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.
The trinuclear cationic complex [Ph4C5(AuPPh3)3+[BF4]- (I) obtained by interaction of C5HPh4AuPPh3 or Ph4C5(AuPPh3)2 with [AuPPh3+[BF4]- in THF was studied by X-ray diffraction. In the presence of benzene, triclinic crystals of the solvate [Ph4C5(AuPPh3)3]+[BF4]-· 2 C6H6 are formed, a = 12.845(6), b = 16.042(8), c = 22.642(11) Å, α = 86.62(4), β = 77.51(4), γ = 76.05(4)°, space group P1, Z = 2, 9494 reflections with I > 2σ (λ(Mo-Kα), θ/2θ scan, 2θ < 46°), with absorption correction R = 0.054. The complex represents a diaurated cation of tetraphenylcyclopentadienyl(triphenylphosphine)gold, containing a triangular Au2C fragment (AuAu 2.820(1) Å) which is bonded to the third Au atom (AuAu 3.021(1) Å), coordinated to the cyclopentadienyl ligand by a bond intermediate between η1(σ) and η3 (AuC 2.21(2), 2.60(2) and 2.71(2) Å).  相似文献   

2.
Nickel-ammonium tetrametaphosphate, Ni(NH4)2P4O12 · 7H2O is triclinic with a = 13.841(3); b = 9.621(5); c = 7.482(2)Å; α = 98.05(4); β = 97.25(4); γ = 103.01(4)°; M = 536.59; V = 947.9Å3; Z = 2; Dx = 1.879 g cm?3; μ = 14.524 cm?1, and space group P1. The crystal structure was solved using 1661 independent reflections measured on a single-crystal diffractometer (Mo). The final R value is 0.056. The two crystallographic independent nickel atoms Ni(1) and Ni(2) are octahedrally coordinated: Ni(1) by four oxygen atoms and two water molecules, Ni(2) by six water molecules. Ni(1), closely connected to two P4O12 rings, forms a complex anion [Ni(P4O12)2(H2O)2]6? which is associated to ammonium polyhedra and [Ni(H2O)6]2+ octahedra. Another interesting feature of this atomic arrangement is the presence of a large channel (10 × 4) Å2 parallel to the c axis. The internal surface of this channel is covered by six zeolitic water molecules.  相似文献   

3.
The reactions (I) Hg2Cl2(s) + Br2(g) and (II) HgCl2(s) + HgBr2(s) have been investigated by an X-ray method. Both the reactions yield two forms of the mixed halide HgClBr, designated as α-HgClBr and β-HgClBr. The cell parameters of the two are as follows:α-HgClBr: a = 6.196 A?, b = 13.12 A?, c = 4.37 A?, z = 4, ? = 5.91 g/cm3. The powder pattern and cell parameters are similar to that of HgCl2. Therefore it is probable that the chlorine atoms, in the linear halogenHghalogen molecules of HgCl2 structure have been replaced by bromines, and since the radius of the bromine atom is larger than that of chlorine, the lattice is larger in this case.β-HgClBr: a = 6.78 A?, b = 13.175 A?, c = 4.17 A?, z = 4, ? = 5.40. These parameters are the same as those reported in the literature for β-Hg(ClBr)2, and its X-ray powder pattern is similar to HgCl2. Therefore this phase also has linear halogenHghalogen molecules but the distribution of Cl and Br atoms is perhaps random.Heating the products (I) and (II) up to the melting point increases the amount of α phase and decreases the β phase, whereas crystallization increases the β phase. DTA study has supported the X-ray findings.  相似文献   

4.
Sc2O2S is hexagonal, P63mmc, a = 3.5196(4) Å, c = 12.519(2) Å, Z = 2, Dc = 3.807 g cm?3, Dm = 4.014 g cm?3, μ(Mo) = 55.51 cm?1. The final R value is 0.038 for 205 symmetry-independent reflections. This scandium oxysulfide has c = 12.52 Å, twice the value found in rare earth oxysulfides. An La2O2S cell combined with its reflection in a (001) mirror gives the Sc2O2S cell.  相似文献   

