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.
黄池宝  任安祥 《化学学报》2007,65(23):2765-2770
报道了具有典型D-A-D型共轭结构的反式2,5-二氰基-1,4-二(4'-甲氧基苯乙烯基)苯(MOS-CN), 2,5-二氰基-1,4-二(4'-二甲胺基苯乙烯基)苯(MAS-CN)和1,4-二(4'-甲氧基苯乙烯基)苯(MOS)的合成. 用核磁、红外和元素分析进行了表征. 测试了紫外吸收光谱、单光子荧光光谱、双光子荧光光谱、双光子吸收系数及双光子吸收截面. 在800 nm的飞秒脉冲激光激发下, 化合物MOS-CN, MAS-CN和MOS分别发出很强的绿色、黄色和蓝色上转换荧光. 化合物MOS-CN, MAS-CN和MOS的最大吸收波长、单光子发射波长、双光子诱导荧光波长、荧光量子产率、双光子吸收系数、双光子吸收截面及双光子荧光寿命各分别是393, 473, 367 nm; 470, 569, 434 nm; 475, 574, 438 nm; 0.12, 0.72, 0.21; 0.8, 5.3, 0.3 cm/GW; 270, 1790, 101 GM; 140 ps, 1.32 ns, 54 ps. MAS-CN的双光子吸收截面是MOS-CN的6.63倍, MOS-CN的双光子吸收截面是MOS的2.67倍, 表明对位氨基显著地提高了化合物的双光子吸收性能, 氰基也较大地提高了双光子吸收截面.  相似文献   

2.
Abstract— The excitation energy transfer processes in the allophycocyanin (APC) monomer and trimer from phycobilisomes of Polysiphonia urceolata were studied using picosecond time-resolved fluorescence isotropic and anisotropic spectroscopy. Based on our experimental results, conclusions could be drawn as follows: (1) After the processes of exciton localization are finished, the localized excitation energy on any chromophore can be transferred to the other chromophores due to the weak couplings between them, and the processes among three p84-phycocyaninbilin (PCB) chromophores in the center of the ring shape of the APC trimer are more important than those of between a84- and p84-PCB chromophores in the same monomer. (2) The decay time constants of 95 ± 5 ps and 40 ± 5 ps components, observed by us in this work, were assigned to the excitation energy transfer or redistribution between α84- and β84-PCB chromophores in the same monomer of the APC trimer and among three β84-PCB chromophores in the center of the ring shape of the APC trimer, respectively. Specifically, the assignment of the decay constants for the 40 ± 5 ps component was different from those of previous results. (3) Based on the model of Debreczeny, and using the fluorescence residual anisotropy r(∞) with a probing wavelength of 650 nm, the angles between the C3 symmetry axis and transition dipoles of α84- and -PCB chromophores were found to be φa84= 67° and φβ84= 148°, respectively, which are in agreement with the prediction of the X-ray crystal structure of APC. (4) The results show that anisotropy decays, observed with the APC trimer, did exhibit a strongly probing wavelength dependence that did not show up in the monomer.  相似文献   

