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 31 毫秒
1.
Sodium perfluoroalkanesulfinates [Cl (CF2)n SO2 Na (1), a , n = 4; b , n = 6; c , n = 8] with the reduction potentials about 0.95—1.00V could be oxidized readily with various oxidizing agents such as Mn (OAc)3 2H2O, Ce (SO4)2, HgSO4 and Co2O3 to generate perfluoroalkyl radicals which added to the olefins RCH ? CHR' to give two kinds of adducts, namely RCH (Rf) CHXR' (3, X ? H; 4, X ? OAc), with good yields depending upon the solvent system used. Different oxidizing agents showed slight variation on the yields of the adducts. The reaction time could be greatly shortened at higher temperature. Thus, this reaction provides a new way for introducing a perfluoroalkyl group into olefinic compounds.  相似文献   

2.
Perfluoroalkytin compounds R(4−n)Sn(Rf)n (R = Me, Et, Bu, Rf = C4F9, n = 1; R = Bu, Rf = C4F9, n = 2, 3; R = Bu, Rf = C6F13, n = 1) have been synthesized, characterized by 1H, 13C, 19F and 119Sn NMR, and evaluated as precursors for the atmospheric pressure chemical vapour deposition of fluorine‐doped SnO2 thin films. All precursors were sufficiently volatile in the range 84–136 °C and glass substrate temperatures of ca 550 °C to yield high‐quality films with ca 0.79–2.02% fluorine incorporation, save for Bu3SnC6F13, which incorporated <0.05% fluorine. Films were characterized by X‐ray diffraction, scanning electron microscopy, thickness, haze, emissivity, and sheet resistance. The fastest growth rates and highest quality films were obtained from Et3SnC4F9. An electron diffraction study of Me3SnC4F9 revealed four conformations, of which only the two of lowest abundance showed close F Sn contacts that could plausibly be associated with halogen transfer to tin, and in each case it was fluorine attached to either the γ‐ or δ‐carbon atoms of the Rf chain. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

3.
Polycrystalline alunite‐d6 KAl3(OD)6(SO4)2, prepared by hydrothermal reaction of Al2(SO4)3, K2SO4 and D2SO4, was studied by neutron powder diffraction performed on the diffractometer E2 (HMI‐BENSC, Berlin). Rietveld refinement of the data set for T = 2 K yielded the crystallographic data: space group R3m, Z = 3, trigonal setting, a = 694.3(1) pm, c = 1722.7(2) pm, N(I/σ(I) > 1) = 172, N(Var.) = 19, Rp = 0.036, wRp = 0.046, RB(I/σ(I) > 1) = 0.020. The deuterium nuclei could be located precisely. Three equivalent O–D bonds with nuclear distances r(O(4)–D) = 96.6(3) pm directed to each of the terminal oxygen atoms of the SO4 groups are found. Partial substitution of K+ by D3O+ was also considered in the refinement procedure. In good agreement with results of other methods a site occupation fraction n(D3O+) = 0.0104 was obtained.  相似文献   

4.
Apparent molar heat capacities C\textp,fC_{{\text{p}},\phi } and volumes VfV_\phi of aqueous trifluoromethanesulfonic (triflic) acid, HCF3SO3 (aq.) were determined with a Picker flow microcalorimeter and vibrating-tube densimeter at temperatures from 283 to 328 K and molalities from 0.05 to 9.5 mol-kg-1. Values of VfV_\phi and C\textp,fC_{{\text{p}},\phi } display a maximum near 0.8 mol-kg-1. VfV_\phi also displays a shallow minimum at ~5 mol-kg-1, while C\textp,fC_{{\text{p}},\phi } continues to decrease smoothly up to the limit of our measurements at 9.5 mol-kg-1. We attribute this behavior to ion–ion interaction between triflate and the hydrated proton to form the aqueous complex H2n+1On+ CF3SO3- (aq.), n = 5. Standard partial molar properties Cpo and Vo are consistent with results obtained from NaCF3SO3 (aq.) and yield values for the triflate anion CF3SO3- (aq.), over this range.  相似文献   

