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 27 毫秒
1.
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   

2.
In maize chloroplasts, the ratio of HCO3 (anion) binding sites to high-affinity atrazine binding sites is unity. In the dark, atrazine noncompetitively inhibits the binding of half of the HCO3 to the photosystem II (PSII) complexes. The inhibition of binding saturates at 5 micromolar atrazine, little inhibition is seen at 0.5 micromolar atrazine, although the high-affinity herbicide binding sites are nearly filled at this concentration. This means that HCO3 and atrazine interact noncompetitively at a specific low-affinity herbicide binding site that exists on a portion of the PSII complexes. Light abolishes the inhibitory effects of atrazine on HCO3 binding. Based on the assumption that there is one high-affinity atrazine binding site per PSII complex, we conclude that there is also only one binding site for HCO3 with a dissociation constant near 80 micromolar. The location of the HCO3 binding site, and the low-affinity atrazine binding site, is not known.  相似文献   

3.
We examined the ability of a soil bacterium, Agrobacterium radiobacter J14a, to degrade the herbicide atrazine under a variety of cultural conditions, and we used this bacterium to increase the biodegradation of atrazine in soils from agricultural chemical distribution sites. J14a cells grown in nitrogen-free medium with citrate and sucrose as carbon sources mineralized 94% of 50 μg of [14C-U-ring]atrazine ml−1 in 72 h with a concurrent increase in the population size from 7.9 × 105 to 5.0 × 107 cells ml−1. Under these conditions cells mineralized the [ethyl-14C]atrazine and incorporated approximately 30% of the 14C into the J14a biomass. Cells grown in medium without additional carbon and nitrogen sources degraded atrazine, but the cell numbers did not increase. Metabolites produced by J14a during atrazine degradation include hydroxyatrazine, deethylatrazine, and deethyl-hydroxyatrazine. The addition of 105 J14a cells g−1 into soil with a low indigenous population of atrazine degraders treated with 50 and 200 μg of atrazine g−1 soil resulted in two to five times higher mineralization than in the noninoculated soil. Sucrose addition did not result in significantly faster mineralization rates or shorten degradation lag times. However, J14a introduction (105 cells g−1) into another soil with a larger indigenous atrazine-mineralizing population reduced the atrazine degradation lag times below those in noninoculated treatments but did not generally increase total atrazine mineralization.  相似文献   

4.
T. Abe  T. Kameya 《Planta》1986,169(2):289-291
Flower formation in 25-d-old seedling of Asparagus officinalis L was increased by soaking the seeds for 12 d in solutions of atrazine (2-chloro-4-ethylamino-6-isopropylamino-1,3,5-triazine) or diuron (3-(3,4-dichlorophenyl)-1,1-dimethylurea), two herbicides known to affect electron flow of photosystem II, from approx. 4% in controls to 37% in treatments with 400 M of atrazine or about 100 M diuron. Both herbicides inhibited plant growth, but treatment of seeds with other growth inhibitors (absicic acid, (2-chloroethyl)trimethylammonium chloride, polyethylene glycol) did not affect flower formation although inhibiting seed germination and plant growth. The increase of flower formation by the two herbicides permits identification of the sex of Asparagus plants at an early developmental stage.  相似文献   

5.
This study was prompted by increasing concerns about ecological damage and human health threats derived by persistent contamination of water and soil with herbicides, and emerging of bio-sensing technology as powerful, fast and efficient tool for the identification of such hazards. This work is aimed at overcoming principal limitations negatively affecting the whole-cell-based biosensors performance due to inadequate stability and sensitivity of the bio-recognition element. The novel bio-sensing elements for the detection of herbicides were generated exploiting the power of molecular engineering in order to improve the performance of photosynthetic complexes. The new phenotypes were produced by an in vitro directed evolution strategy targeted at the photosystem II (PSII) D1 protein of Chlamydomonas reinhardtii, using exposures to radical-generating ionizing radiation as selection pressure. These tools proved successful to identify D1 mutations conferring enhanced stability, tolerance to free-radical-associated stress and competence for herbicide perception. Long-term stability tests of PSII performance revealed the mutants capability to deal with oxidative stress-related conditions. Furthermore, dose-response experiments indicated the strains having increased sensitivity or resistance to triazine and urea type herbicides with I50 values ranging from 6×10−8 M to 2×10−6 M. Besides stressing the relevance of several amino acids for PSII photochemistry and herbicide sensing, the possibility to improve the specificity of whole-cell-based biosensors, via coupling herbicide-sensitive with herbicide-resistant strains, was verified.  相似文献   

