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.
The reaction: F + HCl→ HF (v 3) + Cl (1), has been initiated by photolysing F2 using the fourth-harmonic output at 266 nm from a repetitively pulsed Nd: YAG laser By analysing the time-dependence of the HF(3,0) vibrational chemiluminescence, rate constants have been determined at (296 ± 5) K for reaction (1), k1 = (7.0 ± 0.5) × 10−12 cm3 molecule−1 s−1, and for the relaxation of HF(v = 3) by HCl, CO2, N2O, CO, N2 and O2: kHCl = (1.18 ±0.14) × 10−11 kCO2 = (1.04 ± 0. 13) × 10−12, kN2O = (1.41 ± 0.13) × 10−11 kCO = (2.9 ± 0.3) × (10−12, kN2 = (7.1 ± 0.6) × 10−14 and kO2 = (1.9 ± 0.6) × 10−14 cm3molecule−1s−1.  相似文献   

2.
Saran L  Cavalheiro E  Neves EA 《Talanta》1995,42(12):2027-2032
The highly neutralized ethylenediaminetetraacetate (EDTA) titrant (95–99% as Y4− anion) precipitates with Ag+ cations to form the Ag4Y species, in aqueous medium, which is well characterized from conductometric titration, thermal analysis and potentiometric titration of the silver content of the solid. The precipitate dissolves in excess Y4− to form a complex, AgY3−. Equilibrium studies at 25°C and ionic strength 0.50 M (NaNO3) have shown from solubility and potentiometric measurements that the formation constant (95% confidence level) β1 = (1.93 ± 0.07) × 105 M−1 and the solubility products are KS0 = [Ag +]4[Y4−] = (9.0 ± 0.4) × 10−18 M5 and KS1 = [Ag +]3[AgY3−] = (1.74 ± 0.08) × 10−12 M4. The presence of Na+, rather than ionic strength, markedly affects the equilibrium; the data at ionic strength 0.10 M are: β1 = (1.19 ± 0.03) × 106 M−1, KS0 = (1.6 ± 0.4) × 10−19 M5 and KS1 = (1.9 ± 0.5) × 10−13 M4; at ionic strength tending to zero; β1 = (1.82 ± 0.05) × 107 M−1, KS0 = (2.6 ± 0.8) × 10−22 M5 and KS1 = (5 ± 1) × 10−15 M4. The intrinsic solubility is 2.03 mM silver (I) in 0.50 M NaNO3. Well-defined potentiometric titration curves can be taken in the range 1–2 mM with the Ag indicator electrode. Thermal analysis revealed from differential scanning calorimetry a sharp exothermic peak at 142°C; thermal gravimetry/differential thermal gravimetry has shown mass loss due to silver formation and a brown residue, a water-soluble polymeric acid (decomposition range 135–157°C), tending to pure silver at 600°C, consistent with the original Ag4Y salt.  相似文献   

3.
The collisional quenching of electronically excited germanium atoms, Ge[4p2(1S0)], 2.029 eV above the 4p2(3P0) ground state, has been investigated by time-resolved atomic resonance absorption spectroscopy in the ultraviolet at λ = 274.04 nm [4d(1P10) ← 4p2(1S0)]. In contrast to previous investigations using the ‘single-shot mode’ at high energy, Ge(1S0) has been generated by the repetitive pulsed irradiation of Ge(CH3)4 in the presence of excess helium gas and added gases in a slow flow system, kinetically equivalent to a static system. This technique was originally developed for the study of Ge[4p2(1D2)] which had eluded direct quantitative kinetic study until recently. Absolute second-order rate constants obtained using signal averaging techniques from data capture of total digitised atomic decay profiles are reported for the removal of Ge(1S0) with the following gases (kR in cm3 molecule−1 s−1, 300 K): Xe, 7.1 ± 0.4 × 10−13; N2, 4.7 ± 0.6 × 10−12; O2, 3.6 ± 0.9 × 10−11; NO, 1.5 ± 0.3 × 10−11; CO, 3.4 ± 0.5 × 10−12; N2O, 4.5 ± 0.5 × 10−12; CO2, 1.1 ± 0.3 × 10−11; CH4, 1.7 ± 0.2 × 10−11; CF4, 4.8 ± 0.3 × 10−12; SF6, 9.5 ± 1.0 × 10−13; C2H4, 3.3 ± 0.1 × 10−10; C2H2, 2.9 ± 0.2 × 10−10; Ge(CH3)4, 5.4 ± 0.2 × 10−11. The results are compared with previous data for Ge(1S0) derived in the single-shot mode where there is general agreement though with some exceptions which are discussed. The present data are also compared with analogous quenching rate data for the collisional removal of the lower lying Ge[4p2(1D2)] state (0.883 eV), also characterized by signal averaging methods similar to that described here.  相似文献   

