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.
The relationship between the rate of O2 evolution and the corresponding concentration of System II active centers E has been studied in “State 1” and in “State 2”. No variation of α (fraction of absorbed light captured by System II) could be observed. The phenomena shown by Bonaventura and Myers (Biochim. Biophys. Acta, 189 (1969) 366) may be interpreted by an increase of the apparent equilibrium constant between the two photoreactions II and I when the algae are illuminated at 650 nm.  相似文献   

2.
The applicability of the spectrophotometric determination of histidine according to Ovadi et al. [(1967) Acta Biochim. Biophys. Acad. Sci. Hung.3, 455–458] to a large range of molar ratios of diethylpyrocarbonate (DEP) to histidine requires the use of appropriate differential molar absorptivities dependent on the DEP concentration used. This improved procedure allows a more differentiated aproach to the reactivity of histidine residues in proteins.  相似文献   

3.
4.
The divalent-cation-specific ionophore A23187 is used to define two components of the slow fluorescence quenching of type a spinach chloroplasts: ionophore-reversible and ionophore-resistant quenching. Ionophore-reversible quenching predominates at relatively low light intensities and approaches saturation as light levels are increased. It is sensitive to uncouplers and to 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU) and is dark reversible. At high light intensities the bulk (> 80%) of slow fluorescence quenching is ionophore-resistant. Ionophore-resistant quenching is stimulated by carbonyl cyanide m-chlorophenyl hydrazone (CCCP) at pH 7.6 and by both CCCP and methylamine at pH 9.0. It is insensitive to DCMU and is not reversed in subsequent darkness. Taken together, the two components account for all quenching observed in Type A chloroplasts.Ionophore-reversible quenching is identified with the Mg2+-mediated fluorescence quenching described by Krause (Biochim. Biophys. Acta (1974) 333, 301–313) and by Barber and Telfer (in Membrane Transport in Plants (Dainty, J., and Zimmermann, U., eds.), pp. 281–288, Springer-Verlag, Berlin, 1974). Ionophore-resistant quenching, a first-order process requiring high light, resembles the quenching reported by Jennings et al. (Biochim. Biophys. Acta (1976) 423, 264–274).The resolution of the fluorescence quenching phenomenon into two distinct components reconciles the apparently contradictory observations of these earlier investigations.  相似文献   

5.
Reaction centers of the photosynthetic bacterium Rhodopseudomonas sphaeroides R-26, give rise to large triplet state EPR signals upon illumination at low temperature (11 K). Utilizing monochromatic polarized light to generate the EPR spectra (magnetophotoselection) we have shown that the intensities of the observed triplet signals are strongly dependent upon the wavelength and polarization direction of the excitation. These data can be used to calculate the orientations of the excited transition moments with respect to each other and with respect to the triplet state principal magnetic axes system. Our quantitative approach is to follow the procedure outlined in a previous publication (Frank, H.A., Friesner, R., Nairn, J.A., Dismukes, G.C. and Sauer, K. (1979) Biochim. Biophys. Acta 547, 484–501) where computer simulations of the observed triplet state spectra were employed.The results presented in the present work indicate that the transition moment at 870 nm which is associated with the bacteriochlorophyll ‘special pair’ lies almost entirely along one of the principal magnetic axes of the triplet state. Also, the 870 nm transition moment makes an angle of approx. 60° with the 546 nm transition moment which is associated with a bacteriopheophytin. This latter result is in agreement with previous photoselection studies on the same bacterial species (Vermeglio, A., Breton, J., Paillotin, G. and Cogdell, R. (1978) Biochim. Biophys. Acta 501, 514–530).  相似文献   

6.
Cytochrome b5 was found to bind readily to sonicated vesicles containing as much as 0.8 mol cholesterol per mol egg phosphatidylcholine. This observation conflicts with the suggestion of Enomoto and Sato ((1977) Biochim. Biophys. Acta 466, 136–147) that cholesterol prevents binding of this protein to erythrocyte membranes.  相似文献   

