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 546 毫秒
1.
The degradation of diuron photoinduced by iron(III) in aqueous solution has been investigated with different iron(III) species (monomeric species Fe(OH)2+, dimeric species Fe2(OH)24+ and water-soluble oligomeric species) under monochromatic excitation at 365 nm and under sunlight. The rate of degradation depends on the concentration in Fe(OH)2+, the most reactive species in terms of OH radical formation. The major photoproduct is 3-(3,4-dichlorophenyl)-1-formyl-1-methylurea which represents more than 60% of diuron disappearance. The mechanism only involves the attack by OH radicals arising from iron(III) excited species. The half-lives of diuron when submitted to such a process in the environment were estimated to be 1–2 h and a few days according to the concentration of Fe(OH)2+ (respectively 70% and <10% of total iron(III) concentration).  相似文献   

2.
Paraquat resistance in a biotype of Vulpia bromoides (L.) S. F. Gray   总被引:1,自引:0,他引:1  
A biotype of Vulpia bromoides from a lucerne field in Elmhurst, Victoria, Australia was shown to be resistant to paraquat in pot trials. Application of paraquat at 50 g a.i. ha?1 killed all of the plants of a susceptible Vulpia bromoides biotype but only 6% of the resistant biotype. The LD50 for paraquat of the resistant biotype was five- to sixfold higher than for the susceptible biotype. The resistant biotype was also resistant to the bipyridyl herbicide diquat, but was susceptible to glyphosate and metribuzin. Application of 100 g a.i. ha?1 paraquat at anthesis completely suppressed seed set of the susceptible biotype and reduced that of the resistant biotype by 95%. Seed set by the paraquat-treated resistant biotype, however, showed little reduction in germination. This is the fourth species to have been found to be resistant to bipyridyl herbicides in this field, the others being Hordeum glaucum, H. leporinum and Arctotheca calendula. Résistance au paraquat d'un biotype de Vulpia bromoides (L.) S. F. Gray Un biotype de Vulpia bromoides issu d'un champ de luzerne à Elmhurst, Victoria, Australie s'est révélé résistant au paraquat lors d'essais en pots. Des traitements au paraquat 50 g m.a. ha?1 détruisaient toutes les plantes d'un biotype sensible de V. bromoides mais seulement 6% du biotype résistant. La DLV50 du paraquat pour ce biotype était 5 à 6 fois plus élevée que pour le biotype sensible. Le biotype résistant l'était aussi à l'herbicide bipyridyle diquat, mais était sensible au glyphosate et à la métribuzine. Des traitements au paraquat 100 g m.a. ha?1 au stade anthèse supprimaient complètement la production de graines du biotype sensible et réduisait de 95% celle du biotype résistant. Cependant, le pouvoir germinatif des graines issues du biotype résistant traité, n'était que peu réduit. Après Hordeum glaucum, Hordeum leporinum et Arctotheca calendula, c'est la quatrième espèce trouvée dans ce même champ résistante aux herbicides bipyridyles. Paraquatresistenz eines Biotyps von Vulpia bromoides (L.) S. F. Gray Ein Biotyp von Vulpia bromoides von einem Luzernefeld in Elmhurst, Victoria, Australien, erwies sich in Topfversuchen als paraquatresistent. Mit 50 g AS ha?1 wurden Pflanzen eines empfindlichen Biotyps abgetötet, jedoch nur 6% des resistenten. Die LD50 des resistenten Biotyps für paraquat war 5- bis 6mal höher als die des empfindlichen Biotyps. Der resistente Biotype war auch gegenüber dem Bipyridylherbizid Deiquat résistent, gegenüber Glyphosat und Metribuzin jedoch empfindlich. Nach Anwendung von 100 g AS ha?1 Paraquat vor der Blüte bildeten sich bei dem empfindlichen Biotyp keine Samen aus, bei dem resistenten waren sie um 95% vermindert; die dennoch gebildeten Samen keimten etwas weniger. Dies ist die vierte Pflanzenart, bei der eine Resistenz gegenüber Bipyridylherbiziden beobachtet wurde, die anderen sind Hordeum glaucum. Hordeum leporinum und Arctotheca calendula.  相似文献   

