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 15 毫秒
1.
The silico-phosphate mineral perhamite has been studied using a combination of electron and vibrational spectroscopy. SEM photomicrographs reveal that perhamite morphology consists of very thin intergrown platelets that can form a variety of habits. Infrared spectroscopy in the hydroxyl-stretching region shows a number of overlapping bands which are observed in the range 3581-3078 cm(-1). These wavenumbers enable an estimation to be made of the hydrogen bond distances in perhamite: 3.176(0), 2.880(5), 2.779(6), 2.749(3), 2.668(1) and 2.599(7)A. Intense Raman bands are observed in the region 1110-1130 and 966-996 cm(-1) and are assigned to the SiO(4) and PO(4) symmetric stretching modes. Other bands are observed in the range 1005-1096 cm(-1) and are attributed to the nu(3) antisymmetric bending modes of PO(4). Some low intensity bands around 874 cm(-1) were discovered and remain unclassified. Bands in the low-wavenumber region are assigned to the nu(4) and nu(2) out-of-plane bending modes of the OSiO and PO(4) units. Raman spectroscopy is a useful tool in determining the vibrational spectroscopy of mixed hydrated multi-anion minerals such as perhamite. Information on such a mineral would be difficult to obtain by other means.  相似文献   

2.
A comparison of deuterated and non-deuterated erythrite has been made using a combination of infrared and Raman spectroscopy. Infrared spectrum shows bands at 3442, 3358, 3194 and 3039 cm(-1). The band at 3442 cm(-1) is attributed to weakly hydrogen bonded water and the band at 3039 cm(-1) to strongly hydrogen bonded water. Deuteration results in the observation of OD bands at 2563, 2407 and 2279 cm(-1). The ratio of these bands change with deuteration. Deuteration shows that the strongly hydrogen bonded water is replaced in preference to the weakly hydrogen bonded water. Three HOH bending modes are observed at 1686, 1633, 1572 and DOD bending modes at 1236, 1203 and 1176 cm(-1). Deuteration causes the loss of intensity of the bands at 841, 710 and 561 cm(-1) and new bands are observed at 692, 648 and 617 cm(-1). These three bands are attributed to the water librational modes. Deuteration results in an additional Raman band at 809 cm(-1) with increasing intensity with extent of deuteration. Deuteration results in the shift of Raman bands to lower wavenumbers.  相似文献   

3.
On the basis of new data on the Raman spectra of solid and gaseous 1,1,2-tribromoethane a revised vibrational assignment is suggested.  相似文献   

4.
The N(CH3)4(+) salt of the cis-IO2F3(2-) anion was synthesized from [N(CH 3)4][IO2F2] and excess [N(CH3)4][F] in CH3CN solvent. The [N(CH3)4] 2[IO2F3] salt was characterized by Raman, infrared, and (19)F solid-state MAS NMR spectroscopy. Geometry optimization and calculation of the vibrational frequencies at the DFT level of theory corroborated the experimental finding that the IO2F3(2-) anion exists as a single isomer with a cis-dioxo and mer-trifluoro arrangement. The fluorine atom in IO2F3(2-) that is trans to one of the oxygen atoms is weakly bound with a calculated bond length of 228.1 pm. The IO2F3(2-) anion is only the second example of an AEO 2F 3 species after XeO2F3(-).  相似文献   

5.
The main parameters for precipitation of mixed carbonate materials have been studied by Raman microscopy. These carbonates are compounds of barium, strontium and calcium. It has been shown that the Raman spectrum of a sample is exclusively controlled by its composition, the precipitation parameters do not affect the crystal structure. Even at relatively low levels, the calcium content of a sample can dominate the vibrational frequencies as measured by Raman spectroscopy. Calcium contents greater than 17% show this effect to a considerable degree, and give the broadest or two Raman peaks and thus the least uniform unit cells. The analysis of the lattice modes demonstrates that each Raman shift observed for a mixed carbonate sample corresponds to a specific crystal structure. Some peaks lie within two or three shifts that are observed for different crystal structures.  相似文献   

6.
Tetraphenylarsonium and tetramethylammonium salts of the complex anions Ph3Sn(N3)?2, Ph3Sn(N3)(NCS)?, Me2Sn(N3)2?4 and Ph2Sn(N3)2(NCS)2?2 have been synthesized, and the solid state configuration of the complex anions has been studied by Mössbauer and vibrational spectroscopies. Trigonal bipyramidal structures are advanced for the Ph3SnIV derivatives, with equatorial SnC3 and apical pseudohalide ligands, while the R2SnIV compounds are assumed to be trans-octahedral species. The NCS? ligands are observed to be N-bonded to SnIV. Conductance and PMR (for the Me2SnIV compound) data suggest the presence of the complex anions also in solution phases.  相似文献   

7.
The infrared and Raman spectra of solid salicylaldoxime have been measured both for normal and deuterated substance; the polarized infrared spectra of oriented samples have been also obtained. Assignments have been proposed on the basis of i.r. dichroism, Raman data and isotopic effects. The possible occurrence of Fermi resonance effects in some regions of the i.r. spectra has been also investigated.  相似文献   

