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 46 毫秒
1.
Crystallographic shear (CS) phases occurring in the Nb2O5WO3 and Ta2O5WO3 systems near to WO3 were characterized by X-ray diffraction and high-resolution transmission electron microscopy. The Nb2O5WO3 samples were heated at 1600K. They contained ordered {104} and {001} CS planes and wavy CS which were composed of intergrowths of {104} and {001} CS segments. The composition range over which the {104} CS series extended was from (Nb,W)O2.954 i.e., (Nb,W)65O192, to (Nb,W)O2.942, i.e., (Nb,W)52O153. The composition range over which the {001} CS series extended was from (Nb,W)O2.9375, i.e., (Nb,W)16O47 to (Nb,W)O2.875, i.e., (Nb,W)8O23. The Ta2O5WO3 samples were prepared at 1593, 1623, and 1672K. At lower temperatures ordered {103} CS phases were found, with a composition range extending between (Ta,W)O2.960, i.e., (Ta,W)50O148, to (Ta,W)O2.944, i.e., (Ta,W)36O106. At 1673K ordered {103} CS phases occurred, as did wavy CS composed of intergrowths of {103} and {104} CS segments.  相似文献   

2.
Irradiation of solutions of n5-C5H5W(CO)3R (R  CH3n1-CH2C6H5) in cyclohexane at ca. 310490 nm leads to the formation of [n5-C5H5W(CO)3]2 and methane and of n5-C5H5W5(CO)2(n3-CH2C6H5) and some [n5-C5H5W(CO)3]2, respectively. When the irradiation is carried out in the presence of excess P(C6H5)3, the photoproducts are n5-C5H5W(CO)2[P(C6H5)3]CH3 (R  CH3) and n5-C5H5W(CO)2(n3-CH2C6H5) and trace [n5-C5H5W(CO)3]2 (R  n1-CH2C6H5). Photolysis of the n5-C5H5W(CO)3R in the presence of benzyl chloride affords n5-C5H5W(CO)3Cl (R  CH3) and both n5-C5H5W(CO)2(n3-CH2C2H5) and n5-C5H5W(CO)3Cl (R  n1-CH2C6H5), the relative amounts of the latter products depending on the quantity of added C6H5CH2Cl. Irradiation of n5-C5H5W(CO)3-CH3 in the presence of both P(C6h5)3 and C6H5CH2Cl affords n5-C5H5W(CO)2-[P(C6H5)3]CH3, but no n5-C5H5W(CO)3Cl. It is proposed that the primary photo-reaction in these transformations is dissociation of a CO group from n5-C5H5W-(CO)3R to generate n5-C5H5W(CO)2R, which can either combine with L to form a stable 18 electron complex, n5-C5H5W(CO)2(L)R (L  CO, P(C5H5)3; LR  n3-CH2C6H5), or lose the group R in a competing, apparently slower step. This proposal receives support from the observation that, light intensifies being equal, n5-C5H5W(CO)3CH3 undergoes a considerably faster photoconversion to [n5-C5H5W(CO)3]2 under argon than under carbon monoxide.  相似文献   