5.
Perfluoroalkyl iodide RfI [Rf = (CF2)nO(CF2)2SO2F, n = 2, (a); n = 4, (b); (CF2)4Cl, (c)] reacted with substituted benzene C6H5Y (Y = alkyl, OCH3, CF3, F, Cl, Br, I) in the presence of copper in acetic anhydride to give the corresponding mixture of isomeric disubstituted benzene (RfC6H4Y). The conversion and yield depend on both the amount of copper used and nature of substituent. The likely explanation is that the reaction may involve a free radical process. The perfluoroalkyl radical can be trapped by cyclohexene, isopropylbenzene and styrene. Using DMSO in place of acetic anhydride as a solvent the reaction takes a different course, it is believed that the reaction in DMSO proceeds through a perfluoroalkylcopper intermediate.  相似文献   

6.
The introduction of strong electron-accepting fluorine-containing substituents into the aromatic moiety gives compounds with unique properties. Determination of the electronic nature of the grouping (Rf)2PO in arylbis(perfluoroalkyl)-phosphine oxides has shown that this substituent is comparable with RfSO2, one of the most electron-accepting groups.A general principle is proposed for the construction of the new superstrong electron-accepting substituents by the replacement of oxygen atoms for trifluoromethylsulfonylimino groups. For example, when the oxygen atoms in CF3SO and CF3SO2 groups are replaced by CF3SO2N=, new stable and even more electron-accepting substituents are formed. The grouping CF3S(O)=NSO2CF3 corresponds to two nitro groups. A similar increase in electron-accepting ability is observed in groupings derived from other elements by replacement of oxygen atoms by the CF3SO2N= group.  相似文献   

7.
Contact with SO2 causes almost immediate dissolution of tetraalkylammonium halides, R4NX, (R = CH3 (Me), X = I; R = C2H5 (Et), X = Cl, Br, I; R = C4H9 (nBu), X = Cl, Br), with the formation of an adduct, [R4N]+[(SO2)nX] (n = 1–4). Vapor pressure measurements indicate the proclivity for SO2 uptake follows the order N(CH3)4+ < N(C2H5)4+ < N(C4H9)4+. This trend is in accord with the Jenkins–Passmore volume‐based thermodynamic model. Born–Haber cycles, incorporating the lattice energy and gas phase energy terms, are used to evaluate the energetic feasibility of reactions. Density functional theory calculations (B3PW91; 6‐311+G(3df)) have been used to calculate the energetics of (SO2)nX (X = Cl and Br) anions in the gas phase. The experimental studies show that tetraalkylammonium halides are feasible sorbents for SO2. In order to correlate the theoretical model, experimental enthalpy, Δr and entropy, Δr changes have been determined by the van't Hoff method for the binding of one SO2 molecule to (C2H5)4NCl, resulting in the liquid adduct (C2H5)4NCl · SO2. The structure of the analogous 1:1 bromide adduct, (C2H5)4NBr · SO2, has been determined by single‐crystal X‐ray diffraction (monoclinic, P21/c, a = 9.1409(14) Å, b = 12.3790(19) Å, c = 11.3851(17) Å, β = 107.952(2)°, V = 1225.6(3) Å3). The structure consists of discrete alkylammonium cations, bromide anions and SO2 molecules with short contacts between the anion and SO2 molecules. The (C2H5)4N+ cationadopts a transoid conformation with D2d symmetry, and represents a rare example of a well‐ordered (C2H5)4N+ cation in a crystal structure. The Br anions and SO2 molecules forms a chain, (SO2Br)n, with bifurcated contacts. Non‐bonding electron pairs on the halide anions engage in electrostatic interactions with the sulfur atoms and charge‐transfer interactions with the antibonding S–O orbitals of the bound SO2 moiety. Raman and 17O NMR spectra provide compelling evidence for a charge‐transfer interaction between SO2 molecules and the halide ions.  相似文献   

