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.
MSO4 (M = Zn2+, Cd2+, Hg2+) dissolves in the molten NaNO3—KNO3 eutectic and is decomposed on further heating. The kinetics of decomposition have been studied at different temperatures. The decomposition of CdSO4 and HgSO4 in the eutectic melt obey first-order kinetics whereas the decomposition of ZnSO4 at 420–460°C obeys second-order kinetics. However, at 480°C the decomposition of ZnSO4 obeys first-order kinetics. The mechanism of decomposition has been given as M2+ +SO2?4 +Na+ +K+ +2NO?3 ? (Na,K)SO4 + M2+ +2NO?3 M2+ +NO?3 → MO+NO+2 NO?3 +NO+2 → NO2 + 12O2 Some of the end products have been analysed by X-ray diffraction.  相似文献   

2.
The rate coefficients of the reactions of CN and NCO radicals with O2 and NO2 at 296 K: (1) CN + O2 → products; (2) CN + NO2 → products; (3) NCO + O2 → products and (4) NCO + NO2 → products have been measured with the laser photolysis-laser induced fluorescence technique. We obtained k1 = (2.1 ± 0.3) × 10?11 and k2 = (7.2 ± 1.0) × 10?11 cm3 molecule?t s?1 which agree well with published results. As no reaction was observed between NCO and O2 at 297 K, an upper limit of k3 < 4 × 10?17 cm3 molecule?1 S?1 was estimated. The reaction of NCO with NO2 has not been investigated previously. We measured k4 = (2.2 ± 0.3) × 10?11 cm3 molecule?1 s?1 at 296 K.  相似文献   

3.
The X‐site ion in organic–inorganic hybrid ABX3 perovskites (OHPs) varies from halide ion to bridging linkers like HCOO?, N3?, NO2?, and CN?. However, no nitrite‐based OHP ferroelectrics have been reported so far. Now, based on non‐ferroelectric [(CH3)4N][Ni(NO2)3], through the combined methodologies of quasi‐spherical shape, hydrogen bonding functionality, and H/F substitution, we have successfully synthesized an OHP ferroelectric, [FMeTP][Ni(NO2)3] (FMeTP=N‐fluoromethyl tropine). As an unprecedented nitrite‐based OHP ferroelectric, the well‐designed [FMeTP][Ni(NO2)3] undergoes the ferroelectric phase transition at 400 K with an Aizu notation of 6/mmmFm, showing multiaxial ferroelectric characteristics. This work is a great step towards not only enriching the molecular ferroelectric families but also accelerating the potential practical applications.  相似文献   

4.
Unsaturated 1,6‐dicarbonyls like 2,4‐hexadienedial are ring opening products in the OH initiated photo‐oxidation of aromatic hydrocarbons. In the present study, the photolysis of E,Z‐ and E,E‐2,4‐hexadienedial has been investigated under natural sunlight conditions in a large volume outdoor reaction chamber. In the case of the E,Z‐isomer, an extremely rapid isomerization into the E,E‐form was observed. The photoisomerization frequency, relative to that of NO2, was found to be J(E,Z‐2,4‐hexadienedial)/J(NO2) = (0.148 ± 0.012). A more complex photolysis behavior was observed for E,E‐2,4‐hexadienedial. Here, a fast equilibrium preceded a comparably slow photolysis. For the equilibrium reaction, relative frequencies of J(E,E‐2,4‐hexadienedial → EQUI)/J(NO2) = (0.113 ± 0.009) and J(EQUI → E,E‐2,4‐hexadienedial)/J(NO2) = (0.192 ± 0.016) were obtained, giving an equilibrium constant of K = (0.59 ± 0.07). For the photolysis frequencies, ratios of J(E,E‐2,4‐hexadienedial → products)/J(NO2) = J(EQUI → products)/J(NO2) = (1.22 ± 0.45)·10−2 were determined. Qualitative aerosol measurements during the experiments showed that the photolysis of 2,4‐hexadienedials is a source of secondary organic aerosol. In addition to the photolysis study, OH radical reaction rate constants were determined, values of (7.4 ± 1.9)·10−11 and (7.6 ± 0.8)·10−11 cm3 s−1 were obtained for E,Z‐ and E,E‐2,4‐hexadienedial, respectively. The results indicate that the dominant fate of E,Z‐2,4‐hexadienedial in the atmosphere will be photoisomerization, while for E,E‐2,4‐hexadienedial, both photolysis and OH radical reaction will be important sinks. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 689–697, 1999  相似文献   

