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 627 毫秒
1.
Early stage oxidation of a dilute-depleted uranium-molybdenum alloy was analysed in situ under ultra-high vacuum conditions by AES and XPS. At the equivalent of less than 300 ns at 1-atm O2, U-5Mo oxidizes to form stoichiometric UO2. No molybdenum oxidation is observed. After an oxygen dose of approximately 39 L, the oxide layer approached a limiting thickness of approximately 2.4 nm. The oxidation kinetics followed a logarithmic rate law, with the best fit to the experimental data for the oxide thickness, d, being given by d = 1.26 log(0.12t + 0.56). Changes in oxygen KLL and 1s peak positions associated with transformation from chemisorbed oxygen to metal oxide were observed at similar oxygen doses of 2.3 and 2.6 L O2 by AES and XPS, respectively, which opens up the possibility of using well-characterized XPS chemical information to inform Auger peak shifts.  相似文献   

2.
以VOPO4.2H2O为原料制备了钒磷氧化物催化剂,考察了镍掺杂(1%,2%和5%)对该催化剂的影响.低掺杂量的Ni明显提高了活性晶格氧物种O-的数量,降低了V5 和V4 相的还原峰温.粉末X射线衍射、程序升温还原和化学分析结果表明,高掺杂量的Ni促使V5 物相生成并抑制V4 物相出现.高含量与V5 相关的氧物种会降低正丁烷的转化率,但会提高马来酸酐的选择性.  相似文献   

3.
A polyoxoniobate, [Cu(en)2]4{[Nb6O19H2]K(H2O)5}2 ? (H2en) ? 17H2O (en = ethylenediamine) (1), has been synthesized and characterized by elemental analysis, IR, XPS, TGA, and single-crystal X-ray diffraction. Compound 1 crystallizes in the triclinic system, space group P 1, with a = 12.3533(16) Å, b = 12.7188(16) Å, c = 29.626(4) Å, α = 93.235(2)°, β = 96.094(1)°, γ = 106.098(2)°, V = 4429.0(10) Å3, Z = 2. The polyoxoanion is composed of a Lindqvist-type [Nb6O19H2]6? dimer bi-bridged via two K+. K+ is 10-coordinate with 10 oxygens, three from one [Nb6O19H2]6?, one from a terminal oxygen of another [Nb6O19H2]6? moiety, and the other six from water molecule. Adjacent dimeric polyoxoanions are linked to form an infinite 1-D chain via O–H ··· O hydrogen-bonding interactions which exist between the two water trimers and the dimeric polyoxoanions.  相似文献   

4.
The template synthesis of ethylenediamine ( 1 ) with 2-acetylcyclopentanone ( 2 ) and [Cu(OAc)2 · H2O] ( 5 ) produced [Cu(1-(2-cC5H6(O))C(Me)NCH2)2)] ( 6 ) in 82 % yield. Reaction of 5 with bis(benzoylacetone)diethylenetriamine ( 7 , = L H)[1] gave [Cu(μ-OAc)( L )(H2O)]2 ( 8 ). The solid-state structures of 6 and 8 were determined confirming that 8 possesses intra- and intermolecular hydrogen bonds resulting in a dimer formation. The thermal behavior of 6 – 8 was studied by TG and TG-MS. Under oxygen CuO was formed, whereas under Ar Cu/Cu2O ( 6 ) or Cu ( 8 ) was obtained. Complex 6 was used as CVD precursor for Cu and Cu-oxide deposition (substrate temp., 400–500 °C, N2, 60 mL · min–1; O2, 60 mL · min–1; pressure, 0.87–1.5 mbar). The as-obtained deposits show separated particles of different appearance at the substrate surface as evidenced by SEM. Non-volatile 8 was applied as spin-coating precursor for Cu and CuO formation [conc. 0.25 mol · L–1; volume 0.2 mL; 3000 rpm; depos. time 2 min; heating rate 50 K · min–1; holding time 60 min (Ar), 120 min (air) at 800 °C]. The samples on silicon consist of granulated particles (Ar) or are non-dense with a grainy topography (air). EDX and XPS measurements confirmed the formation of Cu (Ar) or CuO (O2) with up to 13 mol-% C impurity.  相似文献   