4.
UV spectra and kinetics for the reactions of alkyl and alkylperoxy radicals from methyl tert-butyl ether (MTBE) were studied in 1 atm of SF6 by the pulse radiolysis-UV absorption technique. UV spectra for the radical mixtures were quantified from 215 to 340 nm. At 240 nm. σR = (2.6 ± 0.4) × 10−18 cm2 molecule−1 and σRO2 = (4.1 ± 0.6) × 10−18 cm2 molecule−1 (base e). The rate constant for the self-reaction of the alkyl radicals is (2.5 ± 1.1) × 10−11 cm3 molecule−1 s−1. The rate constants for reaction of the alkyl radicals with molecular oxygen and the alkylperoxy radicals with NO and NO2 are (9.1 ± 1.5) × 10−13, (4.3 ± 1.6) × 10−12 and (1.2 ± 0.3) × 10−11 cm3 molecule−1 s−1, respectively. The rate constants given above refer to reaction at the tert-butyl side of the molecule.  相似文献   

5.
Reartes GB  Liberman SJ  Blesa MA 《Talanta》1987,34(12):1039-1042
The acidity constants of benzidine (Bz) in aqueous solutions determined potentiometrically at 25° were Ka1 = (1.11 ± 0.08) × 10−5, Ka2 = (1.45 ± 0.12) × 10−4. The apparent mixed constants in 0.1M sodium nitrate are Ka1 = (5.37 ± 0.28) × 10−6 and Ka2 = (1.14 ± 0.09) × 10−4. The ultraviolet spectra were recorded as a function of pH and analysed with these constants to obtain the absorption spectra of H2Bz2+, HBz+ and Bz; the corresponding wavelengths of maximal absorption are 247, 273 and 278 nm, and molar absorptivities 1.63 × 104, 1.76 × 104 and 2.26 × 104 1.mole−1.cm−1.  相似文献   

6.
Y. Ogata  Y. Kosugi  K. Tomizawa 《Tetrahedron》1970,26(24):5939-5944
Vitamin A,dissolved in liquid paraffin, is stable below room temperature, but suffers oxidative decomposition at 80°, giving its epoxide as the main product. The rate of decomposition of vitamin A (VA) at 80° in the presence of oxygen (partial pressure of p) and a small amount of cobaltous stearate (Co) is expressed as: d[VA]/dt = 3·15 × 10−1:[VA][Co] p1·48 + 1·51 × 10−5[VA]p0·33 + 0·33 × 10−5[VA], where the last term represents the spontaneous thermal decomposition.  相似文献   

7.
Rate constants for the reactions of OH with CH3CN, CH3CH2CN and CH2=CH-CN have been measured to be 5.86 × 10−13 exp(−1500 ± 250 cal mole−1/RT), 2.69 × 10−13 exp(−1590 ± 350 cal mole−1/RT and 4.04 × 10−12 cm3 molecule−1 s−1, respectively in the temperature range 298–424 K. These results are discussed in terms of the atmospheric lifetimes of nitrfles.  相似文献   