3.
Adsorption–desorption characteristics of 2,4-dichlorophenoxyacetic acid (2,4-D) on pure montmorillonite and synthetic chlorite-like complexes [Al(OH)x-montmorillonite complexes, obtained by coating montmorillonite surfaces with different amounts of Al(OH)x] were investigated. The equilibrium adsorption of 2,4-D was described by both Langmuir and Freundlich type isotherms. The extent of adsorption as well as the type of interaction between adsorbate and adsorbent was affected by the nature of incubation buffer and the charge characteristics of supports. At pH 5·6 and in acetate buffer, 2,4-D was negatively adsorbed by montmorillonite and herbicide adsorption capacity increased with increasing amounts of Al(OH)x species loaded on montmorillonite surfaces. When adsorption experiments were performed at the same pH but in phosphate buffer, strong reductions of both the amount of adsorbed pesticide and its affinity for the adsorbents were measured. Evidently, phosphate anions competed strongly with 2,4-D anions for the sorption site on chlorite-like complexes. Furthermore, desorption tests revealed that a large amount (about 60%) of the pesticide was firmly bound to the clay and was not removed even after repeated washings or 24 h exposure to desorption solution. Both electrostatic interactions between the negative COO- moieties of 2,4-D and the positive sites on clays, and ligand exchanges of COO- groups with -OH or water at the clay surface were probably involved in the adsorption process. ©1997 SCI  相似文献   

4.
Two biotypes of the grass weed barley grass (Hordeum glaucum), one resistant and the other susceptible to the herbicide paraquat, were studied along with their F1, F2 and F3 progeny to determine the inheritance of paraquat resistance. The plants were sprayed with 50–200 g a.i. paraquat ha?1. These concentrations killed the susceptible type. The data obtained from segregating populations indicated that paraquat resistance in H. glaucum is controlled by a single nuclear gene with incomplete dominance.  相似文献   

5.
The adsorption of paraquat (1,1′‐dimethyl‐4,4′‐bipyridilium dichloride) and atrazine (6‐chloro‐N 2‐ethyl‐N 4‐isopropyl‐1,3,5‐triazine‐2,4‐diamine) from aqueous solution onto two montmorillonite‐(Ce or Zr) phosphate cross‐linked compounds at different temperatures (288 K and 308 K) has been studied using batch experiments. The adsorption isotherms obtained for paraquat on both adsorbents may be classified as H‐type of the Giles classification, which suggests that paraquat molecules are strongly adsorbed on the samples. For the adsorption of atrazine, L‐type isotherms were obtained for both montmorillonite‐(Ce or Zr) phosphate compounds, which suggests that these compounds have a medium affinity for this herbicide. The increase in temperature from 288 K to 308 K did not have any clear effect on the adsorption process of paraquat on either adsorbent whereas atrazine adsorption decreased slightly as temperature increased, possibly due to a mainly physical process. Fourier transform infrared (FTIR) spectroscopic studies revealed that at the pH generated by the adsorbents, the cationic herbicide interacted to a greater extent with the negatively charged surface of the adsorbents than did atrazine. For both herbicides, the Ce‐montmorillonite adsorbent showed a higher adsorption capacity than the Zr‐montmorillonite adsorbent. © 2000 Society of Chemical Industry  相似文献   

