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 109 毫秒
1.
2.
A full vibrational analysis was recently made for the H3N+CH2CH2N+H3 ion, on the basis of infrared measurements, and a partial vibrational assignment proposed1. The observed fundamentals were interpreted In terms of a C2h, symmetry for the ion. There are thirty-six fundamentals (r - llag + 8Au + 7Bg + 10Bu; Ag and Bg Raman-active only; Au and Bu infrared-active only). In addition to the eighteen infrared-active vibrations, a number of bands was observed which were assigned as formally forbidden Raman-active modes, excited by strong hydrogen bonding causing departure from the exact site symmetry of the species.  相似文献   

3.
The normal Raman and surface‐enhanced Raman scattering (SERS) spectra of flavanthrone and indanthrone were obtained at several excitation wavelengths. The spectral assignments were aided by density functional calculations. Since both molecules have very high symmetry (C2h) including a center of inversion, we expect that the modes of u symmetry will be forbidden in the normal Raman spectrum. However, proximity to the surface causes special SERS enhancement of several of the bu modes, along with somewhat weaker enhancement of the au and bg modes. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

4.
A careful analysis of the Raman spectra of the M′x[M(CN)6]y Prussian blue species has enabled a general model for the interpretation of the ν(CN) vibrational spectra. The spectral patterns are derived from those of the metal ions with local Oh symmetry. Two limiting models are discussed. A ‘localized mode’ model, involving matrix‐isolated species, is in much better accord with the observations than a ‘factor group’ model. The use of the infrared feature as fingerprint of specific M CN M′ units arises from the axis‐specific nature of individual T1u modes. The interpretation of the A1g and Eg Raman features is done in terms of localized vibrations, with involvement of additional energy terms from the lattice motions. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

5.
Abstract

The molybdate‐bearing mineral szenicsite, Cu3(MoO4)(OH)4, has been studied by Raman and infrared spectroscopy. A comparison of the Raman spectra is made with those of the closely related molybdate‐bearing minerals, wulfenite, powellite, lindgrenite, and iriginite, which show common paragenesis. The Raman spectrum of szenicsite displays an intense, sharp band at 898 cm?1, attributed to the ν1 symmetric stretching vibration of the MoO4 units. The position of this particular band may be compared with the values of 871 cm?1 for wulfenite and scheelite and 879 cm?1 for powellite. Two Raman bands are observed at 827 and 801 cm?1 for szenicsite, which are assigned to the ν3(E g ) vibrational mode of the molybdate anion. The two MO4 ν2 modes are observed at 349 (B g ) and 308 cm?1 (A g ). The Raman band at 408 cm?1 for szenicsite is assigned to the ν4(E g ) band. The Raman spectra are assigned according to a factor group analysis and are related to the structure of the minerals. The various minerals mentioned have characteristically different Raman spectra.  相似文献   

6.
The intensities of the Raman lines of anthracene in solution and in the pure crystal at room temperature have been measured using several excitation frequencies (ν0) in the pre-resonance region. The data were used to derive molecular scattering elements αii at each νo for a number of ag fundamentals. While the values of αxx and αyy are constant over the restricted range of excitation used, αzz shows an enhancement as νo moves towards resonance with the first absorption system (I). Overlap factors were determined from the fluorescence spectrum of anthracene in a biphenyl matrix, and these values were used to calculate the contribution αzz(I) due to scattering off the first excited electronic state. Working beyond the oriented-gas assumption improved the fit between calculated and observed enhancement. There is a measurable background contribution to αzz due to scattering off higher excited states and this has an opposite algebraic sign from αzz(I) for all ag fundamentals except possibly for the 394 cm−1 mode.  相似文献   

