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.
Pospíšil  P. 《Photosynthetica》1998,34(3):343-355
The excitation energy of pigment molecules in photosynthetic antennae systems is utilised by photochemistry, partly it is thermally dissipated, and partly it is emitted as fluorescence. Changes in the quantum yield of chlorophyll (Chl) fluorescence reflect the changes in quantum yield of photochemical reaction and thermal dissipation of the excitation energy. Decrease of the Chl fluorescence quantum yield is called the Chl fluorescence quenching. The decrease of the quantum yield that is accompanied by photochemical reactions has been termed the photochemical quenching, and the decrease accompanied by thermal dissipation of the excitation energy is called the non-photochemical quenching. This review deals with mechanisms of the non-photochemical quenching.  相似文献   

2.
The excitation energy of pigment molecules in photosynthetic antennae systems is utilised by photochemistry, partly it is thermally dissipated, and partly it is emitted as fluorescence. Changes in the quantum yield of chlorophyll (Chl) fluorescence reflect the changes in quantum yield of photochemical reaction and thermal dissipation of the excitation energy. Decrease of the Chl fluorescence quantum yield is called the Chl fluorescence quenching. The decrease of the quantum yield that is accompanied by photochemical reactions has been termed the photochemical quenching, and the decrease accompanied by thermal dissipation of the excitation energy is called the non-photochemical quenching. This review deals with mechanisms of the non-photochemical quenching.  相似文献   

3.
4.
The induction and relaxation of non-photochemical quenching (NPQ) under steady-state conditions, i.e. during up to 90 min of illumination at saturating light intensities, was studied in Arabidopsis thaliana. Besides the well-characterized fast qE and the very slow qI component of NPQ, the analysis of the NPQ dynamics identified a zeaxanthin (Zx) dependent component which we term qZ. The formation (rise time 10-15 min) and relaxation (lifetime 10-15 min) of qZ correlated with the synthesis and epoxidation of Zx, respectively. Comparative analysis of different NPQ mutants from Arabidopsis showed that qZ was clearly not related to qE, qT or qI and thus represents a separate, Zx-dependent NPQ component.  相似文献   

5.
Non-photochemical chlorophyll fluorescence quenching (qN) in barley leaves has been analysed by monitoring its relaxation in the dark, by applying saturating pulses of light. At least three kinetically distinct phases to qN recovery are observed, which have previously been identified (Quick and Stitt 1989) as being due to high-energy state quenching (fast), excitation energy redistribution due to a state transition (medium) and photoinhibition (slow). However, measurements of chlorophyll fluorescence at 77 K from leaf extracts show that state transitions only occur in low light conditions, whereas the medium component of qN is very large in high light. The source of that part of the medium component not accounted for by a state transition is discussed.Abbreviations ATP adenosine 5-triphosphate - DCMU 3[3,4-dichlorophenyl]-1,1 dimethylurea - pH trans-thylakoid pH gradient - Fo, Fm room-temperature chlorophyll fluorescence yield with all reaction centres open, closed - Fv variable fluorescence = Fm–Fo - LHC II Light harvesting complex II - PS I, PS II Photosystem I, II - P700, P680 primary donor in photosystem I, II - qP photochemical quenching of variable fluorescence - qN non-photochemical quenching of variable fluorescence - qNe, qNt, qNi non-photochemical quenching due to high energy state, state transition, photoinhibition - qNf, qNm, qNs components of qN relaxing fast, medium, slow - qr quenching of r relative to the dark state - tricine N-tris[hydroxymethyl]methylglycine - r ratio of fluorescence maximum from photosystem II to that from photosystem I at 77 K  相似文献   

6.
The diatom algae, responsible for at least a quarter of the global photosynthetic carbon assimilation in the oceans, are capable of switching on rapid and efficient photoprotection, which helps them cope with the large fluctuations of light intensity in the moving waters. The enhanced dissipation of excess excitation energy becomes visible as non-photochemical quenching (NPQ) of chlorophyll a fluorescence. Intact cells of the diatoms Cyclotella meneghiniana and Phaeodactylum tricornutum, which show different NPQ induction kinetics under high light illumination, were investigated by picosecond time-resolved fluorescence under dark and NPQ-inducing high light conditions. The fluorescence kinetics revealed that there are two independent sites responsible for NPQ. The first quenching site is located in an FCP antenna system that is functionally detached from both photosystems, while the second quenching site is located in the PSII-attached antenna. Notwithstanding their different npq induction and reversal kinetics, both diatoms showed identical NPQ via both mechanisms in the steady-state. Their fluorescence decays in the dark-adapted states were different, however. A detailed quenching model is proposed for NPQ in diatoms.  相似文献   

