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 15 毫秒
1.
《Solid State Sciences》2012,14(5):580-582
Local geometry and bond ionicity around the nitride ions in simple perovskite oxynitrides ATaO2N (A = Ca, Sr, Ba) have been investigated by solid-state magic-angle spinning (MAS) NMR spectroscopy. From all three compounds, fairly sharp 14N NMR peaks were observed, suggestive of the symmetric coordination environment of nitride ions. The 14N chemical shifts of ATaO2N, δ = 269–272 ppm relative to NH4Cl (δ = 0 ppm), are correlated to the bond ionicity, based on the N−Ta bond distances and Ta–N–Ta bond angles determined from the Rietveld refinement of neutron diffraction patterns. The 1H NMR measured for BaTaO2N presented a peak corresponding to H2O, implying that the polycrystalline surface of present oxynitride phases is covered by hydroxide terminals.  相似文献   

2.
Although deemed important to δ18O measurement by on‐line high‐temperature conversion techniques, how the GC conditions affect δ18O measurement is rarely examined adequately. We therefore directly injected different volumes of CO or CO–N2 mix onto the GC column by a six‐port valve and examined the CO yield, CO peak shape, CO–N2 separation, and δ18O value under different GC temperatures and carrier gas flow rates. The results show the CO peak area decreases when the carrier gas flow rate increases. The GC temperature has no effect on peak area. The peak width increases with the increase of CO injection volume but decreases with the increase of GC temperature and carrier gas flow rate. The peak intensity increases with the increase of GC temperature and CO injection volume but decreases with the increase of carrier gas flow rate. The peak separation time between N2 and CO decreases with an increase of GC temperature and carrier gas flow rate. δ18O value decreases with the increase of CO injection volume (when half m/z 28 intensity is <3 V) and GC temperature but is insensitive to carrier gas flow rate. On average, the δ18O value of the injected CO is about 1‰ higher than that of identical reference CO. The δ18O distribution pattern of the injected CO is probably a combined result of ion source nonlinearity and preferential loss of C16O or oxygen isotopic exchange between zeolite and CO. For practical application, a lower carrier gas flow rate is therefore recommended as it has the combined advantages of higher CO yield, better N2–CO separation, lower He consumption, and insignificant effect on δ18O value, while a higher‐than‐60 °C GC temperature and a larger‐than‐100 µl CO volume is also recommended. When no N2 peak is expected, a higher GC temperature is recommended, and vice versa. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

3.
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   

4.
A new mixed-metal polyborate, Na5Li[B12O18(OH)6]·2H2O (1), has been synthesized using solvothermal method and characterized by IR spectroscopy, thermogravimetric analysis, UV–Vis spectroscopy, powder and single-crystal X-ray diffraction, respectively. It crystallizes in the trigonal space group R-3c (No. 167) with unit cell parameters of a = b = 9.6767(6) Å, c = 36.358(5) Å, and Z = 6. Its structure features unprecedented 3D framework constructed from novel honeycomb-shaped inorganic Na-O sheets with unique 12-MR sodium rings and supramolecular polyborate 2D layers of lithium-centered [B12O18(OH)6]6-. UV–Vis spectral characterization indicates that compound 1 is a wide-band-gap semiconductor.  相似文献   

5.
Oxygen permeation through dense ceramic membranes of perovskite-like SrCo0.9−xFe0.1CrxO3−δ (x = 0.01–0.05), Sr1−xyLnxCoO3−δ(Ln = La, Nd, Sm, Gd; x = 0.30–0.35; y = 0–0.10), SrCo1−xTixO3−δ (x = 0.05–0.20) and LaM1−xNixO3−δ (M = Ga, Co, Fe; x = 0–0.6) was studied. The SrCoO3−δ-based solid solutions with cubic perovskite structure were found to exhibit highest permeation fluxes compared to other membranes. However, high thermal expansion coefficients and interaction with gas species such as carbon dioxide may complicate the employment of SrCoO3−δ membranes for oxygen separation membranes. Alternatively, the LaGa1−xNixO3−δ (x = 0.2–0.5) perovskites, having significant permeation fluxes as well as thermal expansion coefficients in the range of (10.8–11.6) × 10−6 K−1, were demonstrated to be suitable as membrane materials at oxygen pressures from 1 × 10−2 to 2 × 104 Pa. Testing oxygen permeation at oxygen partial pressures of 1–60 atm showed that only oxides with a high oxygen deficiency such as SrCo0.85Ti0.15F3−δ possess sufficient oxygen permeation fluxes. The oxygen permeability of perovskites on the basis of LaGaO3 and LaCoO3−δ was found to be negligible at oxygen pressures above 15 atm, caused by low oxygen vacancy concentration and ionic conductivity of such ceramic materials.  相似文献   