7.
Histidine is an important and versatile amino acid residue that plays a variety of structural and functional roles in proteins. Although the Raman bands of histidine are generally weak, histidine in the N‐deuterated cationic form with imidazole Nπ D and Nτ D bonds (N‐deuterated histidinium) gives two strong Raman bands assignable to the C4C5 stretch (νCC) and the Nπ C2 Nτ symmetric stretch (νNCN) of the imidazole ring. We examined the Raman spectra of N‐deuterated histidinium in 12 crystals with known structures. The observed νCC and νNCN wavenumbers were analyzed to find empirical correlations with the conformation and hydrogen bonding. The effect of conformation on the vibrational wavenumber was expressed as a threefold cosine function of the Cα Cβ C4C5 torsional angle. The effect of hydrogen bonding at Nπ or Nτ was assumed to be proportional to the inverse sixth power of the distance between the hydrogen and acceptor atoms. Multiple linear regression analysis clearly shows that the conformational effect on the vibrational wavenumber is comparable for νCC and νNCN. The hydrogen bond at Nπ weakly lowers the νCC wavenumber and substantially raises the νNCN wavenumber. On the other hand, the hydrogen bond at Nτ strongly raises the νCC wavenumber but does not affect the νNCN wavenumber. These empirical correlations may be useful in Raman spectral analysis of the conformation and hydrogen bonding states of histidine residues in proteins. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

8.
A systematic study on lattice dynamics of Mn + 1AlCn (n = 1–3) phases using first‐principle calculations is reported, where the Raman‐active and infrared‐active (IR) modes are emphasized. The highest phonon wavenumber is related to the vibration of C atoms. The ‘imaginary wavenumber’ in the phonon spectrum of Nb3AlC2 contributes to the composition gap in Nb‐Al‐C system (Nb2AlC and Nb4AlC3 do appear in experiments, but there are no experimental reports on Nb3AlC2). The full set of Raman‐active and IR‐active modes in the 211, 312, and 413 Mn + 1AXn phases is identified, with the corresponding Raman and IR wavenumbers. The 211, 312, and 413 Mn + 1AXn phases have 4, 6, and 8 IR‐active modes, respectively. There is no distinct difference among the wavenumber ranges of IR‐active modes for 211, 312, and 413 phases, with the highest wavenumber of 780 cm−1 in Ta4AlC3. The Raman wavenumbers of M2AlC phases all decrease with increasing the d‐electron shell number of transition metal M. However, this case is valid only for the Raman‐active modes with low wavenumbers of M3AlC2 and M4AlC3. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

9.
A quantitative assessment of the Raman spectrum emitted from a coarse‐grained polycrystal of multiferroic BiFeO3 has been carried out by means of a polarized Raman microprobe. The dependence of the intensity of Raman phonon modes has been first theoretically modeled as a function of crystal rotation. Then, the Raman tensor elements have been experimentally determined from the analysis of the Ag and Eg vibrational modes. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

10.
ABSTRACT

The high-resolution infrared spectrum of CHD279Br has been investigated by Fourier transform spectroscopy in the range 700–900?cm?1 at an unapodized resolution of 0.0035?cm?1. This spectral region is characterised by the absorptions of the ν5 (814.5185?cm?1) and ν9 (716.9649?cm?1) fundamental bands, corresponding to H–C–Br deformation and CD2 rocking modes, respectively. The ν5 vibration of symmetry species A gives rise to an a-/c-hybrid band with a predominant a-type component, while the ν9 mode of A′′ symmetry produces a b-type envelope. The spectral analysis resulted in the identification of 5290 (J?≤?63 and Ka?≤?13) and 1657 (J?≤?53 and Ka?≤?12) transitions for ν5 and ν9 bands, respectively. The assigned data were fitted using the Watson’s S-reduced Hamiltonian in the Ir representation and the v5?=?1 and v9?=?1 state parameters up to the quartic centrifugal distortion terms have been obtained. From spectral simulations the dipole moment ratio |Δμa/Δμc| of the ν5 band has been determined to be 1.4?±?0.1 while the intensity ratio between ν5 and ν9 fundamentals has been estimated to have a value of 4.3?±?0.5.  相似文献   

11.
Raman spectra of Cs2NaTmCl6 have been recorded using a diamond anvil cell at ambient temperature. The vibrational energy of each of the Raman-active TmCl6−3 moiety modes increases linearly with pressure. The integrated band areas of the ν1(a1g) and ν2(eg) modes are independent of applied pressure. However, the band area of the ν5(t2g) mode shows an anomalous behaviour, which has been qualitatively interpreted as due to electron-phonon coupling of the aΓ5 electronic state with the Γ15(t2g) vibronic state. This interaction between the coupled states is strongest between ca. 10 and 13 GPa at ambient temperature. The results serve to emphasize the specificity of the occurrence of strong electron-phonon coupling for particular transitions of a given rare earth ion.  相似文献   

