Academic literature on the topic 'Triazine herbicide'

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 'Triazine herbicide.'

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 "Triazine herbicide"

1

Parker, Ethan T., Micheal D. K. Owen, Mark L. Bernards, William S. Curran, Lawrence E. Steckel, and Thomas C. Mueller. "A Comparison of Symmetrical and Asymmetrical Triazine Herbicides for Enhanced Degradation in Three Midwestern Soils." Weed Science 66, no. 5 (September 2018): 673–79. http://dx.doi.org/10.1017/wsc.2018.41.

Full text
Abstract:
AbstractThe triazines are one of the most widely used herbicide classes ever developed and are critical for managing weed populations that have developed herbicide resistance. These herbicides are traditionally valued for their residual weed control in more than 50 crops. Scientific literature suggests that atrazine, and perhaps others-triazines, may no longer remain persistent in soils due to enhanced microbial degradation. Experiments examined the rate of degradation of atrazine and two other triazine herbicides, simazine and metribuzin, in both atrazine-adapted and non-history Corn Belt soils, with similar soils being used from each state as a comparison of potential triazine degradation. In three soils with no history of atrazine use, thet1/2of atrazine was at least four times greater than in three soils with a history of atrazine use. Simazine degradation in the same three sets of soils was 2.4 to 15 times more rapid in history soils than non-history soils. Metribuzin in history soils degraded at 0.6, 0.9, and 1.9 times the rate seen in the same three non-history soils. These results indicate enhanced degradation of the symmetrical triazine simazine, but not of the asymmetrical triazine metribuzin.
APA, Harvard, Vancouver, ISO, and other styles
2

Perry, D. H., J. S. McElroy, F. Dane, E. van Santen, and R. H. Walker. "Triazine-Resistant Annual Bluegrass (Poa annua) Populations with Ser264Mutation Are Resistant to Amicarbazone." Weed Science 60, no. 3 (September 2012): 355–59. http://dx.doi.org/10.1614/ws-d-11-00200.1.

Full text
Abstract:
Amicarbazone is a photosystem II (PSII)-inhibiting herbicide in the triazolinone family, which is similar in mode of action to the triazines. Annual bluegrass is a cool-season weed and has shown resistance to some PSII-inhibiting herbicides. The objective was to evaluate triazine-resistant and -susceptible annual bluegrass populations for potential cross-resistance to amicarbazone. Two triazine-resistant (MS-01, MS-02) and triazine-susceptible (AL-01, COM-01) annual bluegrass populations were treated with amicarbazone, atrazine, and simazine at 0.26, 1.7, and 1.7 kg ai ha−1, respectively. All herbicide treatments controlled the susceptible populations greater than 94% 2 wk after treatment (WAT). No visual injury of MS-01 and MS-02 was observed at any time following herbicide treatment. Quantum yield (ΦPSII) of annual bluegrass was measured 0 to 72 h after application (HAA) to determine the photochemical effects of amicarbazone compared to other PSII inhibitors. ΦPSIIof triazine-susceptible populations was reduced at all measurement times by all three herbicides. However, amicarbazone decreased ΦPSIIof susceptible populations faster and greater than atrazine and simazine at most measurement times. Amicarbazone did not reduce ΦPSIIof the MS-01 population. Amicarbazone significantly reduced ΦPSIIof the MS-02 population during several measurement timings; however, these reductions were short-lived compared to the susceptible populations and no trend in ΦPSIIreduction was observed. Sequencing of thepsbAgene revealed a Ser to Gly substitution at amino acid position 264 known to confer resistance to triazine herbicides. These data indicate amicarbazone efficiently inhibited PSII of susceptible annual bluegrass populations; however, triazine-resistant annual bluegrass populations with Ser264to Gly mutations are cross-resistant to amicarbazone.
APA, Harvard, Vancouver, ISO, and other styles
3

Zainal Abidin, Nurul Auni, Nur Sofiah Abu Kassim, and Noor Hidayah Pungot. "Solid Phase Extraction Method for the Determination of Atrazine and Cyanazine in Water Samples." ASM Science Journal 14 (April 2, 2021): 1–6. http://dx.doi.org/10.32802/asmscj.2020.631.

Full text
Abstract:
Triazine is one of the herbicides group that is widely used in agriculture that acts as an inhibitor for the growth of unwanted weeds in plants. The use of this herbicide on plants is absorbed by the soil and flows into a nearby water system. This research focused on two types of triazines, namely atrazine and cyanazine. This research aims to extract this type of triazine herbicides and to determine their concentration in water samples. It was quantified by using gas chromatography-electron capture detector (GC-ECD). Solid phase extraction (SPE) method was applied to extract herbicides from water samples. The results indicate that all the samples contained atrazine and cyanazine. Studies in the range of 0.5 - 25 mg/L achieved good linearity with good correlation of determination, r2 value of 0.9922 - 0.9982 mg/L. Relative standard deviations (RSD) for triplicate analysis of the samples were less than 10.0%. The limit of detection (LODs) and limit of quantification (LOQs) of cyanazine and atrazine were found, ranging from 3.33 – 6.67 μg/L and 11.09 – 20.10 μg/L, respectively. The recoveries of the triazine herbicides studied in water samples ranged from 82.5% to 107.6%. The developed method exhibited excellent clean-up capability and was successfully applied for determining triazine herbicide residues in water samples.
APA, Harvard, Vancouver, ISO, and other styles
4

GAYNOR, J. D., J. A. STONE, and T. J. VYN. "TILLAGE SYSTEMS AND ATRAZINE AND ALACHLOR RESIDUES ON A POORLY DRAINED SOIL." Canadian Journal of Soil Science 67, no. 4 (November 1, 1987): 959–63. http://dx.doi.org/10.4141/cjss87-091.

