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 46 毫秒
1.
The paper presents the study of selected montmorillonite standards by Raman spectroscopy and microscopy supported by elemental analysis, X-ray powder diffraction analysis and thermal analysis. Dispersive Raman spectroscopy with excitation lasers of 532 nm and 780 nm, dispersive Raman microscopy with excitation laser of 532 nm and 100× magnifying lens, and Fourier Transform-Raman spectroscopy with excitation laser of 1064 nm were used for the analysis of four montmorillonites (Kunipia-F, SWy-2, STx-1b and SAz-2). These mineral standards differed mainly in the type of interlayer cation and substitution of octahedral aluminium by magnesium or iron. A comparison of measured Raman spectra of montmorillonite with regard to their level of fluorescence and the presence of characteristic spectral bands was carried out. Almost all measured spectra of montmorillonites were significantly affected by fluorescence and only one sample was influenced by fluorescence slightly or not at all. In the spectra of tested montmorillonites, several characteristic Raman bands were found. The most intensive band at 96 cm−1 belongs to deformation vibrations of interlayer cations. The band at 200 cm−1 corresponds to deformation vibrations of the AlO6 octahedron and at 710 cm−1 can be assigned to deformation vibrations of the SiO4 tetrahedron. The band at 3620 cm−1 corresponds to the stretching vibration of structural OH groups in montmorillonites.  相似文献   

2.
《Vibrational Spectroscopy》2002,28(2):209-221
Syngenite (K2Ca(SO4)2·H2O), formed during treatment of manure with sulphuric acid, was studied by infrared, near-infrared (NIR) and Raman spectroscopy. Cs site symmetry was determined for the two sulphate groups in syngenite (P21/m), so all bands are both infrared and Raman active. The split ν1 (two Raman+two infrared bands) was observed at 981 and 1000 cm−1. The split ν2 (four Raman+four infrared bands) was observed in the Raman spectrum at 424, 441, 471 and 491 cm−1. In the infrared spectrum, only one band was observed at 439 cm−1. From the split ν3 (six Raman+six infrared) bands three 298 K Raman bands were observed at 1117, 1138 and 1166 cm−1. Cooling to 77 K resulted in four bands at 1119, 1136, 1144 and 1167 cm−1. In the infrared spectrum, five bands were observed at 1110, 1125, 1136, 1148 and 1193 cm−1. From the split ν4 (six infrared+six Raman bands) four bands were observed in the infrared spectrum at 604, 617, 644 and 657 cm−1. The 298 K Raman spectrum showed one band at 641 cm−1, while at 77 K four bands were observed at 607, 621, 634 and 643 cm−1. Crystal water is observed in the infrared spectrum by the OH-liberation mode at 754 cm−1, OH-bending mode at 1631 cm−1, OH-stretching modes at 3248 (symmetric) and 3377 cm−1 (antisymmetric) and a combination band at 3510 cm−1 of the H-bonded OH-mode plus the OH-stretching mode. The near-infrared spectrum gave information about the crystal water resulting in overtone and combination bands of OH-liberation, OH-bending and OH-stretching modes.  相似文献   

3.
A three-step infrared (IR) macro-fingerprint method combining conventional IR spectra, and the secondary derivative spectra with two-dimensional infrared correlation spectroscopy (2D-IR), was developed to analyze Spirulina powder before and after gamma irradiation. In the IR spectra, most of the absorption peaks of samples irradiated at 1, 2.7, 6, and 10.4 kGy had lower intensities than the non-irradiated ones, whereas peaks at 1152, 1078, and 1051 cm−1 were slightly enhanced with irradiation at 2.7, 6, and 10.4 kGy. Their second derivative spectra amplified the differences and revealed that irradiation affected the C=O band of carboxylic acid and esters, and the N–H band of proteins. The peaks at 1746 and 1741 cm−1, and those at 1730 and 1725  cm−1 became two broad peaks. Meanwhile, the three sharp peaks at 1548 cm−1, 1544 cm−1 and 1536 cm−1 changed to two broad peaks at around 1547 and 1534 cm−1 after irradiation at doses higher than 1 kGy. The characteristic IR bands from 1700 cm−1 to 1600 cm−1, which represent the C=O band in proteins, also have different shapes and intensities after irradiation. The finding indicated that irradiation affected the secondary structures of protein which was confirmed by curve fitting results. During the process of increasing the temperature from 50 to 210 °C, the ratio of amide I to II in absorption intensities in the 2D-IR spectra of the irradiated samples varied with different response for different samples. Saccharides in Spirulina powder had a higher thermostability than proteins, but the autopeaks of irradiated samples did show differences from the non-irradiated sample. The intensity of autopeaks at 1012 cm−1 increased dramatically in the irradiated samples while that of peaks at 1053, 1071, and 1083 cm−1 decreased after irradiation. Based on the three-step IR macro-fingerprint method, irradiated Spirulina powder samples were successfully and fast identified and discriminated.  相似文献   

