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 984 毫秒
1.
The 17O-NMR signals of four true C-nitroso compounds 1–4 appear at particularly low field (1550–1265 ppm), whereas the dimers (azodioxy type) resonate at ca. 400 ppm and the ‘isonitroso compounds’ ( ? quinone-oximes; 5 and 6 ) at ca. 250 ppm. S-Nitroso compounds ( ? thionitrites; 8 and 9 ) show shift values of ca. 1300 ppm, not far from C—NO; the NO+ ion is much stronger shielded (474 ppm). The results, together with those for higher-shielded nitroso compounds X—NO (X ? RO, R2N, Cl, O?) are discussed in terms of (a) resonance stabilization through n-donation from X(π-bond order, approximated by the known barriers of rotation around the X—N bond) and of (b) electronic excitation energies ΔE. The latter are approximated by long-wave (symmetry-forbidden) UV/VIS absorptions and confirmed, where available, by the maxima of the curves of circular dichroism (CD); the CD curve of thionitrite 9 has been measured. It is found that the δ(17O) values of X—NO depend both on bond order and on ΔE, which could not be separated. The higher shielding of NO+ compared with X—N?O is explained on the basis of anisotropy effects, which differ between sp and sp2 systems.  相似文献   

2.
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   

3.
The thermal decomposition of nitrocellulose (NC) 12.1% N, has been studied with regard to kinetics, mechanism, morphology and the gaseous products thereof, using thermogravimetry (TG), differential thermal analysis (DTA), IR spectroscopy, differential scanning calorimetry (DSC) and hot stage microscopy. The kinetics of the initial stage of thermolysis ofNC in condensed state has been investigated by isothermal high temperature infrared spectroscopy (IR). The decomposition ofNC in KBr matrix in the temperature range of 142–151°C shows rapid decrease in O?NO2 band intensity, suggesting that the decomposition of NC occurs by the rupture of O?NO2 bond. The energy of activation for this process has been determined with the help of Avrami-Erofe'ev equation (n=1) and is ≈188.35 kJ·mol?1. Further, the IR spectra of the decomposition products in the initial stage of thermal decomposition ofNC, indicates the presence of mainly NO2 gas and aldehyde.  相似文献   

4.
The thermal decomposition of 4‐nitroimidazole catalyzed by Pb(NO3)2 was studied by the combination technique of in situ thermolysis cell with rapid‐scan Fourier transform infrared spectroscopy (thermolysis/RSFT‐IR). The results showed that the decomposition of 4‐nitroimidazole began with the split of the C–NO2 bond in the temperature range of 185–210 °C. The strongly oxidative product NO2 destroyed the instable annulus of 4‐nitroimidazole instantly, all the other C?N, C?C, C–H and N–H bonds of the five membered ring were broken simultaneously, and the detected gas products of 4‐nitroimidazole decomposition were NO2, CO2 and CO.  相似文献   

5.
The first silicon analogues of carbonic (carboxylic) esters, the silanoic thio‐, seleno‐, and tellurosilylesters 3 (Si?S), 4 (Si?Se), and 5 (Si?Te), were prepared and isolated in crystalline form in high yield. These thermally robust compounds are easily accessible by direct reaction of the stable siloxysilylene L(Si:)OSi(H)L′ 2 (L=HC(CMe)2[N(aryl)2], L′=CH[(C?CH2)‐CMe][N(aryl)]2; aryl=2,6‐iPr2C6H3) with the respective elemental chalcogen. The novel compounds were fully characterized by methods including multinuclear NMR spectroscopy and single‐crystal X‐ray diffraction analysis. Owing to intramolecular N→Si donor–acceptor support of the Si?X moieties (X=S, Se, Te), these compounds have a classical valence‐bond N+–Si–X? resonance betaine structure. At the same time, they also display a relatively strong nonclassical Si?X π‐bonding interaction between the chalcogen lone‐pair electrons (nπ donor orbitals) and two antibonding Si? N orbitals (σ*π acceptor orbitals mainly located at silicon), which was shown by IR and UV/Vis spectroscopy. Accordingly, the Si?X bonds in the chalcogenoesters are 7.4 ( 3 ), 6.7 ( 4 ), and 6.9 % ( 5 ) shorter than the corresponding Si? X single bonds and, thus, only a little longer than those in electronically less disturbed Si?X systems (“heavier” ketones).  相似文献   

