Academic literature on the topic 'Diflufenican'

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the lists of relevant articles, books, theses, conference reports, and other scholarly sources on the topic 'Diflufenican.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Journal articles on the topic "Diflufenican"

1

Mawunya, M., I. K. Dzomeku, I. Baba, and M. Abudulai. "Evaluation of the Efficacy of Some Pre-Emergence Herbicides in Cotton (Gossypium Spp.) in Northern Ghana." International Journal of Irrigation and Agricultural Development (IJIRAD) 1, no. 1 (January 24, 2018): 82–90. http://dx.doi.org/10.47762/2017.964x.25.

Full text
Abstract:
A field experiment was conducted at the Savanna Agricultural Research Institute to investigate the effects of pre-emergence herbicides for weed control in cotton during the 2014 and 2015 cropping seasons. The study determined the effects of different rates of three novel cotton herbicides with different formulations on weeds, yield components and yield of seed cotton, applied as pre-emergence. The treatments were laid out in randomized complete block design with four replications and consisted of Diflufenican + Flufenicent and Diflufenican + Flufenicent + Flurtamone as 500SC formulations of each and applied at the rate of 0.3, 0.4, 0.6, and 0.8 l/ha, whilst Diflufenican + Flufenicent + Flurtamone as 200SC formulation was applied at 0.4, 0.5, 0.8, and 1.0L/ha. A reference herbicide + product with active ingredients Promotrin + Metolachlor applied at 3.0l /ha and untreated weedy and weed free checks were included. Results showed that. Diflufenican + Flufenicent + as 200SC applied at 0.8 l/ha gave the lowest weed dry weight (65.0 g/m2) similar to the reference herbicide (189.4 g/m2) and Diflufenican + Flufenicent + Flurtamone as 200SC applied at 0.8 l/ha (144.8 g/m2), whilst weedy check gave highest (548.7 kg/m2). Consequently, Diflufenican+ Flufenicent applied at 0.8 l/ha gave significantly (p<0.05) the highest number of bolls and number of opened bolls than the reference herbicide; but similar to weed free. Diflufenican + Flufenicent +Flurtamone as 500SC, applied at 1.0 l/ha supported tallest plants similar to Diflufenican + Flufenicent + Flurtamone as 200SC at 0.4 and 0.6 l/ha and Diflufenican + Flufenicent as 500SC applied at 0.6 and 0.8 l/ha and better than the reference herbicide. Diflufenican+ Flufenicent as 500SC applied at 0.8L/ha (81.3 %), Diflufenican + Flufenicent + Flurtamone as 500SC at 0.8 l/ha (78.7 %) and the reference herbicide (76.3 %) did better (90 %) than weed free. Generally, early control of weeds with the tested herbicides using 0.4 to 1.0 l/ha minimized weed growth and supported high seed cotton yields.
APA, Harvard, Vancouver, ISO, and other styles
2

Madrigal Monárrez, Ismael, Pierre Benoit, Enrique Barriuso, Benoît Réal, Alain Dutertre, Michel Moquet, Maria Trejo Hernández, and Laura Ortíz Hernández. "Pesticide sorption and desorption from soils having different land use." Ingeniería e Investigación 28, no. 3 (September 1, 2008): 96–104. http://dx.doi.org/10.15446/ing.investig.v28n3.15127.

Full text
Abstract:
This study was carried out within the framework of a multidisciplinary project for evaluating buffer zones for combating pesticide contamination of surface water. Such areas are effective in removing pesticides transported by run-off; however, little information is available about the fate of the pesticides so intercepted. Two herbicides having contrasting properties (isoproturon, moderately hydrophobic (log Kow = 2.5), diflufenican, strongly hydrophobic (log Kow = 4.9)) and isopropylaniline (an isoproturon metabolite) were used for characterising sorption and desorption from soil having three different land uses: grass buffer strip, woodland and cultivated plot. The experiments were carried out in controlled laboratory conditions using isoproturon labelled with 14C in the benzene ring. The results demonstrated that diflufenican and isopropilaniline retention was more significant than isoproturon in three soils. The three molecules’ Kd values revealed that isoproturon and diflufenicanil retention was more important in woodland soil where carbon content was more significant (ZB 0-2: Kd IPU = 15.1 Ls kg-1; Kd DFF = 169.2 Ls kg-1). Isopropilanilina Kd was higher in grass buffer strip soil (BE 0-2: Kd IPA = 53.1 L kg-1). These differences were related to different organic matter content and nature according to the type of land use.
APA, Harvard, Vancouver, ISO, and other styles
3

Stankiewicz-Kosyl, Marta, Mariola Wrochna, Maria Salas, and Stanislaw Waldemar Gawronski. "A strategy of chemical control of Apera spica-venti L. resistant to sulfonylureas traced on the molecular level." Journal of Plant Protection Research 57, no. 2 (June 1, 2017): 113–19. http://dx.doi.org/10.1515/jppr-2017-0015.

