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 46 毫秒
1.
IR data of eight substituted flavanones and their isomeric hydroxychalkones have been recorded in order to assign the various absorption bands and to study the effect of substituents on C=C out-of-plane deformation (400–700 cm?1), C?H out-of-plane deformation (700–1000 cm?1), C?H in-plane deformation (1000–1300 cm?1), C?O stretch (≈1200 cm?1), OCH3 (1200–1300 cm?1), O?H deformation (1300–1400 cm?1), CH3 deformation (1300–1500 cm?1), benzene ring vibration (1400–1600 cm?1) and C=O stretch (≈1650 cm?1). The δC?H (ring A) in 2′,4′-dihydroxy-3-nitrochalkone appears at 826 cm?1 (s), while in the isomeric flavanone it shows up as three bands, viz., 807 (w), 833 (m) and 881 cm?1 (w). This difference principally arises due to the presence of the electron withdrawing nitro substituent. The C=O stretching vibration in flavanones appears at a higher frequency than in the corresponding hydroxychalkones. This is perhaps due to the lack of conjugation in the former class of compounds. Chloro substituents (ring B) in different positions exert differing effects on νC?O. These differences can be rationalized in terms of a field-effect exerted by the chlorine atom.  相似文献   

2.
Bands with primarily v (C=O) and v (C=O) character in the spectra of 4-hydroxycoumarin and its anion were identified by isotopic substitution with either 13C or deuterium. Two bands of each type were found for spectra of 4-hydroxycoumarin in solution in chloroform, dioxane, or dimethylsulfoxide, with v (C=O) at 1704–1733 cm?1 and ~ 1567 cm?1. Two bands, at 1618 and 1559 cm?1, are associated with v (C=C) in the spectrum of crystalline 4-hydroxycoumarin monohydrate, but only a single v (C=O) band at ~ 1655 cm?1 was observed. Anhydrous 4-hydroxycoumarin has v (C=O) bands at ~ 1700 cm?1 and a shoulder at ~ 1670 cm?1. The strong band at 1660 cm?1 in the spectrum of 4-hydroxycommarin anion in dimethylsulfoxide solution is due to a delocalized v (O = C = O) vibration, whereas the band at 1555 cm?1 has partial v (C=C) character and involves C(3) but not C(2), supporting a fully delocalized char structure for the anion. No evidence for the existence of the 2-hydroxychromone tautomer was found, except in the case of anhydrous 4-hydroxycoumarin in the solid state.  相似文献   

3.
The mononuclear complexes of Zn(II), Cd(II) and Hg(II), [Zn(phen-dione)Cl2], [Cd(phen-dione)Cl2] and [Hg(phen-dione)Cl2], where phen-dione?=?1,10-phenanthroline-5,6-dione, have been synthesized and characterized by elemental analysis and IR, 1H?NMR and electronic absorption spectroscopies. The ν(C=O) of coordinated phen-dione ligands in these complexes shows that the phen-dione is not coordinated to metal ion from its C=O sites. Electronic spectra of the complexes show two absorption bands for intraligand transitions. These absorption bands show dependence on the dielectric constant of solvents. These complexes exhibit an intense fluorescence band around 545?nm in DMSO when the excitation wavelengths are 200?nm at room temperature.  相似文献   

4.
The novel mixed ligand complexes [M(bpy)(phen-dione)](PF6)2 (M?=?Zn(II), Cd(II) and Hg(II), bpy?=?2,2-bipyridine and phen-dione?=?1,10-phenanthroline-5,6-dione) have been synthesized and characterized by elemental analysis, IR, 1H NMR and electronic absorption spectroscopies. The ν(C=O) of coordinated phen-dione in these complexes are very similar to the free phen-dione ligand showing that phen-dione is not coordinated to metal ion from its C=O sites. Absorption spectra of the complexes show two absorption bands for intraligand transitions. These absorption bands show dependence to the dielectric constant of solvent. These complexes exhibit an intensive fluorescence band around 535?nm in DMF when the excitation wavelength is 260?nm at room temperature. The fluorescence intensity of these complexes is larger than that of the free ligand.  相似文献   

5.
Cyanoacetaldehyde (NC? CH2CH?O) and its isomer, cyanovinylalcohol (NC? CH?CH? OH), as possible components of the interstellar medium, comets, or planetary atmospheres, exist in equilibrium in the gas phase, although the latter compound is very much in the minority (2 %). The recording and analysis of the gas‐phase infrared spectrum of the former compound within the 4000–500 cm?1 spectroscopic range and the potential presence of the latter isomer, which could be vital for their detection in these media, are reported. CCSD(T) and G4 high‐level ab initio methods, as well as density functional theory calculations, predict the existence of two stable rotamers of cyanoacetaldehyde. The global minimum has a structure with an unusual O‐C‐C‐C dihedral angle (150°) that falls between the antiperiplanar (180°) and anticlinal forms (120°). The second rotamer, which is about 4.0 kJ mol?1 less stable in terms of free energy, has a planar structure that corresponds to the synperiplanar form (O‐C‐C‐C dihedral angle: 0°). The absorption vibrational bands of the two aldehyde rotamers that are present in the mixture lead to a spectrum with a very complex structure in the region of deformation movements, in which several low‐intensity bands overlap. A complete and unambiguous assignment of the experimental spectrum has been achieved by using the calculated harmonic and anharmonic vibrational frequencies.  相似文献   