6.
1,1,1‐Trimethylhydrazinium iodide ([(CH3)3N? NH2]I, 1 ) was reacted with a silver salt to form the corresponding nitrate ([(CH3)3N? NH2][NO3], 2 ), perchlorate ([(CH3)3N? NH2][ClO4], 3 ), azide ([(CH3)3N? NH2][N3], 4 ), 5‐amino‐1H‐tetrazolate ([(CH3)3N? NH2][H2N? CN4], 5 ), and sulfate ([(CH3)3N? NH2]2[SO4]?2H2O, 6 ?2H2O) salts. The metathesis reaction of compound 6 ?2H2O with barium salts led to the formation of the corresponding picrate ([(CH3)3N? NH2][(NO2)3Ph ‐ O], 7 ), dinitramide ([(CH3)3N? NH2][N(NO2)2], 8 ), 5‐nitrotetrazolate ([(CH3)3N? NH2][O2N? CN4], 9 ), and nitroformiate ([(CH3)3N? NH2][C(NO2)3], 10 ) salts. Compounds 1 – 10 were characterized by elemental analysis, mass spectrometry, infrared/Raman spectroscopy, and multinuclear NMR spectroscopy (1H, 13C, and 15N). Additionally, compounds 1 , 6 , and 7 were also characterized by low‐temperature X‐ray diffraction techniques (XRD). Ba(NH4)(NT)3 (NT=5‐nitrotetrazole anion) was accidentally obtained during the synthesis of the 5‐nitrotetrazole salt 9 and was also characterized by low‐temperature XRD. Furthermore, the structure of the [(CH3)3N? NH2]+ cation was optimized using the B3LYP method and used to calculate its vibrational frequencies, NBO charges, and electronic energy. Differential scanning calorimetry (DSC) was used to assess the thermal stabilities of salts 2 – 5 and 7 – 10 , and the sensitivities of the materials towards classical stimuli were estimated by submitting the compounds to standard (BAM) tests. Lastly, we computed the performance parameters (detonation pressures/velocities and specific impulses) and the decomposition gases of compounds 2 – 5 and 7 – 10 and those of their oxygen‐balanced mixtures with an oxidizer.  相似文献   

7.
The halogenotrinitromethanes FC(NO2)3 ( 1 ), BrC(NO2)3 ( 2 ), and IC(NO2)3 ( 3 ) were synthesized and fully characterized. The molecular structures of 1 – 3 were determined in the crystalline state by X‐ray diffraction, and gas‐phase structures of 1 and 2 were determined by electron diffraction. The Hal?C bond lengths in F?, Cl?, and Br?C(NO2)3 in the crystalline state are similar to those in the gas phase. The obtained experimental data are interpreted in terms of Natural Bond Orbitals (NBO), Atoms in Molecules (AIM), and Interacting Quantum Atoms (IQA) theories. All halogenotrinitromethanes show various intra‐ and intermolecular non‐bonded interactions. Intramolecular N ??? O and Hal ??? O (Hal=F ( 1 ), Br ( 2 ), I ( 3 )) interactions, both competitors in terms of the orientation of the nitro groups by rotation about the C?N bonds, lead to a propeller‐type twisting of these groups favoring the mentioned interactions. The origin of the unusually short Hal?C bonds is discussed in detail. The results of this study are compared to the molecular structure of ClC(NO2)3 and the respective interactions therein.  相似文献   

