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 187 毫秒
1.
设计合成了一系列结构新颖的嘧啶联吡唑甲酰胺类化合物5a~5o,其结构均经过1H NM R和MS分析确证。初步生物活性测试结果表明:在有效成分150 g/hm2剂量下苗后茎叶喷雾处理时,化合物(R)-N-[1-(4-氯苯基)乙基]-3-二氟甲基-1-(4,6-二甲氧基嘧啶-2-基)-1H-吡唑-4-甲酰胺(5c)、N-[1-(4-氯苯基)乙基]-1-(4,6-二甲氧基嘧啶-2-基)-N-甲基-3-三氟甲基-1H-吡唑-4-甲酰胺(5i)和N-[1-(4-氯苯基)乙基]-1-(4,6-二甲氧基嘧啶-2-基)-3-三氟甲基-1H-吡唑-4-甲酰胺(5k)对繁缕Stellaria media的抑制率高达90%以上;而同样剂量下苗前土壤喷雾处理时,化合物N-[1-(4-氯苯基)乙基]-3-二氟甲基-1-(4,6-二甲氧基嘧啶-2-基)-1H-吡唑-4-甲酰胺(5b)和5c对繁缕的抑制率达100%。该类结构化合物有望作为除草先导化合物进行开发。  相似文献   

2.
Seventy three isolates of Pythium aphanidermatum obtained from cucumber from four different regions of Oman and 16 isolates of muskmelon from the Batinah region in Oman were characterized for aggressiveness, sensitivity to metalaxyl and genetic diversity using AFLP fingerprinting. Twenty isolates of P. aphanidermatum from diverse hosts from different countries were also included in the study. Most isolates from Oman were found to be aggressive on cucumber seedlings and all were highly sensitive to metalaxyl (EC50 < 0·80 µg mL?1). Isolates from cucumber and muskmelon were as aggressive as each other on both hosts (P > 0·05), which implies a lack of host specialization in P. aphanidermatum on these two hosts in Oman. AFLP analysis of all isolates using four primer–pair combinations resolved 152 bands, of which 61 (~40%) were polymorphic. Isolates of P. aphanidermatum from Oman and other countries exhibited high genetic similarity (mean = 94·1%) and produced 59 different AFLP profiles. Analysis of molecular variance indicated that most AFLP variation among populations of P. aphanidermatum in Oman was associated with geographical regions (FST = 0·118; P < 0·0001), not hosts (FST = –0·004; P = 0·4323). These data were supported by the high rate of recovery (24%) of identical phenotypes between cucumber and muskmelon fields in the same region as compared to the low recovery (10%) across regions in Oman, which suggests more frequent movement of Pythium inoculum among muskmelon and cucumber fields in the same region compared to movement across geographically separated regions. However, recovering clones among regions and different countries may imply circulation of Pythium inoculum via common sources in Oman and also intercontinental spread of isolates.  相似文献   

3.
Prochloraz (N-propyl-N-[2-(2,4,6-trichlorophenoxy)ethyl]-imidazole-1-carboxamide), a recently developed agricultural fungicide, is a potent inducer of microsomal enzymes. Rats fed 7 days with a prochloraz-contaminated diet (2500 ppm) showed an increase in hepatic cytochrome P-450, cytochrome b5, and microsomal protein level; aniline hydroxylase, 7-ethoxycoumarin dealkylase, 7-ethoxyresorufin dealkylase, NADPH-cytochrome c reductase, and epoxide hydrolase were significantly induced. At a lower dose (100 ppm), only an increase in cytochrome P-450 and 7-ethoxyresorufin dealkylase was noticed. As shown with aniline hydroxylase and 7-ethoxycoumarin dealkylase, prochloraz is also a potent inhibitor of drug-metabolizing enzymes. The interaction of prochloraz with hepatic microsomal fraction from rat liver was also studied. Prochloraz binds to oxidized cytochrome P-450 to produce a type II spectral change; the compound also binds to reduced cytochrome P-450. The binding of some ligands (7-ethoxycoumarin, n-octylamine, aniline, and imidazole) to oxidized cytochrome P-450 was determined after induction by prochloraz. Japanese quails (Coturnix coturnix) fed 7 days with a prochloraz-contaminated diet (2000 ppm) showed a dramatic increase in liver weight (2.5-fold) and both hepatic and duodenal cytochrome P-450 (9- and 12-fold, respectively).  相似文献   

