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 375 毫秒
1.
The kinetic regularities of the thermal decomposition of dinitramide in aqueous solutions of HNO3, in anhydrous acetic acid, and in several other organic solvents were studied. The rate of the decomposition of dinitramide in aqueous HNO3 is determined by the decomposition of mixed anhydride of dinitramide and nitric acid (N4O6) formed in the solution in the reversible reaction. The decomposition of the anhydride is a reason for an increase in the decomposition rates of dinitramide in solutions of HNO3 as compared to those in solutions in H2SO4 and the self-acceleration of the process in concentrated aqueous solutions of dinitramide. The increase in the decomposition rate of nondissociated dinitramide compared to the decomposition rate of the N(NO2)2 anion is explained by a decrease in the order of the N−NO2 bond. The increase in the rate constant of the decomposition of the protonated form of dinitramide compared to the corresponding value for neutral molecules is due to the dehydration mechanism of the reaction. For Part 1, see Ref. 1. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 1, pp. 41–47, January, 1998.  相似文献   

2.
Kinetic regularities of thermal decomposition of dinitramide in aqueous and sulfuric acid solutions were studied in a wide temperature range. The rate of the thermal decomposition of dinitramide was established to be determined by the rates of decomposition of different forms of dinitramide as the acidity of the medium increases: first, N(NO2) anions, then HN(NO2)2 molecules, and finally, protonated H2N(NO2)2 + cations. The temperature dependences of the rate constants of the decomposition of N(NO2) (k an) and HN(NO2)2 (kac) and the equilibrium constant of dissociation of HN(NO2)2 (K a) were determined:k an=1.7·1017 exp(−20.5·103/T), s−1,kac=7.9·1016 exp(−16.1·103/T), s−1, andK a=1.4·10 exp(−2.6·103/T). The temperature dependences of the decomposition rate constant of H2N(NO2)2 + (k d) and the equilibrium constant of the dissociation of H2N(NO2)2 + (K d) were estimated:k d=1012 exp(−7.9·103/T), s−1 andK d=1.1 exp(6.4·103/T). The kinetic and thermodynamic constants obtained make it possible to calculate the decomposition rate of dinitramide solutions in a wide range of temperatures and acidities of the medium. In this series of articles, we report the results of studies of the thermal decomposition of dinitramide performed in 1974–1978 and not published previously. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 12, pp. 2129–2133, December, 1997.  相似文献   

3.
This qualitative study examines the response of the novel energetic material ammonium dinitramide (ADN), NH4N(NO2)2, to thermal stress under low heating rate conditions in a new experimental apparatus. It involved a combination of residual gas mass spectrometry and FTIR absorption spectroscopy of a thin cryogenic condensate film resulting from deposition of ADN pyrolysis products on a KCl window. The results of ADN pyrolysis were compared under similar conditions with the behavior of NH4NO3 and NH2NO2 (nitramide), which served as reference materials. NH4NO3 decomposes into HNO3 and NH3 at 182°C and is regenerated on the cold cryostat surface. HNO3 undergoes presumably heterogeneous loss to a minor extent such that the condensed film of NH4NO3 contains occluded NH3. Nitramide undergoes efficient heterogeneous decomposition to N2O and H2O even at ambient temperature so that pyrolysis experiments at higher temperatures were not possible. However, the presence of nitramide can be monitored by mass spectrometry at its molecular ion (m/? 62). ADN pyrolysis is dominated by decomposition into NH3 and HN(NO2)2 (HDN) in analogy to NH4NO3, with a maximum rate of decomposition under our conditions at approximately 155°C. The two vapor phase components regenerate ADN on the cold cryostat surface in addition to deposition of the pure acid HDN and H2O. Condensed phase HDN is found to be stable for indefinite periods of time at ambient temperature and vacuum conditions, whereas fast heterogeneous decomposition of HDN at higher temperature leads to N2O and HNO3. The HNO3 then undergoes fast (heterogeneous) decomposition in some experiments. Gas phase HDN also undergoes fast heterogeneous decomposition to NO and other products, probably on the internal surface (ca. 60°C) of the vacuum chamber before mass spectrometric detection. © 1993 John Wiley & Sons, Inc.  相似文献   

