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.
A photosystem (PS) I holocomplex was isolated from Pleurochloris meiringensis Vischer (Xanthophyceae) using sucrose density centrifugation. This complex exhibited a fluorescence emission maximum at 715 nm, which is in accordance with the long wavelength emission of whole cells. The complex was further dissociated into a core complex and a light-harvesting protein (LHC I). The core protein contains mainly Chl a and β-carotene, is 8.25 times enriched in P700 and has its main emission maximum at 715 nm. Therefore, the longest wavelength emission of P. meiringensis is due to the PS I core, which is in contrast to higher plants. The LHC I differs from LHC II with regard to its polypeptide pattern as well as its spectral properties. The arrangement of antennae is discussed in relation to the regulation of energy transfer between the photosystems.  相似文献   

2.
Abstract—The absolute photoelectron quantum yield spectrum for β-carotene in the wavelength range 180–230 nm is reported. β-Carotene is very photoemissive over this wavelength region with photoelectron quantum yields on the order of 2 × 10--3 electrons per incident photon at 180 nm, 4 × 10--4 at 210 nm, and 3 × 10--5 at 230 nm. At wavelengths longer than 240 nm, β-carotene photoemission dropped off monotonically with increasing wavelength. The photoelectron quantum yield spectrum of β-carotene is very similar to that of chlorophyll a . A photoelectron micrograph of β-carotene deposited on a thin layer of the fatty acid arachidic acid demonstrates the marked photoemission contrast between β-carotene and membrane lipid. Photoelectron micrographs of samples with β-carotene and Chl a in the same field show that prolonged (1 h) illumination in vacuo causes β-carotene to markedly fade while leaving the Chl a photoemissive. This differential bleaching of β-carotene may allow it to be distinguished from Chl in high magnification photoelectron micrographs of photosynthetic membranes.  相似文献   

3.
The influence of UV radiation (253.7 nm) on collagen fluorescence in the absence, and presence, of β-carotene was investigated. It was found that UV radiation of 253.7 nm causes irreversible destruction of tyrosyl and phenylalanyl residues. The fluorescence of collagen (excitation at 275 nm, emission at 305 nm) decreased rapidly during irradiation and a new fluorescence large band at 400–500 nm formed under UV radiation. Smaller changes in the fluorescence of collagen in the presence of β-carotene suggest that it makes collagen less sensitive to the action of UV radiation.  相似文献   

4.
Abstract— Mono- and multilayers of chlorophyll a (Chl a )– lecithin have been prepared on quartz slides, by means of the Blodgett-Langmuir technique, for fluorescence studies. Self-quenching of the Chl a fluorescence has been observed in Chl a -lecithin single layer excited with a laser light at 632.8 nm. The fluorescence yield is reduced by 50% at a concentration of 7 ± 1012 Chl a molecules cm−2. Chl a fluorescence quenching, by adding N,N -distearoyl-1,4-diaminoanthraquinone (SAQ), has been studied. in a single layer, in pure Chl a and also at various dilutions of Chl a in lecithin. The results are explained in terms of a dynamic quenching rather than in terms of a permanent complex formation, at the ground state, between Chl a and SAQ. The fluorescence quenching has been interpreted as the result of an electron transfer from excited Chl a to SAQ, and rate constants of 8.3 ± 10−5 cm2 molecule−1 S−1 and 2.4 ± 10−4 cm2 molecule−1 s−1 have been found for pure diluted Chl a , respectively. Ten per cent of the diluted Chl a fluorescence always remains unquenchable and independent of the quinone concentration. In multilayers, where SAQ and Chl a are in different layers, there is no fluorescence quenching for pure or diluted Chl a even when the chromophores are in two adjacent layers. This happens only if SAQ is not able to diffuse from one layer to another. A minimum value of 22.4 nm has been found for the singlet exciton diffusion length in pure Chl a multilayers.  相似文献   

5.
The fraction FΣ of excited-state oxygen formed as b 1Σg+ was determined for a series of triplet-state photosensitizers in CCl4 solutions. FΣ was determined by monitoring the intensities of (a) O2(b 1Σg+) fluorescence at 1926 nm (O2(b 1Σg+)→O2(a 1Δg) and (b) O2(a 1 Δg) phosphorescence at 1270 nm (O2(a 1Δg) → O2(X3Σg)). Oxygen excited states were formed by energy transfer from substituted benzophenones and acetophenones. The data indicate that FΣ depends on several variables including the orbital configuration of the lowest triplet state and the triplet-state energy. The available data indicate that the sensitizer-oxygen charge transfer (CT) state is not likely to influence FΣ strongly by CT-mediated mixing of various sensitizer-oxygen states.  相似文献   

