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.
A model Pd/Fe2O3 catalyst prepared by the vacuum technique has been studied in the carbon monoxide oxidation in the temperature range of 300–550 K at reagent pressures P(CO)=16 Torr, P(O2)= 4 Torr. It has been shown that the activity of the fresh catalysts is determined by palladium. According to the XPS data, the reduction with carbon monoxide results in the formation of Fe2+ (formally Fe3O4) and appearance of the catalytic activity in this reaction at low temperatures (350 K). High low-temperature activity of the catalyst is supposed to be connected with the reaction between oxygen adsorbed on the reduced sites of the support (Fe2+) and CO adsorbed on palladium (COads) at the metal–oxide interface.  相似文献   

2.
The product distributions for partial oxidation of methane on Fe2(MoO4)3 catalyst were changed remarkably when the oxidant was switched from oxygen to nitrous oxide. When oxygen was used as the oxidant, the main products were HCHO and CO. However, when nitrous oxide was used, the formation of HCHO was greatly suppressed and C2 hydrocarbons (C2H6 and C2H4) were newly produced. The difference in kinetic behaviors between the two reactions using nitrous oxide and oxygen as the oxidant can be explained in terms of the competitive conversions of methyl intermediate into HCHO and C2H6. In the case of nitrous oxide as the oxidant, the adsorbed methyl intermediate would be transformed predominantly into C2H6 due to a low steady-state concentration of the active oxygen species on Fe2(MoO4)3.  相似文献   

3.
A detailed structural analysis on the in situ synthesized β‐Ca3(PO4)2/α‐Fe2O3 composites is demonstrated. Compositional ratios, the influence and occupancy of iron at the β‐Ca3(PO4)2 lattice, oxidation state of iron in the composites are derived from analytical techniques involving XRD, FT‐IR, Raman, refinement of the powder X‐ray diffraction and X‐ray photoelectron spectroscopy. Iron exists in the Fe3+ state throughout the investigated systems and favors its occupancy at the Ca2+(5) site of β‐Ca3(PO4)2 until critical limit, and thereafter crystallizes as α‐Fe2O3 at ambient conditions. Fe3+ occupancy at the β‐Ca3(PO4)2 lattice yields a Ca9Fe(PO4)7 structure that is isostructural with its counterpart. A strong rise in the soft ferromagnetic behavior of β‐Ca3(PO4)2/α‐Fe2O3 composites is obvious that depends on the content of α‐Fe2O3 in the composites. Overall, the diverse level of iron inclusions at the calcium phosphate system with a Ca/P ratio of 1.5 yields a structurally stable β‐Ca3(PO4)2/α‐Fe2O3 composites with assorted compositional ratios.  相似文献   

4.
Photo‐induced atom transfer radical polymerization (ATRP) of methyl methacrylate (MMA) was achieved in poly(ethylene glycol)‐400 with nanosized α‐Fe2O3 as photoinitiator. Well‐defined poly(methyl methacrylate) (PMMA) was synthesized in conjunction with ethyl 2‐bromoisobutyrate (EBiB) as ATRP initiator and FeCl3·6H2O/Triphenylphosphine (PPh3) as complex catalyst. The photo‐induced polymerization of MMA proceeded in a controlled/living fashion. The polymerization followed first‐order kinetics. The obtained PMMA had moderately controlled number‐average molecular weights in accordance with the theoretical number‐average molecular weights, as well as narrow molecular weight distributions (Mw/Mn). In addition, the polymerization could be well controlled by periodic light‐on–off processes. The resulting PMMA was characterized by 1H nuclear magnetic resonance and gel permeation chromatography. The brominated PMMA was used further as macroinitiator in the chain‐extension with MMA to verify the living nature of photo‐induced ATRP of MMA. © 2015 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2015 , 132, 42389.  相似文献   

5.
《分离科学与技术》2012,47(1):97-110
Abstract

The ability of four amorphous Al3+‐ and Fe3+‐doped titanium and zirconium sorbents to separate U(VI) from acidic aqueous solutions (pHinit=3, ionic strength 0.1 M established by NaNO3) was investigated using a batch technique and instrumental neutron activation analysis. All investigated sorbents were found to be chemically stable and remove considerable amounts of uranium from acidic aqueous solutions (pHinit=3). The scanning electron microscopic and powder‐X‐ray diffraction examination of the grains of the two investigated titanium phosphates after contacting the uranium solutions revealed the formation of sodium autunite (Na2(UO2)2(PO4)2 · 6‐8H2O) accompanied, in the case of the Fe3+‐doped titanium phosphate, by iron uranyl phosphate hydroxide hydrate (Fe(UO2)2(PO4)2(OH) · 7H2O). No crystal formation was observed in the cases of uranium sorbed by zirconium phosphates indicating the different sorption mechanism involved.  相似文献   

6.
BACKGROUND: Particles are often too small to be separated from a reaction system and recycled, especially in waste‐water treatment via a catalytic ozonation process. Therefore, the objective of this study was to prepare a magnetic catalyst (SiO2/Fe3O4) that can be recycled by using an external magnetic field. The characteristics of the magnetic catalyst and kinetics of decomposition of Reactive Black 5 (RB5) using a magnetic catalyst/H2O2/O3 have been investigated. Magnetic catalysts (SiO2/Fe3O4) were characterized by X‐ray diffraction (XRD), vibrating sample magnetometry (VSM) and Fourier transform infrared (FTIR) spectroscopy. RESULTS: In the case of the decomposition of RB5 by the SiO2/Fe3O4/H2O2/O3 process, the contribution of the destruction by the indirect oxidation of OH· caused by the chain reaction of H2O, OH?, and H2O2 was found to be larger than that of the direct attack of O3. The kinetic behaviour of the reacting species during the decomposition of RB5 by catalytic ozonation could be well modelled under various experimental conditions. CONCLUSIONS: The paramagnetic behaviour of the prepared SiO2/Fe3O4 led to the formation of the magnetic catalyst SiO2/Fe3O4, which could be separated more easily by the application of a magnetic field. More than 90% of the magnetic catalyst was recovered and easily redispersed in a solution for reuse by mixing with a flat‐bladed turbine. Copyright © 2010 Society of Chemical Industry  相似文献   

7.
The effects of both Al cocatalyst and solvent on catalytic activity in the ethylene polymerization by the (arylmido)(aryloxo)vanadium(V) complex, VCl2(N‐2,6‐Me2C6H3)(O‐2,6‐Me2C6H3) ( 1 ), have been explored in detail. The activity of 5.84×105 kg PE/mol V⋅h (TOF 2.08×107 h−1) has been achieved by 1 /EtAlCl2 catalyst in CH2Cl2 at 0 °C, and the activity in toluene increased in the order: i‐Bu2AlCl>EtAlCl2>Me2AlCl>Et2AlCl> Et2Al(OEt), AlEt3, AlMe3 (negligible activities). Both aluminum alkyl cocatalyst and solvent also affected the catalytic activity and the norbornene (NBE) incorporation in the ethylene/NBE copolymerization using complex 1 , whereas the NBE contents were not strongly affected by the kind of aryl oxide ligand in VCl2(N‐2,6‐Me2C6H3)(OAr) [OAr=O‐2,6‐Me2C6H3 ( 1 ), O‐2,6‐i‐Pr2C6H3 ( 2 ), O‐2,6‐Ph2C6H3 ( 3 )].  相似文献   

8.
The syntheses and crystal structures of Zn(CH4N2O)2(H2O)2·2(NO3) (1) and Co(CH4N2O)2(H2O)2·2(NO3) (2), the first well-characterised metal complexes of formylhydrazine (fh), are described. In both compounds, the fh acts as an N,O-bidentate ligand in a centrosymmetric [M(fh)2(H2O)2]2+ cation, with charge balance supplied by nitrate counter ions. The packing for the two compounds are quite different: in 1, chains of [Zn(fh)2(H2O)2]2+ units are seen in the triclinic unit cell, whereas in the monoclinic structure of 2, sheets of cations occur. This might arise because the conformations of the five-membered chelate rings for the ligands are slightly different, with that for 2 showing a greater degree of puckering.  相似文献   