Full text
Abstract:
Abstract Three populations of silky bent grass (Apera spica-venti L.) were tested – one that is susceptible and two that are resistant to sulfonylureas. This study assessed the efficacy of control by different herbicides in a pot experiment and estimated the molecular status of resistance to sulfonylureas in analysed populations and its effect on the efficacy of different chemical treatments. The three most effective herbicide rotation schemes were: 1) chlorsulfuron + isoproturon, ethametsulfuron + metazachlor + quinmerac, chlorsulfuron + isoproturon; 2) prosulfocarb + diflufenican, ethametsulfuron + quizalofop-p-ethyl, prosulfocarb + diflufenican; 3) diflufenican + flufenacet, quizalofop-p-ethyl, diflufenican + flufenacet. In most cases it was more difficult to destroy 100% of the resistant population from Modgarby where the majority of plants had no mutation in the als gene. In the resistant population from Babin there were significantly more individuals with mutation in the als gene, therefore exhibiting target-site resistance.
APA, Harvard, Vancouver, ISO, and other styles
4

Książek-Trela, Paulina, Ewelina Bielak, Dominika Węzka, and Ewa Szpyrka. "Effect of Three Commercial Formulations Containing Effective Microorganisms (EM) on Diflufenican and Flurochloridone Degradation in Soil." Molecules 27, no. 14 (July 16, 2022): 4541. http://dx.doi.org/10.3390/molecules27144541.

Full text
Abstract:
The aim of this study was to determine the influence of effective microorganisms (EM) present in biological formulations improving soil quality on degradation of two herbicides, diflufenican and flurochloridone. Three commercially available formulations containing EM were used: a formulation containing Bifidobacterium, Lactobacillus, Lactococcus, Streptococcus, Bacillus, and Rhodopseudomonas bacteria and the yeast Saccharomyces cerevisiae; a formulation containing Streptomyces, Pseudomonas, Bacillus, Rhodococcus, Cellulomonas, Arthrobacter, Paenibacillusa, and Pseudonocardia bacteria; and a formulation containing eight strains of Bacillus bacteria, B. megaterium, B. amyloliquefaciens, B. pumilus, B. licheniformis, B. coagulans, B. laterosporus, B. mucilaginosus, and B. polymyxa. It was demonstrated that those formulations influenced degradation of herbicides. All studied formulations containing EM reduced the diflufenican degradation level, from 35.5% to 38%, due to an increased acidity of the soil environment and increased durability of that substance at lower pH levels. In the case of flurochloridone, all studied EM formulations increased degradation of that active substance by 19.3% to 31.2% at the most. For control samples, equations describing kinetics of diflufenican and flurochloridone elimination were plotted, and a time of the half-life of these substances in laboratory conditions was calculated, amounting to 25.7 for diflufenican and 22.4 for flurochloridone.
APA, Harvard, Vancouver, ISO, and other styles
5

Tejada, Manuel. "Evolution of soil biological properties after addition of glyphosate, diflufenican and glyphosate+diflufenican herbicides." Chemosphere 76, no. 3 (July 2009): 365–73. http://dx.doi.org/10.1016/j.chemosphere.2009.03.040.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Gao, Chao, Hongchen Li, Miaochang Liu, Jinchang Ding, Xiaobo Huang, Huayue Wu, Wenxia Gao, and Ge Wu. "Regioselective C–H chlorination: towards the sequential difunctionalization of phenol derivatives and late-stage chlorination of bioactive compounds." RSC Adv. 7, no. 74 (2017): 46636–43. http://dx.doi.org/10.1039/c7ra09939h.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Ueyama, Y., Y. Hashimoto, R. Hasegawa, Y. Horita, and N. Matsuda. "Herbicidal efficacy of diflufenican/trifluralin granule." Journal of Weed Science and Technology 45, Supplement (2000): 224–25. http://dx.doi.org/10.3719/weed.45.supplement_224.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Seefeldt, S., E. Peters, M. L. Armstrong, and A. Rahman. "Crossresistance in chlorsulfuronresistant chickweed (Stellaria media)." New Zealand Plant Protection 54 (August 1, 2001): 157–61. http://dx.doi.org/10.30843/nzpp.2001.54.3714.