3.
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   

4.
Reaction of phenylimido tungsten tetrachloride with MeOH and t-butylamine gave the dimeric complexes [W(NPh)(μ-OMe)(OMe)3]2 and [W(NPh)(μ-OMe)(OMe)2Cl]2. With ethanol [W(NPh)(μ-OEt)(OEt)2Cl]2 was formed whereas isopropyl and neopentyl alcohols gave the monomeric complexes [W(NPh)(OR)4(NH2CMe3)](R = CHMe2, CH2CMe3); t-butanol gave [W(NPh)(OCMe3)3Cl(NH2CMe3)] which could not be converted to [W(NPh) (OCMe3)4]. Further reaction of [W(NPh)(μ-OMe)(OMe)3]2 with o-HOC6H4CH = NC6H3Me2(salim-H) gave the salicylaldimine complex [W(NPh)(OMC)3(salim)]. The products were characterised by analytical data, IR, 1H NMR, 13C NMR and mass spectroscopy. The crystal and molecular structures of the title complexes have been determined from single crystal X-ray diffractometer data. Crystals of [W(NPh)(μ-OMe)(OMe)3]2are triclinic with a = 8.473(7), b = 10.776(5), c = 7.683(Å, α = 102.26(3), β = 102.68(4), γ = 71.13(6)°, space group P1 Crystals of 3) [W(NPh)(OCMe3)3Cl(NH2CMe3) are monoclinic with a = 9.341(2), b = 29.608(7), c = 10.257(2) Å, β = 106.28(2)°, space group, P21/c. Both structures were solved by Patterson and Fourier methods and refined to R = 0.075 for the 1022 observed data of [W(NPh) (μ-OMe)(OMe)3]2 and to R = 0.074. For the 2033 observed data of [W(NPh)(OCMe3)3Cl(NH2CMe3). The former molecule is shown to be a dimer, the two halves of the molecule being related by a centre of symmetry. Both W atoms adopt a distorted octahedral coordination geometry and they are linked by two methoxy bridges. Trans to one of the bridging donors is the phenyl imido group with a WN bond length of 1.61(4) Å; the remaining coordination sites are filled with methoxy groups. The structure of W(NPh)(OCMe3)3 Cl(NH2CMe3) is monomeric with the phenylimido group trans to the NH2CMe3 ligand in a distorted octahedral coordination geometry. Remaining sites are filled with the chloride and 3 OCMe3 ligands. The WN (imido) bond length is 1.71(2) Å, whilst WN(amine) is 2.40(2) Å  相似文献   

5.
The phases occurring in the MnWO, FeWO, CoWO, and NiWO systems at 1373°K have been determined using X-ray diffraction and electron and optical microscopy. Experimentally most attention was given to the MnWO system, where it was found that Mn entered as the Mn2+ ion into the WO3 host matrix and formed a perovskite-related bronze MnxWO3. The highest observed x-value in the bronze is about 0.027. In addition a metastable θw(Mn) oxide with the Mo5O14 structure and a disordered oxide of overall composition approximately (Mn, W)O2.82 were found. The FeWO system was similar to the MnWO system but significant differences occurred in the CoWO and NiWO systems where MxWO3 bronze phases were not observed to form at 1373°K. The stability of the MxWO3 and the θw(M) oxides formed are discussed in terms of the ionic size of the M ions involved. It is suggested that MxWO3 bronzes are metastable if these M ions are small.  相似文献   

6.
Three new uranyl tungstates, A8[(UO2)4(WO4)4(WO5)2] (A=Rb (1), Cs (2)), and Rb6[(UO2)2O(WO4)4] (3), were prepared by high-temperature solid-state reactions and their structures were solved by direct methods on twinned crystals, refined to R1=0.050, 0.042, and 0.052 for 1, 2, and 3, respectively. Compounds 1 and 2 are isostructural, monoclinic P21/n, (1): a=11.100(7), b=13.161(9), , β=90.033(13)°, , Z=8 and (2): , , , β=89.988(2)°, , Z=8. There are four symmetrically independent U6+ sites that form linear uranyl [O=U=O]2+ cations with rather distorted coordination in their equatorial planes. There are six W positions: W(1) and W(2) have square-pyramidal coordination (WO5), whereas W(3), W(4), W(5), and W(6) are tetrahedrally coordinated. The structures are based upon a novel type of one-dimensional (1D) [(UO2)4(WO4)4(WO5)2]4− chains, consisting of WU4O25 pentamers linked by WO4 tetrahedra and WO5 square pyramids. The chains run parallel to the a-axis and are arranged in modulated pseudo-2D-layers parallel to (0 1 0). The A+ cations are in the interlayer space between adjacent pseudo-layers and provide a 3D integrity of the structures. Compounds 1 and 2 are the first uranyl tungstates with 2/3 of W atoms in tetrahedral coordination. Such a high concentration of low-coordinated W6+ cations is probably responsible for the 1D character of the uranyl tungstate units. The compound 3 is triclinic, Pa=10.188(2), b=13.110(2), , α=97.853(3), β=96.573(3), γ=103.894(3)°, , Z=4. There are four U positions in the structure with a typical coordination of a pentagonal bipyramid that contain uranyl ions, UO22+, as apical axes. Among eight W sites, the W(1), W(2), W(3), W(4), W(5), and W(6) atoms are tetrahedrally coordinated, whereas the W(7) and W(8) cations have distorted fivefold coordination. The structure contains chains of composition [(UO2)2O(WO4)4]6− composed of UO7 pentagonal bipyramids and W polyhedra. The chains involve dimers of UO7 pentagonal bipyramids that share common O atoms. The dimers are linked into chains by sharing corners with WO4 tetrahedra. The chains are parallel to [−101] and are arranged in layers that are parallel to (1 1 1). The Rb+ cations provide linkage of the chains into a 3D structure. The compound 1 has many structural and chemical similarities to its molybdate analog, Rb6[(UO2)2O(MoO4)4]. However, the compounds are not isostructural. Due to the tendency of the W6+ cations to have higher-than-fourfold coordination, part of the W sites adopt distorted fivefold coordination, whereas all Mo atoms in the Mo compound are tetrahedrally coordinated. Distribution of the WO5 configurations along the chain extension does not conform to its ‘typical’ periodicity. As a result, both the chain identity period and the unit-cell volume are doubled in comparison to the Mo analog, which leads to a new structure type.  相似文献   