9.
Free acids of the iron substituted heteropoly acids (HPA), H7[(P2W17O61)FeIII(H2O)] (HFe1) and H18[(P2W15O56)2FeIII2(H2O)2] (HFe2) were prepared from the salts K7[(P2W17O61)FeIII(H2O)] (KFe1) and Na12[(P2W15O56)2FeIII4(H2O)2] (NaFe4), respectively. The iron-substituted HPA were adsorbed on to XC-72 carbon based GDLs to form HPA doped GDEs after water washing with HPA loadings of ca. 1 μmol. The HPA was detected throughout the GDL by EDX. Solution electrochemistry of the free acids are reported for the first time in sulfate buffer, pH 1-3. The hydrogen oxidation reaction was catalyzed by KFe1 at 0.33 V, with an exchange current density of 38 mA/cm2. Moderate activity for the oxygen reduction reaction was observed for the iron substituted HPA, which was dramatically improved by selectively removing oxygen atoms from the HPA by cycling the fuel cell cathode under N2 followed by reoxidation to give a restructured oxide catalyst. The nanostructured oxide achieved an OCV of 0.7 V with a Tafel slope of 115 mV/decade. Cycling the same catalysts in oxygen resulted in an improved catalyst/ionomer/carbon configuration with a slightly higher Tafel slope, 128 mV/decade but a respectable current density of 100 mA/cm2 at 0.2 V.  相似文献   

10.
Five Ba(Co1/3Nb2/3)O3 samples sintered at different temperatures (form 1350 to 1550 °C), one Ba(Mg1/3Ta2/3)O3 and a Ba(Mg1/3Nb2/3)O3 sample were examined by Raman scattering to reveal the correlation of the 1:2 ordered perovskite structure with the microwave properties, such as dielectric constant and Q factors. The Ba(Co1/3Nb2/3)O3 sample sintered at 1400 °C, which possesses the highest microwave Q value and the lowest dielectric constant among five Ba(Co1/3Nb2/3)O3 samples, has the narrowest width and the highest frequency of the stretch mode of oxygen octahedron (i.e. A1g(O) near 800 cm−1). We found that the dielectric constant is strongly correlated with the Raman shift of A1g(O) stretch modes, and the width of A1g(O) stretch mode reflects the quality factor Q × f value in the 1:2 ordered perovskite materials. This concludes that the oxygen octahedron play an important role of the material's microwave performance. Based on the results of Q × f values and the lineshapes of A1g(O) stretch mode, we found that the propagation of microwave energy in Ba(Mg1/3Ta2/3)O3 and Ba(Mg1/3Nb2/3)O3 shows weak damping behavior, however, Ba(Co1/3Nb2/3)O3 samples sintered at different temperature exhibit heavily damped behavior.  相似文献   

11.
A novel bis(β‐ketoamino)Ni(II) complex catalyst, Ni{CF3C(O)CHC[N(naphthyl)]CH3}2, was synthesized, and the structure was solved by a single‐crystal X‐ray refraction technique. The copolymerization of norbornene with higher 1‐alkene was carried out in toluene with catalytic systems based on nickel(II) complexes, Ni{RC(O)CHC[N(naphthyl)]CH3}2(R?CH3, CF3) and B(C6F5)3, and high activity was exhibited by both catalytic systems. The effects of the catalyst structure and comonomer feed content on the polymerization activity and the incorporation rates were investigated. The reactivity ratios were determined to be r1‐octene = 0.009 and rnorbornene = 13.461 by the Kelen–Tüdõs method for the Ni{CH3C(O)CHC[N(naphthyl)]CH3}2/B(C6F5)3 system. The achieved copolymers were confirmed to be vinyl‐addition copolymers through the analysis of 1H‐NMR and 13C‐NMR. The thermogravimetric analysis results showed that the copolymers exhibited good thermal stability (decomposition temperature, Tdec > 400°C), and the glass‐transition temperature of the copolymers were observed between 215 and 275°C. The copolymers were confirmed to be noncrystalline by wide‐angle X‐ray diffraction analysis and showed good solubility in common organic solvents. © 2011 Wiley Periodicals, Inc. J Appl Polym Sci, 2012  相似文献   