8.
9.
The infrared spectra of 1-methyl and 2-methyl-5-aminotetrazole have been measured from 4000 to 180 cm−1: polarized spectra have also been obtaine  相似文献   

10.
A spectroscopic study of some radio frequency argon-methane, sulphur, phosphorus and halogen plasmas has been carried out. Their analytical utility was assessed and some temperature measurements in the argon-methane plasma carried out. A number of novel lines is reported in the argon-sulphur plasma spectrum and an excitation process in these plasmas discussed.  相似文献   

11.
Application of infrared and Raman spectroscopy together with hydrogen—deuterium substitution has allowed identification of at least seven of the twelve fundamental vibrational frequencies expected for the double bridge CuBH4 unit in tetrahydroboratobis(triphenylphosphine)copper(I). Limited assignments for the AgBH4 unit in the analogous silver compound, based on infrared data for the hydride deuteride, are also presented.  相似文献   

12.
In this work we report the first Raman spectrum of tris triphenyl phosphine rhodium (I) chloride. The vibrational spectrum of triphenyl phosphine has also been recorded under similar conditions to facilitate comparisons. FT-Raman spectra and diffuse reflectance mid-IR and far-IR spectra are presented and a full skeletal vibrational assignment is made.  相似文献   

13.
The results of an electron diffraction reanalysis, augmented with a combined electron diffraction and vibrational spectroscopic elucidation, of the molecular structure of BiCl3 are reported. The principal parameters arer g (Bi-Cl)=2.424±0.005 å (r =2.417±0.005 å) and <Cl-Bi-Cl=97.5±0.2. They are in excellent agreement with previous electron diffraction analysis [1], utilizing a more limited data range from the same experiment. They are also fully consistent with the expected trends of geometrical variation in the Group V trihalide series. The force fields of BiCl3, determined by normal coordinate analysis and by combined analysis, agree within experimental error.  相似文献   

14.
The high-yield synthesis of Dy3N@C80 (I) opens the possibility of characterizing its molecular and vibrational structures. We report on the structure determination of Dy3N@C80 (I) by X-ray crystallographic study of single crystal of Dy3N@C80.Ni(OEP).2C6H6, revealing a nearly planar Dy3N cluster encapsulated in an Ih-C80 cage. The vibrational structure of Dy3N@C80 (I) is studied by Fourier transform infrared (FTIR) and Raman spectroscopy in combination with force-field calculations. A correlation was found between the antisymmetric metal-nitrogen stretching vibration and the structure of the M3N cluster of M3N@C80 (I) (M = Y, Gd, Tb, Dy, Ho, Er, Tm). Moreover, a stronger interaction between the encaged nitride cluster and the C80 carbon cage was found in the class II M3N@C80 (I) (M = Y, Gd, Tb, Dy, Ho, Er, Tm) than in Sc3N@C80 (I). This study demonstrates that the cluster size plays the dominating role in the structure of the M3N cluster in M3N@C80 (I).  相似文献   

15.
The crystal structure of alunite (K0.72,Na0.28) Al3(SO4)2(OH)6 from El Gnater, central Tunisia, is refined by the Rietveld method. Raman and infrared data of this mineral are also given in order to provide some further information about the mineralogy and chemistry of this alunite. The crystal system is trigonal, space group R $ \bar 3 The crystal structure of alunite (K0.72,Na0.28) Al3(SO4)2(OH)6 from El Gnater, central Tunisia, is refined by the Rietveld method. Raman and infrared data of this mineral are also given in order to provide some further information about the mineralogy and chemistry of this alunite. The crystal system is trigonal, space group R m, with a = 6.9834(4) ? and c = 17.0899(11) ?. Final Rietveld refinement converges to R p = 0.16, R wp = 0.16, and R Bragg = 0.07. In the alkalic site, the occupancy of potassium and sodium is refined to 72 and 28%, respectively. The Raman and infrared spectra are investigated in order to improve previous assignments of the observed frequencies, especially for tetrahedral and octahedral vibration and OH group, which are discussed on the basis of unit-cell group analysis and by comparison with previously observed wavenumbers of natrojarosite and synthetic alunite. The text was submitted by the authors in English.  相似文献   

16.
Intramolecular vibrational energy redistribution (IVR) and vibrational predissociation (VP) from the XH stretching vibrations, where X refers to O or C atom, of aromatic molecules and their hydrogen(H)-bonded clusters are investigated by picosecond time-resolved IR-UV pump probe spectroscopy in a supersonic beam. For bare molecules, we mainly focus on IVR of the OH stretch of phenol. We describe the IVR of the OH stretch by a two-step tier model and examine the effect of the anharmonic coupling strength and the density of states on IVR rate and mechanism by using isotope substitution. In the H-bonded clusters of phenol, we show that the relaxation of the OH stretching vibration can be described by a stepwise process and then discuss which process is sensitive to the H-bonding strength. We discuss the difference/similarity of IVR/VP between the "donor" and the "acceptor" sites in phenol-ethylene cluster by exciting the CH stretch vibrations. Finally, we study the vibrational energy transfer in the isolated molecules having the alkyl chain, namely phenylalcanol (PA). In this system, we measure the rate constant of the vibrational energy transfer between the OH stretch and the vibrations of benzene ring which are connected at the both ends of the alkyl chain. This energy transfer can be called "through-bond IVR". We investigate the three factors which are thought to control the energy transfer rate; (1) "OH <--> next CH(2)" coupling, (2) chain length and (3) conformation. We discuss the energy transfer mechanism in PAs by examining these factors.  相似文献   

