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.
2.
Watching the photosynthetic apparatus in native membranes   总被引:7,自引:0,他引:7  
Over the last 9 years, the structures of the various components of the bacterial photosynthetic apparatus or their homologues have been determined by x-ray crystallography to at least 4.8-A resolution. Despite this wealth of structural information on the individual proteins, there remains an urgent need to examine the architecture of the photosynthetic apparatus in intact photosynthetic membranes. Information on the arrangement of the different complexes in a native system will help us to understand the processes that ensure the remarkably high quantum efficiency of the system. In this work we report images obtained with an atomic force microscope of native photosynthetic membranes from the bacterium Rhodospirillum photometricum. Several proteins can be seen and identified at molecular resolution, allowing the analysis and modeling of the lateral organization of multiple components of the photosynthetic apparatus within a native membrane. Analysis of the distribution of the complexes shows that their arrangement is far from random, with significant clustering both of antenna complexes and core complexes. The functional significance of the observed distribution is discussed.  相似文献   

3.
The first step of photosynthesis in plants is the absorption of sunlight by pigments in the antenna complexes of photosystem II (PSII), followed by transfer of the nascent excitation energy to the reaction centers, where long-term storage as chemical energy is initiated. Quantum mechanical mechanisms must be invoked to explain the transport of excitation within individual antenna. However, it is unclear how these mechanisms influence transfer across assemblies of antenna and thus the photochemical yield at reaction centers in the functional thylakoid membrane. Here, we model light harvesting at the several-hundred-nanometer scale of the PSII membrane, while preserving the dominant quantum effects previously observed in individual complexes. We show that excitation moves diffusively through the antenna with a diffusion length of 50 nm until it reaches a reaction center, where charge separation serves as an energetic trap. The diffusion length is a single parameter that incorporates the enhancing effect of excited state delocalization on individual rates of energy transfer as well as the complex kinetics that arise due to energy transfer and loss by decay to the ground state. The diffusion length determines PSII’s high quantum efficiency in ideal conditions, as well as how it is altered by the membrane morphology and the closure of reaction centers. We anticipate that the model will be useful in resolving the nonphotochemical quenching mechanisms that PSII employs in conditions of high light stress.The first step of photosynthesis is light harvesting, the absorption and conversion of sunlight into chemical energy. In photosynthetic organisms, the functional units of light harvesting are self-assembled arrays of pigment–protein complexes called photosystems. Antenna complexes absorb and transfer the nascent excitation energy to reaction centers, where long-term storage as chemical energy is initiated (1). In plants, photosystem II (PSII) flexibly responds to changes in sunlight intensity on a seconds to minutes time scale. In dim light, under ideal conditions, PSII harvests light with a >80% quantum efficiency (2), whereas, in intense sunlight PSII dissipates excess absorbed light safely as heat via nonphotochemical quenching pathways (3). The ability of PSII to switch between efficient and dissipative states is important for optimal plant fitness in natural sunlight conditions (4). Understanding how PSII’s function arises from the structure of its constituent pigment−protein complexes is a prerequisite for systematically engineering the light-harvesting apparatus in crops (57) and could be useful for designing artificial materials with the same flexible properties (8, 9).Recent advances have established structure−function relationships within individual pigment−protein complexes, but not how these relationships affect the functioning of the dynamic PSII (grana) membrane (10). Electron microscopy and fitting of atomic resolution structures (11) place the pigment−protein complexes in the grana membrane in close proximity, enabling long-range transport. Indeed, connectivity of excitation between different PSII reaction centers has been discussed since 1964 (12), suggesting that the functional unit for PSII must involve a large area of the membrane. Two limiting cases have been used to model PSII light harvesting: The lake model assumes perfect connectivity between reaction centers across the membrane; alternatively, the membrane can be described as a collection of disconnected “puddles” of pigments that each contain one reaction center (1, 13). At present, however, resolving the spatiotemporal dynamics within the grana membrane on the relevant length (tens to hundreds of nanometers) and time (1 ps to 1 ns) scales experimentally is not possible. Structure-based modeling of the grana membrane, however, can access this wide range of length and time scales.The dense packing of the major light-harvesting antenna (LHCII, discs), which is a trimeric complex, and PSII supercomplexes (PSII-S, pills) in the grana membrane is shown in Fig. 1 A and B. PSII-S is a multiprotein complex (14) that contains the PSII core reaction center dimer, along with several minor light-harvesting complexes and LHCII (Fig. 1A, Inset). Electronic excited states in LHCII and PSII-S are delocalized over several pigments (1517), making conventional Förster theory inadequate to describe the excitation dynamics. On the protein length scale, generalized Förster (18, 19) calculations between domains of tightly coupled chlorophylls agree very well with more exact methods [e.g., the zeroth-order functional expansion of the quantum-state diffusion model (ZOFE) approximation to non-Markovian quantum state diffusion (20)] for simulating the excitation population dynamics (21, 22). This agreement suggests that the primary quantum phenomenon involved in PSII energy transfer is the site basis coherence that arises from excited states delocalized across a few (approximately three to four) pigments.Open in a separate windowFig. 1.Accurate simulation of chlorophyll fluorescence dynamics from thylakoid membranes using structure-based modeling of energy transfer in PSII. (A and B) The representative mixed (A) and segregated (B) membrane morphologies generated using Monte Carlo simulations and used throughout this work. PSII-S are indicated by the light teal pills, and LHCII, which are trimeric complexes, are indicated by the light grey-green circles. The segregated membrane forms PSII-S arrays and LHCII pools. As shown schematically in A (Bottom) existing crystal structures of PSII-S (14) and LHCII (24) were overlaid on these membrane patches to establish the locations of all chlorophyll pigments. The light teal and light grey-green dashed lines outline the excluded area associated with PSII-S and LHCII trimers respectively, in the Monte Carlo simulations. The chlorophyll pigments are indicated in green, and the protein is depicted by the grey cartoon ribbon. PSII-S is a twofold symmetric dimer of pigment−protein complexes that are outlined by black lines. LHCII-S (strongly bound LHCII), CP26, CP29, CP43, and CP47 are antenna proteins, and RC indicates the reaction center. The inhomogeneously averaged rates of energy transfer between strongly coupled clusters of pigments were calculated using generalized Förster theory. (C) Simulated fluorescence decay of the mixed membrane (solid black line) and the PSII component of experimental fluorescence decay data from thylakoid membranes from ref. 26 (red, dotted line). Inset shows the lifetime components and amplitudes of the simulated decay as calculated using our model with a Gaussian convolution (σ = 20 ps) (black line) or by fitting to three exponential decays (green bars).Here, we construct a generalized Förster model for the ∼104 pigments covering the few-hundred-nanometer length scale of the grana membrane that correctly incorporates the dynamics occurring within and between complexes on the picosecond time scale. We show how delocalized excited states, or excitons, in individual complexes affect light harvesting on the membrane length scale. The formation of excitons is sufficient to explain the high quantum efficiency of PSII in dim light. The model, by being an accurate representation of the complex kinetic network that underlies PSII light harvesting, provides mechanistic explanations for long-observed biological phenomena and sets the stage for developing a better understanding of PSII light harvesting in high light conditions.  相似文献   