8.
The second-order rate constants of gas-phase Lu(2D3/2) with O2, N2O and CO2 from 348 to 573 K are reported. In all cases, the reactions are relatively fast with small barriers. The disappearance rates are independent of total pressure indicating bimolecular abstraction processes. The bimolecular rate constants (in molecule−1 cm3 s−1) are described in Arrhenius form by k(O2)=(2.3±0.4)×10−10exp(−3.1±0.7 kJmol−1/RT), k(N2O)=(2.2±0.4)×10−10exp(−7.1±0.8 kJmol−1/RT), k(CO2)=(2.0±0.6)×10−10exp(−7.6±1.3 kJmol−1/RT), where the uncertainties are ±2σ.  相似文献   

9.
The phophorescence of biacetyl induced by an energy transfer to biacetyl from triplet benzene produced in the pulse radiolysis of benzene-biacetyl mixtures has been studied. The time required to reach the maximum intensity of phosphorescence, tmax, after the electron pulse, varies as a function of biacetyl pressure at constant benzene pressure (40 torr), which gives the lifetime of triplet benzene τ = (6.7 ± 3.2) × 10−6 s and the rate constant of the energy transfer kC6H6*(T1) + biacetyl = (1.6 ± 0.7) × 10−10 cm3 molecule−1 s−1.  相似文献   

10.
J. Femi Iyun  Ade Adegite 《Polyhedron》1989,8(24):2883-2888
At 25°C, I = 1.0 M (CF3SO3Li++CF3SO3H), [H+] = 0.034–0.274 M and λ = 453 nm, the rate equation for the oxidation of Ti(H2O), 63+ by bromine was found to be: −d/[Br2]T/dt=kK/[Br2][TiIII]/[H+]+K+kK/[Br3][TiIII]/[H++K, where k = 9.2 × 10−3 M −1 s −1 and K = 4.5 × 10−3 M. At [H+] = 1.0 M, [Br] = 0.05–0.4 M, the apparent second-order rate constant decreases as [Br] increases.

The pH-dependence of the oxidation of TiIII-edta by bromine is interpreted in terms of the change in identity of the TiIII-edta species as the pH of the reaction medium changes. The second-order rate constants were fitted using a non-linear least-square computer program with (1/k0edta)2 weighting into an equation of the form: k0edta =k1+k2K1[H+]−1+k3K1K2[H+]−2/1+K1[H+[H+−1+K1K2[H+]−2, with K1 and K2 fixed as earlier determined at 9.55 × 10−3 and 2.29 × 10−9 M, respectively, for the oxidation of bromine. k1=k2=(3.1±0.32)×103M−1s−1 k3=(2.3±0.45)×106N−1s−1.

It is proposed that these electron transfer reactions proceed by univalent changes with the production of Br2.− as a transient intermediate. An outer-sphere mechanism is proposed for these reactions. The homonuclear exchange rate for TiIII-edta+TiIV-edta is estimated at 32 M−1 s−1.  相似文献   


11.
The fluorescence of single crystals of β-9,10-dichloroanthracene at 4.2 K consists solely of excimer emission (τ = 95 ± 5 ns). The absence of monomenc emission shows that excimer formation in this crystal is not a thermally activated process. This result is confirmed by excimer-excimer annihilation studies (γ(4.2 K) = 6 × 10−13, γ(298K) = 3 × 10−12 cm3 s−1).  相似文献   

12.
The reactive Kr+F2 potential energy surface is probed by two-photon, laser-induced chemical bond formation during a Kr+F2 collision. This is compared with the pulsed laser excitation (two-photon) of Kr(2p9) followed by collision with F2 leading to the formation of KrF(B, C). In addition to reporting the excitation spectrum for the two-phonon-induced collision process, these techniques were used to determine quenching rate constants of Kr2F*. Quenching by Xe gives XeF(B, C) with rate constant (1.5±0.2)×10−10 cm3 s−1; the quenching rate constant for F2 is (1.5±0.2)×10−10 cm3 s−1, and the radiative lifetime of Kr2F* is 240±35 ns. The quenching rate constant for the coupled Kr(2p8) and Kr(2p9) levels by F2 is (13±2)×10−10 cm3 s−1.  相似文献   