Full text
Abstract:
In New Zealand chlorsulfuronresistant chickweed was discovered in 1995 Experiments were conducted in the field and glasshouse to determine whether these resistant plants were also cross and multipleresistant to other herbicides normally effective on chickweed A population of chlorsulfuronresistant chickweed growing in an oat crop near Winton in the South Island was treated with 13 different herbicides This population was not controlled by mecoprop methabenzthiazuron and pendimethalin and only partially controlled by bromoxynil ioxynil diflufenican isoproturon and diflufenican bromoxynil in the field trial Two followup glasshouse experiments using plants grown from seeds harvested from the surviving plants confirmed crossresistance to thifensulfuron and susceptibility to all the other herbicides including tribenuron
APA, Harvard, Vancouver, ISO, and other styles
9

Henderson, CWL, and MJ Webber. "Phytotoxicity of several pre-emergence and post-emergence herbicides to green beans (Phaseolus vulgaris)." Australian Journal of Experimental Agriculture 33, no. 5 (1993): 645. http://dx.doi.org/10.1071/ea9930645.

Full text
Abstract:
The phytotoxicity to green beans (Phaseolus vulgaris) of the herbicides metolachlor, pendimethalin, cyanazine, acifluorfen, diflufenican, bentazone, metribuzin, prometryn, terbutryn, methabenzthiazuron, and oxyfluorfen was investigated in 4 experiments on a black earth soil (clay content 40-60%) at Gatton Research Station in southern Queensland during 1990-91. Metolachlor was applied post-sowing and pre-emergence; up to 4 kg a.i./ha did not significantly (P>0.05) affect growth or yields, indicating a considerable safety margin for this herbicide when used at commercial rates. Pendimethalin did not cause significant crop damage when applied in the same manner at rates up to 2.7 kg a.i./ha. Acifluorfen and diflufenican were each applied at 3 or 4 weeks after sowing in 3 experiments. Sensitivity of the bean crop to acifluorfen varied: 0.5 kg a.i./ha did not significantly reduce bean growth or yield in 2 experiments, but 0.11 kg a.i./ha reduced yields by 20% in a third experiment. Application of 0.1-0.12 kg a.i./ha of diflufenican had no adverse effect on beans in 2 experiments, although significant damage was observed in an initial screening experiment. Bentazone applied 3 weeks after sowing had no significant effect on bean yield or growth in 1 experiment; in another, the maximum label rate of 0.96 kg a.i./ha significantly reduced bean growth and yield. Post-emergence application of cyanazine, metribuzin, prometryn, terbutryn, methabenzthiazuron, or oxyfluorfen at rates required for acceptable weed control either killed the bean plants within a few days or resulted in complete yield loss. Levels of damage from these herbicides preclude their use in green beans. Although green beans showed some tolerance to postsowing, pre-emergence application of cyanazine, low rates of 0.75-1 kg a.i./ha reduced yields by 35%. Both metolachlor and pendimethalin appear suitable for pre-emergence use in green beans. Further work on factors affecting phytotoxicity of acifluorfen, diflufenican, and bentazone to green beans is required.
APA, Harvard, Vancouver, ISO, and other styles
10

Dear, B. S., and G. A. Sandral. "The phytotoxicity of the herbicides bromoxynil, pyridate, imazethapyr and a bromoxynil + diflufenican mixture on subterranean clover and lucerne seedlings." Australian Journal of Experimental Agriculture 39, no. 7 (1999): 839. http://dx.doi.org/10.1071/ea98182.