4.
Stabilization of ammonium dinitramide in the liquid phase   总被引:1,自引:0,他引:1  
Andreev  A. B.  Anikin  O. V.  Ivanov  A. P.  Krylov  V. K.  Pak  Z. P. 《Russian Chemical Bulletin》2000,49(12):1974-1976
The kinetics of accumulation of the main products of thermal decomposition of ammonium dinitramide in the melt was investigated. The isotope composition of nitrogen-containing gases evolved by the decomposition of 15NH4N(NO2)2 and NH4 15N(NO2)2 was found. Easily oxidized salts, amines, amides, iodides, and other compounds soluble in the melt interfere with the liquid-phase decomposition of ammonium dinitramide.  相似文献   

5.
The structures of two salts [Co(NH3)6][Rh(NO2)6] (I) and [Co(NH3)6][(NO2)3Rh(μ-NO2)1+x (μ-OH)2−x Rh(NO2)3]·(2−x)(H2O), x = 0.17 (II) are solved. Single crystals of the salts are obtained by the counter diffusion method through the gel of aqueous solutions of [Co(NH3)6]Cl3 and Na3[Rh(NO2)6]. The structure of [Co(NH3)6][Rh(NO2)6] is consistent with the diffraction data for a polycrystalline sample of poorly soluble fine salt formed in the exchange reaction between aqueous solutions of [Co(NH3)6]Cl3 and Na3[Rh(NO2)6]. The structure of [Co(NH3)6][(NO2)3Rh(μ-NO2)1+x (μ-OH)2−x Rh(NO2)3]·(2−x)(H2O), x = 0.17 exhibits the stabilizing effect of a large cation in the formation of novel, unknown previously coordination ions: [(NO2)3Rh(μ-NO2)(μ-OH)2Rh(NO2)3]3− and [(NO2)3Rh(μ-NO2)2(μ-OH)Rh(NO2)3]3−.  相似文献   

6.
The reductions of [Co(CN)5NO2]3−, [Co(NH3)5NO2]2+ and [Co(NH3)5ONO]2+, by TiIII in aqueous acidic solution have been studied spectrophotometrically. Kinetic studies were carried out using conventional techniques at an ionic strength of 1.0 mol dm−3 (LiCl/HCl) at 25.0 ± 0.1 °C and acid concentrations between 0.015 and 0.100 mol dm−3. The second-order rate constant is inverse—acid dependent and is described by the limiting rate law:- k2 ≈ k0 + k[H+]−1,where k=k′Ka and Ka is the hydrolytic equilibrium constant for [Ti(H2O)6]3+. Values of k0 obtained for [Co(CN)5NO2]3−, [Co(NH3)5NO2]2+ and [Co(NH3)5ONO]2+ are (1.31 ± 0.05) × 10−2 dm3 mol−1 s−1, (4.53 ± 0.08) × 10−2 dm3 mol−1 s−1 and (1.7 ± 0.08) × 10−2 dm3 mol−1 s−1 respectively, while the corresponding k′ values from reductions by TiOH2+ are 10.27 ± 0.45 dm3 mol−1 s−1, 14.99 ± 0.70 dm3 mol−1 s−1 and 17.93 ± 0.78 dm3 mol−1 s−1 respectively. Values of K a obtained for the three complexes lie in the range (1–2) × 10−3 mol dm−3 which suggest an outer-sphere mechanism.  相似文献   

