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 656 毫秒
1.
Raman spectra of the polycrystalline l-alanine analogs CH3CH(NH+3)COO?, CH3CH(ND+3)-COO?, CD3CD(NH+3)COO?, and CD3CD(ND+3)COO? have been obtained. A normal coordinate analysis is carried out based on the experimental frequencies of the four isotopic analogs and a 34 parameter valence-type force field defined in terms of local symmetry coordinates. The final refinement, in which five stretching force constants are constrained to fixed values obtained from bond length data, results in an average error of 7 cm?1 (0.9%) for the observed frequencies of the four isotopically substituted molecules. Band assignments are given in terms of the potential energy distribution for local symmetry coordinates. For non-deuterated l-alanine, the vibrations above 1420 cm?1 and below 950 cm?1 may be described as localized group vibrations. By contrast, the eight modes in the middle frequency range, viz. the three skeletal stretching, the COO? symmetric stretching, one NH+3 rocking, the symmetric CH3 deformation, and the two methyne CH deformation vibrations, are very strongly coupled to one another. Some decoupling appears to take place in the perdeutero molecule, and all but five modes can be described as localized group vibrations.  相似文献   

2.
Infrared spectra of gaseous and matrix isolated trans-CH3N:CHCH3, CD3N:CHCH3 and CHN:CDCD3 have been recorded from 4000 to 200 cm?1 under medium resolution. Nearly complete assignments have been established based upon rotational contours, isotope shifts and group frequency correlations. A number of Fermi resonances have been detected mainly related to CH(D) stretching fundamentals. ln particular, the light N-methyl group gives rise to a complicated Fermi multiplet in the CH-stretching region. An examination of Q sub-band structures resulted in the identification of a number of Coriolis resonances, which in most cases involve pairs of nearly coinciding ‘E-type“ methyl deformation fundamentals. The rotational fine structure of three bands has been analysed by means of combination differences providing estimates for effective rotational constants A(u)-12 [B(u)+C(u)]. From a normal coordinate analy a 30 parameter valence force field has been obtained, which allows an approximation of 85 experimental frequencies with an RMS deviation of 7.4 cm?1.  相似文献   

3.
The products of UV photolysis of ternary Ar?CH4(CD4)?F2 mixtures (1:c:c 0,c, c 0=0.001–0.01) at 13–16 K were identified by ESR and FTIR spectroscopy. These products are?CH3 (?CD3) radicals of typesI andII and molecular CH3F?HF complexes. The latter were characterized by the IR bands of the stretching C?F (1003 cm?1) and H?F (3774 cm?1) vibrations. The ESR spectra of radicalsI are asymmetric. The anisotropy of theg-factor (Δg~10?3) of radicalI indicates that the structure of the radicals is nonplanar. The ESR spectrum of the typeII radical is identical to that of matrix-isolated?CH3 (?CD3) radicals with the planar structure (Δg<5·10?5). Under the experimental conditions, the amount of complexes formed in the photolysis is equal to 0.022·c. When the photolysis is ceased, radicalI disappears after ≈103 s and radicalII is stabilized. The limiting concentrations of the stabilized?CH3 and?CD3 radicals are equal to 2·10?2·c and 2·10?3·c, respectively. A mechanism of the formation of the products is suggested. It is based on the assumption that both matrix-isolated CH4 and F2 and their heterodimers CH4?F2 are present in the samples and it takes into account the long-range migration of translationally excited flourine atoms. The CH3F?HF complexes and radicalsI are generated by the photolysis of the CH4?F2 heterodimers. The decay of radicalsI is caused by geminate recombination of proximate F...CH3 pairs. RadicalsII are formed in the reaction of translationally excited fluorine atoms with isolated CH4 (CD4) molecules.  相似文献   