8.
Shizheng Zhu  Ping He 《Tetrahedron》2005,61(23):5679-5685
The thermal decomposition reactions of fluoroalkanesulfonyl azides RfSO2N3 (1) in the presence of various substituted benzene XnC6H6−n [X: CH3 (n=1, 2, 4, 6), OCH3 (n=1, 2), C6H5CH2 (n=1), F, Cl, Br] were studied in detail. The N-aryl fluoroalkanesulfonyl amides [RfSO2NHC6H5−nXn] were produced as the major products. The ortho/para ratio resembled that of an electrophilic aromatic substituted reaction. An ionic π- or σ-complex was postulated as the intermediate for these reactions.  相似文献   

9.
Sol–gel reactions of fluoroalkyl end‐capped trimethoxyvinylsilane oligomer in the presence of low molecular weight aromatic compounds (ArH) such as 1,1′‐bi(2‐naphthol) (BINOL) and 2‐hydroxy‐4‐methoxy benzophenone (HMB) were found to proceed smoothly under alkaline conditions at room temperature to give the corresponding fluorinated oligomeric silica nanocomposites‐encapsulated aromatic compounds (BINOL and HMB) [RF‐(VM‐SiO2)n‐RF/ArH nanocomposites]. UV light irradiation (λmax: 254 nm) toward RF‐(VM‐SiO2)n‐RF/ArH nanocomposites showed that photodegradation of encapsulated ArH was not observed at all in the fluorinated nanocomposites cores, although the parent ArH can exhibit an effective photodegradation behavior under similar conditions. Especially, encapsulated ArH can exhibit no weight loss corresponding to the contents of the aromatic compounds in the fluorinated nanocomposites even after calcination at 800°C. Therefore, fluoroalkyl end‐capped trimethoxyvinylsilane oligomer has high potential for not only the thermal resistance but also the UV resistance fluorinated polymeric materials. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

10.
The thermal reactions of fluoroalkanesulfonyl azides RfCF2SO2N31 with nitrobenzene and its derivatives XC6H4NO2 (X=H, F, Cl, CF3) gave the unexpected N-fluoroalkaneacyl anilides RfCONHC6H4X (X=H, Cl, F, CF3) in addition to fluoroalkanesulfonyl amides RfCF2SO2NH2. Under the same reaction conditions, however, nitrobenzene containing an electron-donating group RC6H4NO2 (R=CH3, OCH3) reacted with 1 affording the corresponding N-fluoroalkanesulfonyl anilides RfCF2SO2NHC6H3(NO2)R. Other electron-poor benzene derivatives, such as benzaldehyde, benzoate, and acetophenone C6H5Y(Y=CHO, COCH3, CO2CH3) all gave the meta-substituted N-fluoroalkanesulfonyl anilides RfCF2SO2NHC6H4Y.  相似文献   

11.
Summary The nickel complexes of Schiff bases formed fromo-aminothiophenol and -dicarbonyl compounds, Ni(o-SC6H4N=CRCR=NC6H4S-o) (R=H, Me, Ph) (1a-c), catalyse the reduction of aromatic nitro compounds by NaBH4. The reduced species [Ni(o-SC6H4N=CHCH=NC6H4S-o)] (2) and [Ni(o-SC6H4NHCH2CH=NC6H4S-o)] (3) were identified as intermediates in the catalytic cycle.  相似文献   

12.
The temperature dependence of the heat capacity C p o = f(T) of palladium oxide PdO(cr.) was studied for the first time in an adiabatic vacuum calorimeter in the range of 6.48–328.86 K. Standard thermodynamic functions C p o(T), H o(T) — H o(0), S o(T), and G o(T) — H o(0) in the range of T → 0 to 330 K (key quantities in different thermodynamic calculations with the participation of palladium compounds) were calculated on the basis of the experimental data. Based on an analysis of studies on determining the thermodynamic properties of PdO(cr.), the following values of absolute entropy, standard enthalpy, and Gibbs function of the formation of palladium oxide are recommended: S o(298.15) = 39.58 ± 0.15 J/(K mol), Δf H o(298.15) = −112.69 ± 0.32 kJ/mol, Δf G o(298.15) = −82.68 ± 0.35 kJ/mol. The stability of Pd(OH)2 (amorph.) with respect to PdO(cr.) was estimated.  相似文献   

