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.
The infrared and Raman spectra of barium oxalate hemihydrate, BaC2O4 · 0.5H2O, were recorded and discussed on the basis of their structural peculiarities and in comparison with the spectra of the previously investigated calcium and strontium oxalates.  相似文献   

2.
Abstract

We report high-pressure Raman scattering spectroscopy and energy dispersive X-ray diffraction investigations on gypsum, CaSO4 · 2H2O, at room temperature in a diamond cell. With increasing pressure, measurements indicate that CaSO4 · 2H2O undergoes two stages of crystalline-state phase transitions at 5 and 9 GPa, and then converts to a disordered phase above 11 GPa. The structures of the three high-pressure phases of gypsum have not been determined yet. These phases are tentatively named as “post-gypsum-I” (PG-I), “post-gypsum-II” (PG-II) and “disordered” according to the sequence of their appearance with pressure.

Gypsum shows anisotropic compressibility along three crystallographic axes with b > c > a below 5 GPa. The difference in the behavior of the two OH stretching modes in gypsum is attributed to the different reduction rate in the hydrogen bonding distances by the anisotropic axial compressibility.  相似文献   

3.
Nickel hydroxides with hierarchical micro-nano structures are prepared by a facile homogeneous precipitation method with different nickel salts (Ni(NO3)2·6H2O, NiCl2·6H2O, and NiSO4·6H2O) as raw materials. The effect of nickel sources on the microstructure and lithium storage performance of the nickel hydroxides is studied. It is found that all the three prepared samples are α-nickel hydroxide. The nickel hydroxides synthesized with Ni(NO3)2·6H2 or NiCl2·6H2O show a similar particle size of 20–30 μm and are composed of very thin nano-sheets, while the nickel hydroxide synthesized with Ni(SO4)2·6H2O shows a larger particle size (30–50 μm) and consists of very thin nano-walls. When applied as anode materials for lithium-ion batteries (LIBs), the nickel hydroxide synthesized with NiSO4·6H2O exhibits the highest discharge capacity, but its cyclic stability is very poor. The nickel hydroxides synthesized with NiCl2·6H2O exhibit higher discharge capacity than the nickel hydroxides synthesized with Ni(NO3)2·6H2O, and both of them show much improved cyclic stability and rate capability as compared to the nickel hydroxide synthesized with Ni(SO4)2·6H2O. Moreover, pseudocapacitive behavior makes a great contribution to the electrochemical energy storage of the three samples. The discrepancies of lithium storage performance of the three samples are analyzed by ex-situ XRD, FT-IR, electrochemical impedance spectroscopy (EIS), and cyclic voltammetry (CV) tests.  相似文献   

4.
5.
The mineral ardealite Ca2(HPO4)(SO4)·4H2O is a ‘cave’ mineral and is formed through the reaction of calcite with bat guano. The mineral shows disorder and the composition varies depending on the origin of the mineral. Raman spectroscopy complimented with infrared spectroscopy has been used to characterise the mineral ardealite. The Raman spectrum is very different from that of gypsum. Bands are assigned to SO42− and HPO42− stretching and bending modes. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

6.
The thermal dehydration of ZnK4(P3O9)2 · 6H2O was studied in the range 25–500°C by thermogravimetric analysis (TGA and DSC) and X‐ray diffraction. We found, based on the TGA and DSC scans, the dehydration of this salt takes place in three stages with a loss of the six water molecules. The infrared and Raman spectra of ZnK4(P3O9)2 · 6H2O have been recorded and interpreted using a factor group analysis. The internal modes are assigned in terms of POP and PO2 structural units using experimental and theoretical IR and Raman frequencies.  相似文献   

7.
Gilalite is a copper silicate mineral with a general formula of Cu5Si6O17 · 7H2O. The mineral is often found in association with another copper silicate mineral, apachite, Cu9Si10O29 · 11H2O. Raman and infrared spectroscopy have been used to characterize the molecular structure of gilalite. The structure of the mineral shows disorder, which is reflected in the difficulty of obtaining quality Raman spectra. Raman spectroscopy clearly shows the absence of OH units in the gilalite structure. Intense Raman bands are observed at 1066, 1083, and 1160 cm?1.