6.
Using N3 species as specific electron acceptor a defined ascorbate radical: AH↔A+H+max=360 nm, =3400 dm3 mol−1 cm−1) is observed. The attack of DMSO+ on vit.E results in a vit.E radical (k=1×109 dm3 mol−1 s−1; λmax=425 nm, =2400 dm3 mol−1 cm−1; 2k=4.7×108 dm3 mol−1 s−1). Vit.E-acetate leads to the formation of a radical cation (vit.E-ac+). β-carotene reacts also with DMSO+ forming a radical cation, β-car+ (k=1.75×108 dm3 mol−1 s−1; λmax=942 nm, =14 600 dm3 mol−1 cm−1), which probably leads to the formation of a dimer radical cation, (β-car)+2 (k=2.5×107 dm3 mol−1 s−1).

Using E.coli bacteria (AB1157) as a model system in vitro it was found that all three vitamins are rather efficient radiation protecting agents. They can also increase the activity of cytostatica, e.g., mitomycin C (MMC), by electron transfer process. The mixture of vit.E-ac and β-car acts contradictory, but adding vit.C to it a strong cooperative enhancement of the MMC activity is observed once again. A relationship between the pulse radiolysis and the radiation biological data is found and discussed. A possible explanation of the previously reported trials concerning the role of vit.E and β-car on the increased occurence of lung and other types of cancer in smokers and drinkers is presented.  相似文献   


7.
The effects of acid and alkali treatment on the light absorption, energy transfer and protein secondary structure of the photosystem II core antenna CP43 and CP47 of spinach were investigated by the absorption spectra, fluorescence emission spectra and ciruclar dichroism spectra. It has been found that acid treatment caused the appearance of absorption characteristic of pheophytin a (Pheo a), whereas alkali treatment induced a new absorption peak at 642 nm. The energy transfer between β-carotene and chlorophyll a (Chl a) in CP43 was easily disturbed by alkali, whereas in CP47 was readily affected by acid. As to the effects on the secondary structure of proetins in CP43 and CP47, effects of acid were far less than those of alkali. Both acid and alkali disturbed the microenvironment of Chl a and interfered exciton interaction between Chl a molecules. It was suggested that acid and alkali affect the light absorption, energy transfer and protein secondary structure of CP43 and CP47 in a differenty way. H+ can permeate into the internal space of α-helix, change Chl a into Pheo a and disturb the microenvironment of pgiments without damaging the secondary structure of protein, whereas OH can induce the protein unfolding at first, then saponify Chl a to chlorophyllide and disturb the microenvironment of pigments.  相似文献   

8.
In plants, excess irradiation can damage the photosynthetic apparatus, although some protective mechanisms exist. The excess energy can be dissipated as thermal energy, and pigments (i.e., carotenoids) also play an important role in protecting the photosynthetic apparatus by epoxidating reactions. Chromatographic analysis of pigment extracts of Gracilaria tenuistipitata shows that zeaxanthin is the major carotenoid in this alga, accounting for up to 82% of total carotenoids. Short-term (55 h) and long-term (10 days) response of the pigments shows that Chl a, β-carotene and zeaxanthin degradation after light increase follows negative exponential trends, while the response of biliproteins is almost linear. Decreasing the irradiance results in a clear saturating response of the synthesis of Chl a and β-carotene after one to two days. Biliprotein synthesis displays a double linear trend, the first one lasting for four days in the cases of both R-phycoerythrin (RPE) and R-phycocyanin (RPC). The response of zeaxanthin is always faster than that of Chl a or biliproteins to changes of irradiance. Our results might indicate the presence of two pools of zeaxanthin in this alga, with different acclimation responses to the changes in the photon flux density.  相似文献   