Full text
Abstract:
Seasonal residues of an acetanilide and triazine herbicide were monitored in ridge, conventional, and zero tillage systems. Alachlor (2-chloro-2′,6′-diethyl-N-(methoxymethyl)acetanilide), and atrazine (2-chloro-4-(ethylamino)-6-(isopropylamino)-s-triazine) residues were less than 8% of the spring application concentration at the end of the growing season. Moldboard plowing in the fall reduced herbicide residues in spring because of soil dilution by plowing to greater than the sampling depth. Ridge tillage systems had higher spring residues apparently because of reduced herbicide dissipation on the drier ridge tops. The higher residues of the triazines on ridge tops may be injurious to triazine sensitive crops. Key words: Herbicide, till-plant, ridge tillage, des-ethyl atrazine
APA, Harvard, Vancouver, ISO, and other styles
5

Shaner, Dale L. "Lessons Learned From the History of Herbicide Resistance." Weed Science 62, no. 2 (June 2014): 427–31. http://dx.doi.org/10.1614/ws-d-13-00109.1.

Full text
Abstract:
The selection of herbicide-resistant weed populations began with the introduction of synthetic herbicides in the late 1940s. For the first 20 years after introduction, there were limited reported cases of herbicide-resistant weeds. This changed in 1968 with the discovery of triazine-resistant common groundsel. Over the next 15 yr, the cases of herbicide-resistant weeds increased, primarily to triazine herbicides. Although triazine resistance was widespread, the resistant biotypes were highly unfit and were easily controlled with specific alternative herbicides. Weed scientists presumed that this would be the case for future herbicide-resistant cases and thus there was not much concern, although the companies affected by triazine resistance were somewhat active in trying to detect and manage resistance. It was not until the late 1980s with the discovery of resistance to Acetyl Co-A carboxylase (ACCase) and acetolactate synthase (ALS) inhibitors that herbicide resistance attracted much more attention, particularly from industry. The rapid evolution of resistance to these classes of herbicides affected many companies, who responded by first establishing working groups to address resistance to specific classes of herbicides, and then by formation of the Herbicide Resistance Action Committee (HRAC). The goal of these groups, in cooperation with academia and governmental agencies, was to act as a forum for the exchange of information on herbicide-resistance selection and to develop guidelines for managing resistance. Despite these efforts, herbicide resistance continued to increase. The introduction of glyphosate-resistant crops in the 1995 provided a brief respite from herbicide resistance, and farmers rapidly adopted this relatively simple and reliable weed management system based on glyphosate. There were many warnings from academia and some companies that the glyphosate-resistant crop system was not sustainable, but this advice was not heeded. The selection of glyphosate resistant weeds dramatically changed weed management and renewed emphasis on herbicide resistance management. To date, the lesson learned from our experience with herbicide resistance is that no herbicide is invulnerable to selecting for resistant biotypes, and that over-reliance on a weed management system based solely on herbicides is not sustainable. Hopefully we have learned that a diverse weed management program that combines multiple methods is the only system that will work for the long term.
APA, Harvard, Vancouver, ISO, and other styles
6

Fuerst, E. Patrick, Charles J. Arntzen, Klaus Pfister, and Donald Penner. "Herbicide Cross-Resistance in Triazine-Resistant Biotypes of Four Species." Weed Science 34, no. 3 (May 1986): 344–53. http://dx.doi.org/10.1017/s0043174500066960.