7.
Thermal behaviour of nickel amine complexes containing SO4 2−, NO3 , Cl and Br as counter ions and ammonia and ethylenediamine as ligands have been investigated using simultaneous TG/DTA coupled with mass spectroscopy (TG/DTA–MS). Evolved gas analyses detected various transient intermediates during thermal decomposition. The nickel ammonium sulphate complex produces NH, N, S, O and N2 species. The nickel ammonium nitrate complex generated fragments like N, N2, NO, O2, N2O, NH2 and NH. The halide complexes produce NH2, NH, N2 and H2 species during decomposition. The ligand ethylenediamine is fragmented as N2/C2H4, NH3 and H2. The residue hexaamminenickel(II) sulphate produces NiO with crystallite size 50 nm. Hexaammine and tris(ethylenediamine)nickel(II) nitrate produce NiO in the range 25.5 nm and 23 nm, respectively. The halide complexes produce nano sized metallic nickel (20 nm) as the residue. Among the complexes studied, the nitrate containing complexes undergo simultaneous oxidation and reduction.  相似文献   

8.
An XRD analysis is used to study the single crystal of [Pd(NH3)4][Rh(NH3)(NO2)5] double complex salt at T = 150(2) K. Crystallographic characteristics are as follows: a = 7.6458(5) ?, b = 9.8813(6) ?, c = 9.5788(7) ?, β = 109.469(2)°, V = 682.30(8) ?3, P21/m space group, Z = 2, d x = 2.553 g/cm3. The geometry of the complex [Rh(NH3)(NO2)5]2− anion is described for the first time: Rh-N(NO2) distances are 2.020(4)–2.060(3) ?, Rh-N(NH3) 2.074(4) ?, N(NO2)-Rh-N(NH3) trans-angle is 178.8(2)°.  相似文献   

9.
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   

10.
1,3,3-Trinitroazetidine (TNAZ) was synthesized using the alternative approach based on the transformation of 3-oximino-1-(p-toluenesulfonyl)azetidine in the reaction with nitric acid through intermediate pseudonitrol. The thermal decomposition of TNAZ in the gas phase, melt and m-dinitrobenzene solution in a wide concentration range (5–80%) was studied by manometry, volumetry, thermogravimetry, IR spectroscopy, and mass spectrometry. In the gas phase in the temperature range from 170 to 220°C the thermal decomposition proceeds according to the first-order kinetic law with the activation energy 40.5 kcal mol?1 and pre-exponential factor 1015.0 s?1. The major gaseous reaction products are N2, NO, NO2, CO2, H2O, and nitroacetaldehyde, and trace amounts of CO and HCN are formed. The rate-determining step of the process is the homolytic cleavage of the N-NO2 bond in the TNAZ molecule. In melt at 170–210 °C the thermal decomposition proceeds with the pronounced self-acceleration and the maximum reaction rates are observed at conversions 53.9–67.4%. The solid decomposition products accelerate the reaction. It is most likely that the autocatalysis of TNAZ decomposition in the liquid phase is due to the autocatalytic decomposition of 1-nitroso-3,3-dinitroazetidine, which is formed by the thermal decomposition of TNAZ. In m-dinitrobenzene TNAZ also decomposes with self-acceleration. The higher the concentration in the solution, the more pronounced the self-acceleration. Additives of picric acid moderately accelerate the thermal decomposition of TNAZ, whereas hexamethylenetetraamine additives exert a strong acceleration.  相似文献   

11.
The effect of ammonium nitrate concentration in the citric acid biosynthesis by Aspergillus niger NC-12 in single-stage continuous cultures with biomass retention was investigated. Experiments were carried out in a BIOMER laboratory fermenter with 5 dm3 working volume. At the initial stage of each cultivation, the substrate in the bioreactor contained 1.5 g NH4NO3 dm−3. After 120 h onwards, the bioreactor was fed continuously at a constant dilution rate of 0.009 h−1. NH4NO3 concentration in the feed was varied from one culture to another, ranging between 0.5 g dm−3 and 2.5 g dm−3. Promising results were obtained when NH4NO3 concentration of 1.5 g dm−3 was used. The observed concentration of citric acid (c P) and yield of citric acid with respect to the introduced sucrose (Y P/S) were 117.88 g dm−3 and 78.59 %, respectively. The efficiency coefficient of citric acid biosynthesis (K ef) was very high, amounting to 83.38. Presented at the 33rd International Conference of the Slovak Society of Chemical Engineering, Tatranské Matliare, 22–26 May 2006.  相似文献   

