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.
Studies were conducted to determine a physiological basis for competitive differences between Senecio vulgaris L. biotypes which are either resistant or susceptible to triazine herbicides. Net carbon fixation of intact leaves of mature plants was higher at all light intensities in the susceptible biotype than in the resistant biotype. Quantum yields measured under identical conditions for each biotype were 20% lower in the resistant than in the susceptible biotype. Oxygen evolution in continuous light measured in stroma-free chloroplasts was also higher at all light intensities in the susceptible biotype than in the resistant biotype. Oxygen evolution in response to flashing light was measured in stroma-free chloroplasts of both biotypes. The steady-state yield per flash of resistant chloroplasts was less than 20% that of susceptible chloroplasts. Susceptible chloroplasts displayed oscillations in oxygen yield per flash typically observed in normal chloroplasts, whereas the pattern of oscillations in resistant chloroplasts was noticeably damped. It is suggested that modification of the herbicide binding site which confers s-triazine resistance may also affect the oxidizing side of photosystem II, making photochemical electron transport much less efficient. This alteration has resulted in a lowered capacity for net carbon fixation and lower quantum yields in whole plants of the resistant type.  相似文献   

2.
Wim F.J. Vermaas  Charles J. Arntzen   《BBA》1983,725(3):483-491
We have analyzed the binding of synthetic quinones and herbicides which inhibit electron transport at the acceptor side of Photosystem II (PS II) of the photosynthetic electron-transport chain in thylakoid membranes. These data show that quinones and PS II-directed herbicides compete for binding to a common binding environment within a PS II region which functions as the Q / PQ oxidoreductase. We observed that (1) synthetic quinones cause a parallel inhibition of electron transport and [14C]herbicide displacement, and (2) herbicide binding is affected both by the fully oxidized and fully reduced form of a quinone. Quinone function and inhibitor binding were also investigated in thylakoids isolated from triazine-resistant weed biotypes. We conclude the following. (1) The affinity of the secondary accepting quinone, B, is decreased in resistant thylakoids. (2) The observation that the equilibrium concentration of reduced Q after transferring one electron to the acceptor side of PS II is increased in resistant as compared to susceptible chloroplasts may be explained both by a decrease in the affinity of PQ for the herbicide / quinone binding environment, and by a decrease of the midpont redox potential of the B / B couple. (3) The binding environment regulating quinone and herbicide affinity may be divided roughly into two domains; we suggest that the domain regulating quinone head-group binding is little changed in resistant membranes, whereas the domain-regulating quinone side-group binding (and atrazine) is altered. This results in increased inhibitory activity of tetrachloro-p-benzoquinone and phenolic herbicides, which are hypothesized to utilize the quinone head-group domain. The two domains appear to be spatially overlapping because efficient atrazine displacement by tetrachloro-p-benzoquinone is observed.  相似文献   

3.
Uptake, translocation and complex formation of 14C-labelled 4-amino-3,5,6-trichloropicolinic acid (picloram) and 2,4-dichlorophenoxyacetic acid (2,4-D) in seedlings of rape (Brassica napus L. cv. Nilla) and sunflower (Helianthus annuus L. var. uniflorus) were studied. Sunflower is susceptible both to 2,4-D and picloram, while rape is susceptible to 2,4-D but more tolerant to picloram. The uptake of the herbicides through the leaves was almost complete in both species. Translocation of 2,4-D into the roots took place more readily than that of picloram. In sunflower about 50 per cent of the applied 2,4-D was extruded through the roots into the nutrient solution after 9 days. In the picloram-treated sunflower most of the activity was found in the aerial parts, while in picloram-treated rape most of the activity still occurred in the treated leaf after 9 days. No activity at all was found in the roots or in the nutrient solution of the picloram-treated rape seedlings. While the major part of 2,4-D always was found in the state of free herbicide, a large fraction of picloram was rapidly bound into water-soluble complexes. This binding was especially pronounced in rape. Separation by paper chromatography showed that different radioactive compounds were formed. Most of these could be hydrolyzed, thereby releasing free herbicide. The results support the hypotheses that complex formation could counteract herbicide translocation and toxicity of auxin herbicides.  相似文献   