7.
The osmotic swelling behavior of water-in-oil-in-water (W/O/W) type emulsion liquid membranes (ELMs) was investigated. Using an optical microscope equipped with a camera, the changes in the size of the W/O/W globules were monitored over a long period of time (up to about 4 h). The osmotic pressure gradient between the internal and external aqueous phases was induced by creating a concentration difference of d-glucose between the two aqueous phases. The results indicate that the swelling ratio, defined as the ratio of globule diameter at time t to globule diameter at t=0, decreases with the increase in ϕW/O(0) (initial volume fraction of internal aqueous phase droplets). The swelling ratio generally increases with the increase in the concentration of surfactant present in the membrane (oil) phase. The permeation coefficient of water also increases with the increase in the surfactant concentration. With the increase in ϕW/O(0) up to about 0.42, the permeation coefficient decreases only slightly. However, with further increase in ϕW/O(0), a sharp reduction in the permeation coefficient occurs. The mechanism of water transfer in ELMs of the present work is reasoned to be the diffusion of hydrated surfactants.  相似文献   

8.
Penicillin potassium salt (penicillin-K) is found to show hydrotrope action, which can increase the solubility of cationic surfactant CTAB in water. Penicillin-K also shows hydrotrope-solubilization action, which makes the W/O and O/W microemulsion more stable and increases the solubilized amount of n-C5H11OH in O/W microemulsion and that of water in W/O microemulsion for CTAB/n-C5H11OH/H2O system. However, in this system, the presence of penicillin-K can decrease the stability of the lamellar liquid crystal phase due to its structure change to bicontinuous, which are proved by the mechanism of its hydrotrope-solubilization action.  相似文献   

9.
Time-of-Flight (TOF) neutron diffraction measurements have been carried out on aqueous 8 mol% sodium acetate solutions in D2O. Scattering cross sections that were observed for sample solutions involving 12C/13C and H/D isotopically substituted acetate ions were used to derive the first-order difference functions, ΔH(Q) and ΔC(Q), and corresponding distribution functions, G H(r;r) and G C(r;r), which describe the environmental structure around the methyl and the carboxyl groups within the acetate ion, respectively. Structural parameters concerning the first hydration shell of the carboxyl group within the acetate ion were obtained through the least squares fit to the observed intermolecular difference function, ΔC inter(Q). The nearest neighbor C O...D W1 (CO: carboxyl carbon atom, DW1: water deuterium atom) distance, r(C O...D W1 ), and the angle, ∠ C O ...D W1 -O W (O W : water oxygen atom), were determined to be 2.63(1) Å and 120(1)°, respectively. The coordination number, n(C O ...D W1 ), was obtained to be 4.0(1). These results are consistent with the hydration structure in which water molecules in the first hydration shell of the carboxyl group are hydrogen-bonded with oxygen atoms of the carboxyl group.  相似文献   