3.
Abstract— Quantum yield and fluorescence polarization determinations on phycobilisomes and their constituent phycobiliproteins show that phycobilisomes are energetically effective macromolecular structures. Energy migration within the phycobilisome to allophycocyanin, the longest wavelength absorbing and emitting phycobiliprotein, was indicated by the predominant allophycocyanin fluorescence emission which was independent of the phycobiliprotein being excited. The high efficiency of the energy migration inside the phycobilisome was reflected by the low polarized fluorescence. Excitation of phycobilisomes in the region of major absorption (500–650 nm) resulted in degrees of fluorescence polarization between +0.02 and –0.02, whereas in isolated phycobiliproteins the values were 2 to 12 times greater. Furthermore, 94–98° of the excitation energy of phycoerythrin was transferred to phycocyanin and allophycocyanin as determined from comparisons of fluorescence spectra of intact and dissociated phycobilisomes. The fluorescence quantum yields of phycobilisomes were about 0.60–0.68, very similar to that of pure allophycocyanin in solution (0.68). Phycobilisomes isolated from Fremyella diplosiphon and Nostoc sp. (blue-gree algae) have respective quantum yields of 0.68 and 0. 65, and those isolated from Porphyridium cruentum (red alga), about 0.60. In Fremyella diplosiphon and Nostoc sp., which showed a striking adaptation to different wavelengths, the phycobilisome quantum yields only varied from 0.68 to 0.67 and from 0.65 to 0. 60, respectively. The mean transfer time, calculated on the basis of experimental results, was about 280 ± 40 ps for transfer of excitation from the phycoerythrin to the phycocyanin layer in phycobilisomes. This time corresponds to the mean number of jumps, about 28, of the excitation in the phycoerythrin layer before it is captured by phycocyanin. These values are in reasonable agreement with the values of 250 ± 30 ps and 25 jumps, calculated on the basis of a phycobilisome model (of Porphyridium cruentum) and Pearlstein's theory of energy migration devised for a three-dimensional photosynthetic unit. It was also shown that Paillotin's theory of energy migration predicts similar values for mean transfer time and mean number of jumps, if one assumes that phycocyanin is a perfect sink for phycoerythrin excitation.  相似文献   

4.
Phycobiliproteins are derived from the photosynthetic apparatus of cyanobacteria and eukaryotic algae. They are composed of a protein backbone to which linear tetrapyrrole chromophores are covalently bound. Furthermore, they are water-soluble highly fluorescent, and relatively stable at room temperature and neutral pH. For this reason, capillary electrophoresis-laser induced fluorescence (CE-LIF) seems the idea method for determination of these important proteins. The effects of buffer additives such as sodium dodecyl sulfate (SDS)and putrescine on the separation of the three major phycobiliprotein types, namely allophycocyanin, phycocyanin, and phycoerythrin, with excitation and emission maxima at 652/660, 615/647, and 565(494)/575 nm, respectively, are considered. Detection limits for these proteins by CE-LIF are some 60-500 times better than by absorbance detection. The development of a fast and sensitive CE-LIF assay such as this is of potential significance to our understand ing of chemical and biological oceanographic processes.  相似文献   

5.
N-ethylpyrrole is one of ethyl-substituted derivatives of pyrrole and its excited-state decay dynamics has never been explored. In this work, we investigate ultrafast decay dynamics of N-ethylpyrrole excited to the S1 electronic state using a femtosecond time-resolved photoelectron imaging method. Two pump wavelengths of 241.9 and 237.7 nm are employed. At 241.9 nm, three time constants, 5.0±0.7 ps, 66.4±15.6 ps and 1.3±0.1 ns, are derived. For 237.7 nm, two time constants of 2.1±0.1 ps and 13.1±1.2 ps are derived. We assign all these time constants to be associated with different vibrational states in the S1 state. The possible decay mechanisms of different S1 vibrational states are briefly discussed.  相似文献   

6.
A series of {(9,9‐dioctylfluorene)0.7?x‐(dibenzothiophene‐S,S‐dioxide)0.3‐[4,7‐bis(2‐thienyl)‐2,1,3‐benzothiadiazole]x} (PFS30‐TBTx), where x represents the minor percentage of the red emitter 4,7‐bis(2‐thienyl)‐2,1,3‐benzothiadiazole (TBT) randomly incorporated into the copolymer backbone, is investigated in order to follow the energy transfer from PFS30 to TBT moieties. The emission of the donor poly[(9,9‐dioctylfluorene)0.7‐(dibenzothiophene‐S,S‐dioxide)0.3 identified by PFS30 and peaking at 450 nm, is clearly quenched by the presence of the red TBT chromophore emitting at 612 nm, with an isoemissive point observed when the spectra are collected as a function of temperature. A plot of the ratio between the TBT and PFS30 emissions as a function of the reciprocal of temperature gives a clear linear trend between 290 and 200 K, with an activation energy of 20 meV and showing a turn over to a non‐activated regime below 200 K. Picosecond time‐resolved fluorescence decays collected at the PFS30 and TBT emission wavelengths, show a decay of the PFS30 emission and a fast build‐in, followed by a decay, of the TBT emission, confirming that the population of the TBT excited state occurs during the PFS30 lifetime(~600 ps). The population of the TBT excited state occurs on a time regime around 150 ps at 290 K, showing an energy barrier of 20 meV that turns over to a non‐activated regime below 200 K in clear agreement with the steady‐state data. The origin of the activation barrier is attributed to the presence of physical and energetic disorder, affected by fast thermal fluctuations that dynamically change the energy landscape and control the exciton migration through the polymer density of states.  相似文献   