4.
2-Acetoxymethyl-1,4-naphthoquinone (2-AcOMeNQ) binds with rapid kinetics and high affinity to the primary quinone QA site of reaction centers from Rhodopseudomonas capsulata. Binding of 2-AcOMeNQ fully restores electron-transfer activity with kinetics that is similar, but not identical, to that seen with ubiquinone-50. When bound at the QA site, 2-AcOMeNQ preferentially labels the L subunit. This preference suggests that 2-AcOMeNQ labels primarily the region of a quinone-binding site that is close to the first isoprenoid unit of the side chain, which is expected from the location and structure of the reaction region of the molecule. In photosystem II particles from Synechococcus sp., 2-AcOMeNQ primarily labels two polypeptides with apparent molecular masses of 38 and 19 kDa. Labeling of only the 38-kDa polypeptide is sufficiently sensitive to 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU) to conclude that it is involved in binding quinones on the acceptor side of photosystem II. Although we have not yet identified the 38-kDa protein, its properties suggest that it is the D2 protein. From the DCMU-sensitive labeling and from homologies to functionally important regions of the bacterial reaction-center subunits, we propose that the 38-kDa protein is intimately involved in binding the cofactors that mediate primary photochemistry.  相似文献   

5.
We consider a model of the photosystem II (PS II) reaction center in which its spectral properties result from weak (approximately 100 cm-1) excitonic interactions between the majority of reaction center chlorins. Such a model is consistent with a structure similar to that of the reaction center of purple bacteria but with a reduced coupling of the chlorophyll special pair. We find that this model is consistent with many experimental studies of PS II. The similarity in magnitude of the exciton coupling and energetic disorder in PS II results in the exciton states being structurally highly heterogeneous. This model suggests that P680, the primary electron donor of PS II, should not be considered a dimer but a multimer of several weakly coupled pigments, including the pheophytin electron acceptor. We thus conclude that even if the reaction center of PS II is structurally similar to that of purple bacteria, its spectroscopy and primary photochemistry may be very different.  相似文献   

