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 150 毫秒
1.
Selected joaquinite minerals have been studied by Raman spectroscopy. The minerals are categorised into two groups depending upon whether bands occur in the 3250 to 3450 cm−1 region and in the 3450 to 3600 cm−1 region, or in the latter region only. The first set of bands is attributed to water stretching vibrations and the second set to OH stretching bands. In the literature, X‐ray diffraction could not identify the presence of OH units in the structure of joaquinite. Raman spectroscopy proves that the joaquinite mineral group contains OH units in their structure, and in some cases both water and OH units. A series of bands at 1123, 1062, 1031, 971, 912 and 892 cm−1 are assigned to SiO stretching vibrations. Bands above 1000 cm−1 are attributable to the νas modes of the (SiO4)4− and (Si2O7)6− units. Bands that are observed at 738, around 700, 682 and around 668, 621 and 602 cm−1 are attributed to O Si O bending modes. The patterns do not appear to match the published infrared spectral patterns of either (SiO4)4− or (Si2O7)6− units. The reason is attributed to the actual formulation of the joaquinite mineral, in which significant amounts of Ti or Nb and Fe are found. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

2.
Raman spectroscopy complemented by infrared spectroscopy was used to characterise both gallium oxyhydroxide (α‐GaO(OH)) and gallium oxide (β‐Ga2O3) nanorods synthesised with and without the surfactants using a soft chemical methodology at low temperatures. Nano‐ to micro‐sized gallium oxyhydroxide and gallium oxide materials were characterised and analysed by both X‐ray diffraction and Raman spectroscopy. Rod‐like GaO(OH) crystals with average length of ∼2.5 µm and width of 1.5 µm were obtained. Upon thermally treating gallium oxyhydroxide GaO(OH) to 900 °C, β‐Ga2O3 was synthesised retaining the initial GaO(OH) morphology. Raman spectroscopy has been used to study the structure of nanorods of GaO(OH) and Ga2O3 crystals. Raman spectroscopy shows bands characteristic of GaO(OH) at 950 and ∼1000 cm−1 attributed to Ga OH deformation modes. Bands at 261, 275, 433 and 522 cm−1 are assigned to vibrational modes involving Ga OH units. Bands observed at 320, 346, 418 and 472 cm−1 are assigned to the deformation modes of Ga2O6 octahedra. Two sharp infrared bands at 2948 and 2916 cm−1 are attributed to the GaO(OH) symmetric stretching vibrations. Raman spectroscopy of Ga2O3 provides bands at 630, 656 and 767 cm−1 which are assigned to the bending and stretching of GaO4 units. Raman bands at 417 and 475 cm−1 are attributed to the symmetric stretching modes of GaO2 units. The Raman bands at 319 and 347 cm−1 are assigned to the bending modes of GaO2 units. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

3.
We measured the Raman spectra of ZnO nanoparticles (ZnO‐NPs), as well as transition‐metal‐doped (5% Mn(II), Fe(II) or Co(II)) ZnO nanoparticles, with an average size of 9 nm. A typical Raman peak at 436 cm−1 is observed in the ZnO‐NPs, whereas Zn1−xMnxO, Zn1−xFexO and Zn1−xCoxO presented characteristic peaks at 661, 665 and 675 cm−1, respectively. These peaks can be related to the formation of Mn3O4, Fe3O4 and Co3O4 species in the doped ZnO‐NPs. Moreover, these samples were analyzed at various laser powers. Here, we observed new vibrational modes (512, 571 and 528 cm−1), which are specific to Mn, Fe and Co dopants, respectively, and ZnO‐NPs did not reveal any additional modes. The new peaks were interpreted either as disorder activated phonon modes or as local vibrations of Mn‐, Fe‐ and Co‐related complexes in ZnO. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

4.
To characterize the local relaxation in the structure of lanthanum silicate oxyapatite materials, six compositions with different cation and oxygen stoichiometries (La8Ba2Si6O26, La9BaSi6O26.5, La10Si5.5Mg0.5O26.5, La9.33SiO26, La9.67SiO26.5 and La9.83Si5.5Al0.5O26.5) were investigated by combining Raman scattering and 29Si and 27Al magic‐angle spinning nuclear magnetic resonance (MAS‐NMR) spectroscopies. Only [SiO4]4− species were evidenced and the hypotheses of [Si2O7]6− and [Si2O9]8− entities were ruled out. Both oxygen excess and cation vacancies induce local distortions in the structure, which leads to nonequivalent [SiO4]4− species, characterized by different 29Si MAS‐NMR signals and by splitting of Raman signals. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