4.
Acetamiprid, (NI-25; (E)-N1-[(6-chloro-3-pyridyl)methyl]-N2-cyano-N1-methylacetamidine), is a novel neonicotinoid insecticide having an N-cyanoacetamidine structure as its characteristic feature. The [1H] and [13C]-NMR spectra indicated the existence of two different structures in acetamiprid at room temperature in solution. The measurement of CH-NOE and C–C coupling constants proved the E-configuration at the cyanoimino group in both existing structures. The [13C] chemical shifts of N-CH3 and the long range C–H coupling in the formamidine analogue of acetamiprid suggested that there exist two conformers generated by the rotation of C–C single bond in the amidine moiety. Dynamic NMR spectra of acetamiprid and the computer simulation of the two-site exchange demonstrated that the two conformers change slowly to each other at room temperature. The conformational analysis by semi-empirical molecular orbital calculations using MNDO-PM3 method predicted four conformers as energy-minimum structures, among which two E-conformers were more stable than Z-conformers. One of the E-conformers in which two methyl groups are in cis configuration was superimposable onto the structure of imidacloprid, which is a known neonicotinoid insecticide having more rigid structure. This E-conformer was assumed as the active conformation of acetamiprid on the basis of the molecular similarity in terms of steric and electrostatic properties. © 1997 SCI  相似文献   

5.
The mechanisms of resistance to pyrethroids were studied in a permethrin-selected (147-R) strain of the house fly, Musca domestica L. Approximately 12-fold synergism was obtained with a mixture of (1R)-trans-permethrin:piperonyl butoxide (1:5) so that the resistance decreased from 97-fold to 22-fold. Tests with the esterase inhibitor S,S,S-tributyl phosphorotrithioate produced very little synergism in either the resistant (R) strain (1.6-fold) or the susceptible (S) strain (1.9-fold). An investigation of the microsomal components revealed that compared to the S strain, the R strain demonstrated twice as much cytochrome P-450 and cytochrome b5 and double the rate of NADPH-cytochrome c reductase activity. In addition, the rate of p-nitroanisole O-demethylation was found to be six times greater in the R strain. An in vivo accumulation study showed that the R strain displayed a decreased rate of penetration of trans-[14C]permethrin. When treated at equitoxic doses the R strain was found to tolerate 50-fold more internal permethrin than the S strain. An in vitro metabolism study indicated that there was no difference between strains in the overall rate of metabolism of trans-[14C]permethrin. The evidence obtained supports the conclusion that several resistance factors are involved but that decreased sensitivity of the nervous system to the action of pyrethroids is the principal mechanism of resistance in the 147-R strain.  相似文献   

6.
Cytoplasmic uptake of carbendazim (methyl benzimidazol-2-yl carbamate; MBC) from an aqueous solution was demonstrated with isolated mesophyll cells. About 2.5% of the labelled MBC (ring-2-[14C]) in the treatment solution (1.85 μg/ml) was taken up in 44 h. When cotyledons of cucumber seedling were treated with either 347 or 36 μg [14C]-MBC/plant 1.11 and 0.13% were extracted, respectively, from the remainder of the plant, 5 days after treatment. Greatest amounts were detected in shoot apices. Likewise, when MBC and benomyl were applied at the dose of 2 μmol, 0.34 and 0.57% were detected in the untreated part of the plant with a bioassay procedure. Foliar application with 347 or 36 μg[14C]-MBC/leaf resulted in the translocation of 1.68 and 0.11% out of the treated area. By scalding the living cells of the petiole translocation was prevented suggesting that long distance movement occurred in the symplast. During a period of 14 days 1.56% of [14C]-MBC applied to cucumber leaves was metabolised and respired as CO2. This degradation was assumed to occur enzymically within the symplast.  相似文献   

