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 31 毫秒
1.
Composition‐tailored Mn1?xRuxO2 2 D nanosheets and their reassembled nanocomposites with mesoporous stacking structure are synthesized by a soft‐chemical exfoliation reaction and the subsequent reassembling of the exfoliated nanosheets with Li+ cations, respectively. The tailoring of the chemical compositions of the exfoliated Mn1?xRuxO2 2 D nanosheets and their lithiated nanocomposites can be achieved by adopting the Ru‐substituted layered manganese oxides as host materials for exfoliation reaction. Upon the exfoliation–reassembling process, the substituted ruthenium ions remain stabilized in the layered Mn1?xRuxO2 lattice with mixed Ru3+/Ru4+ oxidation state. The reassembled Li–Mn1?xRuxO2 nanocomposites show promising pseudocapacitance performance with large specific capacitances of approximately 330 F g?1 for the second cycle and approximately 360 F g?1 for the 500th cycle and excellent cyclability, which are superior to those of the unsubstituted Li–MnO2 homologue and many other MnO2‐based materials. Electrochemical impedance spectroscopy analysis provides strong evidence for the enhancement of the electrical conductivity of 2 D nanostructured manganese oxide upon Ru substitution, which is mainly responsible for the excellent electrode performance of Li–Mn1?xRuxO2 nanocomposites. The results underscore the powerful role of the composition‐controllable metal oxide 2 D nanosheets as building blocks for exploring efficient electrode materials.  相似文献   

2.
The thermodynamic stability of calcium vanadium garnet ferrites Ca3Fe3.5 ? x Ti2x V1.5 ? x O12 upon isomorphic substitution of titanium ions for iron and vanadium ions was studied by the EMF method using ZrO2(Y2O3) ceramic solid electrolyte. Temperature-dependent ΔG 0 was determined. Isomorphic substitutions of titanium ions for iron and vanadium ions in the inequivalent sublattices of the garnet structure in a temperature range of 1100–1483 K cause changes in ΔG 0 due to the entropy and enthalpy components and has a minimal value when x = 0.15.  相似文献   

3.
The Ruddlesden-Popper ferrite Sr3Fe2O6+δ and its titania-doped derivatives Sr3Fe2−xTixO6+δ, where 0<x≤2, have been characterized by X-ray powder diffraction and thermogravimetry. The changes in oxygen content and crystal lattice parameters are consistent with titanium ions entering the solid solution in 4+ oxidation state with octahedral oxygen coordination. Electronic conductivity measurements on polycrystalline Sr3Fe2O6+δ and Sr3Fe0.8Ti1.2O6+δ in the temperature range 750-1000°C and oxygen partial pressures (pO2) varying between 10−20 and 0.5 atm revealed that the predominant partial conductivity of electrons is proportional to pO2−1/4 in the low pO2 region, while the predominant partial contribution of holes to the conductivity is proportional to pO2+1/4 in the high pO2 range. The pressure-independent oxygen ion conductivity is found to decrease with the increase in titanium content. A possible pathway for oxygen ion migration is discussed in relation to disorder in the oxygen sublattice and titanium doping.  相似文献   

4.
The metallic ferromagnetic perovskite-type SrRuO3 (TC  160 K) belongs to the “class” of materials with strongly correlated electrons. Nonetheless a simple ferromagnetism associated with isotropic interactions of low spin Ru4+ ions local moments is far too simple to explain the complex interplay between charge carriers and magnetic interactions. In that sense the suppression of ferromagnetism in isoelectronic Sr1?xCaxRuO3 was tentatively associated to the increased lattice distortion influencing primarily the 4d Ru bandwidths and, hence, the itinerancy and respective populations of the spin-up and spin-down electrons.In order to probe the robustness of the metallic ferromagnetism against electron occupation of 4d Ru orbital we prepared and characterized polycrystalline Sr1?xNaxRuO3 (x = 0.0–0.19) ceramics. The substitution of Sr2+ by Na1+, leading to formally mixed valence Ru4+/Ru5+, induces the decrease of the Curie temperature and spin-wave stiffness, which was determined independently from magnetic and specific heat data. On the other hand the effective paramagnetic moment remains essentially unchanged. All compounds are metallic in a sense of electrical resistivity and thermopower temperature dependence; the low temperature upturn of the electrical resistivity was explained on a base of the weak localization. The metallic nature of the samples is corroborated by Pauli paramagnetism and high Sommerfeld coefficient γ, extracted from the low temperature specific heat, which increases from 30.9 mJ mol?1 K?2 (x = 0.0) to 43.0 mJ mol?1 K?2 (x = 0.19).  相似文献   