12.
Benzyl and trans-cinnamyl alcohols are heterogeneously oxidised to the corresponding aldehydes by O2 in liquid phase at 100 °C and ambient pressure using hydrous binary PdII–M oxides (M=CoIII, FeIII, MnIII and CuII) as catalysts. Modification of PdII oxide with transition metal cations greatly improves the catalytic activity and selectivity to aldehydes, CoIII and FeIII being the most effective promoters. In benzyl alcohol oxidation in toluene solution, the Pd–Co system gives 85–100% selectivity to aldehydes at 53–95% alcohol conversion in 15–60 min reaction time. The catalyst can be re-used without loss of its activity and selectivity. The presence of a certain amount of water in the catalysts is essential for their performance. From TGA, the composition of the optimal Pd–Co catalyst can be approximated as PdO·(0.13–1.0)CoO(OH)·(2–3)H2O. The oxidation of alcohols on Pd–M oxide catalysts is accompanied by transfer hydrogenation and decarbonylation side reactions, which is similar to the oxidation on the palladium metal. This indicates that the oxidation of alcohols on Pd–M oxide catalysts occurs via a dehydrogenation mechanism, with hydrogen being present on the catalyst surface.  相似文献   

13.
A new copper(II) molybdenum(VI) arsenate(III), (C5H5NH)2(H3O)2[(CuO6)Mo6O18(As3O3)2], has been hydrothermally synthesized and characterized by single crystal X-ray diffraction and TG analysis. The compound crystallizes in the monoclinic space group P21/n with a=9.303(2), b=20.731(4), c=12.617(3) Å, β=104.17(3)°, V=2359.3(8) Å3, Z=2 and R1(wR2)=0.0296(0.0683). The structure is composed of pyridinium cations, proton hydrates and [(CuO6)Mo6O18(As3O3)2]4− polyanions. The polyanion framework derives from the Anderson type; the central octahedron is filled up by copper(II) and is capped on both sides by a cyclic As3O6 group.  相似文献   

14.
The temperature manipulation induces the aggregation of Ru2(CO3)43  paddle-wheel precursors and Mn2 + ions in lower temperature ~ 10 °C forming layer structural complex, K[Mn(H2O)4Ru2(CO3)4]·5H2O (1). It composes of new negative layer {Mn(H2O)5Ru2(CO3)4}nn, and magnetic exchanges between spin centers result in ordering below 3.8 K. The observed critical temperature is like the previously reported 3D hetero-metallic carbonates H0.3K0.7Mn[Ru2(CO3)4](H2O)5.5, which demonstrates that it is independent of the interlayer connecting in such heterometallic complexes based on square-grid layer {Ru2(CO3)4}n3n.  相似文献   

15.
A novel cadmium phosphonate compound Na2[Cd2(H2O)3(O3PCH(OH)CO2)2] · 2H2O (1) has been synthesized by hydrothermal reaction at 120 °C and characterized by single-crystal X-ray diffraction as well as with infrared spectroscopy, elemental analysis and thermogravimetric analysis. The structure of compound 1 comprises CdO6 octahedra and CdO7 pentagonal bipyramid connected by [O3PCH(OH)CO2]3− to form a 2D layered structure with a one-dimensional channel system and the charge-compensating Na+ cations being located between two adjacent layers.  相似文献   

16.
Ring-opening polymerization of DL -lactide (LA) has been initiated with the (η3-C3H5)2Sm(μ2-Cl)23-Cl)2Mg(tmed)(η2-Cl)Mg(tmed) complex both in bulk and solution. The effects of reaction conditions, such as reaction time, reaction temperature, and monomer/initiator molar ratio on the polymerization has been discussed. The results showed that (η3-C3H5)2Sm(μ2-Cl)23-Cl)2Mg(tmed)(μ2-Cl)Mg(tmed) was more effective for the polymerization of LA, and high molecular weight of polylactide was obtained by this initiator. The solvent affected the polymerization significantly. The polymerization mechanism was in agreement with the coordination mechanism. The polymer was characterized by FTIR, 1H-NMR, and DSC. © 1999 John Wiley & Sons, Inc. J Appl Polym Sci 73: 2857–2862, 1999  相似文献   