7.
The photoprotective nature of non-photochemical quenching (NPQ) has not been effectively quantified and the major reason is the inability to quantitatively separate NPQ that acts directly to prevent photoinhibition of photosystem II (PSII). Here we describe a technique in which we use the values of the PSII yield and qP measured in the dark following illumination. We expressed the quantum yield of PSII (Φ(PSII)) via NPQ as: Φ(PSII)=qP×(Fv/Fo)/(1+Fv/Fo+NPQ). We then tested this theoretical relationship using Arabidopsis thaliana plants that had been exposed to gradually increasing irradiance. The values of qP in the dark immediately after the illumination period (here denoted qPd) were determined using a previously described technique for Fo' calculation: Fo'(calc.)=1/(1/Fo-1/Fm-1/Fm'). We found that in every case the actual Φ(PSII) deviated from theoretical values at the same point that qPd deviated from a value of 1.0. In an increasing series of irradiance levels, WT leaves tolerated 1000μmolm(-2)s(-1) of light before qP(d) declined. Leaves treated with the uncoupler nigericin, leaves of the mutant lacking PsbS protein and leaves overexpressing PsbS showed a qP(d) reduction at 100, 600 and 2000μmolm(-2)s(-1) respectively, each at an increasing value of NPQ. Therefore we suggest that this simple and timely technique will be instrumental for identifying photoprotective NPQ (pNPQ) and that it is more appropriate than the qE component. Its applications should be broad: for example it will be useful in physiology-based studies to define the optimal level of nonphotochemical quenching for plant protection and productivity.  相似文献   

8.
Cyanobacteria have previously been considered to differ fundamentally from plants and algae in their regulation of light harvesting. We show here that in fact the ecologically important marine prochlorophyte, Prochlorococcus, is capable of forming rapidly reversible non-photochemical quenching of chlorophyll a fluorescence (NPQf or qE) as are freshwater cyanobacteria when they employ the iron stress induced chlorophyll-based antenna, IsiA. For Prochlorococcus, the capacity for NPQf is greater in high light-adapted strains, except during iron starvation which allows for increased quenching in low light-adapted strains. NPQf formation in freshwater cyanobacteria is accompanied by deep Fo quenching which increases with prolonged iron starvation.  相似文献   

9.
10.
Chlorophyll fluorescence quenching induced by low concentrations of m-dinitrobenzene (DNB) is investigated. In intact spinach chloroplasts DNB causes photochemical and non-photochemical quenching. The two forms of quenching are distinguished by applying the saturation pulse method with a new type of modulation fluorometer. Half-maximal photochemical quenching is observed at about 3 micromolar DNB. It is inhibited by 3-(3,4 dichlorophenyl)-1, 1-dimethylurea (DCMU) and by 2,5-dibromo-3-methyl-6-isopropyl-p-benzoquinone (DBMIB). Photochemical quenching by DNB leads to suppression of the I-P transient in a fluorescence induction curve. Upon application of saturating continuous light, the increase of fluorescence yield is separated into a photochemical and a thermal part. DNB causes suppression of only the slowest sub-component of the thermal part, in analogy to the action of Hill reagents. Simultaneous measurements of oxygen exchange rate and fluorescence reveal that a part of DNB induced quenching is accompanied by oxygen uptake. Most DNB-induced non-photochemical quenching is prevented by nigericin and, hence, can be considered energy-dependent quenching. The small component persisting in the presence of nigericin is identical to the one observed with methylviologen and other Hill reagents, likely to be due to static quenching by oxidized plastoquinone. The presented data confirm the original finding of Etienne and Lavergne (Biochim Biophys Acta 283: 268–278, 1972) that low concentrations of DNB selectively affect the thermal component of variable fluorescence. However, while these authors interpreted the quenching by a non-photochemical mechanism, the present investigation emphasizes a photochemical mechanism, in analogy to the effect of electron acceptors or mediators.Abbreviations DBMIB 2,5-dibromo-3-methyl-6-isopropyl-p-benzoquinone - DCMU 3-(3,4-dichlorophenyl)-1, 1-dimethylurea - DNB m-dinitrobenzene - PGA 3-phosphoglycerate - PMS phenazinemethosulphate - PS I and PS II photosystems I and II  相似文献   

