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.
Fluorescence decay measurements were used to investigate the mechanism of the acid dissociation of β-naphthol in the excited singlet state. The solutions of the differential equations involved show that the naphthol fluorescence should decay exponentially while the naphtholate ion decay should involve a difference of exponentials. Comparison between the calculated and the measured decay functions indicates that about 20% of the acid dissociation takes place before the thermalized equilibrium excited state, 1S, is reached.  相似文献   

2.
Effect of the micelles of anionic, cationic and non-ionic surfactants on the fluorescence quenching of 1- and 2-naphthols has been studied in the presence of copper ion. The excited state lifetime, dynamic and static quenching constants for these systems have been determined. Fluorescence quenching in water and SDS micelle is due to the collision of the fluorophore with the quencher with a small static component. The negatively charged naphtholate ions in the excited state are quenched with significantly higher rates than the neutral naphthol molecules, which are located further inside the mesophase. CTAB micelle is less effective than the SDS micelle for fluorescence quenching. The effect of CTAB on water-assisted excited-state deprotonation has been investigated in the presence of ZnSO4. For TX-100 micelle there is negligible quenching even at higher concentration of the quencher.  相似文献   

3.
The endeavor to develop high-performance narrowband blue organic light-emitting diodes (OLEDs) with low efficiency roll-off represents an attractive challenge. Herein, we introduce a hetero-acceptor design strategy centered around the heptagonal diimide (BPI) building block to create an efficient thermally activated delayed fluorescence (TADF) sensitizer. The alignment of a twisted BPI unit and a planar diphenyltriazine (TRZ) fragment imparts remarkable exciton dynamic properties to 26tCz-TRZBPI, including a fast radiative decay rate (kR) of 1.0×107 s−1 and a swift reverse intersystem crossing rate (kRISC) of 1.8×106 s−1, complemented by a slow non-radiative decay rate (kNR) of 6.0×103 s−1. Consequently, 26tCz-TRZBPI facilitates the fabrication of high-performance narrowband pure-blue TADF-sensitized fluorescence OLEDs (TSF-OLEDs) with a maximum external quantum efficiency (EQEmax) of 24.3 % and low efficiency roll-off even at a high brightness level of 10000 cd m−2 (EQE10000: 16.8 %). This showcases a record-breaking external quantum efficiency at a high luminance level of 10000 cd m−2 for narrowband blue TSF-OLEDs.  相似文献   

4.
Relative rate constants k2/k1 for the reaction of vibrationally excited HCl(υ = 1) with oxygen atoms HCl(1)+O → OH+Cl, (k2) to the reaction HCl(0)+O → OH+Cl, (k1) was measured in a flow system (1–2 torr) at room temperature. Vacuum UV resonance fluorescence was used for the detection of the product chlorine atoms, combined with infrared fluorescence to follow the HCl (υ = 1) decay. In this way vibrational relaxation by oxygen atoms could be excluded from the pure chemical reaction rate. The measured ratio k2/k1 is 300 ± 100.  相似文献   

5.
Abstract A simple photochemical model of photosystem II consisting of antenna chlorophyll and a reaction center was used to examine the phenomenon of exciton detrapping, i.e. the transfer of excitation energy from open reaction centers back to the antenna. η, the ratio of the probability of detrapping when the reaction centers are all open, Ψt(o) to the probability when the centers are closed, Ψt(x) was used as a variable parameter to examine the various pathways of energy dissipation in a system in which P, the yield of photochemistry, and R, the ratio of the maximum to the minimum yields of fluorescence, were assumed to be known (e.g. R= 4.0 and P= 0.90). It is shown that η must fall within a range of values between 0 and R (1 –P) and that, for given values of R and P, Ψt(o) and the ratio of the rate constant for photochemistry at the reaction center, kp, to the rate constant for energy transfer back to the antenna, kt, can be determined for any assumed value of η. Even though detrapping occurs at open reaction centers, it is the magnitude of the yield of nonradiative decay at closed reaction centers, Ψa(x) which sets the upper limit on η. Equations for the overall yields of fluorescence and nonradiative decay in the antenna chlorophyll and of nonradiative decay at the reaction center chlorophyll, under conditions of both open and closed reaction centers, were derived in conventional probability terms and in terms of R, P and η. As η increases within its range of permissible values, energy dissipation in the antenna decreases and nonradiative decay at the reaction center increases. The determination of a specific value of η or of the ratio kpkt would require additional information such as the value of the maximum yield of fluorescence and the ratio of the rate constants for fluorescence and nonradiative decay in the antenna chlorophyll. The characteristics of a system in which there is no nonradiative decay in the reaction center (i.e. kd= 0), in which case R (1 –P) = 1.0, were also examined. In this case the yield of detrapping has no influence on energy dissipation in the system. Finally, the question of heterogeneity in PSII was considered. It is suggested that Ψd(x) may be greater in PSIIβ than in PSIIα so that the probability of detrapping could be greater in the PSIIα fraction.  相似文献   

