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 765 毫秒
1.
Singlet oxygen generation is reported from (1) enzymatic reaction and (2) electron transfer reactions of the superoxide anion measured directly with an ultrasensitive near-IR emission spectrophotometer by monitoring the O2(1Δg) → O2 (3Σg?) transition at 1268 nm. Near-IR emission spectra from the myeloperoxidase and lactoperoxidase enzymatic systems show only emission of singlet oxygen at 1268nm. The lipoxygenase/Na–linoleate enzymatic reaction exhibits two emissions, 1268 nm and 1288 nm. The latter emission is identified as originating from a peroxy radical. Spectral and kinetic data giving evidence of singlet oxygen generation is obtained from the reaction of potassium superoxide solubilized by 18-crown-6-ether in acetonitrile with a series of organometallic coordination compounds.  相似文献   

2.
The kinetics of the decomposition reaction of 4‐(4,5‐diphenyl‐1H‐imidazol‐2‐yl)phenyl acetate ( 1 ) in basic alcoholic media was investigated, using a simple fluorescence (FL) spectrophotometric procedure. The process was conveniently studied using FL, since the triphenylimidazole‐derived ester 1 and its reaction products (the corresponding phenol 2 and phenolate 2 ? ) are all highly fluorescent (ΦFL > 37%). By carefully selecting excitation and emission wavelengths, observed rate constants k1 in the order of 10?3 to 10?2 s?1 were obtained from either reactant consumption (λex = 300 nm, λem = 400 nm) or product formation (λex = 350 nm, λem = 475 nm); these were shown to be kinetically equivalent. Intensity‐decay time profiles also gave a residual FL intensity parameter, shown to be associated to the distribution of produced species 2 and 2 ? , according to the basicity of the medium. Studying the reaction in both methanol (MeOH) and isopropanol (iPrOH), upon addition of HO?, provided evidence that the solvent's conjugate base is the active nucleophilic species. When different bases were used (tBuO?, HO?, DBU and TEA), bimolecular rate constants kbim ranging from 4.5 to 6.5 L mol?1 s?1 were obtained, which proved to be non‐dependent on the base pKaH, suggesting specific base catalysis for the decomposition of 1 in alcoholic media.  相似文献   

3.
Spectra of ultraweak chemiluminescence (CL) accompanying auto-oxidation and hydration of cereal products have been measured using single photon counting and cut-off filters. The spectra cover the 380–880 nm spectral range with maxima centred around 600 nm. Analytically pure air-dried carbohydrates like agar, cellulose and nitrocellulose give emission too weak for spectral measurements. The emission from water pure carbohydrates is on average 4–12 times higher and emission spectra are similar to those from cereal products. The effect of free radical scavengers, SOD and O*2 (1Δg)-quenchers on CL spectra indicates a contribution of radical reactions with the participation of excited carbonyls, O2 and excited molecular oxygen dimoles. Moreover, possible mechanisms of chemi-excitation due to a cooperative H-bond formation during the hydration of carbohydrates and/or recombination of trapped radicals and electron-holes are discussed. It is also postulated that the excitation energy transfer to natural sensitizers occuring in cereal products may account for non-specific broad spectra and differences in the intensity of CL. © 1998 John Wiley & Sons, Ltd.  相似文献   

4.
A Circular Dichroic absorption study of the reaction of oxidized pyridine nucleotides with cyanide ions fully confirms the occurence of a very weak Cotton effect around 435 nm in the Circular Dichroic spectrum of the reduced coenzymes and therefore the very faint transition (λmax = 435 nm; ?max ~ 1 M?1 cm?1) from which the Cotton effect originates.  相似文献   

