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 62 毫秒
1.
Recombination rate coefficients of protonated and deuterated ions KrH+, KrD+, XeH+ and XeD+ were measured using Flowing Afterglow with Langmuir Probe (FALP). Helium at 1600 Pa and at temperature 250 K was used as a buffer gas in the experiments. Kr, Xe, H2 and D2 were introduced to a flow tube to form the desired ions. Because of small differences in proton affinities of Kr, D2 and H2 mixtures of ions, KrD+/D3+ and KrH+/H3+ are formed in the afterglow plasma, influencing the plasma decay. To obtain a recombination rate coefficient for a particular ion, the dependencies on partial pressures of gases used in the ion formation were measured. The obtained rate coefficients, αKrD+(250 K) = (0.9 ± 0.3) × 10−8 cm3 s−1 and αXeD+(250 K) = (8 ± 2) × 10−8 cm3 s−1 are compared with αKrH+(250 K) = (2.0 ± 0.6) × 10−8 cm3 s−1 and αXeH+(250 K) = (8 ± 2) × 10−8 cm3 s−1.  相似文献   

2.
A detailed spectroscopic study of the optical characteristics of the tetrahedrally coordinated Cr4+ ion in LiAlO2 and LiGaO2 is given. From absorption and excitation measurements the crystal field parameter Dq and the Racah parameter B were determined to be Dq=1065 cm−1, B=450 cm−1, and Dq/B=2.4 for LiAlO2 and Dq=1055 cm−1, B=428 cm−1, and Dq/B=2.5 for LiGaO2. For the Racah parameter C only a lower limit can be given, i.e. 2417 cm−1 for LiAlO2 and 2667 cm−1 for LiGaO2. Due to the strong crystal field splitting — caused by the low site symmetry — the 3B(3T2) crystal field component is the metastable and thus the emitting level. In the low-temperature absorption and emission spectra the expected three spin–orbit components of the 3B level are found at 8273, 8296, and 8300 cm−1 for Cr4+:LiAlO2 and 8610, 8623, and 8632 cm−1 for Cr4+:LiGaO2. The emission lifetime of Cr4+ in LiAlO2 is 95 μs at 10 K and single exponential. In Mg-codoped LiAlO2 and in LiGaO2 the Cr4+ decay is double exponential. In Cr,Mg:LiAlO2 two centers can be clearly distinguished, while in Cr:LiGaO2 a variety of centers are observed, probably due to different charge compensation processes between Li, Ga, and Cr. The quantum efficiencies at room temperature are 42% for Cr:LiAlO2 and 23% for Cr:LiGaO2. Already at low temperature nonradiative decay processes occur. The temperature dependence of the lifetimes were analyzed with the model of Struck and Fonger. Excited state absorption measurements indicate that in the spectral region of the emission the excited state absorption cross-section is larger than the stimulated emission cross-section. Therefore laser oscillation is unlikely in these systems.  相似文献   

3.
Combining a temperature variable 22-pole ion trap with a cold effusive beam of neutrals, rate coefficients k(T) have been measured for reactions of CO2+ ions with H, H2 and deuterated analogues. The neutral beam which is cooled in an accommodator to TACC, penetrates the trapped ion cloud with a well-characterized velocity distribution. The temperature of the ions, T22PT, has been set to values between 15 and 300 K. Thermalization is accelerated by using helium buffer gas. For reference, some experiments have been performed with thermal target gas. For this purpose hydrogen is leaked directly into the box surrounding the trap. While collisions of CO2+ with H2 lead exclusively to the protonated product HCO2+, collisions with H atoms form mainly HCO+. The electron transfer channel H+ + CO2 could not be detected (<20%). Equivalent studies have been performed for deuterium. The rate coefficients for reactions with atoms are rather small. Within our relative errors of less than 15%, they do not depend on the temperature of the CO2+ ions nor on the velocity of the atoms (k(T) lays between 4.5 and 4.7 × 10−10 cm3 s−1 with H as target, and 2.2 × 10−10 cm3 s−1 with D). For collisions with molecules, the reactivity increases significantly with falling temperature, reaching the Langevin values at 15 K. These results are reported as k = α (T/300 K)β with α = 9.5 × 10−10 cm3 s−1 and β = −0.15 for H2 and α = 4.9 × 10−10 cm3 s−1 and β = −0.30 for D2.  相似文献   