Full text
Abstract:
The cross-resistance of triazine-resistant biotypes of smooth pigweed (Amaranthus hybridusL. # AMACH), common lambsquarters (Chenopodium albumL. # CHEAL), common groundsel (Senecio vulgarisL. # SENVU), and the crop canola (Brassica napusL. var. Atratower) to a selection of herbicides was evaluated at both the whole plant and chloroplast level. The triazine-resistant biotypes of all four species showed a similar pattern of cross-resistance, suggesting that a similar mutation had occurred in each species. The four triazine-resistant biotypes were resistant to injury from atrazine [6-chloro-N-ethyl-N′-(1-methylethyl)-1,3,5-triazine-2,4-diamine], bromacil [5-bromo-6-methyl-3-(1-methylpropyl)-2,4-(1H,3H)pyrimidinedione], and pyrazon [5-amino-4-chloro-2-phenyl-3(2H)-pyridazinone] and were slightly resistant to buthidazole {3-[5-(1,1-dimethylethyl)-1,3,4-thiadiazol-2-yl]-4-hydroxy-1-methyl-2-imidazolidinone}. The triazine-resistant biotypes were more sensitive to dinoseb [2-(1-methylpropyl)-4,6-dinitrophenol]. Triazine-resistant smooth pigweed showed resistance to cyanazine {2-[[4-chloro-6-(ethylamino)-1,3,5-triazin-2-yl] amino]-2-methylpropanenitrile} and metribuzin [4-amino-6-(1,1-dimethylethyl)-3-(methylthio)-1,2,4-triazin-5(4H)-one] with slight resistance to linuron [N′-(3,4-dichlorophenyl)-N-methoxy-N-methylurea] and desmedipham {ethyl [3-[[(phenylamino)carbony] oxy] phenyl] carbamate}. There was little or no resistance to diuron [N′-(3,4-dichlorophenyl)-N,N-dimethylurea], bromoxynil (3,5-dibromo-4-hydroxybenzonitrile), bentazon [3-(1-methylethyl)-(1H)-2,1,3-benzothiadiazin-4(3H)-one 2,2-dioxide], or dicamba (3,6-dichloro-2-methoxybenzoic acid). Parallel studies at the chloroplast level indicated that the degree of resistance to inhibition of photosynthetic electron transport was highly correlated with the degree of resistance to herbicidal injury. This correlation indicates that atrazine, cyanazine, metribuzin, pyrazon, bromacil, linuron, desmedipham, and buthidazole cause plant injury by inhibition of photosynthesis. This correlation also indicates that triazine resistance and cross-resistance at the whole plant level is due to decreased sensitivity at the level of photosynthetic electron transport. Cross-resistance to numerous additional herbicides was evaluated on isolated chloroplast thylakoid membranes and these results are discussed.14C-atrazine was displaced from thylakoid membranes by several herbicides, indicating that these herbicides compete for a common binding site.
APA, Harvard, Vancouver, ISO, and other styles
7

Brooks, David R., Suzanne J. Clark, Joe N. Perry, David A. Bohan, Gillian T. Champion, Les G. Firbank, Alison J. Haughton, Cathy Hawes, Matthew S. Heard, and Ian P. Woiwod. "Invertebrate biodiversity in maize following withdrawal of triazine herbicides." Proceedings of the Royal Society B: Biological Sciences 272, no. 1571 (June 28, 2005): 1497–502. http://dx.doi.org/10.1098/rspb.2005.3102.

Full text
Abstract:
Responses of key invertebrates within Farm Scale Evaluations (FSEs) of maize reflected advantageous effects for weeds under genetically modified herbicide-tolerant (GMHT) management. Triazine herbicides constitute the main weed control in current conventional systems, but will be withdrawn under future EU guidelines. Here, we reappraise FSE data to predict effects of this withdrawal on invertebrate biodiversity under alternative management scenarios. Invertebrate indicators showed remarkably consistent and sensitive responses to weed abundance. Their numbers were consistently reduced by atrazine used prior to seedling emergence, but at reduced levels compared to similar observations for weeds. Large treatment effects were, therefore, maintained for invertebrates when comparing other conventional herbicide treatments with GMHT, despite reduced differences in weed abundance. In particular, benefits of GMHT remained under comparisons with best estimates of future conventional management without triazines. Pitfall trapped Collembola, seed-feeding carabids and a linyphiid spider followed closely trends for weeds and may, therefore, prove useful for modelling wider biodiversity effects of herbicides. Weaker responses to triazines applied later in the season, at times closer to the activity and capture of invertebrates, suggest an absence of substantial direct effects. Contrary responses for some suction-sampled Collembola and the carabid Loricera pilicornis were probably caused by a direct deleterious effect of triazines.
APA, Harvard, Vancouver, ISO, and other styles
8

Gressel, Jonathan, and Lee A. Segel. "Negative Cross Resistance; a Possible Key to Atrazine Resistance Management: A Call for Whole Plant Data." Zeitschrift für Naturforschung C 45, no. 5 (May 1, 1990): 470–73. http://dx.doi.org/10.1515/znc-1990-0528.

Full text
Abstract:
Many photosystem II inhibiting herbicides still inhibit this process in triazine-resistant plants; i.e. they have no cross resistance with atrazine. Five- to twenty-fold lower concentrations of phenolic type herbicidcs and 5-fold less of the active ingredient of pyridate and half as much ioxynil are required to inhibit thylakoid PS II in atrazine-resistant biotypes than in sensitive biotypes; i.e., they even show “negative cross resistance”. Negative cross resistance may be the major reason that atrazine resistance did not evolve where herbicide mixtures were used, when the mixed herbicide (usually a non-PS II inhibiting acetanilide) also controlled triazine-sensitivc weeds. Mathematical modeling in principle allows quantification of the very low field levels of herbicides possessing negative cross resistance that could be mixed with atrazine that would stop or delay the evolution of resistant populations without affecting the maize crop. There are few available actual dose response curves of atrazine-resistant vs. susceptible weeds at the whole plant level for herbicidcs exerting negative cross resistance. Thus, “real situation” modeling cannot be done. Data acquisition is called for so that the model can be extrapolated from the thylakoid to the field.
APA, Harvard, Vancouver, ISO, and other styles
9

Elezovic, Ibrahim, Dragana Bozic, and Sava Vrbnicanin. "Weed resistance to herbicides states: Causes and possibilities of preventive resistance." Pesticidi 18, no. 1 (2003): 5–21. http://dx.doi.org/10.2298/pif0301005e.