5.
The tracer diffusion coefficient, D1O, of oxide ions in LaFeO3 single crystal was determined over the temperature range of 900–1100°C by the gas-solid isotopic exchange technique using 18O as a tracer. For the determination of D1O, the depth profile of 18O was measured by means of a secondary ion mass spectrometer (SIMS). The surface exchange reaction was found to be slow and the surface exchange rate constant, k, was determined together with D1O. It was found that D1O at 950°C is proportional to P?0.58O2, where PO2 is an oxygen pressure. The vacancy mechanism was determined for the diffusion of oxide ions from the PO2 dependence. The vacancy diffusion coefficient, DV, for LaFeO3 was nearly the same as that for LaCoO3 at the same temperature. The activation energy for migration of oxide ion vacancies was 74 kJ · mole?1 for both oxides.  相似文献   

6.
Reaction of methyl 2-acetamido-4,6-O-benzylidene-2-deoxy-α-D-ribo-hexopyranosid-3-ulose with Me3SiCN afforded methyl 2-acetamido-4,6-O-benzylidene-3-C-cyano-2-deoxy-3-O-trimethylsilyl-α-D-allo- Reaction of ethyl 4,6-di-O-acetyl-2,3-anhydro-α-D-mannopyranoside with Me3SiCN gave the corresponding ethyl 4,6-di-O-acetyl-2-C-cyano-2-deoxy-α-D-glucopyranoside. Reaction of methyl 4,6-O-benzylidene-2,3-anhydro-α-D-allopyranoside or methyl 4,6-O-benzylidene-2,3-di-O-tosyl-α-D-glucopyranoside with Me3SiCN at - 75° or - 50° gave the corresponding methyl 6-O-[(R)-cyano phenyl methyl]-α-D-glyco-pyranosides with high or total regio and stereoselectivity.  相似文献   

7.
Combination of in-situ generated monocopperII-substituted Keggin polyoxoanions with copperII-organoamine complexes under hydrothermal conditions results in seven inorganic-organic composite polyoxotungstates [Cu(en)2(H2O)]2{[Cu(en)2][α-PCuW11O39Cl]}·3H2O (1), {[Cu(en)2(H2O)][Cu(en)2]2[α-PCuW11O39Cl]}·6H2O (2), {[Cu(en)2(H2O)]2[Cu(en)2][α-XCuW11O39]}·5H2O (3/4, X=SiIV/GeIV), {[Cu(deta)(H2O)2]2[Cu(deta)(H2O)][α-XCuW11O39]}·5H2O (5/6, X=GeIV/SiIV) and [Cu(dap)2]2{[Cu(dap)2]2[Cu(dap)2][α-PCuW11O39]2} (7) (en=ethylenediamine, dap=1,2-diaminopropane and deta=diethylenetriamine). 1 is an isolated structure whereas 2 is a 1-D chain structure, but both contain [α-PCuW11O39Cl]6− polyoxoanions. 3-6 contain the 1-D linear chains made up of [α-XCuW11O39]6− polyoxoanions in the pattern of -A-A-A- (A=[α-XCuW11O39]6−), while 7 demonstrates the first 1-D zigzag chain constructed from [α-PCuW11O39]210− polyoxoanions via [Cu(en)2]2+ bridges in the pattern of -A-B-A-B- (A=[α-PCuW11O39]210−, B=[Cu(en)2]2+). The successful syntheses of 1-7 can provide some experimental evidences that di-/tri-/hexa-vacant polyoxoanions can be transformed into mono-vacant Keggin polyoxoanions under hydrothermal conditions.  相似文献   

8.
A surface-hopping model is applied to near-resonant electronic energy transfer in the NFBi and O2I systems. Multiple surface crossings occur in NFBi at ca. 8 A, corresponding well with measured transfer cross section of 200 A2. A Landau-Zener model yields the temperature dependence of the thermally averaged cross section for the laser pumping reaction, O*2(a1Δ) + I(2P32) → O2(X3Σ?g) + 1*(2P12).  相似文献   