5.
The maximum value of hysteresis loss EhMAX due to the itinerant-electron metamagnetic (IEM) transition of La(FexSi1−x)13 and the partially substituted compounds La1−zCez(Fe0.86Si0.14)13 and La1−zPrz(Fe0.86Si0.14)13 increases when the magnetocaloric effects (MCEs) become large. It should be noted that the reduction of EhMAX without the decrease of large MCEs is achieved in La1−zCez(Fe0.86Si0.14)13 and La1−zPrz(Fe0.86Si0.14)13. For both the compound systems mentioned above, the critical temperature T0 for the IEM transition decreases and the difference between T0 and the Curie temperature TC becomes larger with decreasing TC. These results are consistent with the magnetic phase diagram of La(Fe0.86Si0.14)13 under hydrostatic pressure. Consequently, the reduction of EhMAX in La1−zCez(Fe0.86Si0.14)13 and La1−zPrz(Fe0.86Si0.14)13 is closely related with the magnetovolume effects.  相似文献   

6.
Raman spectroscopy has been used to study zemannite Mg0.5[Zn2+Fe3+(TeO3)3]4.5H2O and emmonsite Fe23+Te34+O9·2H2O. Raman bands for zemannite and emmonsite, observed at 740 and 650 cm−1 and at 764 and 788 cm−1, respectively, are attributed to the ν1 (TeO3)2− symmetric stretching mode. The splitting of the symmetric stretching mode for emmonsite is in harmony with the results of X‐ray crystallography which shows three non‐equivalent TeO3 units in the crystal structure. Two bands at 658 and 688 cm−1 are assigned to ν3 (TeO3)2− anti‐symmetric stretching modes. Raman bands observed at 372 and 408 cm−1 for zemannite and 397 and 414 cm−1 for emmonsite are attributed to the (TeO3)2−ν2(A1) bending mode. The two Raman bands at 400 and 440 cm−1 for emmonsite are ascribed to the ν4(E) bending modes, while the band at 326 cm−1 is due to the ν2(A1) bending vibration. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

7.
《X射线光谱测定》2004,33(5):321-325
We developed an EPMA mapping method for small AlaFebSic particles in 1050‐H18 aluminum sheet, which is one of the base materials coated by photoresist in advance called PS plate (pre‐sensitized printing plate). In this method, we used the ratios of relative x‐ray intensities, IFe/IAl and IFe/ISi instead of the mass ratios, Fe/Al and Fe/Si, of the main elements which constitute the particles and tried to determine the ratios of relative x‐ray intensities using Monte Carlo calculations. Furthermore, using this developed mapping method, we performed the mapping of small AlaFebSic particles such as Al3Fe (0–3%Si as impurities), Al6Fe (0–1%Si as impurities), α‐AlFeSi(Al8.3Fe2Si) and β‐AlFeSi(Al8.9Fe2Si2) in 1050‐H18 aluminum sheets. We found that the discrimination of these particles was achieved with this mapping method. We confirmed that this method is useful for the mapping of AlaFebSic particles in 1050‐H18 aluminum sheets. Copyright © 2004 John Wiley & Sons, Ltd.  相似文献   

8.
The S3 radical anion is observed in several systems: non‐aqueous polysulfides solutions, doped alkali halides, ultramarine pigments (UP) for which S3 is the blue chromophore and S2 is the yellow one and pigments of zeolite 4A structure. The S3 ion has C2V symmetry, and therefore its three vibrational modes should be observed in the Raman and in IR spectra. In resonance Raman spectroscopy, only the symmetric stretching mode ν1 and the bending mode ν2 have been observed, whereas the anti‐symmetric stretching mode ν3 has never been observed whatever the system. In this work, we confirm that ν3 is not observed in solutions with resonance Raman spectroscopy. However, our investigation of several blue UP, with various concentrations of S2, shows that there is a superposition of two bands at ca 590 cm−1: the first is assigned to ν (S2) and the second to ν3 (S3). With the 457.9 nm excitation line, for which the resonance conditions are simultaneously fulfilled for S2 and S3, the band at ca 590 cm−1 is the sum of the contributions of both ν (S2) and ν3 (S3) vibrations, while, with the 647.1 nm line, which only satisfies the resonance conditions of S3, the band at ca 584 cm−1 must be assigned only to ν3 (S3). Furthermore, ν3 (S3) is observed in green UP and in pigments of zeolite structure. The ν3 vibration of S3, which is observed neither in polysulfide solutions nor in doped alkali halides in resonance Raman conditions, can therefore be observed when this species is inserted into the β‐cages of the sodalite or of the zeolite 4A structures. So, the band at ca 590 cm−1 cannot always be assigned to S2 in these systems. This implies that the concentration of S2 in UP must be reconsidered. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