7.
A laboratory study was conducted to determine the degradation rates and identify major metabolites of the herbicide metsulfuron-methyl in sterile and non-sterile aerobic soils in the dark at 20°C. Both [phenyl-U-14C]- and [triazine-2-14C]metsulfuron-methyl were used. The soil was treated with [14C]metsulfuron-methyl (0.1 mg kg−1) and incubated in flow-through systems for one year. The degradation rate constants, DT50, and DT90 were obtained based on the first-order and biphasic models. The DT50 (time required for 50% of applied chemical to degrade) for metsulfuron-methyl, estimated using a biphasic model, was approximately 10 days (9–11 days, 95% confidence limits) in the non-sterile soil and 20 days (12–32 days, 95% confidence limits) in the sterile soil. One-year cumulative carbon dioxide accounted for approximately 48% and 23% of the applied radioactivity in the [phenyl-U-14C] and [triazine-2-14C]metsulfuron-methyl systems, respectively. Seven metabolites were identified by HPLC or LC/MS with synthetic standards. The degradation pathways included O-demethylation, cleavage of the sulfonylurea bridge, and triazine ring opening. The triazine ring-opened products were methyl 2-[[[[[[[(acetylamino)carbohyl]amino]carbonyl]amino] carbonyl]-amino]sulfonyl]benzoate in the sterile soil and methyl 2-[[[[[amino[(aminocarbonyl)imino]methyl] amino]carbonyl]amino]sulfonyl]benzoate in the non-sterile soil, indicating that different pathways were operable. © 1999 Society of Chemical Industry  相似文献   

8.
In hydroponic experiments, seed-dressing with the herbicide safener 1,8-naphthalic anhydride (NA), significantly enhanced the tolerance of maize, (Zea mays L., cv. Monarque) to the imidazolinone herbicide, AC 263222, (2-[4-isopropyl-4-methyl-5-oxo-2-imidazolin-2-yl]-5-methylnicotinic acid). Uptake, distribution and metabolism studies where [14C]AC 263222 was applied through the roots of hydroponically grown maize plants showed that NA treatment reduced the translocation of radiolabel from root to shoot tissue and accelerated the degradation of this herbicide to a hydroxylated metabolite. Reductions in the lipophilicity and, therefore, mobility of this compound following hydroxylation may account for NA-induced retention of radiolabel in the root system. Hydroxylation of AC 263222 suggested that NA may stimulate the activity of enzymes involved in oxidative herbicide metabolism, such as the cytochrome P450 mono-oxygenases. In agreement with this theory, the cytochrome P450 inhibitor, 1-aminobenzotriazole (ABT), synergized AC 263222 activity and inhibited its hyroxylation in vivo. NA seed-dressing enhanced the total cytochrome P450 and b5 content of microsomes prepared from etiolated maize shoots. Isolated microsomes catalyzed AC 263222 hydroxylation in vitro. This activity possessed the characteristics of a cytochrome P450 mono-oxygenase, being NADPH-dependent and susceptible to inhibition by ABT. Activity was stimulated four-fold following NA seed treatment. Differential NA enhancement of AC 263222 hydroxylase and the cytochrome P450-dependent cinnamic acid-4-hydroxylase (CA4H) activity, suggested that separate P450 isozymes were responsible for each activity. These results indicate that the protective effects of NA result from enhancement of AC 263222 hydroxylation and concomitant reduction in herbicide translocation. This may be attributed to the stimulation of a microsomal cytochrome P450 system. © 1998 SCI.  相似文献   

9.
The distribution resulting from the drenching of soil with a suspension concentrate of [14C]metazoxolon was studied in the laboratory and the field. Penetration of soil columns was increased by (a) increasing the drench volume from 1 to 7.8 litres m?2, (b) changing the original soil moisture content from air-dry to field-capacity, and (c) including 1 % of ‘Renex 30’ surfactant in the drench. Penetration was greatest in soils containing large pores and was reduced when aggregates were broken down by sieving. Leaching the column with 1.56 cm of ‘rain’, 15 h after treatment, did not increase the penetration by metazoxolon. In all experiments, the maximum concentration of metazoxolon occurred in the top 2 cm of soil. Equivalent effects were found when metazoxolon was applied to a poorly-structured sandy clay loam in the field.  相似文献   