7.
Abstract. Fluorescence techniques of various types have been used to study the light-gathering and energy transfer modes for various cryptomonad biliproteins (phycocyanin or phycoerythrins). Analysis of fluorescence polarization and absorption data demonstrates that each cryptomonad biliprotein is composed of at least two distinct types of absorbing chromophore, each attached to the protein through covalent linkages to different polypeptide chains. Examination of the fluorescence emission spectra as a function of excitation at several wavelengths demonstrates that only one of these absorbing chromo-phores is responsible for the fluorescence. This behavior is consistent with a known phenomenon whereby photons are gathered by more than one chromophore and then after radiationless energy transfer are emitted by only one chromophore. Application of Förster dipole-dipole energy transfer theory is made to the study of the mode by which energy absorbed by biliproteins migrates to Chl a. The spectral overlap integral between phycocyanin (Chroomonas sp.) and Chl a is 7.13 ± 10-10cm6mol-1and between phycocyanin and Chl c2 0.25 ± 10-10cm6mol-1. This large difference in overlap suggests, although does not prove, that phycocyanin might transfer energy directly to Chl a without a Chl c2 intermediary. The cryptomonad phycoerythrins may also use this method but a Chl c2 intermediate could not be ruled out for them. Radiationless energy transfer among homogeneous biliproteins is shown to be feasible. All these calculations are based on in vitro spectra and the interpretations extrapolated to the cellular situation, and these tentative conclusions are reached without knowledge of other factors, such as chromophore-chro-mophore orientation and distance, which could greatly influence the energy transfer scheme.  相似文献   

8.
The synthesis and photophysical data of new carbostyrils (quinoline‐2(1H)‐ones) with the longest hitherto observed absorption‐ and emission wavelengths are described. Introduction of 6‐amino, 7‐MeO, and 4‐(CF3) substituents enabled us to rise the absorption and fluorescence maxima up to 414 and 557 nm, respectively. Supported by semi‐empirical and ab initio calculations, the 6,7‐(1,4‐diazine)‐fused carbostyril 23b displayed absorption maxima at up to 440 nm, with quantum yields of up to 0.9 and large Stokes shifts (>100 nm), comparable to the best coumarin chromophores known. The new fluorophore is neither pH‐sensitive between pH 6 and 10 nor susceptible to O2 quenching. At pH 3, the emitted light appears greenish‐white, which arises from three different stages of protonation.  相似文献   

9.
Abstract— 1. R-II is a u.v. induced mutant of cyanidium caldarium which lacks the major accessory pigment phycocyanin. 2. Loss of phycocyanin does not impair photosynthesis or the Hill reaction. 3. The action spectrum for the quantum yield indicates an efficiently operating photosynthetic mechanism which is shifted in peak efficiency toward longer wavelengths when compared to the wild-type spectrum. The red drop is also shifted toward the far red in R-II. 4. Although 30 per cent enhancement was obtained in wild-type Cyanidium, no enhancement was observed when phycocyanin was absent. 5. A pigment absorbing at 720 nm has been observed and may possibly represent a chlorophyll-phycocyanin complex.  相似文献   