5.
The solid‐state, low‐temperature linkage isomerism in a series of five square planar group 10 phosphino nitro complexes have been investigated by a combination of photocrystallographic experiments, Raman spectroscopy and computer modelling. The factors influencing the reversible solid‐state interconversion between the nitro and nitrito structural isomers have also been investigated, providing insight into the dynamics of this process. The cis‐[Ni(dcpe)(NO2)2] ( 1 ) and cis‐[Ni(dppe)(NO2)2] ( 2 ) complexes show reversible 100 % interconversion between the η1‐NO2 nitro isomer and the η1‐ONO nitrito form when single‐crystals are irradiated with 400 nm light at 100 K. Variable temperature photocrystallographic studies for these complexes established that the metastable nitrito isomer reverted to the ground‐state nitro isomer at temperatures above 180 K. By comparison, the related trans complex [Ni(PCy3)2(NO2)2] ( 3 ) showed 82 % conversion under the same experimental conditions at 100 K. The level of conversion to the metastable nitrito isomers is further reduced when the nickel centre is replaced by palladium or platinum. Prolonged irradiation of the trans‐[Pd(PCy3)2(NO2)2] ( 4 ) and trans‐[Pt(PCy3)2(NO2)2] ( 5 ) with 400 nm light gives reversible conversions of 44 and 27 %, respectively, consistent with the slower kinetics associated with the heavier members of group 10. The mechanism of the interconversion has been investigated by theoretical calculations based on the model complex [Ni(dmpe)Cl(NO2)].  相似文献   

6.
Measurements of the rate coefficient of the reaction (O3P) + NO2 → O2 + NO have been made at 296°K and 240°K, using the technique of NO2* chemiluminescent decay. Values of 9.3 × 10?12 cm3 molec?1 sec?1 at 296°K and 10.5 × 10?12 cm3 molec?1 sec?1 at 240°K were obtained, in excellent agreement with the recent results of Davis, Herron, and Huie [1]. The earlier lower values may have resulted from loss of NO2 on surfaces.  相似文献   

7.
The reactions between ceric ammonium nitrate, (NH4)2Ce(NO3)6, (CAN) and the bidentate phosphine oxides, 4,5-bis(diphenylphosphine oxide)-9,9-dimethylxanthene (L1), oxydi-2,1-phenylene bis(diphenylphosphine dioxide) (L2), 1,2-bis(diphenylphosphino)ethane dioxide (L3) and 1,4-bis(diphenylphosphino)butane dioxide, L4 have been investigated. The crystal structures of the molecular Ce(NO3)4L1 ( 1 ), and ionic [Ce(NO3)3L32][NO3]⋅CHCl3 ( 3 ), [Ce(NO3)3L32][NO3] ( 4 ) and the polymeric [Ce(NO3)3L41.5] [NO3] ( 5 ) and the cerium(III) complex [Ce(NO3)2L12][NO3] ( 2 ) are reported. The thermal stability of the complexes has been examined by thermogravimetry with the gaseous decomposition products analysed by infrared spectroscopy. Evolution of CO2 is found for both Ce(III) and Ce(IV) complexes with the later also forming NO2. The formation of the complexes in solution has been studied by 31P NMR spectroscopy and further complexes [Ce(NO3)3L12]+[NO3] and [Ce(NO3)2L13]2+2[NO3] identified in CD3CN solution. The complex ( 1 ) exists as a single molecular species in solution and is stable in dichloromethane whilst ( 3 ) decomposes on standing in both CD2Cl2 and CD3CN to Ce(III) containing species. Complexes of L2 have been identified by solution 31P NMR spectroscopy and these decompose in solution to give Ce(NO3)3L22. This study represents the first structural characterisations of Ce(IV) complexes with bidentate phosphine oxides.  相似文献   

8.
Using laser-induced fluorescence of ozone (to measure the rate of disappearance of O32) and NO2 titration (to determine O atom concentrations), we have determined bimolecular rate constants for the deactivation by O(3P atoms) of ozone in excited stretching and bending modes. These experiments do not distinguish between deactivation by (a) the exchange of vibrational and translational energy or (b) the chemical reaction O3 + O → 2O2. If the non-reactive pathway (a) is assumed to dominate, then O(3P) is 150 times more effective than O2 in deactivating O23. If chemical reaction (b) is dominant, the bimolecular rate constant for O23 + O(3P) is larger by a factor of 150–1500 than that for ground-state ozone.  相似文献   

9.
The reactions Br + NO2 + M → BrNO2 + M (1) and I + NO2 + M → INO2 + M (2) have been studied at low pressure (0.6-2.2 torr) at room temperature and with helium as the third body by the discharge-flow technique with EPR and mass spectrometric analysis of the species. The following third order rate constants were found k1(0) = (3.7 ± 0.7) × 10?31 and k2(0) = (0.95 ± 0.35) × 10?31 (units are cm6 molecule?2 s?1). The secondary reactions X + XNO2X2 + NO2 (X = Br, I) have been studied by mass spectrometry and their rate constants have been estimated from product analysis and computer modeling.  相似文献   