4.
The herbicides trifluralin (alpha,alpha,alpha-trifluoro-2,6-dinitro-N, N-dipropyl-p-toluidine) and diallate (S-[2,3-dichloroallyl] diisopropylthiocarbamate) inhibit electron transport, ATP synthesis, and cytochrome f reduction by isolated spinach (Spinacia oleracea L.) chloroplasts. Both compounds inhibit noncyclic electron transport from H(2)O to ferricyanide more than 90% in coupled chloroplasts at concentrations less than 50 mum. Neither herbicide inhibits electron transport in assays utilizing only photosystem I activity, and the photosystem II reaction elicited by addition of oxidized p-phenylenediamine or 2,5-dimethylquinone is only partially inhibited by herbicide concentrations which block electron flow from H(2)O to ferricyanide. Inhibition of ATP synthesis parallels inhibition of electron flow in all noncyclic assay systems, and cyclic ATP synthesis catalyzed by either diaminodurene or phenazine metho-sulfate is susceptible to inhibition by both herbicides. These results indicate that trifluralin and diallate both inhibit electron flow in isolated chloroplasts at a point in the electron transport chain between the two photosystems.  相似文献   

5.
Wild relatives of genetically engineered crops can acquire transgenic traits such as herbicide resistance via spontaneous crop–wild hybridization. In agricultural weeds, resistance to herbicides is often a beneficial trait, but little is known about possible costs that could affect the persistence of this trait when herbicides are not used. We tested for costs associated with transgenic resistance to glufosinate when introgressed into weedy Brassica rapa . Crosses were made between transgenic B. napus and wild B. rapa from Denmark. F1 progeny were backcrossed to B. rapa and BC1 plants were selected for chromosome numbers similar to B. rapa . Further backcrossing resulted in a BC2 generation that was hemizygous for herbicide resistance. We quantified the reproductive success of 457 BC3 progeny representing six full-sib families raised in growth rooms (plants were pollinated by captive bumblebees). Pollen fertility and seed production of BC3 plants were as great as those of B. rapa raised in the same growth rooms. Segregation for herbicide resistance in BC3 plants was 1:1 overall, but the frequency of resistant progeny was lower than expected in one family and higher than expected in another. There were no significant differences between transgenic and nontransgenic plants in survival or the number of seeds per plant, indicating that costs associated with the transgene are probably negligible. Results from this growth-chamber study suggest that transgenic resistance to glufosinate is capable of introgressing into populations of B. rapa and persisting, even in the absence of selection due to herbicide application.  相似文献   

6.
The present study compares the binding and inhibitory activity of two photosystem II inhibitors: 3-(3,4-dichlorophenyl)-1,1-dimethylurea (diuron [DCMU]) and 2-chloro-4-(ethylamine)-6-(isopropyl amine)-S-triazene (atrazine). Chloroplasts isolated from naturally occurring triazine-susceptible and triazine-resistant biotypes of common groundsel (Senecio vulgaris L.) showed the following characteristics. (a) Diuron strongly inhibited photosynthetic electron transport from H2O to 2,6-dichlorophenolindophenol in both biotypes. Strong inhibition by atrazine was observed only with the susceptible chloroplasts. (b) Hill plots of electron transport inhibition data indicate a noncooperative binding of one inhibitor molecule at the site of action for both diuron and atrazine. (c) Susceptible chloroplasts show a strong diuron and atrazine binding (14C-radiolabel assays) with binding constants (K) of 1.4 × 10−8 molar and 4 × 10−8 molar, respectively. In the resistant chloroplasts the diuron binding was slightly decreased (K = 5 × 10−8 molar), whereas no specific atrazine binding was detected. (d) In susceptible chloroplasts, competitive binding between radioactively labeled diuron and non-labeled atrazine was observed. This competition was absent in the resistant chloroplasts.  相似文献   

