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.
Raman spectra of stearyl alcohol (n-C18H37OH) indicate that vibrational bands in both the skeletal optical (1000-1200 cm?1) and longitudinal acoustical (0–500 cm?1) regions are considerably perturbed by intermolecular hydrogen bonding. The zone interior skeletal stretching mode reflecting phonon dispersion in the v4 branch of the infinite polyethylene chain is found at 1105 cm?1, approximately 20 cm?1 higher than in the corresponding n-alkane. Similarly the single nodal longitudinal acoustical mode (LAM-1) is found shifted by 12 cm?1 to higher frequency when the expected “mass effect” produced by the ? OH group is considered. This shift is further increased to 16 cm?1 at ?100°C indicating a further perturbation on this accordion mode due to the increased strength of the hydrogen bond at the low temperatures. The positions of the higher multinodal vibrations, LAM-3 and LAM-5, are also perturbed by the hydrogen bond but by differing amounts. The observation of a low-frequency Raman-active LAM in polytetrahydrofuran [(? CH2CH2CH2CH2O? )n] is discussed in conjunction with the expected effects of hydrogen bonding at the lamellar surface.  相似文献   

2.
Low-frequency Raman spectra were recorded for α-methyl, ω-methoxy-oligo(oxyethylene)s, C1EmC1 with m in the range 4–25 (i.e., 15–78 atoms). Longitudinal acoustical mode (LAM-1) frequencies were identified and compared with those determined previously for α-hydro, ω-hydroxy-oligo (oxyethylene)s. Nonlinear relationships between LAM-1 frequency and reciprocal chain length were explained as an effect of intermolecular end forces.  相似文献   

3.
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   

4.
Low-frequency Raman spectra are reported for monodisperse methylene-oxyethylene-methylene triblock oligomers. Assignments are made to the longitudinal acoustical modes (LAMs) of the oxyethylene block and the methylene block. The results for LAM-1 can be modeled by the vibrations of rods with perturbing forces and masses. The results for LAM-3 are not so readily modeled.  相似文献   

5.
Lead(II) alkanoates with even chain lengths from octanoate to octadecanoate have been investigated by Raman spectroscopy. In the low frequency region, transverse and longitudinal acoustical modes (TAM, LAM) have been assigned. It was shown that LAM-1 is the vibration of the double chain with the node of the vibration in the Pb2+ layer. A fully extended conformation of the chains in the low temperature phase was confirmed. The frequencies and intensities of the LAMs as compared with those of the alkanes and the fatty acids led to an estimate of the force constant and polarizability of the Pb2+ -COO- bond relative to the C-C bond. The defects at the chain ends were investigated in the ρ(CH3) and v(CC) region. For the intermediate (CM) phase, both the Pb2+ layer distance reduction and the chain length independent enthalpy contributions can be attributed mainly to defects at the chain ends.  相似文献   

6.
It is shown how the shape of the longitudinal acoustic vibration observed in the low-frequency region of the Raman spectra of crystalline polymers can be used to obtain a quantitative distribution of lengths of straight-chain segments associated with polymer lamellas. The procedure is demonstrated for a “solid-state” extrudate of polyethylene and for a bulk-crystallized specimen of the same polymer. Equations relating the shape of the LAM-1 band to the shape of the distribution curve are given. The low intensity observed for the LAM-3 mode relative to LAM-1 is explained quantitatively without recourse to end effects. LAM-5 has been observed for bulk-crystallized polyethylene. For the extrudate we find the distribution of lengths of straight-chain segments to have a tail on the long-length side which is not present for the bulk-crystallized sample. The Raman technique is shown to provide new morphological data unattainable at present by other methods.  相似文献   

7.
Low-Frequency Raman spectra of odd α,ω-dibromo- and α,ω-dihydroxy-n-alkanes were recorded. The longitudinal acoustic mode (LAM-1) frequencies were assigned by references to the published results for n-alkanes and even α,ω-disubstituted n-alkanes and also by taking account of the effects of end intermolecular forces and end-group masses by use of the chain model of Minoni and Zerbi.  相似文献   

8.
《Chemical physics》2001,263(2-3):401-414
We have recorded the dispersed fluorescence and the fluorescence excitation spectra of C60 in toluene matrices at 5 K. Upon excitation with the green Ar+ laser line (λ=514 nm) we obtained for the first time in this matrix well resolved visible fluorescence spectra which we have compared with those observed in other low temperature matrices. Our spectra were interpreted and assigned using theoretical assessments of vibronic activities of transitions between the three lowest excited electronic states 1T1g, 1T2g, 1Gg and the totally symmetric ground state, and on the basis of a single 00 level which has pseudo-Jahn–Teller (JT) components of the three near-degenerate excited states. The fluorescence spectra exhibit prominent JT induced hg(1) progressions, Herzberg–Teller-induced hu and other ungerade mode vibrations, including a very active t1u(4) mode. Excitation wavelength independent bands are assigned to the fluorescence of C60 molecules in toluene microcrystals embedded in the toluene glass whereas excitation wavelength dependent features are interpreted as originating from C60 molecules isolated in the toluene glass itself. These interpretations are supported by the results of spectrally selective detected fluorescence excitation spectra.  相似文献   