13.
The residence-time-dependent desorption during the deposition of polystyrene particles 736 nm in diameter on glass was studied in situ using a parallel-plate flow chamber and automated image analysis. Comparison of successively grabbed images yielded the initial desorption rate coefficient, and final desorption rate coefficient and a relaxation time for the transition from the initial to the final desorption state, i.e. ageing of the bonds. Desorption experiments were performed from suspensions with different potassium nitrate concentrations (1, 10 and 50 mM) and at varying shear rates (15–200 s−1. The initial desorption rate coefficient β0 ranging from 1 × 10−3 to 20 × 10−3s−1, and the final desorption rate coefficient β, ranging from 0.01 × 10−3 to 0.65 × 10−3s−1 were both larger than the desorption rate coefficients calculated neglecting a possible residence time dependence. These desorption rate coefficients, β, ranged from 0.005 × 10−3 to 0.40 × 10−3s−1. The relaxation times, at which the adhesion of the polystyrene particles entered a more irreversible state of adhesion compared with their initial state of adhesion, varied from 100 to 1000 s. The desorption rate coefficients as well as the relaxation time showed major variations with the shear rate and the ionic strength of the suspension. At high ionic strength, the initial and final desorption rate coefficients increase and the relaxation time decreases with increasing shear rate, whereas at low ionic strength the desorption rates decrease and the relaxation time increases with increasing ionic strength. This study provides direct evidence that the interaction forces between adhering particles and a collector surface change over time.  相似文献   

14.
A novel copper(II) thiocyanate complex [Cu(im)2(NCS)2] 1 (im=imidazole) has been prepared and characterized by spectroscopic analysis and crystallographic method. This supramolecular compound exhibits a three-dimensional solid state structure constituted by N–HS hydrogen bonds and π–π stacking interactions. The compound in DMF solutions has a very strong third-order non-linear optical (NLO) behavior with absorption coefficient and refractive index 2=1.18×10−11 mw−1, n2=−9.00×10−16 m2w−1, respectively, and third-order NLO susceptibility χ(3) of 7.00×10−10 esu.  相似文献   

15.
The kinetics of the association reaction of CF3 with NO was studied as a function of temperature near the low-pressure limit, using pulsed laser photolysis and time-resolved mass spectrometry. CF3 radicals were generated by photolysis of CF3I at 248 nm and the kinetics was determined by monitoring the time-resolved formation of CF3NO. The bimolecular rate constants were measured from 0.5 to 12 Torr, using nitrogen as the buffer gas. The results are in very good agreement with recent data published by Vakhtin and Petrov, obtained at room temperature in a higher pressure range and, therefore, the two studies are quite complementary. A RRKM model was developed for fitting all the data, including those of Vakhtin and Petrov and for extrapolating the experimental results to the low- and high-pressure limits. The rate expressions obtained are the following: k1(0) = (3.2 ± 0.8) × 10−29 (T/298)−(3.4±0.6) cm6 molecule−2 s−1 for nitrogen used as the bath gas and k1(∞) = (2.0 ± 0.4) × 10−11 (T/298)(0±1) cm3 molecule−1 s−1. RRKM calculations also help to understand the differences in reactivity between CF3 and other radicals, for the same association reaction with NO.  相似文献   

16.
Nest-shaped cluster [MoOICu3S3(2,2′-bipy)2] (1) was synthesized by the treatment of (NH4)2MoS4, CuI, (n-Bu)4NI, and 2,2′-bipyridine (2,2′-bipy) through a solid-state reaction. It crystallizes in monoclinic space group P21/n, a=9.591(2) Å, b=14.820(3) Å, c=17.951(4) Å, β=91.98(2)°, V=2549.9(10) Å3, and Z=4. The nest-shaped cluster was obtained for the first time with a neutral skeleton containing 2,2′-bipy ligand. The non-linear optical (NLO) property of [MoOICu3S3(2,2′-bipy)2] in DMF solution was measured by using a Z-scan technique with 15 ns and 532 nm laser pulses. The cluster has large third-order NLO absorption and the third-order NLO refraction, its 2 and n2 values were calculated as 6.2×10−10 and −3.8×10−17 m2 W−1 in a 3.7×10−4 M DMF solution.  相似文献   