8.
The electronic influence of substituents on the free enthalpy of rotation around the N? B bond in aminoboranes was investigated in two series of compounds: (a) (CH3)2N?BCl (phenyl-p-X), containing the para-phenyl substituent at the boron atom, and (b) (p-X-phenyl)CH3N?B(CH3)2, containing the para-phenyl substituent at the nitrogen atom of the N? B linkage (X = ? NR2, ? OCH3, ? C(CH3)3, ? Si(CH3)3, ? H, ? F, ? Cl, ? Br, ? I, ? CF3 and ? NO2). By comparing the rotational barriers in corresponding compounds of both series, a reverse effect of the substituents could be observed. Electron-withdrawing substituents in the para position of the phenyl ring increase the ΔGc if the phenyl group is attached to the boron atom; on the other hand, a lower ΔGc is observed if the phenyl ring is bonded to the nitrogen atom of the N? B system. Substitution of the phenyl ring with electron-donating substituents in the paraposition exerts the opposite effect. Within each series of compounds, the differences of ΔGc values [δ(ΔGc) = ΔGc (X) ? ΔGc (X = H)] between substituted and unsubstituted compounds can be explained in terms of inductive and mesomeric effects of the ring substituents and can be correlated with the Hammett σ constant of each substituent. A comparison of the slopes of the plotted lines shows that the influence of the ring substituents is more pronounced in compounds with N-phenyl-p-X than in those with B-phenyl-p-X.  相似文献   

9.
Copper(I) Complexes with 1-Azadiene Chelate Ligands and Their Reaction with Oxygen The reaction of the bidendate 1-azadiene ligands Me2N? (CH2)n? N?CH? CH?CH? Ph with CuX results in the formation of the dimeric compounds [ A CuX]2 and [ B CuX]2 ( A : n = 2, B : n = 3, X: I, Cl). The structure of complex 1 [ A CuI]2 was determined by X-ray crystal structure analysis. 1 consists of two tetrahedrally coordinated Cu atoms connected by two iodo bridges. (Cu? Cu bond length: 261 pm). The ligand Me? N(CH2CH2N?CH? CH?CH? Ph)2 ( C ) reacts with CuX to form the monomeric complexes [ C CuX] ( 5 : X?I, 6 : X?Cl). The crystal structure of 5 shows that the ligand acts as a tridendate ligand. The bond lengths of the CuN(sp2) bonds are significantly shorter than the Cu? N(sp3) distance. Reacting the podand-type ligands N(CH2CH2? N?CH? R)3 ( D : R?Ph, E : R?-CH?CH? Ph) with CuX yields the ionic complexes 7 [ D Cu][CuCl2] and 8 [ E Cu][CuCl2]. 7 was characterized by X-ray analysis which confirmed that D acts as a four-dendate podand ligand. The compounds 1 ? 8 are unreactive towards CO2 but take up O2 even at deep temperatures. At ?78°C the orange-red complex 4 [ B CuCl]2 reacts with O2 in CH2Cl2 to form a deep violet solution, but the primary product of the oxidation could not be isolated. It reacts at room temperature to form the green complex 9 [μ-Cl, μ-OH][ B CuCl]2. The X-ray structure analysis of 9 confirms that a dimeric CuII complex is formed in which both a chloro- and a hydroxo group are bridging the monomeric units. The CuII centers exhibit a distorted tetragonal-pyramidal coordination. The pathway of the reaction with O2 will be discussed.  相似文献   

10.
Polysulfonyl Amines. XLVI. Molecular Adducts of Di(organosulfonyl)amines with Dimethyl Sulfoxide and Triphenylphosphine Oxide. X-Ray Structure Determination of Di(4-fluorobenzenesulfonyl)amine-Dimethyl Sulfoxide(2/1) From equimolar solutions of the respective components in CH2Cl2/petroleum ether, the following crystalline addition compounds were obtained: (X? C6H4SO2)2NH …? OS(CH3)2, where X = H, 4? CH3, 4? Cl, 4? Br, 4? I, 4? NO2 or 3? NO2; [(4? F? C6H4SO2)2NH]2 · (OS(CH)3)2 ( 8 ); (4? I? C6H4SO2)2NH · OP(C6H5)3. A (2/1) complex of (4? F? C6H4SO2)2NH with OP(C6H5)3 could not be isolated. The solid-state structure of the (2/1) compound 8 is compared with the known structure of the (1/1) complex (CH3SO2)2NH · OS(CH3)2. The crystallographic data for 8 at ?95°C are: monoclinic, space group C2/c, a = 2 369.9(13), b = 1 006.8(4), c = 2 772.6(13) pm, β = 110.71(4)°, U = 6.187 nm3, Z = 8. Two N? H …? O hydrogen bonds with N …? O 275 and 280 pm connect the disulfonylamine molecules with the dimethyl sulfoxide molecule. The O atom of the latter has a trigonal-planar environment consisting of the S atom and the two hydrogen bond H atoms.  相似文献   