6.
Adsorption of transformation products of atrazine by soil   总被引:1,自引:0,他引:1  
The adsorption of atrazine and its transformation products, desisopropylatrazine (2-chloro-4-ethylamino-6-amino-l,3,5-triazine), desethyl-atrazine (2-chloro-4-amino-6-isopropylamino-l,3,5-triazine) and hydroxyatrazine (2-hydroxy-4-ethylamino-6-isopropylamino-l,3,5-triazine) to four top-soils was measured. Adsorption coefficients decreased in the order hydroxy atrazine, atrazine, desisopropylatrazine, and desethyl-atrazine: the distribution coefficient between organic matter and water (KOM) ranged from 40 to 100 dm3 kg?1 for atrazine, from 30 to 60 dm3 kg?1 for desisopropylatrazine, from 20 to 50 dm3 kg?1 for desethylatrazine and from 100 to 590 dm3 kg?1 for hydroxy atrazine. Data are discussed in the context of earlier literature.  相似文献   

7.
Rehmannia glutinosa (Gaertn.) Libosch. ex Fisch. & Mey. was very tolerant to paraquat (1,1′-dimethyl-4,4′-bipyridinium). The paraquat concentration required to reduce dry weight of R. glutinosa by 50% (GR50) was 24 mM , whereas a similar effect was obtained with 0·75 mM in Zea mays L. (maize, cv. Dekalb) and Glycine max (L.) Merr. (soybean, cv. Kwangkyo). When 1·5 mM paraquat in 10% aqueous extract of R. glutinosa was applied to maize and soybean, growth inhibition reached 24% and 7%, respectively, of the untreated control. Decreased activity of paraquat due to the extract also occurred in both leaf discs and chloroplasts of soybean. The total amount of [14C]paraquat absorbed into soybean leaves after 48 h was 34%, but it was reduced to 17% when the extract was added. Translocation of [14C]paraquat was also inhibited in the presence of the extract. In thin-layer chromatography (TLC) analysis using various solvent systems, Rf values of [14C]paraquat with the extract differed from those without the extract. The results suggested that the aqueous extract of R. glutinosa contained a substance which could nullify paraquat activity. © 1997 SCI.  相似文献   

8.
A novel method was developed for the simultaneous determination of diquat and paraquat residues in potatoes. Potato tissues were spiked at several levels and extracted with acid using a micro-reflux procedure with 5 g of sample; this was followed by adjusting the hydrolyzate pH to 9 to 10 and using a silica Sep-Pak for rapid clean-up and preconcentration. Aliquots of the final eluate were taken to dryness, dissolved in the h.p.l.c. mobile phase and analyzed as their heptanesulfonate ion-pairs by u.v.-h.p.l.c. (reverse phase column chromatography) at 254 and 313 nm for paraquat and diquat, respectively. A detection limit of approximately 0.05 mg kg?1 dication in a 5-g sample of spiked potato (i.e. 0.25 μg ml?1 final extract) was achieved. Recoveries of 79.5 to 97.6% were obtained at spiking levels of 0.05 to 5.0 mg kg?1 for diquat and paraquat with coefficients of variation not greater than 8.27%. The method was developed and validated using 14C-radiolabelled diquat and paraquat; u.v.-h.p.l.c. recoveries were comparable with recoveries determined by radioassay. Several parameters affecting the extraction, adsorption and chromatography of diquat and paraquat were evaluated. The formula weights of diquat and paraquat were determined and their importance described; they were determined as mono- and tetra-hydrates, respectively.  相似文献   

9.
BACKGROUND: Variations in soil properties with depth influence retention and degradation of pesticides. Understanding how soil properties within a profile affect pesticide retention and degradation will result in more accurate prediction by simulation models of pesticide fate and potential groundwater contamination. Metolachlor is more persistent than other acetanilide herbicides in the soil environment and has the potential to leach into groundwater. Reasonably, information is needed about the dissipation and eventual fate of metolachlor in subsoils. The objectives were to evaluate the adsorption and desorption characteristics and to determine the dissipation rates of metolachlor in both surface and subsurface soil samples. RESULTS: Adsorption of metolachlor was greater in the high‐organic‐matter surface soil than in subsoils. Lower adsorption distribution coefficient (Kads) values with increasing depth indicated less adsorption at lower depths and greater leaching potential of metolachlor after passage through the surface horizon. Desorption of metolachlor showed hysteresis, indicated by the higher adsorption slope (1/nads) compared with the desorption slope (1/ndes). Soils that adsorbed more metolachlor also desorbed less metolachlor. Metolachlor dissipation rates generally decreased with increasing soil depth. The first‐order dissipation rate was highest at the 0–50 cm depth (0.140 week?1) and lowest at the 350–425 cm depth (0.005 week?1). Degradation of the herbicide was significantly correlated with microbial activity in soils. CONCLUSION: Metolachlor that has escaped degradation or binding to organic matter at the soil surface might leach into the subsurface soil where it will dissipate slowly and be subject to transport to groundwater. Copyright © 2009 Society of Chemical Industry  相似文献   