5.
Strontium additions in (La1?x Sr x )1?y Mn0.5Ti0.5O3?δ (x?=?0.15–0.75, y?=?0–0.05) having a rhombohedrally distorted perovskite structure under oxidizing conditions lead to the unit cell volume contraction, whilst the total conductivity, thermal and chemical expansion, and steady-state oxygen permeation limited by surface exchange increase with increasing x. The oxygen partial pressure dependencies of the conductivity and Seebeck coefficient studied at 973–1223?K in the p(O2) range from 10?19 to 0.5?atm suggest a dominant role of electron hole hopping and relatively stable Mn3+ and Ti4+ states. Due to low oxygen nonstoichiometry essentially constant in oxidizing and moderately reducing environments and to strong coulombic interaction between Ti4+ cations and oxygen anions, the tracer diffusion coefficients measured by the 18O/16O isotopic exchange depth profile method with time-of-flight secondary-ion mass spectrometric analysis are lower compared to lanthanum–strontium manganites. The average thermal expansion coefficients determined by controlled-atmosphere dilatometry vary in the range 9.8–15.0?×?10?6?K?1 at 300–1370?K and oxygen pressures from 10?21 to 0.21?atm. The anodic overpotentials of porous La0.5Sr0.5Mn0.5Ti0.5O3?δ electrodes with Ce0.8Gd0.2O2-δ interlayers, applied onto LaGaO3-based solid electrolyte, are lower compared to (La0.75Sr0.25)0.95Cr0.5Mn0.5O3?δ when no metallic current-collecting layers are introduced. However, the polarization resistance is still high, ~2 Ω?×?cm2 in humidified 10?% H2–90?% N2 atmosphere at 1073?K, in correlation with relatively low electronic conduction and isotopic exchange rates. The presence of H2S traces in H2-containing gas mixtures did not result in detectable decomposition of the perovskite phases.  相似文献   

6.
Supported titanium–magnesium catalysts (TMC) comprising isolated and clustered titanium ions in different oxidation states, which are obtained using titanium compounds of different composition (TiCl4, TiCl3?nDBE (DBE – dibutyl ether), [η6–BenzeneTiAl2Cl8]), were synthesized and tested in ethylene polymerization. The state of titanium ions was studied by the ESR method both for the procatalysts and after their interaction with triisobutilaluminum. For identification of ESR‐silent Ti3+ ions and Ti2+ ions, special procedures of additional catalyst treatment with pyridine, water, and chloropentafluorobenzene were used to obtain Ti3+ ions that are observable in ESR spectra. In distinction to numerous earlier works performed with the TiCl4/MgCl2 catalyst comprising after the interaction with AlR3 the Ti3+ surface compounds both as isolated ions and clusters (ESR‐silent), this work considers the [η6–BenzeneTiAl2Cl8]/MgCl2 catalyst (TMC‐3) comprising mainly the isolated Ti2+ ions and a new catalyst TMC‐4 obtained by treating the TMC‐3 with chloropentafluorobenzene. This catalyst comprises only the isolated Ti3+ ions both before and after the interaction with triisobutylaluminum. It was shown that in spite of sharp distinctions between the catalysts under consideration concerning titanium oxidation state and the ratio of isolated Ti3+ ions to clustered ones, all these catalysts produce polyethylenes with similar molecular weights and molecular‐weight distributions. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 6362–6372, 2009  相似文献   

7.
We report the infrared, Raman, and surface‐enhanced Raman scattering (SERS) spectra of triruthenium dipyridylamido complexes and of diruthenium mixed nickel metal‐string complexes. From the results of analysis on the vibrational modes, we assigned their vibrational frequencies and structures. The infrared band at 323–326 cm?1 is assigned to the Ru3 asymmetric stretching mode for [Ru3(dpa)4Cl2]0–2+. In these complexes we observed no Raman band corresponding to the Ru3 symmetric stretching mode although this mode is expected to have substantial Raman intensity. There is no frequency shift in the Ru3 asymmetric stretching modes for the complexes with varied oxidational states. No splitting in Raman spectra for the pyridyl breathing line indicates similar bonding environment for both pyridyls in dpa , thus a delocalized structure in the [Ru3]6–8+ unit is proposed. For Ru3(dpa)4(CN)2 complex series, we assign the infrared band at 302 cm?1 to the Ru3 asymmetric stretching mode and the weak Raman line at 285 cm?1 to the Ru3 symmetric stretching. Coordination to the strong axial ligand CN weakens the Ru‐Ru bonding. For the diruthenium nickel complex [Ru2Ni(dpa)4Cl2]0–1+, the diruthenium stretching mode νRu‐Ru is assigned to the intense band at 327 and 333 cm?1 in the Raman spectra for the neutral and oxidized forms, respectively. This implies a strong Ru‐Ru metal‐metal bonding.  相似文献   