10.
BACKGROUND: Linuron is a globally used phenylurea herbicide, and a large number of studies have been made on the microbial degradation of the herbicide. However, to date, the few bacteria able individually to mineralise linuron have been isolated only from European agricultural soils. An attempt was made to isolate linuron‐mineralising bacteria from Japanese river sediment using a uniquely designed river ecosystem model (microcosm) treated with 14C‐ring‐labelled linuron (approximately 1 mg L?1). RESULTS: A linuron‐mineralising bacterium that inhabits river sediment was successfully isolated. The isolate belongs to the genera Variovorax and was designated as strain RA8. Strain RA8 gradually used linuron in basal salt medium (5.2 mg L?1) with slight growth. In 15 days, approximately 25% of 14C‐linuron was mineralised to 14CO2, with 3,4‐dichloroaniline as an intermediate. Conversely, in 100‐fold diluted R2A broth, strain RA8 rapidly mineralised 14C‐linuron (5.5 mg L?1) and more than 70% of the applied radioactivity was released as 14CO2 within 3 days, and a trace amount of 3,4‐dichloroaniline was detected. Additionally, the isolate also degraded monolinuron, metobromuron and chlorobromuron, but not diuron, monuron or isoproturon. CONCLUSION: Although strain RA8 can grow on linuron, some elements in the R2A broth seemed significantly to stimulate its growth and ability to degrade. The isolate strictly recognised the structural difference between N‐methoxy‐N‐methyl and N,N‐dimethyl substitution of various phenylurea herbicides. Copyright © 2010 Society of Chemical Industry  相似文献   

11.
A new class of agricultural fungicides derived from strobilurins act as respiration inhibitors by binding to mitochondrial cytochrome b. The effects of the strobilurin, kresoxim-methyl, on conidia germination, mycelial growth and the protection of apple leaves from scab development were investigated for two isolates of Venturia inaequalis randomly selected from a culture collection. Inhibition of mycelial growth required relatively high doses of kresoxim-methyl (ED50=1 μg ml-1) for both isolates. In comparison, germination of conidia was highly sensitive for one of the isolates (ED50=0·005 μg ml-1), while the level of inhibition achieved for the second isolate was 60-fold less (ED50=0·3 μg ml-1). As deduced from identical sequences of cytochrome b cDNAs prepared from both isolates, the different responses of germinating conidia to kresoxim-methyl were not caused by differences in the sequence of cytochrome b as the target site for strobilurin action. Strong synergistic effects of salicylhydroxamic acid on kresoxim-methyl inhibitory potency suggested that the reduced kresoxim-methyl sensitivity observed for germinating conidia was caused by interference of the alternative respiratory pathway with inhibitor action. However, this mechanism of target site circumvention in germinating conidia had no adverse effects on the protection of apple leaves from scab infection by kresoxim-methyl. © 1998 Society of Chemical Industry  相似文献   

12.
BACKGROUND: In 2004, field isolates of Botrytis cinerea Pers. ex Fr., resistant to strobilurin fungicides (QoIs), were first found in commercial citrus orchards in Wakayama Prefecture, Japan. Subsequently, QoI‐resistant isolates of this fungus were also detected in plastic strawberry greenhouses in Saga, Ibaraki and Chiba prefectures, Japan. Biological and molecular characterisation of resistant isolates was conducted in this study. RESULTS: QoI‐resistant isolates of B. cinerea grew well on PDA plates containing kresoxim‐methyl or azoxystrobin at 1 mg L?1, supplemented with 1 mM of n‐propyl gallate, an inhibitor of alternative oxidase, whereas the growth of sensitive isolates was strongly suppressed. Results from this in vitro test were in good agreement with those of fungus inoculation tests in vivo. In resistant isolates, the mutation at amino acid position 143 of the cytochrome b gene, known to be the cause of high QoI resistance in various fungal pathogens, was found, but only occasionally. The heteroplasmy of cytochrome b gene was confirmed, and the wild‐type sequence often present in the majority of resistant isolates, indicating that the proportion of mutated cytochrome b gene was very low. CONCLUSION: The conventional RFLP and sequence analyses of PCR‐amplified cytochrome b gene are insufficient for molecular identification of QoI resistance in B. cinerea. Copyright © 2009 Society of Chemical Industry  相似文献   