5.
NARP (Nitric oxide-Ammonia Rectangular Pulse) technique was applied for the in-situ measurement of the number of surface V=O species of V2O5 catalysts. When V2O5 catalyst was exposed to flowing benzene/O2/N2 mixture gas at 573–673 K, the surface concentration of V=O species decreased and finally reached a steady state which was dependent on the [O2]/[Benzene] ratio in the feed gas. It was confirmed by XPS spectra that the decrease in the surface V=O species is attributed to the reduction of the V2O5 surface. Although calcination of V2O5 at above 873 K reduced the surface concentration of V=O species, it recovered after a cyclic reduction-oxidation treatment at 673 K, which modifies the near-surface layer of V2O5. The dependence of the surface V=O species on calcination temperature was correlated with the flatness of the surface and the diffusion of lattice oxygen in near-surface layer.  相似文献   

6.
The pharmaceutically active compound atenolol, a kind of $\beta$-blockers, may result in adverse effects both for human health and ecosystems if it is excreted to the surface water resources. To effectively remove atenolol in the environment, both direct and indirect photodegradation, driven by sunlight play an important role. Among indirect photodegradation, singlet oxygen (1O2), as a pivotal reactive species, is likely to determine the fates of atenolol. Nevertheless, the kinetic information on the reaction of atenolol with singlet oxygen has not been well investigated and the reaction rate constant is still ambiguous. Herein, the reaction rate constant of atenolol with singlet oxygen is investigated directly through observing the decay of the 1O2 phosphorescence at 1270 nm. It is determined that the reaction rate constant between atenolol and 1O2 is 7.0×105 (mol/L)$^{-1}\cdot$s-1 in D2O, 8.0×106 (mol/L)$^{-1}\cdot$s-1 in acetonitrile, and 8.4×105 (mol/L)$^{-1}\cdot$s-1 in EtOH, respectively. Furthermore, the solvent effects on the title reaction were also investigated. It is revealed that the solvents with strong polarity and weak hydrogen donating ability are suitable to achieve high rate constant values. These kinetics information on the reaction of atenolol with singlet oxygen may provide fundamental knowledge to the indirect photodegradation of $\beta$-blockers.  相似文献   

7.
A new organic–inorganic hybrid material constructed from octamolybdate anion and neutral dinuclear copper(I) units, H4{[Cu2(ophen)2]2[Mo8O26]}[Cu2(ophen)2] · H2O (1) (Hophen =2-hydroxy-1,10-phenanthroline), has been prepared under hydrothermal condition and characterized by elemental analysis, IR, XPS, TGA and single-crystal X-ray diffraction. Compound 1 crystallizes in the triclinic system, space group P 1, with a = 9.9091(8), b = 13.3981(8), c = 14.8266(10) Å, α = 84.6310(10), β = 83.0620(10)°, γ = 77.7800(10), V = 1905.0(2) Å3, Z = 1. Compound 1 contains a centrosymmetric polyoxoanion {[Cu2(ophen)2]2[Mo8O26]}4?, in which the β-[Mo8O26]4? is bisupported by two copper(I) coordination groups through the terminal oxygen atoms. The discrete molecules of 1 are extended into a 3-D supramolecular array through C–H ··· O hydrogen bonds and strong aromatic π–π stacking contacts.  相似文献   