Full text
Abstract:
Summary. The effect of the herbicides pyridate, imazethapyr and a bromoxynil + diflufenican mixture on subterranean clover (Trifolium subterraneum L.) (cvv. Trikkala and Karridale) and lucerne (Medicago sativa L.) (cv. Aurora) seedlings was examined in randomised plot field experiments in 2 successive years. Responses were compared against an unsprayed control and a standard bromoxynil application. The herbicides and the rates of product applied were: bromoxynil + diflufenican (0.5, 1.0 L/ha), imazethapyr (0.18, 0.3 L/ha), pyridate (1.0, 3.0 L/ha), and bromoxynil (1.4 L/ha). Weeds were removed by hand from the subterranean clover experiments but not the lucerne experiments. Pyridate and imazethapyr were the least phytotoxic of the herbicides applied on both subterranean clover and lucerne. The bromoxynil + diflufenican mixture was the most phytotoxic, causing severe leaf burn and a depression in herbage biomass in both species. Despite the high level of phytotoxicity by some treatments, none of the herbicides reduced lucerne seedling numbers. Lucerne herbage yields in late spring were higher in most sprayed plots compared with the unsprayed control due to the removal of weed competition. Seed yield responses in subterranean clover due to herbicide application ranged from negative responses up to –21% with pyridate to positive responses up to 92% with the bromoxynil + diflufenican treatment relative to the weed-free, unsprayed control. The positive responses were attributed to increased availability of soil water during seed set in treatments in which herbicides suppressed legume biomass. There was a good correlation in both 1992 (R2 = 0.85–0.89) and 1993 (R2 = 0.63–0.73) between the depression in herbage yield in spring and the increase in seed set relative to the control. Soil water under the subterranean clover cultivar Karridale in spring was highest in the bromoxynil and imazethapyr treatments, which produced a large reduction in biomass, and lowest in the control and pyridate treatments, which had showed the least depression in biomass 60 days after treatment application. Although some herbicides cause a high level of phytotoxicity, their use in weedy subterranean clover–lucerne mixtures is justified in view of the small negative, and potentially large positive, effects on subterranean clover seed yield and the increased lucerne yields later in the season due to weed suppression.
APA, Harvard, Vancouver, ISO, and other styles

Dissertations / Theses on the topic "Diflufenican"

1

Ahmad, Ahmad M. S. "Studies on the biological activity of the herbicide diflufenican." Thesis, Bangor University, 1991. http://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.293902.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Farida, Nihla. "Interaction between atrazine residues in the soil and diflufenican application in peas and lucerne /." Title page, contents and abstract only, 1996. http://web4.library.adelaide.edu.au/theses/09ACM/09acmf244.pdf.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Wightman, Patricia S. "Studies on the mode of action of diflufenican in wheat, barley and selected weed species." Thesis, University of Strathclyde, 1988. http://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.303313.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Knight, Patricia Heather Radcliffe. "Factors affecting the uptake and activity of foliage-applied diflufenican in a selected crop and weed species." Thesis, University of Strathclyde, 1989. http://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.291992.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Haynes, C. A. "The basis of selectivity of Diflufenican in wheat, barley and selected weed species and its fate in the soil environment." Thesis, University of Strathclyde, 1987. http://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.233610.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Madrigal, Monarrez Ismael. "Rétention de pesticides dans les sols des dispositifs tampon, enherbés et boisés : rôle des matières organiques." Paris, Institut national d'agronomie de Paris Grignon, 2004. https://pastel.archives-ouvertes.fr/pastel-00001214.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

PATTY, LAURENT. "Limitation du transfert par ruissellement vers les eaux superficielles de deux herbicides (isoproturon et diflufenicanil) : méthodologie analytique et étude de d'efficacité de bandes enherbées." Université Joseph Fourier (Grenoble), 1997. http://www.theses.fr/1997GRE10057.

Full text
Abstract:
Les travaux presentes dans ce memoire se rapportent a l'etude de la capacite d'un dispositif enherbe experimental, a limiter le transfert par ruissellement de deux herbicides : l'isoproturon (ipu) et le diflufenicanil (dff), qui presentent des proprietes physico-chimiques differentes. Le dispositif, implante a la station i. T. C. F. De la jailliere, a ete suivi au cours de quatre campagnes d'etude en conditions de pluies naturelles, et face a deux ruissellements de plus grande amplitude generes par simulation de pluie. Apres avoir developpe et valide des methodes de dosage sensibles et specifiques de chaque produit et de chaque matrice etudiee, l'analyse des ruissellements collectes a permis de caracteriser le transfert par ruissellement de l'ipu et du dff et d'evaluer l'efficacite des bandes enherbees a reduire ce tranfert. Les resultats provenant des dispositifs implantes dans les stations i. T. C. F. De bignan, plelo et la jailliere, ont confirme et elargi les conclusions a d'autres produits : le lindane, l'atrazine et deux de ses metabolites. Dans les diverses conditions pedoclimatiques, les quantites de produits transferees par ruissellement sont inferieures a 1 % de la dose epandue ; les episodes de ruissellement les plus proches de l'application etant a l'origine des pertes les plus importantes. Les resultats relatifs a l'efficacite des bandes enherbees montrent que differents mecanismes contribuent a la retention de ces residus : l'infiltration de l'eau et des molecules solubilisees (ipu, atrazine) dans le sol, la filtration et la sedimentation des particules solides sur lesquelles sont adsorbes des composes tres peu solubles (dff, lindane) et l'adsorption des molecules solubilisees en surface de la bande. Les dispositifs etudies permettent notamment d'ecreter les pics de concentration des differents composes dans le ruissellement. Les analyses de sol revelent, par ailleurs, que les residus d'ipu et de dff pieges ne s'accumulent pas dans la bande enherbee.
APA, Harvard, Vancouver, ISO, and other styles
8

