and temperature on the phytotoxicity in wild oat of imazamethabenz or fenoxaprop tank-mixed with certain additives or MCPA, The surfactants Agral 90 at 0.5% and Enhance at 0.5% increased imazamethabenz phytotoxicity under both moist and drought conditions. These surfactants had no significant effect on fenoxaprop phytotoxicity regardless of the soil moisture regimes. Fenoxaprop activity was increased by ammonium sulfate [(NH4)(2)SO4] at 1% but only under well watered conditions, 69

Wild oat control with imazamethabenz was also slightly enhanced in a well watered regime by the addition of sodium bisulfate (NaHSO4) at 0.13%. At high temperature (30/20 degrees C) and low temperature (10/5 degrees C), the phytotoxicity of imazamethabenz was increased when tank-mixed with Agral 90 at 0.25% or NaHSO4 at 0.13% compared with that when imazamethabenz was applied alone, if soil moisture was adequate, There was no such increase under conditions of drought and high temperature, (NH4)(2)SO4 at 1% did not significantly affect imazamethabenz performance irrespective of temperature/soil moisture conditions. The phytotoxicity to wild oat of imazamethabenz or fenoxaprop was not changed by tank-mixing with MCPA isooctyl ester at 300 g a.i./ha, regardless of soil moisture levels, The reduced fenoxaprop phytotoxicity in wild oat due to moisture stress was not readily alleviated by the inclusion of selected additives or MCPA in the tank mixture.

Xie, H. S., A. I. Hsiao, et al. (1997). "Influence of drought on graminicide phytotoxicity in wild oat (Avena fatua) grown under different temperature and humidity conditions." Journal of Plant Growth Regulation 16(4): 233-237.://A1997YJ82100008 Controlled environmental experiments were carried out to determine the phytotoxicity of several graminicides on wild oat (Avena futua L.) as influenced by combination of drought and temperature stress or drought and low relative humidity. Compared with unstressed conditions (20/15 degrees C plus adequate soil moisture), imazamethabenz phytotoxicity to wild oat was reduced significantly when plants were exposed to a combination of drought and high temperature (30/20 degrees C) stress. Imazamethabenz phytotoxicity was reduced almost as much by high temperature stress alone as by a combined temperature and drought stress. When herbicides were applied to wild oat plants subjected to drought alone or to drought plus high temperature, the observed reduction in phytotoxicity from greatest to least was: fenoxaprop = diclofop > flamprop > imazamethabenz. Fenoxaprop performance was most inhibited by the combination of drought plus high temperature, although drought alone and to a lesser degree, high temperature alone, inhibited fenoxaprop action. High temperature had an adverse effect on the efficacy of fenoxaprop at lower application rates. Raising fenoxaprop application rates to 400 g ha(-1) overcame the inhibition caused by high temperature alone but only partially alleviated the effect of drought combined with high temperature. When plants were grown under a low temperature regimen the imposition of drought stress had little effect on imazamethabenz phytotoxicity but did reduce fenoxaprop phytotoxicity. At 25/15 degrees C drought reduced the phytotoxicity of fenoxaprop and diclofop greatly but had no significant impact on the performance of any of the herbicides examined, regardless of soil moisture regimen.

Xie, H. S., A. I. Hsiao, et al. (1996). "Influence of water stress on absorption, translocation and phytotoxicity of fenoxaprop-ethyl and imazamethabenz-methyl in Avena fatua." Weed Research 36(1): 65-71.://A1996UC93700008 The influence of water stress on the absorption and translocation of C-14-labelled fenoxaprop-ethyl and imazamethabenz- methyl in Avena fatua L. (wild oat) was studied. The phytotoxicity to A. fatua of both herbicides with a droplet application was also examined under water stress conditions. The absorption of both fenoxaprop-ethyl and imazamethabenz-methyl was seduced by water stress when the plants were harvested within 24 h after herbicide application. Up to 48 h after the application, the translocation out of the treated lamina of both herbicides, based on percentage of applied C-14, was reduced under water stress conditions. When harvested 96 h after herbicide application, however, water stress no longer significantly affected the absorption and translocation of either herbicide. When the herbicides were applied as individual droplets, water stress reduced the phytotoxicity of fenoxaprop-ethyl but not that of imazamethabenz-methyl. It is concluded that the changes in herbicide absorption and translocation may not be the major physiological processes associated with differential whole-plant response of A. fatua to fenoxaprop-ethyl and imazamethabenz-methyl under water stress.