4.
In the present work, blends of poly(ethylene oxide) (PEO), poly(acrylonitrile-co-methyl acrylate) (PANMA) and poly(4-vinylphenol-co-2-hydroxyethyl methacrylate) (PVPh-HEM) were studied by DSC, FTIR and electrochemical impedance spectroscopy (EIS). PEO/PANMA blends were found to be immiscible, while PEO/PVPh-HEM blends are miscible and PVPh-HEM/PANMA exhibits partial miscibility behaviour. The ternary PEO/PANMA/PVPh-HEM blends exhibited miscible compositions for PVPh-HEM and PEO-rich systems. The miscibility observed is a direct consequence of the hydrogen bond interactions among the polymer chains, in which the phenol groups in PVPh-HEM interact with both PEO and PANMA chains. The proton conductivity of a selected membrane based on the ternary blend containing 60% PEO and doped with H3PO4 aqueous solution reached 8 × 10−3 Ω−1 cm−1 at room temperature and 3 × 10−2 Ω−1 cm−1 at 80 °C.  相似文献   

5.
Pulse radiolysis transient UV–visible absorption spectroscopy was used to study the UV–visible absorption spectrum (225–575 nm) of the phenyl radical, C6H5(), and kinetics of its reaction with NO. Phenyl radicals have a strong broad featureless absorption in the region of 225–340 nm. In the presence of NO phenyl radicals are converted into nitrosobenzene. The phenyl radical spectrum was measured relative to that of nitrosobenzene. Based upon σ(C6H5NO)270 nm=3.82×10−17 cm2 molecule−1 we derive an absorption cross-section for phenyl radicals at 250 nm, σ(C6H5())250 nm=(2.75±0.58)×10−17 cm2 molecule−1. At 295 K in 200–1000 mbar of Ar diluent k(C6H5()+NO)=(2.09±0.15)×10−11 cm3 molecule−1 s−1.  相似文献   

6.
The electronic UV–VIS–NIR absorption spectra of single crystalline BaTiO3−δ (BTO) are studied in the temperature range of 102–1173 K in pure oxygen and at conditions of moderate and strong reduction of the material. The strongly reduced crystals are of deep blue colour. The optical spectra of blue BTO are characterised by a strong absorption in the NIR region at around 7000 cm−1, which is attributed to polaronic defects associated with the formation of Ti3+ in the material. This assumption is supported by fits of the spectra using polaronic line shape functions appropriate for disordered systems and also by the electrical conductivity of blue BTO which, in agreement with results from the optical spectra, exhibits an activation energy of 0.20 eV. The EPR spectra of moderately reduced BTO powders show an anisotropic g-factor, which is compatible with the optical spectrum. The temperature dependence of the band gap energy of BTO was found to be given as dEg/dT = −7.21 × 10−4 eV/K.  相似文献   

7.
The characteristics, performance, and application of an electrode, namely, Pt|Hg|Hg2(IBP)2|Graphite, where IBP stands for ibuprofenate ion, are described. This electrode responds to IBP with sensitivity of (58.6 ± 0.9) mV decade 1 over the range 5.0 × 10 5–1.0 × 10 1 mol L 1 at pH 6.0–9.0 and a detection limit of 3.8 × 10 5 mol L 1. The electrode is easily constructed at a relatively low cost with fast response time (within 15–30 s) and can be used for a period of 5 months without any considerable divergence in potentials. The proposed sensor displayed good selectivity for ibuprofen in the presence of several substances, especially concerning carboxylate and inorganic anions. It was used for the direct assay of ibuprofen in commercial tablets by means of the standard additions method. The analytical results obtained by using this electrode are in good agreement with those given by the United States Pharmacopeia procedure.  相似文献   

8.
The mediated oxidation of N-acetyl cysteine (NAC) and glutathione (GL) at the palladized aluminum electrode modified by Prussian blue film (PB/Pd–Al) is described. The catalytic activity of PB/Pd–Al was explored in terms of FeIII[FeIII(CN)6]/FeIII[FeII(CN)6]1− system by taking advantage of the metallic palladium layer inserted between PB film and Al, as an electron-transfer bridge. The best mediated oxidation of NAC and GL on the PB/Pd–Al electrode was achieved in 0.5 M KNO3 + 0.2 M potassium acetate of pH 2. The mechanism and kinetics of the catalytic oxidation reactions of the both compounds were monitored by cyclic voltammetry and chronoamperometry. The charge transfer-rate limiting step as well as overall oxidation reaction of NAC or GL is found to be a one-electron abstraction. The values of transfer coefficients α, catalytic rate constant k and diffusion coefficient D are 0.5, 3.2 × 102 M−1 s−1 and 2.45 × 10−5 cm2 s−1 for NAC and 0.5, 2.1 × 102 M−1 s−1 and 3.7 × 10−5 cm2 s−1 for GL, respectively. The modifying layers on the Pd–Al substrate have reproducible behavior and a high level of stability in the electrolyte solutions. The modified electrode is exploited for hydrodynamic amperometry of NAC and GL. The amperometric calibration graph is linear in concentration ranges 2 × 10−6–40 × 10−6 for NAC and 5 × 10−7–18 × 10−6 M for GL and the detection limits are 5.4 × 10−7 and 4.6 × 10−7 M, respectively.  相似文献   

