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.
S. Ramgopal 《Tetrahedron letters》2007,48(23):4043-4045
Hunsdiecker reactions with α,β-unsaturated carboxylic acids were conducted under solvent-free conditions in the presence of a few drops of HNO3 together with a variety of metal nitrates [Mg(NO3)2, Sr(NO3)2], Al(NO3)3, Ca(NO3)2, Ni(NO3)2, Cd(NO3)2, Zn(NO3)2, Hg(NO3)2, AgNO3, ZrO(NO3)2, UO2(NO3)2, Th(NO3)2] or ammonium nitrate. α,β-Unsaturated aromatic carboxylic acids underwent nitro decarboxylation to afford β-nitro styrenes in moderate to good yields, while α,β-unsaturated aliphatic carboxylic acids underwent decarboxylation to yield the corresponding nitro derivatives.  相似文献   

2.
Osmotic coefficients for Mn(NO3)2, Co(NO3)2, Ni(NO3)2, Cu(NO3)2, Zn(NO3)2, and Mg(NO3)2 in aqueous solution have been determined by the isopiestic method at 25°C, and activity coefficients have been derived. The results agree with the literature data for Zn(NO3)2, while they are significantly different for Co(NO3)2, Cu(NO3)2, and Mg(NO3)2, and those for Mn(NO3)2 and Ni(NO3)2 are new. The concentration dependence of the osmotic coefficients for the bivalent metal nitrates is similar to that for the trifluoroacetates, while it differs from those for the other salts of the same series of metals. The results are discussed in terms of the inner-sphere and outer-sphere association of ions, auxiliary information being derived from the concentration effects in the visible spectra of the coloured metal nitrates.  相似文献   

3.
The solid-liquid equilibria of the quasi-quaternary system H2O-Zn(NO3)2·6H2O-Cu(NO3)2·3H2O-NH4NO3 were studied at 25°C by using a synthetic method based on conductivity measurements. Three isoplethic sections has been established at 25°C and the stable solid phases which appear are: NH4NO3(IV), Zn(NO3)2·6H2O, anhydrous Cu(NO3)2, Cu(NO3)2·3H2O and metastable Cu(NO3)·2.5H2O. Neither double salts, nor mixed crystals are observed at these temperatures and composition range.  相似文献   

4.
Gamma-ray induced decomposition of some divalent nitrates, viz. Mg(NO3)2·6H2O, Ca(NO3)2·4H2O, Sr(NO3)2, Ba(NO3)2, Zn(NO3)2·6H2O, Cd(NO3)2·4H2O, Hg(NO3)2·2H2O, Mn(NO3)2·4H2O, Cu(NO3)2·3H2O and trivalent nitrates, viz. Al(NO3)3·9H2O, Fe(NO3)3·9H2O, Cr(NO3)3·9H2O, Y(NO3)3·6H2O, In(NO3)3·3H2O, La(NO3)3·6H2O, Ce(NO3)3·6H2O, Pr(NO3)3·6H2O, Bi(NO3)3·5H2O has been studied in solid state at room temperature. G(NO 2 ) values (after applying appropriate dose correction) have been found to vary in the range 0.12–3.16 and 0.069–2.15 for divalent and trivalent nitrates respectively. G'-values were calculated by dividing G by the ratio of number of electrons in nitrate ion to the total number of electrons in the nitrate salt. Cation size, its polarizing power, available free space in the crystal lattice and the number and location of water molecules seem to play a dominant role in radiolytic decomposition. For Zn, Sr, In, La and Ce nitrates dose variation studies have been carried out.  相似文献   

5.