10.
Abstract— Resonance-enhanced coherent anti-Stokes Raman (CARS) spectra are reported for trimers of phycoerthrocyanin (PEC) dissolved in H2O and D2O. The CARS spectra are significantly different when recorded with pump wavelength either at 585 nm or 630 nm. By comparison of the 630 nm spectra with those of phycocyanin one can conclude that there is a change in the relative location of the lowest excited states of the β84 and β155 chromophores. Upon additional illumination with 514.5 nm laser light, only the spectra recorded with 585 nm pump wavelength change. This is in accordance with earlier observations that the phycoviolobilin chromophore of the a-subunit exhibits photochromic behavior. The changes in the CARS spectra provide evidence that it is the methine bridge between rings C and D that undergoes the geometrical changes. Furthermore, it is suggested that there are different types of photoinduced rearrangements operative and that the isomeric distribution is different in H2O and D2O.  相似文献   

11.
Picosecond time-resolved fluorescence of photosystem I particles isolated from Synechococcus sp. was recorded in the wavelength range from 680 nm to 736 nm for temperatures of 6°C to 42°C and - 100°C using the single-photon-timing technique. By global analysis of the data we found four contributing lifetime components at the higher temperatures (T1 ' 12 ps, T1= 35 ps, T3 ' 65 ps, T4 ' 1000 ps). We attribute T1 to an energy transfer between two pigment pools, T2 to the charge separation process in the reaction center, component T3 is assigned to aggregate and T4 to uncoupled chlorophyll emission. The corresponding decay-associated spectra are presented. We also applied a target analysis procedure to fit parameters of a kinetic model directly to the data. The resulting rate constants and species-associated spectra are discussed. The data indicate substantial spectral heterogeneity in the antenna with at least three substantially different chlorophyll pools. The overall exciton decay kinetics (by charge separation) is trap-limited.  相似文献   

12.
Visible light (405–615 nm) excitation of carboquone, mitomycin C, and streptonigrin dissolved in dimethylsulfoxide in the presence of oxygen generates superoxide anion radicals (O2?). The quantum yields for these reactions range from 4.2 times 10?2 (carboquone, λ= 615 ± 10 nm) to 7.3 times 10?6 (streptonigrin, λ=545 ± 10 nm). O2? radicals were spin trapped with 5,5-dimethyl-1-pyrroline-1-oxide (DMPO) and identified by electron spin resonance (ESR). The efficiency of DMPO to spin trap O2? in dimethylsulfoxide was determined and indicated that 91% of the O2? present in dimethylsulfoxide is trapped by DMPO. The oxidation of the photoexcited drug molecules occurs via a direct electron transfer to dissolved oxygen in solution. Ultraviolet irradiation (λ= 313 ± 10 nm) of the aminoquinone drug solutions (80% H2O, 20% dimethylsulfoxide) in the presence of peptides results in the decarboxylation of the peptides. In this case the photoexcited drugs are reduced, abstracting an electron from the C-terminal carboxyl group of the peptides. The reaction is specific to the C-terminal amino acid of the peptide. The decarboxylated peptide radicals were spin trapped with 2-methyl-2-nitrosopropane (MNP) and identified by ESR.  相似文献   

13.
Side-chain second-order nonlinear optical polyimides were prepared from four novel chromophore-containing diamines and 4,4′-(hexafluoroisopropylidene) diphthalic anhydride by a traditional two-step process that included a solution polycondensation followed by a chemical imidization. The four diamines were 2,4-di-β-aminoethylamino-6-p-nitrophenylamino-1,3,5-triazine (M1), 4-nitro-4′-[N-(4,6-di-β-aminoethylamino)-1,3,5-triazin-2-yl]amino azobenzene (M2), 2,4-di-p-aminophenylamino-6-p-nitrophenylamino-1,3,5-triazine (M3), and 4-nitro-4′-[N-(4,6- di-4-aminophenylamino)-1,3,5-triazin-2-yl]amino azobenzene (M4). All the polyimides exhibited maximum ultraviolet-visible absorption peaks or shoulders of chromophores at wavelengths below 400 nm, and those based on M1 and M3 were transparent at wavelengths above 450 nm, whereas those based on M2 and M4 were transparent at wavelengths above 550 nm. The polyimides possessed high glass-transition temperatures (Tg's; 218–247 °C) and thermal decomposition temperatures. They were soluble in aprotic solvents such as N-methyl-2-pyrrolidone, N,N-dimethyl acetamide, N,N-dimethyl formamide, and dimethylsulfone. Some were even soluble in common low-boiling-point solvents such as tetrahydrofuran. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 4330–4336, 1999  相似文献   

