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 31 毫秒
1.
Bound water is a major component of biological membranes and is required for the structural stability of the lipid bilayer. It has also been postulated that it is involved in water transport, membrane fusion, and mobility of membrane proteins and lipids. We have measured the fluorescence emission of membrane-bound 1-anilino-8-naphthalenesulfonate (ANS) and the infrared spectra of membranes, both as a function of hydration. ANS fluorescence is sensitive to polarity and fluidity of the membrane-aqueous interface, while infrared absorption is sensitive to the hydrogen bonding and vibrational motion of water and membrane proteins and lipids. The fluorescence results provide evidence of increasing rigidity and/or decreasing polarity of the membrane-aqueous interface with removal of water. The membrane infrared spectra show prominent hydration-dependent changes in a number of bands with possible assignments to cholesterol (vinyl CH bend, OH stretch), protein (amide A, II, V), and bound water (OH stretch). Further characterization of the bound water should allow its incorporation into current models of membrane structure and give insight into the role of membrane hydration in cell surface function.  相似文献   

2.
Abstract— The absorption and fluorescence spectroscopy of natural and model bilayer lipid membranes is reviewed. Basic structural features of biological membranes and the relative advantages of black lipid membranes (BLM) and of liposomes are discussed. Theoretical considerations show that the wavelengths of absorption maxima are affected by the refractive index and dielectric constant of the medium surrounding the chromophore. Techniques of obtaining photoelectric action spectra, direct absorption spectra, and reflection spectra of BLM are described. Polarized spectra can give information about the orientation of membrane constituents and show, for example, that the porphyrin ring of chlorophyll in BLM is tilted at 45 ± 5° to the membrane surface. Absorption maxima of chlorophyll in BLM are compared with solution spectra of various chlorophyll adducts and aggregates. It is concluded that chlorophyll in BLM exists largely as solvated monomer and dimer (or oligomer), depending on concentration, and is not coordinated with water. From the theory of fluorescence spectroscopy it follows that aggregation and the polarity of the environment affect the fluorescence yield and lifetime of a membrane component, and also the wavelength of its emission maximum. The microviscosity of the membrane matrix affects the anisotropy of fluorescence. Techniques of steady-state fluorescence spectroscopy and of fluorescence lifetime measurements are reviewed. Examples of the use of fluorescent probes in membrane studies are given. Certain probes such as anilinonaphthalene sulfonate (ANS) preferentially bind to membrane proteins. The location of a probe in a particular membrane region can be pinpointed from its fluorescence yield and emission maximum. The orientation of the hydrocarbon chains of membrane lipids has been found, from fluorescence polarization of certain probes, to be normal to the membrane surface as postulated a priori on the basis of the lipid bilayer model. Anisotropy of fluorescence shows that elongated probe molecules rotate rapidly about their long axes when surrounded by phospholipids but become immobilized when bound to proteins. Changes in intensity and anisotropy of fluorescence as function of temperature have demonstrated the existence of phase transitions and phase equilibria of membrane lipids. Excimer fluorescence has been used as a measure of the available lipid core volume of membranes. Mechanisms of energy transfer between membrane components are reviewed. The theoretical dependence of energy transfer on distance and orientation for several rigid and fluid membrane models is discussed in terms of the structural information it can provide. Fluorescence sensitization resulting from energy transfer within and across bilayer membranes has been demonstrated in various systems. Quantitative measurement of energy transfer efficiency in BLM has shown that such transfer is about five times more efficient than in solutions at comparable donor-acceptor distances. Lipid membranes can be viewed as structures which maintain their components at high concentrations, in a reactive state, and at favourable orientations.  相似文献   

3.
In mixed monolayers with purified chloroplast glycolipids and other colorless lipids, chlorophyll a fluorescence exhibits a decrease in quantum efficiency with increasing chlorophyll concentration. The fluorescence, which is strongly polarized in dilute films, becomes progressively depolarized as the area fraction of chlorophyll increases, and it is completely depolarized in a pure chlorophyll a monolayer. The observed behavior is consistent with an inductive resonance mechanism of energy transfer among the chlorophyll molecules with a critical transfer distance of 20–90 Å, depending on the model chosen for the energy transfer mechanism. The purified glycolipids–mono-and digalactosyl diglycerides and sulfoquinovodiglyceride–separately form stable, compressible monolayers of the liquid-expanded type on an aqueous subphase and in an atompshere of nitrogen. At maximum compression the three glycolipids occupy areas of 55, 80 and 47 A2-molecule-1, respectively, in the monolayer. Mixed monolayers of chlorophyll a with, separately, the monogalactolipid and the sulfolipid behave upon compression as two-dimensional solutions. The fluorescence polarization at high chlorophyll concentrations in mixed monolayers indicates that several of the lipid diluents facilitate local ordering of the pigment molecules.  相似文献   

