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.
Polyamide 6 has been mechanically destructed in vacuo. At -70°, the ESR spectrum corresponds to the sum of the component spectra of three radicals NH?HCH2, ·CH2NHCO, and ·CH2CONH. After introducing air into the ampoule, the spectrum changes even at -70°; the changes have been studied up to 0°. The spectrum of the peroxy radical ROO· (with line width 1.57 mT, g1 = 2.0089 and g| = 2.0301) predominates.  相似文献   

2.
The standard thermodynamic parameters (Δr G°, Δr H°, TΔr S°) of the reaction of molecular complex formation between 18-crown-6 ether (18C6) and l-phenylalanine (Phe), [Phe18C6], have been obtained from calorimetric titration experiments carried out by means of the microcalorimetric system TAM III (TA Instruments, USA) at T = 298.15 K in water–dimethylsulfoxide (H2O–DMSO) solvents. Results show that the increase of the DMSO concentration in the mixed solvents brings about an increase of the [Phe18C6] complex stability and of the exothermicity of the reaction of complex formation.  相似文献   

3.
The integral enthalpies of solution Δsol H m of L-serine in mixtures of water with acetonitrile, 1,4-dioxane, dimethylsulfoxide (DMSO), and acetone were measured by solution calorimetry at organic component concentrations up to 0.31 mole fractions. The standard enthalpies of solution (Δsol H°), transfer (Δtr H°), and solvation (Δsolv H°) of L-serine from water into mixed solvents were calculated. The dependences of Δsol H°, Δsolv H°, and Δtr H° on the composition of aqueous-organic solvents contained extrema. The calculated enthalpy coefficients of pair interactions of the amino acid with cosolvent molecules were positive and increased in the series acetonitrile, 1,4-dioxane, DMSO, acetone. The results obtained were interpreted from the point of view of various types of interactions in solutions and the influence of the nature of organic solvents on the thermochemical characteristics of solutions.  相似文献   

4.
The electrochemical behaviour of the Au/SCN? (DMSO) interface is studied at 25°C≤T≤65°C, by means of the triangular potential sweep technique and the RDE. Within a relatively limited potential range the main electrode process is:3 SCN?=(SCN)3?+2e The kinetics of this reaction involves a mixed control. At higher potentials the electrodissolution of the base metal takes place. Side reactions occurring there make the overall process rather complex. The electrochemical behaviours of the Au/SCN? (DMSO) and Au/SCN? (ACN) are compared.  相似文献   

5.
The C?D bond stretching vibrations of deuterated dimethyl sulfoxide ([D6]DMSO) and the C2?H bond stretching vibrations of 1,1,1,5,5,5‐hexafluoropentane‐2,4‐dione (hfac) ligand in anion are chosen as probes to elucidate the solvent–solute interaction between chelate‐based ionic liquids (ILs) and DMSO by vibrational spectroscopic studies. The indirect effect from the interaction of the adjacent S=O functional group of DMSO with the cation [C10mim]+ and anion [Mn(hfac)3]? of the ILs leads to the blue‐shift of the C?D stretching vibrations of DMSO. The C2?H bond stretching vibrations in hfac ligand is closely related to the ionic hydrogen bond strength between the cation and anion of chelate‐based ILs. EPR studies reveal that the crystal field of the central metal is kept when the chelate‐based ILs are in different microstructure environment in the solution.  相似文献   

6.
A thick film of aniline-formaldehyde copolymer and PMMA is synthesized via dispersion of aniline-formaldehyde copolymer powder as filler particles in PMMA with two different concentrations. Variation of the complex elastic modulus and mechanical loss factor (tanδ) with temperature is studied. It is observed that the complex elastic modulus decreases with temperature owing to thermal expansion of films. On the other hand, tanδ increases up to a characteristic temperature beyond which it shows a decreasing trend toward melting. Transition temperature T g of sample S1 (pure PMMA) is found to be 80°C. In sample S2 (1 wt % aniline formaldehyde copolymer), the peak of tanδ at a lower temperature (66°C) corresponds to glass transition temperature T g of the PMMA matrix, while the peak of tanδ at a higher temperature (107.8°C) corresponds to T g of a polymer chain restricted by filler particles of aniline-formaldehyde copolymer. A further increase (10 wt % aniline-formaldehyde copolymer) in the concentration of filler particles of aniline-formaldehyde copolymer results in a more compact structure and a shift of T g to a higher temperature, 122.2°C. This shift in the glass transition temperature of thick films of aniline-formaldehyde copolymer and PMMA is dependent upon the concentration of filler particles in the sample.  相似文献   