8.
Thin films of vanadium oxide were grown on vanadium metal surfaces (i) in air at ambient conditions, (ii) in 5 mM H2SO4 (aq), pH 3, (iii) by thermal oxidation at low oxygen pressure (10?5 mbar) at temperatures between 350 and 550 °C and (iv) at near‐atmospheric oxygen pressure (750 mbar) at 500 °C. The oxide films were investigated by atomic force microscopy (AFM), X‐ray photoelectron spectroscopy (XPS), X‐Ray diffraction (XRD) and Rutherford backscattering spectrometry (RBS) and nuclear reaction analysis (NRA). The lithium intercalation properties were studied by cyclic voltammetry (CV). The results show that the oxide films formed in air at room temperature (RT), in acidic aqueous solution, and at low oxygen pressure at elevated temperatures are composed of V2O3. In air and in aqueous solution at RT, the oxide films are ultra‐thin and hydroxylated. At 500 °C, nearly atmospheric oxygen pressure is required to form crystalline V2O5 films. The oxide films grown at pO2 = 750 mbar for 5 min are about 260‐nm thick, and consist of a 115‐nm outer layer of crystalline V2O5. The inner oxide is mainly composed of VO2. For all high temperature oxidations, the oxygen diffusion from the oxide film into the metal matrix was considerable. The oxygen saturation of the metal at 450 °C was found, by XPS, to be 27 at.% at the oxide/metal interface. The well‐crystallized V2O5 film, formed by oxidation for 5 min at 500 °C and 750 mbar O2, was shown to have good lithium intercalation properties and is a promising candidate as electrode material in lithium batteries. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

9.
New dinuclear pentacoordinate molybdenum(V) complexes, [Mo2VO3L2] [L = thiosemicarbazonato ligand: C6H4(O)CH:NN:C(S)NHR′ and C10H6(O)CH:NN:C(S)NHR′; R′ = H, CH3, C6H5) were obtained either by oxygen atom abstraction from MoVIO2L with triphenylphosphine or by using [Mo2O3(acac)4] in the reaction with the corresponding ligands H2L. Crystal and molecular structure of [Mo2O3{C6H4(O)CH:NN:C(S)NHC6H5}2] · CH3CN has been determined by the single‐crystal X‐ray diffraction method.  相似文献   

10.
La3OCl[AsO3]2: A Lanthanum Oxide Chloride Oxoarsenate(III) with a “Lone‐Pair” Channel Structure La3OCl[AsO3]2 was prepared by the solid‐state reaction between La2O3 and As2O3 using LaCl3 and CsCl as fluxing agents in evacuated silica ampoules at 850 °C. The colourless crystals with pillar‐shaped habit crystallize tetragonally (a = 1299.96(9), c = 558.37(5) pm, c/a = 0.430) in the space group P42/mnm (no. 136) with four formula units per unit cell. The crystal structure contains two crystallographically different La3+ cations. (La1)3+ is coordinated by six oxygen atoms and two chloride anions in the shape of a bicapped trigonal prism (CN = 8), whereas (La2)3+ carries eight oxygen atoms and one Cl? anion arranged in the shape of tricapped trigonal prism (CN = 9). The isolated pyramidal [AsO3]3? anions (d(As–O) = 175–179 pm) consist of three oxygen atoms (O2 and two O3), which surround the As3+ cations together with the free, non‐binding electron pair (lone pair) Ψ1‐tetrahedrally (?(O–As–O) = 95°, 3×). One of the three crystallographically independent oxygen atoms (O1), however, is exclusively coordinated by four (La2)3+ cations in the shape of a real tetrahedron (d(O–La) = 236 pm, 4×). These [(O1)(La2)4]10+ tetrahedra form endless chains in the direction of the c axis through trans‐edge condensation. Empty channels, constituted by the lonepair electrons of the Cl? anions and the As3+ cations in the Ψ1‐tetrahedral oxoarsenate(III) anions [AsO3]3?, run parallel to [001] as well.  相似文献   

