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 264 毫秒
1.
The high‐resolution stimulated Raman spectra of the ν2 and ν3 bands of C2H4 have been recorded and analyzed separately by means of the tensorial formalism developed in Dijon and Reims for X2Y4 asymmetric‐top molecules. For the ν2 band, a total of 191 lines were assigned and fitted. We obtained a global root mean square deviation of 1.86 × 10− 3 cm− 1. For the ν3 band analyzed in interaction with the ν6 infrared band, a total of 185 lines were assigned and fitted. We obtained a global root mean square deviation of 1.29 × 10− 3 cm− 1. Both analyses lead to very satisfactory synthetic spectra. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

2.
The high‐resolution stimulated Raman spectra of the ν1/ν5 C–H stretching bands of C2H4 have been recorded and analyzed by means of the tensorial formalism developed in Dijon for X2Y4 asymmetric‐top molecules. A total of 689 lines (428 for ν5 and 261 for ν1) were assigned and fitted as a dyad including Coriolis coupling constants. We obtained a global root mean square deviation of 4.39 × 10− 3 cm− 1 (4.61 × 10− 3 cm− 1 for ν1, 4.25 × 10− 3 cm− 1 for ν5). The nearby 2ν2 band, extrapolated from ν2, was included in the analysis. However, no interaction parameter involving it could be fitted. The analysis is quite satisfactory, although some parts of ν5 are not very well reproduced, probably indicating some yet unidentified resonances. This region is indeed quite dense, with many interacting dark states that cannot be included at present. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

3.
The spectra of the ν1, 2ν1ν1, ν2, 2ν2, and 3ν2ν2 bands of CF4 were obtained with a quasi‐continuous wave stimulated Raman spectrometer. These five bands were studied at a temperature of 135 and 300 K (for the hot bands). The spectrum of ν1 was obtained at a sample pressure of 2 mbar. For the spectra of the other regions, which are much weaker, higher pressures were used. The analysis has been performed thanks to the xtds and spview softwares developed in Dijon for such highly symmetric molecules. Combining the present results with a previous infrared study, we could determine a very accurate value for the C–F equilibrium bond length, i.e. re = 1.31588(6) Å. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

4.
5.
The high‐resolution stimulated Raman spectrum of the 2ν10 band located at 1664.16 cm−1 of C2H4 has been reanalyzed, thanks to the tensorial formalism developed in Dijon for X2Y4 asymmetric‐top molecules. A total of 191 lines were assigned and fitted as a single band without including perturbations such as Fermi or Coriolis coupling constants. We obtained a global root mean square deviation of 8.5 × 10−3 cm−1. Further investigations are required to include interactions with the ν2 and ν7 + ν10 bands. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

6.
We present the first high‐resolution stimulated Raman study of osmium tetroxide (OsO4). Lines from the ν1 totally symmetric stretching fundamental have been assigned. These data together with the infrared assignments of the ν3 band previously recorded (M. Louviot et al., J. Quant. Spectrosc. Radiat. Transfer, 2012, 113, 119–127) allowed a refinement of the analysis of the ν1/ν3 stretching dyad. We found that the ν1 band has an unusual positive isotopic shift of approximately 0.32 cm− 1/amu, which gives further evidence that the stretching dyad should be perturbed by a complex nearby bending band polyad. This work is part of a global effort to analyze all fundamental bands of OsO4 to obtain a more precise experimental value of the ground state bond length for this heavy metal‐containing molecule. The result could serve as a benchmark for high‐level quantum chemistry calculations. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

7.
The Raman spectra of liquid carbon disulfide (CS2) diluted with benzene (C6H6) have been measured. By changing the CS2, the concentration, we found an asymmetric wavenumber shift phenomenon. With decreasing concentration of CS2, the position of the ν1 (655 cm−1) band remains practically unchanged, and the 2ν2 (796 cm−1) band shifts toward higher wavenumbers. To interpret this asymmetric wavenumber shift phenomenon of the Fermi doublet ν1 − 2ν2 in the Raman spectra satisfactorily, we propose a modified Bertran model. The values of the Fermi resonance (FR) parameters of CS2 at different concentrations were calculated using the Bertran equations. In addition, we found the fundamental ν2, which should be independent of the FR interaction, shifted to higher wavenumbers as the concentration decreased. This shift was probably driven by the tuning of the FR. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

8.
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   

9.
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   

10.
The Raman scattering cross section (RSCS) is an important parameter in the applications of Raman spectroscopy to make quantitative analysis. To date, the dependence of the RSCS on concentration has remained unclear. Nitrate aerosols can easily achieve a supersaturated state, which provides a way to obtain the RSCS especially under this state. In this study, Raman spectra of NaNO3 and Mg(NO3)2 solutions are obtained with molar water‐to‐solute ratios (WSRs) ranging from 84.2 to 2.30 and 93.8 to 7.32, respectively. With decreasing WSR, a shift to higher wavenumbers of the symmetric stretching band of nitrate ion, i.e. ν1(NO3), is observed, indicating the formation of various ion pairs. Meanwhile, the area ratio between the strongly and weakly hydrogen‐bonded components of water O H stretching envelope, i.e. ν(H2O), reduces as the WSR decreases, implying the transformation of water molecules from strong hydrogen‐bonding structures to the weak ones. However, a good linear relationship is revealed between the integrated intensity ratio of the ν(H2O) band to ν1(NO3) band and WSR. The results suggest that the RSCSs of NO3 and H2O are insensitive to the structures of both ion pairs and hydrogen‐bonding structures. This observation points to the possibility of conducting quantitative analysis through the area ratio of the ν(H2O) band to the ν1(NO3) band with Raman spectra without considering the formation of ion pairs and the variation of the hydrogen‐bonding structure. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

