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.
This study deals with the reduction of Fe2O3 by H2 in the temperature range of 220-680 °C. It aims to examine the rate controlling processes of Fe2O3 reduction by H2 in the widest and lowest possible temperature range. This is to be related with efforts to decrease the emission of CO2 in the atmosphere thus decreasing its green house effect.Reduction of hematite to magnetite with H2 is characterized by an apparent activation energy ‘Ea’ of 76 kJ/mol. Ea of the reduction of magnetite to iron is 88 and 39 kJ/mol for temperatures lower and higher than 420 °C, respectively. Mathematical modeling of experimental data suggests that the reaction rate is controlled by two- and three-dimensional growth of nuclei and by phase boundary reaction at temperatures lower and higher than 420 °C, respectively.Morphological study confirms the formation of compact iron layer generated during the reduction of Fe2O3 by H2 at temperatures higher than 420 °C. It also shows the absence of such layer in case of using CO. It seems that the annealing of magnetite's defects around 420 °C is responsible for the decrease of Ea.The rate of reduction of iron oxide with hydrogen is systematically higher than that obtained by CO.  相似文献   

2.
The citrate-nitrate gel combustion route was used to prepare SrFe2O4(s), Sr2Fe2O5(s) and Sr3Fe2O6(s) powders and the compounds were characterized by X-ray diffraction analysis. Different solid-state electrochemical cells were used for the measurement of emf as a function of temperature from 970 to 1151 K. The standard molar Gibbs energies of formation of these ternary oxides were calculated as a function of temperature from the emf data and are represented as (SrFe2O4, s, T)/kJ mol−1 (±1.7)=−1494.8+0.3754 (T/K) (970?T/K?1151). (Sr2Fe2O5, s, T)/kJ mol−1 (±3.0)=−2119.3+0.4461 (T/K) (970?T/K?1149). (Sr3Fe2O6, s, T)/kJ mol−1 (±7.3)=−2719.8+0.4974 (T/K) (969?T/K?1150).Standard molar heat capacities of these ternary oxides were determined from 310 to 820 K using a heat flux type differential scanning calorimeter (DSC). Based on second law analysis and using the thermodynamic database FactSage software, thermodynamic functions such as ΔfH°(298.15 K), S°(298.15 K) S°(T), Cp°(T), H°(T), {H°(T)-H°(298.15 K)}, G°(T), free energy function (fef), ΔfH°(T) and ΔfG°(T) for these ternary oxides were also calculated from 298 to 1000 K.  相似文献   

3.
Nickel and iron substituted LaCoO3 with rhombohedrally distorted perovskite structure were obtained in the temperature range of 600-900 °C by thermal decomposition of freeze-dried citrates and by the Pechini method. The crystal structure, morphology and defective structure of LaCo1−xNixO3 and LaCo1−xFexO3 were characterized by X-ray diffraction and neutron powder diffraction, TEM and SEM analyses and electron paramagnetic resonance spectroscopy. The reducibility was tested by temperature programmed reduction with hydrogen. The products of the partial and complete reduction were determined by ex-situ XRD experiments. The replacement of Co by Ni and Fe led to lattice expansion of the perovskite structure. For perovskites annealed at 900 °C, there was a random Ni, Fe and Co distribution. The morphology of the perovskites does not depend on the Ni and Fe content, nor does it depend on the type of the precursor used. LaCo1−xNixO3 perovskites (x>0.1) annealed at 900 °C are reduced to Co/Ni transition metal and La2O3 via the formation of oxygen deficient Brownmillerite-type compositions. For LaCo1−xNixO3 annealed at 600 °C, Co/Ni metal, in addition to oxygen-deficient perovskites, was formed as an intermediate product at the initial stage of the reduction. The interaction of LaCo1−xFexO3 with H2 occurs by reduction of Co3+ to Co2+ prior to the Fe3+ ions. The reducibility of Fe-substituted perovskites is less sensitive towards the synthesis procedure in comparison with that of Ni substituted perovskites.  相似文献   