11.
The reaction of benzyl 2-amino-4,6-O-benzylidene-2-deoxy-α-D -glucopyranoside (HL) with the metal salts Cu(ClO4)2 ⋅ 6 H2O and Ni(NO3)2 ⋅ 6 H2O affords via self-assembly a tetranuclear μ4-hydroxido bridged copper(II) complex [(μ4-OH)Cu4(L)4(MeOH)3(H2O)](ClO4)3 ( 1 ) and a trinuclear alcoholate bridged nickel(II) complex [Ni3(L)5(HL)]NO3 ( 2 ), respectively. Both complexes crystallize in the acentric space group P21. The X-ray crystal structure reveals the rare (μ4-OH)Cu4O4 core for complex 1 which is μ2-alcoholate bridged. The copper(II) ions possess a distorted square-pyramidal geometry with an [NO4] donor set. The core is stabilized by hydrogen bonding between the coordinating amino group of the glucose backbone and the benzylidene protected oxygen atom O4 of a neighboring {Cu(L)} fragment as hydrogen-bond acceptor. For complex 2 an [N4O2] donor set is observed at the nickel(II) ions with a distorted octahedral geometry. The trinuclear isosceles Ni3 core is bridged by μ3-alcoholate O3 oxygen atoms of two glucose ligands. The two short edges are capped by μ2-alcoholate O3 oxygen atoms of the two ligands coordinated at the nickel(II) ion at the vertex of these two edges. Along the elongated edge of the triangle a strong hydrogen bond (244 pm) between the O3 oxygen atoms of ligands coordinating at the two relevant nickel(II) ions is observed. The coordinating amino groups of the these two glucose ligands are involved in additional hydrogen bonds with O4 oxygen atoms of adjacent ligands further stabilizing the trinuclear core. The carbohydrate backbones in all cases adopt the stable 4C1 chair conformation and exhibit the rare chitosan-like trans-2,3-chelation. Temperature dependent magnetic measurements indicate an overall antiferromagnetic behavior for complex 1 with J1=−260 and J2=−205 cm−1 (g=2.122). Compound 2 is the first ferromagnetically coupled trinuclear nickel(II) complex with JA=16.4 and JB=11.0 cm−1 (g1,2=2.183, g3=2.247). For the high-spin nickel(II) centers a zero-field splitting of D1,2=3.7 cm−1 and D3=1.8 cm−1 is observed. The S=3 ground state of complex 2 is consistent with magnetization measurements at low temperatures.  相似文献   

12.
The reaction of (NH4)2PbCl6 and fuming sulfuric acid (65 % SO3) in a sealed glass tube at 250 °C led to colorless single crystals of Pb[S3O10] (orthorhombic, Pbcn, Z = 4, a = 10.9908(4), b = 8.5549(3), c = 8.0130(3) Å, V = 753.42(5) Å3). The compound shows a three‐dimensional linkage of the tenfold oxygen coordinated Pb2+ ions and exhibits the unusual trisulfate anion, [S3O10]2–, that consists of three vertex connected [SO4] tetrahedra. The distances S–O within the S–O–S bridges of the anion are remarkable asymmetric with distances of 155 and 169 pm, respectively. This structural feature is well reproduced by calculations on a PBE0/cc‐pVTZ and a MP2/cc‐pVTZ level of theory. Similar calculations allow also for an inspection of the yet unknown corresponding acid, H2S3O10. Also for this acid non‐symmetric S–O–S bridges are predicted. The thermal behavior of Pb[S3O10] is characterized by the loss of two equivalents of SO3 at low temperature and the decomposition of intermediate Pb[SO4] at higher temperature.  相似文献   