6.
7.
A child heterozygous for the genes for hemoglobins S, C and G alpha Philadelphia presented with a clinical picture similar to sickle cell anemia. Her hemoglobin electrophoretic pattern contained three components with the mobilities of hemoglobins S (35 per cent), C (47 per cent) and a more slowly migrating hybrid G/C molecule (15 per cent). Seven relatives were heterozygous for Hb G beta and Hb S beta and five were heterozygous only for Hb G alpha. Among the latter, three had approximately 30 per cent and two had 40 per cent of Hb G. These proportions are consistent with the hypothesis that the American Negro genome contains two types of chromosomes bearing structural loci for alpha chains, some possessing one Hb alpha locus, others having two loci. Hb G alpha-Philadelphia presumably arose as a mutation on a chromosome with a single locus. Those heterozygotes having 30 per cent and 40 per cent Hb G presumably have two loci and only one locus, respectively, on the homologous chromosome.  相似文献   

8.
J Yl?nne  D A Cheresh  I Virtanen 《Blood》1990,76(3):570-577
The localization of five integrin subunit proteins was studied in human erythroleukemia (HEL) cells spreading on various culture substrata in the presence of 12-O-tetradecanoylphorbol-13-acetate (TPA) and the absence of serum. The cells readily adhered on fibronectin, but TPA was needed for adherence on vitronectin and for the spreading of the cells on both substrata. Indirect immunofluorescence microscopy showed that in the spread cells cultured on vitronectin or fibronectin for 2 hours, beta 1, beta 3, alpha 5, and alpha IIb integrin subunits were localized at focal adhesions as identified by talin-immunoreactivity. The alpha v integrin immunoreactivity was initially found at the focal adhesions when the cells were cultured on vitronectin, but was also found later in cells cultured on fibronectin. The alpha IIb integrin immunoreactivity disappeared from focal adhesions within 24 hours. The alpha 5 and beta 1 integrin immunoreactivities disappeared from the focal adhesions in cells cultured on vitronectin, but not in cells cultured on fibronectin. When the cells were plated on glass substratum in the presence of TPA, they spread much slower than on vitronectin or fibronectin, but some cells showed focal adhesions after only 8 hours in culture. In this case, the alpha v and beta 3 integrin subunits were found at focal adhesions. After TPA treatment, HEL cells deposited thrombospondin-immunoreactive material onto their culture substratum, but synthesis of fibronectin, vitronectin, fibrinogen, or von Willebrand factor was not detected. Thus, the results suggest that TPA would activate several integrin receptors in HEL cells and also stimulate the secretion of thrombospondin, which might be used as an adhesion ligand for the integrin vitronectin receptor alpha v/beta 3 complex.  相似文献   

9.
The catabolism of alpha 2- and beta-very-low-density lipoproteins (VLDL) was studied in normolipidemic and hyperlipidemic subjects to determine whether differences in the catabolism of these subfractions are due to their composition. alpha 2-VLDL (cholesterol/triglyceride ratio, 00.18 +/- 0.06; and apoprotein E/C ratio, 0.27 +/- 0.22, n = 4) and beta-VLDL (cholesterol/triglyceride ratio, 0.67 +/- 0.13; and apoprotein E/C ratio, 1.05 +/- 0.52, n = 4) were isolated from subjects with broad beta disease, iodinated, and injected in five normolipidemic subjects, six with broad beta disease, and five with endogenous hypertriglyceridemia. VLDL, intermediate (IDL) and low-density lipoprotein (LDL) apoprotein (apo)-B radioactivity (tetramethylurea insoluble) following injection of 125I-labeled alpha 2- and beta-VLDL decayed biphasically in all subjects, and this decay in normolipidemic subjects was more rapid than in subjects with broad beta disease (P = 0.004) or endogenous hypertriglyceridemia (P = 0.004 for alpha 2- and P = 0.010 for beta-VLDL). The residence times, however, for the delipidation chain in alpha 2-VLDL were similar in all the subjects and varied from three to six hours. The decay of radioactivity in beta-VLDL in subjects with broad beta disease was much slower (residence time, 36.9 +/- 24.4 hr, n = 7) than in normolipidemic subjects (residence time, 7.56 +/- 4.6 hr, n = 5) or in subjects with endogenous hypertriglyceridemia (residence time, 10.6 +/- 4.65, n = 4). The residence time for alpha 2-VLDL was longer than for beta-VLDL in all subjects, suggesting that alpha 2-VLDL is a precursor to beta-VLDL. To test this directly, iodinated alpha 2-VLDL was injected into a subject with broad beta disease and the radioactivity in the subfractions was followed. The radioactivity from alpha 2-VLDL was transferred into beta-VLDL supporting, the notion that alpha 2-VLDL generated some beta-VLDL. Nicotinic acid treatment of a subject with broad beta disease accelerated the catabolism of alpha 2- and beta-VLDL without changing the VLDL composition.  相似文献   