11.
Two hydrated hydroxy magnesium carbonate minerals brugnatellite and coalingite with a hydrotalcite‐like structure were studied by Raman spectroscopy. Intense bands are observed at 1094 cm−1 for brugnatellite and at 1093 cm−1 for coalingite attributed to the CO32−ν1 symmetric stretching mode. Additional low intensity bands are observed at 1064 cm−1. The existence of two symmetric stretching modes is accounted for in terms of different anion structural arrangements. Very low intensity bands at 1377 and 1451 cm−1 are observed for brugnatellite, and the Raman spectrum of coalingite displays two bands at 1420 and 1465 cm−1 attributed to the (CO3)2−ν3 antisymmetric stretching modes. Very low intensity bands at 792 cm−1 for brugnatellite and 797 cm−1 for coalingite are assigned to the CO32− out‐of‐plane bend (ν2). X‐ray diffraction studies by other researchers have shown that these minerals are disordered. This is reflected in the difficulty of obtaining Raman spectra of reasonable quality and explains why the Raman spectra of these minerals have not been previously or sufficiently described. A comparison is made with the Raman spectra of other hydrated magnesium carbonate minerals. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

12.
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   

13.
Raman spectroscopy has been used to study the arsenate minerals haidingerite Ca(AsO3OH)·H2O and brassite Mg(AsO3OH)·4H2O. Intense Raman bands in the haidingerite spectrum observed at 745 and 855 cm−1 are assigned to the (AsO3OH)2−ν3 antisymmetric stretching and ν1 symmetric stretching vibrational modes. For brassite, two similarly assigned intense bands are found at 809 and 862 cm−1. The observation of multiple Raman bands in the (AsO3OH)2− stretching and bending regions suggests that the arsenate tetrahedrons in the crystal structures of both minerals studied are strongly distorted. Broad Raman bands observed at 2842 cm−1 for haidingerite and 3035 cm−1 for brassite indicate strong hydrogen bonding of water molecules in the structure of these minerals. OH···O hydrogen‐bond lengths were calculated from the Raman spectra based on empirical relations. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

14.
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   

15.
Raman spectroscopy has been used to study the rare‐earth mineral churchite‐(Y) of formula (Y,REE)(PO4) ·2H2O, where rare‐earth element (REE) is a rare‐earth element. The mineral contains yttrium and, depending on the locality, a range of rare‐earth metals. The Raman spectra of two churchite‐(Y) mineral samples from Jáchymov and Medvědín in the Czech Republic were compared with the Raman spectra of churchite‐(Y) downloaded from the RRUFF data base. The Raman spectra of churchite‐(Y) are characterized by an intense sharp band at 975 cm−1 assigned to the ν1 (PO43−) symmetric stretching mode. A lower intensity band observed at around 1065 cm−1 is attributed to the ν3 (PO43−) antisymmetric stretching mode. The (PO43−) bending modes are observed at 497 cm−12) and 563 cm−14). Some small differences in the band positions between the four churchite‐(Y) samples from four different localities were found. These differences may be ascribed to the different compositions of the churchite‐(Y) minerals. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

16.
The pressure dependences of the peaks observed in the micro‐Raman spectra of Prussian blue (Fe4[Fe(CN)6]3), potassium ferricyanide (K3[Fe(CN)6]), and sodium nitroprusside (Na2[Fe(CN)5(NO)]·2H2O) have been measured up to 5.0 GPa. The vibrational modes of Prussian blue appearing at 201 and 365 cm−1 show negative dν/dP values and Grüneisen parameters and are assigned to the transverse bending modes of the Fe C N Fe linkage which can contribute to a negative thermal expansion behavior. A phase transition occurring between 2.0 and 2.8 GPa in potassium ferricyanide is shown by changes in the spectral region 150–700 cm−1. In the spectra of the nitroprusside ion, there are strong interactions between the FeN stretching mode and the FeNO bending and the axial CN stretching modes. The pressure dependence of the NO stretching vibration is positive, 5.6 cm−1 GPa−1, in contrast to the negative behavior in the iron(II)‐meso‐tetraphenyl porphyrinate complex. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