7.
《BBA》1985,808(1):103-111
The competition between various herbicides and plastoquinone, at the QB site of Photosystem II, has been studied by measuring absorption changes between 300 and 360 nm in spinach and chenopod chloroplasts, in response to a train of saturating short xenon flashes. A complex pattern was observable without addition of chemicals interfering with electron flow. The effect of potassium ferricyanide, of hydroxylamine and of valinomycine-K+ permitted to simplify the pattern and, in particular, to observe binary oscillations with flash number, attributable to the functioning of the two-electron gate QB. Herbicides belonging to different classes block electron transfer when added at high concentration. At low concentration, however, inverted binary oscillations become observable. When the chloroplasts have been first oxidized with ferricyanide, this behaviour develops progressively in response to illumination. Varying the herbicide concentration, it appears that the concentration inducing maximum binary oscillations correlates with inhibition of linear electron transfer, within each class of herbicides. Phenolic herbicides induce the largest oscillations and ureas the smallest. The binary oscillations have the spectrum of the plastoquinone anion. The results clearly show that the studied herbicides compete efficiently with QB, but not with QB, at the QB binding site. Atrazine-resistant chenopod chloroplasts still display normal binary oscillations in the absence of herbicide, or in the presence of atrazine alone. They are highly sensitive to DCMU and to i-dinoseb, but no inverted binary oscillations could be observed with these herbicides.  相似文献   

8.
Rapid and widespread evolution of multiple herbicide resistance in global weed species endowed by increased capacity to metabolize (degrade) herbicides (metabolic resistance) is a great threat to herbicide sustainability and global food production. Metabolic resistance in the economically damaging crop weed species Lolium rigidum is well known but a molecular understanding has been lacking. We purified a metabolic resistant (R) subset from a field evolved R L. rigidum population. The R, the herbicide susceptible (S) and derived F2 populations were used for candidate herbicide resistance gene discovery by RNA sequencing. A P450 gene CYP81A10v7 was identified with higher expression in R vs. S plants. Transgenic rice overexpressing this Lolium CYP81A10v7 gene became highly resistant to acetyl-coenzyme A carboxylase- and acetolactate synthase-inhibiting herbicides (diclofop-methyl, tralkoxydim, chlorsulfuron) and moderately resistant to hydroxyphenylpyruvate dioxygenase-inhibiting herbicide (mesotrione), photosystem II-inhibiting herbicides (atrazine and chlorotoluron) and the tubulin-inhibiting herbicide trifluralin. This wide cross-resistance profile to many dissimilar herbicides in CYP81A10v7 transgenic rice generally reflects what is evident in the R L. rigidum. This report clearly showed that a single P450 gene in a cross-pollinated weed species L. rigidum confers resistance to herbicides of at least five modes of action across seven herbicide chemistries.  相似文献   

9.
The mode of action of chemically different herbicides (ureas, pyridazinones, phenylcarbamates, triazines, hydroxyquinolines, hydroxybenzonitriles and dinitrophenols) on photosynthetic electron transport was investigated by measurements of oxygen evolution and thermoluminescence. Depending on the particular herbicide used the thermoluminescence band related to Q (the primary acceptor of Photosystem II) appears at +5, 0 or −14°C. It was shown that these three different peak positions can be ascribed to various redox states of Q, the shifts being due to the binding of herbicides to the chloroplast membrane. Both displacement experiments and additive inhibition of herbicide pairs measured by thermoluminescence and oxygen evolution suggested that the sites of action of these herbicides are on the same protein. However, herbicide treatment of trypsinized chloroplasts showed that there were three different binding sites on the same protein, in agreement with the classification of herbicides into three groups based on thermoluminescence measurements. Our results suggest that the primary and secondary acceptors of Photosystem II (Q and B, respectively) are in close proximity and form a common complex with the herbicide-binding protein within the chloroplast membrane.  相似文献   