5.
Resonance Raman spectra of protocatechuate 3,4-dioxygenase from Pseudomonas aeruginosa have been investigated during the reaction of the enzyme with substrate and oxygen. It is found that the spectrum of the turned-over enzyme is indistinguishable from that of the resting enzyme in the absence of substrate, and is characterized by resonance-enhanced tyrosinate ring vibrational modes at 1263 and 1174 cm?1. In the ternary ESO2 complex, however, the tyrosinate vibrational modes are shifted to 1252 and 1165 cm?1, respectively. There is no evidence for any dioxygen vibrations in the spectra of ESO2 complexes prepared with 16O2, 18O2, and 16O18O in the region between 1300 and 200 cm?1. The results of this resonance Raman study are interpreted to indicate that molecular oxygen is attached only to the substrate (but not iron) in the stable intermediate, and that the concomitant rearrangement at C4 of the substrate induces a substantial change in geometry of the tyrosine residues associated with the iron complex. Furthermore, the optical spectrum of the ESO2 complex (λmax = 520 nm) is dominated by tyrosinate → Fe(III) charge transfer and contains little or no peroxide → Fe(III) charge transfer. These results invalidate the previously advanced analogy in spectral properties between this enzyme and the respiratory protein, oxyhemerythrin.  相似文献   

6.
Chemiluminescence (CL) of 7-hydroxycoumarin, umbelliferone, a strongly fluorescing compound, the precursor of photoactive furocoumarins widely spread in plant kingdom and used in dye lasers has been examined in four oxidative systems producing active oxygen species: A–horseradish peroxidase + H2O2 + buffers pH = 3–10; B? K3Fe (CN)6 + H2O2 + buffers pH = 4–12; C? NaOCI + buffer pH = 3–10.8; and D? HCHO + H2O2 + K2CO3 (Na2CO3), the so-called Trautz–Schorigin reaction. In all these systems a 10–1000-fold increase of CL intensity and quantum yield and an appearance of the blue emission band (λmax = 460–480 nm), typical of the fluorescence of hydroxycoumarin used have been observed. Spectrophotometric and fluorimetric measurements of the systems A–D indicate that the tested hydroxycoumarins - coumarin, umbelliferone and 6-glucoside umbelliferone or esculin – undergo slow (systems A and B) and fast (systems C and D) consumption, leading to the degradation of coumarins. These findings suggest that efficiently fluorescing hydroxy-coumarins may act both as sensitizers – secondary emitters of CL and/or reactants undergoing oxidative degradation.  相似文献   

7.
(1) Aqueous solutions of 1–10 μM ferricytochrome c treated with 100 μM–100 mM H2O2 at pH 8.0 emit chemiluminescence with quantum yield Ф ? 10?9 and absolute maximum intensity Imax ? 105 hv/s per cm3 (λ = 440), and exhibit exponential decay with a rate constant of 0.15 s?1. (2) The emission spectrum of the chemiluminescence covers the range 380–620 nm with the maximum at 460 ± 10 nm. (3) Neither cytochrome c nor haemin fluoresce in the spectral region of the chemiluminescence. In the reaction course with H2O2, a weak fluorescence in the region 400–620 nm with λmax = 465–510 nm (λexc 315–430 nm) gradually arises. This originates from tryptophan oxidation products of the formylkynurenine type or from imidazole derivatives, respectively. (4) Frozen solutions (77 K) of cytochrome c exhibit phosphorescence typical of tryptophan (λexc = 280 nm, λem = 450 nm). During the peroxidation, an additional phosphorescence gradually appears in the range 480–620 nm with λmax = 530 nm (λexc = 340 nm). This originates from oxidative degradation products of tryptophan. (5) There are no red bands in the chemiluminescence spectra of cytochrome c or haemin. This result suggests that singlet molecular oxygen O2(1Δg) is not involved in either peroxidation or chemiluminescence. (6) The haem Fe3+ group and H2O2 appear to be crucial for the chemiluminescence. It is suggested that the generation of electronically excited, light-emitting states is coupled to the production of conformational out-of-equilibrium states of peroxy-Fe-protoporphyrin IX compounds.  相似文献   