4.
We calculated IR, nonresonance Raman spectra and vertical electronic transitions of the zigzag single-walled and double-walled boron nitride nanotubes ((0,n)-SWBNNTs and (0,n)@(0,2n)-DWBNNTs). In the low frequency range below 600 cm−1, the calculated Raman spectra of the nanotubes showed that RBMs (radial breathing modes) are strongly diameter-dependent, and in addition the RBMs of the DWBNNTs are blue-shifted reference to their corresponding one in the Raman spectra of the isolated (0,n)-SWBNNTs. In the high frequency range above ∼1200 cm−1, two proximate Raman features with symmetries of the A1g (∼1355 ± 10 cm−1) and E2g (∼1330 ± 25 cm−1) first increase in frequency then approach a constant value of ∼1365 and ∼1356 cm−1, respectively, with increasing tubes’ diameter, which is in excellent agreement with experimental observations. The calculated IR spectra exhibited IR features in the range of 1200–1550 cm−1 and in mid-frequency region are consistent with experiments. The calculated dipole allowed singlet–singlet and triplet–triplet electronic transitions suggesting a charge transfer process between the outer- and inner-shells of the DWBNNTs as well as, upon irradiation, the possibility of a system that can undergo internal conversion (IC) and intersystem crossing (ISC) processes, besides the photochemical and other photophysical processes.  相似文献   

5.
The infrared spectra of ethylmethylfluorosilane (CH3SiHFCH2CH3) have been recorded as a vapour, liquid and solid at 78 K in the 4000–50 cm−1 range and isolated in an argon matrix at ca. 5 K. Infrared spectra of two different solid phases were obtained after annealing to temperatures of 120 and 130 K, and recooling to 78 K. Although the IR spectra were quite similar in the MIR region, certain differences were noted in the FIR region below 400 cm−1. The most stable conformer MeMe was present after annealing to 130 K, but three bands belonging to MeH were detected after annealing to 120 K. Various infrared bands changed intensity when the argon matrix was annealed to temperatures between 20 and 35 K, and some of these were related to changes in the conformational abundance.Raman spectra of the liquid were recorded at room temperature and at various temperatures between 295 and 153 K. Spectra of an amorphous and annealed solid were recorded at 78 K. In the variable temperature Raman spectra, various bands changed in intensity and were interpreted in terms of conformational equilibria between the three possible conformers. Complete assignments were made for all the bands of the most stable conformer MeMe. From various bands assigned to the three conformers, the conformational enthalpy difference ΔH from MeMe to the intermediate energy conformer MeH was found to be 0.5 kJ mol−1 and to the highest conformer MeF was 0.7 kJ mol−1. At ambient temperature this leads to 39% MeMe, 32% MeH and 29% of the MeF conformer in the liquid.Ab initio calculations in the RHF, MP2, DFT approximations and very accurate G2 calculations were carried out. With one exception, the MeMe conformer had the lowest enthalpy in all these calculations, the MeH had the intermediate and the MeF the highest enthalpy, and the calculations were in good agreement with the measurements.  相似文献   

6.
Behavior of the regularity modes of isotactic polypropylene is analyzed in Raman spectra of a number of random propylene/olefin copolymers. The regularity modes at 809, 841, 973, 998, and 1220 cm−1 decrease in intensity with growth of the content of the incorporated monomer. For the lines at 809, 973, and 1220 cm−1 the rate of intensity damping varies depending on the structure of the incorporated monomer. The type of the incorporated monomer has inconsiderable effect on the evolution of intensity of the regularity bands at 841 and 998 cm−1. Anomalous behavior of the mode at 1220 cm−1 was observed for the propylene/1-butene copolymers.  相似文献   

7.
Raman spectrum of the meso tetraphenylporphine (TPP) deposited onto smooth copper surface as thin film were recorded in the region 200–1700 cm−1. To investigate the effect of meso-phenyl substitution rings on the vibrational spectrum of free base porphyrin, we calculated Raman and infrared (IR) spectra of the meso-tetraphenylporphine (TPP), meso tetramethylporphine (TMP), copper (II)porphine (CuPr) and free base porphine (FBP) at the B3LYP/6-311+G(d,p) level of the density functional theory (DFT). The observed Raman spectrum of the TPP is assigned based on the calculated its Raman spectrum in connection with the calculated spectra of the TMP, CuPr and FBP by taking into account of their corresponding vibrational motions of the Raman modes of frequencies. Results of the calculations clearly indicated that the meso tetraphenyl substitution rings are totally responsible for the observed Raman bands at ∼1593, 1234 and 1002 cm−1. The calculated and observed Raman spectra also suggested that the observed Raman band with a medium intense at 962 cm−1 might result from the surface plasmon effect. Furthermore, the observed Raman bands with medium intense at ∼334 and ∼201 cm−1 are as results of the dimerization or aggregation of the TPP or would be that related to intramolecular interaction. We also calculated IR spectra of these molecules at same level of the theory. To investigate the solvent effect on the vibrational spectrum of porphine, the Raman and IR spectra of the TPP and FBP are calculated in solution phase where water used as solvent. The results of these calculation indicated that there is no any significant effect on the vibrational spectrum of the TPP.  相似文献   

8.
We investigate the nature of bonding and charge states in (U1−yCey)O2 (y = 0.0, 0.2, 0.4, 0.6, 0.8 and 1.0) by Raman spectroscopy. Raman spectrum of UO2 exhibits two prominent bands below 1000 cm−1, a F2g mode at 446 cm−1 and a F1u LO mode at 578 cm−1. As y is increased from 0 to 0.6, the F1u exhibits a large blue shift of 90 cm−1, and from y = 0.6 to 1.0, a red shift of 54 cm−1. We show that our results can be interpreted as arising from anisotropic compression/relaxation of the lattice under Ce substitution and this can give an indication of its charge states. Alternate interpretations have been given in the literature on the effect of substituents and dopants to the Raman spectra of UO2 and CeO2. The present interpretation of chemical stress effects can be taken as another plausible explanation.  相似文献   

9.
《Vibrational Spectroscopy》2007,43(2):288-291
We report significant difference in the Raman spectra of two different kinds of CaB6 single crystals grown from boron purity 99.9% (3N) or 99.9999% (6N), respectively. Our Raman spectra of CaB6 (3N), which are similar to those of previous measurement [N. Ogita, S. Nagai, N. Okamoto, M. Udagawa, F. Iga, M. Sera, J. Akimitsu, S. Kunii, Phys. Rev. B 68 (2003) 224305], show peaks at 781.3 cm−1 (T2g), 1140.1 cm−1 (Eg), and 1283.5 cm−1 (A1g). The Eg mode shows a characteristic double-peak feature due to an additional weak broad peak centered at 1156.0 cm−1. However, the Raman spectra of CaB6 (6N) show sharp peaks at 772.5 cm−1 (T2g), 1137.9 cm−1 (Eg), and 1266.6 cm−1 (A1g). The peak frequencies are down shifted as much as 17 cm−1. In addition, no additional peak feature is observed for the Eg mode so that the mode is symmetric in the case of CaB6 (6N). The X-ray powder diffraction patterns for both CaB6 (3N) and CaB6 (6N) show that the lattice parameters are essentially the same. The majority of the impurity in the 99.9% (3N) boron is assessed to be C. Thus we prepared Ca(B0.995C0.005)6, CaB6 (6N) doped with C, and looked for the difference in the Raman spectra. The Raman spectra of Ca(B0.995C0.005)6 are nearly identical to those of CaB6 (6N), indicating that the difference in the Raman spectra of CaB6 (3N) and CaB6 (6N) is not due to C impurity. However, presence of impurity, even if small amount, seems to be enough to trigger local-structure changes to lower symmetry inducing the difference in Raman spectra of CaB6 (3N) and CaB6 (6N).  相似文献   

10.
Raman and infrared (IR) spectroscopy are complementary spectroscopic techniques. However, measurement of Raman and IR spectra are commonly carried out on separate instruments. A dispersive system that enables both Raman spectroscopy and NIR spectroscopy was designed, built, and tested. The prototype system measures spectral ranges of 2600–300 cm−1 and 752–987 nm for Raman and NIR channels, respectively. A wavelength accuracy better than 0.6 nm and spectral resolution better than 1 nm (14.4 cm−1 for Raman channel) could be achieved with our configuration. The linearity of spectral response was better than 99.8%. The intensity stability of the instrument was found to be 0.7% and 0.4% for Raman and NIR channels, respectively. The performance of the instrument was evaluated using binary aqueous solutions of ethanol and ovalbumin. It was found that ethanol concentrations (2–10%) could be predicted with a root mean squared error of prediction (RMSEP) of 0.45% using Raman peak height at 882.2 cm−1. Quantification of ovalbumin concentration (8–16 g/L) in aqueous solutions and in denatured states yielded RMSEP values of 1.05 g/L and 0.74 g/L, respectively. Using concentration as external perturbation in two-dimensional correlation spectroscopy (2DCOS), heterospectral correlation analysis revealed the relationship between NIR and Raman spectra.  相似文献   

11.
The phosphate mineral series eosphorite–childrenite–(Mn,Fe)Al(PO4)(OH)2·(H2O) has been studied using a combination of electron probe analysis and vibrational spectroscopy. Eosphorite is the manganese rich mineral with lower iron content in comparison with the childrenite which has higher iron and lower manganese content. The determined formulae of the two studied minerals are: (Mn0.72,Fe0.13,Ca0.01)(Al)1.04(PO4, OHPO3)1.07(OH1.89,F0.02)·0.94(H2O) for SAA-090 and (Fe0.49,Mn0.35,Mg0.06,Ca0.04)(Al)1.03(PO4, OHPO3)1.05(OH)1.90·0.95(H2O) for SAA-072. Raman spectroscopy enabled the observation of bands at 970 cm−1 and 1011 cm−1 assigned to monohydrogen phosphate, phosphate and dihydrogen phosphate units. Differences are observed in the area of the peaks between the two eosphorite minerals. Raman bands at 562 cm−1, 595 cm−1, and 608 cm−1 are assigned to the ν4 bending modes of the PO4, HPO4 and H2PO4 units; Raman bands at 405 cm−1, 427 cm−1 and 466 cm−1 are attributed to the ν2 modes of these units. Raman bands of the hydroxyl and water stretching modes are observed. Vibrational spectroscopy enabled details of the molecular structure of the eosphorite mineral series to be determined.  相似文献   

12.
The mineral ettringite has been studied using a number of techniques, including XRD, SEM with EDX, thermogravimetry and vibrational spectroscopy. The mineral proved to be composed of 53% of ettringite and 47% of thaumasite in a solid solution. Thermogravimetry shows a mass loss of 46.2% up to 1000 °C. Raman spectroscopy identifies multiple sulphate symmetric stretching modes in line with the three sulphate crystallographically different sites. Raman spectroscopy also identifies a band at 1072 cm−1 attributed to a carbonate symmetric stretching mode, confirming the presence of thaumasite. The observation of multiple bands in the ν4 spectral region between 700 and 550 cm−1 offers evidence for the reduction in symmetry of the sulphate anion from Td to C2v or even lower symmetry. The Raman band at 3629 cm−1 is assigned to the OH unit stretching vibration and the broad feature at around 3487 cm−1 to water stretching bands. Vibrational spectroscopy enables an assessment of the molecular structure of natural ettringite to be made.  相似文献   