4.
Vibrational spectra in the range 200–3000 cm?1 are reported and assigned for the species (CH3)2PH, (CH3)2PD, (CD3)2PH, (CD3)2PD, CH3CD3PH and CH3CD3PD. The spectra in the range 1020–500 cm?1 are complicated due to the coupling between δPH, ?Me and the skeletal modes of the molecule. Interpretation is only possible through a force field which is markedly different from an earlier one of dimethyl sulphide. This force field predicts uncoupled δPH frequencies of 835 (a) and 909 cm?1 (a), couples PH bending largely to out-of-skeletal plane methyl rocking (?i) and includes a low p¦¦(a) bending constant, a high skeletal bending constant and unusual signs for two interaction constants. In the crystalline phase at 78 K, the two methyl groups are non-equivalent.  相似文献   

5.
Force constants for the internal vibrations involving the metal and for the lattice vibrations of Hg(CH3)X and Hg(CD3)X (X = Cl, Br or I) are calculated on the basis of a D4h7 layer structure. The internal HgX stretching force constants are much lower than for these molecules in solution, but HgC stretching force constants are slightly higher. The HgX and longitudinal translatory force constants within the lattice layer are close in value to the strong and weak HgX bond stretching force constants respectively in the unsymmetrical [Hg(CH3)X2]? complex ions.  相似文献   

6.
Solid complexes of boron decachloro-o-carborane and boron decachloro-m-carborane (B10Cl10C2H2) with some oxygen and nitrogen bases have been investigated by infrared and Raman spectroscopy. Complexes containing CH · · O hydrogen bonds are characterized by a relative CH stretching frequency shift up to 12% and a halfwidth of the νCH band up to 220 cm?1. CH · · N hydrogen bonds, with trimethylamine for example, are stronger with a relative shift of about 18% and ν12 of about 500 cm?1. Triethylamine complexes, however, form a NH+ · · C? proton transfer hydrogen bond while pyridine can give either CH · · N or C? · · +HN hydrogen-bonded adducts depending on the solvent and temperature. The CH · · N?c? · · +HN equilibrium appears to be shifted towards ion-pair formation at considerably smaller enthalpy values compared to the OH · · N?O? · · +HN system. CH and NH stretching frequencies are correlated with the acidity of the donor and the basicity of acceptor molecules.  相似文献   

7.
Seven R-di-(α-pyridyl) hydroperchlorates (R = (1) CH2, (2) NH, (3) CO, (4) (CH2)2, (5) (CH2)3, (6) (CH2)4 and (7) S-S) were prepared and studied in acetonitrile-d3 solutions by NMR and IR spectroscopy. With the hydroperchlorates of compounds 1 and 4, an equilibrium between non-hydrogen-bonded NH+ groups and intramolecular-bonded NH+ groups is present. With compounds 2, 3 and 5–7, the intramolecular hydrogen bonds are formed quantitatively. In compounds 4–7, the potential wells in these intramolecular structurally symmetrical N+H· N ? N · H+N bonds, are double minima. These hydrogen bonds are easily polarizable. With compounds 1–3, the distance between the N atoms given by the steric conditions of the molecules is smaller than with usual linear hydrogen bonds. Therefore, strong bent intramolecular structurally symmetrical hydrogen bonds are found, with relatively narrow single-minimum potential wells. These bonds cause a band in the region 3000–2500 cm?1 instead of the continuum. Thus they are not easily polarizable.  相似文献   

8.
The infrared and Raman spectra of crystalline CH3CONHCH3, CD3CONHCH3, CH3CONHCD3, CD3CONHCD3 and their N deuterated derivatives have been investigated at various temperatures in the 3400-2000 cm1&#x0304; range. A new interpretation of the NH (ND) stretching subbands is given and is based on the assumption of a strong coupling between NH stretching and deformation modes (γ NH and τ CN). Suitable potential functions are derived from experimental data and they allow to calculate the infrared and Raman spectra in the stretching region. The calculated spectra as well as the calculated isotopic and temperature shifts are in good agreement with the observed ones. The potential surface of the bending modes in the excited state has only one rather flat central minimum corresponding to a planar geometry of the amide group which can be contrasted to the four-minimum potential surface of the fundamental state.  相似文献   