13.
Metal Derivatives of Molecular Compounds. VIII. catena-Poly[(2,5,8-trioxanonane-O2,O5) lithium-methylphosphanide] — a Compound with a meso-Helix Structure Studies of Fritz et al. [10] showed methylphosphane to be lithiated at ?60°C in 1,2-dimethoxyethane or bis(2-methoxyethyl) ether solution by stoichiometric amounts of lithium n-butanide in n-hexane. After removing the hydrocarbons almost completely by distillation and cooling the solutions to ?60°C again, colourless square crystals of (1,2-dimethoxyethane-O,O′)lithium ( 1 ) and (2,5,8-trioxanonane-O2,O5)lithium methylphosphanide ( 2 ) precipitate. As shown by an X-ray structure determination (monoclinic, P21/n; a = 805.5(1); b = 1820.6(2); c = 851.5(1) pm; β = 116.76(1)° at ?100 ± 3°C; Z = 4 formula units; R = 0.034) complex 2 forms a polymer which has the shape of an up to now scarcely noted meso-helix. Four-coordinated lithium is bound to two phosphorus (P? Li 252.9 and 253.2 pm; P? Li? P 131.8°; Li? P? Li 132.1°) and to two oxygen atoms (Li? O 203.9 and 206.8; O …? O 270.7 pm; O? Li? O 82.5°) of the inherently tridentate 2,5,8-trioxanonane ligand. As compared to the standard value (185 pm) the P? C distance (187.4 pm) is slightly lengthened. Structure determinations of (2,5,8-trioxanonane-O2,O5,O8) lithium 1-(phenylsulfonyl)alkyl compounds published some years ago [26, 27], allow a comparison of molecular parameters characteristic for the twofold or threefold coordinating chelate ligand.  相似文献   

14.
The new quinary fluoride‐rich rubidium scandium oxosilicate Rb3Sc2F5Si4O10 was obtained from mixtures of RbF, ScF3, Sc2O3 and SiO2 in sealed platinum ampoules after seventeen days at 700 °C. The colourless compound crystallises orthorhombically in space group Pnma with a = 962.13(5), b = 825.28(4), c = 1838.76(9) pm and Z = 4. For the oxosilicate partial structure, [SiO4]4– tetrahedra are connected in (001) by vertex‐sharing to form corrugated unbranched vierer single layers ${2}\atop{{\infty}}$ {[Si4O10]4–} (d(Si–O) = 158–165 pm, ∠(O–Si–O) = 103–114°, ∠(Si–O–Si) = 125–145°) containing six‐membered rings. Similar oxosilicate layers with 63‐net topology are well‐known for the mineral group of micas or in sanbornite Ba2Si4O10. Regarding other systems, identical tetrahedral layers can be found in the synthetic borophosphate Mg(H2O)2[B2P2O8(OH)2] · H2O. The Sc3+ cations are coordinated octahedrally by four F and two O2– anions. These cis‐[ScF4O2]5– octahedra (d(Sc–F) = 200–208 pm, d(Sc–O) = 202–205 pm) share one equatorial and two apical F anions with others to build up slightly corrugated ${1}\atop{{\infty}}$ {[Sc2F${t}\atop{2/1}$ F${v}\atop{6/2}$ O${t}\atop{4/1}$ ]7–} double chains along [010]. These are linked with the oxosilicate layers via two oxygen vertices to construct a three‐dimensional framework with cavities apt to host the three crystallographically independent Rb+ cations with coordination numbers of eleven, twelve and thirteen.  相似文献   

15.
In aqueous solution N, N′-bis-(4-(5)-imidazolylmethyl)-ethylenediamine-cobalt (II) (CoIMEN2+) takes up molecular oxygen giving μ-dioxygen-μ-hydroxo-bis-[N, N′-bis-(4-(5)-imidazolylmethyl)-ethylenediamine]-dicobalt (II). (Co IMEN)2 O2 (OH)3+ is exceptionally stable against irreversible autoxydation to CoIII species. Its absorption spectrum is very similar to that of the known analogous complex (CoTRIEN)2 O2 (OH)3+. The kinetics of formation and dissociation of (CoIMEN)2O2(OH)3+ are studied by spectrophotometry and with an oxygen specific electrode. The rate of the forward reaction is described by vf = [CoIMEN2+]2 · [O2] · (k1 + k2 · [OH?]) with k1 = 9 · 104 M?2 s?1 and k2 = 1 · 1012M?3 S?1, at 25° and I = 0,2. A mechanism including hydroxylated as well as nonhydroxylated intermediates is proposed. Dissociation is preceeded by protonation of the oxygen adduct. At pH 1–2 the rate of dissociation is independent of [H+] and follows first order kinetics: vD = k3 · [(CoIMEN)2O2(OH)3+] with k3 = 2.15 · 10?2 S?1.  相似文献   