9.
One of the most important problems in developing devices for quantum computation is the coupling and dissipation of states by thermal noise. We present a study of a two‐state electric dipole in a crystal coupling to noise from a reservoir. As a realization of such an energy‐dissipating dipole, we report and analyze dielectric loss measurements in single crystal and polycrystalline Al2O3 over the temperature range 70–300 K. We are able to model the dielectric loss in terms of a quasi‐classical model that uses the fluctuation–dissipation theorem. Two key parameters in this model are the crystal oscillator energy and reservoir–lattice coupling constant. In polycrystalline samples, it is assumed that the main effect of structural disorder is a modification of the spectrum of the thermal phonons, so that acoustical vibrations acquire some optical mode character. The temperature dependence of the linewidth of the high dielectric strength infrared (IR) mode at 438 cm?1 and the quasi‐degenerate Raman mode of the k = 0 (418 cm?1) transition are also investigated and are shown to be related simply to the dielectric loss. The model reproduces the unusual temperature dependence of the dielectric loss observed experimentally. The implications for the coupling of quantum mechanical objects to noise and quantum information processing are discussed. © 2005 Wiley Periodicals, Inc. Int J Quantum Chem, 2006  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.
Model compounds for imines formed during the thermal curing of short chain polyimides have been synthesized and characterized. These compounds have imine bonds (C?N) formed by the nucleophilic attack of primary amines on imide carbonyls. The C?N stretching mode appears at 1649–1664 cm?1 in the Raman and infrared spectra of these compounds and the band assigned to the carbonyl mode in an imide ring with an imine bond appears near 1740 cm?1. These compounds have been prepared and characterized to verify the conclusions of a previously reported study in which bands observed in thermally cured short chain polyimides at 1656 and 1742 cm?1 were assigned as the C?N and associated C?O modes, respectively. It has also been confirmed that the C?N stretching mode in the imide model compounds is inherently IR weak and can only be seen if the concentration of imine species is high. © 1993 John Wiley & Sons, Inc.  相似文献   

12.
The influence of temperature, frequency, and straight-chain-length distribution on the low-frequency Raman-active longitudinal-acoustical-mode (LAM) bands of polymers is discussed. Specifically, the effect of these factors on the intensity of LAM-1, on the ratio of intensities of LAM-1 to LAM-k, and on peak position is computed or estimated. Band halfwidth and chain-length-distribution halfwidth are found not to be simply related. It is shown that the integrated intensity of LAM-1 is very nearly proportional to the total length of straight chains.  相似文献   

13.
A Raman band of low frequency, arising from an accordionlike vibration of all-trans \documentclass{article}\pagestyle{empty}\begin{document}$ \rlap{‐‐} ({\rm CH}_2 \rlap{‐‐})_n $\end{document} segments and previously observed in normal paraffins and in polyethylene single crystals, has now also been found in bulk and in cold-drawn polyethylene, both linear and branched. The accordionlike vibration, or longitudinal acoustic mode (LAM), in polyethylene is compared with the LAM in normal paraffins. Whereas the Raman bands corresponding to the third (LAM-3) and higher modes are quite intense in a long-chain paraffin such as n-C94H190, they are so weak in polyethylene as to be unobservable with the apparatus used. This is attributed to the presence of the chain fold in polyethylene. Of the two extreme structural models of the fold here discussed, namely the models of “tight folds” and of “loose loops,” only the latter seems capable of accounting for the weakness of LAM-3 and higher modes in polyethylene. A quantity called “nominal Raman length” is defined as the length of that all-trans n-paraffin that would have the same LAM-1 frequency as the polyethylene sample under consideration. The nominal Raman length is always greater than the average long spacing, deduced from discrete x-ray scattering at small angles after applying a Lorentz correction, and, after allowing for chain tilt, is found equal to the segment length between folds. This can be accounted for by both of the models mentioned. As a test of the theory of surface melting the frequency of the accordion vibration of annealed polyethylene single crystals was measured as a function of temperature up to the melting point; no frequency change with temperature was observable. On the basis of the naive idea that there is complete decoupling of the vibrations in the all-trans chain segment from the disordered (molten) surface layer, one would predict that upon surface melting and the concomitant shortening of the all-trans segment, the LAM-1 frequency should increase. A more careful analysis, taking into account the existence of coupling of the LAM to the surface layer, shows that the outcome of this experiment does not necessarily invalidate the idea of surface melting. Bulk polyethylene samples exposed to 60Co γ-radiation for doses up to 100 Mrad show a slight shift of the Raman band to lower frequencies, whereas no such shift was observed upon absorption of a swelling agent. A search, without success, was made for a longitudinal acoustic mode in polypropylene, poly(vinylidene fluoride), nylon 66, and polyoxymethylene.  相似文献   

14.
The IR spectra (900–100 cm?1) of formamide, acetamide, and thioacetamide in the vapour phase are reported. In the three compounds, the same 5 amino-inversional transitions are assigned, and potential functions for the inversion mode are proposed and discussed.  相似文献   

