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 891 毫秒
1.
The perfluoronitrosocycloalkanes, heptafluoronitrosocyclobutane and nonafluoronitrosocyclopentane, are convenient precursors to a family of new perfluorocycloalkyl(aryl) diazenes. With aniline and o-aminobenzamide, CF2(CF2)xCFNNC(CH)4CH and CF2(CF2)xCFNNC(CH)4CC(O)NH2 (x = 2,3) are formed. Additionally, heptafluoronitrosocyclobutane gives CF2(CF2)2CFNNCCFCFCHCFCF and CF2(CF2)2CFNNC(CH)4CNH2 with 2,3,5,6-tetrafluoroaniline and o-phenylenediamine  相似文献   

2.
In this paper, we present dielectric results for samples of POE (polyethyleneoxyde) divided into three classes. Compounds of low molecular weight (Mw < 1000) show very weak absorption at lower temperatures. At higher temperatures (below the melting point), there appears a very important peak (stronger for lower Mw) which corresponds to the supercooled phase. Compounds of intermediate molecular weight (1000 ≤ Mw ≤ 105) show a β relaxation near 250 K (bisMw ? 3000) which is due to the movement of chain-ends in the amorphous phase. This process increases with the importance of that part of the structure formed of shorter lamellae, thus explaining the marked diminution of this absorption for higher molecular weight compounds (the shorter lamellae cannot be obtained for Mw > 4000). The α relaxation (obtained for T > 300 K) is perturbed by an important conduction; however, we have found a peak and a shoulder at lower frequencies, due to shorter and longer lamellae respectively. The shoulder is also present for high molecular weight compounds (Mw ≥ 0.1–106). At lower temperatures, they display a very broad γ absorption, superposed at higher frequencies with a β relaxation. The nature of this β relaxation is different from that observed for the intermediate molecular weight compounds.  相似文献   

3.
The compound Th0.25 NbO3 melts congruently at 1390°C. Single crystals obtained by slow cooling from the melt are transparent and show uniaxial optical properties. A single-crystal X-ray analysis confirms the tetragonal cell found by Kovba and Trunov from a powder data and gives a = 3.90 Å and c = 7.85 Å. No systematic absence of the hkl reflections is observed on precession films. The relative intensities of the main reflections are characteristic of a perovskite-like arrangement ABO3 whose large dodecahedral A sites are only partly occupied. Several domains have been found in the perovskite-type solid solution (1 ? x) Th0.25NbO3-x NaNbO3. For 0 ? x ? 0.5 the phases have a tetragonal cell with a ? a0 and c ? 2a0 as in pure Th0.25 NbO3. When 0.6 ? x ? 0.8 the corresponding phases crystallize with a small cubic cell (a0 ? 3.9Å), while phases with 0.9 ? x ? 1 have an orthorhombic cell (a ? 212a0, b ? 212a0, c ? a0).  相似文献   

4.
Copolymers of methylmethacrylate (MMA) with phenylvinylketone (PVK) and of styrene (St) with PVK were degraded in solution by irradiation with 25 nsec flashes of 347.1 nm light from a ruby laser. The time dependence of the change of light-scattering intensity (LSI) (due to the decrease of molecular weight) was measured. It was found that the half-life τ12 (LSI) decreases with increasing polymer concentration according to [(τ12(LSI)]?1 ∝ C Polymer and that τ12 (LSI) is proportional to the microviscosity (viscosity of the solvent). This was taken as evidence that τ12 (LSI) corresponds to a diffusion process. τ12 (LSI) decreased with increasing scattering angle, explained by the fact that at low scattering angles the diffusion of long fragments was predominantly monitored.Upon addition of naphthalene to an acetone solution of a copolymer of MMA and PVK, τ12 (LSI) increased with increasing naphthalene concentration and approached a limiting value. This effect was attributed to preferential solvation of the copolymer by naphthalene. From intrinsic viscosity measurements, it could be confirmed that MMA base units are preferentially solvated by an acetone/naphthalene mixture.  相似文献   

5.
Gel permeation chromatography (GPC) and viscometry (V) methods have been combined for determination of long-chain branching in bisphenol-A polycarbonate (PC) by means of a branching factor gv = Mvg1/Mv1, where Mvg1 and Mv1 are the apparent viscosity-average molecular weights calculated from GPC data and from intrinsic viscosities [η] of the samples respectively. A linear dependence of gv on molar % of branching agent has been obtained. The GPC data on PC samples have also been applied for verification of an earlier [η]?M relationship for branched polydisperse polymers.  相似文献   

6.
7.
J. Duflos  G. Queguiner 《Tetrahedron》1985,41(16):3303-3311
The cycloaddition reactions of 6-oxo (2H) cyclohepta (c) pyrroles 4a, b and c with diethyl acetylene dicarboxylate an electron-deficient dienophile of relatively low LUMO energy level, afford the 1:2 adducts 5a, b1-b2 and c respectively The kinetics of the reactions, investigated By 1H NMR spectroscopy, have allowed to precise the mechanism of the addition In particular, new-intermediates, 1:2' adducts 6b and c have been observed and isolated for the compound 6c.  相似文献   

8.
9.
Osmotic pressure measurements on polystyrene (Mn = 396. 000) in trans-decalin for concentrations up to 140 kg m?3 and from 20 to 40 are reported. The θ-temperature is 20.8 . The ratio
?(πc)?cc=0?(πc)?cc=c+
where c+ is the concentration at which a homogeneous segment distribution is assumed to prevail, increases with temperature up to the plateau value of 0.7. From the temperature dependence of the osmotic pressure, the partial molar enthalpy. Δh1, and entropy, Δs1. of mixing are found to be positive. The solvent-solute interaction parameter increases linearly with concentration at all temperatures.  相似文献   

10.
Intensity parameters of Sm3+ in borate glasses were obtained by fitting the oscillator strengths to the Judd-Ofelt formula and a study of energy transfer from gadolinium to samarium was performed. An increase of samarium fluorescence originating from the 4G52 level was observed in the presence of gadolinium, in the concentration range of 0.1–3 wt% samarium with gadolinium constant at 3 wt%. The intensity of samarium fluorescence on excitation at 273 nm increased by one order of magnitude in the presence of gadolinium. From the excitation spectrum of the double-doped glasses (Gd + Sm), it was deduced that energy absorbed by gadolinium is transferred from 6P72 gadolinium levels to the 4P32 and 4P52 samarium levels.The mechanism of this energy transfer was obtained by plotting the energy transfer probabilities as a function of samarium concentration. A linear dependence of η0η (η intensity of gadolinium in the presence of samarium) versus square of concentration of Sm + Gd is obtained. From this it is concluded that the transfer is of electric-multipolar type, mainly dipole-dipole. A small increase (about 10%) of fluorescence of samarium in the presence of gadolinium excited at levels where no energy transfer can take place is attributed to the fact that the quenching of samarium occurring by the cross relaxation (4G526F92) (6H526F92) is suppressed by the presence of gadolinium as seen from concentration dependence of samarium doped glasses compared to double-doped glasses.  相似文献   

11.
Variable temperature 1H NMR spectroscopy has been used in the study of 1,3-intramolecular shifts of the M(CO)5 moiety in complexes of the general formula [M(CO)5L], (M = Cr or w), L = SCH2SCH2SCH2, SCH2SCH2CH2CH2 and SCH(Me)SCH2CH2CH2. For the 1,3,5-trithian complexes precise energy barriers for the process have been obtained by detailed computer simulation of the static and dynamic spectra. Our results suggest that the magnitude of ΔG (298.15 K) for the 1,3-shift is largely dependent upon the skeletal flexibility of the ligand system. In this context we have investigated the X-ray crystal structure of the highly substituted trithian complex [W(CO)5{β-SCH(Me)SCH(Me)SCH(Me)}].  相似文献   

12.
It is well known that the apparent specific volume η2 of a polymer may be expressed by the following relationship: η2= ηm + K/Mn where M?n is the number-average molecular weight of the polymer, ηm the specific volume of the infinite polymer, and K a constant. We have shown that this relationship is valid for low molecular weight polystyrenes (Mn < 4·104) with different end-groups, independently of the nature of the solvent. The K values (and variations with the solvent and with the nature of the end-groups) may be predicted through simple calculations proposed here. We conclude that ηm does not represent the specific volume of the infinite polymer, since we observe a rapid decrease of η2 for increasing M (when Mn < 4·104). The decrease is much greater than expected from the relationship η2 = ? (1/M).  相似文献   

13.
14.
The lanthanum β-alumina phase doped with europium was investigated by X-ray diffraction and fluorescence. This nonstoichiometric phase exists over the composition range: 11Al2O31La2O3 to 14Al2O31La2O3. The unit cell is hexagonal hexagonal with a = 5.560 ± 0.003 Å, c = 22.001 ± 0.003 Å and belongs to the P63mmc space group. X-ray diffraction patterns do not vary between both boundary compositions, but fluorescence spectra show that the structure of the mirror plane in which the lanthanide ions are located is deeply modified. The atomic structure of the mirror plane is of “β-type” (like β(Na) or β(Ag)) for the lower alumina contents; it gradually changes to a “magnetoplumbite type” for higher alumina contents.  相似文献   

15.
The compounds Ba4Fe2S6[S23(S2)13] and Ba3.6Al0.4Fe2S6[S0.6(S2)0.4], designated I and II, were prepared by reacting BaS, Fe, and S powders and Al foils in graphite containers sealed in evacuated quartz ampoules at approximately 1100°C. The crystal structure of I was determined using 1682 independent, nonzero X-ray reflections, while 3589 were used for II. They are triclinic, Al:
a=9.002(2)A?,b=6.7086(8)A?,c=24.658(4)A?α91.49(2)°,
β=105.10(2)°y=90.74(2)°,ψcalc=4.15g/cm3,for I:
a=8.993(6)A?,b=6.708(7)A?,c=24.70(1)A?α91.11(6)°,
β=105.04(6)°y=90.90(9)°,ψcalc=3.90g/cm3,for II:
BaS6 trigonal prisms share edges to form distorted hexagonal rings which form one-dimensional chains leaving two free lateral edges. The chains link in a stairstep manner with the rings offset along the [301] direction. These stairsteps join in a complicated manner to form a three-dimensional network. Fe ions are in two sites forming isolated FeS4 tetrahedra and isolated Fe2S6 dimers by edge-sharing tetrahedra. The Al substitution occurs in the trigonal prisms which have free edges with Al replacing Ba. Room-temperature Mössbauer isomer shifts are 0.20 mm/sec. for I and 0.30 mm/sec for II. These data indicate that upon Al substitution charge compensation occurs by reducing Fe3+. Valence calculations indicate that Fe in edge-sharing tetrahedra are reduced while the Fe in the isolated tetrahedron remains unchanged. The effective charge distribution in the Al substituted compound is approximately Fe3+, Fe2.5+ with electron delocalization across the shared edge. Room temperature electrical resistivity is 105 ohm/cm. The compositions of the crystals are best represented by the formulas [Ba4Fe2S7]23·[Ba4Fe2S6(S2)]13 and [Ba3AlFe2S7]0.4·[Ba4Fe2S7]0.2·[Ba4Fe2S6(S2)]0.4. The replacement of a sulfide by a disulfide ion is thought to be strongly dependent on the sulfur activity during the preparation.  相似文献   

16.
Unstable transition metal compounds formed from hydridosilacyclobutanes are described: 1-methyl-1-silacyclobutane reacts with nonacarbonyldiiron to give the complexes [Fe(CO)4(H){Si(Me)CH2CH2CH2}] and [Fe{CH2CH2CH2Si(H)Me}(CO)4], and with bis(triphenylphosphine)(ethylene)platinum(0) to give [Pt(H)(PPh3)2{Si(Me)CH2CH2CH2}].  相似文献   

17.
trans-7α-carbomethoxy-decal-1-one (2) yields a mixture of the two oxy-esters 6 and 7 on reacting with MeMgX. Ratios 67 were measured for reactions performed in benzene (with X = I) and in THF (with X = Cl). The small variations of the ratios 67 as compared to those obtained in analogous experiments performed with methyl (2-oxo-cyclohexyl)-propionate 4 and methyl 4-methyl-5-oxo-hexanoate 5 suggest that conformational mobility plays a fundamental role in determining the variations of stereospecificity with varying the reaction conditions. Competitive Grignard reactions among 2,4 and 5 show that their reactivities are in the order 4>2>5 (K4k2 = 1.7; k5k2=0.8) when reactions are performed in benzene with X  I and 2>4>5 (k4k2= 0.56; k5k2=0.25) when reactions are performed in THF with X  C1. The experimental data are interpreted in terms of anchimeric assistance given by the ester group to the reactions of the keto group in conformationally mobile δ-keto esters. The occurrence of this effect depends on the reaction conditions which can favour, or not, folded transition states.  相似文献   

18.
19.
20.
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