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 46 毫秒
1.
Pyrene fluorescence quenching by phenylazide derivatives with donor and acceptor substituents has been studied by fluorescence spectroscopy and flash photolysis. The rate constants of quenching (k q) in acetonitrile ((0.2–1.2) × 1010 l mol?1 s?1) are found to be close to a diffusion limit; the rate constants were somewhat higher for perfluoro-substituted arylazides. It is found that k q does not depend on solvent polarity; the formation of the pyrene cation in the course of pyrene fluorescence quenching by tolylazide was not detected. Pyrene fluorescence quenching occurred by an energy-transfer mechanism; this is supported by the coincidence of the quantum yields of the direct and sensitized photodecomposition of tolylazide. As estimated, energy transfer in rigid media occurs at characteristic distances of about 10 Å.  相似文献   

2.
the rate constant for quenching of the triplet state (3B1u) of benzene vapor by nitric oxide has been determined from a study of the flash photosensitization of biacetyl phosphorescence. The rate constant has been found to be (1.08 ± 0.2) × 10?11 cm3 molecule?1 sec?1. Although this value is much larger than expected, it is in agreement with rate constants reported for quenching of some liquid phase aromatics.  相似文献   

3.
Ruthenocene quenches triplet states of organic molecules with energies greater than 24000 cm?1 in benzene solution at a diffusion controlled rate , (6 ± 1) × 109 dm3 mol?1 s?1. For triplets with energies less than this the efficiency of quenching is dependent on the energy of the triplet state being quenched but drops off less acutely than expected for endothermic energy transfer following the Arrhenius equation. This is in agreement with the lowest triplet state of ruthenocene being geometrically distorted as expected from the previously observed large Stokes shift between absorption to and emission from its lowest triplet state. Similarities to ferrocene quenching of triplet states are discussed. Quenching of the triplet state of benzil by ruthenocene does not fall on the smooth curve which exists between the quenching rate constants kq and the energy of the triplet state being quenched. Queching of triplet benzil by ruthenocene is therefore attributed to favourable charge-transfer interactions, also in this case the behaviour is analogous to quenching of triplet methylene-blue by ferrocene where at least a proportion of electron transfer following quenching has been previously established.  相似文献   

4.
The rate constants for oxygen quenching in benzene solution of the triplet states of several organic compounds with relatively high triplet energies have been measured in laser photolysis and pulse radiolysis experiments. The previously observed trend for aromatic hydrocarbons where the quenching rate constants decrease from a limiting value of about one ninth of that expected for a diffusion controlled reaction to lower values for triplet states with increasing triplet energy was not observed for the triplet states of certain aromatic ketones and amines. The higher rate constants observed, e.g. oxygen quenching of triplet N-methyl indole has kQ = 1.4 × 1010 dm3 mol?1 s?1, are interpreted as being due to the presence of low lying triplet charge-transfer states which enhance the efficiency of quenching.  相似文献   

5.
Emission quenching of [Ru(bpy)2(4, 4'-dcbpy)] (PF6)2 (1) by benzenamine,4-[2-[5-[4-[4-dimethylamino]phenyl]-4,5-di-hydro-1-phenyl-1H-pyrazol-3-yl]-ethenyl]-N,N-dimetyl (2) or 1, 5-diphenyl-3-(2-phenothiazine)-2-pyrazoline (3) was observed. Measurements of the emission decay of 1 before and after addition of 2 or 3 by single photon counting technique con-finned the observations. The emission quenching of 1 by 2 or 3 was submitted to Stern-Volmer equation. It was calculated that the quenching rate constants (kq) are 5.5 × 109(mol/L)-1s-1 for 2 and 4.0 × 109(mol/L)-1s-1 for 3, respectively. These results indicated a character of dynamic quenching process. The singlet-state of 2 or 3 was also quenched by 1. The quenching behaviors did not conform to the Stern- Volmer equation and involved both static and dynamic quenching processes. The apparent quenching rate constant (kapp) was calculated to be 3 × 109 (mol/L)-1 for the interaction of excited 2 with 1, and 1.2 × 109 (mol/L)-1 for that of excited 3 wit  相似文献   