13.
The effects of the herbicide isouron and of its plant degradation products designated as metabolite l {N-[5-(1,1-dimethylethyl)-3-isoxazolyl]-N-methylurea} and metabolite 2 {N-[5-(1,1-dimethylethyl)-3-isoxazolyl]-urea} on the metabolism of enzymatically isolated leaf cells of soybean [Glycine max (L.) Merr., cv. Essex] were compared under laboratory conditions. Photosynthesis, protein synthesis, ribonucleic acid synthesis, and lipid synthesis were assayed by the incorporation of NaH14CO3, [14C]-leucine, [14C]-uracil, and [14C]-acetate, respectively, into the isolated cells. Time-course and concentration studies included incubation periods of 30, 60, and 120 min and concentrations of 0.1, 1, 10 and 100 μM of the three herbicides. The urea derivative of isouron (metabolite 2) was the least active of the three compounds. The activity of the mono-methylated derivative of isouron (metabolite 1) was comparable to that of isouron and the sensitivity of the four processes to both chemicals decreased in the order: photosynthesis > ribonucleic acid synthesis > lipid synthesis > protein synthesis. The concentration of isouron that caused a 50% inhibition of photosynthesis of the isolated soybean leaf cells was calculated at 0.51 μM. The effects of isouron and metabolite 1 on photosynthesis, lipid and RNA synthesis appeared to be independent of incubation lime as maximal inhibition occurred within 30 min. Inhibition of protein synthesis by both chemicals was time-dependent, increasing in magnitude with concomitant increases in incubation time.  相似文献   

14.
The fluorescent insect growth regulator 5[[[5-(dimethylamino)-1-naphthalenyl]amino]-1,3-benzodioxole (DNSAB) forms a metabolite complex with house-fly microsomal cytochrome P-450. Formation of the metabolite complex is dependent on the presence of NADPH and O2; NADH supports the reaction at a reduced rate. The presence of antibodies to house-fly cytochrome c (P-450) reductase in reaction mixtures inhibits the complex formation, indicating that the reductase is necessary for transfer of electrons from NADPH to cytochrome P-450 to complete the reaction. In the oxidized form, the metabolite complex has a single absorbance maximum at 431 nm, whereas the reduced form has two absorbance maxima at 426 (major) and 455 nm (minor). The pH of the media affects the extinction of the 426- and 455-nm Soret bands; increased pH decreases the extinction of the 426-nm band and increases the extinction of 455-nm band. Formation of the DNSAB metabolite-cytochrome P-450 complex decreases the amount of CO-reactive cytochrome P-450 by 24%. The metabolite complex is not dissociable by treatment with ferricyanide or by using centrifugation techniques. Dissociation is accomplished by addition of DNSAB to the oxidized metabolite complex. Kinetic analysis of the complex formation gives apparent Km and Vmax values at 2.55 ± 1.0 μM and 1.1 ± 0.4 × 10?2 ΔA min?1 nmol?1 cytochrome P-450, respectively. Addition of juvenile hormone [(E,E)-cis-methyl-10,11-epoxy-7-ethyl-3,11-dimethyl-2,6-tridecadienoate; JH] to the reaction medium competitively inhibits the formation of the metabolite complex giving an inhibition constant of 16 μM. DNSAB synergized the lethal effects of JH against Aedes aegypti larvae threefold; however, JH did not synergize DNSAB. These data suggest that DNSAB may acquire its hormonal qualities by complexing a species of cytochrome P-450 that metabolizes JH, thereby prolonging the in vivo lifetime of this hormone.  相似文献   

15.
The in vivo release of 14CO2 arising from decarbamoylation of [14C]-carbaryl (1-naphthyl methyl-[14C]-carbamate) injected into male and female Periplaneta americana, Leucophaea maderae, Gromphadorhina portentosa and Schistocerca gregaria was measured up to 50 nmol carbaryl/g body weight. The amount of 14CO2 released was proportional to the dose of [14C]-carbaryl injected and was similar for both sexes of each species. The KD50 values for carbaryl for each species did not correlate with their ability to decarbamoylate carbaryl. The degradation of carbaryl by this pathway is not therefore the limiting factor in determining the overall toxic effect of carbaryl in the four species studied. Decarbamoylation is inhibited by nitrogen in all four species. In P. americana the inhibition is virtually complete, in L. maderae, G. portentosa and S. gregaria the inhibition is 84%, 74% and 51% respectively. The results suggest the involvement of two enzymes, one of which is oxygen dependent. Decarbamoylation studies in air suggest the involvement of two enzymes with differing KM values. The involvement of aliesterases and mixed function oxidases in the decarbamoylation pathway(s) have not been established so far with this series of experiments.  相似文献   