6.
One hypothesis of triazine-type herbicide action in photosynthetic material is that the herbicide molecule competes with a secondary quinone acceptor, B, for a binding site at the reaction center of photosystem II. The binding affinity of B has been suggested to change with its level of reduction, being most strongly bound in its semiquinone form. To test this hypothesis, [14C]atrazine binding studies have been carried out under different photochemically induced levels of B reduction in Pisum sativum. It is found that herbicide binding is reduced in continuously illuminated samples compared to dark-adapted samples. Decreased binding of atrazine corresponds to an increase in the semiquinone form of B. With flash excitation, the herbicide binding oscillates with a cycle of two, being low on odd-numbered flashes when the amount of semiquinone form of B is greatest. Treatment with NH2OH was found to significantly decrease the strength of herbicide binding in the dark as well as stop the ability of p-benzoquinone to oxidize the semiquinone form of B. It is suggested that the mode of action of NH2OH is disruption of quinones or their environment on both the oxidizing and reducing sides of photosystem II. Herbicide binding was found to be unaltered under conditions when p-benzosemiquinone oxidation of the reduced primary acceptor, Q, is herbicide insensitive; weak herbicide binding cannot explain this herbicide insensitivity. It is concluded that the quinone-herbicide competition theory of herbicide action is correct. Also, since quinones are lipophilic the importance of the lipid composition of the thylakoid membrane in herbicide interactions is stressed.  相似文献   

7.
A photolabile analog of N-1-naphthylphthalamic acid (NPA), 5′-azido-N-1-naphthylphthalamic acid (Az-NPA), has been synthesized and characterized. This potential photoaffinity label for the plasma membrane NPA binding protein competes with [3H]NPA for binding sites on Curcurbita pepo L. (zucchini) hypocotyl cell membranes with K0.5 = 2.8 × 10−7 molar. The K0.5 for NPA under these conditions is 2 × 10−8 molar, indicating that the affinity of Az-NPA for the membranes is only 14-fold lower than NPA. While the binding of Az-NPA to NPA binding sites is reversible in the dark, exposure of the Az-NPA treated membranes to light results in a 30% loss in [3H]NPA binding ability. Pretreatment of the membranes with NPA protects the membranes against photodestruction of [3H]NPA binding sites by Az-NPA supporting the conclusion that Az-NPA destroys these sites by specific covalent attachment.  相似文献   

8.
Effect of several parameters on inhibition of potato (Solanum tuberosum) invertase by its endogenous proteinaceous inhibitor was determined using homogeneous preparations of both proteins. The inhibitor and invertase formed an inactive complex with an observed association rate constant at pH 4.70 and 37°C of 8.82 × 102 per molar per second and a dissociation rate constant of 3.3 × 10−3 per minute. The inhibitor appeared to bind to invertase in more than one step. Initial interaction (measured by loss of invertase activity) was rapid, relatively weak, readily reversible (Ki of 2 × 10−6 molar) and noncompetitive with substrate at pH 4.70. Initial interaction was probably followed by isomerization to a tighter (Ki of 6.23 × 10−8 molar) complex, which dissociated slowly with a half-time of 3.5 hour. Interaction between enzyme and inhibitor appeared to be of ionic character and essentially pH independent between pH 3.5 and 7.4.  相似文献   

9.
The effects of various calcium ion antagonists and ion transport inhibitors on photosynthetic O2 evolution of corals, isolated zooxanthellae, sea anemone tentacles, and Chlorococcum oleofaciens were measured. Only the phenothiazine drugs were effective at inhibiting photosynthesis. Trifluoperazine, a calcium ion antagonist drug, inhibited at low concentrations, with 10−4 molar and 8 × 10−6 molar completely abolishing photosynthesis in the intact corals and isolated zooxanthellae, respectively. Net photosynthetic O2 evolution of C. oleofaciens was eliminated by concentrations of trifluoperazine as low as 2.8 × 10−5 molar.  相似文献   

10.
The mechanism of action of the p-nitrodiphenyl ether herbicides has remained ambiguous because of conflicting reports in the literature. The diphenyl ether herbicide oxyfluorfen causes a light induced consumption of oxygen which resembles the electron acceptor reaction of paraquat. However, this reaction is not linked to the transport of electrons through photosystem I. This conclusion is based on the observation that the rate of oxygen consumption, in the presence of oxyfluorfen, does not demonstrate a first order rate dependence on light intensity. Using the bleaching of N,N-dimethyl p-nitrosoaniline as a specific detector of singlet oxygen, we demonstrate that oxyfluorfen is a potent generator of this toxic radical. The production of singlet oxygen occurs in the presence of inhibitors of photosynthetic electron transport (oxyfluorfen at 10−4 molar and paraquat) and also under temperature conditions (3°C) which prevent electron transport. This light induced reaction results in oxygen consumption and is the primary cause of lethality for oxyfluorfen. The production of singlet oxygen occurs rapidly and at low herbicide concentrations (10−9 molar). The reaction occurs without photosynthetic electron transport but does require an intact thylakoid membrane.  相似文献   