13.
Summary Aminoalkanesulphonic acids H2N(CH2) n SO3H, (n = 1, 2 or 3) react with phosphonium salts [R2P(CH2OH)2]Cl (R = Ph or Cy, Cy = cyclohexyl) in the presence of Et3N to give the sulphonated aminomethylphosphines [Et3NH] [(R2PCH2)2N(CH2) n SO3] (R = Ph, n = 1, 2 or 3; R = Cy, n = 1). The single crystal X-ray structure of [Et3NH] [(Ph2PCH2)2N(CH2)2SO3] has been determined. Some NiII, PdII, PtII and RhI complexes of the phosphines have been prepared.  相似文献   

14.
The new 22-π, aromatic “pentaplanar” macrocycle, ozaphyrin ( 6 ), has been synthesized by a McMurry coupling of 5,5′-diformyl-4,4′-dipropyl-2,2′-bipyrrole ( 1 ) with 2,5-bis(5-formyl-4-propyl-2-pyrrolyl)furan ( 5 ). This synthetic pathway to ozaphyrin and its characterization by 1H nmr spectroscopy, uv-visible spectroscopy, cyclic voltammetry, and X-ray crystallography are described. The structure consists of layers of planar, staggered macrocycles stacked perpendicular to the α-axis. Ozaphyrin crystallizes with four formula units in the monoclinic space group C52h-P21/n in a cell of dimensions a = 10.481(7) Å, b = 17.353(17) Å, c = 18.726(12) Å, and β = 102.84(5)° (108 K). The structure has been refined on F2 (5171 unique reflections, 411 variables) to Rw(Fo2) = 0.165. The conventional agreement index R(F) is 0.074 for the 3289 reflections have Fo2>2o(Fo2).  相似文献   

15.
Two novel organic amine templated lanthanide sulfates, layer (H3DETA)[Nd(H2O)(SO4)3] (I) and chain-like (H3DETA)[Ho(H2O)2(SO4)3] (II), are hydrothermally synthesized by using diethylenetriamine (DETA) as template, and are structurally characterized by ICP, elemental analysis, X-ray single-crystal diffraction, X-ray powder diffraction, IR, and TG. Compound I is monoclinic with space group P21 and data of unit cell: a = 6.6518(13), b = 10.373(2), c = 11.091(2) ?, β = 93.61(3)°, V = 763.7(3) ?3, ρ c = 2.421 g/cm3, μ(MoK α) = 3.885 mm−1, Z = 2, R 1 = 0.0194 for 3312 reflections with F o > 2σ(F o). The Nd ions are nine-coordinated by one oxygen atom of coordinated water and eight oxygen atoms of sulfates. Compound I displays the layer structure constructed by linking Nd ions with three-linkage SO4 tetrahedra as the bridge (affording one coordinated vertex and coordinated edge). Compound II crystallizes in the monoclinic space group P21/c with unit cell data: a = 6.594(13) ?, b = 14.783(3) ?, c = 16.599(3) ?, β = 93.47(3)°, V = 1614.2(6) ?3, ρ c = 2.454 g/cm3, μ(MoK α) = 5.37 mm−1, Z = 4, R 1 = 0.0259 for 1815 reflections with F o > 2σ(F o). The Ho ions are eight-coordinated by two oxygen atoms of coordinated water and six oxygen atoms of sulfates. The straight chain-like structure of II is attained by alternatively arranging HoO8 polyhedra and two-linkage SO4 tetrahedra (affording two coordinated vertices). The TG results indicate that the two compounds experience three weight losses and lead to distinct residues: Nd2O3 for I and HoO(SO4)0.5 for II.  相似文献   