4.
Iron oxide modified with single- or double-metal additives (Cr, Ni, Zr, Ag, Mo, Mo-Cr, Mo-Ni, Mo-Zr and Mo-Ag), which can store and supply pure hydrogen by reduction of iron oxide with hydrogen and subsequent oxidation of reduced iron oxide with steam (Fe3O4 (initial Fe2O3)+4H2↔3Fe+4H2O), were prepared by impregnation. Effects of various metal additives in the samples on hydrogen production were investigated by the above-repeated redox. All the samples with Mo additive exhibited a better redox performance than those without Mo, and the Mo-Zr additive in iron oxide was the best effective one enhancing hydrogen production from water decomposition. For Fe2O3-Mo-Zr, the average H2 production temperature could be significantly decreased to 276 °C, the average H2 formation rate could be increased to 360.9-461.1 μmol min−1 Fe-g−1 at operating temperature of 300 °C and the average storage capacity was up to 4.73 wt% in four cycles, an amount close to the IEA target.  相似文献   

5.
The CaMnO3−x (x=0 and 0.02) mixed oxide was synthesized from both thermal treatment of the [CaMn(C3H2O4)2(H2O)4] metallo-organic precursor and ceramic method. In the case of the latter method, temperatures of 1350 °C and 50 h were necessary; however, lower temperatures, 800 °C, and shorter reaction times, 15 h, were utilized in the attainment of the mixed oxide from the precursor. As a consequence, the morphology of the different products is clearly different. The samples exhibit three-dimensional antiferromagnetic ordering with TN near 120 K, and a low-dimensional antiferromagnetic ordering at high temperatures. The presence of a ferromagnetic component above TN was also observed in both compounds, it is slightly stronger in the phase prepared by the ceramic route.  相似文献   

6.
The standard molar Gibbs energies of formation of LnFeO3(s) and Ln3Fe5O12(s) where Ln=Eu and Gd have been determined using solid-state electrochemical technique employing different solid electrolytes. The reversible e.m.f.s of the following solid-state electrochemical cells have been measured in the temperature range from 1050 to 1255 K.Cell (I): (−)Pt / {LnFeO3(s)+Ln2O3(s)+Fe(s)} // YDT/CSZ // {Fe(s)+Fe0.95O(s)} / Pt(+);Cell (II): (−)Pt/{Fe(s)+Fe0.95O(s)}//CSZ//{LnFeO3(s)+Ln3Fe5O12(s)+Fe3O4(s)}/Pt(+);Cell (III): (−)Pt/{LnFeO3(s)+Ln3Fe5O12(s)+Fe3O4(s)}//YSZ//{Ni(s)+NiO(s)}/Pt(+);andCell(IV):(−)Pt/{Fe(s)+Fe0.95O(s)}//YDT/CSZ//{LnFeO3(s)+Ln3Fe5O12(s)+Fe3O4(s)}/Pt(+).The oxygen chemical potentials corresponding to the three-phase equilibria involving the ternary oxides have been computed from the e.m.f. data. The standard Gibbs energies of formation of solid EuFeO3, Eu3Fe5O12, GdFeO3 and Gd3Fe5O12 calculated by the least-squares regression analysis of the data obtained in the present study are given byΔfm(EuFeO3, s) /kJ mol−1 (± 3.2)=−1265.5+0.2687(T/K)   (1050 ? T/K ? 1570),Δfm(Eu3Fe5O12, s)/kJ mol−1 (± 3.5)=−4626.2+1.0474(T/K)   (1050 ? T/K ? 1255),Δfm(GdFeO3, s) /kJ mol−1 (± 3.2)=−1342.5+0.2539(T/K)   (1050 ? T/K ? 1570),andΔfm(Gd3Fe5O12, s)/kJ·mol−1 (± 3.5)=−4856.0+1.0021(T/K)   (1050 ? T/K ? 1255).The uncertainty estimates for Δfm include the standard deviation in the e.m.f. and uncertainty in the data taken from the literature. Based on the thermodynamic information, oxygen potential diagrams for the systems Eu-Fe-O and Gd-Fe-O and chemical potential diagrams for the system Gd-Fe-O were computed at 1250 K.  相似文献   