7.
The impact of reaction of galactolipids with ozone on the physicochemical properties of their monolayers was examined. In Megli and Russo (Biochim Biophys Acta, 1778:143–152, 2008), Cwiklik and Jungwirth (Chem Phys Lett, 486:99–103, 2010), Jurkiewicz et al. (Biochim Biophys Acta, 1818:2388–2402, 2012), Khabiri et al. (Chem Phys Lett, 519:93–99, 2012), and Conte et al. (Biochim Biophys Acta, 1828:510–517, 2013), the properties of layers formed from model mixtures composed of chosen lipids and selected oxidation products were studied, whereas in this work, question was raised as to how the oxidation reactions taking place in situ affect the physical properties of the galactolipid layers. So, set experiment should take into account the effect of all reaction products. The mechanical characteristics of monolayers of monogalactosyldiacyl-glycerol (MGDG) and digalactosyldiacylglycerol (DGDG) were determined by Langmuir trough technique, and the electrical properties of liposomes formed from these lipids by measuring their electrophoretic mobility. Considerable loss of galactolipid molecules forming monolayers was found at ozone concentrations (in aqueous medium) higher than 0.1 ppm with a stronger effect measured for MGDG. That goes along with the greater amounts of MDA found in the extracts of oxidized MGDG films compared with DGDG. Based on this, it was concluded that an additional galactose group present in DGDG molecules acts protectively under oxidative conditions. The surface tension of the solutions (of small volume) contacting the oxidized galactolipids films was significantly reduced, indicating the presence of soluble in polar media, surface active reaction products. The presence of α-tocopherol in mixtures with tested galactolipids at a molar ratio of lipid to tocopherol equal to 1.7:1 caused some inhibition of lipid oxidation, reducing the decrease of amount of lipid particles forming the monolayer. Here, also protective effect of α-tocopherol was greater for the MGDG compared to DGDG.  相似文献   

8.
The binding of cytochrome b5 to single-walled liposomes of egg phosphatidylcholine was inhibited by the presence of cholesterol in the lipid bilayer under conditions where a limited amount of liposomes was incubated with the cytochrome. Since similar conditions seem to apply for the binding of cytochrome b5 to erythrocyte ghosts, this observation supports the conclusion of Enomoto and Sato (Enomoto, K. and Sato, R. (1977) Biochim. Biophys. Acta 466, 136–147) that the localization of cholesterol on the outer surface of the ghost membrane prevents the binding of cytochrome b5 to this surface. The finding reported by Roseman et al. (Roseman, M.A., Holloway, P.W. and Calabro, M.A. (1978) Biochim. Biophys. Acta 507, 552–556) that cholesterol did not prevent the cytochrome binding to phosphatidylcholine liposomes in the presence of a large excess of liposomes could be confirmed in the present study, but this does not contradict the abovementioned conclusion.  相似文献   

9.
Ora Canaani  Shmuel Malkin 《BBA》1984,766(3):513-524
Using the photoacoustic technique, state 1-state 2 transitions were studied in an intact leaf by direct monitoring of modulated oxygen evolution, excited by modulated light. States 1 and 2 were characterized by the extent of immediate enhancement of the modulated oxygen evolution — ‘Emerson enhancement’ — and the concomitant fluorescence quenching, resulting from the addition of continuous far-red light (greater than 700 nm), absorbed primarily in Photosystem I (light 1). The extent of Emerson enhancement as well as the saturation curve of this effect by far-red light are very sensitive and quantitative indicators for the ratio of light excitation distributed between Photosystems I and II. The enhancement ratios at 650 nm light, a typical light 2, were in a range 1.4–1.8 in state 1, while values as low as 1.06 were observed in state 2. During the transition from state 2 to state 1, monitored in presence of modulated light 2 and background continuous light 1, the modulated oxygen yield increased considerably, indicating a major increase in excitation flux into Photosystem II. Conversely, with modulated light 2 alone in state 1, the modulated oxygen evolution yield was smaller than in state 2, indicating a decrease of the excitation flux in Photosystem I. In a typical example, of the transition to state 1, the fraction of light absorbed by Photosystem II, β, increased from 0.46 to 0.64, while that absorbed by Photosystem II, α, decreased from 0.43 to 0.36. State 1-state 2 transitions, thus, reflect reciprocal changes in the cross-sections of the two photosystems for light absorption. There is no evidence for the operation of a ‘spill-over’ mechanism. The enhancement effect displayed maxima at 480 and 650 nm, related to chlorophyll-b absorption, as well as another band at 500–550 nm. In a chlorophyll-b-less barley mutant, state 1-state 2 transitions, as monitored by modulated oxygen evolution, were absent, and the resulting enhancement corresponded to state 2. These observations are consistent with the model that the light-harvesting chlorophyll-ab complex plays a role in regulating the distribution of light to the photosystems. It is probable that this complex migrates reversibly in the thylakoid membrane in such a way that it is mainly associated with Photosystem II in state 1, but is more evenly distributed in the two photosystems in state 2.  相似文献   