Full text
Abstract:
Resistance occurs as a result of heritable changes to biochemical processes that enable plant survival when treated with a herbicide. Resistance can result from changes to the herbicides target site such that binding of the herbicide is reduced, or over-expression of the target site may occur. Alternatively, there may be a reduction in the amount of herbicide that reaches the target enzyme through detoxication, sequestration, or reduced absorption of herbicide. Finally, the plant may survive through the ability to protect plant metabolism from toxic compounds produced as a consequence of herbicide action. Herbicide-resistant weeds were predicted shortly after the introduction of herbicides. During the 1970s, many, additional important weed species (e.g., Amaranthus spp., Chennpodium spp., Erigeron canadensis Kochia scoparia, Solanum nigrum, Panicum crus-galli, Senecio vulgaris, Poa annua) were reported to be resistant to triazine herbicides and several other herbicides. Over the last 10 years and now ALS-herbicide-resistant weeds account for the greatest number of resistant species and probably the largest area affected by resistance. In contrast to triazine resistance target-site-based resistance to the ALS-inhibiting herbicides can be conferred by a number of different point mutations. Differences occur in target-site cross-resistance among the different chemical classes of herbicides that inhibit ALS. The differences are related to particular amino acid substitutions that occur within the binding region. Indeed, six different substitutions of Ala, Arg, Glu, Leu, Ser, or Tri for Pro 173 have been observed in different weed populations.
APA, Harvard, Vancouver, ISO, and other styles
10

Burnet, Michael W. M., Orville B. Hildebrand, Joseph A. M. Holtum, and Stephen B. Powles. "Amitrole, Triazine, Substituted Urea, and Metribuzin Resistance in a Biotype of Rigid Ryegrass (Lolium rigidum)." Weed Science 39, no. 3 (September 1991): 317–23. http://dx.doi.org/10.1017/s0043174500072994.

Full text
Abstract:
A biotype of rigid ryegrass (Lolium rigidum G. ♯ LOLRI) has become resistant to amitrole and atrazine after 10 yr of exposure to a mixture of these herbicides. Resistance has also been demonstrated to the chloro-s-triazines: simazine, cyanazine, propazine; the methylthios-triazines: ametryn, prometryn; the substituted ureas: chlortoluron, isoproturon, metoxuron, diuron, fluometuron, methazole; and the triazinone herbicide metribuzin. The biotype remains susceptible to chlorsulfuron, metsulfuron, sulfometuron, sethoxydim, diclofop, fluazifop, glyphosate, carbetamide, and oxyfluorfen. Inhibition of oxygen evolution by atrazine, diuron, and metribuzin was similar in thylakoids isolated from both resistant and susceptible biotypes. Therefore, resistance to the photosystem II inhibitors is not caused by an alteration of the target site of these herbicides. Resistant plants treated with a 3-h pulse of 0.12 mM chlortoluron recover photosynthetic activity more rapidly than susceptible plants. This suggests that the basis for resistance is enhanced metabolism or sequestration of the herbicide within the leaf.
APA, Harvard, Vancouver, ISO, and other styles

Dissertations / Theses on the topic "Triazine herbicide"

1

Blyden, E. R. "Molecular genetics of triazine resistance in Senecio vulgaris L." Thesis, University of Cambridge, 1988. http://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.383076.

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

PEREZ, YAFA. "Toxicologie et ecotoxicologie des triazines utilisees comme herbicides et particulierement des chlorotriazines." Strasbourg 1, 1987. http://www.theses.fr/1987STR10719.

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

Vencill, William K. "Triazine resistance in Chenopodium album and Amaranthus hybridus in Virginia." Thesis, Virginia Polytechnic Institute and State University, 1986. http://hdl.handle.net/10919/94485.

Full text
Abstract:
Studies were conducted to determine the distribution of s-triazine resistant biotypes of common lambsquarters (Chenopodium album L.) and smooth pigweed (Amaranthus hybridus L.) in Virginia. Collections of seed were made from suspected triazine-resistant biotypes of common lambsquarters and smooth pigweed from counties in Virginia which had reported having triazine resistance problems. Triazine resistance was confirmed by measuring chlorophyll fluorescence in the presence of atrazine. For further confirmation of triazine resistance in collected common lambsquarters and smooth pigweed biotypes, greenhouse testing of whole plants and a sinking leaf disc assay were performed. Cross-resistance to another s-triazine, as-triazine, and substituted urea herbicide was also determined for s-triazine-resistant biotypes. These studies have shown triazine- resistant smooth pigweed to be present in 19 counties and common lambsquarters to be present in 8 counties in Virginia. s-Triazine resistant biotypes were found to be resistant to another s-triazine and as-triazine herbicide, but were susceptible to the substituted urea herbicide. Additional studies were initiated to determine the effects of different temperature regimes on triazine-resistant and -susceptible biotypes of common lambsquarters and smooth pigweed from different geographical locations. These studies were conducted at the North Carolina State University Phytotron facility in controlled environment growth chambers. Triazine-resistant common lambsquarters biotypes from Virginia, Maryland, and Switzerland as well as a smooth pigweed biotype from Virginia were examined. Triazine-susceptible biotypes of common lambsquarters and smooth pigweed were included as controls. Shoot height, weight, chlorophyll a and b content, and whole leaf fatty acid content of common lambsquarters and smooth pigweed were determined at 18°/14° C, 26°/22° C, and 36°/26° C. Measurements of shoot height were made at 30 and 63 days after planting. The shoot weight, chlorophyll a and b content, and fatty acid content was determined from plants harvested at 63 days after planting. These data indicate common lambsquarters biotypes from different geographical regions exhibited a differential response to temperature. There was no difference between triazine-resistant and -susceptible biotypes in response to temperature. Differences were detected between triazine-resistant smooth pigweed biotypes which indicated that the susceptible biotypes were more vigorous as indicated by shoot height and weight at lower temperatures than triazine-resistant biotypes of smooth pigweed.
M.S.
APA, Harvard, Vancouver, ISO, and other styles
4

