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 15 毫秒
1.
《Biological Control》2001,20(2):160-166
The herbicides 2,4-D, glyphosate, linuron, and MCPP at rates of 1X (recommended field rate), 0.25X, 0.025X, and 0.0125X were evaluated in vitro for their effects on the rust fungus Puccinia lagenophorae, a biocontrol agent for the annual weed Senecio vulgaris. Herbicides applied at 1X and 0.25X completely prevented aeciospore germination. Glyphosate was toxic even at 0.0125X and 0.025X. Aeciospores germinated in linuron, 2,4-D, and MCPP at 0.025X and 0.0125X at rates similar to the water control. Abnormal germ-tube growth was observed with 2,4-D at 0.25X and 0.025X, with linuron at 0.025X, and with glyphosate at 0.0125X. Further in planta studies were perfomed with two inbred lines of S. vulgaris inoculated with aeciospores of P. lagenophorae and treated with water, linuron, and 2,4-D at 0.025X at different times of application. Quantitative analysis of the infection process revealed that both herbicides reduced spore deposition on the leaves and altered leaf morphology. The herbicides had no effect on disease severity at this low rate although linuron significantly reduced the formation of infection peg. Timing of herbicide application had no influence on the infection process, and the effect of the herbicides on fungal development did not differ between the two plant lines. Thus, the herbicides applied at 0.025X did not increase plant susceptibility to the rust fungus, and the rates of 1X, 0.25X, and 0.025X would prevent, inhibit, or delay fungal development. Therefore, joint application of P. lagenophorae with these herbicides to control S. vulgaris cannot be recommended.  相似文献   

2.
Quantum-chemical calculations using DFT, have been performed to explain the molecular structure antioxidant activity relationship of resveratrol (RSV) (1) analogues: 3,4-dihydroxy-trans-stilbene (3,4-DHS) (2); 4,4′-dihydroxy-trans-stilbene (4,4′-DHS) (3); 4-hydroxy-trans-stilbene (4-HS) (4); 3,5-dihydroxy-trans-stilbene (3,5-DHS) (5); 3,3′-dimethoxy-4,4′-dihydroxy-trans-stilbene (3,3′-DM-4,4′-DHS) (6); 2,4-dihydroxy-trans-stilbene (2,4-DHS) (7) and 2,4,4′-trihydroxy-trans-stilbene (2,4,4′-THS) (8). It was found that all compounds studied were effective antioxidants with the exception of 3, 5-DHS. The high antioxidant activity of both 3, 3′-DM-4, 4′-DHS and 3, 4-DHS may be due to the abstraction of the two hydrogen atoms of the para and ortho-position hydroxyls respectively, to form a quinone structure. Our results revealed that the antioxidant pharmacophore of 2,4-DHS and 2,4,4′-THS, exhibiting higher antioxidant activity than resveratrol, is the 2-hydroxystilbene, rather than 4-hydroxystilbene. Experimental observations were satisfactorily explained and commented.  相似文献   

3.
The prooxidant effect of resveratrol (3,5,4′-trihydroxy-trans-stibene) and its synthetic analogues (ArOH), that is, 3,4,4′-trihydroxy-trans-stibene (3,4,4′-THS), 3,4,5-trihydroxy-trans-stibene (3,4,5-THS), 3,4-dihydroxy-trans-stibene (3,4-DHS), 4,4′-dihydroxy-trans-stibene (4,4′-DHS), 2,4-dihydroxy-trans-stilbene (2,4-DHS), 3,5-dihydroxy-trans-stilbene (3,5-DHS) and 3,5,4′-trimethoxy-trans-stibene (3,5,4′-TMS), on supercoiled pBR322 plasmid DNA strand breakage and calf thymus DNA damage in the presence of Cu (II) ions has been studied. It was found that the compounds bearing ortho-dihydroxyl groups (3,4-DHS, 3,4,4′-THS, and 3,4,5-THS) or bearing 4-hydroxyl groups (2,4-DHS, 4,4′-DHS, and resveratrol) exhibit remarkably higher activity in the DNA damage than the ones bearing no such functionalities. Kinetic analysis by UV-visible spectra demonstrates that the formation of ArOH-Cu (II) complexes, the stabilization of oxidative intermediate derived from ArOH and Cu (II)/Cu (I) redox cycles, might be responsible for the DNA damage. This study also reveals a good correlation between antioxidant and prooxidant activity, as well as cytotoxicity against human leukemia (HL-60 and Jurkat) cell lines. The mechanisms and implications of these observations are discussed.  相似文献   