10.
Kobek, K., Focke, M., Lichtenthaler, H.K., Retzlaff, G. and Würzer, B. 1988. Inhibiton of fatty acid biosynthesis in isolated chloroplasts by cycloxydim and other cyclohexane-1,3-diones. - Physiol. Plant. 72: 492–498.
The effect of the three cyclohexane-1,3-dione herbicides cycloxydim, sethoxydim and clethodim (proposed common name) on the de novo fatty acid biosynthesis of isolated chloroplasts as test system was investigated with intact chloroplasts isolated from sensitive grasses (Poaceae) and tolerant dicotyledonous plants. All three herbicides blocked the de novo fatty acid biosynthesis ([14C]-acetatc incorporation into total fatty acid fraction) in Avena sativa L. cv. Flämingnova chloroplasts in a dose-dependent manner. The I50-values are lower for cycloxydim and clethodim than for sethoxydim. The rate of de novo fatty acid biosynthesis in isolated, intact and photosynthetically active Avena chloroplasts was higher in the light than in the dark, which appeared to be due to the light-dependent regeneration of the cofactors ATP and NADPH. The de novo fatty acid biosynthesis by isolated chloroplasts from the tolerant dicotyledonous species pea ( Pisum savivum L. cv. Kleine Rheinländerin), spinach ( Spinacea oleracea L. cv. Matador) and tobacco ( Nicotiana tabacum L. cv. su/su) was insensitive to the three herbicides. It is assumed that one of the enzymes of the fatty acid biosynthesis is modified in the dicotyledonous plants and not accessible to the cyclohexane-1,3-dione herbicides. In the case of Poa annua L., which as a whole plant is tolerant towards sethoxydim, the tolerance seems not to lie in the chloroplasts but in properties of the cytoplasm, since the isolated chloroplasts are sensitive to the herbicide.  相似文献   

11.
Summary Somatic hybridization between Brassica oleracea ssp. botrytis (cauliflower, 2n=18), carrying the Ogura (R1) male-sterile cytoplasm and B. napus (2n= 38), carrying a male-fertile, atrazine-resistant (ATR) cytoplasm, yielded three hybrids (2n=56) and six cauliflower cybrids (2n=18), which were selected for resistance to the herbicide in vitro. The hybrids and cybrids were male fertile and self-compatible. They contained both chloroplasts and mitochondria from the ATR cytoplasm. We found no evidence for mtDNA recombination in any of the regenerated plants. Selfed progeny of the B. oleracea atrazine-resistant cybrids were evaluated for tolerance to the herbicide in the field. Resistant plants exposed to 0.56–4.48 kg/ha (0.5–4.0 pounds/acre) atrazine in the soil showed no damage at any herbicide level, whereas plants of a susceptible alloplasmic line were severely damaged at the lowest level of herbicide application and killed at all higher levels. These atrazine-resistant cauliflower may have potential horticultural use, especially in fields where atrazine carry over is a serious problem.  相似文献   

12.
13.
Atrazine belongs to the widely used herbicides blocking the electron transport chain in chloroplasts, thus resulting in the generation of active oxygen species. In the present work, we demonstrated that, at low concentrations mimicking residual amounts, atrazine enhanced the susceptibility of Arabidopsis plants to further treatments with the same herbicide applied at the recommended field rate. Arabidopsis thaliana plants were treated three times (at five-day intervals) with 1 µM atrazine. Five days after the last treatment, the plants were sprayed with 5 mM atrazine. Atrazine increased the levels of lipid peroxidation products, hydrogen peroxide, and ion leakage, and caused changes in the activities of antioxidant enzymes, such as superoxide dismutase, guaiacol peroxidase, and catalase.From Fiziologiya Rastenii, Vol. 52, No. 2, 2005, pp. 243–249.Original English Text Copyright © 2005 by Ivanov, Alexieva, Karanov.This article was submitted by the authors in English.This revised version was published online in April 2005 with a corrected cover date.  相似文献   