7.
Hyphenation of thermogravimetric analyzer (TGA) and thermo-Raman spectrophotometer for in situ monitoring of solid-state reaction in oxygen atmosphere forming NiO-Al2O3 catalyst nanoparticles is investigated. In situ thermo-Raman spectra in the range from 200 to 1400 cm−1 were recorded at every degree interval from 25 to 800 °C. Thermo-Raman spectroscopic studies reveal that, although the onset of formation is around 600 °C, the bulk NiAl2O4 forms at temperatures above 800 °C. The X-ray diffraction (XRD) spectra and the scanning electron microscopy (SEM) images of the reaction mixtures were recorded at regular temperature intervals of 100 °C, in the temperature range from 400 to 1000 °C, which could provide information on structural and morphological evolution of NiO-Al2O3. Slow controlled heating of the sample enabled better control over morphology and particle size distribution (∼20-30 nm diameter). The observed results were supported by complementary characterizations using TGA, XRD, SEM, transmission electron microscopy, and energy dispersive X-ray analysis.  相似文献   

8.
Nickel oxide was introduced as a grain growth inhibitor into Ba0.5Sr0.5Co0.8Fe0.2O3 − δ (BSCF) electrode to increase its activity for oxygen reduction reaction. Small amount of NiO can effectively suppress the grain growth of BSCF by spinning the grain boundary and thus increase the electrode surface area, while the particle connection is not obviously affected. Electrochemical impedance spectra of symmetric cells indicate that the electrode activity for oxygen reduction is indeed effectively improved by introducing NiO additive. At 600 °C, area specific resistance of BSCF+5wt.% NiO is about 36.5% lower than that of pure BSCF electrode. Single cell tests also demonstrate an improved cell power output by introducing NiO additive. It implies that the adoption of a sintering inhibitor may be a facile and practical way to increase electrode activity for oxygen reduction.  相似文献   

9.
The iron rich part of the system was examined in the temperature range of 1200-1380 °C in air, with focus on the solid solutions of M-type hexaferrites. Samples of suitable compositions were studied by electronprobe microanalysis (EPMA). Substituted Sr-hexaferrites in the system Sr-La-Co-Fe-O do not follow the 1:1 substitution mechanism of La/Co in M-type ferrites. Due to the presence and limited Co2+-incorporation Fe3+-ions are reduced to Fe2+ within the crystal lattice to obtain charge balance. In all examined M-type ferrites divalent iron is formed, even at 1200 °C. The substitution principle Sr2++Fe3+↔La3++(Fe2+, Co2+) yields to the general substitution formula for the M-type hexaferrite Sr2+1-xLa3+xFe2+x-yCo2+yFe3+12-xO19 (0≤x≤1 and 0≤yx). In addition Sr/La-perovskiteSS (SS=solid solution), Co/Fe-spinelSS, hematite and magnetite are formed. Sr-hexaferrite exhibits at 1200 °C a limited solid solution with small amounts of Fe2+ (SrFe12O19↔Sr0.3La0.7Co0.5Fe2+0.2Fe11.3O19). At 1300 and 1380 °C a continuous solid solution series of the M-type hexaferrite is stable. SrFe12O19 and LaCo0.4Fe2+0.6Fe11O19 are the end members at 1300 °C. The maximum Fe2+O content is about 13 mol% in the M-type ferrite at 1380 °C (LaCo0.1Fe2+0.9Fe11O19).  相似文献   