6.
The effect of Sm doping on CeO2 for its use as a solid electrolyte material for intermediate temperature solid oxide fuel cells (IT-SOFCs) has been explored here. Ce1−xSmxO2−δ (x = 0.1–0.3) samples are successfully synthesized by carbonate co-precipitation method. TG–DTA, XRD, Raman, UV–Vis, FT-IR, SEM and ac-impedance are used for structural and electrical characterization. From the XRD patterns, well-crystalline cubic fluorite structured solid solution is confirmed. Lattice parameters increased with increase in Sm3+ while the crystallite size decreased. The optical absorption spectra exhibits a red shift for Sm3+ doped CeO2. Raman spectra show an intense peak at 463 cm−1, a characteristic peak for doped ceria. SEM shows cluster like particles. Based on ac-impedance data, the total oxygen ionic conductivity is highest for Ce0.8Sm0.2O2−δ in the temperature range of 473–623 K.  相似文献   

7.
The effects of doping the mixed-conducting (La,Sr)FeO3−δ system with Ce and Nb have been examined for the solid-solution series, La0.5−2xCexSr0.5+xFeO3−δ (x = 0–0.20) and La0.5−2ySr0.5+2yFe1−yNbyO3−δ (y = 0.05–0.10). Mössbauer spectroscopy at 4.1 and 297 K showed that Ce4+ and Nb5+ incorporation suppresses delocalization of p-type electronic charge carriers, whilst oxygen nonstoichiometry of the Ce-containing materials increases. Similar behavior was observed for La0.3Sr0.7Fe0.90Nb0.10O3−δ at 923–1223 K by coulometric titration and thermogravimetry. High-temperature transport properties were studied with Faradaic efficiency (FE), oxygen-permeation, thermopower and total-conductivity measurements in the oxygen partial pressure range 10−5–0.5 atm. The hole conductivity is lower for the Ce- and Nb-containing perovskites, primarily as a result of the lower Fe4+ concentration. Both dopants decrease oxide-ion conductivity but the effect of Nb-doping on ionic transport is moderate and ion-transference numbers are higher with respect to the Nb-free parent phase, 2.2 × 10−3 for La0.3Sr0.7Fe0.9Nb0.1O3−δ cf. 1.3 × 10−3 for La0.5Sr0.5FeO3−δ at 1223 K and atmospheric oxygen pressure. The average thermal expansion coefficients calculated from dilatometric data decrease on doping, varying in the range (19.0–21.2) × 10−6 K−1 at 780–1080 K.  相似文献   

8.
The dc and ac electrical conductivity of barium tellurite borate glass doped with Nd2O3 in the composition 50 B2O3- (20-X) BaO- 20TeO2 10 LiF or Li2O where x = 0.5, 1, 1.5 and 2 Nd2O3 were measured in the temperature range 303–648 K and in the frequency range 0.1–100 kHz. The dc and ac conductivities values increase, whereas the activation energy of conductivities decreases with increasing Nd2O3 content in the glasses containing LiF and by the replacement of LiF by Li2O the conductivity was found to decrease with addition of Nd2O3. The electrical conduction has been observed to be due to small polaron hopping at high temperatures. The frequency dependence of the ac conductivity follows the power law σAC (ω) = A ωs. The frequency exponent (s) values (in the range 0.94 and 0.33) decreases with increasing temperature. The dielectric constant and dielectric loss increased with increasing temperature and decreased with increase in frequency for all glasses studied. In LiF glasses, it is observed that, the values of ?\ and tan δ are observed to increase with the addition of Nd2O3 whereas they decrease in the glasses containing Li2O. The electrical modulus formalism has been used for studying electrical relaxation behavior in studied glasses. It is for first time that the Nd2O3 doped barium tellurite borate glasses have been investigated for dc and ac conductivities and dielectric properties over a wide range of frequency and temperature.  相似文献   

