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 421 毫秒
1.
The IR and Raman spectra of gaseous and solid CH3TiX3 and CD3TiX3 species (X = Cl, Br, I) are reported. The gas phase spectra have been recorded between 4000 and 20 cm?1 at pressures of 1 atm and 4 atm at 350 K and the Raman spectra of the solid phase recorded at 4.2 K. Internal rotation barriers and thermodynamic functions have been calculated.  相似文献   

2.
Raman Spectra of the System PCl5? TeCl4 Raman spectra of solid and molten PCl5? TeCl4 mixtures have been recorded. The solid 1:1 mixture contains the compound nPCl4 + (TeCl5?)n. In the molten state there are different fragmentation equilibria between the polymeric (TeCl5)?n, TeCl62?, and lower charged units. The spectra of TeCl4 rich samples indicate different species like Te3Cl13?, Te2Cl102?, and (TeCl5?)n. In such melts the equilibria are shifted to lower charged oligomeres and TeCl5?.  相似文献   

3.
A new series of 9-fluorenone-iodine-alkali iodide complexes (FIMI, M=Na, K, Rb or Cs) has been prepared from chloroform. These complexes yield needle-like crystals with metallic luster along the needle axis. The longitudinal dc conductivities at room temperature are in the order of 10?3 to 10?5 S cm?1. They all have semiconductive properties with activation energies of about 0.4 to 0.7 eV. A significant result is that the conductivity varies systematically with the cation size. A decrease in the cation size results in an increase in the conductivity. Resonance Raman spectra show that iodine is present as a linear chain of symmetric I?3 units in these complexes. Infrared spectra of these complexes are discussed with an attempt to estimate the interaction between the alkali cation and the carbonyl group of 9-fluorenone.  相似文献   

4.
The IR and Raman spectra of arylselenium tribromides X = C6H4SeBr3 where X = H, Br, Cl, CH3 are reported. It is suggested that the solids are possibly tetramers in the solid state formed by donor-acceptor interaction between units RSeBr2+ and Br?. Each selenium atom is in a highly symmetrical or pseudo-symmetrical environment.  相似文献   

5.
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   

6.
Isothermal cells of nickel with diamon windows were used to study various melts and vapours by infrared emission, transmission and reflectance techniques in the 860 to 300 K range with an evacuable Fourier transform spectrometer.IR vapour spectra of AlX3 (X = Cl, Br, I) and GaCl3 in transmission and emission were measured between 700 and 50 cm?1. A comparable signal/noise ratio between the transmission and emission spectra was obtained above 200 cm?1, below 200 cm?1 the transmission spectra had better quality. The dimer and monemer spectra were assigned in terms of D2h and D3h symmetry, respectively, and compared with earlier Raman and IR matrix isolation data.Emission spectra of chloroaluminates AlkAlCl4 (Alk = Li, Na, K, Rb, Cs) were recorded as melts between 1500 and 50 cm?1. Increasing distortion of the tetrahedral AlCl?4 ion in the series Cs<Rb<K<Na<Li was observed. Emission spectra of AlkAl2Cl7 (Alk = Li, Na, K, Cs) indicate D3d symmetry for Al2Cl?7 with a linear Al-Cl-Al bridge as proposed from earlier Raman data. As a demonstration of reflectance technique an IR spectrum of ZnAl2Cl8 at ambient temperature is presented.  相似文献   

7.
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   

8.
Natural berlinite from a heated sedimentary sequence in Cioclovina Cave (Romania) was studied using Raman spectroscopy complemented with infrared techniques. Vibrational data acquired at room temperature were compared with those reported for synthetic berlinite in ambient conditions. The symmetry of the (PO4)3? units is confirmed by the observation of characteristic bands attributed to the ν1(PO4)3? stretching mode, both the ν4 and ν2 bending regions at 500–595 cm?1, and 350–500 cm?1, respectively. The berlinite Raman fingerprint was unambiguously identified at 1111 and 1104 cm?1, confirming the identity of the species and elucidating some controversial reports in the mineralogy field.The vibrational data of natural berlinite relates to its crystallography, and along with the spectra–structure correlation, confirmed an almost ideal natural berlinite crystal.  相似文献   

