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 343 毫秒
1.
Cyanobacterium Plectonema boryanum IU 594 and cyanophage LPP-1 were used as indicator organisms in a bioassay of 16 pesticides. Experiments such as spot tests, disk assays, growth curves, and one-step growth experiments were used to examine the effects of pesticides on the host and virus. Also, experiments were done in which host or virus was incubated in pesticide solutions and then assayed for PFU. P. boryanum was inhibited by four herbicides: 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU), 1,1-dimethyl-3-(alpha, alpha,alpha-trifluoro-m-tolyl)urea ( Fluometeron ), 2-chloro-4-(ethylamino)-6-(isopropylamino)-s-triazine (Atrazine), 2-(ethylamino)-4-(isopropylamino)-6-(methylthio)-s-triazine ( Ametryn ). One insecticide, 2-methyl-2-(methylthio)-propionaldehyde O-( methylcarbamoyl )oxime (Aldicarb), also inhibited the cyanobacterium. Two insecticides inactivated LPP-1, O,O-dimethyl phosphorodithioate of diethyl mercaptosuccinate (malathion) and Isotox . Isotox is a mixture of three pesticides: S-[2-( ethylsulfinyl )ethyl]O,O-dimethyl phosphorothioate ( Metasystox -R), 1-naphthyl methylcarbamate ( Sevin ) and 4,4'-dichloro-alpha- (trichloromethyl) benzhydrom ( Kelthane ). Two pesticide-resistant strains of P. boryanum were isolated against DCMU and Atrazine. These mutants showed resistance to all four herbicides, which indicates a relationship between these phototoxic chemicals. The results indicate that P. boryanum may be a useful indicator species for phototoxic agents in bioassay procedures.  相似文献   

2.
Effect of Selected Herbicides on Bacterial Growth Rates   总被引:1,自引:1,他引:0       下载免费PDF全文
Specific growth rate constants were used to evaluate the effects of selected herbicides on Erwinia carotovora, Pseudomonas fluorescens, and Bacillus sp. Comparison of growth rate constants permitted the identification of either stimulatory or inhibitory effects of these substances. E. carotovora was inhibited by 6,7-dihydrodipyrido(1,2-a:2'-c)pyrazinediium (diquat) and 4-hydroxy-3,5-diiodobenzonitrile (ioxynil) at 25 mug/ml; 1,1'-dimethyl-4,4'-bipyridinium (paraquat) at 50 mug/ml; and pentachlorophenol (PCP) at 10 mug/ml. P. fluorescens was inhibited by paraquat and PCP at 25 mug/ml and by 4-amino-3,5,6-trichloropicolinic acid (picloram) at 50 mug/ml. Stimulation of P. fluorescens was observed with 4-(methylsulfonyl)-2,6-dinitro-N,N-dipropylaniline (nitralin) at 25 mug/ml. The Bacillus species was inhibited by diquat (25 mug/ml), ioxynil (10 mug/ml), and paraquat and PCP (5 mug/ml). No significant effect of 2-chloro-4-(ethylamino)-6-(isopropylamino)-s-triazine (atrazine), 3-(3,4-dichlorophenyl)-1,1-dimethylurea (diuron), alpha,alpha,alpha-trifluoro-2,6-dinitro-N,N-dipropyl-p-toluidine (trifluralin), or 1,1-dimethyl-3-(alpha,alpha,alpha-trifluoro-m-tolyl)urea (fluometuron) on growth rates of the bacteria was observed at 25 and 50 mug/ml.  相似文献   

3.
Atrazine (2-chloro-4-(ethylamino)-6-(isopropylamino)-s-triazine) is one of the most commonly used herbicides to control grasses and weeds. The widespread contamination and persistence of atrazine residues in the environment has resulted in human exposure. Vitamin E is a primary antioxidant that plays an important role in protecting cells against toxicity by inactivating free radicals generated following pesticides exposure. The present study was undertaken to investigate the protective effect of vitamin E against atrazine-induced genotoxicity. Three different methods: gel electrophoresis, comet assay and micronucleus test were used to assess the atrazine-induced genotoxicity and to evaluate the protective effects of vitamin E. Atrazine was administered to male rats at a dose of 300 mg/kg body weight for a period of 7, 14 and 21 days. There was a significant increase (P<0.001) in tail length of comets from blood and liver cells treated with atrazine as compared to controls. Co-administration of vitamin E (100 mg/kg body weight) along with atrazine resulted in decrease in tail length of comets as compared to the group treated with atrazine alone. Micronucleus assay revealed a significant increase (P<0.001) in the frequency of micronucleated cells (MNCs) following atrazine administration. In the animals administrated vitamin E along with atrazine there was a significant decrease in percentage of micronuclei as compared to atrazine treated rats. The increase in frequency of micronuclei in liver cells and tail length of comets confirm genotoxicity induced by atrazine in blood and liver cells. In addition, the findings clearly demonstrate protective effect of vitamin E in attenuating atrazine-induced DNA damage.  相似文献   