9.
Colourless, water- and air-stable single crystals of cerium(III) oxoarsenate(III) Ce[AsO3] were prepared by the reaction of cerium metal (Ce) and arsenic sesquioxide (As2O3) in the presence of cesium chloride (CsCl) as fluxing agent at 750 °C in an evacuated silica ampoule. Ce[AsO3] crystallizes monoclinically (a = 902.89(8), b = 782.54(7), c = 829.68(7) pm, β = 103.393(3)°, Z = 8) in the space group P21/c and is isotypic with α-Pb[SeO3]. There are two crystallographically different Ce3+ positions. (Ce1)3+ is coordinated by nine oxygen atoms (d(Ce–O) = 244–286 pm) and (Ce2)3+ by only eight (d(Ce–O) = 239–273 pm). Both crystallographically different As3+ cations form discrete ψ1 tetrahedra [AsO3]3− (d(As–O) = 174–179 pm), which are attached to the Ce3+ cations via edges and corners. The second monoclinic modification of Ce[AsO3] with the lattice parameters a = 439.32(4), b = 529.21(5), c = 617.34(6) pm and β = 105.369(3)° with Z = 2 was obtained by high-pressure synthesis (11 GPa, 1200 °C) and has both a higher density (6.31 vs. 6.13 g · cm−3) and a higher calculated Madelung part of the lattice energy (15,155 vs. 15,132 kJ · mol−1). It adopts the space group P21/m, crystallizing isotypically with La[AsO3], β-Pb[SeO3], Pb[SO3] (scotlandite) or K[ClO3] and exhibits nine-fold coordinated Ce3+ cations exclusively (d(Ce–O) = 254–287 pm) along with tripodal [AsO3]3− anions (d(As–O) = 175–176 pm). Raman spectroscopy on both phases of Ce[AsO3] shows stretching vibrations between 769 and 731 cm−1 as well as asymmetric vibrations in the range of 659–617 cm−1. The symmetric bending mode vibrations emerge in an interval from 340 to 410 cm−1 and the asymmetric bending modes range between 230 and 290 cm−1.  相似文献   

10.
Stable oxygen isotopic compositions of a coral colony ofPorites lutea obtained on a core allowed the reconstruction of a 56-a (1943–1998) proxy record of the sea surface temperatures. This coral δ18O data are from the east of Hainan Island water (22°20’N, 110°39’E), South China Sea. The relationship between δ18O in the skeletal aragonite carbonate and the sea surface temperature (SST) is SST = -5.36 δ18OPDB-3.51 (r = 0.73,n = 470), dδ18O/d(SST) = -0.187?/ °C; and the thermometer was set at monthly resolution. The 56-a (1943–1998) proxy record of the sea surface temperatures reflected the same change trend in the northern part of South China Sea as the air temperature change trend in China.  相似文献   

11.
The lanthanide complexes derived from (3,5,13,15-tetramethyl 2,6,12,16,21-22-hexaazatricyclo[15.3.I1-17I7-11]cosa-1(21),2,5,7,9,11(22),12,15,17,19-decane) were synthesized. The complexes were found to have general composition [Ln(L)X2·H2O]X, where Ln = La3+, Ce3+, Nd3+, Sm3+ and Eu3+ and X = NO3? and Cl?. The ligand was characterized by elemental analyses, IR, Mass, and 1H NMR spectral studies. All the complexes were characterized by elemental analyses, molar conductance measurements, magnetic susceptibility measurements, IR, Mass, electronic spectral techniques and thermal studies. The ligand acts as a hexadentate and coordinates through four nitrogen atoms of azomethine groups and two nitrogen of pyridine ring. The lanthanum complexes are diamagnetic while the other Ln(III) complexes are paramagnetic. The spectral parameters i.e. nephelauxetic ratio (β), covalency factor (b1/2), Sinha parameter (δ%) and covalency angular overlap parameter (η) have been calculated from absorption spectra of Nd(III) and Sm(III) complexes. These parameters suggest the metal–ligand covalent bonding. In the present study, the complexes were found to have coordination number nine.  相似文献   