9.
Raman spectroscopy was used to study the molecular structure of a series of selected rare earth (RE) silicate crystals including Y2SiO5 (YSO), Lu2SiO5 (LSO), (Lu0.5Y0.5)2SiO5 (LYSO) and their ytterbium‐doped samples. Raman spectra show resolved bands below 500 cm−1 region assigned to the modes of SiO4 and oxygen vibrations. Multiple bands indicate the nonequivalence of the RE O bonds and the lifting of the degeneracy of the RE ion vibration. Low intensity bands below 500 cm−1 are an indication of impurities. The (SiO4)4− tetrahedra are characterized by bands near 200 cm−1 which show a separation of the components of ν4 and ν2, in the 500–700 cm−1 region which are attributed to the distorting bending vibration and in the 880–1000 cm−1 region which are attributed to the symmetric and antisymmetric stretching vibrational modes. The majority of the bands in the 300–610 cm−1 region of Re2SiO5 were found to arise from vibrations involving both Si and RE ions, indicating that there is considerable mixing of Si displacements with Si O bending modes and RE O stretching modes. The Raman spectra of RE silicate crystals were analyzed in terms of the molecular structure of the crystals, which enabled separation of the bands attributed to distinct vibrational units. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

10.
K.C. Chung  F.L. Kwong  Jia Li 《哲学杂志》2013,93(19):1535-1553
The reaction mechanisms between Al and Fe3O4 powders were investigated. Differential thermal analysis revealed that a two-step displacement reaction between Al and Fe3O4 took place during sintering. Initially, the Fe3O4 was converted to amorphous FeO at ~720°C and some of the Al was oxidized to amorphous Al2O3. In the final stage, when the temperature reached ~840°C, crystalline Al2O3 particles were produced in the molten Al–Fe liquid. The effects of cooling rate on the microstructures were studied. When the Al–Fe liquid was furnace-cooled to room temperature, proeutectic Al3Fe plates, plate-like divorced eutectic Al3Fe and Al2O3 particles were in situ formed in the Al(Fe) matrix. While quenching from 700°C, nanometer-sized Al dendrites and Al–Al6Fe eutectic lamellae were produced in the Al matrix. However, when it was rapidly quenched from 900°C, the size of the proeutectic Al3Fe phases was further reduced and Al6Fe nanorods were found in the Al–Al6Fe eutectics. A model was proposed to describe the transformation of the Al–Fe intermetallics during solidification.  相似文献   

11.
The mineral dussertite, a hydroxy‐arsenate mineral with formula BaFe3+3(AsO4)2(OH)5, has been studied by Raman spectroscopy complemented with infrared spectroscopy. The spectra of three minerals from different origins were investigated and proved to be quite similar, although some minor differences were observed. In the Raman spectra of the Czech dussertite, four bands are observed in the 800–950 cm−1 region. The bands are assigned as follows: the band at 902 cm−1 is assigned to the (AsO4)3−ν3 antisymmetric stretching mode, the one at 870 cm−1 to the (AsO4)3−ν1 symmetric stretching mode, and those at 859 and 825 cm−1 to the As‐OM2 + /3+ stretching modes and/or hydroxyl bending modes. Raman bands at 372 and 409 cm−1 are attributed to the ν2 (AsO4)3− bending mode and the two bands at 429 and 474 cm−1 are assigned to the ν4 (AsO4)3− bending mode. An intense band at 3446 cm−1 in the infrared spectrum and a complex set of bands centred upon 3453 cm−1 in the Raman spectrum are attributed to the stretching vibrations of the hydrogen‐bonded (OH) units and/or water units in the mineral structure. The broad infrared band at 3223 cm−1 is assigned to the vibrations of hydrogen‐bonded water molecules. Raman spectroscopy identified Raman bands attributable to (AsO4)3− and (AsO3OH)2− units. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