4.
The average frequency of spontaneous mitotic chromosome aberrations, determined in 12 weed species growing under natural conditions, was 0.4%. Herbicides induced significant increases in this frequency in five species, Ambrosia artemisiifolia L., Pastinaca sativa L., Solidago canadensis L., Solidago nemoralis Ait., and Vicia cracca L. The auxin herbicides — 2,4-D, picloram, and 2,4-D + 2,4,5-T — induced a larger proportion of lagging chromosomes and a smaller proportion of fragmented chromosomes than found among spontaneous aberrations. The non-selective herbicides, simazine and diuron, produced multipolar spindles that were not observed among untreated cells. Clastogenic effects of nonselective herbicides in Pastinaca sativa were short-lived, giving highest frequencies of aberrant cells in June and July and lower aberration rates in August and September. The opposite trend was found in untreated flower buds of this species, suggesting that the spontaneous aberration rate is higher in flower buds from older plants.  相似文献   

5.
Delftia acidovorans MC1071 can productively degrade R-2-(2,4-dichlorophenoxy)propionate (R-2,4-DP) but not 2,4-dichlorophenoxyacetate (2,4-D) herbicides. This work demonstrates adaptation of MC1071 to degrade 2,4-D in a model two-dimensional porous medium (referred to here as a micromodel). Adaptation for 2,4-D degradation in the 2 cm-long micromodel occurred within 35 days of exposure to 2,4-D, as documented by substrate removal. The amount of 2,4-D degradation in the adapted cultures in two replicate micromodels (~10 and 20 % over 142 days) was higher than a theoretical maximum (4 %) predicted using published numerical simulation methods, assuming instantaneous biodegradation and a transverse dispersion coefficient obtained for the same pore structure without biomass present. This suggests that the presence of biomass enhances substrate mixing. Additional evidence for adaptation was provided by operation without R-2,4-DP, where degradation of 2,4-D slowly decreased over 20 days, but was restored almost immediately when R-2,4-DP was again provided. Compared to suspended growth systems, the micromodel system retained the ability to degrade 2,4-D longer in the absence of R-2,4-DP, suggesting slower responses and greater resilience to fluctuations in substrates might be expected in the soil environment than in a chemostat.  相似文献   

6.
Three kinds of diphenyl ether herbicides, 4-nitrophenyl 2,4,6-trichlorophenyl ether (CNP, chlornitrofen), 2,4-dichlorophenyl 3-methoxy-4-nitrophenyl ether (chlomethoxynil) and 2,4-dichlorophenyl 3-methoxycarbonyl-4-nitrophenyl ether (bifenox), were tested for mutagenicity in Salmonella typhimurium YG1026 and YG1021, which have high nitroreductase activity, and also in S. typhimurium TA100 and TA98. CNP and chlomethoxynil showed mutagenicity in S. typhimurium YG1026, without S9 mix, inducing 50 and 304 revertants per μg. These mutagenicities were suppressed by the addition of S9 mix. CNP and chlomethoxynil were also mutagenic to YG1021 with and without S9 mix, and their mutagenicities were lower than those to YG1026. On the other hand, bifenox was mutagenic to YG1026 only with S9 mix, inducing 3.0 revertants per μg. These three herbicides showed no mutagenicity in S. typhimurium TA100 and TA98 either with or without S9 mix.  相似文献   