17.
Raman spectroscopy complimented with supplementary infrared spectroscopy has been used to characterise a synthetic nickel substituted aurichalcite a zinc/nickel hydroxy carbonate, (Zn2+, Cu2+, Ni2+)5(CO3)2(OH)6. XRD patterns show high orientation and indicate the presence of some minor impurities. The diffraction patterns for the Ni-aurichalcite are well correlated with the standard reference patterns. EDAX analyses indicate variations in chemical composition of Zn/Ni ratios of ∼20:1. The symmetry of the carbonate anion in aurichalcite is Cs and is composition dependent. This symmetry reduction results in multiple bands in both the symmetric stretching and bending regions. The intense band for the Ni-aurichalcite at 1070 cm−1 is assigned to the ν1(CO3)2− symmetric stretching mode. Three Raman bands assigned to the ν3(CO3)2− antisymmetric stretching modes are observed for Ni-aurichalcite at 1372, 1480 and 1543 cm−1. Multiple Raman bands are observed in the regions from 800 to 850 cm−1 and 720 to 750 cm−1, and are attributed to ν2 and ν4 bending modes confirming the reduction of the carbonate anion symmetry in the aurichalcite structure. This research proves that nickel containing aurichalcites can be synthesised in the laboratory thus mimicing the natural nickel containing aurichalcites.  相似文献   

18.
The possible stable forms of 3-phenylpropylamine (3-PPA) molecule were experimentally and theoretically studied by infrared and Raman spectroscopy. FT-IR and Raman spectra of 3-PPA were recorded in the regions of 4000–400 cm−1 and 3700–60 cm−1, respectively. The potential energy surface corresponding to the internal rotations of the molecule was investigated by semi-empirical quantum mechanical methods, and appropriate conformers defined with B3LYP hybrid density functional theory method along with the basis sets of different size and type. Results from experimental and theoretical data showed the transtransgauche (TTG) to be the most stable form of a 3-PPA molecule.  相似文献   

19.
The infrared (IR) spectra of Co(HCOO)2 x 2H2O, Ni(HCOO)2 x 2H2O and Cu2Co(Ni)1-x (HCOO)2 x 2H2O mixed crystals (0 < x < or = 0.5) have been recorded and the internal modes of the formate groups and the water molecules are discussed. The analysis of the spectra of the mixed crystals reveals that when copper ions replace cobalt and nickel ions in Co(HCOO) x 2H2O and Ni(HCOO)2 x 2H2O, the Cu2+ ions are localized at the two available positions. However, the occupancy degree of the Me(1) and Me(2) sites by the different cations needs X-ray diffraction (XRD) studies of the single crystals. The new crystal phase Co0.17Cu0.83(HCOO)2 x 2H2O obtained from the Co(HCOO)2 x 2H2O-Cu(HCOO)2 x 2H2O-H2O system at 50 degrees C crystallizes in the monoclinic system with lattice parameters: a = 12.329(4); b = 7.241(2); c = 8.707(5) A and beta = 103.13(3) degrees (SG probably P2/c). The number of the bands corresponding to the uncoupled OD vibrations and the water librations shows that probably more than two water molecules are expected to exist in the structure. Furthermore, it is assumed that the water molecules bonded to the copper ions form stronger hydrogen bonds (stronger Cu-OH2 interaction) than those bonded to the cobalt ions.  相似文献   

20.
Summary From the studies which have been carried out it follows that in Berthollide solid solutions with the tysonite structure there is intensive diffusion of fluoride ions in the vacancies, and this is responsible for the electrical conductivity of the crystals. It follows from an analysis of the NMR data obtained that the fluorine vacancies are produced in the positions corresponding to fluorine atoms of low mobility, confirming structural studies which indicate the preferential formation of vacancies in the FIII positions. At low temperatures, the movement of the highly mobile FI fluorine atoms shows practically no association with the low-mobility FIII positions. Thus in spite of the high concentration of anion vacancies in the specimens studied, the mobility of fluorine at low temperatures does not increase, compared with the ideal tysonite structure. It follows from the temperature dependence of the19F NMR spectra, however, that the presence of vacancies increases the mobility in the subsystems of fluorine atoms with low mobility and increases the exchange between the fluorine atoms with high and low mobilities.L. V. Kirenskii Institute of Physics, Siberian Branch, Academy of Sciences of the USSR. Institute of Crystallography, Academy of Sciences of the USSR. Translated from Zhurnal Strukturnoi Khimii, Vol. 20, No. 4, pp. 622–626, July–August, 1979.  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