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 218 毫秒
1.
The novel λ4-thia-λ5-phospha-h2-manganabicyclo[2.2.1]heptadienes (OC)3Mn[CR2CR2CR2CR2PR12S] (R1 = CH3, C6H5; R2 = CO2CH3, CO2C2H5, CO2C6H11) are formed by action of the activated alkynes R2C  CR2 on the heterocycles [(OC)4MnPR12S]2 via the isolable, five-membered heterometallacyclopentadienes (OC)4MnSPR12C(R2)C(R2). The compound with R1 = CH3 and R2 = CO2CH3 crystallizes in the triclinic space group P1 with Z = 2 and separates quantitatively the thiophene derivative CR2CR2CR2CR2S under CO pressure or by reaction with (NH4)2Ce(NO3)6. The use of various acetylenes and of acetylenes with different alkyl groups yields the unsymmetric substituted manganabicycloheptadienes (OC)3Mn[CR4CR3CR2CR2P(CH3)2S] (R2 = CO2CH3, R3 = R4 = CO2C2H5; R2 = R4 = CO2CH3, R3 = H). With propionic acid methylester the alkyne insertion proceeds regiospecifically. With Raney nickel selective S elimination under ring contraction and formation of the λ4-phospha-h2-manganabicyclo[2.1.1]hexenes (OC)3Mn[CR2CR3CR2CR2PR12] (R1 = CH3: R2 = R3 = CO2CH3, CO2C2H5; R2 = CO2CH3, R3 = H; R1 = C6H5: R2 = R3 = CO2C2H5) occurs. (OC)3Mn[CR2CR3CR2CR2P(CH3)2] (R2 = R3 = CO2CH3) crystallizes in the monoclinic space group P21/m with Z = 2. The IR and NMR spectra of the heterocycles are discussed in detail.  相似文献   

2.
Polymerization experiments with styrene in benzene at 60°, initiated by benzoyl peroxide, covering a wide range of concentration of both monomer and initiator are reported; the results cannot be explained in terms of the classical rate relationship with Rp ∝ [I]12 [M]. Deviations were reflected in unexpected orders of monomer up to [M]1·4 and of initiator down to [I]0·42 when the initiator concentration is increased and monomer concentration is decreased. Based on the concept of primary radical termination, an equation, viz.
lnRp2[I][M]2=ln2fkkdkp2kt?2kprtkikp×Rp[M]2
is proposed. Semi-log plots of Rp2/[I] [M]2 vs Rp/[M]2 show a wide range of linearity; the characteristic constant kprt/kikp and also fk can be obtained from the slope and intercept, respectively, kprt, ki and kp are, respectively, the rate constants of primary radical termination, initiation and propagation and fk is the efficiency of initiation, defined as the fraction of radicals which come out of the solvent cage and take part in initiation, primary radical termination and primary radical recombination. The definition of fk is thus differentiated from the conventional efficiency of initiation. Finally, we have derived an equation which allows determination of the classical efficiency of initiation as a function of [I]/[M]2 and also allows a correction of Rp in handling the above equation by taking into account the small amount of monomer consumed in initiation.  相似文献   

3.
Treatment of [{Ir(COD)(μ-Cl)}2] with excess of the electron-rich olefin [CN(Ar)(CH2)2NAr]2 (abbreviated as (LAr)2, Ar = C6H4Me-p or C6H4OMe-p) affords the ortho-metallated tricycle [Ir(LAr)3], which for Ar = C6H4Me-p (Ia) with HCL yields [Ir(LAr)2(LAr)]Cl (IV); X-ray data show that in IV there is an unexpectedly close Ir?C(o-aryl) contact (2;52(1) Å) involving the “free” LAr which compares with an IrC(o-aryl) distance of 2.09(3) Å in Ia or 2.07(3) Å in the ortho-metallated LAr ligand of complex IV.  相似文献   

4.
Results on the rate of polymerization of acrylic acid by S2O2?8 ion in alkaline and acid conditions are presented. Rp depended upon [S2O2?8]12 and [monomer]32 both in acid and alkaline solutions. The influence of ionic strength, the effect of pH on Rp and the catalytic effect of Ag+ and Cu2+ on the system have been discussed. Suitable mechanisms are proposed.  相似文献   