11.
12.
The components of non-photochemical chlorophyll fluorescence quenching (qN) in barley leaves have been quantified by a combination of relaxation kinetics analysis and 77 K fluorescence measurements (Walters RG and Horton P 1991). Analysis of the behaviour of chlorophyll fluorescence parameters and oxygen evolution at low light (when only state transitions — measured as qNt — are present) and at high light (when only photoinhibition — measured as qNi — is increasing) showed that the parameter qNt represents quenching processes located in the antenna and that qNi measures quenching processes located in the reaction centre but which operate significantly only when those centres are closed. The theoretical predictions of a variety of models describing possible mechanisms for high-energy-state quenching, measured as the residual quenching, qNe, were then tested against the experimental data for both fluorescence quenching and quantum yield of oxygen evolution. Only one model was found to agree with these data, one in which antennae exist in two states, efficient in either energy transfer or energy dissipation, and in which those photosynthetic units in a dissipative state are unable to exchange energy with non-dissipative units.Abbreviations: Fo, Fm room-temperature chlorophyll fluorescence yield with all centres open, closed - Fv variable fluorescence yield - LHC II light-harvesting chlorophyll-protein complex of PS II - PS I, PS II Photosystem I, II - P700, P680 primary donor in Photosystem I, II - QA primary electron acceptor of PS II - Pmax maximum quantum yield of oxygen evolution - qN coefficient of non-photochemical quenching of variable fluorescence - qNe, qNt, qNi coefficient of non-photochemical quenching due to high-energy-state, state transition, photoinhibition - qO coefficient of quenching of dark level fluorescence - qP coefficient of photochemical quenching of variable fluorescence - P intrinsic quantum yield of open PS II reaction centres = s/qP - PS 2 quantum yield of PS = qP × Fv/Fm - S quantum yield of oxygen evolution = rate of oxygen evolution/light intensity  相似文献   

13.
Zeaxanthin has been correlated with high-energy non-photochemical fluorescence quenching but whether antheraxanthin, the intermediate in the pathway from violaxanthin to zeaxanthin, also relates to quenching is unknown. The relationships of zeaxanthin, antheraxanthin and pH to fluorescence quenching were examined in chloroplasts ofPisum sativum L. cv. Oregon andLactuca sativa L. cv. Romaine. Data matrices as five levels of violaxanthin de-epoxidation against five levels of light-induced lumen-proton concentrations were obtained for both species. The matrices included high levels of antheraxanthin as well as lumen-proton concentrations induced by subsaturating to saturation light levels. Analyses of the matrices by simple linear and multiple regression showed that quenching is predicted by models where the major independent variable is the product of lumen acidity and de-epoxidized xanthophylls, the latter as the sum of zeaxanthin and antheraxanthin. The interactions of lumen acidity and xanthophyll concentration are shown in three-dimensional plots of the best-fit multiple regression models. Antheraxanthin apparently contributes to quenching as effectively as zeaxanthin and explains quenching previously not accounted for by zeaxanthin. Hence, we propose that all high-energy dependent quenching is xanthophyll dependent. Quenching requires a threshold lumen pH that varies with xanthophyll composition. After the threshold, quenching is linear with lumen acidity or xanthophyll composition.  相似文献   

14.
The kinetics of non-photochemical quenching (NPQ) of chlorophyll fluorescence was studied in pea leaves at different temperatures between 5 and 25°C and during rapid jumps of the leaf temperature. At 5°C, NPQ relaxed very slowly in the dark and was sustained for up to 30 min. This was independent of the temperature at which quenching was induced. Upon raising the temperature to 25°C, the quenched state relaxed within 1 min, characteristic for qE, the energy-dependent component of NPQ. Measurements of the membrane permeability (ΔA515) in dark-adapted and preilluminated leaves and NPQ in the presence of dithiothreitol strongly suggest that the effect of low temperature on NPQ was not because of limitation by the lumenal pH or the de-epoxidation state of the xanthophylls. These data are consistent with the notion that the transition from the quenched to the unquenched state and vice versa involves a structural reorganization in the photosynthetic apparatus. An eight-state reaction scheme for NPQ is proposed, extending the model of Horton and co-workers (FEBS Lett 579:4201–4206, 2005), and a hypothesis is put forward concerning the nature of conformational changes associated with qE. Electronic supplementary material The online version of this article (doi: ) contains supplementary material, which is available to authorized users.  相似文献   