4.
Atrazine (2-chloro-4 ethylamino-6-(isopropylamino)-s-triazine) is one of the most widely used herbicides. Fourier transform infrared spectroscopy, differential scanning calorimetry and fluorescence polarization of 1,6-diphenyl-1,3,5-hexatriene (DPH) and of its derivative 1-(4-trimethylaminophenyl)-6-phenyl-1,3,5-hexatriene (TMA-DPH) were used to study the interaction of atrazine with dipalmitoyl phosphatidylcholine liposomes used as a model for biological membranes. The results show that atrazine does not perturb the hydrophobic core of the lipid bilayer and suggest that the herbicide localizes near the glycerol backbone of the lipid.  相似文献   

5.
The binding characteristics and the inhibitory power of atrazine and DCMU towards uncoupled electron flow activity were studied in acyl lipid-depleted thylakoid membranes from atrazine-susceptible and-resistant biotypes of Solanum nigrum L. For this purpose, phospholipase A2 from Vipera russelli and the lipase from Rhizopus arrhizus were used to obtain a selective lipid class (phospholipids or galactolipids) depletion which was restricted to the outer monolayer. Neither phospholipid nor galactolipid removal affected the dissociation constant and the number of binding sites of atrazine. In contrast, the dissociation constant of DCMU was increased in phospholipid-depleted thylakoid membranes but remained unchanged after galactolipid depletion. The number of DCMU binding sites decreased significantly after both lipase treatments, but only in the resistant biotype. The inhibitory effectiveness of the herbicide was either decreased or increased (to different extents) depending on the lipid class which was removed from the membrane and on the biotype considered. These results are discussed with reference to the possible conformational changes of the 32 kDa herbicide-binding polypeptide occurring after lipase treatments.Abbreviations Atrazine 2-chloro-4-(ethylamino)-6-(isopropylamino)-s-triazine - BSA bovine serum albumin - DCMU diuron, 3-(3,4-dichlorophenyl)-1,1-dimethylurea - DGDG digalactosyldiacylglycerol - LRa lipase from Rhizopus arrhizus - MGDG monogalactosyldiacylglycerol - PC phosphatidylcholine - PG phosphatidylglycerol - PLA2 phospholipase A2 - R atrazine-resistant - S atrazinesusceptible  相似文献   

6.
Effect of two photosynthetic inhibitor herbicides, atrazine (both purified and formulated) and [3-(3,4-dichlorophenyl)-1,1-dimethyl urea] (DCMU), on the growth, macromolecular contents, heterocyst frequency, photosynthetic O2 evolution and dark O2 uptake of wild type and multiple herbicide resistant (MHR) strain of diazotrophic cyanobacterium A. variabilis was studied. Cyanobacterial strains showed gradual inhibition in growth with increasing dosage of herbicides. Both wild type and MHR strain tolerated < 6.0 mg L(-1) of atrazine (purified), < 2.0 mg L(-1) of atrazine (formulated) and < 0.4 mg L(-1) of DCMU indicating similar level of herbicide tolerance. Atrazine (pure) (8.0 mg L(-1)) and 4.0 mg L(-1) of atrazine (formulated) were growth inhibitory concentrations (lethal) for both wild type and MHR strain indicating formulated atrazine was more toxic than the purified form. Comparatively lower concentrations of DCMU were found to be lethal for wild type and MHR strain, respectively. Thus, between the two herbicides tested DCMU was more growth toxic than atrazine. At sublethal dosages of herbicides, photosynthetic O2 evolution showed highest inhibition followed by chlorophyll a, phycobhiliproteins and heterocyst differentiation as compared to carotenoid, protein and respiratory O2 uptake.  相似文献   