Biagianti, Sylvie. "Contribution à l'étude du foie de juvéniles de muges (Téléostéens, Mugilides), contaminés expérimentalement par l'atrazine (s-triazine herbicide) : approche ultrastructurale et métabolique : intérêt en écotoxicologie." Perpignan, 1990. http://www.theses.fr/1990PERP0084.

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

Andrade, Felipe Nascimento. "Síntese e emprego de polímeros molecularmente impressos em técnicas miniaturizadas acopladas a cromatografia liquida para análises de triazinas e sulfoniluréias em amostras de milho." Universidade de São Paulo, 2015. http://www.teses.usp.br/teses/disponiveis/75/75135/tde-09032016-134923/.

Full text
Abstract:
As triazinas e sulfoniluréias tem sido muito empregadas, devido ao crescente aumento na produção agrícola e aplicação de herbicidas nas lavouras, podendo causar sérios riscos à saúde humana e ao meio ambiente. Uma problemática é a baixa concentração que estes analitos são encontrados, tornando-se necessário o emprego do preparo da amostra para a sua determinação. Com isso, a busca por técnicas de preparo de amostras miniaturizadas, simples, de baixo custo, com menores riscos de contaminações ambientais e baixo consome de solventes, tem grande predominância. Dentre as microtécnicas de preparo de amostra podemos destacar a microextração em dispositivos preenchidos com sorventes (MEPS, do inglês, Microextraction by Packed Sorbent) e a microextração adsortiva em barra (BAμE, do inglês, Bar Adsorptive Microextraction). Outro aspecto desejado no preparo de amostras é a obtenção de uma maior seletividade quanto ao sorvente empregado, quando comparado àqueles convencionais como, por exemplo, sílica modificada (ex. C18), resinas Amberlite XAD, entre outros. Nesse âmbito, o presente trabalho apresenta a síntese de dois polímeros impressos molecularmente (MIP, do inglês, Molecularly Imprinted Polymers) e suas aplicações na adsorção seletiva de moléculas de triazinas e sulfoniluréias, com separação, identificação e quantificação feitas por LC-TOF-ESI e HPLC-DAD. A primeira metodologia desenvolvida foi a síntese dos polímeros impressos empregando o ácido metacrílico e o etileno glicol dimetacrilato. Após a síntese, os polímeros foram caracterizados através da espectroscopia de infravermelho e microscopia eletrônica de varredura. O coeficiente de seletividade para o MIP foi comparado com o coeficiente de seletividade do polímero não impresso (NIP, do inglês, non imprinted polymers) para misturas binárias de atrazina/picloram, atrazina/propanil, bensulfuron/betazon e bensulfuron/prometon onde os valores dos coeficientes de seletividade relativa (k\') obtidos foram de 17,2, 3,2, 10,6 e 8,5. A seguir, foram desenvolvidos dois métodos empregando MEPS para as triazinas e sulfoniluréias, respectivamente. Os métodos validados baseando-se nas recomendações da Agência Nacional de Vigilância Sanitária (ANVISA) e com diretrizes da Comunidade Europeia, apresentaram linearidade, seletividade, precisão, exatidão e recuperação adequadas para as triazinas e sulfoniuréias. Os limites de quantificação obtidos foram da ordem de 5,0-10,0 μg kg-1 para as triazinas e 2,5 μg kg-1 para as sulfoniluréias. Ainda com propósito de determinar esses herbicidas em milho, dois novos métodos de preparo de amostra foram desenvolvidos, empregando a BAμE. A BAμE, foi desenvolvida recentemente e combinada com os polímeros molecularmente impressos, avaliou-se sua determinação para as triazinas e sulfoniluréias. As variáveis da técnica BAμE foram otimizadas e, em seguida, validadas. Os limites de quantificação obtidos foram da ordem de 0,7 μg kg-1 para as triazinas e 0,4 μg kg-1 para as sulfoniluréias. Os métodos propostos foram aplicados com sucesso para determinação de triazinas em diferentes amostras de milho verde, com valores de recuperação satisfatórios na faixa de 80,0 -120,9%.
Triazines and sulfonylureas have been much used due to the growing in agricultural production and application of herbicides on crops, which may cause serious risks to human health and the environment. One problem is the low concentration found of these analytes, making it necessary the use of the sample preparation. Thus, the search for miniaturized sample preparation such as techniques simple, low cost, with less risk of environmental contamination and low solvent consumes has great predominance. Among the sample preparation microtechnology we can highlight the microextraction by packed sorbent (MEPS) and bar adsorptive microextraction (BA?E). Another desired aspect for sample preparation is to obtain a higher selectivity concerning the sorbent employed as compared to their conventional format; for example, modified silica (ex. 18), Amberlite XAD resins, and others. In this context, this paper presents the synthesis of two molecularly imprinted polymers (MIPs) and their applications in the selective adsorption of triazines and sulfonylureas molecules being the separation, identification and quantification steps made by LC-TOF-ESI and HPLC-DAD. The first step developed was the synthesis of imprinted polymers using methacrylic acid and ethylene glycol dimethacrylate. After synthesizing, the polymers were characterized by infrared spectroscopy and scanning electron microscopy. The MIP selectivity coefficient was compared with the NIP (non imprinted polymer) selectivity coefficient employing binary mixtures of atrazine / picloram, atrazine / propanil, bensulfuron / betazon and bensulfuron / prometon where the values of relative selectivity coefficients (k \') obtained were 17.2, 3.2, 10.6 and 8.5. Next, we developed two methods for MEPS triazines and sulfonylureas, respectively. The methods were validated based on the recommendations of the National Health Surveillance Agency (ANVISA) and the European Community directives, and presented linearity, selectivity, precision, accuracy and adequate recovery for the triazines and sulfoniuréias. The quantification limits were obtained in the range of 5.0-10.0 μg kg-1 for triazines and 2.5 μg kg-1 for sulfonylureas. Aiming determining these herbicides in corn samples, two new sample preparation methods were developed, using the BAμE. The recently developed BAμE, was combined with molecularly imprinted polymers to evaluate the determination to triazines and sulfonylureas. The BAμE variables were optimized and validated. The quantification limits were obtained in the range of 0.7 μg kg-1 for triazines and 0.4 μg kg-1 for sulfonylureas. The proposed methods were successfully applied for the determination of triazines in different samples of corn, with satisfactory recovery values in the range of 80.0 -120.9%.
APA, Harvard, Vancouver, ISO, and other styles
6