8.
《Inorganica chimica acta》1988,145(1):117-120
A dimeric ruthenium(II) compound in which two Ru(bpy)3 groups are linked by an amide bonding has been prepared as a model compound to study an energy transfer between Ru(bpy)3 chelates. The nature of the solution luminescence spectrum varied with concentration: the emission maximum appeared at 650 nm for dilute solutions and at 670 nm for concentrated solutions. This concentration dependence has been interpreted in terms of excimers that are formed due to an energy transfer between two Ru(bpy)3 groups in a dimer molecule. The cyclic voltammogram for the Ru3+/Ru2+ reaction is quasireversible: the reaction is governed by a sluggish electron transfer which may be due to an intradimer electronic interaction.  相似文献   

9.
A series of Ca6AlP5O20 doped with rare earths (Eu and Ce) and co‐doped (Eu, Ce and Eu,Mn) were prepared by combustion synthesis. Under Hg‐free excitation, Ca6AlP5O20:Eu exhibited Eu2+ (486 nm) emission in the blue region of the spectrum and under near Hg excitation (245 nm), Ca6AlP5O20:Ce phosphor exhibited Ce3+ emission (357 nm) in the UV range. Photoluminescence (PL) peak intensity increased in Ca6AlP5O20:Eu,Ce and Ca6AlP5O20:Eu, Mn phosphors due to co‐activators of Ce3+ and Mn2+ ions. As a result, these ions played an important role in PL emission in the present matrix. Ca6AlP5O20:Eu, Ce and Ca6AlP5O20:Eu, Mn phosphors provided energy transfer mechanisms via Ce3+ → Eu2+ and Eu2+ → Mn2+, respectively. Eu ions acted as activators and Ce ions acted as sensitizers. Ce emission energy was well matched with Eu excitation energy in the case of Ca6AlP5O20:Eu, Ce and Eu ions acted as activators and Mn ions acted as sensitizers in Ca6AlP5O20:Eu, Mn. This study included synthesis of new and efficient phosphate phosphors. The impact of doping and co‐doping on photoluminescence properties and energy transfer mechanisms were investigated and we propose a feasible interpretation. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

10.
Intermolecular interaction study of human serum albumin (HSA) with two anthraquinones i.e. danthron and quinizarin has been performed through fluorescence, UV-vis and CD spectroscopy along with docking analysis. The titration of drugs into HSA solution brought about the quenching of fluorescence emission by way of complex formation. The binding constants were found to be 1.51 × 104 L mol?1 and 1.70 × 104 L mol?1 at λexc = 280 nm while at λexc = 295 nm, the values of binding constants were 1.81 × 104 L mol?1 and 1.90 × 104 L mol?1 which hinted toward binding of both the drugs in the vicinity of subdomain IIA. Different temperature study revealed the presence of static quenching mechanism. Moreover, more effective quenching of the fluorescence emission was observed at λexc = 295 nm which also suggested that both the drug molecule bind nearer to Trp-214. Thermodynamic parameters showed that hydrophobic interaction was the major force behind the binding of drugs. The UV-vis spectroscopy testified the formation of complex in both the systems and primary quenching mechanism as static one. The changes in secondary structure and α-helicity in both the systems were observed by circular dichroism spectroscopy. Furthermore, molecular docking analysis predicted the probable binding site of drugs in subdomain IIA of HSA molecule. The types of amino acid residues surrounding the drug molecule advocated that van der Waals forces, hydrophobic forces and electrostatic forces played a vital role in the stabilization of drug-protein complex formed.  相似文献   

