首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
R.D. AYLING 《Weed Research》1976,16(5):301-304
The herbicide Tordon 50D (picloram+2,4-D) affected the integrity of the nucleus and cell membranes in Pinus radiata needle segments and caused the swelling of internal chloro plast membranes and the eventual disintegration of the chloroplasts. Tordon 22K (picloram) only affected chloro plast structure. Both herbicides had similar adverse effects on cell membranes and chloroplasts of Eucalyptus viminalis.  相似文献   

2.
The effectiveness of‘Tordon 50-d’(5% a.i. picloram plus 20% a.i. 2,4-D both as the triisopropanolamine salts) and various mixtures of 2,4,5-T and picloram were tested for the control of blackberry (Rubus fruticosus L. agg.) in Victoria, Australia. A high correlation was obtained between the % reduction in live canes and the % kill of crowns 13 months after Rubus procerus P.J. Muell. thickets were sprayed with 2,4,5-T or‘Tordon 50-d'. Counting the number of live canes is, therefore, a convenient method of comparing the efficacy of these herbicides for the control of blackberry. ‘Tordon 50-d’was generally more effective than 2,4,5-T but stimulation of suckering from roots was recorded at one site when low rates of‘Tordon 50-d’were used. It was necessary to add high dose rates of‘Tordon 50-d’to 2,4,5-T before there were worthwhile improvements in weed control.‘Tordon 5–20’(5% a.i. picloram as triisopropanolamine salt plus 20% 2,4,5-T as the ethyl hexyl ester) was only slightly more effective in controlling blackberry than‘Tordon 50-d'. The cost and soil residue problems associated with picloram should limit its use as an additive to 2,4,5-T for the control of blackberry in Australia.  相似文献   

3.
The efficacy of mowing and of herbicides (applied by spot spraying or boom spraying) for the selective control of Cirsium arvense (L) Scop, in Victorian pastures was investigated. The pastures were based on Trifolium repens L. and Lolium perenne L. A general decline in the number of thistle shoots in untreated plots at three out of five sites was recorded.‘Tordon 50-d' (5% a.i. picloram and 20% a.i. 2,4-D as triisopropanolamine salts), MCPB, 2,4-DB and mowing resulted in suppression, but not eradication, during the experimental period. MCPB and 2,4-DB are the most suitable herbicides for selective control. The use of‘Tordon 50-d' is restricted by its effect on clovers, soil persistence and cost. Variable growth of C. arvense, associated with variable summer rainfall in Victoria, increases the difficulty in controlling this weed and necessitates a persistent long term approach.  相似文献   

4.
Losses of picloram (4-aniino-3,5,6-trichloropicolinic acid as the triisopropanolamine salt) and 2,4-D (2,4-dichloropheno-xyacetic acid as the triisopropanolamine salt) from a Tordon 101-sprayed podsol in a Great Lakes forest clearcut were measured for one year following spraying. Less than 1.0 % of the picloram and much less than 0.1% of the 2,4-D applied was lost during seven runoff events. Even an event less than 24 h after spraying failed to release significant quantities of either herbicide. Pertes simultanées de piclorame et de 2,4- D à partir d'un podzol forestier, pendant les tempêtes de pluie Les pertes de piclorame (acide 4-amino-3,5,6-trichloropicoli-nique sous forme de sel de triisopropanolamine) et de 2,4-D (acide 2,4-dichlorophénoxyacétique sous forme de sel de triiosopropanolamine) à partir d'un podzol traité au Tordon 101 dans une coupe à bianc de la forêt des Grands Lacs, ont été mesurées cours de I'année qui a suivi le traitement. Dans sept cas oú il y a eu ruissellement, la perte de piclorame a été inférieure à 1% et celle de 2,4-D très inférieure à 0,1 %, et même dans un cas qui s'est produit mo ins de 24 heures aprés le traitement, it n'a pas pu être trouvé des quantités significatives de pertes pour aucun des deux herbicides. Die laterate Verlagerung von Picloram und 2,4-D in einem Waldpodsol während Regenfätlen Nach der Anwendung von Tordon 101 in einem Kahlschlag (Podsolboden) des Forstgebietes der Grossen Seen, wurde eIn Jahr lang das Verschwinden von Picloram (4-Amino-3,5,6-trichlorpicolinsSure als Triisopropanolaminsalz) und 2,4-D (2,4-Dichlorphenoxyessigsaure als Triisopropanoiaminsalz) gemessen. Weniger als 1% des Piclorams und wesentlich weniger als 0,1% der ausgebrachten 2,4-D-Menge, waren infolge von sieben Regenfällen und dem damit verbundenen abfliessenden Oberflächenwassers verloren gegangen. Sogar ein Regen der weniger als 24 Stunden nach der Spritzung fiel, vermochte keine nennenswerten Mengen von beiden Herbiziden aus dem Boden zu lösen.  相似文献   