Malotaux, Christophe. "Les Triazines-atrazine entre autres, présences dans l'environnement et dans l'eau." Paris 5, 1992. http://www.theses.fr/1992PA05P242.

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

Marchese, Luciana. "Sorção/dessorção e lixiviação do herbicida ametrina em solos canavieiros tratados com lodo de esgoto." Universidade de São Paulo, 2007. http://www.teses.usp.br/teses/disponiveis/64/64135/tde-03092007-154220/.

Full text
Abstract:
Vários estudos têm mostrado os benefícios da aplicação agronômica de lodo de esgoto sobre as propriedades físicas e químicas do solo. No entanto, poucos são aqueles que avaliam o impacto desta prática sobre o comportamento de pesticidas em solos tropicais. O principal objetivo deste estudo foi avaliar o efeito da adição de diferentes fontes de lodo (Ribeirão Preto, Franca e Jundiaí) na sorção / dessorção e lixiviação do herbicida ametrina em solos canavieiros do estado de São Paulo (Neossolo Quartzarênico Órtico Típico (RQ), Latossolo Vermelho Distrófico Típico (LVd), Argissolo Vermelho Eutroférrico Chernossólico (PV) e Latossolo Vermelho Distroférrico (LVdf). Cinco concentrações (de 4,4 a 79,8 mg i.a. L-1) do herbicida foram utilizadas para obter os valores dos coeficientes de sorção de Freündlich (Kf) e dos coeficientes aparentes de sorção médio e para a menor concentração (Kd ap médio e Kd ap [menor]), a qual corresponde à dose de campo recomendada. Para o teste de lixiviação, o método utilizado foi o de lixiviação em colunas de solo (diâmetro = 5 cm e comprimento = 30 cm), utilizando-se três repetições para cada tratamento, sobre as quais foi simulada chuva de 200 mm uniformemente distribuídas durante 48 h, após a aplicação da solução de 14Cametrina na dose de 3,0 kg i.a. ha-1. De forma geral, a sorção da ametrina variou de moderada à alta em todos os tratamentos (2,68 < Kd ap [menor] < 85,71 L kg-1). Solos argilosos com maior teor de matéria orgânica e argilas do tipo 2:1, como é o caso do PV, apresentaram muito maior potencial de sorção da ametrina; enquanto que solos arenosos com baixos teores de matéria orgânica, como foram os casos do LVd e RQ, apresentaram moderado potencial de sorção. A aplicação de lodos menos estabilizados, com biomassa e material orgânico menos recalcitrante e, portanto, com maiores valores de carbono orgânico total, carbono orgânico dissolvido e pH, como é o caso do lodo de Ribeirão Preto, tendeu a diminuir o potencial de sorção da ametrina. Já a adição de lodos mais recalcitrantes, como o de Jundiaí, tendeu a aumentar o potencial de sorção da ametrina, principalmente em solos arenosos devido a sua menor capacidade tampão. Nestes cenários, poderá ocorrer redução da eficácia agronômica ametrina, uma vez que haverá menos produto disponível na solução do solo. A ametrina apresentou baixo potencial de lixiviação (< 1% da quantidade aplicada) em todos os tratamentos, os quais não apresentaram diferenças entre si, sendo que a grande maioria do pesticida (> 95% da quantidade aplicada) ficou retida na camada de 0-10 cm de profundidade da coluna de solo. Isto implica dizer que a ametrina apresenta baixo potencial de contamina águas subterrâneas, mesmo em solos arenosos, como RQ (90% de areia)
Several studies have shown the benefits of applying sewage sludge on the physical and chemical properties of the soils. However, just a few of them evaluates the impact of this practice on the behavior of pesticides in tropical soils. The main goal of this research was to evaluate the effects of applying different sources of sewage sludge (Ribeirão Preto, Franca e Jundiaí) on the sorption / desorption and leaching of ametryne in soils from São Paulo state (Brazil) cultivated with sugarcane (Neossolo Quartzarênico Órtico Típico (RQ), Latossolo Vermelho Distrófico Típico (LVd), Argissolo Vermelho Eutroférrico Chernossólico (PV) e Latossolo Vermelho Distroférrico (LVdf). Five concentrations (4.4 to 79.8 mg a.i. ha-1) of the herbicide were applied to the soil samples to attain the Freündlich sorption coefficients (Kf) and the apparent sorption coefficient for the lower concentration (Kd ap [menor]), which corresponded to the field application rate. For the leaching test the adopted method was the soil leaching columns (diameter = 5 cm and depth = 30 cm), in triplicates, over which a 200 mm rainfall evenly distributed during 48 h was simulated just after ametryne application at the rate of 3.0 kg a.i. ha-1. In general, ametryne sorption ranged from moderate to high in all treatments (2.68 < Kd ap [menor] < 85.71 L kg-1). Clay soils with higher organic matter and 2:1 clay contents, such as the PV, showed much higher sorption potential, whereas sand soils with low organic matter content, such as LVd and RQ, showed moderate sorption potential. The application of less stabilized sewage sludges, with less recalcitrant biomass and organic material and, therefore, with higher organic matter and dissolved organic carbon contents and pH values, such as the Ribeirão Preto, tends to decrease ametryne sorption potential. Otherwise, the addition of more recalcitrant sludges, such as the Jundiaí, tends to enhance its sorption potential, mainly in sand soils due to its lower buffer capacity. In those cases, the agronomic efficacy of ametryne may be reduced since there is lower concentration of the compound available in the soil solution. The ametryne presented low leaching potential in all treatments (< 1% of the applied amount), which were not different among themselves, and that its majority (> 95% of the applied amount) was found at 0 -10 cm soil depth in the column. It implies that ametryne has low potential to contaminate groundwater even in sandy soils, such as the RQ (90% sand)
APA, Harvard, Vancouver, ISO, and other styles
8