7.
Phenoxyalkanoic compounds are used worldwide as herbicides. Cupriavidus necator JMP134(pJP4) catabolizes 2,4-dichlorophenoxyacetate (2,4-D) and 4-chloro-2-methylphenoxyacetate (MCPA), using tfd functions carried on plasmid pJP4. TfdA cleaves the ether bonds of these herbicides to produce 2,4-dichlorophenol (2,4-DCP) and 4-chloro-2-methylphenol (MCP), respectively. These intermediates can be degraded by two chlorophenol hydroxylases encoded by the tfdBI and tfdBII genes to produce the respective chlorocatechols. We studied the specific contribution of each of the TfdB enzymes to the 2,4-D/MCPA degradation pathway. To accomplish this, the tfdBI and tfdBII genes were independently inactivated, and growth on each chlorophenoxyacetate and total chlorophenol hydroxylase activity were measured for the mutant strains. The phenotype of these mutants shows that both TfdB enzymes are used for growth on 2,4-D or MCPA but that TfdBI contributes to a significantly higher extent than TfdBII. Both enzymes showed similar specificity profiles, with 2,4-DCP, MCP, and 4-chlorophenol being the best substrates. An accumulation of chlorophenol was found to inhibit chlorophenoxyacetate degradation, and inactivation of the tfdB genes enhanced the toxic effect of 2,4-DCP on C. necator cells. Furthermore, increased chlorophenol production by overexpression of TfdA also had a negative effect on 2,4-D degradation by C. necator JMP134 and by a different host, Burkholderia xenovorans LB400, harboring plasmid pJP4. The results of this work indicate that codification and expression of the two tfdB genes in pJP4 are important to avoid toxic accumulations of chlorophenols during phenoxyacetic acid degradation and that a balance between chlorophenol-producing and chlorophenol-consuming reactions is necessary for growth on these compounds.  相似文献   

8.
Selection genes are routinely used in plant genetic transformation protocols to ensure the survival of transformed cells by limiting the regeneration of non-transgenic cells. In order to find alternatives to the use of antibiotics as selection agents, we followed a targeted approach utilizing a plant gene, encoding a mutant form of the enzyme acetolactate synthase, to convey resistance to herbicides. The sensitivity of sugarcane callus (Saccharum spp. hybrids, cv. NCo310) to a number of herbicides from the sulfonylurea and imidazolinone classes was tested. Callus growth was most affected by sulfonylurea herbicides, particularly 3.6 μg/l chlorsulfuron. Herbicide-resistant transgenic sugarcane plants containing mutant forms of a tobacco acetolactate synthase (als) gene were obtained following biolistic transformation. Post-bombardment, putative transgenic callus was selectively proliferated on MS medium containing 3 mg/l 2,4-dichlorophenoxyacetic acid (2,4-D), 20 g/l sucrose, 0.5 g/l casein, and 3.6 μg/l chlorsulfuron. Plant regeneration and rooting was done on MS medium lacking 2,4-D under similar selection conditions. Thirty vigorously growing putative transgenic plants were successfully ex vitro-acclimatized and established under glasshouse conditions. Glasshouse spraying of putative transgenic plants with 100 mg/l chlorsulfuron dramatically decreased the amount of non-transgenic plants that had escaped the in vitro selection regime. PCR analysis showed that six surviving plants were als-positive and that five of these expressed the mutant als gene. This report is the first to describe a selection system for sugarcane transformation that uses a selectable marker gene of plant origin targeted by a sulfonylurea herbicide.  相似文献   

9.
Translocation and metabolism of 4-amino-3,5,6-trichloropicolinic acid (picloram) and 2,4-dichlorophenoxyacetic acid (2,4-D) in small plants of aspen (Populus tremula L.) were studied. In most experiments 14C-carboxyl-labelled herbicides were used. Considerable quantities of both herbicides were retained in the treated leaf. Translocation was mainly upwards into the growing shoot tip. Only minute quantities were found in the roots. Injection of the herbicides through a cut stem surface increased the translocation into the roots very little. One of the reasons for the limited downward translocation is considered to be a ready transfer of the herbicides between the phloem and the xylem. Both herbicides were incorporated into complexes from which the active herbicides could be released. However, this complex formation can only partly account for the retention of the herbicides in the treated leaves. The differences in metabolism found between 2,4-D and picloram cannot explain the considerable difference in toxicity between the compounds.  相似文献   