5.
The electrical conductivity of sintered specimens of nonstoichiometric CeO2?x was measured as a function of temperature (750–1500°C) and oxygen pressure (1–10?22 atm). The isothermal compositional dependence of the electrical conductivity of CeO2?x was determined by combining recently obtained thermodynamic data, x = x(PO2, T), with the conductivity data. The compositional and temperature dependence of the electrical conductivity may be represented by the expression
σ=410[x]e?(0.158+x)kT(ohm cm)?1
over the temperature range 750–1500°C and from x = 0.001 to x = 0.1.This expression was rationalized in terms of the following simple relations for (a) the electron carrier concentration
ncece=8xa03
where nCe′Ce is the number of Ce′Ce per cm3 and a0 is the lattice parameter and (b) the electron mobility
μ=5.2(10?2)e?(0.158+x)kT(cm2/V sec)
.  相似文献   

6.
The study of K2NiF4 and perovskite structure type by the “method of invariants” leads to the relationship: (A-X)9 212 ? (A-X)12 = constant, where (A-X)9 and (A-X)12 are the invariant values associated with cation A in coordination number 9 and 12. In the case where A = K+ and X = F?, we propose the relationship:
(K+?F)R = 2.832 R111.4
where R is the coordination number.  相似文献   

7.
It is predicted that a gap of 2μE/312 appears in the dispersion potential for a pair of ground-state (identical) atoms, or molecules, at distance R such that ?δ + (4π?0)?1(μ23R3)(1 ? 3 cos2θ) = 0, where μ is the electric dipole moment, E is the field strength of an intense radiation field and θ is the angle between the polarization direction of the field and R. The detuning term Δ is ω0?ω where ω0 is the transition frequency of the atom or molecule, and ω is that of the field. The gap is of the order of 1011 Hz when the intensity of the radiation field is a few MWcm2.  相似文献   

8.
Benzophenone (BP) in low concentrations (<0.001 mol 1?1) produces a rate enhancing effect in the H2O2-induced bulk photopolymerization of MMA. Rp is proportional to [H2O2]0.4 and [BP]0.4, and kp2k1 at 30° is 1.00 × 10?2 1.mol?1 sec?1. In diluted systems, different solvents produce different kinetic effects, reaction order with respect to monomer being negative for IPA and THF as solvent, positive but <1.0 for benzene and chloroform, 1.2 for acetonitrile, CCl4 and t-butanol and 1.8 for DMA. The variable solvent effect is attributed to modification of the initiation process by the various solvents to different extents. Kinetic analysis of data for bulk photopolymerization gives evidence for primary radical termination and degradative initiator transfer.  相似文献   

9.
The heat capacity of the solid solution Mn3.2Ga0.8N was measured between 5 to 330 K by adiabatic calorimetry. A sharp anomaly with first-order character was detected at TA = (160.5±0.5) K, corresponding to a magnetic rearrangement and a lattice expansion. No sharp anomaly was observed at Tc ≈ 260 K where the magnetic ordering takes place; instead, a smooth shoulder was detected. The thermodynamic functions at 298.15 K are Cp,mR = 15.16, SmoR = 18.57, {Hmo(T)?Hmo(0)}R = 2896 K, ?{Gmo(T)?Hmo(0)}RT = 8.85. At low temperatures the coefficient for the linear electronic contribution to the heat capacity was derived: γ = (0.031±0.003) J·K?2·mol?1. Moreover, the different contributions to the heat capacity were obtained and the electronic origin of the phase transitions was established.  相似文献   

10.
Negative chemical ionisation mass spectrometry is used as a probe to identify reactions between hydrocarbon radicals and cornplexed cobalt(II) centres in the gas phase. Methane NCI mass spectra of a series of cobalt(II) complexes containing O4, O2N2 and N4 donor atom sets are characterised by adduct ions of the form [M + CnH2n+1]? at m/z values above the molecular ion, [M]?. Formation of such ionic species has been rationalised in terms of a one-electron oxidative-addition mechanism involving attack by hydrocarbon plasma-derived alkyl radicals at the metal centre prior to electron capture: CoIILn + R? → RCoIIILne? [CoILn]?. The competing resonance electron attachment reaction: CoIILne? also occurs within the ion source.  相似文献   

11.
Cobaltacyclopentadiene complexes, (η5-C5H5)(PPh3)(CoCR1CR2CR2CR1) (R1, R2 =; Ph, Me, CO2Me), reacted with RH (RH: triethylsilane, thiocresol, dimethyl- and ethylene-thiourea, pyrrole, thiophene) to give diene complexes, (η5-C5H5)(η4-HCR1CR2CR2CR1R)Co, or uncomplexed, highly substituted butadiene derivatives, HCR1CR2CR2CR1R. The reaction with thiourea proceeded catalytically in the presence of excess of diphenylacetylene although turn-over of the catalyst was small.  相似文献   