14.
Treatment with a mixture of formic and hydrobromic acids converted ethyl 3,4-diethyl-5-methyl-pyrrole-2-carboxylate 7a to 3,3′,4,4′-tetraethyl-5,5′-dimethylpyrromethene hydrobromide 8a presumably via the condensation of α-unsubstituted and α-formylpyrrole intermediates 7c and 7e formed in situ. The corresponding 6-cyanohexaalkylpyrromethane 9a was obtained by the addition of hydrogen cyanide to the pyrromethene 8a and was oxidized with bromine to an unstable pyrromethene 10a , an intermediate converted to 1,2,6,7-tetraethyl-3,5-dimethyl-8-cyanopyrromethene–BF2 complex 3 , (PM TEDC), λlas (plastic) 613–639 nm, in a reaction with boron trifluoride etherate. Ethyl 3,4,5-trimethylpyrrole-2-carboxylate 7b was similarly converted to 1,2,3,5,6,7-hexamethyl-8-cyanopyrromethene–BF2 complex 4 , (PM HMC), λlas (plastic) 615–639 nm. Immediately after formation by a condensation between propionyl chloride and 2,4-dimethyl-3-cyanopyrrole 16 , unstable 3,3′,5,5′-tetramethyl-6-ethyl-4,4′-dicyanopyrromethene hydrochloride 17 was treated with boron trifluoride etherate to give 1,3,5,7-tetramethyl -2,6-dicyano-8-ethylpyrromethene–BF2 complex 18 , λlas (ethanol) 540–565 nm.  相似文献   

15.
Abstract— Action spectra for lethality of E. coli WP2s under aerobic and anaerobic conditions. based on final slopes of the survival curves, reveal the absence of oxygen dependence at 313 nm and shorter wavelengths and a strong oxygen dependence (OER of 12 at 334 nm and 16 at 365 nm) at wavelengths longer than 313 nm. Shoulders or small peaks at 340, 365 , 410 and 500 nm suggest the participation of non-DNA chromophores in aerobic lethality at these wavelength ranges.  相似文献   

16.
Molecular dynamics simulation of the relaxation at 300 K of a fully extended polyethylene chain of800 CH_2 units has been carried out by following the changes in morphology, van der Waals energy, radius ofgyration in the sense of mechanics and gyration radius in the sense of Flory, population of trans-conformation and orientation factor. The relaxation went through three stages: (1) relaxation from themorphology of a straight rod of 100 nm length to the molphology close to a random coil of gyration radius5.9 nm in 110 ps; (2) collapse of the morphology of a coil to a highly compact globule close to a sphere ofgyration radius 1.3 nm after 178 ps as the result of intersegmental van der Waals attractive interactions; (3)lateral ordering of the folded chain segments in the globule without appreciable changes in the chaindimension up to 1600 ps, the time limit of present simulation. Nearly complete relaxation of local segmentalorientation was performed much faster than the relaxation of globule chain orientation even for a single chainof low degree of polymerization and at a temperature some 155℃above its T_g. The lateral ordering of thechain segments during the period 178 to 680 ps of the simulation time was found to obey the Avramiequation with an Avrami index of 1 .44.  相似文献   