10.
The binding behaviour at the thylakoid membrane of the radioactively labelled phenolic inhibitors 2-iodo-4-nitro-6-[2′,3′-3H]isobutylphenol and 3,5-diiodo-4-hydroxy[U-14C]benzonitrile (ioxynil) has been studied. As judged from displacement experiments with other herbicides, phenolic herbicides and herbicides as represented best by 3-(3,4-dichlorophenyl)-1,1-dimethylurea have different binding sites at the reducing side of Photosystem II. The binding parameters of phenolic herbicides are not, or only slightly, affected by trypsin treatment of chloroplasts. In chloroplasts, besides free pigments, lipids, and the light-harvesting chlorophyll ab protein complex, a protein of molecular weight 41 000 is radioactively labelled by the photoaffinity label 4-nitro-2-azido-6-[2′,3′-3H]isobutylphenol. The amount of radioactivity bound to the 41 kDa protein is diminished if chloroplasts are incubated with 3-(3,4-dichlorophenyl)-1,1-dimethylurea prior to addition of the photoaffinity label, but not if the 2,4-dinitrophenyl ether of 2-iodo-4-nitrothymol is used instead. These two compounds are characteristic representatives of inhibitiors acting at the reducing or the oxidizing site of plastoquinone, respectively. Based on these results, a model for two different herbicide-binding proteins at the reducing side of Photosystem II is presented.  相似文献   

11.
Effects of Selected Herbicides on Plant Protoplasts   总被引:1,自引:0,他引:1  
Plant protoplasts were released from immature tomato fruits by incubation with a 20% solution of polygalacturonase (Pectinol R-10, Rhom & Haas) dissolved in 0.1 M KCl + 0.1 M MgCl2. In this salt solution the protoplasts remained stabilized for up to 8 h and were used as a source of exposed plasma membrane. Gross responses of protoplasts to selected chemicals and herbicides were recorded photomicroscopically. Paraquat (1,1′-dimethyl-4,4′-bipyridinium ion) treatments resulted in a characteristic response which was different from that of general denaturants (trichloroacetic acid, ethanol, and detergents) and of osmotic shock. Initial phases of the paraquat response were characterized by a segregation of the cytoplasm into isolated areas on the inner membrane surface. The final phase was a rupture of the plasma membrane and collapse of the cell. The herbicides, 2,4′-dinitro-4-trifluoromethyl-diphenylether (preforan); 1,1-dimethyl-3-(α,α,α-trifluoro-m-tolyl)urea (fluometuron); 3-(3-chloro-4-bromophenyl)-1-methoxy-1-methylurea (chlorbromuron); and α,α,α-trifluoro-2,6-dinitro-N-N-dipropyl-p-toluidine (trifluralin) produced no apparent structural effect on the protoplasts.  相似文献   

12.
Uptake, translocation and metabolism of 14C-labelled 4-amino-3,5,6-trichloropicolinic acid (picloram) and 2,4-dichlorophenoxyacetic acid (2,4-D) in seedlings of wheat (Triticum aestivum L.) were studied. The uptake of the herbicides through the upper surface of the first leaf was slow but was almost complete after nine days. Picloram was absorbed faster than 2,4-D. Picloram was also translocated into the stem and the untreated leaves to a greater extent than 2,4-D. Only small fractions of the activity were recovered from the roots and from the nutrient solution. Picloram and 2,4-D formed water-soluble conjugates in the tissues. These conjugates were very labile and hydrolyzed under release of the unchanged herbicides. The isotope from 2,4-D was also incorporated in an insoluble fraction, containing cell walls and proteins. Also from this fraction biologically active 2,4-D could be released by hydrolysis. The formation of the complexes was partly prevented by cycloheximide. It is suggested that herbicide detoxification through complex formation is of importance for the relatively low sensitivity of wheat to auxin herbicides.  相似文献   