6.
Competitive bond dissociation mechanisms for bromoacetyl chloride and 2‐ and 3‐bromopropionyl chloride following the 1[n(O)→π*(C?O)] transition at 234–235 nm are investigated. Branching ratios for C? Br/C? Cl bond fission are found by using the (2+1) resonance‐enhanced multiphoton ionization (REMPI) technique coupled with velocity ion imaging. The fragment branching ratios depend mainly on the dissociation pathways and the distances between the orbitals of Br and the C?O chromophore. C? Cl bond fission is anticipated to follow an adiabatic potential surface for a strong diabatic coupling between the n(O)π*(C?O) and np(Cl)σ*(C? Cl) bands. In contrast, C? Br bond fission is subject to much weaker coupling between n(O)π*(C?O) and np(Br)σ*(C? Br). Thus, a diabatic pathway is preferred for bromoacetyl chloride and 2‐bromopropionyl chloride, which leads to excited‐state products. For 3‐bromopropionyl chloride, the available energy is not high enough to reach the excited‐state products such that C? Br bond fission must proceed through an adiabatic pathway with severe suppression by nonadiabatic coupling. The fragment translational energies and anisotropy parameters for the three molecules are also analyzed and appropriately interpreted.  相似文献   

7.
Spectroscopic evidence for C? H ??? O hydrogen bonding in chloroform ??? acetone [Cl3CH ??? O?C(CH3)2] mixtures was obtained from vibrational inelastic neutron scattering (INS) spectra. Comparison between the INS spectra of pure samples and their binary mixtures reveals the presence of new bands at about 82, 130 and 170 cm?1. Assignment of the 82 cm?1 band to the νO ??? H anti‐translational mode is considered and discussed. In addition, the βC? H mode of CHCl3 at 1242 cm?1 is split in the spectra of the mixtures, and the high‐wavenumber component is assigned to the hydrogen‐bonded complex. The plot of the integrated intensity of this component shows a maximum for x=0.5, in agreement with the 1:1 stoichiometry of the chloroform ??? acetone complex, with a calculated complexation constant of 0.15 dm3 mol?1. Results also show that the complex behaves as an independent entity, that is, despite being weak, such interactions play a key role in supramolecular chemistry.  相似文献   

8.
In this study, the titanyl and vanadyl phthalocyanine (Pc) salts (Bu4N+)2[MIVO(Pc4?)]2? (M=Ti, V) and (Bu3MeP+)2[MIVO(Pc4?)]2? (M=Ti, V) with [MIVO(Pc4?)]2? dianions were synthesized and characterized. Reduction of MIVO(Pc2?) carried out with an excess of sodium fluorenone ketyl in the presence of Bu4N+ or Bu3MeP+ is exclusive to the phthalocyanine centers, forming Pc4? species. During reduction, the metal +4 charge did not change, implying that Pc is an non‐innocent ligand. The Pc negative charge increase caused the C?N(pyr) bonds to elongate and the C?N(imine) bonds to alternate, thus increasing the distortion of Pc. Jahn–Teller effects are significant in the [eg(π*)]2 dianion ground state and can additionally distort the Pc macrocycles. Blueshifts of the Soret and Q‐bands were observed in the UV/Vis/NIR when MIVO(Pc2?) was reduced to [MIVO(Pc . 3?)] . ? and [MIVO(Pc4?)]2?. From magnetic measurements, [TiIVO(Pc4?)]2? was found to be diamagnetic and (Bu4N+)2[VIVO(Pc4?)]2? and (Bu3MeP+)2[VIVO(Pc4?)]2? were found to have magnetic moments of 1.72–1.78 μB corresponding to an S=1/2 spin state owing to VIV electron spin. As a result, two latter salts show EPR signals with VIV hyperfine coupling.  相似文献   