10.
The blood schizontocidal activity of alpha and beta arteether has been compared with that of alpha/beta arteether (a 30:70 mixture of alpha and beta isomers), which is a fast-acting blood schizontocide undergoing phase I clinical trials at the Central Drug Research Institute. Both beta and alpha/beta arteether have comparable activity and are curative at a dose of 5 mg/kg for 3 days against blood-induced Plasmodium cynomolgi B infection in the rhesus monkey; alpha arteether alone is slightly less active, with a 50% cure rate at the above dose.  相似文献   

11.
12.
Molecular heterogeneity of beta thalassaemia in the Italian population   总被引:1,自引:0,他引:1  
S ummary Fifty-one subjects originating from Southern Italy and affected by Cooley's anaemia have been studied in order to define the degree of heterogeneity of β thalassaemia mutations in this high incidence area. Restriction endonuclease mapping has been carried out on genomic DNA by the Southern blot technique both to exclude the existence of gross deletions or rearrangements and to establish the relative frequency of four polymorphic restriction sites (i.e. Gγ and Aγ Hind III, β Ava II and β Bam HI) within the γδβ gene region. In 28 subjects unequivocal linkage of the four polymorphic sites has been determined leading to the identification of seven different chromosome haplotypes, six of which had previously been reported associated with specific β0 and β+ thalassaemia mutations. Globin chain synthesis studies on peripheral blood reticulocytes indicated that subjects carrying the same genotype may behave differently as far as the β chain production is concerned relative to both the a and the non-α chains. Thus, β thalassaemia turns out to be quite heterogeneous even in this limited geographical area. β+ mutations appear to be predominant, particularly those affecting nuclear precursor RNA splicing to mature β globin mRNA.  相似文献   

13.
The yield of energy transfer from photosystem II to photosystem I in Porphyridium cruentum varies from a minimum value of about 0.50 when the photosystem II reaction centers are all open to a maximum value between 0.90 and 0.95 when the centers are all closed.  相似文献   

14.
Fibrinogen fraction I (340 kDa) and fraction II (305 kDa) were isolated by glycine precipitation. The subunit chains of the two fractions were separated, after reduction, by reverse-phase high performance liquid chromatography. The amino acid compositions of the B beta and tau chains of fibrinogen II were identical with those of fibrinogen I. In contrast, the A alpha chains of fibrinogen II were composed of two populations, one comprising homogeneous, intact A alpha chains and the other consisting of heterogeneous, deficient A alpha chains (A alpha' chains) of lengths varying according to the sizes of their COOH-terminal defects. The molar ratio of the A alpha to the A alpha' chains in fibrinogen II was 1.16:1. The amino acid composition and sequence analyses of the TPCK-trypsin peptides derived from the A alpha' chains revealed that the COOH-terminal residues of the A alpha' chains were mainly Asn-269, Gly-297 and Pro-309. These results indicate that the fibrinogen II molecule is asymmetrical and can be represented by the formula (A alpha) (A alpha')(B beta)2(tau)2 and that fibrinogen II cannot be a plasmin degradation product of fibrinogen I.  相似文献   

15.
16.
We have reported telomere attrition in β and α cells of the pancreas in elderly patients with type 2 diabetes, but it has not been explored how the telomere lengths of these islet cells change according to age in normal subjects. To examine the telomere lengths of β and α cells in individuals without diabetes across a wide range of ages, we conducted measurement of the telomere lengths of human pancreatic β and α cells obtained from 104 autopsied subjects without diabetes ranging in age from 0 to 100 years. As an index of telomere lengths, the normalized telomere-centromere ratio (NTCR) was determined for β (NTCRβ) and α (NTCRα) cells by quantitative fluorescence in situ hybridization (Q-FISH). We found NTCRβ and NTCRα showed almost the same levels and both decreased according to age (p < 0.001 for both). NTCRs decreased more rapidly with age and were more widely distributed (p = 0.036 for NTCRβ, p < 0.001 for NTCRα) in subjects under 18 years of age than in subjects over 18 years. There was a positive correlation between NTCRβ and NTCRα only among adult subjects (p < 0.001). In conclusion, the telomeres of β and α cells become shortened with normal aging process.  相似文献   