12.
Although the advantages of online δ18O analysis of organic compounds make its broad application desirable, researchers have encountered NO+ isobaric interference with CO+ at m/z 30 (e.g. 14N16O+, 12C18O+) when analyzing nitrogenous substrates. If the δ18O value of inter‐laboratory standards for substrates with high N:O value could be confirmed offline, these materials could be analyzed periodically and used to evaluate δ18O data produced online for nitrogenous unknowns. To this end, we present an offline method based on modifications of the methods of Schimmelmann and Deniro (Anal. Chem. 1985; 57: 2644) and Sauer and Sternberg (Anal. Chem. 1994; 66: 2409), whereby all the N2 from the gas products of a chlorinated pyrolysis was eliminated, resulting in purified CO2 for analysis via a dual‐inlet isotope ratio mass spectrometry system. We evaluated our method by comparing observed δ18O values with previously published or inter‐laboratory calibrated δ18O values for five nitrogen‐free working reference materials; finding isotopic agreement to within ±0.2‰ for SIGMA® cellulose, IAEA‐CH3 cellulose (C6H10O5) and IAEA‐CH6 sucrose (C12H22O11), and within ±1.8‰ for IAEA‐601 and IAEA‐602 benzoic acids (C7H6O2). We also compared the δ18O values of IAEA‐CH3 cellulose and IAEA‐CH6 sucrose that was nitrogen‐'doped' with adenine (C5H5N5), imidazole (C3H4N2) and 2‐aminopyrimidine (C4H5N3) with the undoped δ18O values for the same substrates; yielding isotopic agreement to within ±0.7‰. Finally, we provide an independent analysis of the δ18O value of IAEA‐600 caffeine (C8H10N4O2), previously characterized using online systems exclusively, and discuss the reasons for an average 1.4‰ enrichment in δ18O observed offline relative to the consensus online δ18O value. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

13.
59Co chemical shifts were computed at the GIAO‐B3LYP level for [Co(CN)6]3?, [Co(H2O)6]3+, [Co(NH3)6]3+, and [Co(CO)4]? in water. The aqueous solutions were modeled by Car–Parrinello molecular dynamics (CPMD) simulations, or by propagation on a hybrid quantum‐mechanical/molecular‐mechanical Born–Oppenheimer surface (QM/MM‐BOMD). Mean absolute deviations from experiment obtained with these methods are on the order of 400 and 600 ppm, respectively, over a total δ(59Co) range of about 18 000 ppm. The effect of the solvent on δ(59Co) is mostly indirect, resulting primarily from substantial metal–ligand bond contractions on going from the gas phase to the bulk. The simulated solvent effects on geometries and δ(59Co) values are well reproduced by using a polarizable continuum model (PCM), based on optimization and perturbational evaluation of quantum‐mechanical zero‐point corrections.  相似文献   

14.
The measurement of the oxygen stable isotope content in organic compounds has applications in many fields, ranging from paleoclimate reconstruction to forensics. Conventional High‐Temperature Conversion (HTC) techniques require >20 µg of O for a single δ18O measurement. Here we describe a system that converts the CO produced by HTC into CO2 via reduction within a Ni‐furnace. This CO2 is then concentrated cryogenically, and 'focused' into the isotope ratio mass spectrometry (IRMS) source using a low‐flow He carrier gas (6–8 mL/min). We report analyses of benzoic acid (C7H6O2) reference materials that yielded precise δ18O measurement down to 1.3 µg of O, suggesting that our system could be used to decrease sample requirement for δ18O by more than an order of magnitude. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

15.
Nitrogen and oxygen isotope ratios (δ15N and δ18O) in nitrates are important in hydrology, oceanography, atmospheric chemistry, and agriculture. This paper reviews current isotope ratio mass spectrometry methods for the determination of nitrogen and oxygen isotopes in nitrates that include graphite combustion, AgNO3-ion exchange, Ba(NO3)2-acetone, two-step chemical conversion, bacterial approaches, and off-axis integral cavity output laser spectroscopy. This paper introduces the principles, processes, advantages, and disadvantages of these procedures. Future studies should focus on the determination of oxygen isotopes. The development of novel spectroscopic or other isotopic methodology may also be new research directions. In addition, in some nitrates, δ17O is more sensitive than traditional δ18O and further development for this isotope is of interest. In addition, the determination of δ17O may be used to more accurately evaluate δ18O.  相似文献   