10.
Following the discovery of a ternary GeW oxide with the Mo5O14 structure, a large number of ternary MWO systems were surveyed to investigate the frequency of occurrence of this structure type. Samples were prepared by heating tungsten oxides and the appropriate ternary element or a suitable compound of the ternary element in evacuated silica ampoules at 1373°K for 1 week. The compositions investigated were close to M0.02W0.98O2.80. Oxides with the Mo5O14 structure were found in many systems across the whole of the periodic table, from Li to Bi. Some aspects of the formation of these phases and the way in which they could affect the course of reduction of WO3 to W metal are discussed.  相似文献   

11.
We have prepared SrFe2/3B1/3O3 (B″=Mo, U, Te, and W) double perovskites in polycrystalline form by ceramic methods. Phases with B″=U, Te and W have been studied by X-ray powder diffraction and the results have been compared with neutron diffraction data available for B″=Mo. At room temperature, the stoichiometric samples crystallize in the tetragonal crystal system (space group I4/m, Z=4). Cell parameters when B″=U, Te and W are a=5.6936(1) Å, c=8.0637(1)Å; a=5.5776(1) Å, c=7.9144(3) Å and a=5.5707(3) Å, c=7.9081(5) Å, respectively.The Mössbauer spectra at room temperature for all compounds show hyperfine parameters belonging to two Fe3+ sites located at lattice positions with different degrees of distortion. This is in agreement with diffraction data that indicate that the series of compounds display different degrees of Fe-site disorder, which increases in the following sequence: Mo<U<Te<W.  相似文献   

12.
IR relative integrated intensities and half-widths of rocking (R) and wagging (W) bands of water in MnCl2 · 2H2O and CoCl2 · 2H2O are presented at 300 K and 120 K. Departure of observed intensity into DW/DR from those predicted by the fixed dipole model is attributed to anisotropic dynamic changes in dipole during these oscillations. A quantity representing the variation of this anisotropy between W and R oscillations is computed and its origin is discussed. An increase by 20% to 50% in both DW and DR on lowering the temperature has also been discussed.  相似文献   

13.
The nuclear spin coupling constants1J(183W13C) and in some cases 2J(183W13C) and 3J(183W13C) are determined for 10 tungsten carbene and 9 tungsten carbyne complexes. 1J is of analytical importance, being characteristically greater for WC than for WC bonds. This is due to different hybridisation at the carbon atom, and provides information about bond angles and polarities of WC and WCR units.Substituents R and R' in (CO)5WCRR' and X(CO)4WCR as well as the halogens X lead to minor changes in 1J. These changes are comparable to those of 1J(13C1H) in correspondingly substituted methanes. Unexpectedly 1J in_ creases with X = Cl, Br, I. 2J(183W13C) though being much smaller than 1J reflects different hydridisation at the β carbon atom.  相似文献   

14.
The crystal structure of KP8W40O136, the tenth member of the series KxP4O8(WO3)2m, has been resolved by three-dimensional single-crystal X-ray analysis. The space group is P21c and the cell parameters are a = 19.589(3) Å, b = 7.5362(4) Å, c = 16.970(3) Å and β = 91.864(14)°. The framework is built up from ReO3-type slabs connected through pyrophosphate groups. The structure is compared to those of the other members of the series: although the ReO3-type slabs show a different type of tilting of the WO6 octahedra, the dispersion of WO distances is always higher for the octahedra linked to one or two P2O7 groups and decreases in proportion as W is farther from these groups. The perovskite cages of the slabs are described and compared to those encountered in the structures of WO3 and of the bronzes AxWO3.  相似文献   