4.
The action of low pH treatment (pH 3.6) known to release Ca2+ from the oxygen-evolving complex in photosystem II (PSII) membranes and to induce Ca2+-revers-ible inhibition of electron transport at the acceptor side of PSII in thylakoid membranes (TM) was compared in PSII membranes and TM. The rate of the inactivation of electron transport by low pH was four times higher in TM than in PSII membranes. Ferricyanide accelerated the inactivation of PSII membranes but decreased it in the case of TM. Low pH treatment also greatly modified the fluorescence induction kinetics in both preparations, but significant differences have been found in the fluorescence induction kinetics of treated TM and PSII membranes. Calcium restored the electron transport activity and fluorescence induction kinetics in PSII membranes and TM, whereas diphenylcarbazide restored these functions only in PSII membranes. The reactivation of Ca-depleted PSII membranes was more effective in the dark, whereas the reactivation of TM required weak light. In the case of PSII membranes subjected to low pH citrate buffer, maximal reactivation was observed at 60 mM Ca2+ but for TM about 10 mM Ca2+ was required and 60 mM fully inhibited electron transport in TM during reactivation. These results indicate that the Ca-dependent inactivation of the acceptor side of PSII in TM after low pH treatment cannot be explained by the extraction of Ca2+ from the oxygen-evolving complex. It is rather suggested that the Ca2+ involved in this inhibition is bound to the acceptor side of the photosystem near to the QA-non-heme iron binding site and may participate in the binding of a polypeptide of the PSII light antenna complex to the PSII reaction center.  相似文献   

5.
The interaction of Momordica charantia (bitter gourd) seed lectin (MCL) with several nucleic acid bases has been investigated by monitoring changes induced in the protein fluorescence by ligand binding. Values of the binding constant, Ka were obtained as 1.1 × 104, 1.56 × 104 and 2.2 × 103 M?1 for adenine, cytosine and uracil, respectively. In addition, binding of 8-anilinonaphthalene 1-sulfonate (ANS) with MCL was investigated by fluorescence spectroscopy. Interaction with MCL at low pH results in a large enhancement of the fluorescence intensity of ANS with a concomitant blue shift in the emission λmax, whereas at neutral and basic pH changes in both fluorescence intensity and emission maximum were very small, clearly suggesting that the MCL–ANS interaction is stronger at lower pH values. When excited at 295 nm in the presence of ANS, the protein fluorescence decreased with a concomitant increase in the emission intensity of ANS, suggesting resonance energy transfer from the tryptophan residues of MCL to ANS. Gel filtration profiles of MCL at pH values 2.0 and 7.4 are similar indicating that the tetrameric nature of MCL is retained even at low pH. Addition of lactose or adenine to MCL–ANS mixture did not alter the change in ANS fluorescence suggesting that lactose, adenine and ANS bind to MCL at independent and non-interacting sites. These results are relevant to understanding the functional role of MCL in the parent tissue.  相似文献   

6.
The peptide hormone adrenocorticotropin and a related peptide were studied in solution and in interaction with a model system of membranes (small unilamellar vesicles of dipalmitoylphosphatidylcholine and 17% dimyristoylphosphatidylglycerol) via fluorescence spectroscopy. In aqueous solution, intramolecular distances between the fluorescent residues R(Tyr2-Trp9) = 9.2 Å and R(Trp9-Tyr23) 18 Å were obtained, in agreement with molecular models. Interaction of the peptide with the negatively charged membrane is evident from the alteration of the Trp photophysical parameters (quantum yield, fluorescence spectra and anisotropy), with a partition constant between the lipidic and aqueous phase of Kp =1–2 times 103. The existence of two populations of Trp in the membrane, which are distinctly accessed by acrylamide, was concluded from the tryptophan fluorescence quenching study; the two fractions are located near the membrane interface as inferred from its fluorescence quenching by the 5-doxylstearate and 16-doxylstearate lipophilic quenchers. This result is further supported by energy transfer experiments to the 3-(9-anthroyloxyl)stearic acid and 12-(9-anthroyloxyl)stearic acid probes.  相似文献   