7.
The effect of three herbicides—DCMU (1,1-dimethyl-3- (3,4-dichlorophenyl) -urea), Simazine (2,4-bis(ethylamino)- 6-chloro-s-triazine), and Atrazine (2-chloro-4-ethylamino-6-iso-propylamino-5-triazine)—on the induction of nitrate reduc–tase and its in vivo activity was studied in detached leaves of Hordeum vulgare L. All increased both extractable nitrate reductase activity and nitrate content. The increases occurred at optimum temperatures for growth and at several concentrations of nitrate. It was also determined that the herbicides did not protect the enzyme against inactivation in vivo. Although the extractable nitrate reductase was greater, the in vivo activity of nitrate reductase was decreased in the presence of the herbicides resulting in a higher internal concentration of nitrate. Since in viva nitrate reduction is dependent upon photosynthesis it is reasonable that reduction is decreased by these known inhibitors of photosynthesis. Hence, the effect of the inhibitors on induction of nitrate reductase activity may be secondary. The higher concentration of nitrate resulting from a decreased rate of in vivo reduction in the presence of the inhibitors could conceivably be responsible for the greater corutent of nitrate reductase.  相似文献   

8.
The effects of DDT, some of its analogs, and selected cyclodiene insecticides on isolated spinach (Spinacea oleracea L.) thylakoids were identified, characterized, and compared to responses induced by selected herbicides. Except for endrin, the insecticides inhibited light-induced electron transport, altered chlorophyll fluorescence transients, and competitively displaced [14C]atrazine [2-chloro-4-(ethylamino)-6-(isopropylamino)-s-triazine], a known photosystem II inhibitor, from the membranes. The insecticides appeared to act at, or near B, the secondary electron acceptor of photo-system II. Binding of DDT and dieldrin was estimated at 900 and 2200 molecules, respectively, per photosynthetic unit (490 chlorophyll molecules). The insecticides also inhibited valinomycin-induced swelling of the thylakoid membrane. Whereas inhibition of electron transport can be attributed to interaction by the insecticides with a proteinaceous component of the thylakoid membrane, interference with the action of valinomycin may involve interaction with lipoidal constituents of the membrane.  相似文献   

9.
A new mutant of the psbA gene conferring resistance to 2-chloro-4-(ethylamino)-6-(isopropylamino)-s-triazine (atrazine) was obtained by selection of photomixotrophic tobacco (Nicotiana tabacum cv Samsun NN) cells. The 264th codon AGT (serine) in the wild psbA gene was changed to ACT (threonine) in these mutant tobacco cells. All other higher plants resistant to atrazine exhibit a change to GGT (glycine) in this codon. Measurements of Hill reaction activity and chlorophyll fluorescence showed that the threonine 264-containing plastoquinone serving as secondary stable electron acceptor of PSII (QB protein) had not only strong resistance to triazine-type herbicides but also moderate resistance to substituted urea-type herbicides. Threonine-type QB protein showed especially strong resistance to methoxylamino derivatives of the substituted urea herbicides. The projected secondary structures of the mutant QB proteins indicate that the cross-resistance of threonine 264 QB protein to triazine and urea herbicides is mainly due to a conformational change of the binding site for the herbicides. However, the glycine 264 QB protein is resistant to only triazine herbicides because of the absence of an hydroxyl group and not because of a conformational change.  相似文献   

10.
Bicarbonate (or carbon dioxide) is required for electron transport in isolated broken pea chloroplasts. The site of action of the bicarbonate ion is between the primary electron acceptor of Photosystem 2, Q, and the plastoquinone pool. After trypsin treatment the Hill reaction with ferricyanide does not require bicarbonate. Photosystem 2 inhibiting herbicides act also at this site. Therefore, a possible interaction of bicarbonate and these herbicides in their effect on photosynthetic electron transport was studied.
The reciprocal of the Hill reaction rate in CO2-depleted chloroplasts was plotted against the reciprocal of added bicarbonate concentration in the absence and in the presence of 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU), 2-methoxy-4,6-bis (ethylamino)-1,3,5-triazine (simeton) or 4,6-dinitro- o -cresol (DNOC). From these Lineweaver-Burk plots we concluded that DCMU and simeton inhibit both bicarbonate binding and Vmax. There is a purely competitive inhibition of bicarbonate binding by DNOC. We suggest that DNOC may exert its inhibition of electron transport by removing bicarbonate from its binding site.
In isolated thylakoid membranes of Synechococcus leopoliensis we did not find a bicarbonate effect nor inhibition by DNOC after Q, indicating that in the thylakoids of this blue-green alga the binding site for bicarbonate and DNOC between Q and plastoquinone is absent.  相似文献   