13.
Two new 4-quinazolone alkaloids have been isolated from seed husks of Zanthoxylum arborescens. Based on their spectroscopic properties they have been assigned structures, 1-methyl-3-(2′-phenylethyl)-1H,3H-quinazoline-2,4-dione and 1-methyl-3-[2′-(4″-methoxyphenyl)ethyl]-1H,3H-quinazoline-2,4-dione. These structural assignments have been confirmed by synthesis. Skimmianine has been obtained from leaf extracts of Z. dimoncillo and Z. caribaeum while skimmianine and scopeletin have been isolated from leaf extracts of Z. fagara.  相似文献   

14.
Two new isoflavonoids, named 6,7,2′-trihydroxy-4′-methoxyisoflavone (1), 7,3′-dimethoxy-5′-hydroxyisoflavone (2), one new norneolignan, named (8S)-2,4-dihydroxy-8-hydroxymethyl-4′-methoxydeoxybenzoin (3); together with six known compounds, methyl 4-hydroxylbenzoate (4), ethyl 4-hydroxybenzoate (5), piceatannol (6), cararosin A (7), 2,4-dihydroxybenzoate (8), and 6,7,4′-trihydroxyisoflavone (9) were isolated from the red heartwood in the rhizomes of Caragana changduensis by using chromatographic methods Their structures were determined by extensive spectroscopic analysis and comparison of their spectral data with previous reported data.  相似文献   

15.
Two dihydrochalcones, 2′,6′-dihydroxy-4′-methoxydihydrochalcone and 2,4′,6′-trihydroxydihydrochalcone have been isolated from leaves of Lindera umbellata.  相似文献   

16.
《农业工程》2020,40(6):492-499
The research and application of natural product herbicides have received considerable attention recently over the world as alternative tools against chemical herbicides for weed control due to many unique properties. A wide variety of compounds shows the broadest spectrum of herbicidal activity were found in Egyptian plant resources including; [6,3′-dihydroxy-3,5,7,4′-tetra methoxy flavone, dihydro-quercetin, 3,6,7,3`,4`-pentamethoxyflavone, quercetagetin 3, 5, 6, 7, 3′, 4′-hexamethyl ether, 6,-4′-dihydroxy-3,7-dimethoxyflavone, 6,4-dihydroxy-3,5,7-trimethoxyflavone, sesquiterpene (Eudesm-4(15), 11(13)-diene-12,5β-oIide) and 3, 5-dicaffeoyl quinic acid] from Jasonia montana, [15-hydroxyisocostic acid, methyl 15- oxo-eudesome-4, 11(13)-diene 12-oate as well as 1α, 9α-dihydroxy-α-cyclocostunolide, isorhamnetin 3-sulfate, isorhamnetin 3-O-rutinoside rhamanetin and epicatechin] from Conyza dioscoridis, [chlorogenic acid, hydroxyl-3-methoxyflavone, quercetin, kaempferol 3β-D-6”-O-cis-cinnamoyl glucoside, kaempferol, mangiferin, coumaroyl glucoside, coumaroyl quince acid, dicaffeoyl quinic acids] from Silverleaf nightshade Solanum elaeagnifolium Cav, [apigenin, matricolone, herniarin and coumarin, apigenin-7-O-4″, 6″-diacetyl glycoside and apigenin 7-O-4–acetyl glycoside] from Matricaria chamomilla, and [kaempferol 3-O-β-(6″-p-coumaroyl glucopyranoside] from Abutilon theophrasti respectively. These constituents are isolated by chromatographic techniques and identified by spectroscopic methods and tested in both pre and post emergence stages of weeds to determine the effective dose and time for use. The natural herbicide isolated from plant or microorganisms are potentially useful as selective, biodegradable, safe to the environment which will provide an alternative natural solution for combating crop weeds. This review focuses on the characteristics of natural product herbicides from Egyptian plants and evaluates against weeds.  相似文献   