11.
The reactions X? + HCR2ONO2 → XH + R2C=O + ?NO2 are very exothermic due to the cleavage of the weak N?O bond and the formation of the energy-intensive C=O bond. The quantum chemical calculation of the transition state of these reactions for X? = Et? and EtO? used as examples showed that they actually proceed in one elementary act as eliminations with concerted fragmentation. The kinetic parameters were estimated within the framework of the intersecting parabolas model; the parameters allow the calculation of the activation energy and rate constant from the enthalpy of the above reaction. For a series of reactions involving the Et?, EtO?, RO?2, and ?NO2 radicals, on the one hand, and a number of alkyl nitrates, on the other, their enthalpies, activation energies, and rate constants were calculated. Based on the data obtained, new kinetic schemes of the chain decomposition of alkyl nitrates involving eliminations with fragmentation were proposed.  相似文献   

12.
The Schiff base enaminones (3Z)‐4‐(5‐ethylsulfonyl‐2‐hydroxyanilino)pent‐3‐en‐2‐one, C13H17NO4S, (I), and (3Z)‐4‐(5‐tert‐butyl‐2‐hydroxyanilino)pent‐3‐en‐2‐one, C15H21NO2, (II), were studied by X‐ray crystallography and density functional theory (DFT). Although the keto tautomer of these compounds is dominant, the O=C—C=C—N bond lengths are consistent with some electron delocalization and partial enol character. Both (I) and (II) are nonplanar, with the amino–phenol group canted relative to the rest of the molecule; the twist about the N(enamine)—C(aryl) bond leads to dihedral angles of 40.5 (2) and −116.7 (1)° for (I) and (II), respectively. Compound (I) has a bifurcated intramolecular hydrogen bond between the N—H group and the flanking carbonyl and hydroxy O atoms, as well as an intermolecular hydrogen bond, leading to an infinite one‐dimensional hydrogen‐bonded chain. Compound (II) has one intramolecular hydrogen bond and one intermolecular C=O...H—O hydrogen bond, and consequently also forms a one‐dimensional hydrogen‐bonded chain. The DFT‐calculated structures [in vacuo, B3LYP/6‐311G(d,p) level] for the keto tautomers compare favourably with the X‐ray crystal structures of (I) and (II), confirming the dominance of the keto tautomer. The simulations indicate that the keto tautomers are 20.55 and 18.86 kJ mol−1 lower in energy than the enol tautomers for (I) and (II), respectively.  相似文献   

13.
In order to study the properties of new energetic compounds formed by introducing nitroazoles into 2,4,6-trinitrobezene, the density, heat of formation and detonation properties of 36 nitro-1-(2,4,6-trinitrobenzene)-1H-azoles energetic compounds are studied by density functional theory, and their stability and melting point are predicted. The results show that most of target compounds have good detonation properties and stability. And it is found that nitro-1-(2,4,6-Trinitrophenyl)-1H-pyrrole compounds and nitro-1-(2,4,6-trinitrop-enyl)-1H-Imidazole compounds have good thermal stability, and their weakest bond is C NO2 bond, the bond dissociation energy of the weakest bond is 222–238 kJ mol−1 and close to 2,4,6-trinitrotoluene (235 kJ mol−1). The weakest bond of the other compounds may be the C NO2 bond or the N N bond, and the strength of the N N bond is related to the nitro group on azole ring.  相似文献   

14.
Mechanistic aspects of the effect of the X and Y substituents (X = Me, H, CF3, CN, Br, Cl, F, OH, NH2; Y = H, NMe2, NH2, CN, NO2) on the carbonyl bond in 4-YC6H4C(O)X compounds are discussed on the basis of the 13C and 17O NMR data.  相似文献   