14.
Salinity stress is a major abiotic stress that affects plant growth and limits crop production. Roots are the primary site of salinity perception, and salt sensitivity in roots limits the productivity of the entire plant. To better understand salt stress responses in canola, we performed a comparative proteomic analysis of roots from the salt-tolerant genotype Safi-7 and the salt-sensitive genotype Zafar. Plants were exposed to 0, 150, and 300 mM NaCl. Our physiological and morphological observations confirmed that Safi-7 was more salt-tolerant than Zafar. The root proteins were separated by two-dimensional gel electrophoresis and MALDI-TOF mass spectrometry was applied to identify proteins regulated in response to salt stress. We identified 36 and 25 protein spots whose abundance was significantly affected by salt stress in roots of plants from the tolerant and susceptible genotype, respectively. Functional classification analysis revealed that the differentially expressed proteins from the tolerant genotype could be assigned to 14 functional categories, while those from the susceptible genotype could be classified into 9 functional categories. The most significant differences concerned proteins involved in glycolysis (Glyceraldehyde-3-phosphate dehydrogenase, Fructose-bisphosphate aldolase, Phosphoglycerate kinase 3), stress (heat shock proteins), Redox regulation (Glutathione S-transferase DHAR1, L-ascorbate peroxidase), energy metabolism (ATP synthase subunit B), and transport (V-type proton ATPase subunit B1) which were increased only in the tolerant line under salt stress. Our results provide the basis for further elucidating the molecular mechanisms of salt-tolerance and will be helpful for breeding salt-tolerant canola cultivars.  相似文献   

15.
Metabolism of the herbicide atrazine by Rhodococcus strains.   总被引:20,自引:8,他引:12       下载免费PDF全文
R Behki  E Topp  W Dick    P Germon 《Applied microbiology》1993,59(6):1955-1959
Rhodococcus strains were screened for their ability to degrade the herbicide atrazine. Only rhodococci that degrade the herbicide EPTC (s-ethyl-dipropylthiocarbamate) metabolized atrazine. Rhodococcus strain TE1 metabolized atrazine under aerobic conditions to produce deethyl- and deisopropylatrazine, which were not degraded further and which accumulated in the incubation medium. The bacterium also metabolized the other s-triazine herbicides propazine, simazine, and cyanazine. The N dealkylation of triazine herbicides by Rhodococcus strain TE1 was associated with a 77-kb plasmid previously shown to be required for EPTC degradation.  相似文献   

16.
Growth and electrophysiological studies in roots of intact diclofop-methyl susceptible and resistant seedlings were conducted to test the hypothesis that the herbicide acts primarily as a proton ionophore. The ester formulation of diclofop, at 0.2 micromolar, completely inhibited root growth in herbicide-susceptible oat (Avena sativa L.) after a 96 hour treatment, but induced only a delayed transient depolarization of the membrane potential in oat root cortical cells. Root growth in susceptible maize (Zea mays L.) seedlings was dramatically reduced by exposure to 0.8 micromolar diclofop-methyl, while the same diclofop-methyl exposure hyperpolarized the membrane potential within 48 hours after treatment. Furthermore, exposure of maize roots to the protonophore, carbonyl cyanide m-chlorophenylhydrazone (CCCP) (50 nanomolar), inhibited growth by only 31%, 96 hours after treatment, while the same CCCP exposure depolarized the resting potential by an average of 32 millivolts. Thus, the protonophore hypothesis cannot account for a differential membrane response to phytotoxic levels of diclofop-methyl in two susceptible species. From the results of others, much of the evidence to support the protonophore hypothesis was obtained using high concentrations of diclofop acid (100 micromolar). At a similar concentration, we also report a rapid (3 minute) diclofop-induced depolarization of the membrane potential in roots of susceptible oat and maize, moderately tolerant barley (Hordeum vulgare L.), and resistant pea (Pisum sativum L.) seedlings. Moreover, 100 micromolar diclofop acid inhibited growth in excised cultured pea roots. In contrast, 100 micromolar diclofop-methyl did not inhibit root growth. Since the membrane response to 100 micromolar diclofop acid does not correspond to differential herbicide sensitivity under field conditions, results obtained with very high levels of diclofop acid are probably physiologically irrelevant. The results of this study suggest that the effect of diclofop-methyl on the membrane potentials of susceptible species is probably unrelated to the primary inhibitory effect of the herbicide on plant growth.  相似文献   