6.
The luminescence induced by the Hg-photosensitized reaction of NH3 was studied by repetitive fast pulsed excitation. From observations of the decay of the luminescence, the emitter of the 290 nm fluorescence was found to be the precursor of the 340 nm emission (the stabilized complex). The first-order decay rate of the stabilized complex was found to be given by 1/τO + k[NH3], where τ0 = 2.3 μs and k = 3.5 × 10?13 cm3 molecule?1 s?1.  相似文献   

7.
The initial oxidative polymerization kinetics of 2,6-dimethylphenol (DMP) catalyzed by a Cu(II)-EDTA complex in water was studied. The initial polymerization rate of DMP (R0) increases with an increase in concentrations of DMP and catalyst. R0 firstly increases with the molar ratio of N/Cu and then decreases. The reaction order with respect to oxygen is 0.1. R0 increases with NaOH concentration and reaches its maximum value at a concentration of 0.50 mol/L. 1/R0 is in direct proportion to 1/[DMP]0, which indicates that the initial polymerization kinetics of DMP in water obeys Michaelis-Menten model. The dissociation rate constant of the intermediate complex (k2) and Michaelis-Menten constant (Km) at various temperatures are calculated. It is found that both k2 and Km increase with an increase in temperature.  相似文献   

8.
The decay of the lowest excited triplet state of 9,10-diazaphenanthrene (DAP) was investigated. A surprisingly high sublevel decay rate (ky = 3.3 × 107 s−1) in biphenyl and fluorene hosts was found at 3.0 K from the decay of a time-resolved EPR (TREPR) signal. This high decay rate was confirmed by a lifetime measurement of the delayed fluorescence using a picosecond laser and photon counting system. The nature of the fast radiationless transition is discussed.  相似文献   

9.
The alkaline fading of malachite green, which is interpreted as parallel first order and second order reactions, has been studied in cetyltrimethylammonium bromide (CTAB) micellar solution at 25°C using spectrophotometry. A micellar catalytic model is proposed in this paper for constant concentration of hydroxideion. For this model, the first order and the second order rate constants in CTAB micellar phase,k1m and k2m have been obtained.

The experimental results indicate that the first order reaction of malachite green cation with water is catalysed by CTAB micelles while the second order reaction of malachite green cation with hydroxide ion is inhibited by CTAB micelles. The first order rate constant in CTAB micellar phase, k1m , is 210 times of that in the bulk phase, but the second order rate constant in CTAB micellar phase, k2m , is 0. 166 time of that in the bulk phase. The results are interpreted mostly in relation to the micellar micropolarity and electrostatic interaction. @Keywords: Micelle, Micellar catalysis, Parallel first order and seond order reactions, Malachite green  相似文献   

10.
It has been shown that by sample application as a pulse the voltage—time relation of a galvanic sensor can generally be described by the equation
where Ua is the voltage output change, Uo the maximum voltage change at open circuit, Ra and R the external and total resistance, k1 and k2=1/RC the rate constants of charging and discharging and C the total capacity of the cell (→condenser model). k1 is related to the over-all reaction of charging and gives no immediate indication to the rate determining step. For instance k1 can be the rate constant of chemisorption or an other first order reaction. In the case of a diffusion controlled reaction it is
where K is the partition and D the diffusion coefficient in the electrolyte, S the electrode surface, d the mean thickness of the electrolyte film, V0 the cell and VL the electrolyte volume above the catalyst. The deduced equation can also be extended to a quite chemical reaction, for instance when the measuring electrode is depolarized by oxygen. k2 has then analogous meanings to k1. The mentioned formulas give some indication to the optimalisation of the cell geometry and the preparation of the electrodes. The experimental treatment includes the response of an alcohol sensor, also the determination of k1 is described and its kinetic meaning is discussed.  相似文献   