10.
The operational parameters of the graphite furnace for electrothermal vaporization inductively coupled plasma mass spectrometry (ETV-ICP-MS), i.e. the internal carrier gas flow rate, the total carrier gas flow rate, the sample pretreatment temperature and the volatilization temperature, are optimized for oligoelement determinations (75As, 9Be, 112Cd, 50Cr, 65Cu, 103Rh, 123Sb). The volatilization temperatures of As and Cr are compared to those obtained by graphite furnace atomic absorption spectrometry (GFAAS). Several modifiers Mg(NO3)2, Pd(NO3)2, Mg(NO3)2/Pd(NO3)2, Ni(NO3)2, KI, (NH4)H2PO4 have been tested using the concentrations recommended for GFAAS. The concentration of Mg(NO3)2 alone and in combination with NaCl has been varied to find the optimal modifier conditions. ETV-ICP-MS signal enhancements by a factor of 10 to 130 respective to those of conventional nebulization have been obtained. The optimized parameters are evaluated by analyzing the water standard reference NIST 1643c and the aqua regia solution of the lake sediment reference material BCR 280.  相似文献   

11.
The operational parameters of the graphite furnace for electrothermal vaporization inductively coupled plasma mass spectrometry (ETV-ICP-MS), i.e. the internal carrier gas flow rate, the total carrier gas flow rate, the sample pretreatment temperature and the volatilization temperature, are optimized for oligoelement determinations (75As, 9Be, 112Cd, 50Cr, 65Cu, 103Rh, 123Sb). The volatilization temperatures of As and Cr are compared to those obtained by graphite furnace atomic absorption spectrometry (GFAAS). Several modifiers Mg(NO3)2, Pd(NO3)2, Mg(NO3)2/Pd(NO3)2, Ni(NO3)2, KI, (NH4)H2PO4 have been tested using the concentrations recommended for GFAAS. The concentration of Mg(NO3)2 alone and in combination with NaCl has been varied to find the optimal modifier conditions. ETV-ICP-MS signal enhancements by a factor of 10 to 130 respective to those of conventional nebulization have been obtained. The optimized parameters are evaluated by analyzing the water standard reference NIST 1643c and the aqua regia solution of the lake sediment reference material BCR 280.  相似文献   

12.
This work analyzed the thermal decomposition of ammonium nitrate (AN) in the liquid phase, using computations based on quantum mechanics to confirm the identity of the products observed in past experimental studies. During these ab initio calculations, the CBS‐QB3//ωB97XD/6–311++G(d,p) method was employed. It was found that one of the most reasonable reaction pathways is HNO3 + NH4+ → NH3NO2+ + H2O followed by NH3NO2+ + NO3 → NH2NO2 + HNO3. In the case in which HNO3 accumulates in the molten AN, alternate reactions producing NH2NO2 are HNO3 + HNO3 → N2O5 + H2O and subsequently N2O5 + NH4+ → NH2NO2 + H2O. In both scenarios, HNO3 plays the role of a catalyst and the overall reaction can be written as NH4+ + NO3 (AN) → NH2NO2 + H2O. Although the unimolecular decomposition of NH2NO2 is thermodynamically unfavorable, water and bases both promote the decomposition of this molecule to N2O and H2O. Thus AN thermal decomposition in the liquid phase can be summarized as NH4+ + NO3 (AN) → N2O + 2H2O.  相似文献   

13.
The literature known, but not fully characterized, silver dinitramide transfer reagents AgN(NO2)2 ( 1 ), [Ag(NCCH3)][N(NO2)2] ( 2 ), and [Ag(py)2][N(NO2)2] ( 3 ) have been investigated by 109Ag, 14N NMR and vibrational spectroscopy (IR, Raman). In addition, the poorly understood [Cu(NH3)4][N(NO2)2)]2 ( 4 ) and [Pd(NH3)4][N(NO2)2]2, ( 5 ) have also been prepared and characterized by 14N NMR and vibrational spectroscopy (IR, Raman). The structures of 2 — 5 have also been determined by X‐ray diffraction.  相似文献   