9.
The infrared spectra (3200-50 cm?1) of gaseous and solid CH3NCS and CD3NCS and the Raman spectra (3200-10 cm?1) of the liquids and solids have been recorded. The spectra have been interpreted on the basis of a “pseudo-symmetric top” with C3v symmetry. An assignment of the fundamental vibrations in both molecules, based on their infrared band contours, depolarization values and group frequencies, is given and discussed. Particularly interesting is the low-frequency region where band maxima were observed at 152 and 80 cm?1 for CH3NCS and 139 and 71 cm?1 for CD3NCS in the infrared spectra of the gases. A normal coordinate analysis has also been carried out based on C3v symmetry. Considerable mixing was found between the CαN stretch and NCS symmetric stretch in both isotopic species. The other normal modes in CH3NCS are reasonably pure but, for the CD3NCS molecule, considerable mixing was found between the CD3 stretches and NCS antisymmetric stretch. The proposed vibrational assignment and the results of the normal coordinate calculations are discussed and compared with the results obtained for similar molecules.  相似文献   

10.
The ion–molecule reactions of CH3NH2+, (CH3)2NH+, and (CH3)3N+ with the respective amines have been investigated at thermal kinetic energies in a high-pressure photoionization mass spectrometer at several wavelengths (energies) in the vacuum ultraviolet. The absolute rate coefficient for proton transfer from (CH3)3N+ to (CH3)3N decreases from 8.2 × 10?10 cm3/molecule · sec at 147.0 nm (8.4 eV) to 4.9 × 10?10 cm3/molecule. sec at 106.7-104.8 nm (11.7 eV). In dimethylamine, the rate coefficient decreases from 11.6 × 10?10 cm3/molecular. sec at 8 4 eV to 10.2 × 10?10 cm3/molecule osec at 11.7 eV, while no significant effect of energy was detected in methylamine. The reactions of several fragment ions are also reported. Experiments were also carried out at pressures up to 0.5 torr in order to investigate the further solvation of CH3NH2+, (CH3)2NH2+, and (CH3)3NH+. It was found that the maximum proton solvation numbers in methyl-, dimethyl-, and trimethyl-amine are 4, 3, and 2, respectively, under these conditions.  相似文献   

11.
A 1:1 reaction of [HO(CH2)3]3P with 4-hydroxy-3-methoxy-cinnamaldehyde (coniferaldehyde) or 3,5-dimethoxy-4-hydroxycinnamaldehyde (sinapaldehyde) in acetone at room temperature affords phosphonium zwitterions of the type R3P+CH(4-O?-Ar)CH2CHO; other phosphines [R = Et, n-Bu, (CH2)2CN, and p-Tol] do not react under the same conditions. In alcohols R??OH(D) [R?? = CD3, Et, (CD3)2CD, s-Bu, HOCH2CH2], the above phosphines (except the cyano-derivative) and those where R = i-Pr, Cy, Me2Ph, MePh2 do react within an equilibrium established between the reactants and the zwitterion-hemiacetal products R3P+CH(4-O?-Ar)CH2CH(OH)(OR??) that are formed as a mixture of two diastereomers. The nature of the phosphine and the alcohol affects the equilibrium and the diastereomeric ratio.  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.
A very recent laser ablation‐molecular beam experiment shows that an Al+ ion can react with a single methylamine (MA, CH3NH2) or dimethylamine (DMA, (CH3)2NH) molecule to form a 1:1 ion–molecule complex Al+[CH3NH2] or Al+[(CH3)2NH)], whereas a dehydrogenated complex ion Cu+[CH3N] or Cu+[C2H5N] is detected, respectively, in the similar reaction for a Cu+ ion. Here, we show a comparative density functional theory study for the reactivities of the Al+ and Cu+ ions toward MA and DMA to reveal the intrinsic mechanism. It is found that the interactions of the Al+ ion with MA and DMA are mostly electrostatic, leading to the direct ion–molecule complexes, Al+? NH2CH3 and Al+? NH( CH3)2, in contrast to the non‐negligible covalent character in the corresponding Cu+‐containing complexes, Cu+? NH2CH3 and Cu+? NH( CH3)2. The general dehydrogenation mechanism for MA and DMA promoted by the Cu+ ion has been shown, and the preponderant structures contributing to the mass spectra of the product ions Cu+[CH3N] and Cu+[C2H5N] are rationalized as Cu+? NHCH2 and Cu+? N( CH2)( CH3). The presumed dehydrogenation reactions are also discussed for the Al+‐containing systems. However, the involved barriers are found to be too high to be overcome at low energy conditions. These results have rationalized all the experimental observations well. © 2009 Wiley Periodicals, Inc. Int J Quantum Chem, 2010  相似文献   