Dang, Hue Thi. "Investigation of herbicide resistance in oriental mustard (Sisymbrium orientale L.) in Australia." Thesis, 2018. http://hdl.handle.net/2440/118239.

Full text
Abstract:
Oriental mustard (Sisymbrium orientale L.), called Indian hedge mustard in Australia, is an important broadleaf weed of southern Australia. It has become more difficult to control in field crops due to the evolution of herbicide resistance. This study investigated the extent of resistance to four different herbicide modes of action, used to control oriental mustard in Australia. Herbicide resistance status was determined in 75 populations collected in southern Australia from 2010 to 2016 with resistance confirmed to herbicides inhibiting acetolactate synthase, photosystem II, phytoene desaturase (PDS) and auxinic herbicides. Populations resistant to PS-II, PDS-inhibitors and auxinic herbicides and two known susceptible populations (S1 and S2) were used to investigate the level of resistance, its mechanism, inheritance and fitness cost associated with resistance. Populations P17 and P18 were 311 and 315-fold, respectively, more resistant to atrazine than the susceptible populations as determined by the comparisons of their LD50 values. However, there was no resistance detected in these populations to diuron. Sequencing of the chloroplastic psbA gene identified a missense mutation of serine 264 to glycine in both herbicide-resistant populations, known to confer high-level of atrazine resistance in other species. P2 and P13 populations were 81 and 67-fold more resistant to 2,4-D at the LD50 level compared to the susceptible populations, respectively. No predicted amino acid modification was detected in sequences of potential target-site genes [Auxin binding protein (ABP), Transport inhibitor response 1 (TIR 1) and Auxin F-box protein 5 (AFB5)]. Further studies showed resistant populations had reduced 2,4-D translocation compared to the susceptible populations. At 72 h after herbicide treatment, 77% of [14C]2,4-D was retained in the treated leaf in the resistant population compared to 32% of [14C]2,4-D retention in the susceptible populations. Studies on inheritance of resistance to PDS-inhibitors confirmed that resistance to diflufenican in P3 population is inherited as a single dominant gene trait. Likewise, resistance to diflufenican and picolinafen in population P40 is also due to a single dominant gene. Resistance to 2,4-D in populations P2 and P13 is inherited as a single partially dominant gene. Populations P3 and P40 were 140 and 237-fold more resistant to the PDS inhibitor diflufenican, respectively, than the susceptible populations. Both populations contained a Leu498-Val substitution in the PDS gene. An additional mutation, Glu-425-Asp, was only detected in P40, where cross-resistance to picolinafen was identified. These results suggest that Leu498 mutation alone can confer a high level of resistance to diflufenican; however, the presence of both Leu498 and Glu425 mutations increased the level of resistance to diflufenican and also conferred resistance to picolinafen. Fitness studies conducted under competition with wheat in the absence of herbicides in pots revealed that the mutant PDS genes in populations P3 and P40 did not impose any fitness costs. This means once a resistant trait occurs in the field, it will persist in the absence of herbicides.
Thesis (Ph.D.) -- University of Adelaide, School of Agriculture, Food and Wine, 2018
APA, Harvard, Vancouver, ISO, and other styles

Book chapters on the topic "Diflufenican"

1

Unger, Thomas A. "Diflufenican." In Pesticide Synthesis Handbook, 20. Elsevier, 1996. http://dx.doi.org/10.1016/b978-081551401-5.50017-4.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Tejada, Manuel, Marina Del, Manuel Tejada, Marina Del, Isidoro Gomez, and Juan Parrado. "Application of Diflufenican Herbicide on Soils Amended with Different Organic Wastes." In Herbicides and Environment. InTech, 2011. http://dx.doi.org/10.5772/13006.

Full text
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography