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.
On the RbNiCrF6 Type. V. New Fluorides of the Type CsBMF6 with B = MnII or NiII and M = Ga, Fe, Rh or Sc, In, Tl, Rh New prepared are the cubic compounds CsNiScF6 (light yellow, a = 10.60 Å); CsNiInF6 (lightyellow, a = 10.64 Å); CsNiTlF6 (light yellow, a = 10.60 Å); CsNiRhF6 (light redbrown, a = 10.37 Å); CsMnGaF6 (pink, a = 10.42 Å); CsMnFeF6 (light green, a = 10.55 Å) and CsMnRhF6 (redbrown, a = 10.58 Å), all RbNiCrF6 type of structure. The Madelung part of lattice energy, MAPLE, is calculated and discussed.  相似文献   

2.
The anhydrous iodates of Cr, Mn, Fe, Co and β-Ni form a single isomorphous family, crystallizing in space group P63 or P6322 with lattice constants typified by Mn(IO3)2 of a = 11.178 ± 0.002, c = 5.035 ± 0.001 Å and four formulas per unit cell; in addition, α-Ni(IO3)2 forms as a second phase. Mn(IO3)2, Fe(IO3)3, α-Ni(IO3)2 and β-Ni(IO3)2 order antiferromagnetically at 6.5, 17.0, 3.5 and 5.0 K, respectively; Cr(IO3)3 and Co(IO3)2 remain paramagnetic to 1.5 K. Below ΘN, a weak ferromagnetic moment develops in the Mn, α-Ni and β-Ni iodates. All the anhydrous iodates generate second harmonics. Co(IO3)2 · 4H2O and β-Ni(IO3)2 · 4H2O crystallize isomorphously in space group P21c, with lattice constants a = 8.370 ± 0.005, b = 6.572 ± 0.007, c = 8.514 ± 0.008 Å, β = 99.8 ± 0.1° for the Co compound. Co(IO3)2 · 2H2O is triclinic, with a = 6.666 ± 0.015, b = 10.991 ± 0.025, c = 4.913 ± 0.011 Å, α = 93.1 ± 0.1, β = 92.1 ± 0.1, γ = 98.9 ± 0.1°, space group P1, and Ni(IO3)2 · 2H2O is orthorhombic, a = 9.14986 ± 0.00008, b = 12.20896 ± 0.00022, c = 6.58353 ± 0.00013Å at 298 K, space group Pbca. The Co iodate hydrates are paramagnetic to 1.5 K; both Ni hydrates are antiferromagnetic, the dihydrate also developing a weak ferromagnetic moment. The lattice spacings of all 11 compounds are presented, 9 with indexing.  相似文献   

3.
Single crystals of CuV2S4 have been made by chemical vapor transport with chlorine as the transport agent. Characterization by an X-ray study, density measurements by the hydrostatic technique and electron microprobe analysis were performed. They showed no deviation from stoichiometry of the compound. Refinement of the structure (space group Fd3m) has been done. Transport properties, magnetic susceptibility and N.M.R. data from 1.5°K to 300°K present discontinuities below 100°K. Observed properties are discussed in view of the previously proposed band model. The relationships between magnetic susceptibility and Knight shifts are included in this paper.  相似文献   

4.
The rate of polymerization of thiophene, at concentrations of catalyst (SnCl4), and thiophene of the same order as was subsequently used in studying the reaction between thiophene and di(chloromethyl)benzene, is of the order of 10-2%/hr at 30°C. There is no significant self-condensation of DCMB under the same conditions. Since the reaction between thiophene and DCMB is complete at 30°C in minutes rather than hours, it is assumed that self-condensation of thiophene or DCMB during the reaction between them will be negligible and should not influence the course of the reaction or the structure of the resulting polymer. Reaction at 30°C is much too fast for convenient study. A temperature of 0°C is more appropriate and was used in subsequent kinetic work. The first two products of the condensation of p-di(chloromethyl)benzene (DCMB) with thiophene have been identified by a combination of mass, infrared, and nuclear magnetic resonance spectroscopy as thenylchloromethylbenzene (TCMB) and dithenylbenzene (DTB). DCMB, TCMB, and DTB have been estimated quantitatively during the course of the reaction by gas-liquid chromatography (GLC), and it has been established that the rates of each of the two reaction steps is first-order with respect to the chloro compound (DCMB and TCMB respectively), thiophene, and SnCl4. Rate constants for these two consecutive reactions were calculated to be k1 = 2.79 × 10-4l.2/mole2-sec, k2 = 6.37 × 10-3l.2/mole2-sec; the corresponding energies of activation are E1 = 7.93 kcal/mole, E2 = 7°67 kcal/mole. These rate constants are appreciably higher than values previously obtained for the corresponding DCMB–benzene reactions.  相似文献   

5.
The structure of H3Os3(CO)9CCH3 has been determined by a combination of nematic-phase PMR and X-ray powder photography; the compound is iso-structural with the analogous H3Ru3(CO)9CCH3, with an osmium to (bridging) hydride proton distance of 1.82 Å and an OsHOs angle of 103°.  相似文献   

6.
Vacuum line kinetic studies of the reaction of p-toluenesulfonyl chloride and benzene or toluene, using aluminum chloride as the catalyst and dichloromethane as the solvent were determined at 25°C by means of gas chromatography. The reaction is first-order in arene, tosyl chloride, and in AlCl3 as catalyst. Noncompetitive results are kT/kB=22±7 with a product sulfone isomer distribution: ortho, 14±1%; meta, 4.3±0.2%; and para 82±1%. With hexadeuteriobenzene kH/kD was determined to be 1.8±0.1. Rate constant ratios and product isomer distributions were also determined competitively: with AlCl3, kT/kB=30±2; % ortho, 13±1; % meta, 4.0±0.5; % para, 84±3; with SbCl5, kT/kB=40±4; % ortho 10.3±0.4; % meta, 4.7±0.2; and % para, 85.0±0.5. The kT/kB ratio for AlCl3 and the meta sulfone product percentages for both AlCl3 and SbCl5 are considerably higher than those reported in the literature. NMR and Raman studies suggest a molecular complex between p-tosyl chloride and AlCl3, with coordination through oxygen as the dominant species and the probable electrophile in CH2Cl2. A reaction mechanism consistent with the kinetic and spectroscopic results is proposed. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet 30: 367–372,1998  相似文献   

7.
Copoly(4,4′-oxanilideterephthalamide—4,4′-phenyleneterephthalamide) (A-202/PPD) was synthesized by reaction of 4,4′-diaminooxanilide, p-phenylenediamine, and terephthaloyl chloride in organic solvents. Copolymer inherent viscosities in H2SO4 as high as 10.3 were obtained. Isotropic copolymer solutions (4%—5% concentration) of A-202/40%–80% PPD were spun to fibers with tenacity/elongation/modulus at 1% extension in the 13–14 gpd/1.5%–2%/700–1000 gpd range. Oxamide and amide stabilities in 98–100% H2SO4 and 20% oleum were compared. Poly(4,4′-oxanilideterephthalamide) (A-202), A-202/PPD copolymers, and poly(4,4′-phenyleneterephthalamide) (PPT) were unstable in 20% oleum, but all proved relatively stable in 100% H2SO4. However, the oxamide linkage proved less stable than the amide linkage in 98% H2SO4. A-202 and A-202/PPD copolymers formed stable anisotropic spinning solutions in 1% oleum at 10–20% concentrations. Dynamic mechanical analyses (Vibron) showed no glass transition temperature (Tg) below 200°C. Dilatometric measurement of A-202/50% PPD revealed a Tg at 257°C. Differential thermal analyses of A-202/40–80% PPD exhibited endotherms at 470–480°C. Thermogravimetric analyses showed no significant weight loss below 400°C.  相似文献   