7.
Abstract— Incubating spinach chloroplasts with various chaotropic agents results in inhibition of photosynthetic electron transport between water and Photosystem II similar to the inhibition caused by washing chloroplasts with a high concentration of Tris buffer. Partial restoration of NADP photoreduction and fluorescence of variable yield is achieved by adding hydroquinone or Mn2+, either of which donates electrons to Photosystem II in the inhibited chloroplasts. The inhibitory treatments cause the release of Mn from its bound state in the chloroplast, thus allowing the measurement of the ESR signal of Mn2+. The ESR measurement is used to follow the photooxidation of Mn2+ as it donates electrons to photosystem II.  相似文献   

8.
The photoreaction of the furan-side moiety of 5-methoxypsoralen (5-MOP) with thymidine used as a DNA model compound was investigated in the dry state. Under these conditions, two main fluorescent photoadducts were formed and isolated by HPLC. The two modified nucleosides were characterized as the two cis-syn diastereoisomers of furan-side monoadducts of 5-MOP to thymidine on the basis of spectroscopic measurements including UV, fluorescence, 1H-NMR and circular dichroism analysis. The identification and quantification of the latter photoproducts within naked DNA exposed to photoexcited 5-MOP were achieved by enzymatic digestion completed by HPLC separation and fluorescence detection. Similarly, the two cis-syn furan-side monoadducts were found to be formed in the DNA of Saccharomyces cerevisiae cells after incubation with 5-MOP and subsequent exposure to 365 nm at an incident dose of 38.4 kJ m?2. Under these conditions, the rate of induction of two diastereoisomeric photoadducts was as low as one modification per 106 and 2 × 105 bases, respectively.  相似文献   

9.
Homogenous selective assay of albumin (ALB) in clinical sera was tested with 1-anilinonaphthalene-8-sulphonate (ANS) as Förster-resonance-energy-transfer (FRET) acceptor of tryptophan residues and biosensor of ALB. Between the excitation at 280 and 350 nm, the ratio of the fluorescence at 470 nm of free ANS in ethanol was about 1.9 while that of the complexes of ALB and ANS was about 3.9, supporting FRET in complexes of ANS and ALB. ANS below 1.0 mM saturated one site of ALB with Kd of about 0.13 μM in 20 mM sodium phosphate buffer at pH 7.0. For selective assay of ALB, 0.30 μM ANS was used to quantify fluorescence of the complexes at 470 nm under the excitation at 280 nm. ALB from 1.8 to 25 nM was quantified, whose lower limit was below 1% than that by bromocresol green assay while one-third than that by immunoturbidimetric assay. Globular proteins at comparable levels gave negligible signals. This new method showed reasonable resistance to other interfering substances in clinical sera. Quantities of ALB in clinical sera by this method were consistent with those by bromocresol green assay and immunoturbidimetric assay. Hence, homogenous assay of ALB with ANS as FRET biosensor was effective.  相似文献   

10.
The S-center radical (ANS·) of sodium 1-anilino-8-naphthalene sulfate (ANS) generated by photoinduced charge transfer in ANS/CdS and ANS/CdS/β-cyclodextrin(β-CD) systems has been studied by using spin trapping electron spin resonance techniques, UV-visible spectroscopic methods, and fluorescence spectroscopic methods. It was found that the S-centered radical (ANS·) was produced by the charge transfer reaction between the ground state ANS and the positive hole h+(CdS) from the valence band of CdS colloids, by the charge transfer from the excited singlet state 1ANS* to the conduction band of CdS colloids, or by both in the ANS/CdS and ANS/CdS/β-CD systems. The ESR signal intensity of the spin adduct (5,5′-dimethyl-1-pyrroline-N-oxide (DMPO)–ANS)·, which is formed from ANS· trapped by DMPO, in the latter system is 15 times stronger than that in the former system. The apparent association constants between ANS and CdS colloids in the absence and presence of β-CD determined from fluorescence quenching experiments are 1097 and 1606 M−1, respectively. From ESR and fluorescence results, it is estimated that the efficiency of photoinduced charge transfer from ANS to CdS colloids in the ANS/CdS/β-CD system is 12.5 times that in the ANS/CdS system.  相似文献   

11.
–Small and immunoaffinity-purified large phytochrome (Pr and Pfr) show a so-called anomalous fluorescence with λemmax= 470 and 440 nm, respectively, when irradiated within the blue absorption band. Model studies indicate that this emission arises from a dipyrromethenone partial structure produced by a nucleophilic addition to the central carbon C-10 of the bilindione chromophore. The anomalous emitter of phytochrome is thus similar to one bilirubin conformer which has previously been found to contribute to the absorption and emission of the bile pigment.  相似文献   

12.
Abstract— Phycobilisomes from the blue-green alga Nostoc sp. are known to contain the phycobiliproteins: c-phycoerythrin (c-PE), c-phycocyanin (c-PC) and four forms of allophycocyanin (APC I, II, III, and B). We have made a detailed study of the effects of the intensity of a single 6 ps excitation pulse on the decay kinetics and the yield of fluorescence in the individual isolated phycobiliproteins at pH 7 and 23°C. The risetime of the fluorescence of c-PE, c-PC and APC was > 12 ps. We found that the decay of the fluorescence was exponential at intensities of 1014 photons/cm2 in all the phycobiliproteins; the lifetimes being 1552 ± 31ps for c-PE, 2111 ± 83ps for c-PC, 1932 ± 165ps for APC I, 1870 ± 90ps for APC II, 1816 ± 88ps for APC III, (1869 ± 62ps for the averaged APC's I, II, and III), and 2667 ± 233 ps for APC B. We also found that the fluorescence decay became non-exponential in c-PE at excitation intensities < 1014 photons/cm2, but was exponential for all the other phycobiliproteins even at a pulse intensity of 1015 photons/cm2. The relaxation times of c-PE and c-PC decreased with excitation intensity above 1014 photons/cm2. For c-PE and c-PC the relative fluorescence vs excitation intensity was readily described by a relationship derived for a model in which exciton–exciton annihilation occurs. In APC the fluorescence yield and relaxation time were only slightly dependent on the excitation intensity. The results are interpreted to indicate the occurrence of singlet–singlet annihilation intramolecularly among the several phycobilin chromophores within the individual phycobiliprotein molecules in solution. The s to f transfer time is less than 12ps in c-PC.  相似文献   

13.
Abstract— Phase-fluorometer measurements of the fluorescence lifetime, τ, from chlorophyll in Chlorella, Bishop's 8 and 11 Scenedesmus mutants, sugarbeet leaf and chloroplast fragments demonstrate that: τ is independent of modulation frequency at 27 and 14 mc. in the experimental-wavelength range from 650 to 735 nm (with blue or blue-green excitation); with Chlorelfa and chloroplast fragments τ rises hyperbolically with intensity to τmax about 2 nsec and 0·7 nsec respectively; DCMU poisoned Chlorella and sugarbeet leaf as well as the mutants have τ values near 2 nsec; the lifetime-incident intensity relationship for Chlorella and chloroplast fragments is quantitatively similar to the incident-intensity dependence of fluorescence yield and oxygen evolution and thus supports the hypothesis that these three measuring variables are controlled by the concentration of ‘open’ trapping systems; τ is independent of emission wave-length to suggest that fluorescence is dominated by a single chlorophyll species. The reaction velocity-lifetime correlation indicates that fluorescence behavior is directly controlled by system II.  相似文献   

14.
-We have carried out a very detailed study, using fluorescence and optical flash photolysis techniques, of the photoreduction of methyl viologen (MV2+) by the electron donor ethylene diamine tetraacetic acid (EDTA) in aqueous solution sensitized by the dye acridine orange (AOH+). A complete mechanism has been proposed which accounts for virtually all of the known observations on this reaction. This reaction is novel in that both the triplet and the singlet state of AOH+ appear to be active photochemically. We have shown that mechanisms previously proposed for this reaction are probably incorrect due to an artifact. At pH 7 the fluorescence quantum yield φs of AOH+ is 0.26 ± 0.02 and the fluorescence lifetime is 1.8 ± 0.2 ns. φs is pH dependent and reaches a maximum of 0.56 at pH 4. The fluorescence of AOH+ is quenched by MV2+ at concentrations above 1 mM and the quenching obeys Stern-Volmer kinetics with a quenching rate constant of (1.0 ± 0.1) × 1010M?1 s?1. The quenching of the AOH+ excited singlet state by MV2+ almost certainly returns the AOH+ to its ground state with no photochemistry occurring. EDTA also quenches the fluorescence of AOH· with Stern-Volmer kinetics but with a smaller rate constant (6.4 ± 0.5) × 108M?1s?1 at pH 7. In this case the quenching is reactive resulting in the formation of semireduced AOH. In the presence of MV2+, flash irradiation of AOH+ does result in the reversible formation of the semireduced MV? which absorbs at 603 nm. We attribute this to a photochemical reaction of the triplet state of AOH+ with MV2+. The initial quantum yield for formation of MV? (φMV:)0 was found to be constant at 0.10 ± 0.05 for [MV2+] from 5 × 10?5 to 1.0 × 10?3 with [AOH+] = 8 × 10?6M. Previous workers had found that (φMV:)0 appears to decrease with decreasing [AOH+]; however, on careful investigation, we found this was most probably due to quenching of the triplet state of AOH+ by trace amounts of oxygen. When EDTA is added to a mixture of AOH + and MV2+ at pH 7, the photochemical formation of MV? becomes irreversible as the [EDTA] is increased. The quantum yield for the irreversible formation of MV? exceeds 0.10 becoming as large as 0.16 for [EDTA] = 0.014M. This fact requires that an alternative photochemical process must be operative and we present evidence that this is a reaction of EDTA with the excited singlet state of AOH+ to produce the semi-reduced AOH- which then reacts with MV2+ to produce MV?. The full kinetic scheme was tested by computer simulation and found to be totally consistent. This also enabled the processing of a full set of rate constants. When colloidal PtO2 was added to the optimal mixture [EDTA] = 3.4 × 10?2M; [MV2+] = 5 × 10?4M; [AOH+] = 4 × 10?5M; pH6 H2 gas was produced at a rate of 0.2μmol H2h?1. Thus, acridine orange should serve as an effective sensitizer in reactions designed to use solar energy to photolyze water.  相似文献   

15.
Abstract— The binding parameters of bicarbonate to the thylakoid membrane at different formate concentrations have been established [Stemler and Murphy (1983) Photochem. Phorobiol. 38, 701–707]. Based on these parameters, predictions could be made concerning the effects of bicarbonate and formate on photosynthetic electron flow. In this work these effects of various concentrations of bicarbonate and formate are measured and compared to predictions from the binding study. Electron flow is measured between QA and QB (the primary and secondary quinone acceptors) and QB and the plastoquinone pool. Also, these same concentration effects are determined for silicomolybdate supported oxygen evolution. It is found that the results of the bicarbonate binding study are in good agreement with the concentration dependence determined for the quinone reactions, as well as the silicomolybdate reaction. The bicarbonate concentrations required for half-maximal effects are approximately 100 μM, 300 μM and 1.3 mM in the presence of 0, 20 mM and 100 mM formate, respectively. It is concluded that a hierarchy of possible electron flow rates exist. The slowest rates occur when formate is bound. A substantially higher rate occurs when neither formate nor bicarbonate (< 2 μM) are present, but only chloride is present. The highest rates of electron flow occur when bicarbonate is bound. The QA- QB→ Qa Qb? Qa? Qb2– PQ → Qa Qb- PQ2–, and the silicomolybdate reactions all have the same concentration dependence on formate and bicarbonate. From this it is concluded that a single binding site for formate and bicarbonate affect all of these reactions. The possibility that multiple sites exist with approximately equal affinities for bicarbonate cannot be excluded.  相似文献   