12.
The fine structure of the fundamental vibrational bands and some combination tones of fullerite C60 in its IR absorption and reflection spectra, as well as in Raman spectra, has been studied. This structure is due to the overlapping components of Davydov and isotopic splittings and the removal of vibrational degeneracy with symmetry lowering. It is shown that for IR F u (i) bands (i = 1–4) and low-frequency H g (1) and A g (1) bands in the Raman spectrum the splittings at room temperature exceed those for the low-temperature phase. The enhancement of intermolecular interaction at elevated temperatures is explained by the nonequilibrium vibrational excitation of the medium as a result of nonlinear interaction of vibrational modes and by the change in the electronic states.  相似文献   

13.
Raman spectroscopy has been used to characterise four natural halotrichites: halotrichite FeSO4.Al2(SO4)3. 22H2O, apjohnite MnSO4.Al2(SO4)3.22H2O, pickingerite MgSO4.Al2(SO4)3.22H2O and wupatkiite CoSO4.Al2(SO4)3.22H2O. A comparison of the Raman spectra is made with the spectra of the equivalent synthetic pseudo‐alums. Energy dispersive X‐ray analysis (EDX) was used to determine the exact composition of the minerals. The Raman spectrum of apjohnite and halotrichite display intense symmetric bands at ∼985 cm−1 assigned to the ν1(SO4)2− symmetric stretching mode. For pickingerite and wupatkiite, an intense band at ∼995 cm−1 is observed. A second band is observed for these minerals at 976 cm−1 attributed to a water librational mode The series of bands for apjohnite at 1104, 1078 and 1054 cm−1, for halotrichite at 1106, 1072 and 1049 cm−1, for pickingerite at 1106, 1070 and 1049 cm−1 and for wupatkiite at 1106, 1075 and 1049 cm−1 are attributed to the ν3(SO4)2− antisymmetric stretching modes of ν3(Bg) SO4. Raman bands at around 474, 460 and 423 cm−1 are attributed to the ν2(Ag) SO4 mode. The band at 618 cm−1 is assigned to the ν4(Bg) SO4 mode. The splitting of the ν2, ν3 and ν4 modes is attributed to the reduction of symmetry of the SO4 and it is proposed that the sulphate coordinates to water in the hydrated aluminium in bidentate chelation. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

14.
Uranopilite, [(UO2)6(SO4)O2(OH)6(H2O)6](H2O)8, the composition of which may vary, can be understood as a complex hydrated uranyl oxyhydroxy sulfate. The structure of uranopilite from different locations has been studied by Raman spectroscopy at 298 and 77 K. A single intense band at 1009 cm−1 assigned to the ν1 (SO4)2− symmetric stretching mode shifts to higher wavenumbers at 77 K. Three low‐intensity bands are observed at 1143, 1117 and 1097 cm−1. These bands are attributed to the (SO4)2− ν3 anti‐symmetric stretching modes. Multiple bands provide evidence that the symmetry of the sulfate anion in the uranopilite structure is lowered. Three bands are observed in the region 843 to 816 cm−1 in both the 298 and 77 K spectra and are attributed to the ν1 symmetric stretching modes of the (UO2)2+ units. Multiple bands prove the symmetry reduction of the UO2 ion. Multiple OH stretching modes prove a complex arrangement of OH groupings and hydrogen bonding in the crystal structure. A series of infrared bands not observed in the Raman spectra are found at 1559, 1540, 1526 and 1511 cm−1 attributed to δ UOH bending modes. U‐O bond lengths in uranyl and O H/dotbondO bond lengths are calculated and compared with those from X‐ray single crystal structure analysis. The Raman spectra of uranopilites of different origins show subtle differences, proving that the spectra are origin‐ and sample‐dependent. Hydrogen‐bonding network and its arrangement in the crystal structure play an important role in the origin and stability of uranopilite. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

15.
Polarised IR and Raman spectra of Na3Li(MoO4)2· 6H2O single crystal were measured. Discussion of the results is based on the factor group approach for the trigonal R 3c(C3v6) space group with Z = 2. The assignment of the observed bands was performed on the basis of their polarisation behaviour and literature data. The obtained results for the spontaneous Raman scattering were used in the analysis of the stimulated Raman spectra of the material studied—a new Raman laser crystal. The promoting modes of the stimulated effect were identified. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