13.
Results of inelastic neutron scattering (INS), infra-red (IR), Raman and 1H NMR spectroscopy used for investigations on the l-asparagine dynamics are reported. The crystallographic structure and experimental vibrational spectra are compared with those calculated by the DFT methods applied to the solid state. Very good conformity of the experimental and theoretical structures has been found. The NH3+ torsional vibration mode is observed in the INS spectra at 494 cm−1, while the bands assigned to the vibrations of the strong NH⋯O hydrogen bonds are observed at 2849, 2650, and 2480 cm−1 in the IR spectrum. A 1H NMR investigation has been carried out at 26.75 MHz in the temperature range 150–300 K. For l-asparagine the activation energy needed for the NH3+ group reorientation is equal 5.6 kcal/mol.  相似文献   

14.
The samples of dibarium magnesium orthoborate Ba2Mg(BO3)2 were synthesized by solid-state reaction. The X-ray diffraction (XRD) patterns and Raman spectra of the samples were collected. Electronic structure and vibrational spectroscopy of Ba2Mg(BO3)2 were systematically investigated by first principle calculation. A direct band gap of 4.4 eV was obtained from the calculated electronic structure results. The top valence band is constructed from O 2p states and the low conduction band mainly consists of Ba 5d states. Raman spectra for Ba2Mg(BO3)2 polycrystalline were obtained at ambient temperature. The factor group analysis results show the total lattice modes are 5Eu + 4A2u + 5Eg + 4A1g + 1A2g + 1A1u, of which 5Eg + 4A1g are Raman-active. Furthermore, we obtained the Raman active vibrational modes as well as their eigenfrequencies using first-principle calculation. With the assistance of the first-principle calculation and factor group analysis results, Raman bands of Ba2Mg(BO3)2 were assigned as Eg (42 cm−1), A1g (85 cm−1), Eg (156 cm−1), Eg (237 cm−1), A1g (286 cm−1), Eg (564 cm−1), A1g (761 cm−1), A1g (909 cm−1), Eg (1165 cm−1). The strongest band at 928 cm−1 in the experimental spectrum is assigned to totally symmetric stretching mode of the BO3 units.  相似文献   

15.
We measured 785 nm excited Raman and infrared spectra of pentacene-d14. The observed spectra were assigned on the basis of the Raman and infrared spectra calculated by the density functional theory (DFT) method at the B3LYP/6⬜311 + G** level. We measured 785 nm excited Raman spectrum of a pentacne-d14:C60 bulk heterojunction film. The spectrum was assigned on the basis of the wavenumber shifts upon deuteration of pentacene. The assignments of the 1462 and 493 cm↙1 Ag bands of C60 were confirmed. The 511, 453, and 256 cm↙1 bands, which were observed only in pentacene:C60 bulk heterojunction films, did not show large deuteration shifts. This result indicates that the 511, 453, and 256 cm↙1 bands are attributed to activation of the silent modes of C60 due to symmetry lowering.  相似文献   

16.
A phenoxonium cation intermediate was successfully observed in situ by laser Raman spectroscopy during the electro-oxidation of 4-methoxyphenol in a lithium perchlorate/nitromethane electrolyte solution as shown by the Raman bands at 1665 cm−1 and 1615 cm−1. These conditions are those used for the electrosynthesis of dihydrobenzofurans by a formal [3+2] cycloaddition when the oxidation is carried out in the presence of an olefin.  相似文献   