10.
Mn/Fe mixed oxide solids doped with Al2O3 (0.32-1.27 wt.%) were prepared by impregnation of manganese nitrate with finely powdered ferric oxide, then treated with different amounts of aluminum nitrate. The obtained samples were calcined in air at 700-1000 °C for 6 h. The specific surface area (SBET) and the catalytic activity of pure and doped precalcined at 700-1000 °C have been measured by using N2 adsorption isotherms and CO oxidation by O2. The structure and the phase changes were characterized by DTA and XRD techniques. The obtained results revealed that Mn2O3 interacted readily with Fe2O3 to produce well-crystallized manganese ferrite (MnFe2O4) at temperatures of 800 °C and above. The degree of propagation of this reaction increased by Al2O3-doping and also by increasing the heating temperature. The treatment with 1.27 wt.% Al2O3 followed by heating at 1000 °C resulted in complete conversion of Mn/Fe oxides into the corresponding ferrite phase. The catalytic activity and SBET of pure and doped solids were found to decrease, by increasing both the calcination temperature and the amount of Al2O3 added, due to the enhanced formation of MnFe2O4 phase which is less reactive than the free oxides (Mn2O3 and Fe2O3). The activation energy of formation (ΔE) of MnFe2O4 was determined for pure and doped solids. The promotion effect of aluminum in formation of MnFe2O4 was attributed to an effective increase in the mobility of reacting cations.  相似文献   

11.
TG experiments on the hydrogen reduction of α-Fe2O3 were carried out to elucidate the influence of the preparation history of the oxide on its reactivity. α-Fe2O3 samples were prepared by the thermal decomposition of seven iron salts in a stream of oxygen, air or nitrogen at temperatures of 500–1200°C for 1 h. Thirteen metal ions such as Cu2+, Ni2+, etc. were used as doping agents. The reactivity of the oxide was indicated by the initial reduction temperature (Ti. α-Fe2O3 prepared at lower temperatures showed lower Ti values and the reduction proceeded stepwise (Fe2O3 → Fe3O4 → Fe). Ti values increased with the rise in the preparation temperature of the oxide. The oxides prepared at higher temperatures showed that two reduction steps (Fe2O3 → Fe3O4 → Fe) proceed simultaneously. the preparation in oxygen gave higher Ti than that in air or nitrogen. The doping by metal ions, except Ti4+, lowered the Ti of α-Fe2O3. The Cu2+ ion showed the lowest Ti, while Ti4+ showed the highest Ti and the inhibition effect.The reduction process was expressed by two equations; Avrami—Erofeev's equation for α-Fe2O3 → Fe3O4 and Mampel's equation for Fe3O4 → Fe.  相似文献   

12.
The chemical reactivity of La2NiO4+δ and nickel metal or nickel oxide with fast oxide-ion conductor La2Mo2O9 is investigated in the annealing temperature range between 600 and 1000 °C, using room temperature X-ray powder diffraction. Within the La2NiO4+δ/La2Mo2O9 system, subsequent reaction is evidenced at relatively low annealing temperature (600 °C), with formation of La2MoO6 and NiO. The reaction is complete at 1000 °C. At reverse, no reaction occurs between Ni or NiO and La2Mo2O9 up to 1000 °C. Together with a previous work [G. Corbel, S. Mestiri, P. Lacorre, Solid State Sci. 7 (2005) 1216], the current study shows that Ni-CGO cermets might be chemically and mechanically compatible anode materials to work with LAMOX electrolytes in solid oxide fuel cells.  相似文献   

13.
A series of spinel-type CoxNi1−xFe2O4 (x = 0, 0.2, 0.4, 0.5, 0.6, 0.8, 1.0) magnetic nanomaterials were solvothermally synthesized as enzyme mimics for the eletroctrocatalytic oxidation of H2O2. X-ray diffraction and scanning electron microscope were employed to characterize the composition, structure and morphology of the material. The electrochemical properties of spinel-type CoxNi1−xFe2O4 with different (Co/Ni) molar ratio toward H2O2 oxidation were investigated, and the results demonstrated that Co0.5Ni0.5Fe2O4 modified carbon paste electrode (Co0.5Ni0.5Fe2O4/CPE) possessed the best electrocatalytic activity for H2O2 oxidation. Under optimum conditions, the calibration curve for H2O2 determination on Co0.5Ni0.5Fe2O4/CPE was linear in a wide range of 1.0 × 10−8–1.0 × 10−3 M with low detection limit of 3.0 × 10−9 M (S/N = 3). The proposed Co0.5Ni0.5Fe2O4/CPE was also applied to the determination of H2O2 in commercial toothpastes with satisfactory results, indicating that CoxNi1−xFe2O4 is a promising hydrogen peroxidase mimics for the detection of H2O2.  相似文献   