15.
Time-resolved fluorescence on oligomers of the main light-harvesting complex from higher plants indicate that in vitro oligomerization leads to the formation of a weakly coupled inter-trimer chlorophyll-chlorophyll (Chl) exciton state which converts in tens of ps into a state which is spectrally broad and has a strongly far-red enhanced fluorescence spectrum. Both its lifetime and spectrum show striking similarity with a 400ps fluorescence component appearing in intact leaves of Arabidopsis when non-photochemical quenching (NPQ) is induced. The fluorescence components with high far-red/red ratio are thus a characteristic marker for NPQ conditions in vivo. The far-red emitting state is shown to be an emissive Chl-Chl charge transfer state which plays a crucial part in the quenching.  相似文献   

16.
《Journal of bryology》2013,35(3):171-177
Abstract

Vascular plants are typically endohydric and are killed by drying beyond 30% relative water content. Bryophytes are ectohydric and are typically desiccation tolerant (DT). Mosses in open sun-exposed habitats show major electron flow to oxygen and high levels of non-photochemical quenching (NPQ) in chlorophyll fluorescence measurements. This has been regarded as a main source of photoprotection for these plants. The aim of the work described in this paper was to explore the rate and extent of relaxation of this quenching, and to seek evidence of its nature and consequences. Sequences of measurements were made during illumination at various intensities and a subsequent dark period. Light-response curves were constructed using dithiothreitol (DTT) as an inhibitor of violaxanthin de-epoxidase to provide additional evidence of the proportion of NPQ mediated by the xanthophyll cycle. The relaxation curves were fitted by exponential decay curves. A double-exponential fit to curves for the sun-adapted species gave a fast phase with a halflife of ca 6–16 seconds, and a slow phase with a halflife of ca 100–300 seconds. Shade species were best fitted by single-exponential curves. A persistent offset remained of ca 5–23% of the pre-darkening NPQ. Light-response curves for several species showed NPQ reduced in the presence of DTT to similar proportions of the control. Around 70–95% of NPQ in the bryophytes investigated relaxed with a halflife of ca 2–5 minutes. The fast phase of the double-exponential fit is consistent with likely rates of decay of the trans-thylakoid pH gradient and re-epoxidation of zeaxanthin. This leads to the same conclusion as the effect of DTT in depressing NPQ. The contrast in physiology between bryophytes and vascular plants reflects the different selection pressures facing leaf cells of poikilohydric plants and the mesophyll cells of vascular plants, and their divergent evolutionary histories since the mid-Palaeozoic.  相似文献   

17.
Diatoms are especially important microorganisms because they constitute the larger group of microalgae. To survive the constant variations of the light environment, diatoms have developed mechanisms aiming at the dissipation of excess energy, such as the xanthophyll cycle and the non-photochemical chlorophyll (Chl) fluorescence quenching. This contribution is dedicated to the relaxation of the latter process when the adverse conditions cease. An original nonlinear regression analysis of the relaxation of non-photochemical Chl fluorescence quenching, qN, in diatoms is presented. It was used to obtain experimental evidence for the existence of three time-resolved components in the diatom Phaeodactylum tricornutum: qNf, qNi and qNs. qNf (s time-scale) and qNs (h time-scale) are exponential in shape. By contrast, qNi (min time-scale) is of sigmoidal nature and is dominant among the three components. The application of metabolic inhibitors (dithiothreitol, ammonium chloride, cadmium and diphenyleneiodonium chloride) allowed the identification of the mechanisms on which each component mostly relies. qNi is linked to the relaxation of the ΔpH gradient and the reversal of the xanthophyll cycle. qNs quantifies the stage of photoinhibition caused by the high light exposure, qNf seems to reflect fast conformational changes within thylakoid membranes in the vicinity of the photosystem II complexes.  相似文献   

