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 281 毫秒
1.
Partial substitution of cations and anions in perovskite-type materials is a powerful way to tune the desired properties. The systematic variation of the cations size, the partial exchange of O2− for N3− and their effect on the size of the optical band gap and the thermal stability was investigated here. The anionic substitution resulted in the formation of the orthorhombic perovskite-type oxynitrides Mg0.25Ca0.65Y0.1Ti(O,N)3, Ca1-xYxZr(O,N)3, and Sr1–xLaxZr(O,N)3. A two-step synthesis protocol was applied: i) (nano-crystalline) oxide precursors were synthesized by a Pechini method followed by ii) ammonolysis in flowing NH3 at T = 773 K (Ti) and T = 1273 K (Zr), respectively. High-temperature synthesis of such oxide precursors by solid–state reaction generally resulted in phase separation of the different A-site cations. Changes of the crystal structures were investigated by Rietveld refinements of the powder XRD data, thermal stability by DSC/TG measurements in oxygen atmosphere, oxygen and nitrogen contents by O/N analysis using hot gas extraction technique, and optical band gaps by photoluminescence spectroscopy. By moving from Mg0.25Ca0.65Y0.1Ti(O,N)3 via Ca1–xYxZr(O,N)3 to Sr1–xLaxZr(O,N)3, the degree of tilting of the octahedral network is reduced, as observed by an increase in the BXB angles caused by the simultaneously increasing effective ionic radius of the A-site cation(s). In general, increasing substitution levels on the A-site (Y3+ and La3+) are accompanied by an enhanced replacement of O2− by N3−. In all three systems, this anionic substitution resulted in a reduction of the optical band gap by approximately 1 eV (Ti) and up to 2.1 eV (Zr) compared to the respective oxides. For Mg0.25Ca0.65Y0.1Ti(O,N)3 an optical band gap of 2.2 eV was observed, appropriate for a solar water splitting photocatalyst. The Zr-based oxynitrides required a by a factor of 2 higher nitrogen contents to significantly reduce the optical band gap and the measured values of 2.9 eV–3.2 eV are larger compared to the Ti-based oxynitride. Bulk thermal stability was revealed up to T = 881 K. In general, the thermal stability decreased with increasing substitution levels due to an increasing deviation from the ideal anionic composition as demonstrated by O/N analysis.  相似文献   

2.
Defect-pyrochlores based on the formulation CsM0.5W1.5O6 (M=Ti, Ti/Zr, Zr and Hf) have been studied using neutron diffraction and magic-angle spinning nuclear magnetic resonance (MAS NMR). The results show that structural changes are linearly linked to the change in ionic-radius for the B-site, e.g. the unit cell changes from 10.2763 Å for CsTi0.5W1.5O6 to 10.3820 Å for CsZr0.4W1.6O6. Changes in the NMR chemical shift correlate with the change in electronegativity on the B-site, and show the presence of only one Cs crystal site.  相似文献   

3.
《Mendeleev Communications》2023,33(1):135-137
The local environment of thorium in murataite ceramics (Al,Ca,Ti,Mn,Fe,Zr,Th)Ox and ThO2(001) crystalline film on Si(100) substrate as a reference was explored by X-ray absorption spectroscopy (XAS) for the first time. It was found that Th4+ is located in the center of a cube formed by 8 oxygen atoms [r(Th–O) = 2.37 ± 0.03 Å] in murataite ceramics and ThO2 film. The Th4+ second coordination sphere [r(Th–M) ≈ 3.5 Å] in murataite is represented by 3d metals: titanium, iron or manganese  相似文献   