16.
A rat, given a single oral dose of [14C] cymoxanil, 1-(2-cyano-2-methoxyimino-[2-14C]-acetyl)-3-ethylurea, eliminated 91% of the radioactivity within 72 h. The urine contained 71%, the faeces 11%, and the expired air about 7% of the radiolabel; no 14C residue was found in the internal organs. Greater than 70% of the radioactivity in the urine was identified. The major metabolite was characterised as glycine, both free and conjugated, as hippuric acid and phenylaceturic acid [N-(phenylacetyl)-glycine], and probably in the form of polypeptides of low molecular weight. The other metabolites identified included 2-cyano-2-methoxyiminoacetic acid, 2-cyano-2-hydroxyiminoacetic acid and 1-ethylimidazolidine-2, 4, 5-trione. The minor metabolites included succinic acid and 2-oxoglutaric acid which indicated reincorporation of metabolic 14C. Cymoxanil, as such, was not detected in the urine.  相似文献   

17.
The capacity of several strains of root-colonizing bacteria to suppressPythium aphanidermatum, Pythium dissotocum and root rot was investigated in chrysanthemums grown in single-plant hydroponic units containing an aerated nutrient solution. The strains were applied in the nutrient solution at a final density of 104 CFU ml−1 and 14 days later the root systems were inoculated withPythium by immersion in suspensions of 104 zoospores ml−1 solution. Controls received no bacteria, noPythium, or one of thePythium spp. but no bacteria. Strain effectiveness was estimated based on percent roots colonized byPythium and area under disease progress curves (AUDPC). In plants treated respectively withPseudomonas (Ps.)chlororaphis 63-28 andBacillus cereus HY06 and inoculated withP. aphanidermatum, root colonization by the pathogen was 83% and 72% lower than in the pathogen control, and AUDPC values were reduced by 61% and 65%. ForP. dissotocum, the respective strains reduced root colonization by 87% and 91%, and AUDPC values by 70% and 90%. In plants treated respectively withPs. chlororaphis Tx-1 andComamonas acidovorans C-4-7-28, root colonization byP. aphanidermatum was 84% and 80% lower than in the controls and AUDPC values were reduced by 66% and 57%; these strains did not suppressP. dissotocum. Burkholderia gladioli C-2-74 andC. acidovorans OCR-7-8-38, respectively, suppressed colonization of roots byP. dissotocum by 74% and 86%, and reduced AUDPC values by 60% and 70%, but were ineffective againstP. aphanidermatum. C. acidovorans OCR-7-8-39 reduced colonization and AUDPC values ofP. aphanidermatum by 57% and 42%, respectively.Pseudomonas corrugata 13,Ps. fluorescens 15 and JZ12, and three additional strains ofC. acidovorans were weakly or nonsuppressive againstP. aphanidermatum. Strains that reduced AUDPC values forP. aphanidermatum orP. dissotocum when applied at 104 CFU ml−1 were 11%–39% less effective at 103 CFU ml−1. Four tested strains (Ps. chlororaphis 63-28,Ps. chlororaphis Tx-1,B. cereus HY06, andB. gladioli C-7-24) in most instances suppressed root colonization and lowered AUDPC values ofP. aphanidermatum when applied at 14, 7 or 0 days before inoculation, but reduction of the respective variables was generally greater when the strains were applied at 14 days (63%–87% and 75%–78%) or 7 days (44%–47% and 31%–88%) than at 0 days (14%–31% and 23%–62%) before inoculation.Ps. chlororaphis Tx-1,Ps. chlororaphis 63-28 andB. cereus HY06 significantly suppressedP. aphanidermatum whether the temperature of the nutrient solution was high (32°C) or moderate (24°C). Taken together, the observations suggest thatPs. chlororaphis 63-28,B. cereus HY06,Ps. chlororaphis Tx-1,B. gladioli C-2-74 andC. acidovorans OCR-7-8-38 have the potential for controlling Pythium root rot in hydroponic chrysanthemums. http://www.phytoparasitica.org posting Jan. 24, 2007.  相似文献   