9.
To study the role of the long-wavelength chlorophylls (Chl) in photosystem I (PSI), the action spectra of P700 photooxidation at 293 and 77 K have been measured for PSI trimeric and monomeric complexes isolated from Spirulina platensis. The long-wavelength Chls which absorb in the region 710dash740 nm transfer excitation energy to the reduced P700 with the same efficiency as bulk antenna Chls, causing the oxidation of P700. The relative quantum yield of P700 photooxidation is about unity (293-77 K) even under the direct excitation of Chl absorbing at 735 nm (Chl735). At 77 K Chl735 exhibits a fluorescence band at 760 nm (F760) whose intensity is quenched under illumination of the PSI trimeric complexes from Spirulina. The relative quantum yield of F760 quenching is not dependent on the wavelength of excitation in the region 620–750 nm. Since the value of the overlap integral between the band of F760 and the absorption band of the cation radical of P700 (P700+) is higher than that of the P700 band, it is suggested that Chl735 transfers energy to P700+ more efficiently than to reduced P700; energy transfer to P700+ causes the quenching of F760. A linear relationship between the photooxidation rate of P700 and the fraction of P700+ at 293 K indicates that the energy exchange between PSI subunits of the trimer is negligible. Thus, the antenna of PSI trimers of Spirulina is organized in separate photosynthetic units.  相似文献   

10.
研究了黄豆黄素和黄豆黄苷在不同pH条件下的吸收光谱和荧光光谱, 从分子结构的角度解释了二者呈现不同光谱特征的原因. 黄豆黄素分子基本无荧光. 在弱碱性时, 黄豆黄素分子发生7-OH质子的电离, 导致吸收光谱中320 nm的吸收峰红移至348 nm. 采用pH-光度法测得7-OH质子的电离常数pKa1=7.08±0.04. 黄豆黄素一价阴离子呈现较强荧光, 最大激发和发射波长λex/λem分别为334 nm/464 nm, 荧光量子产率为0.049. 在碱性溶液中, 黄豆黄素4'-OH质子电离, 导致吸收光谱中254 nm的吸收峰红移至260 nm, 电离常数pKa2=9.96±0.01. 黄豆黄苷分子基本无荧光. 在碱性条件下, 黄豆黄苷分子的4'-OH质子发生电离, 导致吸收光谱中256 nm的吸收峰红移至 280 nm, 电离常数pKa=9.81±0.03. 黄豆黄苷阴离子基本无荧光, 但在热碱性条件下发生γ-吡喃酮环裂解反应而产生较强荧光, λex/λem为288 nm/388 nm, 裂解产物的荧光量子产率为0.056. 虽然, 黄豆黄苷与黄豆黄素是苷与苷元的关系, 但黄豆黄苷不能在热碱性条件下通过糖苷水解转变为黄豆黄素, 二者的荧光增强机理存在本质不同.  相似文献   

11.
Abstract. The efficiency of singlet-singlet energy transfer was studied in bilayer lipid membrane vesicles (liposomes) for the following donor-acceptor systems: (1) p -terphenyl (TP) and diphenyloctatetraene (DPO); (2) DPO and chlorophyll a (Chl a ); and (3) β-carotene and Chl a. The energy transfer efficiency φDA was measured by sensitized fluorescence of the acceptor. Fractional quenching of the donor φQ was found from the donor fluorescence in absence and presence of the acceptor. For TP-DPO and for DPO-Chl a , the transfer efficiency increased with increasing acceptor concentration but was essentially independent of the donor concentration. No energy transfer from β-carotene to Chl a could be detected. In liposomes, φDA differed only slightly from φQ at all donor and acceptor concentrations, thus demonstrating the absence of any appreciable energy losses. For solutions of the same donor-acceptor pairs in cyclohexane φQ was considerably larger than φDA. The difference represents energy lost, principally by internal conversion, due to collisional quenching. The principal function of the lipid membrane appears to be the suppression of such losses. In addition, the rate of energy transfer in lipid membranes is about double that in solutions (at the same intermolecular distance) due to more favorable orientation.  相似文献   

12.
A noise elimination technique was applied to the determination of binary diffusion coefficients D12 from the response curves having extremely low absorbance intensities in impulse response methods under supercritical conditions of carbon dioxide. The effectiveness of this technique was experimentally examined for the analyses of response curves through both the curve-fitting and the moment methods in two cases: the chromatographic impulse response method for phenol and β-carotene with a polymer-coated capillary column, and the Taylor dispersion method for acetone with an uncoated capillary column. Unreliable D12 values were obtained from the moment method of the response curves at lower absorbance intensities, even treated with noise elimination. The curve-fitting method with the noise elimination treatment was quite effective for determining the D12 values accurately, and was valid at the lowest absorbance intensities, on the order of 10−4 absorbance unit of UV-Vis multi-detector, corresponding to the smallest quantity of the solute, i.e. 6×10−5, 6×10−6, and 5×10−2 μ mol for phenol, β-carotene, and acetone, respectively, under conditions studied. Infinite dilution regions for binary diffusion coefficients were obtained by injecting various amounts: the binary diffusion coefficients showed constant values at concentrations less than 0.6, 0.004, and 0.08 mol m−3 for phenol, β-carotene, and acetone, respectively, in supercritical carbon dioxide at 313.2 K and 16–18 MPa.  相似文献   