10.
Sprangletop (Leptochloa chinensis L. Nees) is a serious grass weed in direct‐seeded rice cropping systems in Thailand. One population of sprangletop, BLC1, was found to be resistant to fenoxaprop‐p‐ethyl at 62‐fold the concentration of a susceptible biotype, SLC1. This study elucidated the inheritance of resistance to fenoxaprop‐p‐ethyl in this sprangletop BLC1 genotype. The reaction to the herbicide at 0.12–2.4 mg ai L?1 was determined in the seedlings of self‐pollinated resistant BLC1, susceptible SLC1 and SLC1 that had been allowed to cross‐pollinate with BLC1. At 0.24 mg ai L?1, all the seedlings of SLC1 were killed, while 99% of BLC1 survived, along with 5% of the cross‐pollinated SLC1 seedlings, which were considered to be putative F1 hybrids. The root and shoot lengths of the F1 hybrids in 0.24 mg ai L?1 of fenoxaprop‐p‐ethyl, relative to those in the absence of the herbicide, were close to or the same as the resistant parent, indicating that the resistance is a nearly complete to complete dominant trait. One‐hundred‐and‐forty‐one of the F2‐derived F3 families were classified by their response to the herbicide at 0.24 and 0.48 mg ai L?1 into 39 homozygous susceptible : 72 segregating : 30 homozygous resistant, fitted with a 1:2:1 ratio at χ2 = 1.21 and P = 0.56, indicating that the resistance to fenoxaprop‐p‐ethyl in the sprangletop BLC1 genotype is controlled by a single gene.  相似文献   

11.
Summary. Some characteristics of the interaction of paraquat with a dioctahedral montmorillonite have been studied by a combination of adsorption, radioisotope exchange, and X-ray diffraction techniques. The clay exhibited a high affinity for paraquat, since when less than 50 me/100 g were absorbed no paraquat was detected in the solution (<0·07 ppm), irrespective of the saturating cation, and in the presence or absence of 3·5 M ammonium nitrate solution. Interlamellar adsorption of paraquat was detected by the occurrence, even in an aqueous environment, of the 12·6 A c-spacing associated with the adsorption of pyridinium ions. When 14C-labelled paraquat was added to a clay containing less than 50 me/100 g of adsorbed paraquat, complete self-exchange was observed within 24 hr. Therefore at all levels of adsorption, paraquat held in the adsorbed phase must be in equilibrium with a solution concentration. Interaction entre le paraquat et le sol; adsorption par un mineral argileux gonflant Résumé. Quelques caractères de l'interaction entre le paraquat et une montmorillonite dioctahèrique ont étéétudiés par une combinaison des techniques suivantes: adsorption, échanges de radioisotopes et diffraction des rayons X. L'argile manifesta une grande affinité pour le paraquat: lorsque moins de 50 me/100 g furent absorbés, le paraquat ne fut plus décelé cans la solution (<0.07 ppm), indépendamment du cation saturant, et en présenee ou non d'une solution 3,5 M de nitrate d'ammonium. L'adsorption interlamellaire du paraquat fut détectée par l'occurence, même dans un environnement aqueaux, d'un espacement selon l'axe c de 12,6 A associé avec l'adsorption d'ions pyridinium. Quand du paraquat marqué au 14C fut ajoutéà une argiie contenant moins de 50 me/100 g de paraquat adsorbé, un autoéchange complet fut observé dans les 24 heures. En conséquence, á tous les niveaux d'adsorption, le paraquat retenu dans la phase adsorbée doit être en êquilibre avec une concentration de la solution. Die Wechselwirkung zwischen Paraquat und Boden: Adsorption an ein aufweit-bares Schichtgitter-Tonmineral Zusammenfassung. Mit Hilfe einer Kombination von Adsorptions- und Radio-isotopenaustauschtechniken und Röntgenstrukturanalyse wurden einige Charakteristika der Wechseiwirkungen zwischen Paraquat und einem diocatedrischen Montmorillonit untersucht. Der Ton besass eine hohe Affinität fur Paraquat, da in der Lösung kein Paraquat mehr nachzuweisen war (<0,07 ppm), wenn weniger als 50 mval/100 g adsorbiert worden waren. Dies war sowohl von dem entsprechenden Kation als auch von der Anwesenheit oder Abwesenheit einer 3,5 M Ammoniumnitratlosung unabhangig. Interlamellare Adsorption von Paraquat wurde anhand der sogar in wassrigem Milieu stattfindenden C-Aufweitung (12,6 Å), verbunden mit einer Adsorption von Pyridiniumionen, festgestellt. Wurde 14C-markiertes Paraquat einem Boden, der weniger als 50 mval/100 g adsorbiertes Paraquat enthielt, zugegeben, so trat innerhalb von 24 Stunden vollkommener Selbstaustausch ein. Das adsorbierte Paraquat muss daher bei alien Adsorptionsmengen in einem Gleichgewicht mit einer umgebenden Paraquatlosung stehen.  相似文献   

