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 31 毫秒
1.
We have applied laser Raman scattering to the measurement of the density of UF6 vapor. A linear relation between scattering intensity and density is shown. We have also measured the absolute Raman scattering cross section of the 665 cm−1 line of UF6 by the substitution technique and obtained a value of 1.02±0.09×10−29 cm2/sterad at 488 nm. Measurements of the line position and the depolarization ratio of this line are also reported. Work supported by the U. S. Energy Research and Development Administration.  相似文献   

2.
Coherent Stokes and anti-Stokes Raman scattering are used to study the ν1 and ν2 spectral band profiles of UF6 and SF6. Most of the observed SF6 “hot” bands are assigned, leading to evaluations of the anharmonicity constants Xij: X12 = ?(2.80 ± 0.30) cm?1, X14 = ?(1.00 ± 0.15) cm?1, X15 = ?(1.00 ± 0.15) cm?1. For UF6, a tentative assignment of the “hot” bands is made: X12 = ?(1.80 ± 0.30) cm?1, X13 = ?(1.60 ± 0.30) cm?1, X14 = ?(0.20 ± 0.10) cm?1, X15 = ?(0.25 ± 0.10) cm?1, and X16 = ?(0.10 ± 0.05) cm?1. Parameters such as the vibration-rotation coupling constants are determined. For SF6: α = (7 ± 2) × 10?5 cm?1 for the ν2 band and α = ?(1.02 ± 0.01) 10?4 cm?1 for the ν1 band. The calculated spectral profiles of the coherent Stokes or anti-Stokes spectra, which are in good agreement with experimental results, give values for the resonant and nonresonant parts of the susceptibility in both molecules. They also show, in some cases, the influence of neighboring combination bands.  相似文献   

3.
Using a cw argon-ion laser, we have measured the spontaneous Raman scattering cross section, σR, and linewidth in atomic thallium vapor. For 4880-Å excitation, σR = 1.6 × 10?27 cm2. We observe no pressure broadening of the Raman line at vapor pressures to 100 torr.  相似文献   

4.
Raman scattering experiments on K2Pt(CN)4Br0.3 · 3H2O are reported between 5 and 300 K as a function of temperature. A line of A1 symmetry detected at 44 cm?1 shows interesting temperature dependent properties. It is concluded from a comparison of the frequency, symmetry, and scattering intensity of this line with theoretical predictions that the excitation concerned represents the amplitude mode of the charge density wave (the line observed in infrared absorption being the phase mode). No Peierls transition is observed, but the results are consistent with a Peierls distortion present at all temperatures. The findings are correlated with inelastic neutron scattering and infrared studies. Finally, the CN stretching modes at 2189 and 2173 cm?1 and the water mode at 3490 cm?1 are studied as a function of temperature.  相似文献   

5.
In this work, the Raman scattering of melamine was studied under high pressure up to 60 GPa. The behavior of the most intensive peaks of the Raman spectrum of melamine, 677 cm?1 and 985 cm?1 modes, and their line widths do not show any phase transition or indication of formation of sp 3 bonds. Comparing the behavior of the line width of the Raman peaks of graphite under pressure and that of melamine leads us to conclude that the s-triasine (C–N) ring is more rigid than the C–C graphite ring. High pressure results with melamine suggest that the direct phase transition g-C3N4 to dense C3N4 phase should occur above 60 GPa.  相似文献   

6.
The Raman and infrared active long wavelength phonons of a GaS single crystal were studied at different temperatures in the 10–600 cm?1 range. Properly polarized Raman spectra could be obtained with the 4880 Å exciting line and the previous assignment of the E1g modes controversed recently could be confirmed. Infrared spectra were recorded in the 30–600 cm?1 region. The vibrational frequencies of the crystal were also calculated using a method developed by Wieting and six new frequencies corresponding to infrared and Raman inactive modes have been proposed.We have observed that the degree of leakage of scattered intensity in unallowed polarizations increases when the wavelength of the exciting line moves off the exciton absorption front. The phonon at 74 cm?1 was particularly sensitive and the question of the antiresonant behaviour of this compound is raised.  相似文献   