11.
A strongly fluorescing 7-hydroxycoumarin (umbelliferone, U) oxidized in dilute (10 μmol/L-0, 1 mol/L) aqueous solution with CIO? or CIO? + H2O2 (but not with H2O2 alone) produces a strong chemiluminescence (CL). Light emission kinetics depends on the pH of solution (4.0–10.5) and the reaction has a low activation energy Ea = 31 ± 2 kJ/mol (285–310 K). The spectrum covers the fluorescence of umbelliferone (400–550 nm, λmax 460nm). No red emission typical of 1Δg, 1Σ+g (O2)2 is observed either in the umbelliferone +CIO? or the umbelliferone +CIO? + H2O2 solution. The possible mechanism of CL and concomitant degradative oxidation of umbelliferone is discussed.  相似文献   

12.
This article focuses on the effect of monovalent cation doping on the optical properties of rare earth (RE = Eu3+, Tb3+) co-doped Ca14Zn6Al10O35 which has been synthesized by a low temperature combustion method. Crystalline phase of the Ca14Zn6Al10O35 phosphor was examined and confirmed by X-ray diffraction measurement. Under near-ultraviolet light excitation Eu3+-doped Ca14Zn6Al10O35 phosphor exhibit characterization of Eu3+ emission bands that are located at a maximum wavelength (λmax) of approximately 470 nm and other peaks centred at 593 nm and 615 nm, respectively. With Tb3+-doped Ca14Zn6Al10O35 phosphor showing a green emission band centred at 544 nm under near-ultraviolet range. Furthermore, we studied the energy transfer process in Eu3+/Tb3+pair and enhancement in photoluminescence (PL) intensity with doping different charge compensation. Here we obtained the optimum PL emission intensity of the phosphor in broad and intense visible spectral range which may be significant for the fabrication of white light emitting diodes (WLEDs).  相似文献   

13.
Simple acidification of aqueous alkaline peroxynitrite quantitatively generates singlet (1Δg) molecular oxygen, detected and quantitated spectroscopically (1270 nm). This observation provides a chemical basis for physiological cytotoxicity of ONOO? generated in the diffusion - controlled reaction of cellular NO? and O. The experiments consist of (i) chemical generation of ONOO? from NO? gas and KO2 powder in alkaline aqueous solution; (ii) absorption spectral identification of ONOO? in the near-UV with maximum at 302 nm; (iii) spectroscopic identification of 1O2 by its emission band at 1200–1340 nm with maximum at 1275 nm; and (iv) quantitation of 1O2 generated in ONOO?/H+ reaction by comparison of the chemiluminescence intensity at 1270 nm with that from H2O2/OCl? reaction that generates 1O2 with unit efficiency at alkaline pH. 1O2 was generated with unit efficiency with respect to ONOO? concentration by the ONOO?/H+ reaction.  相似文献   

14.
Alongside rare‐earth metals, Ni, Fe, Co, Cu are some of the critical materials that will be in huge demand thanks to growth in clean‐energy sector. Herein scrap stainless steel wires (SSW) from worn‐out tires are employed as a support material for catalyst integration in the hydrogen evolution reaction (HER). In addition, SSW by corrosion engineering is exercised as an in situ formed freestanding robust electrode for the oxygen evolution reaction (OER). By superficial corrosion of SSW, inherent active species are unmasked in the form of Ni/FeOOH nanocrystallites displaying efficient water oxidation by reaching 500 mA cm?2 at low overpotential (η500) of 287 mV in 1 m KOH. Similarly, cathode scrap SSW with active (alloy) coatings of MoNi4 catalyzes the HER at η‐200 = 77 mV, with a low activation energy (Ea = 16.338 kJ mol?1) and high durability of 150 h. Promisingly, when used in industrial conditions, 5 m KOH, 343 K, these electrodes demonstrate abnormal activity by yielding high anodic and cathodic current density of 1000 mA cm?2 at η = 233 mV and η = 161 mV, respectively. This work may inspire researchers to explore and reutilize high‐demand metals from scrap for addressing critical material shortfalls in clean‐energy technologies.  相似文献   