8.
Three cobalt model molecular compounds, Co‐cubane ([Co43‐O)4(µ‐OAc)4py4]), Co‐trimer ([Co33‐O)(µ‐OAc)6py3]PF6), and Co‐dimer ([Co2(μ‐OH)2(µ‐OAc)(OAc)2py4]PF6), are investigated as water oxidation reaction (WOR) catalysts, using electrochemical, photochemical, and photoelectrochemical methodologies in phosphate electrolyte. The actual species contributing to the catalytic activity observed in the WOR are derived from the transformation of these cobalt compounds. The catalytic activity observed is highly dependent on the initial compound structure and on the particular WOR methodology used. Co‐cubane shows no activity in the electrochemical WOR and negligible activity in the photochemical WOR, but is active in the photoelectrochemical WOR, in which it behaves as a precursor to catalytically active species. Co‐dimer also shows no activity in the electrochemical WOR, but behaves as a precursor to catalytically active species in both the photochemical and photoelectrochemical WOR experiments. Co‐trimer behaves as a precursor to catalytically active species in all three of the WOR methodologies.  相似文献   

9.
The apparent molal volume φv, expansibility φE, compressibility φK, and heat capacity φc of NaCl were measured in urea-water mixtures, as a function of salt (<1.5m) and urea (<13m) concentrations at 25°C. At a fixed urea concentration, the transfer functions from H2O to 3m urea are linear functions of the NaCl aquamolality. At a fixed salt aquamolality, (0.1m), the sign of the transfer functions is in the direction of a decrease in the structure-breaking effect, and the absolute values of the transfer functions tend to level off at high urea concentrations (13m). The functions φv, φE, φK, φc, and (?φv/?T)p were measured for the sodium halides and alkali, bromides (chlorides in the case of φK) at a fixed salt aquamolality of 0.1m and fixed urea molality of 3m. The corresponding transfer functions from H2O to 3m urea are opposite those from H2O to D2O and similarly are relatively independent of ionic size. This suggests that urea, shows no specific interaction affinity for ions and that the overall number of water molecules influenced by the ions is relatively constant for all alkali halides. The lithium halides are an exception in that Li+ seems to have hardly any structure-breaking effect.  相似文献   

10.
On the RbNiCrF6 Type. III. New Fluorides of the Type CsZnMF6 (M = Al, Ga, In, Tl, Sc, Ti, V, Mn, Cu, Rh) Cubic compounds are CsZnGaF6 [3] (colourless, a = 10.29 Å); CsZnInF6 (colourless, a = 10.58 Å); CsZnTlF6 (colourless, a = 10.62 Å); CsZnScF6 (colourless, a = 10.58 Å); CsZnTiF6 (lightblue, a = 10.50 Å); CsZnVF6 (lightgreen, a = 10.43 Å); CsZnMnF6 (redbrown, a = 10.40 Å); CsZnCuF6 (light brown, a = 10.24 Å); CsZnRhF6 (redbrown, a = 10.41 Å), all RbNiCrF6 type of structure, in addition non cubic: CsZnAlF6 (colourless). The Madelung part of lattice energy, MAPLE, is calculated and discussed.  相似文献   

11.
The complex H2Os33-NCH3)(CO)12 in decalin at 198°C in 35% yield. Crystal data for the former obtained at ?158°C are: orthorhombic, space group Pmcn, a 14.113(2), b 6.605(1), c 17.683(4) », Z = 4, Dc 3.44 g cm?3. The hydrogen atoms are related by symmetry. The position of the unique hydrogen atom has been refined. It is observed asymmetrically bridging (closer to the unique Os atom) the longer edge of the isosceles triosmium triangle. The hydrogen atoms are out of the trimetal plane away from the triply-bridging nitrogen atom.Crystal data for the tetraosmium complex at 25°C are: monoclinic, space group C2/c a 30.818(9), b 8.463(2), c 16.621(2), », β 108.90(2)°, Z = 8, Dc 3.75 g cm?3. The four osmium atoms form a distorted tetrahedral framework capped by the nitrogen atom of the methylnitrene group on the face containing the three longer OsOs separations.  相似文献   