5.
Summary. High concentrations of 2,4-D and picloram interfered with the downward movement of 14c-assimilates infield-grown vines. The interference in translocation was appreciably greater with picloram than it was with 2,4-D, Although basipetal translocation was retarded, translocation within the treated shoots continued from the vegetative part to the clusters. Translocation of 2,4-D appeared to follow the same route as 14c-assimilatcs for the most part. Formative effects were absent on untreated grape shoots although the adjacent shoots treated with 2,4-D or picloram on the same cordons were killed; however, formative effects were evident on some of the stump sprouts which developed after the vines were harvested. The malformed leaves on the stump sprouts were twelve or more nodes from the base of the shoots, while 14c was in the more basal leaves.
Thompson Seedless (Sultanina) rootings treated with 20 000 ppm 2,4-D or picloram transported less 14C to the roots than did the controls. Treatment with either herbicide resulted in a marked increase in the labelling of the stems.
Effet du 2,4-D et du pichlorame sur la migration de métabolites marqués au 14C dans Vitis vinifera L.  相似文献   

6.
R. GROVER 《Weed Research》1968,8(3):226-232
Summary. Effective dosages of picloram (4–amino-3,5,6–trichloropicolinic acid) required to reduce fresh weight of sunflowers ( Helianthus annuus L. var. Menonite) were determined for seven Saskatchewan soils under controlled environmental conditions. The relationships between ED50 value and clay content, organic matter content and cation exchange capacity were evaluated by correlation and regression analysis for possible usefulness in predicting dose requirements. ED50 values were also determined for Weyburn loam and in culture solutions adjusted to various pH levels.
There Was no significant correlation between ED50 values of picloram and soil clay content or cation exchange capacity. ED50 values were highly correlated with soil organic matter content, and they increased as the soil pH was lowered or raised from pH 6.5. The increase in ED50 values in the acidic range was attributed to adsorption of the unionized molecules of picloram on the organic matter in the soil. The increase in ED50 values in the alkaline range may be mainly due to reduced uptake of the ionized acid by plant roots.  相似文献   

7.
The response of Achillea millefolium L. to herbicides was measured to determine the effectiveness of the current recommendations and to test alternative herbicides. Five plots at each of the three replicate sites were selected and randomly treated with one of the four herbicides: dicamba/2,4‐D, glyphosate, metsulfuron‐methyl and triclopyr/picloram. After 12 months, the recommended treatment, dicamba/2,4‐D, did not cause a significant reduction in aerial biomass compared with untreated controls and the number of flowering stems was significantly increased in treated plots. Metsulfuron‐methyl and triclopyr/picloram caused a significant reduction in A. millefolium aerial biomass but did not affect the number of flowering stems. Glyphosate produced a significant reduction in aerial biomass and was less effective. These results suggest that the current recommendation for A. millefolium might be improved. Site‐specific effects such as aspect may also influence the effectiveness of herbicides on this species.  相似文献   

8.
A model system consisting of chemically isolated cuticular membranes placed on agar was used to study the penetration of various formulations of 14C-labelled clopyralid, fluroxypyr, triclopyr, picloram, and 2,4-D into and through cuticular membranes. Clopyralid, commercially formulated as the acid, or 1-decyl ester was rapidly absorbed after 12 h by isolated cuticles of Euonymus fortunei. There was less absorption of the monoethanolamine and potassium salt formulations when compared to the acid and 1-decyl ester. However, in terms of the absorbed 14C-activity that partitioned into the agar, there was no difference between the acid and salt formulations with approximately 90% being partitioned after 48 h. Conversely, the 1-decyl ester formulation of clopyralid was retained in the cuticle; less than 5% of the absorbed fraction was recovered in the agar after 48 h. When the acid forms of clopyralid, fluroxypyr, triclopyr, picloram, and 2,4-D were compared, there was little or no difference in absorption among the herbicides. However, the 14C-activity from clopyralid partitioned the most (90%) and triclopyr the least (50%) into the agar. When ester formulations of clopyralid, fluroxypyr, and triclopyr were compared, at least 95% of the 14C-activity was absorbed 24 h after application. However, of the amount absorbed, significantly more of the butoxyethyl ester of triclopyr (36%) partitioned into the agar than did the 1-decyl ester of clopyralid (6%) or the 1-methylheptyl ester of fluroxypyr (5%). Differential retention of various herbicide formulations in this model system may explain, in part, the differences in absorption and translocation among radiolabelled clopyralid formulations observed in previous research (Kloppenburg & Hall, 1990).  相似文献   