12.
BACKGROUND: Chlorantraniliprole, a novel anthranilic diamide insecticide, was recently introduced into the United States where rice–crayfish crop rotations are practiced to control rice water weevil (Lissorhoptrus oryzophilus Kuschel) infestations. Chlorantraniliprole has high margins of mammalian safety and excellent insecticidal efficacy, but its toxicity to non‐target crayfish is uncertain. In this study, the acute toxicity of chlorantraniliprole to the red swamp crayfish Procambarus clarkii Girard was determined using aquatic and feeding assays. RESULTS: The aquatic 96 h median lethal toxicity (LC50) data indicate that technical‐grade chlorantraniliprole is highly toxic (US EPA category) to crayfish with an LC50 of 951 µg L?1 (95% CL = 741–1118 µg L?1). A no observed effect concentration (NOEC) of 480 µg L?1 was recorded. Neither the 36 day chronic feeding study, where crayfish fed on chlorantraniliprole‐treated rice seed in aquaria, nor the 144 h acute feeding test, where crayfish fed on rice seeds treated with chlorantraniliprole, produced mortality or abnormal behavior. CONCLUSION: Chlorantraniliprole is three orders of magnitude less acutely toxic to P. clarkii than lambda‐cyhalothrin and etofenprox, two pyrethroid insecticides also used in rice, and is less likely to cause acute crayfish toxicity in rice pond ecosystems. Based on acute toxicity data, the use of chlorantraniliprole should be more compatible with rice–crayfish crop rotations than pyrethroids. Copyright © 2010 Society of Chemical Industry  相似文献   