15.
Polysulfonyl Amines. VII. Aliphatic Trisulfonyl Amines The compounds N(SO2R1)2(SO2R2) with R1 = R2 = CH3 ( 2a ), R1 = R2 = C2H5 ( 2b ) and R1 = CH3, R2 = C2H5 ( 2c ) are prepared by cleavage of aminostannanes (CH3)3SnN(SO2R1)2 with sulfonyl chlorides R2SO2Cl. A simple synthesis of 2a from AgN(SO2CH3)2 and CH3SO2Cl is described. From the vibrational spectra of 2a , evidence is obtained for a planar NS3 group in this compound. X-ray structure determinations of 2b and HN(SO2C2H5)2 ( 3 ) are reported. In 2b , the NS3 group is approximately planar (S? N? S bond angles 119.0 ± 0.6°, sum of bond angles at N 356.9°); the S? N bond lengths of ca. 173 pm indicate a bond order of 1. In compound 3 , the nitrogen atom has a planar coordination (S? N? S angle 125.3°, sum of bond angles at N 359.3°), the S? N bond lengths of ca. 165 pm correlate with a bond order of 1.3? 1.4.  相似文献   

16.
A procedure for preparing Li4[Ru(NO2)6]…12H2O and K4[Ru(NO2)6] is described. We have carried out an X-ray diffraction study of polycrystals (DRON-UM1 diffractometer, R=192mm, CuKα radiation, Ni filter) and single crystals (CAD-4 automatic diffractometer, MoKα radiation, graphite monochromator, ω/2? scan mode). Crystal data for Li4[Ru(NO2)6]…12H2O are a=11.749(9), c=16.807(12) Å, space group $I\bar 4$ , V=2320.0 Å3, Z=4, dcalc=1.778 g/cm3; for K4 [Ru(NO2)6]: a=8.595(1) Å, α=53.23 (2)o, space group $R\bar 3$ , V=367.0 Å3, Z=1, dcalc=2.414 g/cm3. The structures are composed of [Ru(NO2)6]4? comple anions, alkaline metal cations, and crystallization water molecules. In both compounds, the coordination polyhedra of Ru are nearly regular octahedra formed by nitrite nitrogen atoms. The Ru-N bond lengths are 2.054–2.102 Å. The NO2 groups have virtually the same geometric characteristics: the N?O distances are within 1.196–1.316 Å and the O?N?O angles are 112.9–117.4o. The coordination environment of alkaline metals by oxygen atoms is discussed.  相似文献   

17.
The syntheses and reactivity of the two N‐heterocyclic carbene (NHC)→ silylene complexes 2 and 4 have been investigated. The latter are easily accessible by reaction of the zwitterionic, N‐heterocyclic silylene LSi: 1 [L=Ar‐N‐C(=CH2)CH?C(Me)‐N‐Ar, Ar=2,6‐iPr2C6H3] with 1,3,4,5‐tetramethylimidazol‐2‐ylidene and 1,3‐diisopropyl‐4,5‐dimethylimidazol‐2‐ylidene, respectively. While compound 2 undergoes facile rearrangement above ?20 °C to give the unsymmetrical N‐heterocyclic silylcarbene 3 , the derivative 4 remains unchanged even after boiling in benzene. The remarkable reactivity of 3 and 4 towards cyclohexylisocyanide has been examined which leads in a unique series of C? H, Si? H, and C? N bond activations to the new triaminosilanes 5 and 6 , respectively. The novel compounds 3 , 4 , 5 , and 6 were fully characterized by 1H, 13C, and 29Si NMR spectroscopy, EI‐MS, elemental analysis, and single‐crystal X‐ray diffraction.  相似文献   