16.
Abstract

The FTIR and FT Raman spectra of benzylidene aniline, and o-hydroxybenzylidene o-hydroxyaniline compounds in the solid state in the wavenumber (1800-200 cm?1) are recorded. An assignment for nearly all fundamentals are proposed. Comparison of the spectra of trans stilbene and benzylidene aniline reveals that v N-Ph stretch for the latter compound is situated at 1368 cm?1 in the IR spectra with medium intensity. for o-hydroxybenzylidene o-hydroxyaniline, the stretching modes v N-Ph, and v C-Ph are observed at 1356 and 1226 cm?1 respectively. the two v O-Ph are observed as intense bands in the IR spectra at 1245 and 1278 cm?1, respectively. the FTIR spectra of the o-hydroxybenzylidene o-hydroxyaniline complexes with Cu(II) and Ni(II) metal ions are also recorded and assigned.  相似文献   

17.
The IR spectrum of the ν8 band of monodeutero methyl fluoride has been recorded with a resolution of 0.004 cm?1 using a Fourier transform spectrometer. The ν8 vibration, of A′′ symmetry species, gives rise to a pure c-type envelope centred at 1298.58 cm?1, resembling a perpendicular band of a prolate symmetric rotor. From the spectral analysis it has been found that ν8 is coupled with ν4 for the low Ka sublevels while for the higher ones other vibrational states, directly or indirectly, are also interacting. For satisfactory reproduction of the rovibrational data of ν8, the Watson's Hamiltonian in the Ir representation has been implemented with interaction parameters, which included first- and second-order a-, b-type and c-type Coriolis couplings with ν4 and ν9, respectively. Being these states also interacting with other levels, the effective spectroscopic parameters of the analysed band up to fourth order have been obtained from a data set formed by the six fundamentals located below 1500 cm?1.  相似文献   

18.
J. S. Singh 《Pramana》2008,70(3):479-486
Laser Raman (200–4000 cm−1) and IR (200–4000 cm−1) spectra of 5-aminouracil were recorded in the region 200–4000 cm−1. Assuming a planar geometry and Cs point group symmetry, it has been possible to assign all the 36 (25a′ + 11a″) normal modes of vibration for the first time. The two NH bonds of the NH2 group appear to be equivalent as the NH2 stretching frequencies satisfy the empirical relation proposed for the two equivalent NH bonds of the NH2 group. The two NH2 stretching frequencies are distinctly separated from the CH/NH ring stretching frequencies. A strong and sharp IR band at 3360 cm−1 could be identified as the anti-symmetric NH2 mode whereas the band at 3290 cm−1 with smaller density could be identified as the symmetric NH2 stretching mode. All other bands have also been assigned different fundamentals/overtones/combinations.   相似文献   

19.
First and second‐order Raman spectra of B6O and their dependence on the wavelength of the excitation line from IR (infrared) to deep UV (ultraviolet) has been studied. The first‐order Raman spectra contain 11 well‐resolved lines of the 12 expected modes 5 A1g + 7 Eg (space group R‐3m, point group D3d). The second‐order Raman spectra contains eight lines that are resolved only in the case of the 244‐nm excitation line. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

20.
The complex orientation dependence in space of Raman active vibrations in the orthorhombic structure of polyethylene (PE) is discussed in terms of Raman tensor elements as intrinsic physical parameters of the lattice. Building upon the symmetry assignment of these vibrational modes, we systematically studied, from both theoretical and experimental viewpoints, the changes of polarized intensity for the Ag and the B2g + B3g vibrational modes with respect to PE molecular orientation. After explicitly expanding the Raman selection rules associated with the Ag and the B2g + B3g modes, introducing them into general expressions of the orientation distribution function, and validating them by means of a least‐square fitting procedure on experimental data, we compare here two mesostructural models for a highly crystallized and self‐aligned PE fiber structure. Stereological arguments are shown concerning the arrangement of orthorhombic fibrils in such a sample that unfold the correct values of five independent Raman tensor elements for orthorhombic PE. Copyright © 2010 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