9.
Measurements are reported of the surface-enhanced Raman of 12CN and 13CN (and of isotopically labelled mixtures) adsorbed at silver electrodes. The spectra are shown to arise from a complex species whose coordination number does not change with electrode potential. This species is probably a [Ag(CN2)]? entity having C2v symmetry; at very negative potentials a reduced form of this complex [Ag(CN)2]2? coexists with the formally Ag1 species at the surface. The shifts in band position are interpreted in terms of changes in the bond character of the adsorbed CN? species. The spectrum of water coadsorbed with CN? is also markedly dependent on the charge density of the adsorbed CN? groups.  相似文献   

10.
We report the infrared, Raman, and surface‐enhanced Raman scattering (SERS) spectra of triruthenium dipyridylamido complexes and of diruthenium mixed nickel metal‐string complexes. From the results of analysis on the vibrational modes, we assigned their vibrational frequencies and structures. The infrared band at 323–326 cm?1 is assigned to the Ru3 asymmetric stretching mode for [Ru3(dpa)4Cl2]0–2+. In these complexes we observed no Raman band corresponding to the Ru3 symmetric stretching mode although this mode is expected to have substantial Raman intensity. There is no frequency shift in the Ru3 asymmetric stretching modes for the complexes with varied oxidational states. No splitting in Raman spectra for the pyridyl breathing line indicates similar bonding environment for both pyridyls in dpa , thus a delocalized structure in the [Ru3]6–8+ unit is proposed. For Ru3(dpa)4(CN)2 complex series, we assign the infrared band at 302 cm?1 to the Ru3 asymmetric stretching mode and the weak Raman line at 285 cm?1 to the Ru3 symmetric stretching. Coordination to the strong axial ligand CN weakens the Ru‐Ru bonding. For the diruthenium nickel complex [Ru2Ni(dpa)4Cl2]0–1+, the diruthenium stretching mode νRu‐Ru is assigned to the intense band at 327 and 333 cm?1 in the Raman spectra for the neutral and oxidized forms, respectively. This implies a strong Ru‐Ru metal‐metal bonding.  相似文献   

11.
The bifluoride ion FHF?, (and FDF?), has been substitutionally isolated within single crystal samples of several alkali halides. Infrared and Raman spectra of these crystals have been studied at variable temperature and pressure. The infrared absorptions are strong, whereas the Raman is weak. At low temperatures the bands are very sharp with halfwidths less than 1 cm?1. On applying pressure, ν3 increases in frequency whereas ν2 decreases. On reducing temperature, ν3 decreases in frequency whereas ν2 increases. Hence the effect of volume contraction is overridden in the temperature dependent case. The deuterated spectra confirm that the bifluoride ion is well isolated within the alkali halide matrix.  相似文献   

12.
The infrared spectra of CH3CH2I, CD3CH2I, and CH3CD2I of the vapors and the solids at 170°C have been recorded from 4000-200 cm?1. The Raman spectra of the liquids and vapors have also been recorded and depolarization values have been measured. Assignment of the eighteen fundamental vibrations has been based on depolarization values, band contours, group-frequency correlations, and normal coordinate calculations. A critical discussion of the CH stretching assignments in CH3CH2X molecules is presented.  相似文献   

13.
[Co(NH3)6][AuX4]X2 binary complex salts, where X = Cl? (I) and Br? (II), have been obtained and defined by element, X-ray diffraction, and thermal analyses and by IR, Raman, and electron spectroscopy. The compounds are isostructural. Their structural units are the [Co(NH3)6]3+ complex cations, the [AuX4]? complex anions, and the X? anions. The plane square environment of the gold atom is completed to an elongated bipyramid by two halide ions lying at distances Au...Cl 3.245 Å for I and Au...Br 3.362 Å for II. The thermolysis products of I and II are pure gold and cobalt metal powders when thermolysis is performed under hydrogen and a mixture of metallic gold with cobalt halide in a reaction under an inert atmosphere.  相似文献   