12.
The electrical conductivity of polycrystalline strontium titanate with (SrTi = 0.996, 0.99, and 0.98 was determined for the oxygen partial pressure range of 100 to 10?22 atm and the temperature range of 850–1050°C. These data were found to be similar to that obtained for the sample with ideal cationic ratio. The observed data were proportional to the ?16 power of oxygen partial pressure for PO2 < 10?15atm, proportional to P?14O2 for the pressure range 10?8–10?15 atm, and proportional to P+14O2 for PO2 > 10?4atm. The deviation from the ideal Sr-to-Ti ratio was found to be accommodated by neutral vacancy pairs, (V″Sr V″0. The results indicate that the single-phase field of strontium titanate extends beyond 50.505 mole% TiO2 at elevated temperatures.  相似文献   

13.
The molecular structure of the dinuclear complex [(η6-benzene)Mo(μ-η6: η4-azulene)Cr(CO)3 was determined by an X-ray diffraction study. The reaction of [(η6-azulene)(η6-benzene)Mo] with [RhCI(CO)2]2 gives the salt [η6-benzene)Mo(μ-η6: η4-azulene)Rh(CO)2]+[RhCl2(CO)2]?, the structure of which was also characterized by X-ray crystallography. The cation of this salt can also be synthesized from [(η6-azulene)(η6-benzene)Mo] and [Rh(CO)2]+. The fluxionality of the cation was studied by temperature-dependent 1H-NMR measurements. The complex [(η6-azulene)(η6-benzene)Mo] reacts with Fe2(CO)9 to give the dinuclear complex [(η6-benzene)Mo(μ-η5 : 3-azulene)Fe(CO)3], as confirmed by X-ray diffraction.  相似文献   

14.
The electrical conductivity of polycrystalline CaTiO3 was measured over the temperature range 800–1100°C while in thermodynamic equilibrium with oxygen partial pressures from 10?22 to 100 atm. The data were found to be proportional to the ?16th power of the oxygen partial pressure for the oxygen pressure range 10?16 – 10?22 atm, proportional to P?14O2 for the oxygen pressure range 10?8 – 10?15 atm, and proportional to P+14O2 for the oxygen pressure range greater than 10?4 atm. The region of linearity where the electrical conductivity varies as ?14th power of PO2 increased as the temperature was decreased. The observed data are consistent with the presence of small amounts of acceptor impurities in CaTiO3. The band-gap energy (extrapolated to zero temperature) was estimated to be 3.46 eV.  相似文献   

15.
Poly[2-methoxy-4,6-di(p,p′-isopropylidene diphenyloxy)-s-triazine] has been synthesized by interfacial polycondensation of 2-methoxy-4,6-dichloro-s-triazine with bisphenol A. The conditions have been optimized to prepare a sample of both high solubility and molecular weight. The polymer has been fractionated by fractional precipitation and the fractions characterized by viscometry, osmometry and gel permeation chromatography. Mark-Houwink-Kuhn-Sakurada (MHKS) expressions have been developed for three solvents. The values of unperturbed dimensions (<R2>0/M)12, solvent-polymer interaction parameter (B and χ1) and conformational parameter (σ) have been computed by applying the two-parameter theories of excluded volume developed by Fixman, Kurata, Stockmayer and Roig and by Flory, Fox and Schaefgen.  相似文献   

16.
N-Benzoyloxyimides such as N-benzoyloxysuccinimide (NBzS) were found to act as photoinitiators for acrylonitrile (AN), methyl acrylate and methyl methacrylate. Polymerization rate (Rp) for the photopolymerization of AN with NBzS in dimethylformamide at 30° was expressed thus;
Rp - k [NBZS]0.25 [AN1.3
The overall activation energy was 27.6 kJ/mol. A spin trapping study on the photolysis of NBzS revealed formation of acyl and phenyl radicals, indicating that simultaneous scissions of the NO and OCO bonds in NBzS occur under irradiation. An initiation mechanism is proposed and discussed.  相似文献   