16.
Abstract The short-term adaptation of intact leaves to an increase in light intensity was studied by an analysis of chlorophyll fluorescence and oxygen evolution monitored by photoacoustics. An increase in light intensity led to an oxygen “gush”. This “gush” was followed by a large (up to 120%) biphasic increase in the yield of oxygen evolution characterized by a fast phase (T = 0.5–2 min) and a slow phase (T = 4–20 min). The fast phase of the increase in oxygen yield was coupled to a decrease of fluorescence, whereas the slow phase was accompanied by a parallel fluorescence increase. A comparison of fluorescence parameters with oxygen yield indicates that the slow phase of the increase in oxygen yield was coupled to an increase in the antenna size of photosystem II. The slow phase was not inhibited by the uncoupler Nigericin but it was absent in chlorophyll-b-less barley mutants dencient in the light harvesting chlorophyll a/b protein complex of photosystem II (LHC II). These experiments indicate that changes in the LHC II mediated energy distribution, which occur in the time-range of several minutes, are involved in the adaptation to changing light intensities. Moreover, electrophoretic analysis of 32P orthophosphate labeled leaf discs adapted to low and high light intensities suggests that the slow phase of the increase in oxygen evolution involves dephosphorylation of the 25 kDa polypeptide of LHC II, by a small extent of 12%. The trigger for the slow phase of the increase in oxygen yield does not involve the oxidation of the plastoquinone pool. It was found that in response to the increased light intensity, the plastoquinone pool became more reduced as judged by model calculations. Experiments with the uncoupler Nigericin suggest that the control of the slow phase of adaptation to increased light intensity was also not exerted by the pH gradient across the thylakoid membrane. The similarities between the adaptation to increased light intensity and the state II to state I transition suggest that both adaptation phenomena involve LHC II dephosphorylation possibly triggered by the cytochrome b6/f complex.  相似文献   

17.
Abstract— The short-term adaptation of intact leaves to an increase in light intensity was studied by an analysis of chlorophyll fluorescence and oxygen evolution monitored by photoacoustics. An increase in light intensity led to an oxygen “gush”. This “gush” was followed by a large (up to 120%) biphasic increase in the yield of oxygen evolution characterized by a fast phase (T = 0.5–2 min) and a slow phase (T = 4–20 min). The fast phase of the increase in oxygen yield was coupled to a decrease of fluorescence, whereas the slow phase was accompanied by a parallel fluorescence increase. A comparison of fluorescence parameters with oxygen yield indicates that the slow phase of the increase in oxygen yield was coupled to an increase in the antenna size of photosystem II. The slow phase was not inhibited by the uncoupler Nigericin but it was absent in chlorophyll-b-less barley mutants deñcient in the light harvesting chlorophyll a/b protein complex of photosystem II (LHC II). These experiments indicate that changes in the LHC II mediated energy distribution, which occur in the time-range of several minutes, are involved in the adaptation to changing light intensities. Moreover, electrophoretic analysis of 32P orthophosphate labeled leaf discs adapted to low and high light intensities suggests that the slow phase of the increase in oxygen evolution involves dephosphorylation of the 25 kDa polypeptide of LHC II, by a small extent of 12%. The trigger for the slow phase of the increase in oxygen yield does not involve the oxidation of the plastoquinone pool. It was found that in response to the increased light intensity, the plastoquinone pool became more reduced as judged by model calculations. Experiments with the uncoupler Nigericin suggest that the control of the slow phase of adaptation to increased light intensity was also not exerted by the pH gradient across the thylakoid membrane. The similarities between the adaptation to increased light intensity and the state II to state I transition suggest that both adaptation phenomena involve LHC II dephosphorylation possibly triggered by the cytochrome b6/f complex.  相似文献   

18.
Abstract: A variety of naturally occurring photosensitizers of plant origin were tested for their ability to cause ultraviolet-A (UVA)-dependent inhibition of photosynthetic 14CO2 fixation in leaf disks of Pisum sativum L. At 0.1 mM concentrations and 60 min UVA irradiation, α-terthienyl was strongly inhibitory, harmine and sanguinarine inhibited to a lesser degree, and curcumin, 8-methoxypsoralen and nordihydroguaiaretic acid had no effect under the conditions tested. Alpha-terthienyl + UVA treatments that fully inhibited 14CO2 fixation had no effect on fresh weight, chlorophyll or protein content of the tissue. Chloroplast malate dehydrogenase and glyceraldehyde-3-phosphate dehydrogenase were inhibited 45% and 29%, respectively, by α-terthienyl + UVA treatment. Electron transport from H2O to the reducing side of photosystem I was inhibited to a similar extent, suggesting interference with the reductive activation of chloroplast enzymes. Alpha-terthienyl + UVA-treated tissue exhibited a seven-fold increase in leakage of labeled photosynthate into the external medium. Treated leaf disks showed no ability to recover 14CO2 fixing ability over a 24 h period. These results indicate photosensitized damage at the level of the thylakoid membranes resulting in partial loss of electron transport capability and more general damage to chloroplast and cell membranes.  相似文献   