Xie, H. S., W. A. Quick, et al. (1994). "Effect of Drought and Formulation on Wild Oat (Avena-Fatua) Control with Imazamethabenz and Fenoxaprop." Crop Protection 13(3): 195-200.://A1994NH86700004 Experiments were conducted in the greenhouse to determine the effects of drought, rehydration and herbicide formulation on the phytotoxicity to wild oat of imazamethabenz and fenoxaprop. Under both well-watered and drought conditions, the activity of imazamethabenz with the liquid concentrate formulation tended to be higher than the suspension concentrate formulation, and the activity of the soluble powder formulation was enhanced by the surfactant Agral 90. Long-term drought did not reduce imazamethabenz phytotoxicity to wild oat with any formulations examined, whereas the phytotoxicity of three fenoxaprop formulations, fenoxaprop-ethyl, fenoxaprop-p-ethyl and fenoxaprop-p-ethyl plus safener, was decreased by long-term drought, Fenoxaprop-ethyl activity was reduced by drought stress imposed before spraying, after spraying or, most significantly, both before and after spraying. The adverse effect of drought on fenoxaprop-ethyl activity was still evident even when daily watering was resumed 8 to 48 h before spraying, although plants rehydrated in this way were more susceptible to fenoxaprop-ethyl than plants where rehydration was deferred until after spraying.

Xie, H. S. S., B. C. Caldwell, et al. (1995). "Spray Deposition of Fenoxaprop and Imazamethabenz on Wild Oat (Avena-Fatua) as Influenced by Environmental-Factors." Weed Science 43(2): 179-183.://A1995RB70200004 The effect of soil moisture, temperature, and light intensity on the spray deposition of fenoxaprop and imazamethabenz applied to wild oat plants was examined by using fluorescent tracer dye, Based on either biomass or total leaf area, the apparent deposition of the two herbicides diminished in the following order: shading > low temperature greater than or equal to drought greater than or equal to ''optimum'' > high temperature, The enhanced phytotoxicity of both herbicides under shading could be associated with increased spray deposition; and reduced fenoxaprop phytotoxicity under high temperature stress could be related to reduced deposition, Changes in spray deposition were attributed mainly to differences in herbicide interception due to altered plant morphology. Reduced retention for both herbicides was exhibited only in the plants grown at high temperature, Under ''optimum'' conditions, fenoxaprop phytotoxicity was directly associated with leaf orientation and thus with the proportion of projected leaf area at the time of herbicide spraying, Given similar application conditions, spray deposition of fenoxaprop and imazamethabenz on wild oat could be estimated by determining the ratio of the projected leaf area, as measured by an image analyzer to the total leaf area.

Yaduraju, N. T. and V. S. Mani (1987). "The Influence of Delayed Planting and Seed Bed Preparation on the Competition of Wild Oats in Wheat." Indian Journal of Agronomy 32(3): 299-301.://A1987T284100034

70

Yang, Q., L. Hanson, et al. (1999). "Genome structure and evolution in the allohexaploid weed Avena fatua L-(Poaceae)." Genome 42(3): 512-518.://000080857700018 Allohexaploid wild oat, Avena fatua L. (Poaceae; 2n = 6x = 42), is one of the world's worst weeds, yet unlike some of the other Avena hexaploids, its genomic structure has been relatively little researched. Consequently, in situ hybridisation was carried out on one accession of A. fatua using an 18S-25S ribosomal DNA (rDNA) sequence and genomic DNA from A. strigosa (AA-genome diploid) and A. clauda (CC-genome diploid) as probes. Comparing these results with those for other hexaploids studied previously: (i) confirmed that the genomic composition of A. fatua was similar to the other hexaploid Avena taxa (i.e., AACCDD), (ii) identified major sites of rDNA on three pairs of A/D-genome chromosomes, in common with other Avena hexaploids, and (iii) revealed eight chromosome pairs carrying intergenomic translocations between the A/D- and C-genomes in the accession studied. Based on karyotype structure, the identity of some of these recombinant chromosomes was proposed, and this showed that some of these could be divided into two types, (i) those common to all hexaploid Avena species analysed (3 translocations) and (ii) one translocation in this A. fatua accession not previously observed in reports on other hexaploid Avena species. If this translocation is found to be unique to A. fatua, then this information, combined with more traditional morphological data, will add support to the view that A. fatua is genetically distinct from other hexaploid Avena species and thus should retain its full specific status.