18.
To get an insight to the mechanism of the zeaxanthin-dependent non-photochemical quenching in photosystem II of photosynthesis, we probed the interaction of some xanthophylls with excited chlorophyll-a by trapping both pigments in micelles of triton X-100. Optimal distribution of pigments among micelles was obtained by proper control of the micelle concentration, using formamide in the reaction mixture, which varies the micellar aggregation number over three orders of magnitude. The optimal reaction mixture was obtained around 40% (v/v) formamide in 0.2-0.4% (v/v) triton X-100 in water. Zeaxanthin in the micellar solution exhibited initially absorption and circular dichroism spectral features corresponding to a J-type aggregate. The spectrum was transformed over time (half-time values vary-an average characteristic figure is roughly 20 min) to give features representing an H-type aggregate. The isosbestic point in the series of spectral curves favors the supposition of a rather simple reaction between two pure J and H-types dimeric species. Violaxanthin exhibited immediately stable spectral features corresponding to a mixture of J-type and more predominately H-type dimers. Lutein, neoxanthin and beta-carotene did not show any aggregated spectral forms in micelles. The spectral features in micelles were compared to spectra in aqueous acetone, where the assignment to various aggregated types was established previously. The specific tendency of zeaxanthin to form the J-type dimer (or aggregate) could be important for its function in photosynthesis. The abilities of five carotenoids (zeaxanthin, violaxanthin, lutein, neoxanthin and beta-carotene) to quench chlorophyll-a fluorescence were compared. Zeaxanthin, in its two micellar dimeric forms, and beta-carotene were comparable good quenchers of chlorophyll-a fluorescence. Violaxanthin was a much weaker quencher, if at all. Lutein and neoxanthin rather enhanced the fluorescence. The implications to non-photochemical quenching process in photosynthesis are discussed.  相似文献   

19.
Under natural conditions, photosynthesis has to be adjusted to fluctuating light intensities.Leaves exposed to high light dissipate excess light energy in form of heat at photosystem II(PSII) by a process called non-photochemical quenching(NPQ). Upon fast transition from light to shade, plants lose light energy by a relatively slow relaxation from photoprotection. Combined overexpression of violaxanthin de-epoxidase(VDE), PSII subunit S(PsbS) and zeaxanthin epoxidase(ZEP) in tobacco accelerates ...  相似文献   

20.
Ivanov B  Edwards G 《Planta》2000,210(5):765-774
 Non-photochemical quenching of chlorophyll fluorescence (NPQ) and quantum yield of photosystem II (PSII) were studied with intact mesophyll chloroplasts of maize (Zea mays L.) during the initial minutes of illumination using the pulse-modulated chlorophyll fluorescence technique. Non-photochemical quenching was rapidly reversible in the dark at any point during illumination, which is indicative of energy-dependent dissipation of energy (mediated via thylakoid ΔpH changes and ascorbate-dependent synthesis of zeaxanthin). In chloroplasts suspensions including 15 mM ascorbate in the medium, with addition of oxaloacetate and pyruvate, the PSII yield, rate of reduction of oxaloacetate and phosphorylation of pyruvate reached a maximum after approximately 2 min of illumination. Under these conditions, which promote phosphorylation and a decreased ΔpH across the thylakoid membrane, NPQ rose to a maximum after 2–3 min of illumination, dropped to a minimum after about 6 min, and then increased to a steady-state level. A rather similar pattern was observed when leaves were illuminated following a 30-min dark period. Providing chloroplasts with higher levels of ascorbate (60 mM), prevented the transient drop in NPQ. Anaerobic conditions or addition of potassium cyanide caused a decrease in PSII yield, providing evidence for operation of the ascorbate-dependent Mehler-peroxidase reaction. These conditions also strongly suppressed the transient drop in NPQ. Dithiothreitol, an inhibitor of violaxanthin de-epoxidase, caused a large drop in NPQ even in the presence of high levels of ascorbate. The results suggest that the decline of NPQ occurs in response to an increase in lumen pH after initiation of phosphorylation, that this decline can be suppressed by conditions where ascorbate is not limiting for violaxanthin de-epoxidase, and that the increase of NPQ after such a decline is the result of development of energy dissipation in PSII reaction centers. Received: 13 August 1999 / Accepted: 17 September 1999  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