15.
Diorganothallium transition metal complexes of the general formula R2Tl—MLn with MLn = M(CO)2LCp (M = Mo, W; L = CO, PPh3) are obtained by protolytic reactions, redistribution reactions or by methatetic reactions, and are characterized spectroscopically and by chemical reactions. For MLn = Cr(CO)3Cp, Fe(CO)2Cp and Co(CO)4 R3Tl and Tl(MLn)3 can always be isolated. In the case of Me2Tl—M(CO)3Cp (M = Mo, W) variable temperature NMR measurements gave evidence for a symmetrisation—redistribution equilibrium 3 R2Tl—MLn—MLn ? R3Tl + Tl(MLn, which generally determines the stability of the diorganomthallium transition metal complexes.  相似文献   

16.
The dithiocarbene complex W(CO)5[C(SCH3)2 reacts with tertiary phosphines, PPh2CH3, PPh(CH3)2, P(C2H5)3 and P(OCH3)3 to form the phosphorane complexes W(CO)5[CH3S)2C-PR3] and with HPPh2 to form the phosphine complex W(CO)5[PPh2[CH(SCH3)2]. Kinetic studies of both types of reactions show that their rates are first order each in W(CO)5[C(SCH3)2] and in the phosphorus ligand. A mechanism involving rate determining phosphorus attack at the carbene carbon followed by rapid rearrangement to the product is consistent with this rate law. Rate constants for the reactions increase with increasing nucleophilicities of the phosphines: P(OCH3)3 < PPh2H < PPh2CH3 ? PPh(CH3)2 < P(C2H5)3. The ΔH values decrease (P(OCH3)3 > PPh2H > PPh2(CH3) > PPh(CH3)2 > P(C2H5)3) as the nucleophilicities of the phosphines increase. The ΔS values (≈-30 e.u.) remain essentially constant for all the reactions. The cyclic dithiorcarbenes W(CO)5[CS(CH2)nS], wheren- 3 or 4, react with PPh2(CH3) to form the cyclic phosphorane complexes, W(CO)5[S(CH2)nSC-PPh2(CH3)]. The 6- and 7- membered cyclic dithiocarbenes also react with PPh2H to form the phosphine complexes, W(CO)5 {PPh2- [CS(CH2)nS(H)]}.  相似文献   

17.
We describe the preparation and structural characterization of four In-containing perovskites from neutron powder diffraction (NPD) and X-ray powder diffraction (XRPD) data. Sr3In2B″O9 and Ba(In2/3B1/3)O3 (B″=W, U) were synthesized by standard ceramic procedures. The crystal structure of the W-containing perovskites and Ba(In2/3U1/3)O3 have been revisited based on our high-resolution NPD and XRPD data, while for the new U-containing perovskite Sr3In2UO9 the structural refinement was carried out from high-resolution XRPD data. At room temperature, the crystal structure for the two Sr phases is monoclinic, space group P21/n, where the In atoms occupy two different sites Sr2[In]2d[In1/3B2/3]2cO6, with a=5.7548(2) Å, b=5.7706(2) Å, c=8.1432(3) Å, β=90.01(1)° for B″=W and a=5.861(1) Å, b=5.908(1) Å, c=8.315(2) Å, β=89.98(1)° for B″=U. The two phases with A=Ba should be described in a simple cubic perovskite unit cell (S.G. Pmm) with In and B″ distributed at random at the octahedral sites, with a=4.16111(1) Å and 4.24941(1) Å for W and U compounds, respectively.  相似文献   