8.
DSC was used for heat capacity measurements of pure RuO2 in the temperature range from 300 to 1170 K of solid solutions corresponding to the compositions of (Ti1−x Rux )O2 (x ≤0.15 and x ≥0.85) and in the temperature range from 300 to 1550 K of pure TiO2. The analysis of experimental data obtained within ±2% of accuracy has shown that the characteristic temperatures representing the harmonic lattice vibrations do not strongly depend on the chemical composition x . It was demonstrated that non-harmonic heat capacity is strongly correlated to x. The existence of additional excess heat capacity was observed with the mixed oxide solid solution samples of low Ru content and explained by the defect formation model. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

9.
E.P.R. and Ligand Field Spectra of Chlorovanadates(IV). The Crystal Structure of PPh4[VxTi2–xCl9] (x = 0.15) Black, moisture-sensitive crystals of PPh4[VxTi2–xCl9] (x = 0.15) are formed by the reaction of titanium tetrachloride and PPh4[VCl5] in dichloromethane. Its EPR and ligand field spectra as well as those of PPh4[VCl5] and (PPh4)2[V2Cl9][VCl5] · CH2Cl2 were recorded. In the mixed crystals of PPh4[V0.15Ti1.85Cl9], the existence of [VTiCl9]? ions consisting of trigonally distorted, face sharing octahedra can be proven by the spectra. The spectra of the compounds with [VCl5]? ions can only be explained when a significant Jahn-Teller distortion of the trigonal bipyramids is assumed; this distortion was not detected in the crystal structure determination of (PPh4)2[V2Cl9][VCl5] · CH2Cl2. The crystal structure of PPh4[V0.15Ti1.85Cl9] was determined by X-ray diffraction (2588 independent observed reflexions, R = 0.044). Crystal data: triclinic, space group P1 , a = 1090.4, b = 1217.4, c = 1287.7 pm, α = 73.19°, β = 69.87°, γ = 82.15°, Z = 2. The compound consists of PPh4 and [V0.15Ti1.85Cl9]? ions. In the anions, Ti and V atoms are distributed statistically in the two face sharing octahedra.  相似文献   

10.
Investigations on Electronically Conducting Oxide Systems. XVI. Solid Solutions and Conductivity in the System MgTi2O5? Ti3O5 Solid solution formation is reported for the system Mg1–xTi2–xIVTi2xIIIO5. With increasing x there is at room temperature a transition from the orthorhombic pseudobrookite structure to the monoclinic low-temperature modification of Ti3O5. The X-ray diffraction pattern results are supported by DSC measurements, electrical and magnetic investigations. The tendency of Ti? Ti pair formation in the low-temperature Ti3O5 structure is accompanied by a drop of the activation energy for electrical conductivity and a decreasing susceptibility at high TiIII concentrations.  相似文献   

11.
Ruthenocuprate microstructures and Ru valences have been studied. Electron microscopy reveals short-range order of the RuO6 octahedra rotations into a √2a×√2a×c supercell in Pb2RuSr2Cu2O8Cl. However, reanalysis of neutron diffraction data gives no significant difference between the populations of the rotation states, showing that the coherence length is very short (<100 Å). The Ru valence estimated from the XANES spectrum of Pb2RuSr2Cu2O8Cl is ∼5, in keeping with the physical properties of this material which show that there is essentially no Ru-Cu charge transfer. The Ru valence in doped Ru1−xMxSr2GdCu2O8 (M=Sn, Nb) is ∼4.8 in all samples, verifying a previous rigid band analysis of the charge distribution in these materials.  相似文献   