9.
Conyza bonariensis is a major weed infesting zero‐tilled cropping systems in subtropical Australia, particularly in wheat and winter fallows. Uncontrolled C. bonariensis survives to become a problem weed in the following crops or fallows. As no herbicide has been registered for C. bonariensis in wheat, the effectiveness of 11 herbicides, currently registered for other broad‐leaved weeds in wheat, was evaluated in two pot and two field experiments. As previous research showed that the age of C. bonariensis, and to a lesser extent, the soil moisture at spraying affected herbicide efficacy, these factors also were investigated. The efficacy of the majority of herbicide treatments was reduced when large rosettes (5–15 cm diameter) were treated, compared with small rosettes (<5 cm diameter). However, for the majority of herbicide treatments, the soil moisture did not affect the herbicide efficacy in the pot experiments. In the field, a delay in herbicide treatment of 2 weeks reduced the herbicide efficacy consistently across herbicide treatments, which was related to weed age but not to soil moisture differences. Across all the experiments, four herbicides controlled C. bonariensis in wheat consistently (83–100%): 2,4‐D; aminopyralid + fluroxypyr; picloram + MCPA + metsulfuron; and picloram + high rates of 2,4‐D. Thus, this problem weed can be effectively and consistently controlled in wheat, particularly when small rosettes are treated, and therefore C. bonariensis will have a less adverse impact on the following fallow or crop.  相似文献   

10.
The effects of picloram formulated with a nonionic surfactant (X-77) and of the surfactant alone on the ultrastructure of leaf cells of velvet mesquite (Prosopis velutina Woot.) and catclaw [Acacia greggii var. ariznica (Gray) Isely] were examined The surfactant induced temporary protrusions from chloroplasts in both species. A proliferation of rough endoplasmic reticulum (RER) was noted in velvet mesquite within 8h of application of the herbicide and in catclaw with in 27 h. By 72h after treatment. both species exhibited distortions of organelles with more severe symptoms in catclaw. the species more sensitive to the herbicide Leaf abscission occurred subsequently and was more pronounced in catclaw than in mesquitel It is known that RER proliferation is induced by ethylene and that ethylene evolution is stimulated by picloram. The present study suggests that the interaction between these two chemicals was similar in the two plant species studied.  相似文献   

11.
G. W. IVENS 《Weed Research》1971,11(2-3):150-158
Summary. In a series of foliar spray treatments on the deciduous Acacia drepanolobium Sjöstedt, a mixture of picloram+2,4-D lulled a large proportion of the trees in May, the latter part of the principal rainy season, moderate numbers in November-December, the subsidiary rains, and small numbers during the two intervening dry seasons. An ester formulation of 2,4,5-T was generally less effective, but similar seasonal variations in kill occurred. At the time of maximum susceptibility, the rate of shoot extension growth was declining and new leaves were fully extended, but still succulent.
With the evergreen Tarchonanthus camphoratus L., the effectiveness of spraying was related directly to the growth rate of the shoots. The percentage kill was highest following application of picloram+2,4-D in May-July, the late long rains and first half of the long dry season, but low during the November-December rains and short dry season. T. camphoratus is susceptible to foliar herbicide treatment over a longer period than A. drepanolobium , presumably because of its different seasonal growth characteristics.  相似文献   