13.
Raman scattering and relaxed fluorescence is observed upon cw laser excitation resonant with the lower vibrational manifold of the X(1Σo+u) → B(3Πo+u) transition of matrix isolated Br2 at liquid He temperatures. The excitation profile of the relaxed fluorescence maps out the resonances, but neither detectable enhancement of Raman scattering nor resonance fluorescence is observed.  相似文献   

14.
We describe a simple model for P680 sensitized excitation in photosynthesis. Chl a fluorescence quenching effects observed when water is added to Chl a solutions in acetone are shown to be the result of resonant transfer between acetone solvates of monomeric Chl a, Chl a·Ac, and dimers of Chl a dihydrate. The presence of (Chl a·2H2O)2 is evidenced by a 678 nm difference absorbance (ΔA band obtained on conversion of a 680 nm absorption shoulder to polycrystalline Chl a precipitate, (Chl a·H2O)n. The equilibration between (Chl a·2H2O)2 and Chl a·Ac as a principal mechanism for Chl a·Ac fluorescence quenching is supported by theoretical fits of the data.  相似文献   

15.
Single and multiple photon processes are identified in the 193 nm excimer laser photolysis of CS2. CS(X1Σ+, υ = 1 to 5, J = 5 to 45) is observed by dye laser induced fluorescence of the A1Π ↔ ; X1Σ+ transition, following the single photon 193 nm photolysis of CS2. Multiple photon 193 nm generation of CS fragment emission from 620 to 170 nm is also reported. A partial assignment of the emission spectrum identifies fluorescence from the CS A′1Σ+ and A1Π states.  相似文献   

16.
Abstract— The order of inhibition of the photooxidation of chlorophyll a in ethanol and ethanol-benzene is as follows: β-carotene, α-tocopherol, benzoquinone, DABCO, menadione, cholesterol and KI. The quenching of singlet oxygen by β-carotene occurs by a collisional quenching mechanism with a diffusion-controlled rate of 1.7 × 1010 M -1 s-1. Photodecomposition of Chi a is faster in ethanol-D2O than in ethanol-H2O. Photoirradiation (660 nm) of the peridinin-Chl a -protein complex, a photosynthetic light-harvesting pigment isolated from marine dinoflagellates, did not show any photo-decomposition of its Chi a in H2O or D2O, even after an extended period (12 h) of irradiation. However, the carotenoid, peridinin, in the photosynthetic antenna pigment was photobleached (ca. 10%) during the irradiation. We conclude that the singlet oxygen formed as a result of the Chi photosensitization is immediately quenched by the low-lying triplet state of four peridinin molecules (per Chl a ) bound within the same protein crevice. The carotenoid thus effectively protects Chl a from photodynamic damage, providing a direct proof for the protective role of carotenoids in the photosynthetic pigment complex.  相似文献   

17.
Ya-Huei Gan  Jian-Cheng Lee  Fung-E. Hong   《Polyhedron》2006,25(18):3555-3561
A palladium dimer with a cobalt-containing phosphine ligand, {(μ-PPh2CH2PPh2)Co2(CO)4(μ,η-(tBu)2PCCC6H4-κC1)Pd(μ-Cl)}2 (3), was prepared from the reaction of its monomer precursor, (μ-PPh2CH2PPh2)Co2(CO)4(μ,η-(tBu)2PCCC6H4-κC1)Pd(μ-OAc) (2), with LiCl. The crystal structure of 3, determined by X-ray diffraction methods, revealed a doubly chloride-bridged palladium dimeric conformation. Suzuki coupling reactions of bromobenzene with phenylboronic acid were carried out catalytically using these two novel palladium complexes 2 and 3 as catalyst precursors. Factors such as the molar ratio of substrate/catalyst, reaction temperature, base and solvent that might affect the catalytic efficiencies were investigated. As a general rule, the performance is much better by employing 3 than 2 as the catalyst precursor.  相似文献   