11.
The temperature dependence of the fluorescence and fluorescence excitation spectra of all-trans diphenyl hexathene (DPH) and octatetraene (DPO) in six solvents confirms the S1(1Ag*) and S2(1Bu*) state assignment, and determines their energy difference ΔE. The S1 fluorescence rate parameter kF depends on ΔE, the solvent refractive index n, the S2 (n = 1) fluorescence rate parameter kF20 (2.23 × 108 s?1 for DPH, 2.33 × 108 s?1 for DPO), and the S2-S1 coupling matrix element V (745 cm?1 for DPH, 500 cm?1 for DPO). The S1 fluorescence is induced by 1Bu*-1Ag* potential interaction (PI), via a bu vibrational mode (≈ 900 cm?1), and not by vibronic coupling. The main S1 radiationless transition, rate parameter kR, is thermally-activated internal rotation through an angle θ about the central ethylenic bond(s). The PI distorts the S1 (θ) potential surface and thus influences kR.  相似文献   

12.
The photophysics of the complex forming reaction between Quin-2 and Ca2+ were investigated using steady-state and time-resolved fluorescence measurements. The fluorescence decay traces were analyzed with global compartmental analysis yielding the following values for the rate constants at room temperature in aqueous solution with EGTA as Ca2+ buffer: k01= 8.6 times 108 s?1, k21= 1 times 1011M?1 s?1, k02= 8.8 times 107 s?1, k12= 4 times 104 s?1. k01 and k02 denote the respective deactivation rate constants of the Ca2+ free and bound forms of Quin-2 in the excited state. The constant k21 represents the second-order rate constant of binding of Ca2+ and Quin-2 in the excited state while k12 is the first-order rate constant of dissociation of the excited Ca2+:Quin-2 complex. From the estimated values of k12 and k21 the dissociation constant Kd* in the excited state was calculated. It was found that pKd* (6.4) is slightly smaller than pKd (7.2). There was no interference of the excited-state complex forming reaction with the determination of Kd. Intracellular Ca2+ concentrations can thus accurately be determined from fluorometric measurements using Quin-2 as Ca2+ indicator.  相似文献   

13.
The use of microemulsion in flow-injection systems is considered, with particular reference to the determination of C6C10 primary amines by derivatization with o-phthal-aldehyde and 2-mercaptoethanol. Microemulsions provide an ideal interface for reactions between a water-soluble reagent and a non-aqueous sample. The effect of reaction conditions (R value, droplet concentration, pH and flow rate) on the fluorescence signal is discussed and calibration data for three primary alkylamines are given.  相似文献   

14.
A tunable dye laser has been used to excite single vibronic features in the low-pressure vapor of CrO2Cl2. The fluorescence spectrum, fluorescence excitation spectrum and time-resolved fluorescence decay are discussed. It is shown that the active ν′4 and ν″4 modes are the same frequency in the gas phase, thus collapsing the sequence congestion normally observed in gas-phase spectra. This degeneracy makes impossible the excitation of single vibronic levels. It is shown that the fluorescence lifetime of the excited state in all except the vibrationally cold level is severely shortened by unimolecular radiationless decay. This radiationless rate is strongly dependent upon the partitioning of energy into various excited-state modes. The radiative lifetime of the vibrationally cold excited state is (1.34 ± 0.08) μs and the apparent bimolecular quenching rate is (5.9 ± 0.2) × 10?10 cm3/molecules. No evidence of emission from the lowest-energy excited electronic state recently reported by Spoliti [J. Mol. Spectrosc. 52 (1973) 146] is observed.  相似文献   