6.
7.
Pyrazolo[3,4‐b]quinoline derivatives are reported to be highly efficient organic fluorescent materials suitable for applications in light‐emitting devices. Although their fluorescence remains stable in organic solvents or in aqueous solution even in the presence of H2O, halide salts (LiCl), alkali (NaOH) and weak acid (acetic acid), it suffers an efficient quenching process in the presence of protic acid (HCl) in aqueous or ethanolic solution. This quenching process is accompanied by a change in the UV spectrum, but it is reversible and can be fully recovered. Both steady‐state and transient fluorescence spectra of 1‐phenyl‐3,4‐dimethyl‐1H‐pyrazolo‐[3,4‐b]quinoline (PAQ5) during quenching are measured and analyzed. It is found that a combined dynamic and static quenching mechanism is responsible for the quenching processes. The ground‐state proton‐transfer complex [PAQ5 ??? H+] is responsible for static quenching. It changes linearly with proton concentration [H+] with a bimolecular association constant KS=1.95 M ?1 controlled by the equilibrium dissociation of HCl in ethanol. A dynamic quenching constant KD=22.4 M ?1 is obtained by fitting to the Stern–Volmer equation, with a bimolecular dynamic quenching rate constant kd=1.03×109 s?1 M ?1 under ambient conditions. A change in electron distribution is simulated and explains the experiment results.  相似文献   

8.
The quenching rate for O(1S) by O(3P) into the O(1D) + O(1D) channel is calculated using a theoretically calculated spin-orbit coupling matrix element. An upper bound to the rate is found to be 2.0 × 10?14 cm3 s?1 which is much smaller than the experimental value. The low value of the rate constant is the result of a spin-orbit coupling matrix element of about 3 cm?1 at the relevant curve corssing.  相似文献   

9.
Fluorescence decays have been measured for pyrenesulfonate (PyS?) in the presence of anthraquinone-1-sulfonate (AQ1S?), for pyrene (Py) at various concentrations, and Py in the presence of AQ1S? in 5 mmol·dm?3 didodecyldimethylammonium bromide (DDABr) vesicle solutions. Reaction rate constants were found to increase in the order Py*–Py < PyS?*–AQ1S? < Py*–AQ1S?. Fluorescence decays have also been measured in 5 mmol·dm?3 didodecyldimethylammonium chloride (DDACl) solutions; the results showed micellar solution behavior, and micellar aggregation numbers and quenching rate constants were obtained. The quenching rate constants in vesicles and micelles were converted to the two dimensional rate constants having the same units. Two dimensional diffusion coefficient and reaction distance were obtained for the PyS?*–AQ1S? system on the DDABr vesicle surface.  相似文献   

10.
The heme in horseradish peroxidase (HRP) was replaced by phosphorescent Pt‐mesoporphyrin IX (PtMP), which acted as a phosphorescent marker of oxygen quenching and allowed comparison with another probe, Pd‐mesoporphyrin IX (Khajehpour et al. (2003) Proteins 53, 656–666). Benzohydroxamic acid (BHA), a competitive inhibitor of the enzyme, was also used to monitor its effects on phosphorescence quenching. With the addition of BHA, in the presence of oxygen, the phosphorescence intensity of the protein increased. In contrast, the addition of BHA, in the absence of oxygen, reduced the phosphorescence intensity of the protein. Kd= 18 μM when BHA binds to PtMP‐HRP. The effect of BHA can be explained by two factors: ( 1 ) BHA reduces the accessibility of O2 to the protein interior and ( 2 ) BHA itself quenches the phosphorescence. Consistent with this, the oxygen quenching of the phosphorescence of PtMP‐HRP gave a quenching constant of kq= 234 mm Hg?1 s?1 in the absence of BHA and kq= 28.7 mm Hg?1 s?1 in the presence of BHA. The quenching rate of BHA is 4000 s?1. The relative quantum yield of the phosphorescence of the Pt derivative is about six times that of the Pd derivative, whereas the phosphorescence lifetime is approximately eight times shorter. The high quantum yield and suitable lifetime make Pt‐porphyrins appropriate as sensors of O2 diffusion and flexibility in heme proteins.  相似文献   