13.
DISTRIBUTION AND BREAKDOWN OF PARAQUAT IN SOIL   总被引:1,自引:0,他引:1  
Summary. The degradation of paraquat dichloride in soil by Lipomyces starkeyi Lod. and Rij, was investigated spectrophotometrically and by 14C-labelled carbon dioxide emission. Breakdown occurred within the first 96 hr of incubation in cultures of soil of high organic content. Adsorption isotherms showed that soils rich in organic matter had a greater total adsorption of paraquat from solutions of concentration in excess of 250 ppm. Soil high in organic matter was fractionated by ultrasonic disintegration into its organic and inorganic components. Paraquat degradation occurred in cultures containing the former but not the latter. Slow transfer of 14C-labelled paraquat adsorbed on the organic components was shown to take place through a dialysis membrane to the much stronger adsorption sites on the inorganic components placed outside the membrane. It appears that decomposition of paraquat in soil by L. starkeyi occurs while the herbicide is weakly adsorbed on the organic soil components. Over a period of time (96 hr in one instance), slow transfer to the inorganic soil components rendered the paraquat unavailable for microbial degradation. Dispersion et decomposition du paraquat dans le sol Résumé. La décomposition du paraquat-dichlorure dans le sol par Lipomyces starkeyi Lod. ci Rij a étéétudiée par spectrophotométrie et par l'émission de gaz carbonique marqué au 14C. La décomposition survint durant les premières 96 heures d’incubation dans les cultures de sols à forte teneur en matière organique. Les isothermes d'adsorption montrè-rent que les sols riches en matière organique présentaient une d'adsorption totale du paraquat plus grande à partir de solutions dont la concentration dépassait 250 ppm. Les sols à forte teneur en matière organique furent fractionnés par désintégration aux ultra-sons en leurs composants organiques et minéraux. La dégradation du paraquat se manifesta dans les cultures contenant les premiers, mais pas dans celies contentant les seconds. Une lente migration du paraquat marqué au 14C adsorbé sur les composants organiques fut observée par dialyse à travers une membrane vers les lieux de plus forte adsorption dans les composants minéraux placés de l'autre côté de la membrane. II apparait que la décomposition du paraquat dans le sol par L. starkeyi se manifeste lorsque l'herbicide est faiblement adsorbé sur les composants organiques du sol. Au-delà d'une certaine durée (96 heures dans un cas) la lente migration vers les composants minéraux du sol mit le paraquat à l'abri de la dégradation microbienne. Verteilung und Abbau von Paraquat im Boden Zusammenfassung. Der Abbau von Paraquat-chlorid im Boden durch Lipomyces starkeyi Lod, und Rij wurde sowohl spektrophotometrisch als auch durch Messung der 14C-Kohlendioxydabgabe untersucht, Der Abbau erfolgte innerhalb der ersten 96 Stunden nach dcr Inkubation in Bodenkulturen mit hohem organischem Anteil. Die Adsorptions-isotherm en zeigten, dass Boden mit hohem Anteil an organischer Substanz mehr Paraquat aus einer Lösung mit einer Konzentration von mehr als 250 ppm adsorbierten. Boden mit hohem Anteil an organischer Substanz wurde durch Ultraschalldesintegration in seine organischen und anorganischen Bestandteilv zcrlegt. Abbau von Paraquat erfolgte in Kulturcn mit den organischen Bestand teil en, nicht jedoch in Kulturen mit anorganischen. Durch eine Dialysemembran erfolgte ein langsamer Ubertritt von 14C-markiertem Paraquat, das an die organischen Bestandteile adsorbiert war, an die viel starker adsorbierenden anorganischen Bestaudteile, die sich ausserhatb der Membran befanden. Bei schwacher Adsorption des Herbizids an organische Bod en best and teile erfolgt offensichtlich ein Abbau von Paraquat durch L. starkeyi. Mit der Zeit (in einem Fall innerhalb von 96 Stunden) wird Paraquat durch einen langsamen Ubertritt an die anorganischen Bodenbestandteile fiir Mikroorganismen unzuganglich.  相似文献   