16.
Peroxynitrates (RO2NO2), in particular acyl peroxynitrates (R = R′C(O) with R′ = alkyl), are prominent constituents of polluted air. In this work, a systematic study on the thermal decomposition rate constants of the first five members of the series of homologous R′C(O)O2NO2 with R′ = CH3 ( =PAN), C2H5, n‐C3H7, n‐C4H9, and n‐C5H11 is undertaken to verify the conclusions from previous laboratory data (Grosjean et al., Environ. Sci. Technol. 1994, 28, 1099–1105; Grosjean et al., Environ. Sci. Technol. 1996, 30, 1038–1047; Bossmeyer et al., Geophys. Res. Lett. 2006, 33, L18810) that the longer chain peroxynitrates may be considerably more stable than PAN. Experiments are performed in a temperature‐controlled, evacuable 200 L‐photoreactor made from quartz. n‐Acyl peroxynitrates are generated by stationary photolysis of mixtures of molecular bromine, O2, NO2, and the corresponding parent aldehydes, highly diluted in N2. Thermal decomposition of R′C(O)O2NO2 is initiated by the addition of an excess of NO. First‐order decomposition rate constants k1 of the reactions R′C(O)O2NO2 (+M) → R′C(O)O2 + NO2 (+M) are derived at 298 K and a total pressure of 1 bar from the measured loss rates of R′C(O)O2NO2, correcting for wall loss of R′C(O)O2NO2 and several percentages of reformation of R′C(O)O2NO2 by the reaction of R′C(O)O2 radicals with NO2. With increasing chain length of R′, k1(298 K) slightly decreases from 4.4 × 10?4 s?1 (R′ = CH3) to 3.7 × 10?4 s?1 (R′ = C2H5), leveling off at (3.4 ± 0.1) × 10?4 s?1 for R′ = n‐C3H7, n‐C4H9, and n‐C5H11. Temperature dependencies of k1 were measured for CH3C(O)O2NO2 and n‐C5H11C(O)O2NO2 in the temperature range 289–308 K, resulting in the same activation energy within the statistical error limits (2σ) of 0.9 and 1.5 kJ mol?1, respectively. A few experiments on n‐C6H13C(O)O2NO2, n‐C7H15C(O)O2NO2, and n‐C8H17C(O)O2NO2 were also performed, but the results were considered to be unreliable due to strong wall loss of the peroxynitrate and possible complications caused by radical‐sinitiated side reactions.  相似文献   

17.
The interaction between oxygen and polycrystalline palladium (Pd(poly)) at \(P_{O_2 } \) = 2.6 × 10?6–10 Pa and T = 300–1300 K was studied by the thermal desorption (TD) method. The interaction between O2 and Pd(poly) is governed by the O2 pressure and the sample temperature. At low pressures of \(P_{O_2 } \) (≤1.3 × 10?5 Pa), O2 is chemisorbed dissociatively on the Pd(poly) surface. During chemisorption, the Oads-surface bond energy and the O2 sticking coefficient gradually decrease as the surface coverage θ increases. At \(P_{O_2 } \) ≥ 10?2 Pa and T ≤ 500 K, after the saturation of the Oads layer (θ ~ 0.5), Oads atoms penetrate under the surface layer of the metal to form surface palladium oxide. At \(P_{O_2 } \) ≥ 1 Pa and T > 500 K, after the saturation of the surface oxide film 2 ML in thickness (n ~ 2), Oads atoms penetrate into the oxide film and then into the subsurface palladium layer and diffuse deep into the metal bulk. As a result, the oxygen uptake at 700 K is n ~ 50. Upon heating, the surface oxides decompose, desorbing O2, which gives rise to a low-temperature TD peak with T max = 715 K. The release of oxygen inserted in the subsurface layers of palladium shows itself as a distinct high-temperature TD peak with T max ≥ 750 K.  相似文献   