Dupont, Stephane. "Bound (nonextractable) residues of triazine herbicides in soybean and canola plants." Thesis, University of Ottawa (Canada), 1989. http://hdl.handle.net/10393/21091.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Oketunde, Olukayode Felix. "The adsorptive behaviour of two triazine herbicides in three Nigerian soils." Thesis, National Library of Canada = Bibliothèque nationale du Canada, 1997. http://www.collectionscanada.ca/obj/s4/f2/dsk2/ftp04/mq22372.pdf.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Olson, B. M. "Spectroscopic study of interactions between s-triazine herbicides and humic substances." Thesis, University of Aberdeen, 1986. http://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.374630.

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

Books on the topic "Triazine herbicide"

1

Langland, Michael J. Nutrient and triazine-herbicide concentrations in streams of the Chickies Creek Basin, south-central Pennsylvania, during low-flow conditions. [Lemoyne, Pa.]: U.S. Dept. of the Interior, U.S. Geological Survey, 1996.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
2

Langland, Michael J. Nutrient and triazine-herbicide concentrations in streams of the Chickies Creek Basin, south-central Pennsylvania, during low-flow conditions. [Lemoyne, Pa.]: U.S. Dept. of the Interior, U.S. Geological Survey, 1996.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
3

Langland, Michael J. Nutrient and triazine-herbicide concentrations in streams of the Chickies Creek Basin, south-central Pennsylvania, during low-flow conditions. [Lemoyne, Pa.]: U.S. Dept. of the Interior, U.S. Geological Survey, 1996.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
4

Langland, Michael J. Nutrient and triazine-herbicide concentrations in streams of the Chickies Creek Basin, south-central Pennsylvania, during low-flow conditions. [Lemoyne, Pa.]: U.S. Dept. of the Interior, U.S. Geological Survey, 1996.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
5

Langland, Michael J. Nutrient and triazine-herbicide concentrations in streams of the Chickies Creek Basin, south-central Pennsylvania, during low-flow conditions. [Lemoyne, Pa.]: U.S. Dept. of the Interior, U.S. Geological Survey, 1996.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
6

Langland, Michael J. Nutrient and triazine-herbicide concentrations in streams of the Chickies Creek Basin, south-central Pennsylvania, during low-flow conditions. [Lemoyne, Pa.]: U.S. Dept. of the Interior, U.S. Geological Survey, 1996.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
7

Langland, Michael J. Nutrient and triazine-herbicide concentrations in streams of the Chickies Creek Basin, south-central Pennsylvania, during low-flow conditions. [Lemoyne, Pa.]: U.S. Dept. of the Interior, U.S. Geological Survey, 1996.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
8

Langland, Michael J. Nutrient and triazine-herbicide concentrations in streams of the Chickies Creek Basin, south-central Pennsylvania, during low-flow conditions. [Lemoyne, Pa.]: U.S. Dept. of the Interior, U.S. Geological Survey, 1996.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
9

Ballantine, Larry G., Janis E. McFarland, and Dennis S. Hackett, eds. Triazine Herbicides: Risk Assessment. Washington, DC: American Chemical Society, 1998. http://dx.doi.org/10.1021/bk-1998-0683.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

1944-, Ballantine Larry Gene, McFarland Janis E. 1956-, and Hackett Dennis S. 1950-, eds. Triazine herbicides: Risk assessment. Washington, DC: American Chemical Society, 1998.