14.
Infrared photodissociation spectra of (CH3NH2) n clusters were measured fromn=2 ton=6 near the monomer absorption of the C-N stretching mode at 1044 cm?1 using a cw-CO2 laser. The clusters were size-selected by scattering from a helium beam. The spectrum of cold dimers shows a red (1038 cm?1) and a blue (1048 cm?1) shifted peak which is attributed to the non-equivalent position of the C-N in the open dimer structure. The larger clusters exhibit only one peak between 1045.4 cm?1 and 1046.0 cm?1 caused by the equivalent position of the C-N in the cyclic structures of the larger clusters. Structure calculations confirm these results. Secondly, the mixed complexes C2H4-CH3COCH3 and C2H4-(CH3COCH3)2 were investigated. The dimer spectrum, measured around the monomer frequency of the out-of-plane bending mode of C2H4 at 949 cm?1, shows two peaks at 946.2 cm?1 and 961.3 cm?1. This splitting is attributed to two different isomers that are found in configuration calculations. A similar behaviour is found for the trimer.  相似文献   

15.
Silica from leached chrysotile fibers (SILO) was silanized with trialkoxyaminosilanes to yield inorganic–organic hybrids designated SILx (x=1–3). The greatest amounts of the immobilized agents were quantified as 2.14, 1.90, and 2.18 mmol g−1 on SIL1, SIL2, and SIL3 for –(CH2)3NH2,–(CH2)3NH(CH2)2NH2, and –(CH2)3NH(CH2)2NH(CH2)2NH2 groups attached to the inorganic support. The infrared spectra for all modified silicas showed the absence of the Si–OH deformation mode, originally found at 950 cm−1, and the appearance of asymmetric and symmetric C–H stretching bands at 2950 and 2840 cm−1. Other important bands associated with the organic moieties were assigned to νas(NH) at 3478 and νsym(NH) at 3418 cm−1. The NMR spectrum of the solid precursor material suggested two different kinds of silicon atoms: silanol and siloxane groups, between −90 and 110 ppm; however, additional species of silicon that contain the organic moieties bonded to silicon at −58 and −66 ppm appeared after chemical modification. These modified silicas showed a high adsorption capacity for cobalt and copper cations in aqueous solution, in contrast to the original SILO matrix, confirming the unequivocal anchoring of silylating agents on the silica surface.  相似文献   

16.
The far IR (450–480 cm?1) and Raman (3200–3230 cm?1) spectra of (CH3)3 NGaCl3 have been recorded in the solid state and interpreted in detail on the basis of C3 molecular symmetry. A modified valence force field model is used to calculate the frequencies and potential energy distribution of the adduct. The calculated force constants of the adduct are compared with those previously reported for the free Lewis acid and the free Lewis base moieties, and the observed differences ascribed to geometrical changes of the uncomplexed species on adduct formation and explained on the basis of the VSEPR model and non-bond interactions. Extensive coupling is observed between the GaN stretching mode and the NC3 symmetric stretching and the NC3 symmetric deformational modes. Strong coupling interaction is also found between the GaCl3 antisymmetric stretch and the NC3 antisymmetric deformation. The calculated value of 2.50 mdyn Å?1 for the GaN stretching force constant in (CH3)3NGaCl3 is larger than any of those previously determined in complexes such as (CH3)3NGaH3 (2.43 mdyn Å?1), (CH3)3NGa(CH3)3 (1.61 mdyn Å?1), and H3NGa(CH3)3 (1.08 mdyn Å?1). The observed variations in the magnitudes of the stretching force constants of the donor—acceptor dative bond is found to be consistent with the estimated relative stabilities of this series of adducts.  相似文献   