9.
Deactivation rate constants of spin-orbital excited Br atoms in the reactions Br(2P12) + O2 → Br(2P32) + O2 (k1), and Br(2P12) + NO → Br(2P32) + NO (k4) have been measured with a photodissociative IBr laser on the electronic transition 2P12?2P32 in the Br atom (λ = 2.7 μm). The values obtained are (6.4 ± 1.8) × 10?14 cm3 s?1 and (1.9 ± 0.6) × 10?12 cm3 s?1, respectively. Comparison with published data leads to the conclusion that, contrary to a widely accepted point of view, the high rate constants for the quenching of excited halogen atoms are due to resonant energy transfer processes and not to the paramagnetic nature of the quencher.  相似文献   

10.
The dianion [RuIr4(CO)15]2- has been obtained by reductive carbonylation of mixtures of Ir4(CO)12 and RuCl3 · χ H2O, and the bis(triphenylphosphine)-iminium salt has been characterized by single-crystal X-ray diffraction techniques. Crystal data: [((C6H5)3P)2N]2[RuIr4(CO)15], space group P1 (Z  2), a  11.425(3), b  14.141(2), c  25.979(5) Å, α  84.55(1), β  83.53(2), γ  80.71(2)°. The mixed-metal cluster has a structure with an elongated trigonal bipyramidal array of metal atoms in which Ru occupies an apical position. The anion is unstable in vacuum or in an N2 atmosphere yielding predominantly another mixed-metal species which is not as yet fully characterized. Upon reexposure to CO, this latter species is converted back to [RuIr4(CO)15]2-, plus additional products.  相似文献   

11.
The tracer diffusion coefficient, D1O, of oxide ions in LaCoO3 single crystal was determined over the temperature range of 700–1000°C by a gas-solid isotopic exchange technique using 18O tracer. For the determination, two methods, the gas phase analysis and the depth profile measurement, were employed. Under an oxygen pressure of 34 Torr, the temperature dependence of D1O in LaCoO3 was expressed by
D1O(cm2·sec?1) = 3.63 × 104exp? (74 ± 5)kcal · mole?1RT
D1O at 950°C was found to be proportional to P?0.35O2. The diffusion of oxide ions occurs through a vacancy mechanism. The activation energy for the migration of oxide ion vacancies was estimated as 18 kcal · mole?1.  相似文献   

12.
The reduction of permanganate by oxalate in the presence of manganese(II) ion in acidic media is described. All reactions were run at 525 nm and constant ionic strength 1.0 M. The reaction was found to obey the rate expression —d[MnO4-]dt = k [Mn2+] [C2O42-]2 [MnO4-] [H+]-2 = k' [MnC2O4] [MnO4-]. The values of k and k' were shown to be 5.4 × 104 M-1 s-1 and 8.2 × 104 M-1 s-1, respectively. Reaction rate methods for the determination of manganese(II) and oxalic acid are reported. The rate of disappearance of permanganate was monitored automatically and related directly to manganese-(II) and oxalic acid concentrations. Manganese(II) in the ranges 1–10 × 10-4 M and 1–10 × 10-3 M and oxalic acid in the range 0–20 μg ml-1 can be determined very rapidly with a precision of 1–2%.  相似文献   

13.
Conditions for the quantitative coulometric titration of iodide and iodine with electrolytically generated hypobromite in the presence of borax buffer have been established. Iodide and iodine are oxidized to iodate. The method, with biamperometric indication of the equivalence point, was successfully applied for a wide range of iodide concentrations (6.21–2115μg with reliability intervals of ±0.21–±11μg) and iodine concentrations (24.26–3311μg with reliability intervals of ±0.36–±11.7μg). The determinations are accurate and sensitive even in the presence of large amounts of bromides and chlorides (Br?I?= 1.2·106 and Cl?I?=4.0·103), as well as in the presence of oxidizing agents such as IO3?, BrO3? and CrO42? (IO3?/I2)=3.2·105, IO3?/I2=3.1·103, BrO-3/I2=1.1·104 and CrO2-4/I2=1.0·104, as was confirmed by statistical tests. The oxidation mechanism under the conditions of coulometric titrations is discussed.  相似文献   