15.
W.L. Butler  M. Kitajima 《BBA》1975,396(1):72-85
A model for the photochemical apparatus of photosynthesis is presented which accounts for the fluorescence properties of Photosystem II and Photosystem I as well as energy transfer between the two photosystems. The model was tested by measuring at ?196 °C fluorescence induction curves at 690 and 730 nm in the absence and presence of 5 mM MgCl2 which presumably changes the distribution of excitation energy between the two photosystems. The equations describing the fluorescence properties involve terms for the distribution of absorbed quanta, α, being the fraction distributed to Photosystem I, and β, the fraction to Photosystem II, and a term for the rate constant for energy transfer from Photosystem II to Photosystem I,kT(II→I). The data, analyzed within the context of the model, permit a direct comparison of α andkT(II→I) in the absence (?) and presence (+) of Mg2+:α/?α+= 1.2andk/?T(II→I)k+T(II→I)= 1.9. If the criterion thatα + β = 1 is applied absolute values can be calculated: in the presence of Mg2+,a+ = 0.27 and the yield of energy transfer,φ+T(II→I) varied from 0.065 when the Photosystem II reaction centers were all open to 0.23 when they were closed. In the absence of Mg2+? = 0.32 andφT(II→I) varied from 0.12 to 0.28.The data were also analyzed assuming that two types of energy transfer could be distinguished; a transfer from the light-harvseting chlorophyll of Photosystem II to Photosystem I,kT(II→I), and a transfer from the reaction centers of Photosystem II to Photosystem I,kt(II→I). In that caseα/?α+= 1.3,k/?T(II→I)k+T(II→I)= 1.3 andk/?t(II→I)k+(tII→I)= 3.0. It was concluded, however, that both of these types of energy transfer are different manifestations of a single energy transfer process.  相似文献   

16.
Microalgal pigment composition, photosynthetic characteristics, single-cell absorption efficiency (Qa(λ)) spectra, and fluorescence-excitation (FE) spectra were determined for platelet ice and benthic communities underlying fast ice in Mc Murdo Sound, Antarctica, during austral spring 1988. Measurements of spectral irradiance (E(λ)) and photosynthetically active radiation (PAR) as well as samples for particulate absorption measurements were taken directly under the congelation ice, within the platelet layer, as profiles vertically through the water column, and at the benihic surface. Light attenuation by.sea ice, algal pigments, and particulates reduced PAR reaching the platelet ice layer to 3%(9–33 fimol photons m-2-?s-1) of surface values and narrowed its spectral distribution to a band between 400 and 580 nm. Attenuation by the water column further reduced PAR reaching the sea floor (28–m depth) to 0.05% of surface levels (< 1 μmol photons m-2 s-1), with a spectral distribution dominated by 470–580–nm wavelengths. The photoadaptive index (I) for platelet ice algae (5.9–12.6 μmol photons m-2.s-1) was similar to ambient PAR, indicating that algae had acclimated to their light environment (i.e. the algae were light-replete). Maximum Qa(λ) at the blue absorption peak (440 nm) was 0.63, and enhanced absorption was observed from 460–500 nm and was consistent with observed high cellular chlorophyll (chi) c:chl a and fucoxanthin: chl a molar ratios (0.4 and 1.2, respectively). Benthic algae were light-limited despite the maintenance of very low Ik values (4–11 μmol photons.m-2.s-1). Extremely high fucoxanthin: chi a ratios (1.6) in benthic algae produced enhanced green light absorption, resulting in a high degree of complementation between algal absorption and ambient spectral irradiance. Qa(λ) values for benthic algae were maximal (0.9) between 400 and 510 nm but remained >0.35 even at absorption minima. Strong spectral flattening, a characteristic of intense pigment packaging, was also apparent in the Qa(λ) spectra for benthic algae. FE and Qa(λ) spectra were similar in shape for platelet ice algae, indicating that the efficiency at which absorbed energy was transferred to photosystem II (PSII) was independent of wavelength. Fluorescence emission by benthic algae was greatest for the 500–560–nm excitation wavelengths, suggesting that most energy absorbed by accessory pigments was transferred to PSII. These results suggest that under ice algae employ complementary pigmentation and maximize absorption efficiency as adaptive strategies to low-light stress. Regulating the distribution of absorbed energy between PSI and PSII may be an adaptive response to the restricted spectral distribution of irradiance.  相似文献   