4.
The vibrational (infrared and Raman) spectroscopy is used in order to identify and characterize the following amphibole minerals with general formula W0–1X2Y5Z8O22(OH)2 (W = Na, K; X = Na, Ca; Y = Mg, Fe2+, Fe3+, Al; Z = Si, Al) originating from the localities in the Republic of Macedonia: glaucophane, Na2(Mg,Fe2+)3(Fe3+,Al)2Si8O22(OH)2; tremolite–actinolite, Ca2(Mg,Fe2+)5Si8O22(OH)2; hornblende (Na,K)0–1Ca2(Mg,Fe2+,Fe3+,Al)5(Si,Al)8 O22(OH)2 and arfvedsonite, NaNa2(Mg,Fe2+)4(Fe3+,Al)Si8O22(OH)2. The chemical composition of these minerals is not necessarily fixed. It is due to the possibility to form solid solution series with other minerals being their end-members (for example, tremolite–ferro-actinolite series, Ca2Mg5Si8O22(OH)2–Ca2Fe2+5Si8O22(OH)2). In this context, it is shown that the intensity and especially the number of the IR bands in the ν(OH) region could serve as a tool for exact mineral identification. Namely, it is based on the presence of different Y cations in various octahedral sites (M1 and M3), which is manifested by different spectral view. On the other hand, the expressed similarities in the 1300–370 cm−1 (IR) and 1200–100 cm−1 regions (Raman) of the spectra are observed due to their common structural characteristics (double chains of SiO4 tetrahedra). Thus, the bands in this region are tentatively prescribed mostly to the vibrations of the SiO4 tetrahedra. The results of our study are compared with the corresponding literature data for the analogous mineral species originating all over the world.  相似文献   

5.
Partially deuterated Ca3Al2(SiO4)y(OH)12−4y-Al(OH)3 mixtures, prepared by hydration of Ca3Al2O6 (C3A), Ca12Al14O33 (C12A7) and CaAl2O4 (CA) phases in the presence of silica fume, have been characterized by 29Si and 27Al magic-angle spinning-nuclear magnetic resonance (MAS-NMR) spectroscopies. NMR spectroscopy was used to characterize anhydrous and fully hydrated samples. In hydrated compounds, Ca3Al2(OH)12 and Al(OH)3 phases were detected. From the quantitative analysis of 27Al NMR signals, the Al(OH)3/Ca3Al2(OH)12 ratio was deduced. The incorporation of Si into the katoite structure, Ca3Al2(SiO4)3−x(OH)4x, was followed by 27Al and 29Si NMR spectroscopies. Si/OH ratios were determined from the quantitative analysis of 27Al MAS-NMR components associated with Al(OH)6 and Al(OSi)(OH)5 environments. The 29Si NMR spectroscopy was also used to quantify the unreacted silica and amorphous calcium aluminosilicate hydrates formed, C-S-H and C-A-S-H for short. From 29Si NMR spectra, the amount of Si incorporated into different phases was estimated. Si and Al concentrations, deduced by NMR, transmission electron microscopy, energy dispersive spectrometry, and Rietveld analysis of both X-ray and neutron data, indicate that only a part of available Si is incorporated in katoite structures.  相似文献   

6.
Nanocrystallinity has been detected in the X-ray absorption spectra of transition metal and rare-earth oxides by (i) removal of d-state degeneracies in the (a) Ti and Sc L3 spectra of TiO2 and LaScO3, respectively, and (b) O K1 spectra of Zr(Hf)O2, Y2O3, LaScO3 and LaAlO3, and by the (ii) detection of the O-atom vacancy in the O K1 edge ZrO2–Y2O3 alloys. Spectroscopic detection is more sensitive than X-ray diffraction with a limit of ∼2 nm as compared to >5 mm. Other example includes detection of ZrO2 nanocrystallinity in phase-separated Zr(Hf) silicate alloys.  相似文献   

7.
Ma X  Li Y 《Analytica chimica acta》2006,579(1):47-52
This paper describes a rapid, accurate and precise method for the determination of trace Fe, Hf, Mn, Na, Si and Ti in high-purity zirconium dioxide (ZrO2) powders by inductively coupled plasma atomic emission spectrometry (ICP-AES). The samples were dissolved by a microwave-assisted digestion system. Four different digestion programs with various reagents were tested. It was found that using a mixture of sulfuric acid (H2SO4) and ammonium sulfate ((NH4)2SO4), the total sample dissolution time was 30 min, much shorter than that required for conventional digestion in an opening system. The determination of almost all of the target analytes suffered from spectral interferences, since Zr shows a line-rich atomic emission spectrometry. The wavelet transform (WT), a recently developed mathematical technique was applied to the correction of spectral interference, and more accurate and precise results were obtained, compared with traditional off-peak background correction procedure. Experimental work revealed that a high Zr concentration would result in a significant decrease in peak height of the analyte lines, which was corrected by standard addition method. The performance of the developed method was evaluated by using synthetic samples. The recoveries were in the range of 87-112% and relative standard deviation was within 1.1-3.4%. The detection limits (3σ) for Fe, Hf, Mn, Na, Si and Ti were found to be 1.2, 13.3, 1.0, 4.5, 5.8 and 2.0 μg g−1, respectively. The results showed that with the microwave-assisted digestion and the WT correction, the detection limits have improved by a factor of about 5 for Fe, 4 for Mn and Ti, 3 for Si, and 2 for Hf and Na, respectively, in comparison with conventional open-system digestion and off-peak correction. The proposed technique was applied to the analysis of trace elements above-mentioned in three types of ZrO2 powders.  相似文献   