17.
Equilibria between aluminium(III), pyrocatechol (1,2-dihydroxybenzene, H2L) and OH were studied in 0.6 M Na(Cl) medium at 25°C. The measurements were performed as emf titrations (glass electrode) within the limits 1.5 ≤ − log[H+] ≤ 9; 0.0005 ≤ B ≤ 0.015 M; 0.006 ≤ C ≤ 0.03 M and 2 ≤ C/B ≤ 30 (B and C stand for the total concentrations of aluminium(III) and pyrocatechol respectively). All data can be explained with a main series of complexes: A1L+, log β−2,1,1 = − 6.337 ± 0.005; A1L2, log β−4,1,2 = −15.44 ± 0.017 and A1L33−, log β−6,1,3 = − 28.62 ± 0.024 together with two minor species: Al(OH)L22−, log β−5,1,2 = − 23.45 ± 0.079 and Al3(OH)3L3, log β−9,3,3 = − 29.91 ± 0.066. Of the two, the latter probably is a type of average composition complex principally occurring at low C/B quotients. The first acidity constant for pyrocatechol as determined in separate experiments is log β−1,0,1 = − 9.198 ± 0.001. The standard deviations given are 3σ(log β p,q,r). Data were analyzed with the least squares computer program LETAGROPVRID. In a model calculation using kaolinite as solid phase, we compared the complexation ability of this system with that of the system Al3+-OH-salicylic acid, reported earlier in this series.  相似文献   

18.
Inam R  Somer G 《Talanta》1998,46(6):1347-1355
The polarographic reduction of lead in the presence of selenite gives rise to an additional peak corresponding to the reduction of lead (Pb) on adsorbed selenium (Se) on mercury at −0.33 V. The selenium and lead content can be determined using this peak by the addition of a known amount of one of these ions first and then the second ion. The linear domain range of lead is 5.0×10−7–2.0×10−5 M and for selenium 5.0×10−7–1.0×10−5 M. Using this method 4.90×10−7 M Se(IV) and 1.47×10−6 M Pb(II) in a synthetic sample could be determined with a relative error of +2.0% and 1.8%, respectively (n=4). A recovery test after acid digestion for a synthetic sample was 97% for selenium and 96.5% for lead. The method was applied to 1 ml of digested blood, and 328±23 μg l−1 Se(IV) and 850±62 μg l−1 Pb(II) could be determined with a 90% (n=5) confidence interval.  相似文献   

19.
The cross section for the quenching of NH(c 1Π, ν = 0) by HN3 was measured by using a pulsed laser technique. A single rotational level of NH(c 1Π, ν = 0) was formed by exciting NH(a 1Δ, ν = 0) with a frequency doubled dye laser. NH(a1Δ) was produced by photolyzing HN3 with a XeCl excimer laser. The time profiles of the NH(c-a) fluorescence were measured at various pressures of HN3. Experiments were performed both in the presence and in the absence of He buffer gas. In the absence of He, the NH radicals were found to be translationally hot; the average velocity was 3800±600 m s−1. The quenching cross sections for the translationally hot and thermalized NH(c) radicals by HN3 were determined to be (28±5) × 10−16 and (85±3) × 10−16 cm2, respectively. No rotational level dependence could be observed in the quenching of the hot NH(c) radicals.  相似文献   

20.
This Letter reports the first kinetic study of 2-butoxy radicals to employ direct monitoring of the radical. The reactions of 2-butoxy with O2 and NO are investigated using laser-induced fluorescence (LIF). The Arrhenius expressions for the reactions of 2-butoxy with NO (k1) and O2 (k2) in the temperature range 223–311 K have been determined to be k1=(7.50±1.69)×10−12×exp((2.98±0.47) kJmol−1/RT) cm3 molecule−1 s−1 and k2=(1.33±0.43)×10−15×exp((5.48±0.69) kJmol−1/RT) cm3 molecule−1 s−1. No pressure dependence was found for the rate constants of the reaction of 2-butoxy with NO at 223 K between 50 and 175 Torr.  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