7.
Pressure broadening of the R(6) manifold in ν3 of UF6 at 628.32 cm?1 has been measured at pressures of 0–30 torr of Ar in a long-path cell at 199 ± 1 K. We obtain a pressure-broadening coefficient of 9.0 ± 1.5 MHz/torr (FWHM), which corresponds to 7.3 ± 1.2 MHz/torr at 300 K. The UF6-Ar optical collision diameter is 7.4 ± 0.6 Å; this is larger than the hard-sphere kinetic theory collision diameter of 5.2 Å obtained from the diffusion coefficient and implies rapid rotational relaxation.  相似文献   

8.
Low-temperature experiments of Raman scattering and heat capacity have been performed in a B2O3 glass, pressure quenched from 1200 °C in order to obtain the density as largest as possible (ρ = 2373 kg/m3). When compared to those of compacted B2O3 glasses having smaller density, the Raman spectrum of this glass exhibits a strong decrease of the intensities of the Boson peak and the band at 808 cm?1, both the features being determined by the decrease of the boroxol ring population. Moreover, the Boson peak exhibits a large shift to 68 cm?1 (from 26 cm?1 observed in normal vitreous B2O3). The high atomic packing of the glassy network also leads to a marked decrease of the excess heat capacity over the Debye T3-behaviour characterizing the crystal. The density g(ν) of low-frequency vibrational states has been assessed by using the low-frequency Raman intensity to determine the temperature dependence of the low-temperature heat capacity. The observations performed over a wide range of glass densities are compared to the predictions of theoretical models and computer simulations explaining the nature of the Boson peak. Consistency with the results of a simulation study concerning the vibrations of jammed particles leads to evaluate a nanometre length scale which suggests the existence of poorly packed domains formed from several connected boroxols. These soft regions are believed to be the main source of low-frequency optic-like vibrations giving rise to the Boson peak.  相似文献   

9.
ABSTRACT

The mixed metal fluorides containing alkali metals have a range of important applications in optical and electronic devices. Raman spectrums of two such fluorides were examined. Raman spectrum of KCuF3 at 300 K exhibited bands at 261, 295, 363, 468, 519, and 549 cm?1, indicating site symmetry (orthorhombic) lower than the tetragonal symmetry as observed from the powder X-ray diffraction pattern. Cubic KNiF3 showed bands at 410, 468, and 657 cm?1. The first two bands were attributed to the second-order phonon scattering, and the band at 657 cm?1 was assigned to two-magnon peak.  相似文献   

10.
This paper reports on the use of phonon spectra obtained with laser Raman spectroscopy for the uncertainty concerned to the optical phonon modes in pure and composite ZnO1?x (Cr2O3) x . Particularly, in previous literature, the two modes at 514 and 640 cm?1 have been assigned to ZnO are not found for pure ZnO in our present study. The systems investigated for the typical behavior of phonon modes with 442 nm as excitation wavelength are the representative semiconductor (ZnO)1?x (Cr2O3) x (x = 0, 5, 10 and 15 %). Room temperature Raman spectroscopy has been demonstrated polycrystalline wurtzite structure of ZnO with no structural transition from wurtzite to cubic with Cr2O3. The incorporation of Cr3+ at most likely on the Zn sub-lattice sites is confirmed. The uncertainty of complex phonon bands is explained by disorder-activated Raman scattering due to the relaxation of Raman selection rules produced by the breakdown of translational symmetry of the crystal lattice and dopant material. The energy of the E 2 (high) peak located at energy 53.90 meV (435 cm?1) due to phonon–phonon anharmonic interaction increases to 54.55 meV (441 cm?1). A clear picture of the dopant-induced phonon modes along with the B 1 silent mode of ZnO is presented and has been explained explicitly. Moreover, anharmonic line width and effect of dislocation density on these phonon modes have also been illustrated for the system. The study will have a significant impact on the application where thermal conductivity and electrical properties of the materials are more pronounced.  相似文献   

