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.
Vapour pressure measurements have been carried out on the complexes W(CO)it6-x (NCCH3x(x=1,2,3) and Mo(CO)it6-x(NCCH3x(x=1,3) employing the Knudsen effusion technique. The following enthalpies of sublimation, ΔH298sub(kJ mole?1), have been determined from vapour pressure data: W(CO)5(NCCH3)=98.1±2.0; W(CO) 4 (NCCH3)2=131.0±6.0; W(CO)3(NCCH33=103.4±6.0; Mo(CO)5(NCCH3)=105.8± 5.6; and Mo(CO)3(NCCH3)3=111.3±3.0.  相似文献   

2.
Second virial coefficients, A2, intrinsic viscosities, [η], and solvation preferential coefficients, γ, for the ternary systems n-hexane, HEX, (1)/butanone, MEK, (2)/poly(dimethylsiloxane), PDMS, (3) and n-heptane, HEP, (1)/MEK (2)/PDMS (3) have been determined at 20.0°. Binary interaction parameters, g2, have also been measured by light scattering. Inversion in γ at u2 ? 0.15 in the HEX system and at u2 ? 0.22 in the HEP system takes place. The inversion points are accompanied by smooth maxima in A2 and in [η]. Both systems show a weak cosolvent character. Theoretical γ's derived, according to Flory-Huggins and Flory-Prigogine-Patterson, are compared with experimental values. The evaluation of the ternary interaction potential, gT, and its dependence on system composition allow evaluation of the binary interaction potentials and their dependence on polymer concentration.  相似文献   

3.
A detailed analysis of the chemiluminescence emission (CL) from poly(styrene-b-ethylene-co-butylene-b-styrene), SEBS, was carried out. A phenol-phosphite stabilization system based on Irgafos 168 and Irganox 1330, was studied. The kinetic analysis of the CL profile under nitrogen shows a first-order reaction for the decay of chemiluminescence. The activation energy shows different values as a function of temperature, showing that different reactions are involved in the thermal degradation of the SEBS. The CL decay rate correlates well with the amount of the phosphite, Irgafos 168, and confirms the activity of this stabilizer as radical chain-breaking antioxidant in these copolymers.The isothermal analysis of CL under oxygen allows evaluation of the oxidation state, as well as the efficiency of the antioxidants. Good correlations are found between the CL parameters and concentration of Irgafos 168. Several factors suggest that oxidation begins in the interfacial region. Spectral analysis of the chemiluminescence shows the presence of different types of hydroperoxides.Finally, the characterization of the SEBS copolymers by differential scanning calorimetry reveals an order-disorder transition, assigned to aggregates that behave as paracrystalline regions.  相似文献   

4.
Nanowires of an iodine containing Pb-Sb-sulfosalt have been synthesized by chemical vapor transport. Their structure was studied using high-resolution transmission electron microscopy and X-ray powder diffraction. The lattice parameters show values equal to a=4.9801(4) nm, b=0.41132(8) nm (with two-fold superstructure), c=2.1989(1) nm and β=99.918(6)°. These parameters and the results of a multislice simulation are in good agreement with the mineral pillaite, Cu0.10Pb9.16Sb9.84S22.94Cl1.06O0.5 (space group C2/m, a=4.949(1) nm, b=0.41259(8) nm, c=2.1828(4) nm, and β=99.62(3)°). Microprobe and EDX analyses yielded a chemical composition of Cu0.507(5)Pb8.73(9)Sb8.15(8)I1.6S20.0(2) which is close to natural pillaite but contains no oxygen and iodine instead of chlorine. The structure of the investigated material is based on chains of M-S polyhedra (M=Pb or Sb) typical for the architecture of sulfosalts implying iodine atoms in trigonal prismatic coordination with Pb atoms from the M-S polyhedra of neighboring chains. The [010] superstructure of the specimen was found to be unstable under electron beam irradiation with a rapid decrease of the b lattice parameter from 0.8 to 0.4 nm within 5 min.  相似文献   

5.
The solubility properties of the zinc chloride amino acids(Leu/Try/ Val/Thr) water systems at 25 ℃ in the whole concentration range have been investigated by equilibrium method, the corresponding phase diagrams and refractive index curves were constructed. The results indicate the formation of the congruently soluble compound of Zn(Leu)Cl2(A) in Leu system and the incongruently soluble compounds of Zn(Try)Cl2•1/2H2O(B)、 Zn(Val)Cl2 •H2O(C)、 Zn(Val)2Cl2(D)、 Zn(Thr)Cl2•H2O(E) and Zn(Thr)2Cl2(F) in Try/Val/Thr systems, which have not been reported in literature. According to the phase equilibrium result, six solid complexes have been prepared in water, and their compositions have been determined by chemical analysis which are consistent with the results of phase diagrams.  相似文献   