The Raman band at 853 cm?1 is assigned to the –SiO3 symmetrical stretching vibration and the low-intensity Raman bands at 914, 953, and 964 cm?1 may be ascribed to the antisymmetric SiO stretching vibrations. An intense Raman band at 673 cm?1 with a shoulder at 663 cm?1 is assigned to the ν4 Si-O-Si bending modes. Raman spectroscopy complemented with infrared spectroscopy enabled a better understanding of the molecular structure of gilalite.  相似文献   

8.
The metastable phase of well-faceted, hexagonal, prism-like molybdenum oxide hydrate (MoO3·0.55H2O) was successfully synthesized by evaporating molybdic acid solution prepared through cation membrane electrolysis of Na2MoO4·2H2O aqueous solution. The obtained crystals were characterized by X-ray diffraction (XRD), thermogravimetric (TG), scanning electronic microscopy (SEM) and photoluminescence (PL) spectrophotometry. The as-prepared MoO3·0.55H2O rods were of 2–4 μm in width and 5–12 μm in length. The MoO3·0.55H2O microrods displayed photoluminescence properties at room temperature and were transformed into stable orthorhombic α-MoO3 after air annealing at 380 °C. Moreover, the influence of temperature factor on the phase transformation process, morphology and photoluminescence properties of MoO3·0.55H2O was investigated in detail.  相似文献   

9.
The minerals of the mixite group—zálesíite CaCu6[(AsO4)2(AsO3OH)(OH)6]·3H2O from abandoned uranium deposit Zálesí, Czech Republic and calciopetersite CaCu6[(PO4)2(PO3OH)(OH)6]·3H2O from a quarry near Domašov na Bystřicí, northern Moravia, Czech Republic—were studied by Raman and infrared spectroscopy. The observed bands were assigned to the stretching and bending vibrations of (AsO4)3− and (AsO3OH)2− ions in zálesíite, and (PO4)3− and (PO3OH)2− in calciopetersite, and to molecular water, hydroxyl ions, and Cu‐(O,OH) units in both minerals. O H···O hydrogen‐bond lengths in zálesíite and calciopetersite structures were calculated with Libowitzky's empirical relation. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

10.
In this report, we extended the works of Rizzato et al. [Angew. Chem. Int. Ed. 49, 7440 (2010)] on the nature of O–H···Pt hydrogen bond in trans-[PtCl2(NH3)(N–glycine)]·H2O(1·H2O) complex, by computational study of O–H···Pt interaction in [NBu4][Pt(C6F5)3(8-hydroxyquinaldine)], with emphasis on charge transfer effect in this interaction of platinum(II) and hydrogen atom. According to the crystallographic geometry reported by José María Casas et al., [NBu4][Pt(C6F5)3(8-hydroxyquinaldine)] possesses one O–H···Pt hydrogen bridging interaction, similar to the case in trans-[PtCl2(NH3)(N–glycine)]·H2O(1·H2O) complex. On the basis of topological criteria of electron density, we characterised this O–H···Pt interaction. Charge transferred between platinum(II) and σ*O–H orbital in this complex was calculated by using NBO method. The stabilised energy associated to charge transfer was estimated using a direct proportionality, that is 2–3 eV per electron transferred. Charge transfer effects in O–H···Pt hydrogen bonds were studied for these two complexes. Our results indicate that the interaction of O–H···Pt is closed–shell in nature with significant charge transfer, and that charge transfer effect is not negligible in the interaction of O–H···Pt. The second conclusion is different from the result of Rizzato et al.  相似文献   

11.
A comparative analysis has been carried out on the Raman spectra of FeSO4·nH2O (n = 1, 4, 7) including the 2D‐analogs. The effects of changing the degrees of hydration have been found from the lattice, SO42− internal, and H2O internal modes. Increasing degrees of hydration shift the intense ν1(SO4) peak to lower wavenumbers and reduce the amount of splitting on the ν3(SO4) peaks. Some of the water librational bands cause the broadening of the ν4(SO4) peaks in FeSO4·7H2O and the ν2(SO4) peaks in FeSO4·7D2O. The ν2(H2O) band in FeSO4·H2O is red‐shifted in excess of 100 cm−1 relative to the unperturbed H2O band. Between 240 and 190 K and between 140 and 90 K in the spectra of FeSO4.4H2O, two potential phase transitions have been identified from the changes in the lattice and water‐stretching regions. The resolution of the ν1(H2O) and ν3(H2O) bands in FeSO4·4H2O and FeSO4·H2O also improved sharply at low temperatures. The capability of distinguishing various forms of FeSO4 hydrates unambiguously makes the Raman technique a potential analytical tool for the identification of sulfate minerals on planetary surfaces. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