12.
The liana, hiptage ( Hiptage benghalensis ), is currently invading the wet tropics of northern Queensland and remnant bushland in south-eastern Queensland, Australia. Trials using seven herbicides and three application methods (foliar, basal bark, and cut stump) were undertaken at a site in north Queensland (158 700 hiptage plants ha−1). The foliar-applied herbicides were only effective in controlling the hiptage seedlings. Of the foliar herbicides trialed, dicamba, fluroxypyr, and triclopyr/picloram controlled >75% of the treated seedlings. On the larger plants, the cut stump applications were more effective than the basal bark treatments. Kills of >95% were obtained when the plants were cut close to ground level (5 cm) and treated with herbicides that were mixed with diesel (fluroxypyr and triclopyr/picloram), with water (glyphosate), or were applied neat (picloram). The costings for the cut stump treatment of a hiptage infestation (85 000 plants ha−1), excluding labor, would be $A14 324 ha−1 using picloram and $A5294 ha−1 and $A2676 ha−1, respectively, using glyphosate and fluroxypyr. Foliar application using dicamba for seedling control would cost $A1830 ha−1. The costs range from 2–17 cents per plant depending on the treatment. A lack of hiptage seeds below the soil surface, a high germinability (>98%) of the viable seeds, a low viability (0%) of 2 year old, laboratory-stored fruit, and a seedling density of 0.1 seedlings m−2 12 months after a control program indicate that hiptage might have a short-term seed bank. Protracted recolonization from the seed bank would therefore be unlikely after established seed-producing plants have been controlled.  相似文献   

13.
In controlled environmental studies, a marked difference was observed between the growth pattern of tomato and eastern black nightshade plants that received doses of 2,4-D ranging from 28 to 952 g a.e. ha?1. The highest dose of 2,4-D reduced the dry weight of eastern black nightshade and tomato by approximately 15 and 50%, respectively, when compared with controls. Although the height of both species was reduced by all doses of 2,4-D, eastern black nightshade plants produced secondary shoots, which compensated for any potential loss in dry weight that otherwise may have occurred. Tomato plants did not produce secondary shoots. After application of 14C-2,4-D to tomato and eastern black nightshade, the pattern of 14C absorption and translocation was similar in both plant species. However, there was significantly more radioactivity recovered in tomato (72%) than in eastern black nightshade (52%) plants, 72 h after treatment. Assay radioactivity in the nutrient solution of hydroponically grown plants indicated that 7·0 and 27·9% of the applied radioactivity was exuded from the roots of tomato and eastern black nightshade, respectively, within 72 h after treatment. Assay of plant extracts by thin layer chromatography revealed that the amount of radioactivity that remained as unaltered 2,4-D was 73 and 49% in tomato and eastern black nightshade, respectively, 72 h after treatment. Thus the greater tolerance of eastern black nightshade appeared to be due to greater rates of 2,4-D metabolism and/or greater rates of herbicide elimination by root exudation.  相似文献   

14.
Diaeretiella rapae (M’Intosh.) is an endoparasite of the turnip aphid,Lipaphis erysimi (Kalt.) with a wide geographical distribution. The four Pesticides Nogos 50 EC at 600 ml per acre, Dimecron 50 WSC, Monofos 40 WSC and Tamaron 600 SL at 500 ml per acre, used in practice for the control of this pest, were tested for their side-effects onD. rapae. The parasitoid was reared on pottedBrassica napus plants infested with aphid under laboratory conditions (22±2°C, 60–70% rel. hum., 16 h light and 8 h dark). In one test, adult female parasitoids were exposed to fresh pesticide residues on glass plates and in another test, the pupae within aphid mummies were directly sprayed. The results revealed that Dimecron 50 WSC, Nogos 50 EC and Monofos 40 WSC were harmful causing 100% mortalitv toD. rapae followed by Tamaron 600 SL (97% moderately harmful) after 24 hours of application, compared to no mortality in control, where only water was sprayed. Directly spraying of pupae with Dimecron 50 WSC and Nogos 50 EC reduced adult parasitoid emergence to 9 and 7%, respectively, Monofos 40 WSC and Tamaron 600 SL to 3% compared to 78% emergence for the control within one week of treatment. The results showed that none of the tested pesticides was safe toD. rapae and according to the International Organization for Biological Control (IOBC) further testing under semi-field and field conditions is recommended.  相似文献   

15.
Influence of picloram on Cirsium arvense (L.) Scop, control with glyphosate   总被引:1,自引:0,他引:1  
Low rates of picloram in mixture with glyphosate provided a rapid enhancement of the onset of injury to the shoots of Cirsium arvense (Canada thistle or creeping thistle) under field (0.07+1.0 and 0.07+1.5 kg ha?1) and greenhouse (0.035+0.42 and 0.07+0.84 kg ha?1) conditions. Picloram slightly reduced the amount of 14C-glyphosate absorbed at 24 and 48 but not 72 h after treatment. Movement of 14C-glyphosate from the treated leaves to the shoot apex, remainder of the shoot and roots was reduced in the presence of picloram. Necrosis of the treated leaves above the treated spots was evident, presumably indicating acropetal movement of either or both herbicides. With the picloram + glyphosate mixtures there was increased shoot regrowth over glyphosate alone at 1 year after treatment under field, and with certain mixtures at 18 days and 4 weeks after treatment under greenhouse conditions. Following application of the mixtures, accumulation of glyphosate in the shoots may be responsible for the enhanced onset of shoot injury while failure of enough glyphosate to translocate to, and cause death of, the roots may be responsible for the increased shoot regrowth over glyphosate alone.  相似文献   