17.
Intramolecular F?rster-type excitation energy transfer (FRET) processes in a series of first-generation polyphenylene dendrimers substituted with spatially well-separated peryleneimide chromophores and a terryleneimide energy-trapping chromophore at the rim were investigated by steady-state and time-resolved fluorescence spectroscopy. Energy-hopping processes among the peryleneimide chromophores are revealed by anisotropy decay times of 50--80 ps consistent with a FRET rate constant of k(hopp) = 4.6 ns(-1). If a terryleneimide chromophore is present at the rim of the dendrimer together with three peryleneimide chromophores, more than 95% of the energy harvested by the peryleneimide chromophores is transferred and trapped in the terryleneimide. The two decay times (tau(1) = 52 ps and tau(2) = 175 ps) found for the peryleneimide emission band are recovered as rise times at the terryleneimide emission band proving that the energy trapping of peryleneimide excitation energy by the terryleneimide acceptor occurs via two different, efficient pathways. Molecular- modeling-based structures tentatively indicate that the rotation of the terryleneimide acceptor group can lead to a much smaller distance to a single donor chromophore, which could explain the occurrence of two energy-trapping rate constants. All energy-transfer processes are quantitatively describable with F?rster energy transfer theory, and the influence of the dipole orientation factor in the F?rster equation is discussed.  相似文献   

18.
Fluorescence resonance energy transfer (FRET) from Coumarin 153 (C153) to Rhodamine 6G (R6G) in a secondary aggregate of a bile salt (sodium deoxycholate, NaDC) is studied by femtosecond up-conversion. The emission spectrum of C153 in NaDC is analysed in terms of two spectra-one with emission maximum at 480 nm which corresponds to a non-polar and hydrophobic site and another with maximum at ∼530 nm which arises from a polar hydrophilic site. The time constants of FRET were obtained from the rise time of the emission of the acceptor (R6G). In the NaDC aggregate, FRET occurs in multiple time scales — 4 ps and 3700 ps. The 4 ps component is assigned to FRET from a donor (D) to an acceptor (A) held at a close distance (R DA ∼ 17 ?) inside the bile salt aggregate. The 3700 ps component corresponds to a donor-acceptor distance ∼48 ?. The long (3700 ps) component may involve diffusion of the donor. With increase in the excitation wavelength (λ ex) from 375 to 435 nm, the relative contribution of the ultrafast component of FRET (∼4 ps) increases from 3 to 40% with a concomitant decrease in the contribution of the ultraslow component (∼3700 ps) from 97 to 60%. The λ ex dependence is attributed to the presence of donors at different locations. At a long λ ex (435 nm) donors in the highly polar peripheral region are excited. A short λ ex (375 nm) ‘selects’ donor at a hydrophobic location.  相似文献   

19.
Transient UV/Vis absorption spectroscopy is used to study the primary dynamics of the ring‐A methyl imino ether of phycocyanobilin (PCB‐AIE), which was shown to mimic the far‐red absorbance of the Pfr chromophore in phytochromes (R. Micura, K. Grubmayr, Bioorg. Med. Chem. Lett.­ 1994, 4, 2517–2522 ). After excitation at 615 nm, the excited electronic state is found to decay with τ1=0.4 ps followed by electronic ground‐state relaxation with τ2=1.2 and τ3=6.7 ps. Compared with phycocyanobilin (PCB), the initial kinetics of PCB‐AIE is much faster. Thus, the lactim structure of PCB‐AIE seems to be a suitable model that could not only explain the bathochromic shift in the ground‐state absorption but also the short reaction of the Pfr as compared to the Pr chromophore in phytochrome. In addition, the equivalence of ring‐A and ring‐D lactim tautomers with respect to a red‐shifted absorbance relative to the lactam tautomers is demonstrated by semiempirical calculations.  相似文献   

20.
Nanostructured TiO2 crystals were synthesized by gel to crystalline conversion. The crystals obtained were anatase form of titania averaging in 30 nm particles with an intrinsic band gap of 3.1 eV. The photocatalytic behavior was evaluated for the bactericidal effect in water, contaminated with the indicator organism Escherichia coli. The 100% photoinactivation of E. coli was achieved within 60 min with suspended nano-TiO2. The catalytic activity of synthesized nanosample was observed to be 2.6 times more than that of commercial TiO2 sample referred to as Degussa P-25. The photoinactivation of E. coli was tested with irradiation source of different wavelengths to substantiate the influence of particle size and nano crystallinity on electronic band structure. The photoactivity of nano titania enhanced to 1.625 times when the source of irradiation shifted from 360 to 400 nm while Degussa P-25 showed no change.  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