11.
Pyrophosphorylytic kinetic constants (S0.5, Vmax) of partially purified UDP-glucose- and ADP-glucose pyrophosphorylases from potato tubers were determined in the presence of various intermediary metabolites. The S0.5 of UDP-glucose pyrophosphorylase for UDP-glucose (0.17 millimolar) or pyrophosphate (0.30 millimolar) and the Vmax were not influenced by high concentrations (2 millimolar) of these substances. The most efficient activator of ADP-glucose pyrophosphorylase was 3-P-glycerate (A0.5 = 4.5 × 10−6 molar). The S0.5 for ADP-glucose and pyrophosphate was increased 3.5-fold (0.83 to 0.24 millimolar) and 1.8-fold (0.18 to 0.10 millimolar), respectively, with 0.1 millimolar 3-P-glycerate while the Vmax was increased nearly 4-fold. The magnitude of 3-P-glycerate stimulation was dependent upon the integrity of key sulfhydryl groups (−SH) and pH. Oxidation or blockage of −SH groups resulted in a marked reduction of enzyme activity. Stimulations of 3.1-, 2.9-, 4.8-, and 9.5-fold were observed at pH 7.5, 8.0, 8.5, and 9.0, respectively, in the presence of 3-P-glycerate (2 millimolar). The most potent inhibitor of ADP-glucose pyrophosphorylase was orthophosphate (I0.5 = 8.8 × 10−5. molar). This inhibition was reversed with 3-P-glycerate (1.2 × 10−4 molar), resulting in an increased I0.5 value of 1.5 × 10−3 molar. Likewise, orthophosphate (7.5 × 10−4 molar) caused a decrease in the activation efficiency of 3-P-glycerate (A0.5 from 4.5 × 10−6 molar to 6.7 × 10−5 molar). The significance of 3-P-glycerate activation and orthophosphate inhibition in the regulation of α-glucan biosynthesis in Solanum tuberosum is discussed.  相似文献   

12.
Conditions are described for isolating functional phycobilisome-thylakoid vesicles from the red alga Porphyridium cruentum. Phycobilisome-thylakoid vesicles were prepared by brief sonication and centrifugation in a medium containing 0.5 molar sucrose, 0.5 molar potassium phosphate, and 0.3 molar sodium citrate (pH 7.0). They required ferricyanide as an oxidant and had O2 evolution rates (about 450 micromoles O2 per hour per milligram chlorophyll) higher than whole cells (about 250 micromoles O2 per hour per milligram chlorophyll). Energy transfer to photosystem II chlorophyll was evident from a high F695 nanometer (−196 C) emission peak. Preparations could be stored for over 24 hours and were considerably more stable than those from the cyanobacterium Anabaena variabilis (Katoh T, E Gantt 1979 Biochim Biophys Acta 546: 383-393). In electron micrographs of negatively stained material, the active thylakoid vesicles were found covered by closely spaced phycobilisomes on their external surface. The phycobilisome number in negatively stained vesicles was 450 per square micrometer, which was in the same range as the 400 per square micrometer observed in surface sections. A cell containing 1.5 × 10−6 micrograms phycoerythrin and 1.3 × 10−6 micrograms chlorophyll was found to contain 5 to 7 × 105 phycobilisomes on a thylakoid area of 1.1 to 1.6 × 103 square micrometers.  相似文献   