12.
Trigonal Planar CuX3-Groups in Cu2Mo6X14, X = Cl, Br, I Cu2Mo6Cl14 (I), Cu2Mo6Br14 (II) and Cu2Mo6I14 (III) were synthesized by thermal treatment of corresponding mixtures of copper(I) and molybdenum(II) halides. The crystal structures were determined by single crystal X-ray analyses. I and II show isotypism, cubic, Pn3 (no. 201, sec. setting), Z = 4, I: a = 12.772(3) Å, II: a = 13.350(2) Å. III shows a new structural type, orthorhombic, Pbca (No. 61), Z = 4, a = 16.058(3) Å, b = 10.643(2) Å, c = 16.963(3) Å. Trigonal planar CuX3 units were found in I? III. Structural behaviour relations are discussed, especially with regard to ionic conductivity.  相似文献   

13.
The ion–molecule reactions of CH3NH2+, (CH3)2NH+, and (CH3)3N+ with the respective amines have been investigated at thermal kinetic energies in a high-pressure photoionization mass spectrometer at several wavelengths (energies) in the vacuum ultraviolet. The absolute rate coefficient for proton transfer from (CH3)3N+ to (CH3)3N decreases from 8.2 × 10?10 cm3/molecule · sec at 147.0 nm (8.4 eV) to 4.9 × 10?10 cm3/molecule. sec at 106.7-104.8 nm (11.7 eV). In dimethylamine, the rate coefficient decreases from 11.6 × 10?10 cm3/molecular. sec at 8 4 eV to 10.2 × 10?10 cm3/molecule osec at 11.7 eV, while no significant effect of energy was detected in methylamine. The reactions of several fragment ions are also reported. Experiments were also carried out at pressures up to 0.5 torr in order to investigate the further solvation of CH3NH2+, (CH3)2NH2+, and (CH3)3NH+. It was found that the maximum proton solvation numbers in methyl-, dimethyl-, and trimethyl-amine are 4, 3, and 2, respectively, under these conditions.  相似文献   

14.
Contributions to the Properties of Titanates with Ilmenite Structure. I. Aspects of a High Temperature Phase Transition in NiTiO3 NiTiO3 (Ilmenite structure) shows an anomalous increase of the electrical conductivity in the temperature range between 1 250°C and 1 290°C. The temperature dependence of the Seebeck coefficient is also found to be anomalous in this region. DTA experiments are consistent with a phase transition; the transition enthalpy was estimated to ΔuHm = 17 ± 3 kJmol?1. NiTiO3 powders and single crystals were quenched from 1 350°C and 1 200°C to room temperature; in principle no differences of the lattice parameters or atomic positions could be detected. The determination of the distribution of cations using x-ray powder methods failt because of strong texture effects. Structure refinements with single crystal methods suggest an ordered Ilmenite structure independent of the quenching temperature. The results are in agreement with a reversible order-disorder transition of higher order. It seems to be impossible to quench the high temperature phase by conventional methods.  相似文献   

15.
The same secondary ferrocenylisopropylcarbenium ion was formed from treatment with concentrated H2SO4 at 10°C of either ferrocenylisopropylcarbinol or ferrocenylmethyldimethylcarbinol and the same secondary ferrocenyldiphenylmethylcarbenium ion was obtained when either ferrocenyldiphenylmethylcarbinol or ferrocenylmethyldiphenylcarbinol was treated with CF3COOH at 5°C. The results indicate the occurrence of 1,2-hydride shifts converting tertiary to secondary carbocations, thus providing a novel demonstration of the extraordinary stability of α-ferrocenyl substituted carbocations.  相似文献   