17.
The activation of molecular oxygen by alkaline hemin (ferriprotoporphyrin IX) has been studied. In the presence of reductant nicotineamide adenine dinucleotide (NADH) or nicotineamide adenine dinucleotide phosphate (NADPH) and organic substrate, aniline, hemin activates oxygen to the hydroperoxide anion (HO2?) and subsequently mediates insertion of active oxygen into the benzene ring of the substrate to form p-aminophenol, with a high degree of regiospecificity. Oxygen activation does not occur in the absence of aniline. Stoichiometry of the reaction indicates that two electrons are required per molecule of oxygen activated or atom of oxygen inserted into the substrate aromatic ring system. Direct measurements of H2O2 and of the pKa for maximum rate of p-aminophenol formation (11.7 ± 0.1) indicate participation of the hydroperoxide anion as the active oxygen species in the rate-determining step of the insertion reaction. Powerful scavengers of the hydroxyl radical (OH′) have little effect on the formation of H2O2 or p-aminophenol by the system. Superoxide dismutase (10?7 mol dm?3) inhibited both p-aminophenol and H2O2 formation, when added to the system immediately prior to initiation of the reaction. Studies involving N-phenylhydroxylamine indicate that aromatic ring hydroxylation is occurring directly and not by rearrangement of an N-hydroxylated intermediate. Implications of hemin-mediated hydroxylation reactions for those of enzymatic mixed function oxidase activity are discussed.  相似文献   

18.
The electronic spectrum of CuII(dps)2 in CH3CN with dps=3,5-diisopropylsalicylate shows a ligand field absorption at λmax=711 nm (ε=140 M−1 cm−1), and a phenolate to Cu(II) ligand-to-metal charge transfer (LMCT) band at λmax=428 nm (ε=950). LMCT excitation of CuII(dps)2 leads to the reduction of Cu(II) to Cu(I). Copper(II) disappears with φ=2.8×10−3 at λirr=436 nm.  相似文献   

19.
Procedures and conditions have been established such that the unstable enzyme-bound flavin intermediate produced in the bacterial luciferase reaction can be isolated as approximately 70% of the flavin product, the remaining being the final product, FMN. The structure of the intermediate is proposed to be that of a luciferase-bound 4a,5-dihydroflavin-4a-hydroxide. The intermediate has a half-life of 33 min at 2°C and decays spontaneously to give H2O and luciferase-bound FMN with an activation enthalpy of about 120 kJ/mol. It has an absorption spectrum (λmax = 360 nm) that is consistent with the proposed structure, and a fluorescence emission (λmax = 485 nm) that matches the bioluminescence emission closely.  相似文献   

20.
It has been shown that direct excitation of NADH (or NADPH) in aqueous medium at 254 nm, or at wavelengths longer than 320 nm (where only the reduced nicotinamide moiety absorbs), leads to generation of NAD+ (or NADP+). The reaction proceeds both in the presence and absence of oxygen. Under aerobic conditions the reaction is accompanied by formation of H2O2 at a level equimolar with that of the NADH present in solution. On irradiation at wavelengths longer than 320 nm, conversion of NADH to enzymatically active NAD+ is about 75%. Under analogous irradiation conditions, the dimers (NAD)2 and (NADP)2 undergo disproportionation to NAD+ and NADP+, respectively, to the extent of 90%. Both physicochemical and enzymatic criteria were employed to formulate mechanisms for the photooxidation of NADH and the photodisproportionation of the dimer (NAD)2.  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