12.
Calcium phosphate glasses, in which part of calcium oxide was replaced by lanthanum oxide, were prepared by using the conventional melt quench method. The structures of xLa2O3 · (50-x)CaO · 50P2O5 (x = 0, 1, 3, 6, 12 mol%) samples were investigated by X-ray diffraction (XRD), Raman spectrum, Fourier transform infrared spectrum (FTIR), and differential scanning calorimetry (DSC). The results show that lanthanum oxide acts as network modifier in the network space of glass structure. The glass formation occurs at an O/P ratio of about 3.0–3.12. At larger values, the crystalline phases calcium pyrophosphate (Ca2P2O7) and calcium lanthanum phosphate [Ca9La(PO4)7] are detected in the samples. Raman and FTIR spectra indicate that the structure of lanthanum-free sample is chain P–O–P bond metaphosphate–based Q2 units. Glass structure will change to Q2 and Q1 units when the lanthanum oxide content is less then 6 mol%. When lanthanum oxide content increases to 9 and 12 mol% more nonbridging oxygen in the glass, resulting in the depolymerization of the phosphate network, the network of glass transforms to Q2, Q1, and Q0 units mixture. Based upon DSC results, Tg slightly decreases because of the depolymerization of microstructure. Endothermal peak of DSC curves indicate that crystal phases separate out from vitreous body with the addition of lanthanum oxide content.  相似文献   

13.
In this paper seventeenth century gilded paintings from the crypt of Sant Benet de Bages, a medieval monastery in the Catalonia region of Spain, near Barcelona, have been studied. Cross sections from two different gilded decorations were studied by means of optical microscopy and electron microscopy and EDS to determine the statigraphy and elemental composition, and by means of FTIR coupled to a microscope to determine the binding media associated to each layer. These preliminary results demonstrated that gilded decorations were made by the application of a gold foil on a mordant substrate on a gypsum base, while the mouldings of the vaults seem to be gilded on a bol with a glaze on top of the gold leaf. It is interesting to notice that the first remained unaltered, while the gilded vault mouldings look almost black, due to the darkening of the organic material. To elucidate the causes involved in the darkening of the sample from the vaults a set of synchrotron μXRD and μFTIR experiments have been carried out on these samples at the ESRF (ID18F and ID21, respectively). High brightness and small spot working conditions revealed the development and distribution of calcium oxalates in the binding media, which seem to be responsible for the darkening. Results point out the fact that weddellite (CaC2O4·2H2O) is the phase formed in those layers where organic material has also been identified or at least it would be supposed to be by bibliographic sources and not necessarily those superficial as it would have been suggested due to the similarities with patinas formation. PACS 78.30-j; 68.37.Hk; 61.10.Nz  相似文献   

14.
Tunable diode laser absorption spectroscopy sensors for detection of CO, CO2, CH4 and H2O at elevated pressures in mixtures of synthesis gas (syngas: products of coal and/or biomass gasification) were developed and tested. Wavelength modulation spectroscopy (WMS) with 1f-normalized 2f detection was employed. Fiber-coupled DFB diode lasers operating at 2325, 2017, 2290 and 1352 nm were used for simultaneously measuring CO, CO2, CH4 and H2O, respectively. Criteria for the selection of transitions were developed, and transitions were selected to optimize the signal and minimize interference from other species. For quantitative WMS measurements, the collision-broadening coefficients of the selected transitions were determined for collisions with possible syngas components, namely CO, CO2, CH4, H2O, N2 and H2. Sample measurements were performed for each species in gas cells at a temperature of 25 °C up to pressures of 20 atm. To validate the sensor performance, the composition of synthetic syngas was determined by the absorption sensor and compared with the known values. A method of estimating the lower heating value and Wobbe index of the syngas mixture from these measurements was also demonstrated.  相似文献   

15.
The measurements of the rotational spectrum of 1,4-dioxane-water complex, performed by pulsed jet Fourier transform microwave spectroscopy, have been extended to the 7–18.5 GHz frequency region and to two additional isotopologues, 1,4-dioxane···H2 17O and 1,4-dioxane···HOD. The effective orientation of water in the complex has been obtained from the 17O quadrupole coupling constants.  相似文献   