16.
A method is described for the analysis of mixtures of picloram with 2,4-D or 2,4,5-T in commercial formulations. The method involves the esterification of the acids in sealed tubes at 105°C using boron trifiuoride-methanol reagent, followed by extraction, and then estimation by gas chromatography using flame ionisation detection.  相似文献   

17.
Effect of low temperatures on 2,4-D behaviour in maize plants   总被引:3,自引:0,他引:3  
The foliar surface of 4-leaf maize plants was found to be poorly wettable and retained 106 μl g?1 dry matter when sprayed with a U46D (2,4-D formulation) blank. The third leaf retained 141 μl g?1. A 7-day cold spell (17/9°C) increased retention per unit dry matter by 53% (135% on the third leaf). Cold stress lowered epicuticular wax quantity by 29% on the third leaf. Contact angles of formulated 2,4-D lay between 115 and 125° and were not significantly affected by cold stress. 2,4-D rapidly entered into maize third leaf (66% in 24 h) but migration from it was less than 1.5%. 2,4-D was readily degraded in maize (80% in 72 h). The most abundant metabolite was probably an ester conjugate; little of the hydroxy derivatives were found. Cold stress reduced 2,4-D degradation, and 72 h after treatment the amount of undegraded 2,4-D was 78% higher in cold-stressed maize plants. It was concluded that 2,4-D selectivity in maize results from low spray retention per unit dry matter and active degradation of penetrated herbicide. Cold stress affects both factors.  相似文献   

18.
T. J. MUZIK 《Weed Research》1965,5(3):207-212
Summary. One leaf of each of several tomato plants 5–6 weeks old maintained at 55, 70 or 85° F was dipped in 0.05 M amitrole or 0.005 M 2,4-D solution. The treated leaf was removed 1 week later. Scions taken from untreated plants were grafted to the treated plants at intervals up to 4 months after treatment.
Amitrole-treated plants maintained at 55° F were killed within 1 month but those at 70 and 85° F recovered.
Plants treated with 2,4-D showed typical symptoms when maintained at 70 and 85° F but no symptoms at 55° F. Scions grafted on plants maintained at 55, 70 and 85° F for 60 days after 2,4-D treatments exhibited typical symptoms. Grafts made more than 60 days following the 2,4-D application did not exhibit symptoms except a slight malformation on the plants held at 55° F.
New growth on the scions grafted up to 103 days after amitrole treatment on plants which had been maintained at 70 and 85° F exhibited typical amitrole symptoms (i.e. white foliage) but scions grafted after this lime did not. New shoots developing from buds on the treated plants also produced green growth 115 days after treatment but produced white foliage prior to that time.
Effet de la température sur l'activi'ité et la persistance de l'aminotriazole et du 2,4-D  相似文献   

19.
Summary. The growth has been tested in vitro of selected species of fungi, yeasts, bacteria, actinomycetes and green algae in the presence of the acetic and α-butyric acid forms of MCPA, 2,4–D and 2,4,5–T. The results show that at concentrations approximating to field rates of application, these chemicals were harmless. At much higher concentrations (above 500 ppm), however, the α-butyric acids were found to be highly toxic whereas the acetic acids has no effect on growth.  相似文献   

20.
Summary. Picloram is decarboxylated to a negligible extent in roots of skeleton weed: applied to leaves it kills roots better than 2,4-D; it penetrates leaves better than 2,4-D and disrupts the translocation system more, does not leak considerably into the soil from roots by the time of appropriate harvests, yet is not as well translocated as 2,4-D. Under such conditions its ability to kill roots to a greater depth is attributable to its greater intrinsic toxicity.
There is some evidence that under certain field conditions useful movement or uptake of picloram may occur via the soil.
Recherches sur Chondrilla juncea L . avee le plclorame et le 2,4-D.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号