18.
Treatment of [Ru2(CO)4(MeCN)6][BF4]2 or [Ru2(CO)4(μ-O2CMe)2(MeCN)2] with uni-negative 1,1-dithiolate anions via potassium dimethyldithiocarbamate, sodium diethyldithiocarbamate, potassium tert-butylthioxanthate, and ammonium O,O′-diethylthiophosphate gives both monomeric and dimeric products of cis-[Ru(CO)22-(SS))2] ((SS)=Me2NCS2 (1), Et2NCS2 (2), tBuSCS2 (3), (EtO)2PS2 (4)) and [Ru(CO)(η2-(Me2NCS2))(μ,η2-Me2NCS2)]2 (5). The lightly stabilized MeCN ligands of [Ru2(CO)4(MeCN)6][BF4]2 are replaced more readily than the bound acetate ligands of [Ru2(CO)4(μ-O2CMe)2(MeCN)2] by thiolates to produce cis-[Ru(CO)22-(SS))2] with less selectivity. Structures 1 and 5 were determined by X-ray crystallography. Although the two chelating dithiolates are cis to each other in 1, the dithiolates are trans to each other in each of the {Ru(CO)(η2-Me2NCS2)2} fragment of 5. The dimeric product 5 can be prepared alternatively from the decarbonylation reaction of 1 with a suitable amount of Me3NO in MeCN. However, the dimer [Ru(CO)(η2-Et2NCS2)(μ,η2-Et2NCS2)]2 (6), prepared from the reaction of 2 with Me3NO, has a structure different from 5. The spectral data of 6 probably indicate that the two chelating dithiolates are cis to each other in one {Ru(CO)(η2-Et2NCS2)2}fragment but trans in the other. Both 5 and 6 react readily at ambient temperature with benzyl isocyanide to yield cis-[Ru(CO)(CNCH2Ph)(η2-(SS))2] ((SS)=Me2NCS2 (7) and Et2NCS2 (8)). A dimerization pathway for cis-[Ru(CO)22-(SS))2] via decabonylation and isomerization is proposed.  相似文献   

19.
This work aims to identify of non-reversible structural changes induced in β-lactoglobulin by permeation through porous ultrafiltration membranes. The evaluation of these structural changes is performed using a fluorescence methodology, which combines the use of three different, complementary, fluorescence techniques: steady-state fluorescence, picosecond time-resolved fluorescence and steady-state fluorescence anisotropy. The identification of the nature of the structural changes induced upon permeation is possible through comparison of the fluorescence responses obtained for β-lactoglobulin solutions collected after permeation (permeates and retentates) with those induced by chemical (addition of Guanidine hydrochloride, GndHCl) and thermal denaturation of β-lactoglobulin.

The fluorescence approach used allowed to identify irreversible losses of structural integrity of β-lactoglobulin in the permeates, while β-lactoglobulin retentates seemed to be unaffected by the ultrafiltration process.

The mechanisms that regulate the structural alterations of β-lactoglobulin and the magnitude of these alterations depend on the protein to membrane pore size ratio, λ, being more substantial at higher λ (severe pore constriction). Under these conditions (permeation with a 10 kDa membrane) the structural changes induced in the proteins are dictated by the high shear stress at the membrane pore walls. The increase of the membrane cut-off (30 kDa membrane) induces a decrease in the magnitude of the shear stress and the effect of protein–membrane chemical interactions becomes noticeable.  相似文献   


20.
Norfloxacin, 1-ethyl-6-fluoro-1,4-dihydro-4-oxo-7-(1-piperazinyl)-3-quinoline carboxylic acid (NORH), reacts with aluminium(III) ion forming the strongly fluorescent complex [Al(HNOR)]3+, in slightly acidic medium. The complex shows maximum emission at 440 nm with excitation at 320 nm. The fluorescence intensity is enhanced upon addition of 0.5% sodium dodecylsulphate. Fluorescence properties of the Al-NOR complex were used for the direct determination of trace amounts of NOR in serum. The linear dependence of fluorescence intensity on NOR concentration, at a NOR to Al concentration ratio of 1:10, was found in the concentration range 0.001–2 μg/ml NOR with a detection limit of 0.1 ng/ml. The ability of aluminium (III) ion to form complexes with NOR was investigated by titrations in 0.1 M LiCl medium, using a glass electrode, at 298 K, in the concentration range: 2 × 10−4 ≤ [Al] ≤ 8 × 10−4; 5 × 10−4 ≤ [NOR] ≤ 9 × 10−4 mol/dm3; 2.8 ≤ pH ≤ 8.3. The experimental data were explained by the following complexes and their respective stability constants, log(β ± σ): [Al(HNOR)], (14.60 ± 0.05); [Al(NOR)], (8.83 ± 0.08); [A1(OH)3(NOR)], (−14.9 ± 0.1), as well as several pure hydrolytic complexes of A13+. The structure of the [Al(HNOR)] complex is discussed, with respect to its fluorescence properties.  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