Yeung, E. C., C. C. Chinnappa, et al. (1987). "Seed Detachment in Avena-Fatua L." Weed Research 27(6): 391-396.://A1987L205200001

Youngs, V. L. and D. M. Peterson (1973). "Protein Distribution in Oat (Avena-Sterilis L) Kernel." Crop Science 13(3): 365-367.://A1973Q094300023

Zand, E. and H. J. Beckie (2002). "Competitive ability of hybrid and open-pollinated canola (Brassica napus) with wild oat (Avena fatua)." Canadian Journal of Plant Science 82(2): 473-480.://000175508800031 Zand, E. and Beckie, H. J. 2002. Competitive ability of hybrid and open-pollinated canola (Brassica napus) with wild oat (Avena fatua). Can. J. Plant Sci. 82: 473-480. The competitiveness of three hybrid and three open-pollinated canola cultivars against two wild oat populations was determined under controlled environment conditions at two plant densities and five canola:wild oat ratios (100:0, 75:25, 50:50, 25:75, 0:100). Analysis of replacement series and derivation of relative crowding coefficients (RCC), based on shoot dry weight or leaf area, indicated that hybrid canola cultivars were twice as competitive than open-pollinated cultivars when weed interference was relatively high (i.e., high plant density and vigorous wild oat growth). Little difference in competitiveness among cultivar types was apparent when weed interference was lower. The results of this study suggest that hybrid canola cultivars may be best suited for use in an integrated weed management program, particularly for farmers of organic or low input cropping systems.

Zavas, T., L. Symeonidis, et al. (1991). "Differential Response to Al-Toxicity of 2 Populations of Avena-Sterilis L." Journal of Agronomy and Crop Science-Zeitschrift Fur Acker Und Pflanzenbau 167(4): 277-284.://A1991GQ82800010 Negative correlation between Al concentrations and root and shoot growth was detected in two populations of Avena sterilis L. investigated [one from a bauxite area (1) and the other (2) from a pasture area]. Al solubility and Al content of plant tissue depend on pH levels of nutrient solution. Population-1 although in all treatments contained greater amounts of Al in its roots, was grown better than population-2, suggesting that population-1 is more tolerant than population-2. Al content of shoot and root of both populations was greater at pH 10.0 than at pH 4.5, a fact that may indicate that Al ions in alkaline medium are preferentially absorbed than Al ions in acid medium. The better growth of both populations observed in all Al concentrations at pH 10.0, where Al content of plant tissues was greater, may indicate that Al forms predominant in acid nutrient medium are more harmful than Al forms in alkaline nutrient medium.

Zemanek, J. and A. Ludvova (1988). "The Effect of Crop Rotations and Herbicides on the Occurrence of Wild Oat (Avena-Fatua L) in Long-Continued Stationary Trials." Rostlinna Vyroba 34(4): 349-356.://A1988N281300002

Zemanek, J. and J. Mikulka (1985). "The Control of Wild Oat (Avena Fatua L) by Herbicides and Crop Rotations." Rostlinna Vyroba 31(4): 395-400.://A1985AGT2300008

Zorner, P. S., R. L. Zimdahl, et al. (1984). "Sources of Viable Seed Loss in Buried Dormant and Non-Dormant Populations of Wild Oat (Avena-Fatua L) Seed in Colorado." Weed Research 24(2): 143-150.://A1984SH73600010

Zwar, J. A. and R. Hooley (1986). "Hormonal-Regulation of Alpha-Amylase Gene-Transcription in Wild Oat (Avena-Fatua L) Aleurone Protoplasts." Plant Physiology 80(2): 459-463.://A1986A192500029