This study investigated the reactive dissolution of nitric oxide (NO) and nitrogen dioxide (NO2) mixtures in deionized water. The dissolution study was carried out in a flat surface type gas–liquid reaction chamber utilizing a gas flow-pattern resembling plasma jets which are often used in biomedical applications. The concentration of NO and NO2 in the gas mixtures was varied in a broad range by oxidizing up to 800 ppm of nitric oxide in Ar carrier gas with variable amount of ozone. The production of nitrite (NO2?) and nitrate (NO3?) in the water was proportional to treatment time up to 50 min. The concentration of NO3? was a power function of gas phase NO2 while the concentration of NO2? increased approximately linearly with gas phase NO2. The formation of NO2? and NO3? could be described by reactions between dissolved NO2 and NO in the water while the production rate was determined by diffusion-limited mass transport of nitrogen oxides to the bulk of the liquid. At higher NO2 concentrations, the formation of dinitrogen tetraoxide (N2O4) increased the formation rate of NO2? and NO3?. The identified mass transport limitation by diffusion suggests that convection of water created by the gas jet is insufficient and dissolution of nitrogen oxides can be increased by additional mixing. In respect of practical applications, the ratio of NO2? /NO3? in water could be varied from 0.8 to 5.3 with treatment time and gas phase NO2 and NO concentrations.

  相似文献   

6.
Solid electrolyte cells for detecting NO2 with Ba(NO3)2 or Sr(NO3)2 partially replaced by γ-Al2O3 as solid electrolytes have been studied. The cell tension depends on the NO2- as well as on the NO-concentration. Investigations of the establishment of the NO2-NO-O2-equilibrium by the catalytic effects of the used electrode materials Pt and Au have shown that a decomposition of NO2 below 400°C cannot be expected.  相似文献   

7.
The reactions of naphthalene in N2O5? NO3? NO2? N2? O2 reactant mixtures have been investigated over the temperature range 272–297 K at ca. 745 torr total pressure and at 272 K and ca. 65 torr total pressure using long pathlength Fourier transform infrared absorption spectroscopy. 2,3-Dimethyl-2-butene was added to the reactant mixtures at 272 K to rapidly scavenge the NO3 radicals both initially present in the added N2O5 and formed from the thermal decomposition of N2O5 during the reactions. The data obtained in the presence and absence of added 2,3-dimethyl-2-butene showed that napthalene undergoes initial reaction with the NO3 radical to form an NO3-naphthalene adduct, which either rapidly decomposes back to the reactants (at a rate of ca. 5 × 105 s?1 at 298 K) or reacts exclusively with NO2 to form products. When NO3 radicals, N2O5 and NO2 are in equilibrium, this overall process is kinetically equivalent to reaction of naphthalene with N2O5, and previous kinetic and product studies have indeed assumed the reactions of naphthalene and alkyl-substituted naphthalenes in N2O5? NO3? NO2? air mixtures to be with N2O5, and not with NO3 radicals.  相似文献   

8.
Crystalline NO2[Fe(NO3)4] was obtained by dehydration of a solution of Fe(NO3)3 in 100 % HNO3 and subsequent sublimation. NO2[Zr(NO3)5] was synthesized by reaction of ZrCl4 with N2O5 followed by sublimation in vacuum. X‐ray single crystal structure determination showed both compounds to consist of nitronium cations, NO2+, and nitratometalate anions. N‐O distances in the linear NO2+ cations are in the range of 1.08—1.13Å. In both [Fe(NO3)4] and [Zr(NO3)5] anions, all nitrate groups are coordinated bidentately with average M‐O distances 2.134 and 2.293Å, respectively. Taking into account the position of N atoms around the M atoms, the arrangement of nitrate groups can be described as tetrahedral for the Fe complex and trigonal‐bipyramidal for the Zr complex. There are four shortest N(nitronium)····O(nitrate group) contacts with average distances of 2.705 and 2.726Å in NO2[Fe(NO3)4] and 2.749Å in NO2[Zr(NO3)5]. Nitronium pentanitratohafnate is isotypic to the zirconium complex.  相似文献   

9.
Reactions of F2NC(NO2)3 with metal fluorides (KF and CsF) in DMF yield a substitution product of the fluorine atom for one nitro group, F2NC(NO2)2F. The reaction of F2NC(NO2)3 with LiBr in ethanol or DMF affords Br(NO2)C=NF rather than the expected bromo derivative F2NC(NO2)2Br.  相似文献   