9.
A detailed study of the electrochemical reduction of diacetylbenzene A in aqueous medium between Ho = −5 and pH 14 is presented. The reactants are strongly adsorbed, so that the reactions are of a surface nature. From Ho = −5 to pH 6, a global 2e reduction yielding an enediol-type intermediate occurs. Analysis using the theory of the square schemes with protonations at equilibrium shows that, up to pH 4, the reaction is controlled by the first electron uptake, the paths being successively H+e and eH+. The elementary electrochemical surface rate constants are 9.6 × 107 s and 1.2 × 106 s for AH+ and A respectively. From pH 6 to 14, a le adsorption wave, corresponding to the formation of (a) monoradical(s), appears and is followed by a le wave due to the reduction of the radical(s). A dimerization occurs, due to the coupling A + AH, as in the case of the monocarbonyl compounds. The rate of this surface process, kd = 5 × 1013 cm2 mol−1 s−1, is markedly smaller than the rate of the homogeneous reaction obtained in alkaline ethanol by Savéant et al. for the coupling of the radicals of benzaldehyde, benzophenone and acetophenone.  相似文献   

10.
The diffusion of strontium and zirconium in single crystal BaTiO3 was investigated in air at temperatures between 1000 °C and 1250 °C. Thin films of SrTiO3, deposited by spin coating a precursor solution and thin films of zirconium, deposited onto the sample surfaces by sputtering, were used as diffusion sources. The diffusion profiles were measured by SIMS depth profiling on a time-of-flight secondary ion mass spectrometer (ToF-SIMS). The diffusion coefficients of strontium and zirconium were given by DSr = 3.6 × 102.0±4.4 exp[−(543 ± 117) kJ mol−1/(RT)] cm2 s−1 and DZr = 1.1 × 101.0±2.1 exp[−(489 ± 56) kJ mol−1/(RT)] cm2 s−1. The results are discussed in terms of different diffusion mechanisms in the perovskite structure of BaTiO3.  相似文献   

11.
Heterogeneous electrocatalytic reduction of hydrogen peroxide (H2O2) by C60 is reported for the first time. C60 is embedded in tetraoctylammonium bromide (TOAB) film and is characterized by scanning electron microscopy and cyclic voltammetry. Electrocatalytic studies show that the trianion of C60 mediates the electrocatalytic reduction of H2O2 in aqueous solution containing 0.1 M KCl. Application of such film modified electrode as an amperometric sensor for H2O2 determination is also examined. The sensor shows a fast response within 1 s and a linear response is obtained (R = 0.9986) in the concentration range from 3.33 × 10−5 to 2.05 × 10−3 mol L−1 for H2O2, with the detection limit of 2 × 10−5 mol L−1 and the sensitivity of 1.65 μA mM−1. A good repeatability and stability is shown for the sensor during the experiment.  相似文献   

12.
The spectra and kinetic behavior of solvated electrons (esol) in alkyl ammonium ionic liquids (ILs), i.e. N,N-diethyl-N-methyl-N-(2-methoxyethyl)ammonium bis(trifluoromethanesulfonyl)imide (DEMMA-TFSI), N,N-diethyl-N-methyl-N-(2-methoxyethyl)ammonium tetrafluoroborate (DEMMA-BF4), N,N,N-trimethyl-N-propylammonium bis(trifluoromethanesulfonyl)imide (TMPA-TFSI), N-methyl-N-propylpiperidinium bis(trifluoromethanesulfonyl)imide (PP13-TFSI), N-methyl-N-propylpyrrolidinium bis(trifluoromethanesulfonyl)imide (P13-TFSI), and N-methyl-N-butylpyrrolidinium bis(trifluoromethanesulfonyl)imide (P14-TFSI) were investigated by the pulse radiolysis method. The esol in each of the ammonium ILs has an absorption peak at 1100 nm, with molar absorption coefficients of 1.5–2.3×104 dm3 mol−1 cm−1. The esol decayed by first order with a rate constant of 1.4–6.4×106 s−1. The reaction rate constant of the solvated electron with pyrene (Py) was 1.5–3.5×108 dm3 mol−1 s−1 in the various ILs. These values were about one order of magnitude higher than the diffusion-controlled limits calculated from measured viscosities. The radiolytic yields (G-value) of the esol were 0.8–1.7×10−7 mol J−1. The formation rate constant of esol in DEMMA-TFSI was 3.9×1010 s−1. The dry electron (edry) in DEMMA-TFSI reacts with Py with a rate constant of 7.9×1011 dm3 mol−1 s−1, three orders of magnitude higher than that of the esol reactions. The G-value of the esol in the picosecond time region is 1.2×10−7 mol J−1. The capture of edry by scavengers was found to be very fast in ILs.  相似文献   