14.
The photodissociation of methyl iodide in various matrices at low temperature was studied. The observed Raman spectra excited by 514.5 nm laser radiation showed that there were two different photolytically produced iodine species isolated in the matrices after illumination by a medium pressure mercury lamp. One species which was dominant at lower iodine concentrations and exhibited a progression with an ωe of 201 cm?1, belonged to the matrix isolated iodine monomer (I2). The other species, which was dominant at higher iodine concentrations with an ωe of approximately 180 cm?1, belonged to the iodine aggregate ((I2)n). Five progressions of resonance Raman or resonance fluorescence of these two species were also observed in the other matrices. The iodine aggregate in the methyl iodide matrix at 77 K was formed in a crystalline structure, while the photolytically generated iodine aggregate from CH3I/Ar (2/3) matrix at 10 K, after illumination with a mercury lamp, was in amorphous form. The rearrangement of photolytically produced iodine aggregate in methyl iodide matrix was observed as a function of the duration of illumination. Local heating effects of the laser radiation might induce the iodine monomer to aggregate in matrices. The photodissociation mechanism of methyl iodide in matrices is also proposed.  相似文献   

15.
Two species of iodide ions (I3? and I5?) are found in iodine—nylon 6 complexes. Orientation of I5? arrays (most likely I2/I3? complex) along the polymer chain and I3? ions perpendicular to the chain axis in uniaxially drawn films and in films with planar orientation suggests that there is and intrinsic relation between the direction of iodide ion arrays and nylon 6 chains. When an unoriented film of nylon 6 in the amorphous or the α crystalline form is treated with an aqueous solution of iodine—potassium iodide, the I3? species in the resulting iodine—nylon complex lie in planes parallel to the surface of the film, and I2/I3? units are oriented normal to the surface of the film. The γ form obtained by desorbing the iodine from this complex shows considerable uniaxial rientation with the nylon chains oriented perpendicular to the plane of the film; this orientation is maintained during the γ to α transition. It is proposed that the iodine-induced orientation of the nylon 6 chains is due to the nucleating effects of the iodide ion species as the iodine diffuses unidirectionally into the film.  相似文献   

16.
Ruthenium(III) Phthalocyanines: Synthesis and Properties of Di(halo)phthalocyaninato(1?)ruthenium(III) Di(halo)phthalocyaninato(1?)ruthenium(III), [Ru(X)2Pc?] (X = Cl, Br, I) is prepared by oxidation of [Ru(X)2Pc2?]? (Cl, Br, OH) with halogene in dichloromethane. The magnetic moment of [Ru(X)2Pc?] is 2,48 μB (X = Cl) resp. 2,56 μB (X = Br) in accordance with a systeme of two independent spins (low spin RuIII and Pc?: S = 1/2). The optical spectra of the red violet solution of [Ru(X)2Pc?] (Cl, Br) are typical for the Pc? ligand with the “B” at 13.5 kK, “Q1” at 19.3 kK and “Q2 region” at 31.9 kK. Sytematic spectral changes within the iron group are discussed. The presence of the Pc? ligand is confirmed by the vibrational spectra, too. Characteristic are the metal dependent bands in the m.i.r. spectra at 1 352 and 1 458 cm?1 and the strong Raman line at 1 600 cm?1. The antisymmetric Ru? X stretch (vas(Ru? X)) is observed at 189 cm?1 (X = I) resp. 234 cm?1 (X = Br). There are two interdependent bands at 295 and 327 cm?1 in the region expected for vas(Ru? Cl) attributed to strong interaction of vas(Ru? Cl) with an out-of-plane Pc? tilting mode of the same irreducible representation. Only the symmetric Ru? Br stretch at 183 cm?1 is selectively enhanced in the resonance-Raman(RR) spectra. The Raman line at 168 cm?1 of the diiodo complex is assigned to loosely bound iodine. The broad band at 978 cm?1 in the RR spectra of the dichloro complex is due to an intraconfigurational transition within the electronic ground state of low spin RuIII split by spin orbit coupling.  相似文献   