18.
Oxo-centered mixed-valence trinuclear iron dicarboxylic acid complex iron fumarate [Fe3O(O2CCH=CHCO2)3(H2O)3nH2O (n = 18-19), have been synthesized firstly. Variable temperature Mössbauer spectroscopy has been carried out to elucidate the rate of intramolecular electron transfer. It was found that the complex showed a temperature dependent mixed-valence state. At low temperature two quadrupole split doublets were observed corresponding to high spin Fe(III) and high spin Fe(II) state and a complete averaged valence state was observed at about 270 K. At temperatures between 200 and 270 K the spectra showed relaxation effect as the electron transfer rate was comparable to the Mössbauer timescale (-10–7 s).  相似文献   

19.
The desorption of oxygen from polycrystalline palladium (Pd(poly)) was studied using temperature-programmed desorption (TPD) at 500–1300 K and the amounts of oxygen absorbed by palladium (n) from 0.05 to 50 monolayers. It was found that the desorption of O2 from Pd(poly), which occurred from a chemisorbed oxygen layer (Oads), in the release of oxygen from a near-surface metal layer in the course of the decomposition of PdO surface oxide, and in the release of oxygen from the bulk of palladium (Oabs), was governed by repulsive interactions between Oads atoms and the formation and decomposition of Oads-Pd*-Oabs structures (Pd* is a surface palladium atom). At θ ≤ 0.5, the repulsive interactions between Oads atoms (ɛaa = 10 kJ/mol) resulted in the desorption of O2 from Pd(poly) at 650–950 K. At 0.5 ≤ n ≤ 1.0, the release of inserted oxygen from a near-surface palladium layer occurred during TPD in the course of the migration of Oabs atoms to the surface and the formation-decomposition of Oads-Pd*-Oabs structures. As a result, the desorption of O2 occurred in accordance with a first-order reaction with a thermal desorption (TD) peak at T max ∼ 700 K. At 1.0 ≤ n ≤ 2.0, the decomposition of PdO surface oxide occurred at a constant surface cover-age with oxygen during TPD in the course of the formation-decomposition of Oads-Pd*-Oabs structures. Because of this, the desorption of O2 occurred in accordance with a zero-order reaction at low temperatures with a TD peak at T max ∼ 675 K. At 1.0 ≤ n ≤ 50, oxygen atoms diffused from deep palladium layers in the course of TPD and arrived at the surface at high temperatures. As a result, O2 was desorbed with a high-temperature TD peak at T > 750 K.  相似文献   

20.
Investigations into Tin(IV) Alkoxides. II. Isolation and Characterization of the Compound Sn3O(OiBu)1010 · 2i-BuOH. The First Example of a Partially Hydrolized Tin(IV) Alkoxide The partial hydrolysis product Sn3O(OiBu)10 · 2i-BuOH was obtained by slow hydrolysis of the reaction product of tin tetrachloride with sodium isobutoxide. The compound forms colourless, moisture sensitive crystals, which easily release the coordinated solvent molecules in dry air. Its crystal and molecular structure has been determinated by single crystal X-ray diffraction. The compound crystallizes in the triclinic space group P1 with a = 1363.5(7), b = 1462.7(10), c = 1637.7(7) pm, α = 95.40(5)°, β = 96.79(4)°, γ = 102.12(5)° and Z = 2. The crystal structure consists of discrete, trimeric molecules with octahedrally coordinated tin atoms which are connected to each other corresponding to the formulation Sn33-O)(μ2-OiBu)3(O1Bu)7 · (i-BuOH)2 by three isobutoxide groups bridging two metal atoms and a single threefold bridging oxygen atom  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