17.
The reaction of Hg(NO3)2·H2O with bis(diphenylphosphino)amine, Ph2PNHPPh2 (dppam), produces [Hg2(HN2O3)2(NO3)2(dppam)2] (1), which is the first complex containing the monoprotonated trioxodinitrate anion. 1 has been fully characterized by an X-ray structure analysis, NMR spectroscopy (199Hg{1H}, 31P{1H}, 1H), FAB mass spectrometry, IR spectroscopy, elemental analysis and melting point. 1 is the unique example of a face-to-face complex of five-coordinate Hg(II) centres with dppam including the first structural characterization of a rare tautomer of the monoprotonated trioxodinitrate anion.  相似文献   

18.
A dinuclear peroxotungstate, K2[{W(O)(O2)2(H2O)}2(μ‐O)]⋅2 H2O, exhibits high catalytic performance for the epoxidation of various allylic alcohols with only one equivalent of hydrogen peroxide at 305 K in water solvent. The effectiveness of this system is evidenced by high chemo‐, regio‐, and diastereoselectivity, and stereospecificity for the epoxidation of allylic alcohols. Furthermore, products/catalyst separation can be easily carried out by simple extraction and the catalyst recovered can be reused with the maintenance of the catalytic performance.  相似文献   

19.
We rely on a hierarchical approach to identify the low‐lying isomers and corresponding global minima of the pentagonal dodecahedron (H2O)20 and the H3O+(H2O)20 nanoclusters. Initial screening of the isomers is performed using classical interaction potentials, namely the Transferable Interaction 4‐site Potential (TIP4P), the Thole‐Type Flexible Model, versions 2.0 (TTM2‐F) and 2.1 (TTM2.1‐F) for (H2O)20 and the Anisotropic Site Potential (ASP) for H3O+(H2O)20. The nano‐networks obtained with those potentials were subsequently refined at the density functional theory (DFT) with the Becke‐3‐parameter Lee–Yang–Parr (B3LYP) functional and at the second order Møller–Plesset perturbation (MP2) levels of theory. For the pentagonal dodecahedron (H2O)20 it was found that DFT (B3LYP) and MP2 produced the same global minimum. However, this was not the case for the H3O+(H2O)20 cluster, for which MP2 produced a different network for the global minimum when compared to DFT (B3LYP). The low‐lying networks of H3O+(H2O)20 correspond to structures having 9 ‘free’ OH bonds and the hydronium ion on the surface of the nanocluster. The IR spectra of the various networks are further analysed in the OH stretching (‘fingerprint’) region and the various bands are assigned to structural arrangements of the underlying hydrogen bonding network. © 2012 Canadian Society for Chemical Engineering  相似文献   

20.
In this work, a close correlation between variations of critical temperature (Tc) and the hole concentration of (Bi1.6Pb0.4Sr2Ca2Cu3O10+δ)1-x(Fe3O4)x systems was studied. The (Bi1.6Pb0.4Sr2Ca2Cu3O10+δ)1-x(Fe3O4)x samples were fabricated using the solid-state reaction method, where x ranged from 0 to 0.2. The Tc values of the samples deduced from magnetization versus temperature measurement gradually decreased with increasing the Fe3O4 content (x). To investigate a possible reason for the observed decrease in the values of Tc in the samples, the valence state of copper (V) and the hole concentration (p) in all samples were examined by analyzing the Cu K-edge and Cu L2,3-edge X-ray absorption near edge structure (XANES) spectra. The values of V and p monotonously decreased with the increase in Fe3O4 doping content (x) and agreed with the behavior of Tc. The existence of Fe3+ was confirmed by analyzing Fe L2,3-edge XANES. Hence, Fe3+ ions possibly entered the lattice structure of the samples and filled the holes in CuO2 planes. The degradation of superconductivity in Fe3O4 doped samples was then explained.  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