13.
Atrazine, a herbicide widely used in corn production, is a frequently detected groundwater contaminant. Nine gram-positive bacterial strains able to use this herbicide as a sole source of nitrogen were isolated from four farms in central Canada. The strains were divided into two groups based on repetitive extragenic palindromic (rep)-PCR genomic fingerprinting with ERIC and BOXA1R primers. Based on 16S ribosomal DNA sequence analysis, both groups were identified as Nocardioides sp. strains. None of the isolates mineralized [ring-U-14C]atrazine. There was no hybridization to genomic DNA from these strains using atzABC cloned from Pseudomonas sp. strain ADP or trzA cloned from Rhodococcus corallinus. S-Triazine degradation was studied in detail in Nocardioides sp. strain C190. Oxygen was not required for atrazine degradation by whole cells or cell extracts. Based on high-pressure liquid chromatography and mass spectrometric analyses of products formed from atrazine in incubations of whole cells with H218O, sequential hydrolytic reactions converted atrazine to hydroxyatrazine and then to the end product N-ethylammelide. Isopropylamine, the putative product of the second hydrolytic reaction, supported growth as the sole carbon and nitrogen source. The triazine hydrolase from strain C190 was isolated and purified and found to have a Km for atrazine of 25 μM and a Vmax of 31 μmol/min/mg of protein. The subunit molecular mass of the protein was 52 kDa. Atrazine hydrolysis was not inhibited by 500 μM EDTA but was inhibited by 100 μM Mg, Cu, Co, or Zn. Whole cells and purified triazine hydrolase converted a range of chlorine or methylthio-substituted herbicides to the corresponding hydroxy derivatives. In summary, an atrazine-metabolizing Nocardioides sp. widely distributed in agricultural soils degrades a range of s-triazine herbicides by means of a novel s-triazine hydrolase.  相似文献   

14.
A 2-gram fresh weight inoculum of bromegrass (Bromus inermis Leyss. culture BG970) cell suspension culture treated with 7.5 × 10−5 molar abscisic acid (ABA) for 7 days at 25°C survived slow cooling to −60°C. Over 80% of the cells in ABA treated cultures survived immersion in liquid N2 after slow cooling to −40 or −60°C. In contrast, a 6-gram fresh weight inoculum only attained a hardiness level of −28°C after 5 days of ABA treatment. Ethanol (2 × 10−2 molar) added to the culture medium at the time of ABA addition, inhibited the freezing tolerance of bromegrass cells by 25°C. A 6-gram inoculum of both control and ABA treated bromegrass cells altered the pH of the medium more than a 2-gram inoculum. ABA inhibited the increase in fresh weight of bromegrass by 20% after 4 days. Both control and ABA (10−4 molar) treated alfalfa cells (Medicago sativa L.) grown at 25°C hardened from an initial LT50 of −5°C to an LT50 of −23°C by the third to fifth day after subculture. Thereafter, the cells dehardened but the ABA treated cells did not deharden to the same level as the control cells. ABA inhibited the increase in fresh weight of alfalfa by 50% after 5 days.  相似文献   

15.
Salicylic and acetylsalicylic acids applied on excised leaves of Cassia fasciculata modify the dark-induced (scotonastic) and light-induced (photonastic) leaflet movements. They inhibit the scotonastic movements in a dose-dependent manner from 1 × 10−4 to 1 × 10−3 molar and they promote the photonastic movements at an optimum concentration of 5 × 10−4 molar. These results suggest that these phenolic compounds do not act specifically on the K+ uptake, which was shown to be inhibited by their action on other materials.  相似文献   

16.
Shi Y  Wang T  Li Y  Darmency H 《Genetics》2008,180(2):969-975
Developing genetically modified crop plants that are biologically contained could reduce significantly the potential spread of transgenes to conventional and organic crop plants and to wild or weedy relatives. Among several strategies, the hereditary mode of transmission of transgenes, whether dominant, recessive, or maternal, could play a major role in interspecific gene flow. Here we report on the gene flow between foxtail millet (Setaria italica), an autogamous crop, and its weedy relative, S. viridis, growing within or beside fields containing the three kinds of inherited herbicide resistance. Over the 6-year study, in the absence of herbicide selection, the maternal chloroplast-inherited resistance was observed at a 2 × 10−6 frequency in the weed populations. Resistant weed plants were observed 60 times as often, at 1.2 × 10−4 in the case of the nuclear recessive resistance, and 190 times as often, at 3.9 × 10−4 in the case of the dominant resistance. Because the recessive gene was not expressed in the first-generation hybrids, it should be more effective than dominant genes in reducing gene flow under normal agricultural conditions where herbicides are sprayed because interspecific hybrids cannot gain from beneficial genes.  相似文献   