16.
ABSTRACT

Apachite, Cu9Si10O29 · 11H2O, is a mineral named after the American Indian Apache tribe. Raman and infrared spectroscopy have been used to characterize the molecular structure of apachite. The structure of the mineral shows disorder, which is reflected in the difficulty in obtaining quality Raman spectra. Raman spectroscopy clearly shows the presence of OH units in the apachite structure, which attests the formula to be not correct. Both Raman and infrared spectroscopy show the presence of water in the apachite structure. Different water molecules are present with different hydrogen bonding strengths. A suggested formula might be Cu9Si10O23(OH)12 · 5H2O.

The Raman band at 967 cm?1 is assigned to the –SiO3 symmetrical stretching vibration and the bands at 997 and 1096 cm?1 are assigned to the ν3 –SiO3 antisymmetric stretching vibrations. An intense Raman band at 673 cm?1 with a shoulder at 663 cm?1 is assigned to the ν4 Si-O-Si bending modes. Raman spectroscopy complemented with infrared spectroscopy enabled a better understanding of the molecular structure of apachite.  相似文献   

17.
18.
Specific features of complexation in solutions of a strong dibasic acid in the H2SO4–2-pyrrolidone (Pyr) system (in the range of compositions of 0–100% H2SO4) are studied using multiple frustrated total internal reflection IR spectroscopy. The conclusions drawn on the structure of the complexes formed in such solutions are confirmed by quantum-chemical calculations of the mPyr · nH2SO4 (m, n = 1, 2) heteroassociates and by comparison of their calculated and measured vibrational spectra. It is found that, in the investigated solutions, four types of acid–base complexes, with various degrees of proton transfer in the OHO bridge, are formed: (AHA) anions with quasi-symmetric H-bonds, solvated by acid molecules, or entering into the composition of PyrH+ · (AHA) ion pairs; quasi-ion pairs with incomplete proton transfer to the base molecule of 1: 1 and 2: 2 compositions; and 2Pyr · H2SO4 complexes with two O–H···O bridges of molecular type. The main differences in the mechanisms of the acid–base interactions in the H2SO4–Pyr system as compared to the CH3SO3H–Pyr system result from the participation of two OH-groups of H2SO4 molecule in these interactions. Therefore, two types of quasi-ion pairs and complexes of 2Pyr · H2SO4 composition are formed.  相似文献   

19.
The addition reaction of CH2OO + H2O CH2(OH)OOH without and with X (X = H2CO3, CH3COOH and HCOOH) and H2O was studied at CCSD(T)/6-311+ G(3df,2dp)//B3LYP/6-311+G(2d,2p) level of theory. Our results show that X can catalyse CH2OO + H2O → CH2(OH)OOH reaction both by increasing the number of rings, and by adding the size of the ring in which ring enlargement by COOH moiety of X inserting into CH2OO···H2O is favourable one. Water-assisted CH2OO + H2O → CH2(OH)OOH can occur by H2O moiety of (H2O)2 or the whole (H2O)2 forming cyclic structure with CH2OO, where the latter form is more favourable. Because the concentration of H2CO3 is unknown, the influence of CH3COOH, HCOOH and H2O were calculated within 0–30 km altitude of the Earth's atmosphere. The results calculated within 0–5 km altitude show that H2O and HCOOH have obvious effect on enhancing the rate with the enhancement factors are, respectively, 62.47%–77.26% and 0.04%–1.76%. Within 5–30 km altitude, HCOOH has obvious effect on enhancing the title rate with the enhancement factor of 2.69%–98.28%. However, compared with the reaction of CH2OO + HCOOH, the rate of CH2OO···H2O + HCOOH is much slower.  相似文献   

20.
FePO4·xH2O/graphene oxide (FePO4·xH2O/GO) composites were prepared by a facile chemical precipitation method. Using the as-prepared FePO4·xH2O/GO and LiOH·H2O as precursors and followed by carbothermal reduction, LiFePO4/graphene composites were obtained. Scanning electron microscope (SEM) images indicated that the graphene had very good dispersity and uniformly attached to the LiFePO4 particles. The conductive framework of graphene improved the electrochemical properties of the composites. The composites deliver high initial discharge capacity of 163.4 mAh g?1 as well as outstanding rate performance.  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