12.
The decomposition of compounds Y[CH2C(NO2)2X]2 (X=NO2 and F; Y=CH2C(O)O and OCH2O) in the liquid phase (melt, solution) was found to proceedvia the same mechanism (homolytic cleavage of the C−N bond) as in the gas phase. Some stabilizing effects of the Oβ atom and independence of the gas evolution rate constant (measured by the yield of final products) on the number of the −C(NO2)2X groups were found and interpreted. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 12, pp. 2455–2458, December, 1998.  相似文献   

13.
The thermal decompositions of polycrystalline samples of [Ni(NH3)6](NO3)2 were studied by thermogravimetric analysis with simultaneous gaseous products of the decomposition identified by a quadruple mass spectrometer. Two measurements were made for samples placed in alumina crucibles, heated from 303 K up to 773 K in the flow (80 cm3 min?1) of Ar 6.0 and He 5.0, at a constant heating rate of 10 K min?1. Thermal decomposition process undergoes two main stages. First, the deamination of [Ni(NH3)6](NO3)2 to [Ni(NH3)2](NO3)2 occurs in four steps, and 4NH3 molecules per formula unit are liberated. Then, decomposition of survivor [Ni(NH3)2](NO3)2 undergoes directly to the final decomposition products: NiO1+x, N2, O2, nitrogen oxides and H2O, without the formation of a stable Ni(NO3)2, because of the autocatalytic effect of the formed NiO1+x. Obtained results were compared both with those published by us earlier, by Farhadi and Roostaei-Zaniyani later and also with the results published by Rejitha et al. quite recently. In contradiction to these last ones, in the first and second cases agreement between the results was obtained.  相似文献   

14.
The extraction of uranyl nitrate by the novel extractant N,N’-dimethyl-N,N’-dioctylsuccinylamide (DMDOSA) from aqueous nitric/nitrate solutions was investigated. The effects of concentration of HNO3 and DMDOSA on the U(VI) extraction distribution was studied. The extraction mechanism was established and the stoichiometry of the main extracted species was confirmed to be UO2(NO3)2·2DMDOSA. The value of ΔH of the extraction is −23.9±1.7 kJ·mol−1. A IR spectral study of the U(VI) extracted species was also made.  相似文献   

15.
The rates of heat release in the nitrogen dioxide—n-decane system at a molar ratio of nitrogen oxides ton-decane (β) from 2.4·10−3 to 3.1 and gaseous volumes per mole ofn-decane (V(g)) equal to 0.05–4.5 were studied in the 55.2–92.8 °C temperature range. The initial rate of the process is determined by the interaction of NO2 withn-decane. The equilibrium constants of dissociation of N2O4 inn-decane and Henry's constants of NO2 and N2O4 in ann-decane solution were determined by complex analysis of the thermodynamic equilibrium in the NO2n-decane system and dependences of the initial rates onV(g) and β. The experimentally observed self-acceleration of the process in the region of high β and lowT values was suggested to be due to the reaction of N2O4 with intermediate oxidation products. The rate constants of the reaction of NO2 withn-decane were compared with analogous values determined in its mixtures with HNO3 solutions. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 10, pp. 1789–1794, October, 1997.  相似文献   

16.
Precursor compounds [Co(NH3)6][Rh(NO2)6] and [Co(NH3)6][Co(NO2)6], solid solutions [Co(NH3)6] [Rh(NO2)6]1−x [Co(NO2)6] x , and solid solutions Na3[Rh1−x Co x (NO2)6] were synthesized and studied by IR spectroscopy and elemental, X-ray phase, X-ray diffraction, and thermogravimetric analyses. X-ray phase analysis was employed to examine products of thermal decomposition of precursors in the atmospheres of hydrogen and helium. Catalysts with a Co-Rh active system, supported by ZrO2, were prepared and tested in the reaction of steam conversion of ethanol.  相似文献   