16.
The G-value for D2 production has been measured for C6D12 and for gem-, cis-, and trans-C6H10D2. The G-values for the unimolecular process are 0.24, 0.023, 0.010 and 0.004, resp. In C6D12 two bimolecular processes can be distinguished: one includes ‘hot’ D-atoms, e. g. a species that shows no isotope effect for the hydrogen abstraction yielding D2 and HD, resp.; its G-value is 1.22. The other process goes via thermal D-atoms, it has a G-value of 1.97 and gives an isotope effect for abstraction: kD/kH = 0.075 (10°C). In the partially deuteriated cyclohexanes, only the former process can be observed. Its G-value depends on the irradiation temperature and phase, is independent of the position of the D-atoms in the molecule and amounts to 0.035 (10°C), 0.028 (?40°C solid), 0.040 (?40°C liquid with 25% hexane). The following conclusions can be drawn: the unimolecular detachment shows no intramolecular isotope effect in partially deuteriated molecules and comes only from the gem-, cis- and trans-positions. Most of the G-values for the D2 and HD production by different reaction paths can be cross-checked by G-values from different mixtures. The agreement between these values is excellent.  相似文献   

17.
Hydrates of Barium Chloride. X-ray, Thermoanalytical, Raman, and I.R. Data In the system BaCl2? H2O the hydrates BaCl2 · 2 H2O, BaCl2 · 1 H2O, BaCl2 · 1/2 H2O, and BaCl2 · uH2O were obtained. X-ray powder data, i.r. and Raman spectra, as well as thermoanalytical measurements (TG, DTA) are reported. BaCl2 · 1 H2O and BaCl2 · 1/2 H2O, which are both isotype with the corresponding hydrates of SrCl2, were prepared by dehydration of BaCl2 · 2 H2O or by back hydration of anhydrous BaCl2 with the calculated amounts of water. BaCl2 · uH2O (u ≈? 1) is formed as the primary product by the reaction of anhydrous BaCl2 with water vapour at room temperature. Preparation methods of salt hydrates by controlled back hydration of the anhydrous salts are reported.  相似文献   

18.
Gas-phase electron-diffraction methods have been used to determine the molecular structure of bis(difluorophosphino)ether, F2POPF2. Most of the geometrical parameters are strongly correlated due to overlapping peaks in the radial distribution curve. In the structure that fits the experimental data most closely, the P-F and P-O bond lengths are 159.7 ± 0.4 and 153.3 ± 0.6 pm respectively, and the POP angle is 2.53 ± 0.02 rad (145°). The conformation is such that the molecule has no symmetry elements other than I (point group C1). In other refinements somewhat longer P-O and shorter P-F distances were obtained.  相似文献   

19.
Samples of low-density polyethylene (LDPE), high-density polyethylene (HDPE), and tetratetracontane (n-C44H90) free from additives were heated in air at temperatures between 120 and 180°C. As a comparison, “as received” HDPE containing unspecified additives has also been included. The structural changes have been studied with gel chromatography, viscometry, infrared spectroscopy, differential scanning calorimetry, and gravimetric measurements. LDPE, HDPE, and n-C44H90 follow the same course of thermooxidative degradation when they are free from additives and present in the molten state. Both molecular-diminishing and enlargement reactions occur. At temperatures below 150°C molecular enlargement is not observed until after rather long exposure times, whereas at higher temperatures enlargement occurs immediately. The difference is because “peroxide curing” becomes increasingly important above 150°C, whereas ester formation is operating at all temperature levels. Degradation below Tm is restricted to the amorphous phase that results in a different degradation pattern. In accelerated testing work extrapolations of the Arrhenius type in the prediction of structural change are thus not justified, even within the actual narrow temperature range. Neither are changes in commonly used standards like carbonyl content justified as a measure of the changes; for example, in mechanical properties. The stabilizer in the unpurified HDPE not only influences the induction period but also the course of the thermooxidative degradation.  相似文献   

20.
Zinc dimethyl-, methylphenyl-, and diphenylphosphinate have been prepared by the reaction of Zn(C2H3O2)2·2H2O with the appropriate acid or salt and found to exist in amorphous and crystalline forms. Zinc methylphenylphosphinate, the most tractable, exhibits many of the physical attributes of polymeric materials, both in solution and bulk form, supporting our earlier suggestion that these compounds are double-bridged coordination polymers. Thermogravimetric analysis indicates initial decomposition temperatures of 440, 425, and 490°C., respectively, for the three polymers, but long-term studies show initiation of decomposition at somewhat lower temperatures.  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