14.
New niobium oxynitrides containing either magnesium or silicon were prepared at 1000 °C by ammonia nitridation of oxide precursors obtained via the citrate route. The products had rock-salt type crystal structures. Crystallinity was improved by annealing in 0.5 MPa N2 and the final compositions were (Nb0.95Mg0.05)(N0.92O0.08) at 1500 °C and (Nb0.87Si0.090.04)(N0.87O0.13) at 1200 °C. The magnesium and oxide ions partially co-substitute the niobium and nitride ions in the octahedral sites of the δ-NbN lattice, respectively. Silicon ions were also successfully doped together with oxide ions into the rock-salt type NbN lattice. The Si doped product exhibited relatively large displacement at the octahedral sites and was accompanied by a small amount of cation vacancies. Superconductivity was improved by annealing to obtain critical temperatures/volume fractions of Tc=17.6 K/100% for Mg- and Tc=16.2 K/95% for the Si-doped niobium oxynitrides.  相似文献   

15.
The structure and crystal phase of the nanocrystalline powders of Ni1−xZnxFe2O4 (0 ≤ x ≤ 0.5) mixed ferrite, synthesized by ethylene glycol mediated citrate sol-gel method, were characterized by X-ray diffraction and microstructure by transmission electron microscopy. Further studies by Fourier transform infrared spectroscopy were also conducted. Moreover, DC electrical properties of the prepared nanoparticles were studied by DC conductivity measurements. The response of prepared Ni1−xZnxFe2O4 mixed ferrites to different reducing gases (ethanol, hydrogen sulfide, ammonia, hydrogen and liquefied petroleum gas) was investigated. In particular, Ni0.6Zn0.4Fe2O4 composition exhibited high response to 100 ppm ethanol gas at 300 °C. Incorporation of palladium further improved the response, selectivity and response time of Ni0.6Zn0.4Fe2O4 to ethanol gas with the blue shift in the operating temperature by 25 °C.  相似文献   

16.
Subsolidus phase relationships in the In2O3-WO3 system at 800-1400°C were investigated using X-ray diffraction. Two binary-oxide phases—In6WO12 and In2(WO4)3—were found to be stable over the range 800-1200°C. Heating the binary-oxide phases above 1200°C resulted in the preferential volatilization of WO3. Rietveld refinement was performed on three structures using X-ray diffraction data from nominally phase-pure In6WO12 at room temperature and from nominally phase-pure In2(WO4)3 at 225°C and 310°C. The indium-rich phase, In6WO12, is rhombohedral, space group (rhombohedral), with Z=1, a=6.22390(4) Å, α=99.0338(2)° [hexagonal axes: aH=9.48298(6) Å, c=8.94276(6) Å, aH/c=0.9430(9)]. In6WO12 can be viewed as an anion-deficient fluorite structure in which 1/7 of the fluorite anion sites are vacant. Indium tungstate, In2(WO4)3, undergoes a monoclinic-orthorhombic transition around 250°C. The high-temperature polymorph is orthorhombic, space group Pnca, with a=9.7126(5) Å, b=13.3824(7) Å, c=9.6141(5) Å, and Z=4. The low-temperature polymorph is monoclinic, space group P21/a, with a=16.406(2) Å, b=9.9663(1) Å, c=19.099(2) Å, β=125.411(2)°, and Z=8. The structures of the two In2(WO4)3 polymorphs are similar, consisting of a network of corner sharing InO6 octahedra and WO4 tetrahedra.  相似文献   