18.
Phosphorane Iminato Complexes of Sulfur. Syntheses and Crystal Structures of [O3SS(NPPh3)2] · CH3CN, [SO(NPPh3)2], and [SCl(NPMe3)2]Cl The title compounds have been prepared by the reaction of Me3SiNPPh3 with SO2 and SOCl2, respectively, and by the reaction of Me3SiNPMe3 with S2Cl2. They form colourless, moisture sensitive crystals, which were characterized by IR spectroscopy and by crystal structure determinations. [O3SS((NPPh3)2)] · CH3CN : Space group Pca21, Z = 4, structure solution with 4016 observed unique reflections, R = 0.050. Lattice dimensions at ?60°C: a = 1865.1, b = 1168.4, c = 1569.0 pm. The compound has a zwitterionic structure with a S? S bond length of 218.2 pm and bond lengths S? N of 161.2 and P? N of 160.1 pm. [SO(NPPh3)2] : Space group P21/c, Z = 4, structure solution with 2854 observed unique reflections, R = 0.113. Lattice dimensions at ?50°C: a = 1173.1, b = 1585.6, c = 1619.2 pm, b? = 98.13°. The compound forms monomeric molecules, in which the positions of S and N atoms are disordered in two positions. The bond lengths are S? N 166 pm and P? N 163 pm in average. [SCl(NPMe3)2]Cl : Space group P1 , Z = 2, structure solution with 2416 observed unique reflections, R = 0.038. Lattice dimensions at 20°C: a = 613.2, b = 1030.3, c = 1111.4 pm, α = 88.48°, b? = 88.01°, γ = 83.10°. The compound forms ions [SCl(NPMe3)2]+ and Cl?. In the cation the sulfur atom is ?-tetrahedrally coordinated with a long S? Cl distance of 246.9 pm and bond lengths S? N of 155.3 pm and P? N of 164.3 pm in average.  相似文献   

19.
The kinetic regularities of the heat release during the thermal decomposition of liquid NH4N(NO2)2 at 102.4–138.9 °C were studied. Kinetic data for decomposition of different forms of dinitramide and the influence of water on the rate of decomposition of NH4N(NO2)2 show that the contributions of the decomposition of N(NO2)2 and HN(NO2)2 to the initial decomposition rate of the reaction at temperatures about 100 °C are approximately equal. The decomposition has an autocatalytic character. The analysis of the effect of additives of HNO3 solutions and the dependence of the autocatalytic reaction rate constant on the gas volume in the system shows that the self-acceleration is due to an increase in the acidity of the NH4N(NO2)2 melt owing to the accumulation of HNO3 and the corresponding increase in the contribution of the HN(NO2)2 decomposition to the overall rate. The self-acceleration ceases due to the accumulation of NO3 ions decreasing the equilibrium concentration of HN(NO2)2 in the melt. For Part 2, see Ref. 1. Translated fromIzvestiya Akademii Nauk, Seriya Khimicheskaya, No. 3, pp. 395–401 March 1998.  相似文献   

20.
A new class of bidentate, aza‐based phosphinic amide ligands of the type RN(H)P(?O)(2‐py)2 (2‐py = 2‐pyridyl) was synthesized within minutes via a one‐pot process including Staudinger reaction of an organic azide (RN3) with 2‐pyridylphosphines, followed by partial, unprecedented hydrolysis under loss of one aromatic substituent. The structure of the unusual‐hydrolysis product H2C?CH(CH2)9N(H)P(?O)(2‐py)2 ( 5a ) was characterized by IR, 1H‐ and 31P‐NMR, as well as by X‐ray crystal‐structure analysis (Figure). The tetrahedral P‐atom was found to be surrounded by a trigonal‐pyramidal arrangement of the substituents. To gain insight into the formation of these novel phosphinic amides, a series of intermediate iminophosphoranes, H2C?CH(CH2)9N?P(Ar)n(2‐py)3 ? n (n = 0–3), compounds 1a – 1f , were synthesized, and their hydrolyses were studied. All tested compounds followed the classical hydrolysis route of P?N cleavage under acidic conditions. Sequential hydrolysis to 5a – 5d only occurred under either basic conditions or in wet MeCN as solvent. Notably, H2C?CH(CH2)9N?P(C6H5)(4‐MeO‐2‐py)2 ( 1c ) was hydrolyzed at a much slower rate compared to its analogue 1b lacking the MeO group. On the contrary, the halogenated compounds H2C?CH(CH2)9N?P(4‐X‐C6H4)3 ( 1f,g ) (X = F, Cl) were hydrolyzed at a notably faster rate relative to the non‐halogenated congener 1e (X = H).  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