17.
Preparation and Vibrational Spectra of trans-[Pt(acac)2X2] (X ? Cl, Br, I, SCN, SeCN, N3) By electrolytical oxidation of [Pt(acac)2] in presence of chloride or bromide, dissolved in dichlormethane, trans-[Pt(acac)2X2], X ? Cl, Br, are formed. On treatment of trans-[Pt(acac)2I2] with silver pseudohalides trans-[Pt(acac)2X2], X ? SCN, SeCN, N3, are obtained. Beside the nearly persistent bands of coordinated acetylacetonate in the Raman spectra the intensive and sharp symmetric, in the IR spectra the corresponding antisymmetric stretching vibration of the X? Pt? X axis is observed. The observance of the rule of mutual exclusion proves the complexes to belong to point group D2h. From the resonance Raman spectrum of trans-[Pt(acac)2I2] for vs (Pt? I), Ag, the harmonic frequency ω1 = 142.45 cm?1 and the inharmonicity constant x11 = 0.48 cm?1 is calculated. In the Raman spectrum of trans-[Pt(acac)2Cl2] vs (Pt? Cl) is splitted by the isotops 35Cl/37Cl into the triplet 340, 335, 330 cm?1 giving the force constant fPtCl = 2.01 N/cm.  相似文献   

18.
In this research, we have used vibrational spectroscopy to study the phosphate mineral kosnarite KZr2(PO4)3. Interest in this mineral rests with the ability of zirconium phosphates (ZP) to lock in radioactive elements. ZP have the capacity to concentrate and immobilize the actinide fraction of radioactive phases in homogeneous zirconium phosphate phases. The Raman spectrum of kosnarite is characterized by a very intense band at 1,026?cm?1 assigned to the symmetric stretching vibration of the PO4 3? ??1 symmetric stretching vibration. The series of bands at 561, 595 and 638?cm?1 are assigned to the ??4 out-of-plane bending modes of the PO4 3? units. The intense band at 437?cm?1 with other bands of lower wavenumber at 387, 405 and 421?cm?1 is assigned to the ??2 in-plane bending modes of the PO4 3? units. The number of bands in the antisymmetric stretching region supports the concept that the symmetry of the phosphate anion in the kosnarite structure is preserved. The width of the infrared spectral profile and its complexity in contrast to the well-resolved Raman spectrum show that the pegmatitic phosphates are better studied with Raman spectroscopy.  相似文献   

19.
We present a detailed study of Raman spectroscopy and photoluminescence measurements on Li‐doped ZnO nanocrystals with varying lithium concentrations. The samples were prepared starting from molecular precursors at low temperature. The Raman spectra revealed several sharp lines in the range of 100–200 cm?1, which are attributed to acoustical phonons. In the high‐energy range two peaks were observed at 735 cm?1 and 1090 cm?1. Excitation‐dependent Raman spectroscopy of the 1090 cm?1 mode revealed resonance enhancement at excitation energies around 2.2 eV. This energy coincides with an emission band in the photoluminescence spectra. The emission is attributed to the deep lithium acceptor and intrinsic point defects such as oxygen vacancies. Based on the combined Raman and PL results, we introduce a model of surface‐bound LiO2 defect sites, that is, the presence of Li+O2? superoxide. Accordingly, the observed Raman peaks at 735 cm?1 and 1090 cm?1 are assigned to Li? O and O? O vibrations of LiO2.  相似文献   

20.
The Raman scattering of nonplanar (form 2) poly(vinylidene fluoride) (PVF2) is described. Unique Raman bands not observed in the infrared spectra are found at 2973, 1437, 1327, 1198, and 1059 cm?1. Band assignments are discussed by comparing infrared and Raman spectra of form 2 PVF2.  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