8.
The use of direct current arc atomic emission spectrometry (DC-arc-AES) with a CCD spectrometer for the direct determination of the trace impurities Al, Ca, Cr, Cu, Fe, Mg, Mn, Na, Ni, Si, Ti, and Zr in three well characterized boron carbide powders is described. The detection limits obtained by the procedure were found to be between 0.2 (Mg) and 25 (Na) ??g?g?1 for the above elements. Three boron carbide powder samples with trace element concentrations between 0.9 (Cu) and 934 (Si) ??g?g?1 for Al, Ca, Cr, Cu, Fe, Mg, Mn, Na, Ni, Si, Ti, and Zr ?? including the standard reference material ERM?-ED102 ?? were analyzed by DC-arc-AES. The relative standard deviations for 9 measurements when using 5.0?±?0.3?mg of the respective samples were found to vary from 6.2 to 27% for Al and Cu, respectively. The trace elements Al, Ca, Cr, Cu, Fe, Mn, Ni, Si, Ti and Zr could be determined in the standard reference material and their concentrations determined by DC-arc AES were found to be between 89 and 116% of the accepted values. Fe and Ti were determined by DC-arc AES in the three boron carbide samples as well as in Al2O3, BN, SiC, coal fly ash, graphite and obsidian rock. The correlation coefficients of the plots of the net intensities versus the accepted values over the concentration ranges from 18 to 1750 and from 6 to 8000???g?g?1 are 0.999 and 0.990 for Fe and Ti, respectively.
Figure
Coupling of DC arc to a CCD spectrometer  相似文献   

9.
Our recent extensive research on Lewis acid catalysts with a weak base for the cationic polymerization of vinyl ethers led to unprecedented living reaction systems: fast living polymerization within 1–3 s; a wide choice of metal halides containing Al, Sn, Fe, Ti, Zr, Hf, Zn, Ga, In, Si, Ge, and Bi; and heterogeneously catalyzed living polymerization with Fe2O3. The use of added bases for the stabilization of the propagating carbocation and the appropriate selection of Lewis acid catalysts were crucial to the success of such new types of living polymerizations. In addition, the base‐stabilized living polymerization allowed the quantitative synthesis of star‐shaped polymers with a narrow molecular weight distribution via polymer‐linking reactions and the precision synthesis and self‐assembly of stimuli‐responsive block copolymers. © 2007 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 45: 1801–1813, 2007.  相似文献   

10.
The present paper mainly studied the phase formation and reaction pathway of the Al–Ti–Si system in detail by thermal analysis combined with XRD and SEM observations. The phase formation sequence in Al–Ti–Si system from starting mixtures to final products with increasing temperature can be described as following: Al(l) + Ti(s) + Si(s) → (Al–Si)(l) + Ti(s) + Si(s) → Ti(Al,Si)3(s) + Si(s)Ti5(Si,Al)3 + Al(l). More importantly, the solubility of Si in Ti(Al,Si)3 decreased gradually while that of Al in Ti5(Si,Al)3 increased with temperature increasing, suggesting the transportation of Si atoms from intermediate aluminides Ti(Al,Si)3 to final stable silicides Ti5(Si,Al)3 and hence further confirming the formation of Ti5(Si,Al)3 at the expense of Ti(Al,Si)3.  相似文献   