9.
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   

10.
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   

11.
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   

12.
Raman spectra have been measured for aqueous Al2(SO4)3 solutions from 25 to hydrothermal conditions at 184°C under steam saturation. The Raman spectrum at 184°C contained four polarized bands in the S–O stretching wavenumber range, which suggest that a new sulfato complex, where sulfate acts as a bridging ligand (possibly bidentate or tridentate), is formed in solution, in addition to a 1:1 aluminium(III) sulfato complex, where sulfate is monodentate, which is the only ion pair identified at room temperature. Under hydrothermal conditions, it was possible to observe the hydrolysis of aluminium(III) aqua ion by measuring the relative intensity of bands due to SO2? 4 and HSO4 ?, according to the coupled equilibrium reaction [Al(OH2)6]3+ + SO4 2? ? [Al(OH2)5OH]2+ + HSO4 ?. The precipitate in equilibrium with the solution at 184°C could be characterized as a hydronium alunite, (H3O)Al3(SO4)2(OH)6, by chemical analysis, X-ray diffraction, and Raman and infrared spectroscopy.  相似文献   

13.
Depolarization ratios ρ of the Raman bands due to CH3 stretching at 2907 cm?1 and the Si? O skeletal mode at 491 cm?1 have been measured in polydimethylsiloxane gum as a function of temperature from 100°C to ?45°C. Below 0°C the changes in p have been interpreted in terms of the formation of helical regions in the gum. The enthalpy of helix formation ΔH has been determined as 3200 ± 600 cal/mole. An upper limit on the entropy change, ΔS, of 16 ± 3 e.u./mole and minimum values of helix content at different temperatures have been found. The Raman spectrum of crystalline polydimethylsiloxane is presented.  相似文献   

14.
Reaction of Diphenoxyphosphorylchloride with N,N-disubstituted Ureas – Formation of Phosphorylated Biuret Compounds N′,N′-disubstituted N-diphenoxyphosphorylureas, (PhO)2P(O)? NH? CO? NR1R2 (R1 = R2 = Et, 1 ; n-Pr, 2 ; n-Bu, 3 ; i-Bu, 4 ; R1 = Me and R2 = Ph, 5 ) as well as phosphorylated biuret compounds, (PhO)2P(O)? NH? CO? NH? CO? NR1R2 are obtained in the reaction of diphenoxyphosphorylchloride with N,N-disubstituted ureas and triethylamine. The biuret derivatives are formed via (PhO)2P(O)NCO. Their yield rises if the reaction is carried out without amine. The X-ray crystal structure analysis of (PhO)2P(O)? NH? CO? NH? CO? NPr2, 8 , shows that dimers exist in the crystal with intermolecular as well as intramolecular hydrogen bonds. The framework formed by atoms P? N1? C1(O4)? N2? C2(O5)? N3(C3)C6 is planar. The existence of a rotation barrier along the bond C2–N3 was detected by NMR spectroscopy.  相似文献   

15.
The carbonyl bands of 2-formyl and 3-formyl-furans [18O] are determined and explained. The magnitude of the inductive and mesomeric effects in 2-acyl and 3-acyl-thiophenes and furans is discussed on the basis of a comparison of the frequencies and the intensities of the v(CO) and v(CC) bands. The conformational preference of some aroyl-furans and aroyl-formyl-furans and the rotation around the C(Ar)-C(O) bonds are related to these parameters.  相似文献   

16.
Acyl- and Alkylidenephosphines. XXXII. Di-cyclohexoyl- and Diadamant-1-oylphosphine – Keto-Enol Tautomerism and Structure Lithium dihydrogenphosphide · DME (1) [12] and cyclo-hexoyl or adamant-1-oyl chloride react in a molar ratio of 3:2 to give lithium di-cyclo-hexoylphosphide · DME and the corresponding diadamant-1-oylphosphide.2THF (1) resp. Treatment of these two compounds with 85% tetrafluoroboric acid. diethylether adduct yields di-cyclo-hexoyl- ( 1b ) and diadamant-1-oylphosphine ( 1c ). In nmr spectroscopic studies 1b over a range of 203 to 343 K, a strong temperature dependence of the keto-enol equilibrium is found; thermodynamic data characteristic for the formation of the enol tautomer (ΔH0 = ?4.3 kJ. mol?1; ΔS0 = ?9.2 J. mol?1. K (?1) are compared of 1,3-diketones. The enol tautomer of diadamant-1-oylphosphine ( E-1c ) as obtained from a benzene solution in thin colourless plates, crystallizes in the monoclinic space group P21/c {a = 722.2(2); b = 1085.5(4); c = 2434.8(5) pm; ß = 96.43(2)° at –100 ± 3°C; Z = 4}. An X- ray structure analysis (Rw = 0.033) shows bond lengths and angles to be almost identical within the enolic system (P? C 179/180; C? O 130/129; C? C(adamant-1-yl) 152/153 pm; C? P? C 99°; P? C? O 124°/124°; P? C? C 120°/120°; C? C? O 116°/116°. The geometry of the very strong, but probably asymmetric O‥H‥O bridge is discussed (O? H 120/130, O‥O 245 pm).  相似文献   