13.
The double-perovskite Sr2NiMoO6−δ (SNMO) was investigated as an anode material of a solid oxide fuel cell (SOFC). With a 300 μm thick La0.9Sr0.1Ga0.8Mg0.2O3−σ (LSGM) disk as electrolyte and Ba0.5Sr0.5Co0.8Fe0.2O3−δ as the cathode, the SNMO anode showed power densities of 819 mW cm−2 in hydrogen at 1123 K. Moreover, there was no buffer layer between anode and electrolyte, which would reduce design techniques and save design cost. After test no chemical reaction was discovered between anode and electrolyte. The anode exhibited good conductivity and the value was around 60 S cm−1 in H2. Also it had almost linear thermal expansion from room temperature to 1253 K and the average thermal expansion coefficient was about 12.14 × 10−6 K−1, which was quite close to that of La0.9Sr0.lGa0.8Mg0.2O3 (12.17 × 10−6 K−1) electrolyte.  相似文献   

14.
To evaluate the contribution of local pulsed heating of light-absorbing microregions to biochemical activity, irradiation of Escherichia coli was carried out using femtosecond laser pulses (λ = 620 nm, τp=3 × 10−13 s, fp = 0.5 Hz, Ep = 1.1 × 10−3J cm−2, Iav = 5.5 × 10−4 W cm−2, Ip = 109 W cm−2) and continuous wave (CW) laser radiation (λ = 632.8 nm, I = 1.3 W cm−2). The irradiation dose required to produce a similar biological effect (a 160%–190% increase in the clonogenic activity of the irradiated cells compared with the non-irradiated controls) is a factor of about 103 lower for pulsed radiation than for CW radiation (3.3 × 10−1 and 7.8 × 102 J cm−2 respectively). The minimum size of the microregions transiently heated on irradiation with femtosecond laser pulses is estimated to be about 10 Å, which corresponds to the size of the chromophores of hypothetical primary photoacceptors—respiratory chain components.  相似文献   

15.
l-Phenylalaninium maleate (abbreviated as LPM) chemical formula C9H12NO2+·C4H3O4, a new organic nonlinear optical crystal was grown by slow evaporation technique. Transparent needle shaped crystal of dimensions 7 mm × 1 mm × 0.5 mm was obtained. Single crystals of LPM have been subjected to X-ray diffraction analysis to estimate the lattice parameters and the space group. The powder X-ray diffractogram of the crystal has been recorded and the reflections from various planes are identified. The XRD studies confirm the crystalline nature. The qualitative analysis on the crystal has been carried out using Fourier transform infrared (FTIR) and Fourier transform Raman (FTRaman) spectral measurements. The presence of hydrogen and carbon in the grown crystal was confirmed by using proton and carbon nuclear magnetic resonance (NMR) spectral analyses. Optical behaviour of the crystal was investigated using UV–vis spectroscopy. The thermal stability of the crystal was analysed with the aid of thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC). The nonlinear optical (NLO) property of the crystal was tested by Nd:YAG laser source.  相似文献   

16.
Recombination of HCO+ and DCO+ ions with electrons was studied in afterglow plasma. The flowing afterglow with Langmuir probe (FALP) apparatus was used to measure the recombination rate coefficients and their temperature dependencies in the range 150–270 K. To obtain a recombination rate coefficient for a particular ion, the dependencies on partial pressures of gases used in the ion formation were measured. The variations of αHCO+(T) and αDCO+(T) seem to obey the power law: αHCO+(T) = (2.0 ± 0.6) × 10−7 (T/300)−1.3 cm3 s−1 and αDCO+(T) = (1.7 ± 0.5) × 10−7 (T/300)−1.1 cm3 s−1 over the studied temperature range.  相似文献   