18.
The reactions of the halogenoalkyl compounds [Cp(CO)3W{(CH2)nX}] (Cp = η5-C5H5; n = 3-5; X = Br, I) and [Cp(CO)2(PPhMe2)Mo{(CH2)3Br}] with the nucleophiles Z = CN and gave compounds of the type [Cp(CO)3W{(CH2)nZ}] for the tungsten compounds, whilst cyclic carbene compounds were obtained from the reactions of the molybdenum compound. The reactions of [Cp(CO)3W{(CH2)nBr}] (n = 3, 4) and [Cp(CO)2(PPhMe2)Mo{(CH2)3Br}] with gave [Cp(CO)3W{(CH2)nONO2}] and [Cp(CO)2(PPhMe2)Mo{(CH2)3ONO2}], respectively. The reaction of [Cp(CO)3W{(CH2)nBr}] with AgNO2 gave [Cp(CO)3W{(CH2)nNO2}]. In the solid state the complex [Cp(CO)3W{(CH2)3NO2}] crystallizes in a distorted square pyramidal geometry. In this molecule the nitropropyl chain deviates from the ideal, all-trans geometry as a result of short, non-hydrogen intermolecular N-O?O-N contacts. The reactions of the heterobimetallic compounds [Cp(CO)3W{(CH2)3}MLy] {MLy = Mo(CO)3Cp, Mo(CO)3Cp and Mo(CO)2(PMe3)Cp; Cp = η5-C5(CH3)5} with PPh3 and CO were found to be totally metalloselective, with the ligand always attacking the metal site predicted by the reactions of the corresponding monometallic analogues above with nucleophiles. Thus the compounds [Cp(CO)3W{(CH2)3}C(O)MLz] {MLz = Mo(CO)2YCp, Mo(CO)2YCp and Mo(CO)Y(PMe3)Cp; Y = PPh3 or CO} were obtained. Similarly, the reaction of [Cp(CO)2Fe{(CH2)3}Mo(CO)2(PMe3)Cp] with CO gave only [Cp(CO)2Fe{(CH2)3C(O)}Mo(CO)2(PMe3)Cp]. Hydrolysis of the bimetallic compound, [Cp(CO)3W(CH2)3C(O)Mo(CO)(PPh3)(PMe3)Cp], gave the carboxypropyl compound [Cp(CO)3W{(CH2)3COOH}]. Thermolysis of the compound [Cp(CO)2Fe(CH2)3Mo(CO)3(PMe3)Cp] gave cyclopropane and propene, indicating that β-elimination and reductive processes had taken place.  相似文献   

19.
The Group VIB complexes M(CO)5L have been synthesized for the cases L= cis-1,2-diisopropyldiazene (c-DIPD) with M = Cr, Mo, W, L = trans-1,2-diisopropyldiazene (t-DIPD) with M = Or, W, and L, = 1,2-diisopropylhydrazine (DIPH) with M = Cr, W. Failure to obtain any bimetallic complexes is discussed in terms of steric interactions of these and related complexes. The significance of diazene ligand geometry is demonstrated by the differences in properties of the c-DIPD and t-DIPD complexes. The available evidence indicates that cis diazenes are better ligands than their trans isomers. Complex stability decreases in the order W > Cr > Mo and c-DIPD > t-DIPD. Infrared, visible, and NMR spectra are interpreted in terms of bonding in the complexes. A 30–60 cm?1 reduction of v(NN) of the diazenes upon coordination is attributed to metal-ligand π-bonding with c-DIPD being a better π-acceptor than t-DIPD. The NMR spectra of the c-DIPD complexes are temperature dependent and show that the M(CO)5 moiety undergoes coordination site exchange between the two nitrogen atoms. No exchange occurs in the t-DIPD complexes. Coalescence temperatures of 10, ?48, and 60°C were recorded for the Cr, Mo, and W complexes of c-DIPD respectively, with the Gibbs free energy barriers of 15.0, 11.5 and 15.0 kcal/ mol. A comparison with exchange in other M(CO)5(cis-diazene) complexes is made and the role of the diazene structure on the reaction rate is discussed. The M(CO)5(DIPH) (M = Cr, W) complexes have been oxidized by H2O2/Cu2+ and by activated MnO2 to DIPD complexes in low yield. The tungsten DIPH complex gives only W(CO)5(t-DIPD) but the chromium system gives predominantly Cr(CO)5(c-DIPD).  相似文献   

20.
Single crystal NaxWO3 elastic constants have been measured for x = 0.522, 0.628, 0.695, and 0.74 using Brillouin light scattering. The elasticity of NaxWO3 is found to be similar to that of ReO3. Calculations indicate that an observed decrease in c11 with increasing sodium concentration results from perturbation associated with an increasing lattice constant of strong, covalent WO bonds.  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