14.
The decay of pernitric acid (HO2NO2) in the presence of excess nitric oxide has been studied in a 5800-liter, Teflon-lined chamber over the temperature range 254 to 283 K at 1 atm pressure of N2 by Fourier transform infrared spectroscopy. A heterogeneous reaction of NO2 and H2O2 was used to generate HO2NO2 with less than 20% HNO3 and less than 5% NO2 present as impurities. The HO2NO2 had lifetimes of 5 to 20 h in our chamber, presumably determined by heterogeneous loss to the walls. Two paths have been proposed for the reaction of NO2 with HO2:
(1), NO2 + HO2 → HONO + O2 (2). In this study the ratio k1/k2 was calculated to be greater than 103 throughout the temperature range studied. The homogeneous unimolecular decay of the HO2NO2, reaction (3), was investigated by adding excess NO in order to remove HO2 by reaction (4).
(3), HO2 + NO → NO2 + OH (4). The rate constant k3 was determined to be 1.4 × 1014 exp(?20700 ± 500/RT)s?1. The thermal decomposition lifetimes of HO2O2 at 1 atm total pressure calculated from k3 are 12 s at 298 K, 5 min at 273 K and 1 month at 220 K. Implications of these results for the role of pernitric acid in the lower and upper atmosphere are discussed.  相似文献   

15.
High-resolution spectra of the NO2 continuum emission produced from the reaction NO + O3 → NO2 + O2 have been investigated to detect any possible emission from O2(1Δg) at 1270 nm or O2(1Σ+g) at 762 nm. The photolysis of O3/O2 mixtures at 253.7 nm, which produces both states of O2 with known quantum efficiency, has been used as an internal standard. From the results it is concluded that less than 1/300 and 1/200 of the NO + O3 reactive collissions result in production of O2(1Δg) or O2(1Σ+g), respectively, at room temperature.  相似文献   

16.
The polarized absorption spectra of pyrazine (h4 and d4) in single benzene crystals have been measured at high resolution at 2 K. Vibrational assignments in the 1Ag → 1B3u (nπ*) transition have been confirmed and considerably extended. The quartic potential component of hydrogen bending mode :Oa has been found to originate primarily from vibronic coupling with the nearby 1B2u (ππ*). The coupling between in-plane 6a and out-of-plane 10a modes is described theoretically, and leads to further spectral assignments. Other out-of-plane modes 4 and 5 are identified and shown to have combination defects with 6a. A quartic component found for the out-of-plane ring bending mode 4 could not be explained by vibronic coupling.  相似文献   

17.
NH(A3Π → X3Σ?) and OH(A2Σ+ → X2Π) chemiluminescences from the reaction of CH(X2Π) with NO and O2, respectively, have been observed at room temperature. From the decay of such emissions we have measured the rate constants for these two reactions: kNO = (2.5 ± 0.5) × 10?10 and kO2 = (8 ± 3) × 10?11 cm3 molecule ?1 s?1, which are in agreement with previously reported rates determined by direct CH(X) detection using, laser-induced fluorescence. This indicates that a four-centered mechanism generating these excited species is operative in both reactions. The CH generation from 266 nm photolysis of CHBr3 has also been investigated via analysis of CH* emissions.  相似文献   

18.
Rate constants have been determined for the reaction OH + NO2 (+ N2) → HNO3 (+ N2), using time-resolved resonance absorption to follow the removal of OH radicals produced by flash photolysis of HNO3. The measurements cover the ranges: 220 ? T ? 358 K and 3.2 × 1017 ? [N2] ? 4.0 × 1018 molecule cm?3.  相似文献   

19.
Based on an FTIR-product study of the photolysis of mixtures containing Br2? CH3CHO and Br2? CH3CHO? HCHO in 700 torr of N2, the rate constant for the reaction Br + CH3CHO → HBr + CH3CO was determined to be 3.7 × 10?12 cm3 molecule?1 s?1. In addition, the selective photochemical generation of Br at λ > 400 nm in mixtures containing Br2? CH3CHO? 14NO2 (or 15NO2)? O2 was shown to serve as a quantitative preparation method for the corresponding nitrogen-isotope labeled CH3C(O)OONO2 (PAN). From the dark-decay rates of 15N-labeled PAN in large excess 14NO2, the rate constant for the unimolecular reaction CH3C(O)OO15NO2 → CH3C(O)OO + 15NO2 was measured to be 3.3 (±0.2) × 10?4 s?1 at 297 ± 0.5 K.  相似文献   

20.
The reaction of CF3 with NO2 was studied at 296 ± 2K using two different absolute techniques. Absolute rate constants of (1.6 ± 0.3) × 10−11 and (2.1 −0.3+07) × 10−11 cm3 molecule−1 s−1 were derived by IR fluorescence and UV absorption spectroscopy, respectively. The reaction proceeds via two reaction channels: CF3 + NO2 → CF2O + FNO, (70 ± 12)% and CF3 + NO2 → CF3O + NO, (30 ± 12)%. An upper limit of 11% for formation of other reaction products was determined. The overall rate constant was within the uncertainty independent of total pressure between 0.4 to 760 torr. © 1996 John Wiley & Sons, Inc.  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