10.
The primary stage of the decomposition of compounds RN(NO2)CH2C(NO2)2X is the homolytic cleavage of the C?NO2 bond, at X=NO2 and N?NO2 bond at X=F. The inductive effect of substituents decreases the dissociation energies of the C?N and N?N bonds by 1–2 kcal mol?1. Kinetic effects caused by the spatial interaction of groups and by stepwise decomposition of polyfunctional compounds are described.  相似文献   

11.
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   

12.
NO3-type and NO2-type adsorbed species are formed on Cu-ZSM-5 together with adsorbed O species at 523 K in the decomposition of NO accompanied by the evolution of N2, N2O, and NO2. NO3-type adsorbed species formed on Cu-ZSM-5 is well reduced with CH4 around 570 - 600K, while NO2-type adsorbed species formed on Cu-ZSM-5 was less active on the reduction with CH4.  相似文献   

13.
The reactions of PtF5, O2PtF6, PtF6, IrF6, NOPtF6, NO2PtF6, NOIrF6 and NO2IrF6 with NOF and NO2F have been examined under a variety of conditions. The relative ease of synthesis of NOPtF6, (NO)2PtF6, NOIrF6 and NO2IrF6 and the conversion of NOIrF6 to NOPtF6 with PtF6 confirms the order of strong oxidizing properties of PtVIF6, IrVIF6, PtVF-6 and IrVF-6. The NO+2 ion is intrinsically unstable with respect to the elimination of oxygen in Pt (IV) and Ir (IV) fluorometallate salts and accordingly there is serious doubt about earlier claims for the synthesis of the (NO+)(NO+2)PtF6 salt. No evidence for the (NO2)2PtF6 could be found.Raman spectral data for NOBF4, NO2BF4, NO2AsF6, NOPtF6, NO2PtF6, (NO)2PtF6, NOIrF6, NO2IrF6 and (NO)2IrF6 are presented and analyzed. The NO+2 ion appears to be linear in all of the compounds and the absence of the Raman forbidden ν2 fundamental indicates little if any anion-cation interaction at least of the type that leads to a permanent distortion of the cation. In the spectra of all of the nitryl salts, including NO2BF4, a low frequency band at about 140 cm-1 is clearly observed, the intensity and shape of which is a function of the anion. The band probably reflects an unknown lattice dynamic process. No such bands are evident in the spectra of NO+ and (NO+)2 salts.  相似文献   

14.
The amino acid l ‐phenylalanine has been cocrystallized with d ‐2‐aminobutyric acid, C9H11NO2·C4H9NO2, d ‐norvaline, C9H11NO2·C5H11NO2, and d ‐methionine, C9H11NO2·C5H11NO2S, with linear side chains, as well as with d ‐leucine, C9H11NO2·C6H13NO2, d ‐isoleucine, C9H11NO2·C6H13NO2, and d ‐allo‐isoleucine, C9H11NO2·C6H13NO2, with branched side chains. The structures of these 1:1 complexes fall into two classes based on the observed hydrogen‐bonding pattern. From a comparison with other l :d complexes involving hydrophobic amino acids and regular racemates, it is shown that the structure‐directing properties of phenylalanine closely parallel those of valine and isoleucine but not those of leucine, which shares side‐chain branching at Cγ with phenylalanine and is normally considered to be the most closely related non‐aromatic amino acid.  相似文献   

15.
We observe vibrationally excited NO2 from photodissociation of CH3NO2 and 2-C3H7NO2 by means of laser induced fluorescence. This approximate method shows very large vibrational excitation in all frequencies of NO2. The result is interpreted as an indirect predissociation.  相似文献   

16.
Two new Sillén–like layered lead fluoride nitrates, PbNaF2NO3 and Pb2OFNO3, have been prepared at 300 °C. PbNaF2NO3 is a structural analog of alkaline earth – bismuth oxyhalides, BaBiO2X and SrBiO2X (X = Cl – I), but not the corresponding nitrates, SrBiO2NO3 or BaBiO2NO3. Pb2OFNO3 is analogous to the corresponding halides, Pb2OFX (X = Cl, Br, I). Both structures belong to orthorhombic symmetry and demonstrate Na/Pb and O/F ordering, respectively. A continuous solid solution is formed between PbNaF2NO3 and Pb2OFNO3 which demonstrates neither cation nor anion ordering; the structure of intermediate composition Pb1.5Na0.5F1.5O0.5NO3 was refined in tetragonal symmetry and is indeed very close to that of PbBiO2NO3 and CaBiO2NO3. In PbNaF2NO3, the O:F ratio may be varied to a slight extent, PbNaF2-2yOyyNO3, which also breaks the Na – Pb cation ordering. Analogous fluoride halides could not be prepared. Structural analogies to lead, bismuth, and antimony oxyhalides are discussed.  相似文献   

17.
By codestillation of the appropriate mixtures we could prove that NOF or NO2F resp. and JF7 react to yield instable compounds NOF·JF7 and NO2·JF7 which exhibit IR-bands at 2260 and 2360 cm-1 resp. indicating NO+ and NO2+ ions and a band at 600 cm-1 characteristic for JF8-. Complete dissociation in the vapour phase is proven by IR-spectra. Vapour pressure data of NO+JF6- are given. A 1:1 mixture of NOJF8 and NOJF6 or NO2JF8 and NO2JF6 is obtained when NO and NO2 are combined with JF7.A new method for the estimation of the pressures of highly corrosive gases is described.  相似文献   

18.
The crystal packing of 1‐iodo‐3‐nitro­benzene, C6H4INO2, is formed by planar mol­ecules which are linked by I⋯I and NO2⋯NO2 interactions. In the case of 1‐iodo‐3,5‐di­nitro­benzene, C6H3IN2O4, the NO2 groups are not exactly coplanar with the benzene ring and the mol­ecules form sheets linked by NO2⋯NO2 interactions. In contrast with 4‐iodo­nitro­benzene, the crystal structures of both title compounds do not form highly symmetrical I⋯NO2 intermolecular interactions.  相似文献   

19.
To elucidate the effects of NO2 and H2O molecules on the surface conductivity of hydrogenated diamond film, models of various adsorbates containing different molecular ratio of NO2 and H2O on hydrogenated diamond (100) surfaces were constructed. The adsorption energies, equilibrium geometries of adsorbates, density of states, and atomic Mulliken populations were studied by using first‐principles method. The results showed that H2O molecule in the adsorbate could weaken the interactions between the adsorbates and hydrogenated diamond surface significantly. Compared with H2O molecule, NO2 molecule relaxes more dramatically when adsorbed on hydrogenated diamond surface. In addition, density of states for C(100):H–2NO2, C(100):H–NO2, and C(100):H–NO2 + H2O systems are very similar to each other, which indicates an obvious peak at valence band maximum level for all the three samples. It can be attributed to mainly single occupied molecule orbital of NO2 molecule and slightly C–H bond of C(100):H substrate. When the adsorbates contain one NO2 and two H2O molecules, the peak shifts slightly into valence band, but its intensity increases significantly. All the samples exhibit p‐type surface conductivity when adsorbed with pure NO2 molecules, and the surface conductivity remains as H2O molecules added into the NO2 adsorbate layer. However, for oxygenated diamond surface, very week interactions generate between diamond surface and various adsorbates. All the oxygenated diamond (100) surfaces with various adsorbates containing different NO2 and H2O molecules on it exhibit an insulating property.  相似文献   

20.
Electrical conductivities were measured for the ternary systems Y(NO3)3+La(NO3)3+H2O, La(NO3)3+Ce(NO3)3+H2O, La(NO3)3+Nd(NO3)3+H2O, and their binary subsystems Y(NO3)3+H2O, La(NO3)3+H2O, Ce(NO3)3+H2O, and Nd(NO3)3+H2O at (293.15, 298.15 and 308.15) K. The measured conductivities were used to test the generalized Young’s rule and the semi-ideal solution theory. The comparison results show that the generalized Young’s rule and the semi-ideal solution theory can yield good predictions for the conductivities of the ternary electrolyte solutions, implying that the conductivities of aqueous solutions of (1:3 + 1:3) electrolyte mixtures can be well predicted from those of their constituent binary solutions by the simple equations.  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