11.
Rare-earth (RE) doped Ba(Zr,Ti)O3 (BZT) thin films were prepared by rf-magnetron sputtering from a Ba0.90Ln0.067Zr0.09Ti0.91O3 (Ln=La, Nd) target. The films were deposited at a substrate temperature of 600 °C in a high oxygen pressure atmosphere. X-ray diffraction (XRD) patterns of RE-BZT films revealed a 〈001〉 epitaxial crystal growth on Nb-doped SrTiO3, 〈001〉 and 〈011〉 growth on single-crystal Si, and a 〈111〉-preferred orientation on Pt-coated Si substrates. Scanning electron microscopy (SEM) showed uniform growth of the films deposited, along with the presence of crystals of about half-micron size on the film's surface. Transmission electron microscopy (TEM) evidenced high crystalline films with thicknesses of about 100 nm for 30 min of sputtering. Electron-probe microanalysis (EPMA) corroborated the growth rate (3.0-3.5 nm/min) of films deposited on Pt-coated Si substrates. X-ray photoelectron spectroscopy (XPS), in depth profile mode, showed variations in photoelectron Ti 2p doublet positions at lower energies with spin-orbital distances characteristic of BaTiO3-based compounds. The XPS analysis revealed that lanthanide ions positioned onto the A-site of the BZT-perovskite structure increasing the MO6-octahedra distortion (M=Ti, Zr) and, thereby, modifying the Ti-O binding length. Polarization-electric field hysteresis loops on Ag/RE-doped BZT/Pt capacitor showed good ferroelectric behavior and higher remanent polarization values than corresponding non-doped system.  相似文献   

12.
A new layered carbide, [Zr0.72(3)Y0.28(3)]Al4C4, has been synthesized and characterized by X-ray powder diffraction, transmission electron microscopy and energy dispersive X-ray spectroscopy (EDX). The atom ratios [Zr:Y] were determined by EDX, and the initial structure model was derived by the direct methods, and further refined by Rietveld method. The crystal is trigonal (space group , Z=1) with lattice dimensions of a=0.333990(5) nm, c=1.09942(1) nm and V=0.106209(2) nm3. This compound shows an intergrowth structure with [Zr0.72Y0.28C2] thin slabs separated by Al4C3-type [Al4C4] layers. It is a new member with l=1 and m=1 of the homologous series, the general formula of which is (MC)l(T4C3)m (l=1, 2 and 3, m=1 and 2, M=Zr, Y and Hf, T=Al, Si and Ge).  相似文献   

13.
Aromatic aldehydes and aryl isocyanates do not react at room temperature. However, we have shown for the first time that in the presence of catalytic amounts of group(IV) n-butoxide, they undergo metathesis at room temperature to produce imines with the extrusion of carbon dioxide. The mechanism of action has been investigated by a study of stoichiometric reactions. The insertion of aryl isocyanates into the metal n-butoxide occurs very rapidly. Reaction of the insertion product with the aldehyde is responsible for the metathesis. Among the n-butoxides of group(IV) metals, Ti(OnBu)4 (8aTi) was found to be more efficient than Zr(OnBu)4 (8aZr) and Hf(OnBu)4 (8aHf) in carrying out metathesis. The surprisingly large difference in the metathetic activity of these alkoxides has been probed computationally using model complexes Ti(OMe)4 (8bTi), Zr(OMe)4 (8bZr) and Hf(OMe)4 (8bHf) at the B3LYP/LANL2DZ level of theory. These studies indicate that the insertion product formed by Zr and Hf are extremely stable compared to that formed by Ti. This makes subsequent reaction of Zr and Hf complexes unfavorable.  相似文献   

14.
Conduction band edge d-states are compared for complex oxides: (i) mixed tetravalent–trivalent ZrO2–Y2O3 alloys, (ii) tetravalent Zr(Hf)O2–TiO2 alloys, and (iii) trivalent La scandate and aluminate. Low Y2O3 content cubic ZrO2–Y2O3 alloys display two crystal-field split 4d-features in O K1 spectra. Alloys with higher Y2O3 content, as well as Zr(Hf)O2–TiO2 alloys display increased d-state multiplicity. O K1 spectra of perovskite-structured LaScO3 and LaAlO3 indicate Jahn–Teller d-state term-splittings with contributions from both trivalent atomic species.  相似文献   