6.
Phase analytical investigations in the system magnesium-iridium-indium revealed the magnesium-rich intermetallics Ir3.30(1)Mg17.96(4)In0.74(4) and Ir3Mg17.1(1)In1.9(1). The samples were prepared from the elements via induction melting in glassy carbon crucibles in a water-cooled sample chamber and subsequent annealing. Both intermetallics were investigated by X-ray powder and single-crystal diffraction: C2/c, Z=4, a=979.1(1), b=2197.4(2), , β=105.79(1)°, wR2=0.0434, 3076 F2 values, 108 variables for Ir3.30(1)Mg17.96(4)In0.74(4), and a=983.39(8), b=2211.4(2), , β=105.757(6)°, wR2=0.0487, 3893 F2 values, and 115 variables for Ir3Mg17.1(1)In1.9(1). Both compounds show solid solutions. In Ir3.30(1)Mg17.96(4)In0.74(4), the indium site shows an occupancy by 69.9(4)% In+30.1(4)% Ir, and one magnesium site has a small mixed occupancy with indium, while nine atomic sites in Ir3Mg17.1(1)In1.9(1) show Mg/In mixing with indium occupancies between 1.2(3)% and 14.8(3)%. The relatively complex crystal structure is of a new type. It can be explained by a packing of coordination number 10 and 12 polyhedra around the iridium atoms. The crystal chemical peculiarities and chemical bonding in both intermetallics is briefly discussed.  相似文献   

7.
The 63Cu quadrupole, the 63Cu hyperfine and 31P hyperfine coupling tensors of the mixed-ligand chelate (n-Bu4N)[63Cu(mnt)(H(et)dtp)] doped into the corresponding nickel complex have been measured by ENDOR spectroscopy. Conclusion are drawn concerning the symmetry of the Cu complex. The 63Cu quadrupole tensors and the 31P hyperfine tensors can be understood assuming a “transannular Cu 3ditx2 -ity2-P 3s overlap”.  相似文献   

8.
Likely candidates for the lowest potential energy minima of (C60)nCa2+, (C60)nF and (C60)nI clusters are located using basin-hopping global optimisation. In each case, the potential energy surface is constructed using the Girifalco form for the C60 intermolecular interaction, an averaged Lennard–Jones C60–ion interaction, and a polarisation potential, which depends on the first few non-vanishing C60 multipole polarisabilities. We find that the ions generally occupy the interstitial sites of a (C60)n cluster, the coordination shell being tetrahedral for Ca2+ and F. The I ion has an octahedral coordination shell in the global minimum for (C60)6I, however for 12  n  8 the preferred coordination geometry is trigonal prismatic.  相似文献   

9.
Single crystals of KxMg(8+x)/3Sb(16−x)/3O16 (x≈1.76) with a hollandite superstructure were grown. Ordering schemes for guest ions (K) and the host structure were confirmed by the structure refinement using X-ray diffraction intensities. The space group is I4/m and cell parameters are a=10.3256(6), c=9.2526(17)Å with Z=3. Superlattice formation is primarily attributed to the Mg/Sb occupational modulation in the host structure. Mg/Sb ratios at two nonequivalent metal sites are 0.8977/0.1023 and 0.1612/0.8388. Two types of the cavity are seen in the tunnel, where parts of K ions deviate from the cavity center along the tunnel direction. Probability densities for K ions in the two cavities are different from each other, which seems to have arisen from the Mg/Sb modulation.  相似文献   

10.
The thermal behavior of kaolinite–urea intercalation complex was investigated by thermogravimetry–differential scanning calorimetry (TG–DSC), X-ray diffraction (XRD), and fourier transform infrared spectroscopy (FTIR). In addition, the interaction mode of urea molecules intercalated into the kaolinite gallery was studied by means of molecular dynamics simulation. Three main mass losses were observed at 136 °C, in the range of 210–270 °C, and at 500 °C in the TG–DSC curves, which were, respectively, attributed to (1) melting of the surface-adsorbed urea, (2) removal of the intercalated urea, and (3) dehydroxylation of the deintercalated kaolinite. The three DSC endothermic peaks at 218, 250, and 261 °C were related to the successive removals of intercalated urea with three different distribution structures. Based on the angle between the dipole moment vector of urea and the basal surface of kaolinite, the three urea models could be described as follows: (1) Type A, the dipole moment vector is nearly parallel to the basal surface of kaolinite; (2) Type B, the dipole moment vector points to the silica tetrahedron with the angle between it and the basal surface of kaolinite ranging from 20°to 40°; and (3) Type C, the dipole moment vector is nearly perpendicular to the basal surface of kaolinite. The three distribution structures of urea molecules were validated by the results of the molecular dynamics simulation. Furthermore, the thermal behavior of the kaolinite–urea intercalation complex investigated by TG–DSC was also supported by FTIR and XRD analyses.  相似文献   

11.
The rare earth-nickel-indides Tm2Ni1.896(4)In, Tm2.22(2)Ni1.81(1)In0.78(2), Tm4.83(3)Ni2In1.17(3), and Er5Ni2In were synthesized from the elements by arc-melting and subsequent annealing for the latter three compounds. Three indides were investigated by X-ray powder and single crystal diffraction: Mo2FeB2 type, P4/mbm, Z=2, a=731.08(4), c=358.80(3) pm, wR2=0.0201, 178 F2 values, 13 variables for Tm2Ni1.896(4)In, a=734.37(7), c=358.6(1) pm, wR2=0.0539, 262 F2 values, 14 variables for Tm2.22(2)Ni1.81(1)In0.78(2), and Mo5SiB2 type, I4/mcm, a=751.0(2), c=1317.1(3) pm, wR2=0.0751, 317 F2 values, 17 variables for Tm4.83(3)Ni2In1.17(3). X-ray powder data for Er5Ni2In revealed a=754.6(2) and c=1323.3(5) pm. The Mo2FeB2 type structures of Tm2Ni1.896(4)In and Tm2.22(2)Ni1.81(1)In0.78(2) are intergrowths of slightly distorted CsCl and AlB2 related slabs, however, with different crystal chemical features. The nickel sites within the AlB2 slabs are not fully occupied in both indides. Additionally In/Tm mixing is possible at the center of the CsCl slab, as is evident from the structure refinement of Tm2.22(2)Ni1.81(1)In0.78(2). The Mo5SiB2 type structures of Tm4.83(3)Ni2In1.17(3) and Er5Ni2In can be considered as an intergrowth of distorted CuAl2 and U3Si2 related slabs in an ABAB′ stacking sequence along the c-axis. Again, one thulium site shows Tm/In mixing. The U3Si2 related slab has great structural similarities with the Mo2FeB2 type structure of Tm2Ni1.896(4)In and Tm2.22(2)Ni1.81(1)In0.78(2). The crystal chemical peculiarities and chemical bonding in these intermetallics are briefly discussed.  相似文献   

12.
The homocoupling reaction between the conjugated n-(2-chloroethenyl)pyridine; n, 2-, 3- and 4- (or quinoline; n, 2- and 4-) mediated by zero-valent nickel complexes at room temperature affords to the corresponding 1,4-diaryl-1,3-butadiene, always as the 1E,3E stereoisomer. The yield in 1,4-diaryl-1,3-butadiene increases with the nickel catalyst and hence, the active zero-valent nickel catalyst is not regenerated during the homocoupling reaction.The stereospecific synthesis of (1Z,3Z)-1,4-di(4′-pyridyl)-1,3-butadiene stereoisomer was efficiently carried out by partial hydrogenation of the appropriate 1,4-di(4′-pyridyl)-1,3-butadiyne.  相似文献   

13.
Using density functional theory (DFT) method with 6-31G* basis set, we have carried out the optimizing calculation of geometry, vibrational frequency and thermodynamical stability for (AlN) n + and (AlN) n + (n=1–15) clusters. Moreover, their ionic potential (IP) and electron affinity (EA) were discussed. The results show that the electrical charge condition of the cluster has a relatively great impact on the structure of the cluster and with the increase of n, this kind of impact is reduced gradually. There are no Al-Al and N-N bonds in the stable structure of (AlN) n + or (AlN) n -, and the Al-N bond is the sole bond type. The magic number regularity of (AlN) n + and (AlN) n - is consistent with that for (AlN) n , indicating that the structure with even n such as 2, 4, 6, ... is more stable. In addition, (AlN10 has the maximal ionization power (9.14 eV) and the minimal electron affinity energy (0.19 eV), which manifests that (AlN)10 is more stable than other clusters.  相似文献   

14.
Racemic α-methylbenzyl vinyl ether was copolymerized with optically active (S)-(-)- or (R)-(+)-N-(α-methylbenzyl)maleimide using 2,2′-azobisisobutyronitrile in order to examine the possibility of stereoelective radical polymerization of vinyl-type racemic monomers. The resulting copolymers were found to have almost alternating sequences of the two kinds of monomeric units. The non-polymerized α-methylbenzyl vinyl ether, recovered from the copolymerization system, showed an optical activity of opposite sign to the optically active comonomer used, indicating clearly that the co-polymerization process is stereoelective. It was confirmed that α-methylbenzyl vinyl ether preferentially incorporated in the copolymer has the same absolute configuration as the optically active N-substituted maleimide.  相似文献   