12.
The oxygen partial pressure (P(O2)) dependence of in situ X-ray absorption spectroscopy (XAS) at the Co and Fe K edges was measured simultaneously and continuously at 900 and 1000 K. These experiments, which were performed during reduction, changing P(O2) from 1 to 10?4 atm, were used to investigate each valence related to Co and Fe in (La0.6Sr0.4)(Co0.2Fe0.8)O3?δ (LSCF). The absorption edge shift of the Co K edge was more than twice that of the Fe K edge at 1000 K during reduction. For quantitative analysis, X-ray absorption near-edge structure spectroscopy was carried out at the Co and Fe K edges; the results indicated that the Co valence decreased more easily than the Fe valence; that is, the oxygen preferentially left from the oxygen sites around Co.  相似文献   

13.
The growth and characterization of high‐quality ultrathin Fe3O4 films on semiconductor substrates is a key step for spintronic devices. A stable, single‐crystalline ultrathin Fe3O4 film on GaAs(001) substrate is obtained by post‐growth annealing of epitaxial Fe film with thicknesses of 5 and 12 nm in air. Raman spectroscopy shows a high ability to convincingly characterize the stoichiometry, epitaxial orientation and strain of such ultrathin Fe3O4 films. Polarized Raman spectroscopy confirms the unit cell of Fe3O4 films is rotated by 45° to match that of the Fe (001) layer on GaAs, which results in a built‐in strain of − 3.5% in Fe3O4 films. The phonon strain‐shift coefficient(−126 cm−1) of the A1g mode is proposed to probe strain effect and strain relaxation of thin Fe3O4 films on substrates. It can be used to identify whether the Fe layer is fully oxidized to Fe3O4 or not. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

14.
A new member of the family of garnets with fast lithium ion conduction has been found with the composition Li7La3Hf2O12. The anion arrangement corresponds to the oxygen framework in garnets, e.g., in Ca3Fe2Si3O12. Hafnium is coordinated octahedrally while the lanthanum environment can be described as a distorted cube. Lithium occupies a large number of positions with tetrahedral, trigonal planar, and metaprismatic coordination. Li7La3Hf2O12 shows a lithium bulk ion conductivity of 2.4 × 10−4 Ω−1 cm−1 at room temperature with an activation energy of 0.29 eV.  相似文献   

15.
The molecular structure of the mineral pecoraite, the nickel analogue of chrysotile of formula Ni3Si2O5(OH)4, was analysed by a combination of Raman and infrared spectroscopies. A comparison is made with the spectra of the minerals nepouite and chrysotile and a synthetic pecoraite. Pecoraite is characterised by OH stretching vibrations at 3645 and 3683 cm−1 attributed to the inner and inner surface hydroxyl stretching vibrations. Intense infrared bands at around 3288 and 3425 cm−1 are assigned to the stretching vibrations of water strongly hydrogen‐bonded to the surface of the pecoraite. The asbestos‐like mineral is characterised by SiO stretching vibrations at 979, 1075, 1128 and 1384 cm−1, OSiO chain vibrations at 616 and 761 cm−1 and Ni O(H) vibrations at 397 and 451 cm−1. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

16.
The Raman spectrum of the symmetric stretching vibration (ν1) of liquid carbon tetrachloride observed at 295 K and reported repeatedly over the last 80 years clearly shows four of the five more abundant isotopomers at 440–470 cm−1. At the lower energy end of this spectrum, additional intensity due to isotopomeric contributions from the symmetric stretch for v = 1 → 2 (hotbands) partially overlaps the prominent v = 0 → 1 features, and accounts for about 18% of the integrated intensity at 295 K in agreement with theory. When these two patterns are modeled and subtracted from the experimental spectrum, a feature underlying almost exactly the C35Cl4 (v = 0 → 1) band at 462.5 cm−1 becomes apparent. We propose that this feature is the ν3 − ν4 difference band. Observations at lower temperatures, and of the combination bands, and the polarized Raman spectra are consistent with this hypothesis. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