18.
Seedlings of Solanum scabrum Mill. and Solanum ptycanthum Dun. were treated with [14C]ethalfluralin (N-ethyl-α,α,α-trifluoro-N-(methylallyl)-2,6-dinitro-p-toluidine) and [14C]trifluralin (α,α,α-trifluoro-2,6-dinitro-N,N-dipropyl-p-toluidine) supplied in nutrient solution to determine the basis for differences in response by these two species to these two herbicides. Plants of S. scabrum absorbed more [14C]ethalfluralin and [14C]trifluralin than plants of S. ptycanthum. During the first 24 h, S. scabrum seedlings, but not S. ptycanthum seedlings absorbed more [14C]ethalfluralin than did plants treated with [14C]trifluralin. More [14C]ethalfluralin than [14C]trifluralin was found in the shoots of plants of both species. Seventy-two hours after treatment with [14C]herbicides, the conversion to water-soluble metabolites was greater for [14C]ethalfluralin than for [14C]trifluralin. In the shoots of plants from both species an average of nearly 55% of the 14C recovered was found in the water-soluble fraction following [14C]ethalfluralin treatment whereas an average of only 40% was found in the water-soluble fraction following [14C]trifluralin treatment.  相似文献   

19.
When the petioles of detached tobacco leaves (10–17 cm2) were incubated in aqueous solutions containing [14C]metalaxyl, uptake of the fungicide was dependent on the temperature and photoperiod. Detached leaves took up 78% more [14C]metalaxyl at 26°C than at 16°C. The rate of uptake in the light at 21°C was linear, but after an additional 20h in the dark, there was only twice as much fungicide in the leaves. Different sized leaves contained the same amount of fungicide per cm2 area. Uptake by detached leaves of the 14C-labelled anilide lactones ofurace and RE-26940 [2-methoxy-N-(tetrahydro-2-oxo-3-thienyl)acet-2′,6′-xylidide] was similar to that of metalaxyl. At the concentration of metalaxyl (66 ng ml?1) that controlled blue mould (Peronospora tabacina) on detached tobacco leaves, the amount of fungicide in the leaves was found to be 7.25 ng. Autoradiography showed that the distribution of [14C]metalaxyl in detached leaves after incubation for 23h was uniform, although higher concentrations of the label were present in the smaller veins of the leaves.  相似文献   

20.
The photodegradation of fenpropathrin [(RS)-α-cyano-3-phenoxybenzyl 2,2,3,3-tetramethylcyclopropanecarboxylate] ( I ), in water, on soil and on plant foliage, was investigated using 14C-preparations labelled separately at the cyano group, cyclopropyl-C1 or in the benzyl ring. On exposure to sunlight, I was photodecomposed with initial half-lives of >6 weeks in distilled water, 6.0 weeks in humic acid aqueous solution, 2.7 weeks in river water, 1.6 weeks in sea water and 0.5 of a day in 2% aqueous acetone. A triplet photosensitiser, acetone, together with naturally occurring substances in river and sea water, including humic acid, enhanced the photodegradation of I . On three kinds of soil, I was rapidly photodegraded with initial half-lives of 1–5 days, whereas it was fairly photostable on a mandarin orange leaf. The photoreactions involved were: decarboxylation, hydration of the cyano group to carboxamide, cleavage of the ester or the diphenyl ether linkage, hydrolysis of the carboxamide group to carboxyl, and hydroxylation at either or both of the gem dimethyl groups. The predominant reactions in water were decarboxylation, ester bond cleavage and photo-induced evolution of [14C] carbon dioxide from the [14C] cyano label; on soil, hydration or ester bond cleavage predominated. The hydration was also of importance in river and sea water. Decarboxylation did not occur on soil and plant foliage.  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