14.
From measurements of the heats of iodination of CH3Mn(CO)5 and CH3Re(CO)5 at elevated temperatures using the ‘drop’ microcalorimeter method, values were determined for the standard enthalpies of formation at 25° of the crystalline compounds: ΔHof[CH3Mn(CO)5, c] = ?189.0 ± 2 kcal mol?1 (?790.8 ± 8 kJ mol?1), ΔHof[Ch3Re(CO)5,c] = ?198.0 ± kcal mol?1 (?828.4 ± 8 kJ mo?1). In conjunction with available enthalpies of sublimation, and with literature values for the dissociation energies of MnMn and ReRe bonds in Mn2(CO)10 and Re2(CO)10, values are derived for the dissociation energies: D(CH3Mn(CO)5) = 27.9 ± 2.3 or 30.9 ± 2.3 kcal mol?1 and D(CH3Re(CO)5) = 53.2 ± 2.5 kcal mol?1. In general, irrespective of the value accepted for D(MM) in M2(CO)10, the present results require that, D(CH3Mn) = 12D(MnMn) + 18.5 kcal mol?1 and D(CH3Re) = 12D(ReRe) + 30.8 kcal mol?1.  相似文献   

15.
The reaction of {[UO2(HCOO)2(H2O)]} with diaza-18-crown-6 (DA18C6 = C12H26O4N2) in aqueous ethanol in the presence of formic acid yields the complexes {[DA18C6H2]·[UO2(HCOO)3]2} (I), [DA18C6H2]·[UO2(HCOO)4] (II), and [DA18C6H2]·(HCOO)2·(H2O)2 (III). The complexes are characterized using IR spectroscopy, chemical analysis, and powder X-ray diffraction. From the comparison of the structural and spectral characteristics of [DA18C6H2]·An2·(H2O)2n (where An = Cl?,NO 3 ? ,HCOO?,HSO 4 ? ; n = 0.1), correlations are derived between the conformation of the [DA18C6H2]2+ units and the conformation-sensitive frequencies. On the basis of these correlations, the conformations of the N+CCO and OCCO units were determined in the diazonia cations of compounds I and II and in [DA18C6H2]·[UO2(NO3)4]; the latter was prepared previously by reacting [UO2(NO3)2(H2O)2]·(H2O)4 with DA18C6 in ethanol in the presence of nitric acid.  相似文献   

16.
Microcalorimetric measurements at 520–523 K of the heats of thermal decomposition and of iodination of bis-(benzene)molybdenum and of bis-(toluene)tungsten have led to the values (kJ mol?): ΔHof[Mo(η-C6H6)2, c] = (235.3 ± 8) and ΔHof[W(η6-C7H8)2, c] = (242.2 ± 8) for the standard enthalpies of formation at 25°C. The corresponding ΔHof(g) values, using available and estimated enthalpies of sublimation, are (329.9 ± 11) and 352.2 ± 11) respectively, from which the metalligand mean bond-dissociation enthalpies, D(Mo—benzene) = (247.0 ± 6) and D(W—toluene) = (304.0 ± 6) kJ mol?1, are derived.  相似文献   

17.
The synthesis and mechanism of formation of phosphonium salts of the type [R3P+CFXY]Z? (where X = F, Cl, Br; Y = Br, Cl; Z = Br, Cl), bis-phosphonium salts of the type [R3P+CF2P+R3]2Br?, and phosphoranium salts of the type [R3P+C?FP+R3]X? (X = Br, Cl) will be presented. The applicability of these substrates in the generation of useful nucleophilic or electrophilic synthetic intermediates will be discussed.  相似文献   

18.
The opto-acoustic spectrum of I2 in the presence of various quenching gases — NO, O2, CH3I, SO2, C3HS, N2, and He — has been studied. Of these, the I2/O2 spectrum is quite different due to the near-resonant energy transfer I(2P12) + O2(3Σ) → I(2P32) + O2(IΔ), wherein the resistance of the O2((IΔ) species to collisional relaxation severely distorts the acoustic signal. The photochemical production of excited 2P12 iodine atoms commences at wavelengths considerably longer than the dissociation limit of the I2B? state.  相似文献   

19.
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   

20.
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   

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