16.
A new environmental cell allowing for the independent synchronous collection of the near- and mid-infrared spectra (12,000–600 cm−1) in the diffuse reflection and attenuated total reflection (ATR) modes, respectively, is reported. The cell is employed to study in real time the dehydration of the phyllosilicate mineral sepiolite, Mg8Si12O30(OH)4(OH2)4·wH2O, in both its natural form and after in situ deuteration at ambient. The spectra are obtained under dynamic purging with dry N2 and compared to those of the same material conditioned over saturated salt solutions. Sepiolite is an important industrial mineral with a modulated structure of alternating tunnels and ribbons. Its mild drying is associated with pronounced vibrational spectral changes due to the removal of surface and zeolitic H2O and the concomitant structural relaxation of the ribbons. Detailed assignments are provided for the fundamental, combination and overtone spectrum of H2O confined in the tunnels of sepiolite, SiOH groups on the external surface of the particles, and Mg3OH groups in the 2:1 ribbons. The spectra are discussed in comparison to those of palygorskite (modulated phyllosilicate with narrower ribbons and tunnels), talc (trioctahedral magnesian phyllosilicate without modulation) and high-surface area silica. It is demonstrated that sepiolite exhibits three discrete states of zeolitic hydration at ambient temperature: Besides the previously known hydrated (w = 7–8) and dry (w = 0–1) states which dominate the spectra above 30% and below 3% relative humidity, respectively, a hitherto unknown intermediate (w = 4–5) is found in the 3–10% range. The new state is most conveniently identified in the near-infrared by a ν02 Mg3O-H stretching mode at 7205 cm−1 (ν01 = 3686 cm−1, X = 83.5 cm−1) and a characteristic H2O combination band at 5271 cm−1 (D2O: 3908 cm−1).  相似文献   

17.
Nd3+ doped H3BO3–PbO–TeO2–RF (R = Li, Na and K) glasses were prepared through melt quenching technique. Optical absorption and near infrared (NIR) fluorescence spectra were recorded at room temperature. The spectral intensities were analyzed in terms of the Judd–Ofelt (J–O) parameters (Ωλ = 2, 4, 6). The covalency effect of Nd–O bond on the J–O parameters was estimated from the relative absorbance ratio (R) between 4I9/2  4F7/2 and 4I9/2  4S3/2 transitions. The effect of Nd–O covalency on the Ω4 and Ω6 intensity parameters as well as on the spontaneous emission probabilities (AR) was discussed. Lomheim and Shazer hybrid method was applied to determine the fluorescence branching ratios (βR) of each emission transition from the 4F3/2 metastable level to its lower lying levels. The evaluated total radiative transition probabilities (AT), stimulated emission cross-sections (σe) and gain bandwidth parameters (σe × ΔλP) were compared with the earlier reports.  相似文献   

18.
We have conducted a systematic 57Fe Mössbauer study on BaR(Cu0.5Fe0.5)2O5+δ double perovskites with various oxygen contents and rare-earth elements (R=Lu, Yb, Y, Eu, Sm, Nd, and Pr). In samples based on R=Lu, Yb, Y, Eu, Sm the oxygen content remained at δ≈0, upon reductive or oxidative heat treatments under normal pressure. The larger rare-earth elements, i.e. Nd or Pr, readily allowed for continuous oxygen content tuning up to δ≈0.3. By employing high-pressure heat treatments higher oxygen contents were achieved for all samples. The Néel temperature of the samples was found to decrease with increasing amounts of oxygen entering the lattice. In high-pressure oxygenated samples the decrease was less severe indicating that despite the incorporation of oxygen a large amount of Fe still remains in the high-spin trivalent state. By using charge-neutrality arguments together with the relative intensities of the Mössbauer spectral components the average valences of Fe and Cu were obtained. Oxygenation under normal pressure led to a corresponding increase of the valence of Fe, while Cu remained divalent. Upon high-pressure heat treatment equal amounts of Fe3+ and Cu2+ were found to be oxidized to Fe5+ and Cu3+, respectively.  相似文献   

19.
20.
A comparative study on the oxidation and charge compensation in the AxCoO2−δ systems, A=Na (x=0.75, 0.47, 0.36, 0.12) and Li (x=1, 0.49, 0.05), using X-ray absorption spectroscopy at O 1s and Co 2p edges is reported. Both the O 1s and Co 2p XANES results show that upon removal of alkali metal from AxCoO2−δ the valence of cobalt increases more in LixCoO2−δ than in NaxCoO2−δ. In addition, the data of O 1s XANES indicate that charge compensation by oxygen is more pronounced in NaxCoO2−δ than in LixCoO2−δ.  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