10.
We have investigated the influence of chloroplast organization on the nature of chemical reductive titrations of Photosystem II fluorescence decay kinetics in spinach chloroplasts. Structural changes of the chloroplast membrane system were induced by varying the ionic environment of the thylakoids. A single-photon timing system with picosecond resolution monitored the kinetics of the chlorophyll a fluorescence emission. At all ionic concentrations studied, we have observed biphasic potentiometric titration curves of fluorescence yield; these have been interpreted to be suggestive of electron acceptor Q heterogeneity (Karukstis, K.K. and Sauer, K. (1983) Biochim. Biophys. Acta 722, 364–371; Cramer, W.A. and Butler, W.L. (1969) Biochim. Biophys. Acta 172, 503–510). A direct relation is observed between the Em value of the low-potential component of Q and the Mg2+ concentration of the chloroplast suspending medium. We have attributed these midpoint potential variations to the thylakoid structural rearrangements involved in cation-regulated grana stacking. Ionic effects on the fluorescence decay kinetics at the redox transitions are discussed in terms of the heterogeneity of Photosystem II units (α- and β-centers) and the mechanism of deexcitation at a closed reaction center (fluorescence or nonradiative decay).  相似文献   

11.
Three distinct states can be identified for cells of the green alga Chlorella vulgaris; State 1 and State 2 obtained by preillumination in far-red and red light, respectively, and the dark state obtained by dark-adaptation. Addition of the inhibitor DCMU to algal cells leads to an initial rapid increase in chlorophyll-a fluorescence reflecting the closure of Photosystem II traps. This, in the case of dark and state-2-adapted algae is followed by a slow light-dependent increase to a fluorescence yield typical of State-1-adapted cells. Measurements of low temperature (77 K) emission spectra indicate that the low fluorescence yields of dark and State-2-adapted algae reflect similar balances in excitation-energy distribution between the two photosystems. In both cases, the balance favours PS I and the slow fluorescence increase seen in the poisoned algae reflects a redressing of this balance in favour of PS II. The low fluorescence yield of State-2-adapted algae is thought to be associated with the phosphorylation of chlorophyll a/b light-harvesting protein (Biochim. Biophys. Acta (1983) 724, 94–103). Measurements of the uncoupler and ATPase sensitivity of the light-dependent increases seen in DCMU-poisoned cells indicate that the low fluorescence yield of dark-adapted algae is of different origin. Evidence is presented showing that the light-driven changes in excitation-energy distribution seen in green algae involve two distinct processes; a low-intensity, wavelenght-independent change reflecting simple light/dark changes and a higher intensity, wavelength-dependent change reflecting State 1/State 2 adaptation. The former changes appear to be associated with changes in the local ionic environment within the algal chloroplast, whilst the latter appear to reflect changes in the phosphorylation state of chlorophyll a/b light-harvesting protein.  相似文献   

12.
13.
In 1972, Lieb, W. R.; Stein, W. D. (Biochim. Biophys. Acta 265, 187–207) in their review of sugar transport in human erythrocytes concluded that the conventional two-state carrier model was inconsistent with the experimental data available at that time. Since then, other papers have appeared which question the validity of the model. In this paper, we give a brief derivation of the equations describing the two-state carrier model, and analyze the predictions of the model in the classical experiments, i.e. zero-trans, infinite-cis, and equilibrium exchange. We show that the estimate of the half saturatiion constant of 2.8 mM for glucose at the inner face of the human red cell membrane for the infinite-cis procedure reported by Hankin, B. L., Lieb, W. R. and Stein, W. D ((1972) Biochim. Biophys. Acta 288, 114–126) is unreliable. We note that all of the other experimental findings are consistent with the asymmetric carrier model.  相似文献   

14.
15.
Streaming of actomyosin solutions observed in glass capillary tubes has been attributed to the energy-tranducing interaction of actin and myosin in the presence of ATP (A. Oplatka and R. Tirosh, Biochim. Biophys. Acta 1973, 305, 684–688). Other results have indicated that this is a physical phenomenon of the system used (Y. Matsutake, J. Sagara, T. Yamada, and H. Shimizu, Biochim. Biophys. Acta 1975, 405, 347–352). Results presented here using a microcapillary system relatively free of artifacts confirm the indication that the structure of the capillary system used is responsible for the appearance of streaming in the solutions studied and further delineate the nature of the phenomenon.  相似文献   