17.
We purified a secreted fungal laccase from filtrates of Gaeumannomyces graminis var. tritici cultures induced with copper and xylidine. The active protein had an apparent molecular mass of 190 kDa and yielded subunits with molecular masses of 60 kDa when denatured and deglycosylated. This laccase had a pI of 5.6 and an optimal pH of 4.5 with 2,6-dimethoxyphenol as its substrate. Like other, previously purified laccases, this one contained several copper atoms in each subunit, as determined by inductively coupled plasma spectroscopy. The active enzyme catalyzed the oxidation of 2,6-dimethoxyphenol (Km = 2.6 × 10−5 ± 7 × 10−6 M), catechol (Km = 2.5 × 10−4 ± 1 × 10−5 M), pyrogallol (Km = 3.1 × 10−4 ± 4 × 10−5 M), and guaiacol (Km = 5.1 × 10−4 ± 2 × 10−5 M). In addition, the laccase catalyzed the polymerization of 1,8-dihydroxynaphthalene, a natural fungal melanin precursor, into a high-molecular-weight melanin and catalyzed the oxidation, or decolorization, of the dye poly B-411, a lignin-like polymer. These findings indicate that this laccase may be involved in melanin polymerization in this phytopathogen’s hyphae and/or in lignin depolymerization in its infected plant host.  相似文献   

18.
Genome-wide association studies (GWASs) of follicular lymphoma (FL) have previously identified human leukocyte antigen (HLA) gene variants. To identify additional FL susceptibility loci, we conducted a large-scale two-stage GWAS in 4,523 case subjects and 13,344 control subjects of European ancestry. Five non-HLA loci were associated with FL risk: 11q23.3 (rs4938573, p = 5.79 × 10−20) near CXCR5; 11q24.3 (rs4937362, p = 6.76 × 10−11) near ETS1; 3q28 (rs6444305, p = 1.10 × 10−10) in LPP; 18q21.33 (rs17749561, p = 8.28 × 10−10) near BCL2; and 8q24.21 (rs13254990, p = 1.06 × 10−8) near PVT1. In an analysis of the HLA region, we identified four linked HLA-DRβ1 multiallelic amino acids at positions 11, 13, 28, and 30 that were associated with FL risk (pomnibus = 4.20 × 10−67 to 2.67 × 10−70). Additional independent signals included rs17203612 in HLA class II (odds ratio [ORper-allele] = 1.44; p = 4.59 × 10−16) and rs3130437 in HLA class I (ORper-allele = 1.23; p = 8.23 × 10−9). Our findings further expand the number of loci associated with FL and provide evidence that multiple common variants outside the HLA region make a significant contribution to FL risk.  相似文献   

19.
PS II membrane fragments produced from higher plant thylakoids by Triton X-100 treatment exhibit strong photoinhibition and concomitant fast degradation of the D1 protein. Involvement of (molecular) oxygen is necessary for degradation of the D1 protein.The herbicides atrazine and diuron, but not ioxynil, partly protect the D1 protein against degradation. Binding of atrazine to the D1 protein is necessary to protect the D1 polypeptide, as shown with PS II membrane fragments from an atrazine-resistant biotype of Chenopodium album which are protected by diuron not by atrazine.Abbreviations atrazine 2-chloro-4-ethylamino-6-isopropylamino-1,3,5-triazine - Chl chlorophyll, diuron - (DCMU) 3-(3,4-dichlorophenyl)-1,1-dimethylurea - DMBQ 2,5-dimethyl-p-benzoquinone - DCIP 2,6-dichlorophenol indophenol - DPC diphenylcarbazide - ioxynil 4-cyano-2,6-diiodophenol - kb binding constant - Mes 4-morpholinoethanesulfonic acid - P-680 reaction-center chlorophyll a of photosystem-II - PAGE polyacrylamide gel electrophoresis - PS II photosystem-II - QA and QB primary and secondary quinone electron acceptors - Z electron donor to the photosystem-II reaction center - SDS sodium dodecylsulfate - Tricine N-2-hydroxy-1,1-bis(hydroxymethyl)ethylglycine  相似文献   

20.
Cobalamin is essential for growth of Euglena gracilis and photosynthesis. Methylcobalamin in Euglena chloroplasts (Y Isegawa, Y Nakano, S Kitaoka, 1984 Plant Physiol 76: 814-818) functions as a coenzyme of methionine synthetase. The requirement of cobalamin for photosynthesis appeared remarkably high in Euglena grown under the dark-precultured condition. The required amount of cobalamin for normal photosynthetic activity was 7.4 × 10−11 molar, while 7.4 × 10−10 molar cobalamin was required for normal growth. The lowered photosynthetic activity in cobalamin-limited cells was restored 20 hours after feeding cyanocobalamin or methionine to cobalamin-limited cells. Lowering of photosynthetic activity was due to loss of photosystem I activity. This photosynthetic activity was recovered after supplementation by methionine or cobalamin. The results suggest that methionine serves for the stabilization of photosystem I. This paper is the first report of the physiological function of cobalamin in chloroplasts of photosynthetic eukaryotes.  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