15.
The absorption spectra of 11 solid-state EDA-complexes of unsaturated hydrocarbon donors with TCNE as acceptor have been studied in the FIR. region (40–400 cm?1). The dominant feature in the spectra is the appearance of a new intense absorption in the region 80–100 cm?1 which is assigned to the intra charge-transfer lattice mode. The (idealized) force constant for this lattice mode is found to be roughly constant for most of the EDA-complexes and is thus believed to be a measure of the total interaction of donor and acceptor within the crystal site. Force constants differing in magnitude from those of the majority of complexes are traced back to a particular structure of the donor component and/or the EDA-crystal as a whole. A recently proposed reassignment of the Raman active fundamentals of pure TCNE is questioned in view of the present findings.  相似文献   

16.
The microwave spectrum of propionyl chloride has been investigated in the region 18.0–40.0 GHz, and transitions due to a cis conformer have been assigned. This form has a heavy atom planar configuration and the methyl group and the carbonyl oxygen atom are cis to each other. Using the substitution structures of propionic acid and acetyl chloride as molecular models for the propionyl chloride molecule, good agreement is found between observed and calculateò effective rotational constants. For the 35Cl species satellite spectra assigned to the first four excited states of the C-C torsional mode have been observed together with the first excited state of the methyl torsional mode. The ground state spectrum has also been assigned for the 37Cl species. Relative intensity measurements yielded the lowest C-C torsional vibration frequency of 86 ± 10 cm?1. The CH3 internal rotation frequency was found to be 197 cm?1. Nuclear quadrupole coupling constants were determined for the ground state of the 35Cl and 37Cl species. From observed A-E splittings of bQ-branch transitions of the first excited state of the methyl torsional mode a barrier to internal rotation was estimated to be V3 = 2480 ± 40 cal mol?1 (867 ± 14 cm?1).  相似文献   

17.
Temperature‐dependent 1H and 13C‐NMR spectra of the title compounds are presented. Coalescence effects are discussed and assigned to dynamic process—the interconversion of bicyclic system. The free energies of activation covered the range 39–52 kJ/mol. The dioxepane ring adopts twist‐chair (TC) conformation. GIAO/DFT calculation of isotropic shieldings for the set of low‐energy conformations showed that only one conformer is present at 298 K in solution that matched well with experimental data. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

18.
氯化癸铵的低频拉曼光谱研究   总被引:1,自引:0,他引:1  
在290—340 K的温度范围内考察了氯化癸铵的低频拉曼光谱, 指认了210 cm~(-1)附近的谱带为分子烷基链的纵向声子振动带, 低于100 cm~(-1)的谱带为晶格振动带, 结果表明DeAC晶体存在313 K和321 K两个结构相变。在313 K以下的低温相, DeAC分子以完全有序的全反式链构象存在, 在313—321 K的中间相, 分子烷基链出现了旁式构象, 高温相与中间相相比主要是分子链横向堆积有序度明显降低, 而分子链构象特征无明显差异。  相似文献   

19.
Antimony(III)-butoxo Compounds Tri(tert.-butoxo)stiban 1 reacts with antimony(III) chloride dependent from the molar ratio to yield Sb(OtBu)2Cl 2 and Sb4Cl7O(OtBu)3 3 resp ‥ 3 are discrete molecules with a cage structure. 2 is associated by oxygen bridges. The vibrational spectra of 1–3 are assigned.  相似文献   

20.
Dynamic mechanical analysis, coupled with polarized step-scan FTIR transmission spectroscopy, has been used to monitor the submolecular motional behavior of uniaxially oriented polyamide 11. The dynamic in-phase spectra depend upon the morphology of the samples as well as on the polarization direction of the infrared radiation. The lineshape features of the dynamic in-phase spectra and their relationship to sample deformation are analyzed on the basis of changes of the internal coordinates, the reorientation movement of several functional groups, and the thickness change of the film during the stretching cycle. Dynamic infrared spectra are helpful for deconvolution of overlapping bands on the basis of their different responses to the external perturbation, which sometimes cannot be resolved well by derivative spectroscopy or curve-fitting analysis. The lineshape features have been used to follow microstructural changes after isothermal heat treatment. Near the N H stretching frequency, two bands at 3270 cm−1 and 3200 cm−1 are resolved and analyzed in terms of Fermi resonance between the N H stretching fundamental mode and the overtone and combination modes of the amide I and II vibrations. The dynamic response of the N H stretching mode correlates with the modulation of hydrogen bond strength in uniaxially oriented PA-11. After thermal treatment at the highest temperature (190°C), the dynamic response in this region is mainly caused by the modulation of crystals. In amide I region, three bands at 1680 cm−1, 1648 cm−1, and 1638 cm−1 are separated and assigned to hydrogen bond-free, hydrogen-bonded amorphous, and hydrogen-bonded crystalline regions, respectively. The dynamic responses of the hydrogen-bonded regions are more sensitive to external perturbation. Two components are found in the amide II region, and the band at 3080 cm−1 is assigned to the overtone resonance of the component with perpendicular polarization. © 1998 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 36: 2895–2904, 1998  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