15.
Jahn–Teller (J–T) term-split states in nanocrystalline transition metal and trivalent rare earth elemental and complex oxides reduce the band gap, and tunnelling barrier height at interfaces with crystalline Si substrates. These states are identified by x-ray absorption spectroscopy and spectroscopic ellipsometry. Alloys for suppression of J–T d-state degeneracy removal are identified as: (i) non-crystalline Zr/Hf silicates and Si oxynitrides and (ii) ZrO2–Y2O3 alloys with high concentrations of randomly distributed O-vacancies that promote cubic crystalline symmetry.  相似文献   

16.
We have employed aliovalent A-site cation substitution, LaIII-for-SrII, to dope the Sr(Fe0.5Ta0.5)O3 perovskite oxide with electrons. Essentially single-phase samples of (Sr1−xLax)(Fe0.5Ta0.5)O3 were successfully synthesized up to x≈0.3 in a vacuum furnace at 1400 °C. The samples were found to crystallize (rather than with orthorhombic symmetry) in monoclinic space group P21/n that accounts for the partial ordering of the B-site cations, Fe and Ta. With increasing La-substitution level, x, the degree of Fe/Ta order was found to increase such that the La-richest compositions are best described by the B-site ordered double-perovskite formula, (Sr,La)2FeTaO6. From Fe L3 and Ta L3 XANES spectra it was revealed that upon electron doping the two B-site cations, FeIII and TaV, are both prone to reduction. Magnetic susceptibility measurements showed spin-glass type behaviour for all the samples with a transition temperature slightly increasing with increasing x.  相似文献   

17.
The reaction of Group 4 metal alkoxides ([M(OR)4]) with the potentially bidentate ligand, 2-hydroxy-pyridine (2-HO-(NC5H4) or H-PyO), led to the isolation of a family of compounds. The products isolated from the reaction of [M(OR)4] [where M = Ti, Zr, or Hf; OR = OPri (OCH(CH3)2), OBut (OC(CH3)3), or ONep (OCH2C(CH3)3] under a variety of stoichiometries with H-PyO were identified by single crystal X-ray diffraction as [(OPri)2(PyO-κ2(O,N))Ti(μ-OPri)]2 (1), [(ONep)2Ti(μ(O)-PyO-κ2(O,N))2(μ-ONep)Ti(ONep)3] (2), [(ONep)2Ti(μ(O)-PyO-κ2(O,N))(η1(N),μ(O)-PyO)(μ-O)Ti(ONep)2]2 (2a), [H][(PyO-κ2(O,N))(η1(O)-PyO)Ti(ONep)3] (3), [(OR)2Zr(μ(O)-PyO-κ2(O,N))2(μ-OR)Zr(OR)3] (OR = OBut (4), ONep (5)), [(OR)2Zr(μ(O,N)-PyO-κ2(O,N))2(μ(O,N)-PyO)Zr(OR)3] (OR = OBut (6), ONep (7)), [[(OBut)2Zr(μ(O)-PyO-(κ2(N,O))(μ(O,N)-PyO)2Zr(OBut)](μ3-O)]2 (6a), [[(ONep)(PyO-κ2(N,O))Zr(μ(O,N)-PyO-κ2(N,O))2(μ(O)-PyO-κ2(N,O))Zr(ONep)](μ3-O)]2 (7a), [(OBut)(PyO-κ2(O,N))Zr(μ(O)-PyO-κ2(O,N))2((μ(O,N)-PyO)Zr(OBut)3] (8), [(OBut)2Hf(μ(O)-PyO-κ2(N,O))2(μ-OBut)Hf(OBut)3] (9), [(OR)2 M(μ(O)-PyO-κ2(N,O))2(μ(O,N)-PyO)M(OR)3] (OR = OBut (10), ONep (11)), and [(ONep)3Hf(μ-ONep)(η1(N),μ(O)-PyO)]2Hf(ONep)2 (12)·tol. The structural diversity of the binding modes of the PyO led to a number of novel structure types in comparison to other pyridine alkoxy derivatives. The majority of compounds adopt a dinuclear arrangement (1, 2, 411) but oxo-based tetra- (2a and 7a), tri- (12), and monomers (3) were observed as well. Compounds 112 were further characterized using a variety of analytical techniques including Fourier Transform Infrared Spectroscopy, elemental analysis, and multinuclear NMR spectroscopy.  相似文献   