11.
The linear absorption of CO2 laser radiation in SF6, WF6, and UF6 has been measured by using optoacoustic detection techniques. Absolute absorption coefficients per Torr as low as 1 × 10?7 cm?1 Torr?1 in a 2-cm active path length could be measured by taking advantage of calibration measurements performed with SF6.  相似文献   

12.
Gilalite is a copper silicate mineral with a general formula of Cu5Si6O17 · 7H2O. The mineral is often found in association with another copper silicate mineral, apachite, Cu9Si10O29 · 11H2O. Raman and infrared spectroscopy have been used to characterize the molecular structure of gilalite. The structure of the mineral shows disorder, which is reflected in the difficulty of obtaining quality Raman spectra. Raman spectroscopy clearly shows the absence of OH units in the gilalite structure. Intense Raman bands are observed at 1066, 1083, and 1160 cm?1.

The Raman band at 853 cm?1 is assigned to the –SiO3 symmetrical stretching vibration and the low-intensity Raman bands at 914, 953, and 964 cm?1 may be ascribed to the antisymmetric SiO stretching vibrations. An intense Raman band at 673 cm?1 with a shoulder at 663 cm?1 is assigned to the ν4 Si-O-Si bending modes. Raman spectroscopy complemented with infrared spectroscopy enabled a better understanding of the molecular structure of gilalite.  相似文献   

13.
The time-dependent change in the concentration of UF6 monomers populated in the ground state was monitored in a supersonic Laval nozzle flow with an infrared diode-laser spectrometer in which the frequency of the laser beam was fixed at the = 1 0 transition (627.7 cm–1) of the 3 vibrational mode of the238UF6 monomer. The concentration of UF6 monomers in the ground state increased immediately after a single shot from a Raman laser tuned to a vibrational mode of UF6 clusters (614.8 cm–1) was applied to the gas in the nozzle. Subsequently, this concentration leveled off and slowly returned to the previous level. These results indicate that the population of UF6 monomers in the ground state increased as a result of the predissociation of UF6 clusters vibrationally excited with Raman laser radiation. It is demonstrated that one can utilize this procedure for vibrational predissociation of UF6 clusters as a technique to increase the concentration of UF6 monomers in an irradiation zone for molecular laser isotope separation of uranium.  相似文献   

14.
The Raman spectrum of uranium tetrafluoride (UF4) is unambiguously characterized with multiple Raman excitation laser sources for the first time. Across different laser excitation wavelengths, UF4 demonstrates 16 distinct Raman bands within the 50–400 cm−1 region. The observed Raman bands are representative of various F–F vibrational modes. UF4 also shows intense fluorescent bands in the 325–750 nm spectral region. Comparison of the UF4 spectrum with the ZrF4 spectrum, its crystalline analog, demonstrates a similar Raman band structure consistent with group theory predictions for expected Raman bands. Additionally, a demonstration of combined scanning electron microscopy and in situ Raman spectroscopy microanalytical measurements of UF4 particulates shows that despite the inherent weak intensity of Raman bands, identification and characterization are possible for micron‐sized particulates with modern instrumentation. The published well‐characterized UF4 spectrum is extremely relevant to nuclear materials and nuclear safeguard applications. Published 2016. This article is a U.S. Government work and is in the public domain in the USA. Journal of Raman Spectroscopy published by John Wiley & Sons Ltd.  相似文献   

15.
We have studied ion mobility in a Li0.03Na0.97Ta0.4Nb0.6O3 solid solution by its Raman spectra. It has been revealed that, as the temperature of the solution is increased to approach the point of the phase transition to a state with a high conductivity with respect to lithium, the lines with frequencies at 77, 118, and 142 cm?1, which refer, respectively, to librations of oxygen octahedra Nb(Ta)O6 as a whole and vibrations of Li and Na ions in octahedra, considerably broaden, decrease in intensity, and smear into the wing of the Rayleigh line. Remaining lines are preserved in the spectrum. We have observed that the width of the line with a frequency of 118 cm?1 depends exponentially on temperature, while the width of the line with a frequency of 142 cm?1 changes linearly with it, which makes it possible to attribute to the line with the frequency of 118 cm?1 to vibrations of Li+ cations, whereas the line with the frequency of 142 cm?1 should be attributed to vibrations of Na+ cations in AO12 cuboctahedra. The average lifetime of Li+ ions in equilibrium positions and the jump barrier have been estimated to be ~8 × 10?12 s and ~20 kJ/mol, respectively. This agrees well with the data in the literature on measurements of electric conductivity.  相似文献   