14.
BACKGROUND: Twenty‐three isolates of Metarhizium anisopliae (Metschnikof) Sorokin and M. acridum (Driver & Milner) JF Bischoff, Rehner & Humber from non‐aphid host insects around the globe were evaluated for their aphid biocontrol potential, which is not well known. RESULTS: The apterous adults of green peach aphid Myzus persicae (Sulzer) were exposed to the fungal sprays of 11.5, 99 and 1179 conidia mm?2 and blank control in three leaf‐dish bioassays. All the tested isolates except one were proven to be infective to the aphid species at 21 ± 1 °C and 14:10 h light:dark photoperiod, causing corrected mortalities of 10.1–95.3% at the high spore concentration. The data from ten isolates causing > 50% mortality at the high concentration were found to fit a time–concentration–mortality model well, yielding parameters for the estimates of their LC50 and LT50 that vary with post‐spray time and spore concentration respectively. Four isolates of M. anisopliae (ARSEF 759, 4132, 2080 and 576) had LC50 values of 44–80 conidia mm?2 on day 8 and LT50 values of 4.9–6.8 days at 100 conidia mm?2, with 91–98% of the killed aphids being well mycotised after death. CONCLUSION: The Metarhizium infectivity to M. persicae differs greatly among the tested isolates. The four mentioned isolates with desired virulence and sporulation potential are excellent candidates for microbial control of aphids. Copyright © 2010 Society of Chemical Industry  相似文献   

15.
The rate of photosynthesis in intact leaves attached to plants of an atrazine-resistant cultivar of Brassica napus L. was compared with that in leaves of an atrazine-susceptible, rather isogenic cultivar, in a number of irradiance series. The resistant cultivar had been obtained from the susceptible cultivar by crossing the resistant biotype of Brassica campestris L. with the susceptible cultivar of B. napus, and using the fourth back-cross for comparison. At a natural CO2 concentration (360 cm3 m?3) the rate of photosynthesis of leaves of the atrazine-resistant cultivar was significantly (17–21%) less than that of leaves of the susceptible cultivar, over the whole range of irradiances up to 250 W m?2 At the enriched CO2 concentration of 1500 cm?3 m?3 this reduction in rate amounted to 9–12%.  相似文献   

16.
BACKGROUND: The parasitoid Trichogramma ostriniae (Pang and Chen) is a major natural enemy of many lepidopterans, but only a few studies have been conducted on the compatibility of biological and chemical controls. In this study, the selectivity of 30 insecticides to T. ostriniae adults was evaluated. RESULTS: Among the seven classes of chemicals tested, organophosphates and carbamates had the highest intrinsic toxicity to the wasp, with LC50 values ranging from 0.032 (0.029–0.038) to 2.38 (1.91–3.15) mg AI L?1. They are followed by phenylpyrazoles, avermectins, neonicotinoids and pyrethroids, which induce variable toxicity responses, with LC50 values ranging from 0.14 (0.11–0.21) to 56.67 (48.94–67.24) mg AI L?1, from 2.57 (1.85–4.28) to 4.48 (3.34–6.83) mg AI L?1, from 2.48 (1.80–4.03) to 503.6 (457.6–557.5) mg AI L?1 and from 5.44 (3.95–8.84) to 104.2 (92.48–119.7) mg AI L?1 respectively. The insect growth regulators (IGRs) exhibited least toxicity to the parasitoid. CONCLUSION: Risk quotient analysis classifies neonicotinoids, avermectins, pyrethroids, IGRs and phenylpyrazoles (with the exception of butane‐fipronil and fipronil) as safe agents to the parasitoid, but categorises organophosphates and carbamates as slightly to moderately toxic or dangerous to T. ostriniae. Copyright © 2012 Society of Chemical Industry  相似文献   

17.
Ephestia elutella larvae in diapause were exposed at 25°C to methyl bromide at 12 mg litre?1 for 3.5 or 7.5 h and then immediately exposed to a lower concentration. The minimum effective concentration (that at which Haber's rule, concentration × time = k, a constant for mortality, still applied) was about 3 mg litre?1 in tests with no previous exposure toa high concentration, but it was about 2.5 mg litre?1 for individuals surviving a 3.5 h exposure to 12 mg litre?1, and was about 1.6 mg litre?1 for those surviving a 7.5 h exposure to 12 mg litre?1. These exposures to 12 mg litre?1, respectively, killed 2–20% and 50–75% of larvae exposed, and hence the smaller the proportion of survivors of exposure to a high concentration, the lower the minimum effective concentration needed against them. Thus the low concentration persisting at the end of a practical fumigation should contribute significantly to the success of the treatment and be much more effective than any similar low concentration present soon after the introduction of gas.  相似文献   