11.
The fluorescence of pyrene and five other aromatic hydrocarbons solubilized with caffeine in aqueous solution was quenched by 0.001–1mol dm?3 sulfuric acid; this quenching did not occur in the absence of caffeine. Exceptionally a non-alternative hydrocarbon, fluoranthene, showed an increase of the fluorescence intensity by the acid. The observed quenching rate was explained by the rate of energy transfer from the aromatic molecules to protonated caffeine in the complex.  相似文献   

12.
The concept of quenching pairs is proposed for a system of randomly distributed centers which can quench one another through an energy cross-relaxation mechanism. The formula for the quenching rate is derived for the case of an exciton migrating over the centers. Comparison is made with experimental data for cross-relaxational quenching of the vibrationally excited molecule CH3CCl3. The micro-interaction parameters are estimated to be CDD = 2 × 10?33 cm6/s and CDA = 3 × 10?35 cm6/s.  相似文献   

13.
Concentration effect of silver nanoparticles (AgNPs) on the photophysical properties of ciprofloxacin (Cip) have been investigated using optical absorption and fluorescence techniques. When performed AgNPs solution was added to the Cip solution, metal-enhanced fluorescence intensity and a blue-shift of 20 nm in the maximum emission spectra of Cip has been observed. The enhanced intensity of this system is strongly dependent on the AgNPs concentration and largest at the 6.0 × 10?6 mol L?1. With increase of AgNPs concentration, quenching of fluorescence is observed. Stern–Volmer quenching constants have been calculated at four temperatures. The results show the quenching constants are directly correlated with temperature. It indicates the quenching mechanism is the dynamic quenching in nature rather than static quenching. From which we determined the activation energy for the quenching of Cip-AgNPs to be about 31.1 kJ mol?1. In addition, in the presence of optimum AgNPs concentration, a sensitive fluorimetric method for the determination of ciprofloxacin at the range 5.0 × 10?7–3.0 × 10?5 mol L?1 and the detection limit of 2 × 10?8 mol L?1 in solution is proposed.  相似文献   

14.
Hydroxyaryl alkyl tellurides are effective antioxidants both in organic solution and aqueous biphasic systems. They react by an unconventional mechanism with ROO. radicals with rate constants as high as 107 M ?1 s?1 at 303 K, outperforming common phenols. The reactions proceed by oxygen atom transfer to tellurium followed by hydrogen atom transfer to the resulting RO. radical from the phenolic OH. The reaction rates do not reflect the electronic properties of the ring substituents and, because the reactions occur in a solvent cage, quenching is more efficient when the OH and TeR groups have an ortho arrangement. In the presence of thiols, hydroxyaryl alkyl tellurides act as catalytic antioxidants towards both hydroperoxides (mimicking the glutathione peroxidases) and peroxyl radicals. The high efficiency of the quenching of the peroxyl radicals and hydroperoxides could be advantageous under normal cellular conditions, but pro‐oxidative (thiol depletion) when thiol concentrations are low.  相似文献   

15.
Free radical fragments produced in the photoinduced electron transfer from triethylamine (TEA) to excited pyrenebutyltrimethylammonium (*PyBu+) lead to 1-vi-nyl-2-pyrrolidinone (VP) and 2-hydroxyethyl methacry-late (HEMA) polymerization. Experiments carried out in water/acetonitrile solvent mixtures showed that the polymerization rate of VP increases upon increasing the water content, whereas the polymerization rate of HEMA follows the opposite trend. These results are interpreted in terms of the strong dependence on the solvent properties of the photochemical behavior of PyBu+ in the presence of the amine or monomers. Thus, the *PyBu+ quenching by VP is almost negligible in both solvents (water and acetonitrile). Whereas, the *PyBu+ quenching rate constant by HEMA in water is 4 times 109M?l s?1 and decreases four orders of magnitude in acetonitrile. The quenching of *PyBu+ by TEA in aqueous solutions is controlled by hydrogen-bonding interactions between water molecules and the amine. Quantum yields of the pyrene radical anion (φPy) also strongly depend on the water content, decreasing from 0.28 to 0.015 upon going from acetonitrile to water.  相似文献   