17.
The hydrogen abstraction reaction of 1,1,1,2-tetrafluoroethane (HFC-134a) by chlorine radical is investigated by theoretical calculations. Equilibrium geometries and harmonic vibrational frequencies of the reactants, transition state, and products are calculated using high-level ab initio methods. Rate constants of forward and backward reactions for the temperatures from 200 to 1000 K are calculated using classical transition state theory with Eckart tunneling correction, fitted in the expressions kf (T) = 1.19 × 10−23T3.93exp (−1110/T), and kb (T) = 8.86 × 10−24T3.32exp (−959/T) cm3 molecule−1 s−1 for forward and backward reactions, respectively, and are in reasonable agreement with the available experimental values.  相似文献   

18.
A recently developed experimental and theoretical procedure is used in order to calculate the magnitude and anisotropy of interaction between a lanthanide and a 3d-metal ion. The general formula of the molecular compounds is [Ln(H2O)3(dmf)4(μ-CN)Fe–(CN)5] · nH2O where 1  n  1,5 and dmf = N,N′-dimethylformamide, abbreviated as [LnFe] from now on. The main parts of this procedure are (a) the evaluation of the effective g-parameters of the lanthanide ion with the help of EPR measurements. (b) The use of dual mode EPR spectroscopy to define the anisotropic exchange interactions with the help of an anisotropic Hamiltonian model. (c) Use of the same magnetic model to fit magnetization and susceptibility data in order to verify the EPR findings.It was possible to define some trends concerning the exchange components of the [DyFe] dimer according to which the antiferromagnetic isotropic exchange constant is smaller than 4 cm−1 and the anisotropic components are [DexcEexc] = [6(1), 0.0] cm−1. Also for the case of [TmFe] and [YbFe] dimers the antiferromagnetic isotropic exchange constant is smaller than 0.3 cm−1 while the anisotropic components are [DexcEexc] = [12.0, 0.0] cm−1 and [DexcEexc] = [0.4(1), 0.0] cm−1, respectively.  相似文献   

19.
Degradation of polyoxyethylene chain of non-ionic surfactant (TritonX-100) by chromium(VI) has been studied spectrophotometrically under different experimental conditions. The reaction rate bears a first-order dependence on the [Cr(VI)] under pseudo-first-order conditions, [TritonX-100]  [Cr(VI)] in presence of 1.16 mol dm−3 perchloric acid. The observed rate constant (kobs) was 3.3 × 10−4 to 3.5 × 10−4 s−1 and the half-life (t1/2) was 33–35 min for chromium(VI). The effects of total [TritonX-100] and [H+] on the reaction rate were determined. Reducing nature of non-ionic TritonX-100 surfactant is found to be due to the presence of –OH group in the polyoxyethylene chain. It was observed that monomeric and non-ionic micelles of TritonX-100 were oxidized by chromium(VI). When [TritonX-100] was less than its critical micelle concentration (cmc) the kobs values increased from 0.76 × 10−4 to 1.5 × 10−4 s−1. As the [TritonX-100] was greater than the cmc, the kobs values increases from 2.1 × 10−4 to 8.2 × 10−4 s−1 in presence of constant [HClO4] (1.16 mol dm−3) at 40 °C. A comparison was made of the oxidative degradation rates of TritonX-100 with different metal ion oxidants. The order of the effectiveness of different oxidants was as follows: permanganate > diperiodatoargentate(III) > chromium(VI) > cerium(IV).  相似文献   

20.
Room temperature rate coefficients and product distributions are reported for the reactions initiated in D2O with dications of the alkaline-earth metals Mg, Ca, Sr and Ba. The measurements were performed with a selected-ion flow tube (SIFT) tandem mass spectrometer and electrospray ionization (ESI). Mg2+ reacts with water by a fast electron transfer leading to charge separation with a rate coefficient of 1.4 × 10−9 cm3 molecule−1 s−1. Ca2+ reacts with D2O in a first step to form the adduct Ca2+(D2O), with an effective bimolecular rate coefficient of 2.3 × 10−11 cm3 molecule−1 s−1, which then undergoes rapid charge separation by deuteron transfer to form CaOD+ and D3O+ in a second step with k = 7.9 × 10−10 cm3 molecule−1 s−1. The CaOD+ ion reacts further by clustering up to five more D2O molecules. Sr2+ clusters up to eight D2O molecules and Ba2+ up to seven D2O molecules, with the first addition of D2O being rate determining in each case and the last addition being distinctly slower, as might be expected from a transition in the occupation of the added water molecules from an inner to an outer hydration shell.  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