15.
Geometrical isomerization of fac-Mo(CO)3L3 (L = P(OPh)3, P(OMe)3, P(OEt)3) to the mer form and that of cis-Mo(CO)4L2 (L = P(OPh)3, P(OMe)3, PPh2(OMe)) to the trans form were observed in CH2Cl2 at room temperature in the presence of a catalytic amount of Me3SiOSO2CF3 (TMSOTf). Crossover experiments suggest that a ligand dissociation is not involved in the isomerization. A catalytic cycle involving an interaction of the silicon atom in Me3Si+ with one oxygen in P(OR)3 ligands has been proposed. The first isolation and the X-ray structure analysis were attained for mer-Mo(CO)3{P(OPh)3}3 through the TSMOTf-assisted isomerization of fac-Mo(CO)3{P(OPh)3}3.  相似文献   

16.
Type-II band engineered quantum dots (CdTe/CdSe(core/shell) and CdSe/ZnTe(core/shell) heterostructures) are described. The optical properties of these type-II quantum dots are studied in parallel with their type-I counterparts. We demonstrate that the spatial distribution of carriers can be controlled within the type-II quantum dots, which makes their properties strongly governed by the band offset of the comprising materials. This allows access to optical transition energies that are not restricted to band gap energies. The type-II quantum dots reported here can emit at lower energies than the band gaps of comprising materials. The type-II emission can be tailored by the shell thickness as well as the core size. The enhanced control over carrier distribution afforded by these type-II materials may prove useful for many applications, such as photovoltaics and photoconduction devices.  相似文献   

17.
Supersonic jet expansions of mixtures of nitric oxide with either nitrous oxide or carbon dioxide have been investigated over a wide range of relative concentrations. Mixed molecular cluster ions of the form (NO) m + (N2O)n and (NO) m + (CO2)n are detected following non-resonant two-photon ionization. Over a wide range of intermediate concentrations, the cluster ion distributions (NO) 3 + (N2O)n and (NO) 3 + (CO2)n with n30 are significantly more intense than clusters containing other numbers of nitric oxide molecules. The extra abundance of these species is attributed to their especially stable structures and several possible forms are discussed. An intriguing possibility involves a stable cyclic nitric oxide trimer (or ion) when combined with nitrous oxide or carbon dioxide clusters.  相似文献   

18.
The title compound crystallizes in the orthorhombic space group P212121 with 4 molecules in the unit cell (cell dimensions: a 9.778(2), b 10.639(2) and c 12.423(4) Å). The structure was solved by means of the heavy atom method. The rhodium atom is linked to both olefinic double bonds. The terpene carbonyl group does not participate in coordination to rhodium. Unlike the endocyclic olefinic group, which is approximately perpendicular to the coordination plane of rhodium, the exocyclic Cz.sbnd;C double bound shows a considerable deviation from this arrangement. The π-complexation of carvone with rhodium proceeds diastereospecifically. The absolute configuration of (+)-carvone is 4S in agreement with the assignment derived by indirect chemical correlation.  相似文献   

19.
Two new potassium vanadium phosphates have been prepared and their structures have been determined from analysis of single crystal X-ray data. The two compounds, K3(VO)(V2O3) (PO4)2(HPO4) and K3(VO)(HV2O3)(PO4)2(HPO4), are isostructural, except for the incorporation of an extra hydrogen atom into the nearly identical frameworks. The structures consist of a three-dimensional network of [VO]n chains connected through phosphate groups to a [V2O3] moiety. Magnetic susceptibility experiments indicate that in the case of the di-hydrogen compound, there are no significant magnetic interactions between the three independent vanadium (IV) centers. Crystal data: for K3(VO)(V2O3)(PO4)2 (HPO4), Mr = 620.02, orthorhombic space group Pnma (No. 62), a = 7.023(4) Å, b = 13.309(7) Å, c = 14.294(7) Å, V = 1336(2) Å3, Z = 4, R = 5.02%, and Rw = 5.24% for 1238 observed reflections [I > 3σ(I)]; for K3(VO)(HV2O3)(PO4)2(HPO4), Mr = 621.04, orthorhombic space group Pnma (No. 62), a = 6.975(3) Å, b = 13.559(7) Å, c = 14.130(7) Å, V = 1336(1) Å3, Z = 4, R = 6.02%, and Rw = 6.34% for 1465 observed reflections [I > 3σ(I)].  相似文献   

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