17.
Y2Si2O7 is an intriguing material combining a complex structural polymorphism with several important technological applications. Raman spectra were experimentally determined for most of the seven known modifications of Y2Si2O7 except the form ε, and in the case of β, γ, δ and ζ for the first time. The error‐prone procedure of mode assignment to the measured Raman bands, usually done by comparison with similar or related structures, has been replaced by quantum chemical calculations of the spectra of the polymorphs. Various functionals were evaluated considering the agreement of the calculated modes with the experimental data. The average and maximum deviations between calculated and experimental spectra are ± 8 cm−1 and 20 cm−1, respectively. Assignments of most of the observed bands to vibrational modes are given. The relationship between selected Raman bands, Si O and Y O polyhedra stretching and bending modes, and the crystal structures are discussed. Y2Si2O7 offers the possibility to study the relationship between structural and spectral changes in a chemically fixed system. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

18.
A series of iron- and/or aluminium-doped apatite-type lanthanum silicates (ATLS) La9.83Si6 ‐ x ‐ yAlxFeyO26 ± δ (x = 0, 0.25, 0.75, and 1.5, y = 0, 0.25, 0.75, and 1.5) were synthesized using the mechanochemical activation (MA), solid state reaction (SSR), Pechini (Pe) and sol-gel (SG) methods. The total conductivity of the prepared materials was measured under air in the temperature range 600-850 °C using 4-probe AC impedance spectroscopy. Its dependence on composition, synthesis method, sintering conditions and powder particle size was investigated. It was found that for electrolytes of the same composition, those prepared via mechanochemical activation exhibited the highest total specific conductivity, which was improved with increasing Al- and decreasing Fe-content. The highest conductivity value at 700 °C, equal to 2.04 × 10− 2 S cm− 1, was observed for the La9.83Si5Al0.75Fe0.25O26 ± δ electrolyte. La9.83Si4.5Fe1.5O26 ± δ electrolyte samples synthesized using the Pechini method exhibited higher conductivity when sintered conventionally than when spark-plasma sintering (SPS) was used.  相似文献   

19.
Tellurites may be subdivided according to formula and structure. There are five groups based upon the formulae (a) A(XO3), (b) A(XO3)·xH2O, (c) A2(XO3)3·xH2O, (d) A2(X2O5) and (e) A(X3O8). Raman spectroscopy has been used to study the tellurite minerals teineite and graemite; both contain water as an essential element of their stability. The tellurite ion should show a maximum of six bands. The free tellurite ion will have C3v symmetry and four modes, 2A1 and 2 E. Raman bands for teineite at 739 and 778 cm−1 and for graemite at 768 and 793 cm−1 are assigned to the ν1 (TeO3)2− symmetric stretching mode while bands at 667 and 701 cm−1 for teineite and 676 and 708 cm−1 for graemite are attributed to the ν3 (TeO3)2− antisymmetric stretching mode. The intense Raman band at 509 cm−1 for both teineite and graemite is assigned to the water librational mode. Raman bands for teineite at 318 and 347 cm−1 are assigned to the (TeO3)2−ν2(A1) bending mode and the two bands for teineite at 384 and 458 cm−1 may be assigned to the (TeO3)2−ν4(E) bending mode. Prominent Raman bands, observed at 2286, 2854, 3040 and 3495 cm−1, are attributed to OH stretching vibrations. The values for these OH stretching vibrations provide hydrogen bond distances of 2.550(6) Å (2341 cm−1), 2.610(3) Å (2796 cm−1) and 2.623(2) Å (2870 cm−1) which are comparatively short for secondary minerals. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

20.
The kaolinite‐like phyllosilicate minerals bismutoferrite BiFe3+2Si2O8(OH) and chapmanite SbFe3+2Si2O8(OH) have been studied by Raman spectroscopy and complemented with infrared spectra. Tentatively interpreted spectra were related to their molecular structure. The antisymmetric and symmetric stretching vibrations of the Si O Si bridges, δ SiOSi and δ OSiO bending vibrations, ν (Si Oterminal) stretching vibrations, ν OH stretching vibrations of hydroxyl ions, and δ OH bending vibrations were attributed to the observed bands. Infrared bands in the range 3289–3470 cm−1 and Raman bands in the range 1590–1667 cm−1 were assigned to adsorbed water. O H···O hydrogen‐bond lengths were calculated from the Raman and infrared spectra. Copyright © 2009 John Wiley & Sons, Ltd.  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