Find full text
APA, Harvard, Vancouver, ISO, and other styles

Book chapters on the topic "Triazine herbicide"

1

Bonora, S., G. Fini, and A. Torreggiani. "Interaction of Herbicide Triazine Derivatives with Model Membranes." In Spectroscopy of Biological Molecules, 415–16. Dordrecht: Springer Netherlands, 1995. http://dx.doi.org/10.1007/978-94-011-0371-8_191.

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

Lawruk, Timothy S., Charles S. Hottenstein, James R. Fleeker, Fernando M. Rubio, and David P. Herzog. "Factors Influencing the Specificity and Sensitivity of Triazine Immunoassays." In Herbicide Metabolites in Surface Water and Groundwater, 43–52. Washington, DC: American Chemical Society, 1996. http://dx.doi.org/10.1021/bk-1996-0630.ch004.

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

Koskinen, W. C., J. S. Conn, and B. A. Sorenson. "Fate of a Symmetric and an Asymmetric Triazine Herbicide in Silt Loam Soils." In Herbicide Metabolites in Surface Water and Groundwater, 125–39. Washington, DC: American Chemical Society, 1996. http://dx.doi.org/10.1021/bk-1996-0630.ch011.

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

Wackett, Lawrence P. "Evolution of New Enzymes and Pathways: Soil Microbes Adapt tos-Triazine Herbicide." In ACS Symposium Series, 37–48. Washington, DC: American Chemical Society, 2003. http://dx.doi.org/10.1021/bk-2004-0863.ch004.

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

Panneels, P., A. Van Moer, P. Reimer, P. Salis, A. Chouhiat, R. Lannoye, and Figeys H. "Fluorescence Investigation of DCMU and S-Triazine Herbicide Activity in Crop and Weed Species." In Progress in Photosynthesis Research, 827–30. Dordrecht: Springer Netherlands, 1987. http://dx.doi.org/10.1007/978-94-017-0516-5_175.

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

Edmunds, Andrew J. F. "Triazine Herbicides." In Bioactive Heterocyclic Compound Classes, 21–38. Weinheim, Germany: Wiley-VCH Verlag GmbH & Co. KGaA, 2012. http://dx.doi.org/10.1002/9783527664412.ch2.

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

Muir, D. C. G. "Triazine Herbicides." In Mass Spectrometry in Environmental Sciences, 423–35. Boston, MA: Springer US, 1985. http://dx.doi.org/10.1007/978-1-4613-2361-7_19.

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

Gianessi, Leonard P. "Benefits of Triazine Herbicides." In ACS Symposium Series, 1–8. Washington, DC: American Chemical Society, 1998. http://dx.doi.org/10.1021/bk-1998-0683.ch001.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Karu, A. E., Robert O. Harrison, D. J. Schmidt, C. E. Clarkson, J. Grassman, M. H. Goodrow, A. Lucas, B. D. Hammock, J. M. Van Emon, and R. J. White. "Monoclonal Immunoassay of Triazine Herbicides." In ACS Symposium Series, 59–77. Washington, DC: American Chemical Society, 1990. http://dx.doi.org/10.1021/bk-1990-0451.ch006.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Giddings, Jeffrey M., and Lenwood W. Hall. "The Aquatic Ecotoxicology of Triazine Herbicides." In ACS Symposium Series, 347–56. Washington, DC: American Chemical Society, 1998. http://dx.doi.org/10.1021/bk-1998-0683.ch027.

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

Conference papers on the topic "Triazine herbicide"

1

Autullo, Mattia, Mauro Mennuni, Mauro Giustini, Marcello Giomini, Francesco Lopez, Antonia Mallardi, and Gerardo Palazzo. "Triazine herbicides determination in water with an optical biosensor." In 2009 3rd International Workshop on Advances in sensors and Interfaces. IEEE, 2009. http://dx.doi.org/10.1109/iwasi.2009.5184798.

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

Scutariu, Roxana-Elena, Vasile Ion Iancu, Gheorghe Nechifor, Gabriel-Lucian Radu, Marius Simion, and Marcela Niculescu. "MEMBRANE FILTRATION EFFICIENCY ON TRIAZINE HERBICIDES IN ORGANIC AND AQUEOUS SOLUTIONS." In International Symposium "The Environment and the Industry". National Research and Development Institute for Industrial Ecology, 2018. http://dx.doi.org/10.21698/simi.2018.fp48.

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

Reports on the topic "Triazine herbicide"

1

Nutrient and triazine-herbicide concentrations in streams of the Chickies Creek Basin, south-central Pennsylvania, during low-flow conditions. US Geological Survey, 1996. http://dx.doi.org/10.3133/wri964073.

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

Occurrence of phosphorus, other nutrients, and triazine herbicides in water from the Hillsdale Lake basin, Northeast Kansas, May 1994 through May 1995. US Geological Survey, 1997. http://dx.doi.org/10.3133/wri974019.

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