15.
The laser flash photolysis of ozone at ≈ 6000 Å has been used to generate a clean kinetic source of ground state atomic oxygen, O(3P). The decay of O(3P) due to reaction with O3 was monitored via resonance fluorescence at 1300 Å, under static reaction cell conditions. Over the temperature range of 220–353°K, the bimolecular rate constant, k1, could be expressed in Arrhenius form as: k1 = (2.02 ± 0.19) × 10?11 exp[-(4522 ± 210 kcal/mole)/RT]. Units are in cm3molec?1 sec-1. A comparison of the results from this work with other recent investigations, indicates that the reliability of k1 is now probably as good as 10–15% over nearly 300 degrees.  相似文献   

16.
The applications of the reverse pulse method are extended in order to distinguish between different paths of electrodimerization. The method for determination of rate constants of the chemical reaction coupled with charge transfer (ec2) is further developed. Measurements of Kd vs. tp provide valuable information for the determination of rate constants.The reverse pulse method is applied in the study of the electrodimerization of 1-alkyl-4-t-butylpyridinium ions (1-alkyl=?CH3; ?CH2CH3; ?CH(CH 3)2) in acetonitrile. It is found that the electrodimerization mechanism is consistent with the sequence: Py+ + e = Py ? 12 Py2. The formal potential Eo′ is measured directly for 1-CH(CH3)2, and indirectly for the other pyridium ions. All values are identical within experimental error (Eo′ = ?1.835 V vs. Ag/Ag+).The values of kd and km are measured. With an increase in the size of the l-alkyl group, kd decreases and km increases. This is consistent with the expected “stabilization” of the radicals due to the increase of the l-alkyl group size which hinders the dimerization site (2- and 6-positions).  相似文献   

17.
18.
Time-resolved and steady-state fluorescence studies of proflavine in aqueous solution are presented. The observation of a monoexponential fluorescence decay with a time constant decreasing with increasing pH and the presence of an anomalous red-shift in the fluorescence spectrum as a function of pH indicate the existence of a complex proton-transfer mechanism in the excited state. A reaction scheme is proposed and the corresponding proton-transfer rates are evaluated. An excited-state pK value of 12.85 is obtained for the equilibrium between the cationic form of proflavine and the same form dissociated at an amino group.  相似文献   

19.
The kinetic mechanism for the formation and decay of HCO(0,0,0) following flashlamp excitation (10 μs pulse width) into the 1A″ → 1A′ absorption transition of gas phase acetaldehyde (0.2 Torr) was examined by time-resolved intracavity laser detection (TRMD) and by phosphorescence lifetime measurements. The HCO radical was found to appear primarily in the vibrationless level reaching a maximum concentration about 250 μs after the excitation of acetaldehyde. The formation rate of HCO(0,0,0) was observed to be insensitive to an order of magnitude change in the number of collisions of excited-state acetaldehyde with either argon, cyclohexane, or the cell wall. Contrastingly, the decay rate of HCO exhibited a strong dependence on the collisional environment. The rate constants for HCO(0,0,0) decay by collisions with acetaldehyde, argon, and cyclohexane and by reaction with O2 were measured by TRILD. The rate constant for O2, quenching of 3A″ phosphorescence was also obtained.The potential for HCO(0,0,0) being either a primary or secondary dissociation product is considered in the formulation of a kinetic mechanism describing both the formation and decay behavior observed. Evidence is presented in support of a mechanism in which (1) HCO(0,0,0) is formed by the thermal reaction between acetyl radicals. CH3CO, and ground-state acetaldehyde after excited-state acetaldehyde undergoes primary dissociation to CH3CO, and (2) HCO(0,0,0) decays principally by collisionally-induced dissociation at the cell wall.  相似文献   

20.
The kinetics of the gas phase reaction between O(3P) atoms and N2O5 have been examined in a discharge-flow mass spectrometer at 4.5 torr (N2) total pressure. At O(3P) concentrations (5–10) × 1014 molecules/cm3, the decay of N2O5 was very small and only slightly greater than the data scatter. From these data, upper limits to the rate constant of this reaction was obtained at 223 and 300 K: k223 and k300 ? 3 × 10?16 cm3/molecule s.  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