17.
《Vibrational Spectroscopy》2010,52(2):283-288
The far-infrared and Raman spectra of binuclear molecules [Me2AuX]2 (X = Cl, Br, I) and [Me2Au(OOCR)]2 (R = Me, CF3, But, Ph) in the 600–70 cm−1 region are reported. The experimentally measured vibrational frequencies of [Me2AuX]2 are in a good agreement with density functional theory predictions. The Au…Au vibrational interactions predicted to be in the 270–60 cm−1 region of [Me2AuX]2 far-IR and Raman spectra have been observed. The Raman-active Au…Au vibrations of the [Me2Au(OOCR)]2 molecules were found to be in the same region as those of [Me2AuX]2. The Au–X stretching modes were observed between 100 and 250 cm−1 in accordance with the DFT predictions. Their frequencies in the IR spectra of [Me2AuX]2 increase in the sequence I < Br < Cl while the AuC2 stretching frequencies decrease in the same order. This fact might be an evidence of the decreasing covalent character of the gold-halogen bridges. The Au–O stretching bands of dimethylgold(III) carboxylates have been observed in the 500–250 cm−1 region, and Au–C stretching frequencies of both [Me2AuX]2 and [Me2Au(OOCR)]2 compounds have been found between 600 and 500 cm−1.  相似文献   

18.
Raman spectra of coquandite Sb6O8(SO4)·(H2O) were studied, and related to the structure of the mineral. Raman bands observed at 970, 990 and 1007 cm?1 and a series of overlapping bands are observed at 1072, 1100, 1151 and 1217 cm?1 are assigned to the SO42? ν1 symmetric and ν3 antisymmetric stretching modes respectively. Raman bands at 629, 638, 690, 751 and 787 cm?1 are attributed to the SbO stretching vibrations. Raman bands at 600 and 610 cm?1 and at 429 and 459 cm?1 are assigned to the SO42? ν4 and ν2 bending modes. Raman bands at 359 and 375 cm?1 are assigned to O–Sb–O bending modes. Multiple Raman bands for both SO42? and SbO stretching vibrations support the concept of the non-equivalence of these units in the coquandite structure.  相似文献   

19.
《Vibrational Spectroscopy》2007,43(2):366-379
Variable temperature DRIFTS has been used to investigate temperature variation of the number and population of the different environments occupied by water on unmilled kaolin and samples ball milled for 3, 10 and 30 min. Increasing the milling time resulted in structural damage of the kaolin and an increase in the amount of hydrogen bonded water indicated by bands in the OH stretching and bending regions. Curve fitting of the spectra, collected at 50 °C intervals in the temperature range 25–500 °C, established that the intensity of the bands diminished as the temperature was increased but also revealed bands that were more stable to high temperatures or were generated as the sample temperature was increased. Bands at 3750, 3386 and 3200 cm−1 and 1680, 1650, 1634 and 1600–1580 cm−1 were identified in the OH-stretching and bending regions, respectively. In particular a band at 1670 cm−1 has been attributed to strongly hydrogen bonded water which acts to hold the deformed kaolin sheets together. Upon aging the samples the intensity of this band decreased and was replaced by a band at 1630 cm−1. Boehmite was tentatively identified as a product of the milled kaolin.  相似文献   

20.
Raman spectra of mineral peretaite Ca(SbO)4(OH)2(SO4)2·2H2O were studied, and related to the structure of the mineral. Raman bands observed at 978 and 980 cm?1 and a series of overlapping bands observed at 1060, 1092, 1115, 1142 and 1152 cm?1 are assigned to the SO42? ν1 symmetric and ν3 antisymmetric stretching modes. Raman bands at 589 and 595 cm?1 are attributed to the SbO symmetric stretching vibrations. The low intensity Raman bands at 650 and 710 cm?1 may be attributed to SbO antisymmetric stretching modes. Raman bands at 610 cm?1 and at 417, 434 and 482 cm?1 are assigned to the SO42? ν4 and ν2 bending modes, respectively. Raman bands at 337 and 373 cm?1 are assigned to O–Sb–O bending modes. Multiple Raman bands for both SO42? and SbO stretching vibrations support the concept of the non-equivalence of these units in the peretaite structure.  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