17.
Phenoxyalkanoic herbicides such as 2,4‐dichlorophenoxyacetate (2,4‐D), 2,4‐dichlorophenoxybutyrate (2,4‐DB) or mecoprop are widely used to control broad‐leaf weeds. Several bacteria have been reported to degrade these herbicides using the α‐ketoglutarate‐dependent, 2,4‐dichlorophenoxyacetate dioxygenase encoded by the tfdA gene, as the enzyme catalysing the first step in the catabolic pathway. The effects of exposure to different phenoxyalkanoic herbicides in the soil bacterial community and in the tfdA genes diversity were assessed using an agricultural soil exposed to these anthropogenic compounds. Total community bacterial DNA was analysed by terminal restriction fragment length polymorphism of the 16S rRNA and the tfdA gene markers, and detection and cloning of tfdA gene related sequences, using PCR primer pairs. After up to 4 months of herbicide exposure, significant changes in the bacterial community structure were detected in soil microcosms treated with mecoprop, 2,4‐DB and a mixture of both plus 2,4‐D. An impressive variety of novel tfdA gene related sequences were found in these soil microcosms, which cluster in new tfdA gene related sequence groups, unequally abundant depending on the specific herbicide used in soil treatment. Structural analysis of the putative protein products showed small but significant amino acid differences. These tfdA gene sequence variants are, probably, required for degradation of natural substrate(s) structurally related to these herbicides and their presence explains self‐remediation of soils exposed to phenoxyalkanoic herbicides.  相似文献   

18.
(±)-(2Z,4E)-5-(1′,2′-epoxy-2′,6′,6′-trimethylcyclohexyl)-3-methyl-2,4-pentadienoic acid was metabolized by Cercospora cruenta, which has the ability to produce (+)-abscisic acid (ABA), to give (±)-(2Z,4E)-xanthoxin acid, (±)-(2Z,4E)-5′-hydroxy-1′,2′-epoxy-1′,2′-dihydro-β-ionylideneacetic acid, (±)-1′,2′-epoxy-1′,2′-dihydro-β-ionone and trace amounts of ABA.  相似文献   

19.
The aim of this study was to enrich and characterise bacterial consortia from soils around a herbicide production plant through their capability to degrade the herbicides 4-(2,4-dichlorophenoxy) butyric acid (2,4-DB) and 4-(4-chloro-2-methylphenoxy) butyric acid (MCPB). Partial 16S rRNA gene sequencing revealed members of the genera Stenotrophomonas, Brevundimonas, Pseudomonas, and Ochrobactrum in the 2,4-DB- and MCPB-degrading communities. The degradation of 2,4-DB and MCPB was facilitated by the combined activities of the community members. Some of the members were able to utilise other herbicides from the family of chlorophenoxyalkanoic acids. During degradation of 2,4-DB and MCPB, phenol intermediates were detected, indicating ether cleavage of the side chain as the initial step responsible for the breakdown. This was also verified using an indicator medium. Repeated attempts to amplify putatively conserved tfd genes by PCR indicated the absence of tfd genes among the consortia members. First step cleavage of the chlorophenoxybutyric acid herbicides is by ether cleavage in bacteria and is encoded by divergent or different tfd gene types. The isolation of mixed cultures capable of degrading 2,4-DB and MCPB will aid future investigations to determine both the metabolic route for dissimilation and the fate of these herbicides in natural environments.  相似文献   

20.
The 1,3-dipolar cycloaddition of azomethine ylides derived from acenaphthenequinone and α-amino acids viz. sarcosine, phenylglycine, 1,3-thiazolane-4-carboxylic acid and proline to a series of 2,6-bis[(E)-arylmethylidene]cyclohexanones afforded novel spiro-heterocycles chemo-, regio- and stereoselectively in quantitative yields. These compounds were screened for their in vitro activity against Mycobacterium tuberculosis H37Rv (MTB) using agar dilution method. Two compounds, 4-(2,4-dichlorophenyl)-5-phenylpyrrolo(spiro[2.2″]acenaphthene-1″-one)spiro[3.2′]-6′-(2,4-dichlorophenylmethylidene)cyclohexanone (4i) and spiro[5.2″]acenaphthene-1″-onespiro[6.2′]-6′-(2,4-dichlorophenylmethylidene)cyclohexanone-7-(2,4-dichlorophenyl)tetrahydro-1H-pyrrolo[1,2-c][1,3]thiazole (5i) display maximum activity in vitro with a MIC value of 0.40 μg/mL against MTB and were 4 and 15.6 times more potent than ethambutol and pyrazinamide, respectively.  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