18.
The adsorption of carbofuran on soils from water‐methanol mixtures has been evaluated by batch shake testing. Two uncontaminated soils having different physicochemical properties were used in these experiments. The volume fraction of methanol in the liquid phase (fs) was varied from 0.25 to 1.0. Higher adsorption of carbofuran was observed in medium black (silt loam) soil than in alluvial (sandy loam) soil; calculated values of the Freundlich constant (Km) and distribution coefficient (Kd) showed that adsorption of carbofuran in both soils decreased with increase in fS values. The decreased carbofuran adsorption in methanol–water mixtures meant a greater potential of ground‐water contamination through leaching from potential sites. The data have been used to evaluate the co‐solvent theory for describing adsorption of carbofuran in methanol–water mixtures. The aqueous phase partition coefficient Kdw (mol g−1) normalized with respect to foc and the aqueous phase adsorption constant Kw for carbofuran were evaluated by extrapolating to fS = 0. © 2000 Society of Chemical Industry  相似文献   

19.
BACKGROUND: This study was initiated to search for fungal candidates for microbial control of brown planthopper (BPH) Nilaparvata lugens Stål, to which little attention has been paid in the past two decades. RESULTS: Thirty‐five isolates of Metarhizium anisopliae (Metschnikoff) Sorokin and M. flavoviride Gams & Rozsypal from different host insects worldwide were bioassayed for their lethal effects against third‐instar BPH nymphs at 25 °C and a 14:10 h light:dark photoperiod at ca 1000 conidia mm?2. On day 9 post‐treatment, mortality attributable to mycosis ranged from 6.5 to 64.2% and differed significantly among the tested isolates with no apparent relationship to their host origin. Only two BPH‐derived M. anisopliae isolates from the Philippines (ARSEF456) and Indonesia (ARSEF576) killed > 50% of the nymphs. Both isolates were further bioassayed for time–concentration–mortality responses of the nymphs to the sprays of 19–29, 118–164 and 978–1088 conidia mm?2 in repeated bioassays. The resultant data fitted a time–concentration–mortality model very well. Their LC50 values were estimated as 731 and 1124 conidia mm?2 on day 7 and fell to 284 and 306 conidia mm?2, respectively, on day 10. CONCLUSION: The two M. anisopliae isolates are potential biocontrol agents of BPH for further research. This is the first report of the lethal effects of global Metarhizium isolates on the rice pest. Copyright © 2008 Society of Chemical Industry  相似文献   

20.
HALDOR FYKSE 《Weed Research》1976,16(5):309-316
Research on Sonchus arvensis L. III. Metabolism of MCPA Roots of Sonchus arvenis L. were injected with 14C-MCPA at the time of planting of the root sections and when the plants had leaves 3 cm, 5–7 cm and 12–15 cm long. After extraction with 70% ethanol and separation by thin layer chromatography, three components, termed 1, 2 and 3, were detected by scanning. Component 2 gave a chromatograph identical to MCPA and was probably the non-metabolized residue of the injected material. Component 3 appeared faster than component 1 but, when heated with 1 N HCl or 1 N NaOH, both of them yielded component 2 (MCPA). In this latter reaction the conversion of component 3 was again the fastest. The older the plants the more rapid was the metabolism of MCPA, the highest metabolic rate occurring in the main roots. In secondary roots and leaves this process occurred more slowly. Dormant roots were capable of metabolizing MCPA quite as well as non-dormant roots. Raising the temperature from 7° to 23°C accelerated the metabolic process. The liberation of 14C02 proceeded very slowly. During a period of 4 days only 1.3% of the total radio-activity was released as 14CO2. In biotests, with Raphanus sativus L. as test species, component 3 appeared to be as phytotoxic as component 2 (MCPA), whereas the more stable component 1 exhibited very low phytotoxicity.  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