11.
In order to elucidate the mode of action of some herbicides, effect of several anilide type herbicides on the respiration of yeast cells was studied. The results obtained were as follows: 1) DCPA (3,4-dichloropropionanilide) and DCMU (3-(3,4-dichlorophenyl)-1,1- dimethylurea), the powerful inhibitors of the Hill reaction in photosynthesis, inhibited the oxygen uptake of yeast cells at low concentrations. 2) DCPA and DCMU inhibited the enzymic reduction of cytochrome-c by the yeast cell-free preparation, but not the reduction of dye. 3) The oxidation of cytochrome-b was inhibited in the yeast cells treated with DCPA or DCMU.  相似文献   

12.
Growth of Anacystis in high light in the presence of sublethal concentrations of DCMU-type inhibitors leads to an increased synthesis of phycocyanin paralleled by a reduced rate of 35S methionine incorporation into the D1 protein compared to the high light controls, as is characteristic for naturally-induced shade phenotype. On the contrary, sun phenotype is characterized by a low rate of antenna synthesis, but a high rate of 35S methionine incorporation into the D1 protein.Room temperature excitation spectra of 684 nm fluorescence emission clearly demonstrate the participation of the extraordinarily high concentration of phycocyanin in artificially shade-adapted cells in excitation energy transfer to chlorophyll.It could be shown that the development of shade-type appearance is not simply the consequence of an imbalance in electron transport, since an addition of thiosulphate to cultures growing in high light in the presence of DCMU-type inhibitors can only partially prevent or revert the change from sun to artificial-herbicide-induced-shade phenotype. This is regarded as evidence that the dynamic herbicide-binding D1 protein itself may play a role as a light meter in the process of natural shade adaptation, the rate of its degradation and resynthesis possibly giving the signal for the adaptive reorganization of the photosynthetic apparatus. The chain of signal transduction remains to be established.Abbreviations atrazine 2-chloro-4-(ethylamino)-6-(isopropylamino)-s-triazine - chl chlorophyll - D1 reaction center polypeptide carrying the secondary plastoquinone electron acceptor of PS II - DCMU 3-(3,4-dichlorophenyl)-1,1-dimethylurea - PAGE polyacrylamide gel electrophoresis - PAR photosynthetically active radiation - PC phycocyanin - PCC Pasteur Culture Collection - PS photosystem - QB secondary plastoquinone electron acceptor of PS II - SAUG Sammlung von Algenkulturen am Pflanzenphysiologischen Institut der Universtität Göttingen - SDS sodium dodecyl sulphate Dedicated to Professor Wilhelm Menke on the occasion of his 80th birthday.  相似文献   

13.
Stem cuttings of alligatorweed [Alternanthera philoxeroides (Mart.) Griseb.] were subjected to various light and chemical inhibitor treatments to obtain information about the physiological nature of the hypoxic quiescence induced by dark submergence. White or red light at 40 μE m?2 s?1 stimulated growth from submerged stem cuttings but far-red at 5 μE M?2 s?1 did not. Photo-system II inhibitors, such as 3-(3,4-dichlorophenyl) 1,1-dimethylurea (DCMU) at 1.4 × 10?5M or 2-chloro-4,6-bis(ethylamino)-s-triazine (simazine) at 10?5M, completely inhibited the growth that normally occurs in a submerged state under continuous white light at 40 μE m?2 s?1. These concentrations of DCMU or simazine did not reduce nonphotosynthetic growth from underwater nodes of emersed stem cuttings partially exposed to air in the light for 1 week. Hydrogen peroxide at 50 mg/1 added every other day partially relieved the simazine-induced inhibition of growth from submerged, illuminated cuttings. These data indicated that sprouting and early growth of submerged, illuminated alligatorweed depended on the oxygen produced by photosystem II to support respiration and to overcome hypoxic quiescence.  相似文献   