16.
Perfluoroorganoantimony(III) (Rf)nSbCl3? n and -antimony(V) (Rf)nSbCl5? n (n = 1,2,3, Rf = C6F5) compounds are screened for the first time for biological activity. The compounds exhibited significant in vitro antitumor activity against MCF-7 (human breast cancer) cell line and antibacterial activity against three pathogenic bacteria: Pseudomonas aeruginosa, Staphylococcus aureus, and Klebsiella pneumoniae. They also showed antifungal activity against Aspergillus flavus and Aspergillus niger as well as insecticidal activity against cockroach (Periplanata americana), housefly (Musca domestica), Tobacco caterpillar (Spodoptera litura), and spider mite (Tetranychus urticae). These studies suggest a better biocidal activity for the pentafluorophenyl antimony halides compared to the corresponding phenyl analogues.  相似文献   

17.
The enthalpies of formation (ΔH f o) for 24 hydrocarbon radicals (R?), mainly polycyclic aromatic radicals with the complex structure, were determined from the published data on bond dissociation energies. The ΔH f o values of the corresponding molecules were calculated, in the majority of cases, by the macroincrement method. Calculations by the group contribution method were performed. Some ΔH f o(R?) values were compared to those calculated by the additive-group method. Calculations were performed, and the conjugation energies of the radicals were discussed. The errors of determination of the ΔH f o(R?) values found were estimated. Due to this work, the database for ΔH f o values of hydrocarbon radicals was increased more than by 25%.  相似文献   

18.
Poly(benzyl ether) dendrimers with o-, m-, and p-isomers of dialkoxybenzene at their focal points [o-, m-, and p-(Gn)2Ar], having generation numbers (n) of 0–3, were synthesized. 1H NMR pulse relaxation times (T1) of the exterior MeO groups of o- and m-(Gn)2Ar (n = 0–3) all remained in the range of 0.92–1.43 s. In sharp contrast, an exceptionally short T1 value (0.23 s) was observed for p-(G3)2Ar. Although their absorption spectral profiles were slightly different from one another, an essential difference was observed for their fluorescence properties. When the generation number was increased, the fluorescence efficiency of o-(Gn)2Ar increased, but that of p-(Gn)2Ar decreased, whereas m-(Gn)2Ar exhibited a relatively small change in the fluorescence efficiency. Fluorescence depolarization studies showed a highly efficient intramolecular energy migration in p-(G3)2Ar as compared with o-(G3)2Ar and m-(G3)2Ar. © 2003 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 41: 3524–3530, 2003  相似文献   

19.
Four novel diorganotin(IV) complexes with general formula R2SnL (R = nBu, PhCH2) were synthesized from diorganotin dichlorides and binary Schiff‐bases (H2L) containing N2O2 donor atoms in the presence of sodium ethoxide. The Schiff bases were prepared by reactions of o‐phenylenediamine with 3‐tert‐butyl‐2‐hydroxy‐5‐methylbenzaldehyde (H2L1) and salicylaldehyde (H2L2) respectively. The compounds were characterized by elemental analyses, IR, and NMR spectroscopy. The solid‐state crystal structure of the compound nBu2SnL1 was determined by single‐crystal structural analysis.  相似文献   

20.
全氟烷基磺酰亚胺盐催化芳香化合物硝化反应的研究   总被引:8,自引:0,他引:8  
利用系列全氟烷基磺酰亚胺盐[M(NPf2)n]作为一种新型的Lewis酸催化剂,用于催化芳香化合物与等摩尔65% (mm)硝酸的硝化反应. 通过考察不同的催化剂、反应时间、反应温度和反应介质效应等因素对甲苯硝化的影响,以及比较1 mol% Yb[N(C4F9SO2)2]3催化不同结构的取代芳烃硝化反应的效果,表明全氟烷基磺酰亚胺盐不仅具有环境友好和原子经济的特点,而且是一类比常规Lewis酸更有效的、芳香化合物硝化反应的催化剂.  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