17.
Polysulfonyl Amines. XLII. An Aquasilver(I) Complex with an Ag(m?-H2O)2Ag Structural Unit: Characterization of Aqua(1,1,3,3-tetraoxo-1,3,2-benzodithiazolido)silver(I) by X-Ray Diffractometry and Thermal Analysis The title compound C6H4(SO2)2NAg · H2O, where C6H4(SO2)2Nº is the anion of 1,2-benzenedisulfonimide, crystallizes in the monoclinic space group C2/m with (at ?95°C) a = 1 129,7(3), b = 1 196.1(3), c = 810.7(2) pm, β = 124.25(2)°, V = 0.9055 nm3, Z = 4, Dx = 2.524 Mg m?3. The crystal packing consists of [Ag(m?-H2O)2Ag{m?-C6H4(SO2)2N}2]n bands with crystallographic mirror symmetry, associated into layers by H-bonds with O(W)—O(S) 289.7 pm. The Ag(m?-H2O)2Ag moiety forms a planar four-membered ring with Ag? O(W)? Ag 97.3°, O(W)? Ag? O(W) 82.7° and Ag°Ag 372.1 pm. In the Ag{C6H4(SO2)2N}2Ag′ unit, the anions act as tridentate (N, 1-O, 3-O)-ligands: One is N-bonded to Ag and O,O-chelated to Ag′, the other N-bonded to Ag′ and O,O-chelated to Ag. The silver atoms are (O4N)-pentacoordinate, with nitrogen in the apical position of a distorted square pyramid [Ag? N 223.6, Ag? O(W) 247.8, Ag? O(S) 259.4 pm]. The thermochemical behaviour of the hydrate was investigated by thermal analysis and calorimetry. Water is only released at temperatures above 220°C. The dehydration enthalpy at 298 K is + 13.9 kJ mol?1.  相似文献   

18.
The electronic properties of four divinylanthracene‐bridged diruthenium carbonyl complexes [{RuCl(CO)(PMe3)3}2(μ? CH?CHArCH?CH)] (Ar=9,10‐anthracene ( 1 ), 1,5‐anthracene ( 2 ), 2,6‐anthracene ( 3 ), 1,8‐anthracene ( 4 )) obtained by molecular spectroscopic methods (IR, UV/Vis/near‐IR, and EPR spectroscopy) and DFT calculations are reported. IR spectroelectrochemical studies have revealed that these complexes are first oxidized at the noninnocent bridging ligand, which is in line with the very small ν(C?O) wavenumber shift that accompanies this process and also supported by DFT calculations. Because of poor conjugation in complex 1 , except oxidized 1+ , the electronic absorption spectra of complexes 2+ , 3+ , and 4+ all display the characteristic near‐IR band envelopes that have been deconvoluted into three Gaussian sub‐bands. Two of the sub‐bands belong mainly to metal‐to‐ligand charge‐transfer (MLCT) transitions according to results from time‐dependent DFT calculations. EPR spectroscopy of chemically generated 1+ – 4+ proves largely ligand‐centered spin density, again in accordance with IR spectra and DFT calculations results.  相似文献   

19.
The purpose of this study is to propose a correlation between IR spectra and the urea fraction of waterborne polyurethanes (PUs) to investigate the side reaction, that is, isocyanate–water reaction, during polymerization. This method is based on the decomposition of the spectrum in amide I range, that is, 1600–1800 cm?1, where the bands of interest overlap. Several individual bands present in this region were resolved by employing Fourier self‐deconvolution (FSD) and Gaussian curve‐fitting techniques, and the intensity ratio of urethane's C?O to urea's C?O was determined. To realize some quantitative measurements, a calibration curve was established with some polyurethane‐urea samples, characterized by 1H NMR, which were used as standards. The concentration ratios of urethane groups to urea groups were determined from 1H NMR. A good correlation was evidenced between IR and 1H NMR measurements. Moreover, waterborne PUs were prepared by miniemulsion polymerization of IPDI with diols. From quantitative IR analysis, it was shown that a vinylic monomer, as a solvent of polyaddition, restrained the isocyanate–water reaction, and this side reaction was influenced by the hydrophilicity of the vinylic monomers. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 2433–2444, 2008  相似文献   

20.
The thermolysis under argon of various polysiloxane resins containing D, T, DH, or TH units was investigated using thermogravimetric analysis combined with mass spectroscopy (TG/MS analysis) and solid-state 29Si-NMR. Redistribution reactions involving the exchange of Si? C/Si? O bonds or Si? H/Si? O bonds were evidenced in addition to the exchange of Si? O/Si? O bonds reported to date. These reactions significantly modify the initial siloxane units and lead to an escape of volatile silanes or siloxanes. The exchange of Si? H/Si? O bonds takes place at lower temperatures (300°C) than the exchange of Si? C/Si? O bonds (500°C).  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