18.
This work reports a comparative study of the catalytic behaviour for a series of metallocenes derived from Ti, Zr, Hf and Nb, which after activation with methylaluminoxane can polymerize ethylene. Results show that the Zr metallocene with a  (CH3)2Si Bridge presents the highest activity, and the metallocenes based on Hf and Nb do not show any significant activity under the tested conditions.  相似文献   

19.
Titanomagnetite (Fe3−xTixO4) nanoparticles were synthesized by room temperature aqueous precipitation, in which Ti(IV) replaces Fe(III) and is charge compensated by conversion of Fe(III) to Fe(II) in the unit cell. A comprehensive suite of tools was used to probe composition, structure, and magnetic properties down to site-occupancy level, emphasizing distribution and accessibility of Fe(II) as a function of x. Synthesis of nanoparticles in the range 0 ? x ? 0.6 was attempted; Ti, total Fe and Fe(II) content were verified by chemical analysis. TEM indicated homogeneous spherical 9-12 nm particles. μ-XRD and Mössbauer spectroscopy on anoxic aqueous suspensions verified the inverse spinel structure and Ti(IV) incorporation in the unit cell up to x ? 0.38, based on Fe(II)/Fe(III) ratio deduced from the unit cell edge and Mössbauer spectra. Nanoparticles with a higher value of x possessed a minor amorphous secondary Fe(II)/Ti(IV) phase. XANES/EXAFS indicated Ti(IV) incorporation in the octahedral sublattice (B-site) and proportional increases in Fe(II)/Fe(III) ratio. XA/XMCD indicated that increases arise from increasing B-site Fe(II), and that these charge-balancing equivalents segregate to those B-sites near particle surfaces. Dissolution studies showed that this segregation persists after release of Fe(II) into solution, in amounts systematically proportional to x and thus the Fe(II)/Fe(III) ratio. A mechanistic reaction model was developed entailing mobile B-site Fe(II) supplying a highly interactive surface phase that undergoes interfacial electron transfer with oxidants in solution, sustained by outward Fe(II) migration from particle interiors and concurrent inward migration of charge-balancing cationic vacancies in a ratio of 3:1.  相似文献   

20.
The ternary stoichiometric perovskite compounds, Na0.75Ln0.25Ti0.5Nb0.5O3 (Ln=La, Pr, Nd, Sm, Eu, Gd, Tb, Dy, Ho, Er, Tm) are intermediate members of the NaNbO3-Na0.5Ln0.5TiO3 solid solution series. The compounds were synthesized by standard ceramic methods at 1300 °C followed by annealing at 800 °C and quenching to ambient conditions. Rietveld analysis of the powder X-ray diffraction patterns shows that the compounds with Ln ranging from Pr to Tm adopt the orthorhombic space group Pbnm (ab≈√2ap; c≈2ap; Z=4) and the GdFeO3 structure. In contrast, Na0.75La0.25Ti0.5Nb0.5O3 adopts the orthorhombic space group Cmcm (abc≈2ap; Z=4). All cations located at the A- and B-sites are disordered in these compounds. The unit cell parameters and cell volumes of the compounds decrease regularly with increasing atomic number of the Ln cation. The Pbnm compounds with Ln from Sm to Tm have A-site cations in eight-fold coordination. A-site cations in the Pr and Nd compounds are considered to be in ten-fold coordination. Analysis of the crystal chemistry of the Pbnm compounds shows that B-site cations enter the second coordination sphere of the A-site cations for compounds with Ln from Tb to Tm as the A-B intercation distances are less than the maximum A-IIO(2) bond lengths. The [111] tilt angles of the (Ti,Nb)O6 polyhedra in the Pbnm compounds increase with increasing atomic number from 11.1° to 15.8° and are less than those observed in lanthanide orthoferrite and orthoscandate perovskites. These data are considered as relevant to the sequestration of lanthanide fission products in perovskite and the structure of lanthanide-bearing perovskite-structured minerals.  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