17.
According to X-ray diffraction data, the STK catalyst is a mixture of Fe2O3 and Cr2O3. The temperature-programmed reduction spectrum exhibited two reduction peaks: one, with T max = 250°C, corresponds to the reduction process Cr2O3 → CrO and the other, with T max = 360°C, corresponds to the reduction Fe2O3 → Fe3O4. The results of thermal desorption measurements suggest that the individual adsorption of oxygen on the surface of the STK catalyst is low; in this case (according to IR-spectroscopic data), an atomic form is the main species. Surface nitrite-nitrate complexes are formed upon the adsorption of NO. Nitrite and nitrate complexes desorbed at maximum rates at 105 and 160°C, respectively. Unlike the NTK-10-1 catalyst, the NO species, which desorbed at high temperatures (250–400°C), was absent from the surface of STK. Propane adsorbed at room temperature to form surface compounds containing an acetate group. The interaction of propane with the surface of the STK catalyst at reaction temperatures resulted in strong surface reduction.__________Translated from Kinetika i Kataliz, Vol. 46, No. 4, 2005, pp. 550–558.Original Russian Text Copyright © 2005 by Tret’yakov, Burdeinaya, Zakorchevnaya, Matyshak, Korchak.  相似文献   

18.
Nanocrystalline single-phase samples of Zn1−xNixFe2O4 ferrites (0<x<1) have been obtained via a soft-chemistry method based on citrate-ethylene glycol precursors, at a relatively low temperature (650 °C). The influence of the nickel and zinc contents as well as that of heat treatments were investigated by means of X-ray powder diffraction, Brunauer-Emmett-Teller (BET) surface area, scanning electron microscopy (SEM) and Fourier Transform Infrared (FTIR) Spectroscopy. Higher Ni content increases the surface areas, the largest one (∼20 m2/g) being obtained for NiFe2O4 annealed at 650 °C for 15 h. For all compositions, the surface area decreases for prolonged annealing at 650 °C and for higher annealing temperatures. Those results were correlated to the particle size evolution; the smallest particles (∼50 nm) observed in the NiFe2O4 sample (650 °C, 15 h) steadily increase as Ni ions were replaced by Zn, reaching ∼100 nm in the ZnFe2O4 sample (650 °C, 15 h). For all the Zn1−xNixFe2O4 samples and, whatever the heat treatments was, the FTIR spectra show two fundamental absorption bands in the range 650-400 cm−1, characteristics of metal vibrations, without any superstructure stating for cation ordering. The highest ν1-tetrahedral stretching, observed at ∼615 cm−1 in NiFe2O4, shifts towards lower values with increasing Zn, whereas the ν2-octahedral vibration, observed at 408 cm−1 in NiFe2O4, moves towards higher wavenumbers, reaching 453 cm−1 in ZnFe2O4.  相似文献   

19.
The miscibility of TbBaMn2O5+x and TbBaMn2O5.5−y has been investigated at 100-600 °C using in situ powder neutron diffraction. No miscibility is observed, and the two phases remain oxygen stoichiometric (x,y=0) at 600 °C. Structure refinement results show that neither material undergoes a phase transition in this temperature range. TbBaMn2O5 is Mn2+/Mn3+ charge ordered and any charge melting transition is >600 °C. This symmetry-broken charge ordering is remarkably robust in comparison to that in other oxides.  相似文献   

20.
Nb4W13O47, a member of the solid solution series Nb8−nW9+nO47 (0?n?5), crystallizes in a threefold superstructure of the tetragonal tungsten bronze structure. While the oxidation of this reduced phase at TOX=1200 °C leads to a separation into the thermodynamically stable phases, lower oxidation temperatures result in products that comprise new structural elements and ordering variants. The characterization of the oxidation products obtained at TOX=1000 °C was performed by scanning transmission electron microscopy applying a high-angle annular dark field detector. At the selected imaging conditions (Z contrast), not only the metal positions are revealed by this technique but valuable additional information about the elemental distribution can be obtained simultaneously.  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