12.
Na27Ru14O48 has been synthesized in air at 700 °C. The composition and crystal structure of the phase were determined by single crystal X-ray diffraction. The triclinic crystal structure contains isolated planar Ru7O24 plaquettes made from seven edge-sharing RuO6 octahedra. The complex Na:Ru ratio is a result of tilting of the plaquettes to disrupt the packing of nominally hexagonal close packed planes made of Na ions and RuO6 octrahedra. Resistivity measurements show that the material is semiconducting with an activation energy of 0.53 eV. The observed magnetic moment of 3.11 μB per Ru is lower than the expected spin only value, but is within the range seen in other compounds and is too large to indicate that the fundamental magnetic entities are the isolated Ru7O24 plaquettes. Small, reproducible deviations in the Curie-Weiss behavior occur below 200 K and the onset of a broad magnetic transition is seen between 40 and 32 K.  相似文献   

13.
We have investigated the electronic structure of Sr(Ca)Ru1−xCrxO3 using the full potential linearized augmented plane wave method by different approximation such as LSDA and LSDA+U. The LSDA calculation suggest that Cr4+-Ru4+ hybridization is responsible for the high Curie temperature TC in SrRu1−xCrxO3, but it cannot completely describe its physical behavior. Our LSDA+U DOS results for SrRu1−xCrxO3 clearly establishes renormalization of the intra-atomic exchange strength at the Ru sites, arising from the Cr-Ru hybridization. The antiferromagnetic coupling of Cr3+ with Ru4+,5+ lattice increases the screening, which is consistent with the low magnetic moment of the Ru ions. The more distorted Ca-based compounds as compared to the Sr-based systems shows that the hybridization mechanism is not relevant for these compounds. The bigger exchange splitting of Ru 4d and Cr 3d at the Fermi level with Ru4+,5+ and Cr3+,4+ orbital occupancies of CaRu0.75Cr0.25O3 in the LSDA+U calculation, compared with that of the LSDA calculation, shows that repulsion between electrons tend to keep the localized spins from overlapping. The low screening of the Ru t2g electrons increases TC in the Ca-based systems, which is consistent with the both high Ru exchange splitting and magnetic moment. The insulating behavior of the high Cr-doped systems can be explained by considering the Ru4++Cr4+Ru5++Cr3+ charge transfer.  相似文献   

14.
The magnetic, electronic, and structural properties of the solid solutions LaxSr1−xRuO3 and LaxCa1−xRuO3 have been studied by 99Ru Mössbauer spectroscopy and other techniques. The LaxCa1−xRuO3 phases are reported for the first time and have been shown by powder X-ray diffraction measurements to be orthorhombically distorted perovskites. Electrical resistivity measurements on compacted powders show that all the phases are metallic with p 10−3, ohm-cm. Progressive substitution of Sr2+ by La3+ in ferromagnetic SrRuO3 leads to a rapid collapse of the magnetic hyper-fine splitting at 4.2°K. For x = 0.25 some ruthenium ions still experience a magnetic field but for 0.4 x 0.75 only single, narrow resonance lines are observed, consistent both with the complete removal of the ferromagnetism and with the presence of an averaged ruthenium oxidation state in each phase, i.e., Lax3+Sr1−x2+Ru(4−x)+O3 rather than Lax3+Sr1−x2+Rux3+Ru1−x4+O3. LaRuO3 and CaRuO3 both give essentially single-line spectra at 4.2°K, indicating that the ruthenium ions in these oxides are not involved in long-range antiferromagnetic order but are paramagnetic. The solid solutions LaxCa1−xRuO3 (0 < x 0.6) give sharp symmetrical singlets with chemical isomer shifts (relative to the Ru metal) which move progressively from the value characteristic of Ru4+ (−0.303 mm sec−1) toward the value for Ru3+ (−0.557 mm sec−1), consistent with the presence of intermediate ruthenium oxidation states in these phases also.  相似文献   

15.
Sodium ruthenium(III,IV) oxide Na1−x Ru2O4 was synthesized by the solid state reaction of Na2CO3 and RuO2 in inert atmosphere and characterized by X-ray powder diffraction, electron diffraction, and high-resolution transmission electron microscopy. The compound crystallizes in the CaFe2O4-type structure (space group Pnma, Z = 4, a = 9.2641(7) Å, b = 2.8249(3) Å, c = 11.1496(7) Å). Double rutile-like chains of the RuO6 octahedra form a three-dimensional framework, whose tunnels contain sodium cations. The structure contains two crystallographically independent sites of ruthenium atoms randomly occupied by the RuIII and RuIV cations. The superstructure with the doubled b parameter found for one of the samples under study using electron diffraction is caused, probably, by ordering of the Ru cations in the rutile-like chains. The Na1− x Ru2O4 compound exhibits temperature-independent paramagnetism with χ0 = 1.9·10−4 cm3 (mole of Ru−1). Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 10, pp. 1655–1660, October, 2006.  相似文献   