16.
The spectral characteristics of the ‘655 nm’ band of cytochrome oxidase were found to be affected by ligands of the binuclear centre, including formate and chloride, and by the resting/pulsed transition. The band titrated with near n=1 characteristics at a midpoint of about 400 mV, in contrast to haem a3, which exhibits strong redox interaction and a titration range at significantly lower potential. Thus, although the total reduced-oxidised difference spectrum of haem a3, shows a trough at about 655 nm, this characteristic is absent in the low potential region. The 655 nm feature may arise from a charge transfer band of ferric high-spin haem a3, which is modulated by the redox state of CuB, as suggested by Beinert et al. [(1976) Biochim. Biophys. Acta 423, 339–355].  相似文献   

17.
A new type of multi-color PAM chlorophyll fluorometer (Schreiber et al. 2012) was applied for measurements of photodamage to photosystem II (PS II) in optically thin suspensions of Chlorella (200 μg Chl l?1) in the presence of 1 mM lincomycin. An action spectrum of the relative decrease of F v/F m in the 440–625 nm range was measured, which not only showed the expected high activity in the blue, but at a lower level also substantial activity above 540 nm. With the same dilute suspension, a PS II absorption spectrum was derived via measurements of the O-I1 rise kinetics induced by differently colored strong light at defined incident quantum flux densities. After normalization of the two spectra at 625 nm, the relative extent of photodamage at 440–480 nm proved substantially higher than absorption by PS II, whereas the two spectra were close to identical in the 540–625 nm range. Hence, overall photodamage to PS II appears to consist of two components, one of which is due to light absorbed by PS II pigments, whereas the other one is likely to involve direct light absorption by Mn in the oxygen-evolving complex (Hakala et al. Biochim Biophys Acta 1706:68–80, 2005). Based on this rationale, an action spectrum of the Mn mechanism of photodamage was deconvoluted from the overall action spectrum, declining steeply above 480 nm. An almost identical Mn-spectrum was derived by another approach with the PAR of the various colors being adjusted to give identical rates of PS II turnover, PAR II. The tentative, basic assumption of negligibly small contribution of the Mn mechanism to photodamage above 540 nm was supported by supplementary measurements using an external 665 nm lamp. 665 nm not only gave about two times PS II turnover as compared to 625 nm, but also about two times photodamage.  相似文献   

18.
The rise kinetics of the absorption changes induced at 515 nm and 480 nm by a flash were studied using two types of xenon flashes of different durations. The ‘slow’ rise of the absorption change (t12 = 15–20 μs) observed by Cox and Delosme (1978 C.R. Acad. Sci. (Paris) Sér. D 282, 775–778) and Joliot P., Delosme, R. and Joliot, A. ((1977) Biochim. Biophys. Acta 459, 47–57) was found to be due to double hits occurring in the reaction centers of System I during the flash.The turnover kinetics of the reaction centers of System I after a short flash were studied by a double flash method. They are in agreement with a second order reaction between P+-700 and its electron donor.  相似文献   

19.
In Cryptomonas rufescens (Cryptophyceae), phycoerythrin located in the thylakoid lumen is the major accessory pigment. Oxygen action spectra prove phycoerythrin to be efficient in trapping light energy.The fluorescence excitation spectra at ?196°C obtained by the method of Butler and Kitajima (Butler, W.L. and Kitajima, M. (1975) Biochim. Biophys. Acta 396, 72–85) indicate that like in Rhodophycease, chlorophyll a is the exclusive light-harvesting pigment for Photosystem I.For Photosystem II we can observe two types of antennae: (1) a light-harvesting chlorophyll complex connected to Photosystem II reaction centers, which transfers excitation energy to Photosystem I reaction centers when all the Photosystem II traps are closed. (2) A light-harvesting phycoerythrin complex, which transfers excitation energy exclusively to the Photosystem II reaction complexes responsible for fluorescence at 690 nm.We conclude that in Cryptophyceae, phycoerythrin is an efficient light-harvesting pigment, organized as an antenna connected to Photosystem II centers, antenna situated in the lumen of the thylakoid. However, we cannot afford to exclude that a few parts of phycobilin pigments could be connected to inactive chlorophylls fluorescing at 690 nm.  相似文献   

20.
A low-molecular-weight acid phosphatase which contains iron   总被引:3,自引:0,他引:3  
A simple, large scale purification procedure for the violet “phosphoprotein phosphatase” from beef spleen [Glomset, J. and Porath, J., Biochim. Biophys. Acta, 39, 1 (1960)] has been devised. The procedure involves a 24-hr extraction of a tissue homogenate at pH 3, ammonium sulfate fractionation, CM-cellulose chromatography, and Sephadex G-75 chromatography. The highly purified enzyme shows a constant ratio of both enzymatic activity and absorbance at 550 nm to iron concentration. The data strongly indicate that the enzyme contains iron, and moreover, support a 1:1 molar ratio of iron to enzyme.  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