17.
Alkylation of 8-dimethylamino-1-methylamino-4-nitronaphthalene in the CD3I/KOH/DMSO system afforded a 4-nitro derivative containing the N(CD3)Me group in position 1. Direct proof of the structure of the intramolecular hydrogen bond in solutions of monoprotonated 4-R-1,8-bis(dimethylamino)naphthalenes was obtained for the first time. 1H NMR study revealed that the chelated NH proton is shifted to the N(8) atom for R = NO2 and NH3 + and to the N(1) atom for R = NH2. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 1, pp. 159–162, January, 2006.  相似文献   

18.
Unequal intensities of the Λ-doublet components were observed in the CH(A2Δ-X2Π) emission following the multiphoton dissociation of (CH3)2CO, (CH3)2S and CH3NO2 by an ArF laser (193 nm). The power dependence of the emission intensity was estimated to be cubic (3.1±0.2) when the laser power was below ≈ 8×1017 photons cm?2 pulse?1. The Λ-doublet populations depended on the rotational quantum number N′ and the preferred level changed at N′ = 20. A similar behavior was observed for the CD(A2Δ) from (CD3)2CO. Rotational distributions show bimodal behavior, having a hump around N′ = 13 in CH(A2Δ) and N′ = 11 in CD(A2Δ). These trends indicate that the CH(A2Δ) is produced through multiple processes where stepwise mechanisms are operative via either CH2 or CH3, or both radicals as intermediates.  相似文献   

19.
The potential of near infrared (NIR) spectroscopy in characterization of organically modified clay minerals is introduced. Selected organo-clays, possibly perspective fillers in clay polymer nanocomposites, were prepared from Na-montmorillonite and different surfactants containing octylammonium chain(s), hexadecylammonium chain(s) or a benzene ring with or without a reactive double bond. Based on the stretching (ν) and bending (δ) vibrations observed in the middle IR (MIR) region, the first overtone (2νXH) and combination (ν + δ)XH modes of XH groups (X = O, C, N) are identified. The effect of larger alkylammonium cations on the vibrations of Si-O and OH bonds in montmorillonite layers is observed. The changes in the intensity of the (ν + δ)H2O band near 5250 cm−1 allows for comparison of the amount of water adsorbed on the montmorillonite surface. The water content decreases with the size of the organic cation reflecting increasing hydrophobicity of the montmorillonite surface. The NIR region shows the 2νCH3 and 2νCH2 bands in the 5900-5500 cm−1 region, an upward shift is observed for the complex band due to 2νCH(Ar) of aromatic benzene ring. The NIR spectra are extremely useful in identification of NH2+, NH+ and vinyl groups, which are difficult to recognize in the MIR spectra of organo-clays due to overlapping with other absorption bands. The intense bands corresponding to overtones and combination vibrations of NH3+ and NH2+ groups are found in the 6600-6050 cm−1 and 5000-4600 cm−1 regions, the (ν + δ)NH+ is unambiguously identified near 4750 cm−1. The characteristic band assigned to 2νCH2 in H2CC is detected near 6130 cm−1.  相似文献   

20.
Quantum calculations at the MP2/aug‐cc‐pVDZ level are used to analyze the SH···N H‐bond in complexes pairing H2S and SH radical with NH3, N(CH3)3, NH2NH2, and NH2N(CH3)2. Complexes form nearly linear H‐bonds in which the S? H covalent bond elongates and shifts its stretching frequency to the red. Binding energies vary from 14 kJ/mol for acceptor NH3 to a maximum of 22 kJ/mol for N(CH3)3 and N(CH3)2NH2. Analysis of geometric, vibrational, and electronic data indicate that the SH···N interaction involving SH is slightly stronger than that in which the closed‐shell H2S serves as donor. © 2010 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  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