16.
The interaction between a water-soluble cationic fluorescent conjugated polymer (WCFP) and bovine serum albumin (BSA) was studied using UV?CVis absorption, fluorescence and circular dichroism spectroscopies. The results show that the fluorescence of BSA is strongly quenched by the WCFP under physiological conditions (pH?=?7.4). The quenching mechanism was found to be static, which was confirmed by the quenching rate constant (Kq) and UV?CVis absorption spectra. The thermodynamic parameters (?H ??, ?S ?? and ?G ??) calculated from the complexation constant, determined according to Lineweaver?CBurk equations are 38.6?kJ·mol?1, 228?J·mol?1·K?1 and ?29.4?kJ·mol?1 at 298?K. The principal interaction was proposed to be electrostatic.  相似文献   

17.
Quenching of Ru(bpy)32+ electrochemiluminescence (ECL) by Cl?, Br?, and I? ions was studied as a function of halide concentration in a bipolar electrochemical cell. All of the halides investigated showed similar qualitative behavior: above a critical concentration, ECL intensity was found to decrease linearly as the halide ion concentration was increased, due to dynamic quenching of Ru(bpy)32+ ECL. Stern‐Volmer slopes (KSV) of 0.111±0.003, 4.2±0.3, and 6.2±0.3 mM?1 were measured for Cl?, Br? and I?, respectively. The magnitude of KSV correlates with halide ion oxidation potential, consistent with an electron transfer quenching mechanism. Using the bipolar platform described herein, aqueous, halide‐containing solutions could be quantified rapidly using the sequential standard addition method. The lower detection limit is determined by a complex mechanism involving the competitive electrooxidation of halide ions and the ECL co‐reactants, as well as the passivation of the surface of the bipolar electrode, and was found to be 0.20±0.01, 0.08±0.01 and 10±1 mM, respectively, for I?, Br?, and Cl?. The performance of the bipolar ECL quenching assay is comparable to previously published fluorescence quenching methods for the determination of halide ions, while being much simpler and less expensive to implement.  相似文献   

18.
The quenching of toluene fluorescence by cis-6-phenyl-2-hexene has been studied to determine the rate of singlet “energy hopping” in dilute solutions of alkylbenzenes. The singlet lifetime date have been analyzed by the Stern-Volmer method to give the quasi-isoenergetic rate, kq, as 1.2 × 1010 M?1?1. The result is consistent with an excimer formation dissociation mechanism for alkylaromatic singlet energy transfer in dilute solution.  相似文献   

19.
The mechanism of eosin-sensitized photoreduction of benzil with 1-benzyl-1,4-dihydronicotinamide — a model compound of NAD(P)H and the behavior of the excited states of eosin have been investigated. The effect of anthracene as a diffusion-controlled quencher of the photoreaction indicates that both excited triplet state and an unquenchable excited singlet state of eosin participated in the sensitized photoreaction. From the Stern-Volmer plot of quantum yield vs. anthracene concentration, the triplet reaction rate constant has been calculated to be 0.78 × 108 L M?1S?1 while the singlet reaction rate constant determined from quenching of eosin fluorescence by benzil is equal to 7.2 × 109 L M?1S?1. The singlet and triplet quantum yields are also determined to be 0.09 and 0.18 respectively. Since both the singlet and triplet energies of eosin are lower than that of benzil, energy transfer sensitization is not feasible. It is proposed that electron transfer from the excited eosin to benzil is responsible for the initiation.  相似文献   

20.
Excitation of a mixture of dimethylethylamine (DEMA) and benzene in n-hexane at 222 nm primarily produces excited amine, while at 261 nm excited benzene predominantly results. The fluorescence spectra appreciably overlap. With 222 nm excitation, DEMA fluorescence is quenched by benzene at the diffusion-controlled rate; this quenching results with nearly unit efficiency in sensitized benzene fluorescence. With 261 nm excitation, some sensitized DEMA fluorescence is observed: the rate constant for tins process is ≈ 2.6 × 109 M?1 s?1.  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