16.
Anodic TiO2 films with a doping of Ru ions on the surface were coincidently prepared by facile one-step anodization for application of the electrode in electrolysis. We found that the amount of Ru ions on the surface of the oxide could be determined based on the applied potential in KRuO4 electrolyte, which provided negative RuO4 ions through dissociation. Overpotentials for the evolution of both O2 and Cl2 were greatly reduced when the coincidently-Ru-doped TiO2 was employed. The Ru-doped electrode prepared at 60 V showed the highest electrocatalytic activity due to the largest amount of Ru incorporation in the oxide.  相似文献   

17.
La3Ru3O11 was prepared by the reaction of La2O3, RuO2, and NaClO3 in a KCl flux under vacuum at 950°C. The crystal structure of this new cubic KSbO3 derivative oxide was determined from single-crystal X-ray diffraction data collected on an automated diffractometer with Mo radiation. Principal crystallographic data: Cubic, space group Pn3; a = 9.451(2), Å; V = 844.2Å3; dX = 7.049 g cm?3. Final discrepancy indices R = 0.036, Rw = 0.042. La3Ru3O11 is isomorphous with Bi3Ru3O11, but is notably different in showing no direct bonding between ruthenium atoms; the closest RuRu contact in this new oxide is 2.994(1) Å.  相似文献   

18.
Single crystals of the title compounds were prepared using a BaCl2 flux and investigated by X-ray diffraction methods using MoKα radiation and a charge coupled device (CCD) detector. The crystal structures of these two new compounds were solved and refined in the hexagonal symmetry with space group P63/mmc, a=5.851(1) Å, c=25.009(5) Å, ρcal=4.94 g cm−3, Z=2 to a final R1=0.069 for 20 parameters with 312 reflections for Ba5Ru2Cl2O9 and space group , a=5.815(1) Å, c=14.915(3) Å, ρcal=5.28 g cm−3, Z=1 to a final R1=0.039 for 24 parameters with 300 reflections for Ba6Ru3Cl2O12. The structure of Ba5Ru2Cl2O9 is formed by the periodic stacking along [001] of three hexagonal close-packed BaO3 layers separated by a double layer of composition Ba2Cl2. The BaO3 stacking creates binuclear face-sharing octahedra units Ru2O9 containing Ru(V). The structure of Ba6Ru3Cl2O12 is built up by the periodic stacking along [001] of four hexagonal close-packed BaO3 layers separated by a double layer of composition Ba2Cl2. The ruthenium ions with a mean oxidation degree +4.67 occupy the octahedral interstices formed by the four layers hexagonal perovskite slab and then constitute isolated trinuclear Ru3O12 units. These two new oxychlorides belong to the family of compounds formulated as [Ba2Cl2][Ban+1RunO3n+3], where n represents the thickness of the octahedral string in hexagonal perovskite slabs.  相似文献   

19.
About Ba6Ru2PtO12Cl2 Single crystals of Ba6Ru2PtO12Cl12 were prepared by a BaCl2 flux and investigated by X-ray methods (D? P3 M1; a = 5,805; c = 15.006 Å; Z = 1). The characteristic face shared M3O12-octahedratriples show an ordered (Ru/Pt/Ru) occupation. Calculation of the Coulomb term of lattice energy support the charge distribution (5+/4+/5+) ions engage three point sites with different coordinations. The connection to other compounds are discussed.  相似文献   

20.
Solid solutions of as-batch composition (Ru1?x Nb x )Sr2(Sm1.4Ce0.6)Cu2O10?δ (the Ru,Nb)-1222 phase), where x = 0.0, 0.25, 0.50, 0.75, or 1.00, have been synthesized and characterized by X-ray diffraction. A correlation is proposed between the refined composition of the Ru-1222 and Nb-1222 phases and their structural features. With increasing oxygen concentration in the Ru-1222 phase, the superconducting transition temperature increases from T c = 28 to T c = 34 K. The composition and magnetic properties of the Ru-1222 phase are affected by the batch composition: unlike in Ru + RuO2 mixtures, the presence of ruthenium in the batch decreases the oxygen proportion and increases the magnetic ordering temperature T m; the phase of as-batch composition NbSr2(Sm1.4Ce0.6)Cu2O10?δ is paramagnetic.  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