16.
The scattering cross section of the Raman-active phonons at 156 cm?1 (Eg) and 169 cm?1 (F2g) in the ferromagnetic semiconductor CdCr2Se4 (Tc=130 K) has been measured as a function of incident photon energy between 1.55 and 2.81 eV, both in the ferromagnetic and paramagnetic phases. The resonance curve peaks sharply near 2 eV and shows a broadening for temperatures below the Curie point. The relative line intensities change significantly with photon energy. The results show that the concept of spin-dependent Raman scattering in the ferromagnetic spinels has to be revised in terms of exchange-splitting-induced resonant Raman scattering.  相似文献   

17.
Raman scattering is performed to access phase stability in the boron-implanted Hg0.7Cd0.3Te with fluences ranging from 1 × 1012 to 1 × 1015 cm?2. Threshold fluence for the formation of an amorphous phase is invoked here using Thomas–Fermi statistical model. Asymmetric broadening and red shift of the Raman active HgTe-like LO phonon mode are observed with varying fluencies. Electrical properties such as sheet carrier concentration and mobility are also changed with the fluence and reach their saturated values beyond threshold fluence of 5 × 1013 cm?2. Threshold fluence for the formation of amorphous phase is also validated by the Raman measurements and electrical transport properties in the implanted layers. The excess free energy of 6.8 kJ/mole is found corresponding to the threshold fluence for phase transition.  相似文献   

18.
Vibronic spectra are reported for lead sulfide in argon, krypton and SF6 matrices at low temperatures. Emission stimulated by laser line irradiation of PbS is observed from the v′ = 0 level of three electronic states lying at about 14 500, 18 500 and 21 500 cm?1 above the ground state. Emission is also observed from an excited state of Pb2S2 at about 17 000 cm?1. In addition, the laser radiation gives rise to the vibrational Raman spectrum of PbS in argon at 423.2 cm?1 and to a very weak Raman band at 297 ± 2 cm?1 which we attribute to Pb2S2.The effects of temperature on the matrix spectra, of matrix material on the band origins, and of matrix concentration on the vibrational relaxation process, and the apparent degrees of coupling among the electronic states have all been examined. The electronic absorption spectrum of PbS in Ar is reported and the matrix data are compared with available information on gaseous PbS.  相似文献   

19.
The Raman spectra of some selected hydrocarbons and freon gases have been measured and are reported here. The observed peak intensities of the strongest line for each hydrocarbon molecule relative to that of the 2331 cm?1 line of molecular nitrogen range from 14·5 for benzene to 2·6 for 2-pentene. The Raman cross sections of the C-H stretching bands relative to the 2331 cm?1 nitrogen line range from 46 for hexane to 1·2 for acetylene.  相似文献   

20.
The absolute Raman scattering cross section (σRS) for the 1584‐cm−1 band of benzenethiol at 897 nm (1.383 eV) has been measured to be 8.9 ± 1.8 × 10−30 cm2 using a 785‐nm pump laser. A temperature‐controlled, small‐cavity blackbody source was used to calibrate the signal output of the Raman spectrometer. We also measured the absolute surface‐enhanced Raman scattering cross section (σSERS) of benzenethiol adsorbed onto a silver‐coated, femtosecond laser‐nanostructured substrate. Using the measured values of 8.9 ± 1.8 × 10−30 and 6.6 ± 1.3 × 10−24 cm2 for σRS and σSERS respectively, we calculate an average cross‐section enhancement factor (EF) of 0.8 ± 0.3 × 106. Copyright © 2009 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