17.
Raman spectroscopy complemented with infrared spectroscopy has been used to study the rare‐earth‐based mineral decrespignyite [(Y,REE)4Cu(CO3)4Cl(OH)5· 2H2O] and the spectrum compared with the Raman spectra of a series of selected natural halogenated carbonates from different origins including bastnasite, parisite and northupite. The Raman spectrum of decrespignyite displays three bands at 1056, 1070 and 1088 cm−1 attributed to the CO32− symmetric stretching vibration. The observation of three symmetric stretching vibrations is very unusual. The position of the CO32− symmetric stretching vibration varies with the mineral composition. The Raman spectrum of decrespignyite shows bands at 1391, 1414, 1489 and 1547 cm−1, whereas the Raman spectra of bastnasite, parisite and northupite show a single band at 1433, 1420 and 1554 cm−1, respectively, assigned to the ν3 (CO3)2− antisymmetric stretching mode. The observation of additional Raman bands for the ν3 modes for some halogenated carbonates is significant in that it shows distortion of the carbonate anion in the mineral structure. Four Raman bands are observed at 791, 815, 837 and 849 cm−1, which are assigned to the (CO3)2−ν2 bending modes. Raman bands are observed for decrespignyite at 694, 718 and 746 cm−1 and are assigned to the (CO3)2−ν4 bending modes. Raman bands are observed for the carbonate ν4 in‐phase bending modes at 722 cm−1 for bastnasite, 736 and 684 cm−1 for parisite and 714 cm−1 for northupite. Multiple bands are observed in the OH stretching region for decrespignyite, bastnasite and parisite, indicating the presence of water and OH units in the mineral structure. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

18.
Infrared spectra of 1,2‐bis(trifluorosilyl)ethane (SiF3CH2CH2SiF3) were obtained in the vapour and liquid phases, in argon matrices and in the solid phase. Raman spectra of the compound as a liquid were recorded at various temperatures between 293 and 270 K and spectra of an apparently crystalline solid were observed. The spectra revealed the existence of two conformers (anti and gauche) in the vapour, liquid and in the matrix. When the vapour was chock‐frozen on a cold finger at 78 K and annealed to 150 K, certain weak Raman bands vanished in the crystal. The vibrational spectra of the crystal demonstrated mutual exclusion between IR and Raman bands in accordance with C2h symmetry. Intensity variations between 293 and 270 K of pairs of various Raman bands gave ΔH(gauche—anti) = 5.6 ± 0.5 kJ mol−1 in the liquid, suggesting 85% anti and 15% gauche in equilibrium at room temperature. Annealing experiments indicate that the anti conformer also has a lower energy in the argon matrices, is the low‐energy conformer in the liquid and is also present in the crystal. The spectra of both conformers have been interpreted, and 34 anti and 17 gauche bands were tentatively identified. Ab initio and density functional theory (DFT) calculations were performed giving optimized geometries, infrared and Raman intensities and anharmonic vibrational frequencies for both conformers. The conformational energy difference derived in CBS‐QB3 and in G3 calculations was 5 kJ mol−1. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

19.
The electronic (UV‐vis) and resonance Raman (RR) spectra of a series of para‐substituted trans‐β‐nitrostyrenes were investigated to determine the influence of the electron donating properties of the substituent (X = H, NO2, COOH, Cl, OCH3, OH, N(CH3)2, and O) on the extent of the charge transfer to the electron‐withdrawing NO2 group directly linked to the ethylenic (C = C) unit. The Raman spectra and quantum chemical calculations show clearly the correlation of the electron donating power of the X group with the wavenumbers of the νs(NO2) and ν (C = C)sty normal modes. In conditions of resonance with the lowest excited electronic state, one observes for X = OH and N(CH3)2 that the symmetric stretching of the NO2, νs(NO2), is the most substantially enhanced mode, whereas for X = O, the chromophore is extended over the whole molecule, with substantial enhancement of several carbon backbone modes. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

20.
The B‐band resonance Raman spectra of 2(1H)‐pyridinone (NHP) in water and acetonitrile were obtained, and their intensity patterns were found to be significantly different. To explore the underlying excited state tautomeric reaction mechanisms of NHP in water and acetonitrile, the vibrational analysis was carried out for NHP, 2(1D)‐pyridinone (NDP), NHP–(H2O)n (n = 1, 2) clusters, and NDP–(D2O)n (n = 1, 2) clusters on the basis of the FT‐Raman experiments, the B3LYP/6‐311++G(d,p) computations using PCM solvent model, and the normal mode analysis. Good agreements between experimental and theoretically predicted frequencies and intensities in different surrounding environments enabled reliable assignments of Raman bands in both the FT‐Raman and the resonance Raman spectra. The results indicated that most of the B‐band resonance Raman spectra in H2O was assignable to the fundamental, overtones, and combination bands of about ten vibration modes of ring‐type NHP–(H2O)2 cluster, while most of the B‐band resonance Raman spectra in CH3CN was assigned to the fundamental, overtones, and combination bands of about eight vibration modes of linear‐type NHP–CH3CN. The solvent effect of the excited state enol‐keto tautomeric reaction mechanisms was explored on the basis of the significant difference in the short‐time structural dynamics of NHP in H2O and CH3CN. The inter‐molecular and intra‐molecular ESPT reaction mechanisms were proposed respectively to explain the Franck–Condon region structural dynamics of NHP in H2O and CH3CN.Copyright © 2015 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