17.
The extraction of Pu(IV) and Th with tridodecylamine—xylene mixtures from about 6M nitric acid soil leach solutions was studied as a function of the chemical composition of the aqueous phase (iron and calcium concentration, acidity) and the amine concentration in the extractant. No correlation was found between the partition coefficients of Pu(IV) and Th and the composition parameters mentioned above at any of the amine concentrations examined. The slope, in a bilogarithmic plot, of the partition coefficients versus the amine concentrations was found to be close to 2 for Pu(IV) as well as Th in pure 6.5M nitric acid solution, thus indicating the presence of the complexes Pu(NO3) 6 2− and Th(NO3) 6 2− in the extract. When the pure nitric acid solution was replaced by soil leach solutions of similar molarity in HNO3, the slope remained 2 for Pu(IV), but changed to 1.5 for Th. A possible reason for this slope yielded by Th may be the coexistence of the complexes Th(NO3) 6 2− and Th(NO3) 5 in the extraction phase. Presented at the 4th SAC Conference on Analytical Chemistry, Birmingham 1977.  相似文献   

18.
Methods were developed for the controlled thermal synthesis of high-spin cubane-like pivalates {MII 43−OR)4} (M = Co or Ni; R = H or Me) starting from mono-and polynuclear complexes. The solid-state thermal decomposition of the known pivalate clusters [MII 43−OMe)4−(μ2−OOCBut)22−OOCBut)2(MeOH)4] and the new clusters [M4II3)−OH41−OOCBut)3−(μ−(NH2)2C6H2Me2)31−(NH2)2C6H2Me2)3]+(OOCBut)− (M = Co or Ni) was studied by differential scanning calorimetry and thermogravimetry. The thermolysis of cubane-like CoII and NiII pivalates is a destructive process. The phase composition of the decomposition products is determined by the nature of coordinated ligands and the structural features of the metal core.  相似文献   

19.
The properties of supported bimetallic Rh-Co/ZrO2 catalysts in ethanol steam reforming into hydrogen-containing gas were studied. The particles of Rh-Co solid solutions on the catalyst surface were prepared by the thermal decomposition of the double complex salt [Co(NH3)6][Rh(NO2)6] and the solid solution Na3[RhCo(NO2)6]. It was found that the bimetallic Rh-Co/ZrO2 catalysts exhibited high activity in the reaction of ethanol steam reforming. The equilibrium composition of reaction products was attained at 500–700°C and a reaction mixture space velocity of 10000 h−1.  相似文献   

20.
The kinetics of nucleophilic substitution of pyridine in bis-cationic [Pt(L)(py)]2+ complexes (L=SNS, NNN, NSN) [SNS=bis(methylthiomethyl)pyridine, NNN=bis(2-pyridylmethyl)amine, NSN=bis(2-pyridylmethyl)sulphide] by a series of nucleophiles (Cl, Br, I, N3, (C2H5)2S, NH3, thiourea (tu), NO2, C5H10NH, SeCN, SCN, CN when L=SNS; Cl, Br, I, N3, (C2H5)2S, SCN, NH3, NO2 when L=NNN; Br, N3, NO2, NH3, C5H10NH when L=NSN) have been measured in MeOH at 25 °C, μ =0.1 mol dm−3 (LiClO4 or LiCF3SO3). The logarithms of the second-order rate constants calculated at μ=0, log k° 2, do not follow the dependence upon the n° Pt scale. In particular, the reactivity of the biphilic reagents tu, SeCN, SCN and, to a lesser extent, NO 2, towards these doubly charged substrates is largely lower than expected on the basis of the n° Ptscale. There are good linear relationships between logk° 2 for the bis-cationic substrate [Pt(SNS)(py)]2+, chosen as the standard, and log k° 2 for the same reactions with [Pt(NNN)(py)]2+, [Pt(NSN)(py)]2+ and other double charged complexes previously studied. A new wide nucleophilicity scale based on [Pt(SNS)(py)]2+, that is appropriate to all the bis-cationic substrates, is here proposed  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