14.
[14C]Atrazine (2-chloro-4-[ethylamino]-6-[isopropylamino]-s-triazine) and [14C]glyphosate (N-[phosphonomethyl]glycine) were xylem fed to sunflower shoots at 100 micromolar for 1 hour in the light, then placed in the dark at 100% relative humidity for 1, 4, 7, or 10 hours. The distribution of atrazine and glyphosate between shoot parts, in the leaves, and between the aoplast and symplast of the leaf was determined. The apoplastic concentrations and distribution patterns of atrazine and glyphosate in the leaves were evaluated using a pressure dehydration technique, our results were compared to the previously reported distribution patterns of the naturally occurring apoplastic leaf solutes, and the apoplastic dye PTS (trisodium 3-hydroxy-5,8,10-pyrenetrisulfonate). The pattern of atrazine and glyphosate distribution in the shoot, and between the leaf apoplast and symplast, was found to reflect the potential of these herbicides to enter the shoot symplast. The results of this study are discussed with respect to current theories of xenobiotic transport in plants, and have been found to be consistent with the intermediate permeability hypothesis for xenobiotic transport.  相似文献   

15.
Photosynthetic acclimation to temperature and irradiance was studied in the filamentous, non-heterocystous cyanobacterium Plectonema boryanum UTEX 485. Growth rates of this cyanobacterium measured at ambient CO2 were primarily influenced by temperature with minimal effects of irradiance. Both growth temperature and irradiance affected linolenic (18:3) and linoleic acid (18:2) levels in the four major lipid classes in an independent but additive manner. In contrast, photosynthetic acclimation was not due to either growth temperature or irradiance per se, but rather, due to the interaction of these environmental factors. P. boryanum grown at low temperature and moderate irradiance mimicked cells grown at high light. Compared to cells grown at either 29 degrees C/150 micromol m(-2) s(-1) (29/150) or 15/10, P. boryanum grown at either 15/150 or 29/750 exhibited: (1) reduced cellular levels of Chl a and phycobilisomes (PBS), and concomitantly higher content of an orange-red carotenoid, myxoxanthophyll; (2) higher light saturated rates (Pmax) when expressed on a Chl a basis but lower apparent quantum yields of oxygen evolution and (3) enhanced resistance to high light stress. P. boryanum grown at 15/150 regained normal blue-green pigmentation within 16 h after a temperature shift to 29 degrees C at a constant irradiance of 150 micromol m(-2) s(-1). DBMIB and KCN but not DCMU and atrazine partially inhibited the change in myxoxanthophyll/Chl a ratio following the shift from 15 to 29 degrees C. We conclude that P. boryanum responds to either varying growth temperature or varying growth irradiance by adjusting the ability to absorb light through decreasing the cellular contents of Chl a and light-harvesting pigments and screening of excessive light by myxoxanthophyll predominantly localized in the cell wall/cell membrane to protect PSII from over-excitation. The possible role of redox sensing/signalling for photosynthetic acclimation of cyanobacteria to either temperature or irradiance is discussed.  相似文献   

16.
Nitrogenase Activity and Photosynthesis in Plectonema boryanum   总被引:3,自引:1,他引:2       下载免费PDF全文
Nitrogen-starved Plectonema boryanum 594 cultures flushed with N(2)/CO(2) or A/CO(2) (99.7%/0.3%, vol/vol) exhibited nitrogenase activity when assayed either by acetylene reduction or hydrogen evolution. Oxygen evolution activities and phycocyanin pigments decreased sharply before and during the development of nitrogenase activity, but recovered in the N(2)/CO(2) cultures after a period of active nitrogen fixation. Under high illumination, the onset of nitrogenase activity was delayed; however, the presence of 3-(3, 4-dichlorophenyl)-1, 1-dimethylurea (DCMU) eliminated this lag. Oxygen was a strong and irreversible inhibitor of nitrogenase activity at low (>0.5%) concentrations. In the dark, low oxygen tensions (0.5%) stimulated nitrogenase activity (up to 60% of that in the light), suggesting a limited but significant respiratory protection of nitrogenase at low oxygen tensions. DCMU was not a strong inhibitor of nitrogenase activity. A decrease in nitrogenase activity after a period of active nitrogen fixation was observed in the N(2)/CO(2-), but not in the A/CO(2-), flushed cultures. We suggest that this decrease in nitrogenase activity is due to exhaustion of stored substrate reserves as well as inhibition by the renewed oxygen evolution of the cultures. Repeated peaks of alternating nitrogenase activity and oxygen evolution were observed in some experiments. Our results indicate a temporal separation of these basically incompatible reactions in P. boryanum.  相似文献   