19.
The ultraviolet fluorescence of the purple membrane of H. halobium and its apomembrane was characterized by measuring emission spectra, polarization, decay lifetimes and the changes induced by pH and temperature. The fluorescence quantum yields of the two membranes are 0.024 × 0.003 and 0.17 × 0.03, respectively. The emission, which shows lifetimes in the 0.4 to 4 ns range, was assigned to heterogeneous populations of emitters, consisting, probably, of two tryptophans in the purple membrane and seven or eight residues in the apomembrane. Acrylamide quenching experiments showed that the accessibility of this neutral quencher to the fluorophors is reduced greatly in both membranes. Fluorimetric methods were also used in an attempt to monitor the purple complex reconstitution process. It was concluded that the fluorescence quantum yields of any monomers, dimers and trimers present in the partially reconstituted membranes should be very similar.
Finally, based on the spectroscopic results and on specific folding patterns of the seven α-helical regions of bacteriorhodopsin (Stoeckenius and Bogomolni, 1982), it is proposed that Trp 137, Trp 138 (and perhaps Trp 10) of the protein molecule are the most plausible fluorophors in the purple membrane. It is also suggested that the protein in the apomembrane takes a more open configuration which is permeable to small ions and molecules.  相似文献   

20.
Abstract— Effect of red-light irradiation on the medium pH at 10d?C was measured and compared among unbuffered solutions of the 121-kDa native pea (Pisum sativum cv. Alaska) phytochrome and its 114- and 62-kDa fragments in a red-light-absorbing form (Pr), all of which converted to far-red-light-absorbing form (Pfr) on red-light irradiation. Red-light irradiation induced alkalinization in the solutions of the phytochrome and the fragments in the pH range 6.6-7.2 and 6.2-7.8, respectively. The amount of protons taken up by the 121-kDa phytochrome was less than one half of that of the 114-kDa fragment. Red-light irradiation induced acidification in the solutions of the 114- and the 62-kDa fragments above pH 7.8. In the solutions of the 121-kDa phytochrome, however, the irradiation induced no pH change at pH 7.2-8.2, and only a slight acidification at pH 8.2-8.7, which may be ascribed to a small amount of contamination from the 114-kDa fragment. All these red-light-induced pH changes were reversible following exposure to far-red light. The 7-kDa polypeptide(s) of the native 121-kDa phytochrome, which is lacking in the 114-kDa fragment, thus, prohibited proton transfer between phytochrome and the medium. A red-light-induced pH change was also measured in unbuffered solutions of the 39-kDa fragment of the phytochrome and of the 114-kDa fragment in the presence of 0.8 mM soyasaponin I. The 39-kDa fragment showed partially photoreversible conversion between a spectral form having an absorption maximum at 659 nm (P659) and a bleached form, P***. The 114-kDa fragment in the presence of the saponin showed a photoreversible conversion between P65V and Pb,. Exposure of P659 from the 39-kDa fragment and from the 114-kDa fragment in the presence of the saponin to red light, caused acidification of the medium in the pH range 6.8-8.8 and 7.2-9.0, respectively, but no change at pH 6.2-6.8 and 6.4-7.2, respectively. The acidification of the latter was reversible following a far-red-light irradiation, but that of the former was only partially photoreversible. Proton uptake of phytochrome was inhibited by tryptic degradation to the 39-kDa fragment and also by the presence of the saponin. Only proton release was observed during the photoconversion from P659 and P***hl. It is suggested that a phytochrome molecule has possible site(s) for both proton release and for uptake and that the proton release reaction may be correlated to the photoconversion process(es) prior to the bleached intermediate (I***) of phytochrome.  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