17.
Photosystem II (PSII) catalyzes the light-driven oxidation of water and the reduction of plastoquinone; the oxidation of water occurs at a cluster of four manganese. The PSII CP43 subunit functions in light harvesting, and mutations in the fifth luminal loop (E) of CP43 have established its importance in PSII structure and/or assembly [Kuhn, M. G. & Vermaas, V. F. J. (1993) Plant Mol. Biol. 23, 123-133]. The sequence A(350)PWLEPLR(357) in luminal loop E is conserved in CP43 genes from 50 organisms. To map important posttranslational modifications in this sequence, tandem mass spectrometry (MS/MS) was used. These data show that the indole side chain of Trp-352 is posttranslationally modified to give mass shifts of +4, +16, and +18 daltons. The masses of the modifications suggest that the tryptophan is modified to kynurenine (+4), a keto-/amino-/hydroxy- (+16) derivative, and a dihydro-hydroxy- (+18) derivative of the indole side chain. Peptide synthesis and MS/MS confirmed the kynurenine assignment. The +16 and +18 tryptophan modifications may be intermediates formed during the oxidative cleavage of the indole ring to give kynurenine. The site-directed mutations, W352C, W352L, and W352A, exhibit an increased rate of photoinhibition relative to wild type. We hypothesize that Trp-352 oxidative modifications are a byproduct of PSII water-splitting or electron transfer reactions and that these modifications target PSII for turnover. As a step toward understanding the tertiary structure of this CP43 peptide, structural modeling was performed by using molecular dynamics.  相似文献   

18.
Two symmetrically positioned redox active tyrosine residues are present in the photosystem II (PSII) reaction center. One of them, TyrZ, is oxidized in the ns-micros time scale by P680+ and reduced rapidly (micros to ms) by electrons from the Mn complex. The other one, TyrD, is stable in its oxidized form and seems to play no direct role in enzyme function. Here, we have studied electron donation from these tyrosines to the chlorophyll cation (P680+) in Mn-depleted PSII from plants and cyanobacteria. In particular, a mutant lacking TyrZ was used to investigate electron donation from TyrD. By using EPR and time-resolved absorption spectroscopy, we show that reduced TyrD is capable of donating an electron to P680+ with t1/2 approximately equal to 190 ns at pH 8.5 in approximately half of the centers. This rate is approximately 10(5) times faster than was previously thought and similar to the TyrZ donation rate in Mn-depleted wild-type PSII (pH 8.5). Some earlier arguments put forward to rationalize the supposedly slow electron donation from TyrD (compared with that from TyrZ) can be reassessed. At pH 6.5, TyrZ (t1/2 = 2-10 micros) donates much faster to P680+ than does TyrD (t1/2 > 150 micros). These different rates may reflect the different fates of the proton released from the respective tyrosines upon oxidation. The rapid rate of electron donation from TyrD requires at least partial localization of P680+ on the chlorophyll (PD2) that is located on the D2 side of the reaction center.  相似文献   

19.
Plants and green algae have a low pH-inducible mechanism in photosystem II (PSII) that dissipates excess light energy, measured as the nonphotochemical quenching of chlorophyll fluorescence (qE). Recently, nonphotochemical quenching 4 (npq4), a mutant strain of the green alga Chlamydomonas reinhardtii that is qE-deficient and lacks the light-harvesting complex stress-related protein 3 (LHCSR3), was reported [Peers G, et al. (2009) Nature 462(7272):518–521]. Here, applying a newly established procedure, we isolated the PSII supercomplex and its associated light-harvesting proteins from both WT C. reinhardtii and the npq4 mutant grown in either low light (LL) or high light (HL). LHCSR3 was present in the PSII supercomplex from the HL-grown WT, but not in the supercomplex from the LL-grown WT or mutant. The purified PSII supercomplex containing LHCSR3 exhibited a normal fluorescence lifetime at a neutral pH (7.5) by single-photon counting analysis, but a significantly shorter lifetime at pH 5.5, which mimics the acidified lumen of the thylakoid membranes in HL-exposed chloroplasts. The switch from light-harvesting mode to energy-dissipating mode observed in the LHCSR3-containing PSII supercomplex was sensitive to dicyclohexylcarbodiimide, a protein-modifying agent specific to protonatable amino acid residues. We conclude that the PSII-LHCII-LHCSR3 supercomplex formed in the HL-grown C. reinhardtii cells is capable of energy dissipation on protonation of LHCSR3.  相似文献   

20.
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