17.
KCN-treated spinach chloroplasts, their photosystem I being ineffective, exhibit a single reaction site for silicomolybdate. Using this heteropolyanion as electron acceptor, photosynthetic oxygen evolution is partially inhibited by ureas, triazines, or phenylpyridazinone herbicides, their inhibitory effect depending on the concentration of silicomolybate. Labelled atrazine attached to isolated chloroplast material is competitively replaced by silicomolybdate in the same manner as e.g. ureas complete with a triazine herbicide. – It is concluded (1) that silicomolybdate is bound and reduced at the herbicide-binding protein, and (2) that the inhibition of silicomolybdate reduction by herbicides such as DCMU is due to loss of reaction sites for silicomolybdate.  相似文献   

18.
Mechanism of paraquat tolerance in perennial ryegrass   总被引:6,自引:3,他引:3  
Abstract The mechanism of paraquat tolerance was investigated in lines of perennial ryegrass (Lolium perenne L.) which had been selected for resistance to the herbicide. Uptake, metabolism and translocation of paraquat were studied. Susceptible cultivars and a tolerant line were not found to differ in uptake of radioactive paraquat applied to the leaf surface or supplied to the cut ends of excised leaves. Distribution of herbicide within leaf tissue was similar in tolerant and susceptible plants and no metabolites of 14C-paraquat were detected in tolerant or susceptible plants treated with sub-lethal concentrations of the herbicide. Autoradiography and quantitative determinations showed much variation in translocation of 14C-paraquat out of treated leaves of intact plants, but the variation was not related to the degree of susceptibility to the herbicide. It is concluded that paraquat tolerance in perennial ryegrass is unlikely to depend upon reduced uptake, enhanced metabolism or altered translocation of the herbicide.  相似文献   

19.
A series of Chlamydomonas reinhardii mutants were selected for resistance to the herbicides atrazine, bromacil, and diuron. Four of these have reduced herbicide binding to the thylakoid membranes and show the non-Mendelian inheritance pattern characteristic of chloroplast genes. These mutants show a variety of selective alterations in binding of the three herbicides. These changes account for the observed patterns of in vivo cross-resistance. Analyses of chloroplast gene recombination indicate that these four mutations are in the same gene. Overall, the results suggest that this gene codes for a protein component of the herbicide binding site. One of the mutants has slow phototrophic growth and altered electron transport as has been observed in atrazine-resistant higher plant varieties, but the others are normal in these respects. The slow growth characteristic of this mutant seems to be the consequence of the same mutation which confers herbicide resistance.

The mutants isolated also include a large number which achieve resistance by some secondary mechanism. These are all nuclear gene mutations, and represent numerous loci. They also show a variety of patterns of cross-resistance, but the mechanisms behind them have not yet been investigated.

  相似文献   

20.
Absorption from culture solution of the herbicides 2, 4-dichlorophenoxyacetic acid (2, 4-D) and 3-(p-chlorophenyl)-1, 1-dimethylurea (Monuron) by excised barley (Hordeum vulgare L.) roots was studied to determine whether absorption was due to an active or a passive mechanism. Herbicide absorption was followed at low temperature, under anaerobic conditions, and in the presence of metabolic inhibitors and compounds of structure similar to that of the herbicide. Total absorption was divided into two phases, exchangeable and nonexchangeable herbicide, by washing the roots for 1 hour following absorption. Absorption of both exchangeable and non-exchangeable 2, 4-D appeared to depend on a supply of metabolic energy which suggests that an active mechanism may be involved. A possible conclusion is that 2, 4-D is absorbed by roots by an adsorption mechanism and that energy is required to maintain the integrity of the absorbing surfaces of the cell. In contrast, absorption of Monuron was independent of an energy supply. It is concluded that the bulk of the Monuron absorbed was taken up passively by diffusion.  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