17.
The objectives of this research were to observe plant response to vegetable oil sprays and to learn if vegetable oils — sunflower, soybean, linseed, or camelina — can replace petroleum oil as an herbicide adjuvant. Vegetable oils were sprayed on grain sorghum [Sorghum bicolor (L.) Moench.] and sunflower (Helianthus annuus L.) at 4 7 L/ha and were neither harmful nor beneficial to the crops. When used as an adjuvant postemergence with 1.68 kg/ha of atrazine [2-chloro-4-(ethylamino)-6-(isopropylamino)-s-triazine], 2.3 L/ha of vegetable oil were as effective in weed control as either 2.3 or 9.4 L/ha of petroleum oil. All atrazine treatments gave complete control of dicotyledonous weeds. But vegetable oil adjuvants with atrazine at 1.68 kg/ha gave significantly more grass weed control than atrazine alone at 2.24 kg/ha in two of five trials and were equally good in the other trials. The cost of the adjuvant is less than that of the atrazine replaced, and the initial herbicide residue in the soil is reduced by 25%.  相似文献   

18.
Series of phenoxypyrimidines and phenoxytriazines were prepared to be evaluated as herbicides. Among them, 2-(2,6-dichlorophenoxy)-pyrimidine (XV), 2-phenoxy-4,6-dimethyl- pyrimidine (XVII), 2-(3-methyl-4-chlorophenoxy)-4,6-bis(ethylamino)-5-triazine (LIV), 2-(2,4-dichlorophenoxy)-4,6-bis(ethylamino)-s-triazine (LVIII), and 2-(2,6-dichlorophenoxy)-4,6-bis(ethylamino)-s-triazine (LX) showed high pre-emergent herbicidai activity to radish. On the other hand, 2-chloro-4-(2,6-dichlorophenoxy)-6-methylpyrimidine (XXX) revealed high efficiency to millet. Some structure-activity relationship is discussed.  相似文献   

19.
Studies with Eudorina elegans L. were done to provide additional information on the effect of phenyl urea herbicides on phytoplankton. Colonies were grown in various concentrations of DCMU, 3(3,4-dichlorophenyl)-1,1-dimethylurea. DCMU at 10-5 m induced an algicidic response. DCMU at 10-7 M and 10-9 M caused a significant reduction in the growth of colonies. Photosynthesis was significantly inhibited at all concentrations of DCMU. A rapidly growing population of algae treated with 10-7 m and 10-9 m DCMU showed a reduced motile/non-motile balance of colonies.  相似文献   

20.
To gain new insights into the function of photosystem II (PSII) herbicides DCMU (a urea herbicide) and bromoxynil (a phenolic herbicide), we have studied their effects in a better understood system, the bacterial photosynthetic reaction center of the terbutryn-resistant mutant T4 of Blastochloris (Bl.) viridis. This mutant is uniquely sensitive to these herbicides. We have used redox potentiometry and time-resolved absorption spectroscopy in the nanosecond and microsecond time scale. At room temperature the P(+)(*)Q(A)(-)(*) charge recombination in the presence of bromoxynil was faster than in the presence of DCMU. Two phases of P(+)(*)Q(A)(-)(*) recombination were observed. In accordance with the literature, the two phases were attributed to two different populations of reaction centers. Although the herbicides did induce small differences in the activation barriers of the charge recombination reactions, these did not explain the large herbicide-induced differences in the kinetics at ambient temperature. Instead, these were attributed to a change in the relative amplitude of the phases, with the fast:slow ratio being approximately 3:1 with bromoxynil and approximately 1:2 with DCMU at 300 K. Redox titrations of Q(A) were performed with and without herbicides at pH 6.5. The E(m) was shifted by approximately -75 mV by bromoxynil and by approximately +55 mV by DCMU. As the titrations were done over a time range that is assumed to be much longer than that for the transition between the two different populations, the potentials measured are considered to be a weighted average of two potentials for Q(A). The influence of the herbicides can thus be considered to be on the equilibrium of the two reaction center forms. This may also be the case in photosystem II.  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