17.
The kinetics of the interaction of hexaaquochromium(III) ion with potassium octacyanomolybdate(IV) have been studied using conductance and spectrophotometric data. The mechanism of the reaction is discussed and the effect of H+ ion and the ionic strength on the rate of the reaction determined. The reaction is found to be pseudo-first order with respect to potassium octacyanomolybdate(IV) and inverse first order with [H3O+]. The rate of the reaction increases with increase in ionic strength and temperature. Activation parameters have been calculated using the Arrhenius equation and have the values ΔE* = 1.3 × 102 kJ mol?1, ΔH* = 129 kJ mol?1, ΔS* = ?315 e.u., ΔF* = 2.3 × 102 kJ and A = 1.5 × 10?3. The mechanism proposed is based on ion-pair formation and the rate equation obtained is given by: kobs=[kKE[H3O+]+k′K′kEkh][Mo(CN)84?][H3O+]+kh+[KE[H3O+]+K′Ekh][Mo(CN)84?]  相似文献   

18.
Binary systems consisting of dimethylaniline-N-oxide (DMOA) and some α-halo-carbonyl compounds, such as phenacyl halide and α-halo-acetic acid ester, were found to induce radical polymerization of vinyl monomers. Bromo-derivatives showed higher initiating activities than chloroderivatives. The polymerization of methyl methacrylate (MMA) with DMAO and phenacyl bromide (PBr) was investigated kinetically. The polymerization rate (Rp) was expressed as follows; Rp = k[DMAO]12[PBr]12[MMA].The overall activation energy of the polymerization was calculated to be 42.7 kJ mol?1. No noticeable chain-transfer from the polymer radical to DMAO or PBr was observed. The benzoyl radical was trapped by 2-nitroso-2-methylpropane, a spin trapping reagent, in the reaction of DMAO and PBr. The u.v. spectrum of poly (MMA) obtained suggests that the polymer contains end-groups similar to acetophenone and DMA. From the results, an initiation mechanism for the polymerization has been proposed and discussed.  相似文献   

19.
Compounds MIIMeIVF6 requently undergo phase transitions from the cubic ordered ReO3 to the trigonal LiSbF6 structure when lowering the temperature. In case of a strongly Jahn-Teller unstable cation in the MII position additional phases may occur. Results of powder neutron-diffraction studies on CaSnF6, FeZrF6, and CrZrF6 at different temperatures are reported. The high-temperature phases have the space group Fm3m; the F? ligands are either statistically displaced from the MIIMeIV directions or undergo a strong thermal motion perpendicular to these directions (?MIIFMeIV: 165–180°). The thermal ellipsoids of the CrF bonds are strongly indicative of a dynamical Jahn-Teller effect in addition. In the low-temperature phases of CaSnF6 and FeZrF6 (space group R3) the ?MIIFMeIV is more distinctly bent (?155–160°). CrZrF6 undergoes two reversible phase transitions, which are determined to occur at 415 ± 5 K (cubic → tetragonal, dynamic to static Jahn-Teller distortion of CrF6 octahedra and 150 ± 10 K (tetragonal → (pseudo)monoclinic).  相似文献   

20.
The new phosphine, PBut2Bui (L), was prepared from But2PCl and LiBui. PPh2Bui (L′) was prepared from Ph2PCl and LiBui. Treatment of [PtCl2(NCBut)2] with L′ gives [PtCl2L′2] which does not cyclometallate even on prolonged boiling in 2-methoxyethanol. In contrast, [PtCl2(NCBut)2] reacts with PBut2Bui in boiling 2-methoxyethanol to give the cyclometallated complex [Pt2Cl2(PBut2CH2-CHMeCH2)2] (II, X = Cl). The corresponding bromide, iodide and acetylacetonate were prepared. With PPh3 II (X = Cl) gives [PtCl(PBut2CH2CHMeCH2)(PPh3)] which with NaBH4 gives [PtH(PBut2CH2CHMeCH2)(PPh3)]. Na2PdCl4 with L (2 mol equivalents) gave trans-[PdCl2L2], which was converted into trans-[Pd(NCS)2-L2] by metathesis with KSCN. Treatment of Na2PdCl4 with L (1 mol equivalent) gave [Pd2Cl4L2], which on heating in 2-methoxyethanol gave [Pd2Cl2(PBut2CH2-CHMeCH2)2], as a mixture of syn- and anti-isomers. The complexes trans-[PdCl2-L′2] and [Pd2Cl4L′2] were also prepared. 1H- and 31P NMR data are given.  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