7.
Temperature Dependent Phase Transitions of Sodium-rich Amalgams Starting from thermoanalytical measurements the temperature dependent phase transitions of the microcrystalline, sodium rich amalgams Na 8 Hg3 (? Na2.67Hg) and Na3Hg are investigated using a fast, position sensitive X-Ray detector in combination with a modified counter Guinier system. Near the melting point (64°C) Na8Hg3 shows two phase transitions. At 54°C α-Na 8 Hg3 forms a slightly distorted monoclinic defect variant of the cubic Li3Bi-structure (β-Na8Hg3). At 62°C this modification transforms into the corresponding undistorted cubic phase (γ-Na 8 Hg3). The increase of symmetry results mainly from the fact, that small deviations of the mercury atoms from the positions of an ideal close packing vanish due to strong thermal vibrations. The “literature-melting point” of Na3Hg (33°C) corresponds to a phase transition from α-Na3Hg (hexagonal, modified Na3As type) into β-Na3Hg (monoclinic distorted Li3Bi type, corresponding to β-Na 8 Hg3). The actual melting point of Na3Hg is 59°C.  相似文献   

8.
Complexation of the 18-crown-6 ether (18C6) with glycine (Gly) in mixed H2O-DMSO solvents with the composition of 0.1, 0.2, and 0.25 mole fraction of DMSO (T = 298.15 K) was studied calorimetrically. Thermodynamic characteristics of the reaction of the formation of the molecular Gly18C6 complex (Δr G°, Δr H°, TΔr S°) were calculated from the calorimetric data. It was established that the change in the stability of the Gly18C6 complex is mainly determined by the predominance of the enthalpy component of the Gibbs energy over the entropy component. It was shown during the analysis of the enthalpy contributions of the reagents to the enthalpy of the reaction of the formation of Gly18C6 that the change in the enthalpy of the reaction upon a change the solvent composition was due to changes in the solvation state of 18C6.  相似文献   

9.
The crystal and molecular structure of the orthorhombic form of 1,8-dinitronaphthalene, C10H6N2O4, has been reinvestigated at 22°C and examined at 97°C using single-crystal diffractometer data in order to throw some light on the structural aspects of the polymorphism of the title compound. The space group of the crystal is P212121 with a = 11.375 (1), b = 14.974 (5), c = 5.388 (1) Å at 22°C and a = 11.475 (1), b = 15.002 (1), c = 5.425 (6) Å at 97°C. Z = 4 at both temperatures. The data, collected with CuKα radiation, were refined by full-matrix least squares to R = 0.049 for 1087 reflexions (22°C) and R = 0.061 for 1078 reflexions (97°C). The structure was found to be in agreement with that determined by Z. A. Akopian, A. I. Kitaigorodskii, T. J. Struchkov (Zh. Strukt. Khim.6, 729 (1965)). Changes occurring on heating to a temperature of 97°C, which is close to the transition temperature (of cpc type) 100–105°C, suggest that the possible mechanism of the phase transition is that of the displacive transformation of secondary coordination. Anistotropic temperature factors of the majority of atoms increased by ca. 50% with no particular direction of atom vibrations being distinguished.  相似文献   

10.
Steric effects on proton transfer from, and to, hydroxylic oxygen have been studied in a series of seventeen α-methyl and a-benzyl cyclohexanols in anhydrous DMSO, under both acid and base catalysed conditions, using dynamic MNR techniques. The protonation rate constants (k1 ? 106 M-1 s-1 at 25°C) obey a Taft-Ingold relationship, containing only a steric contribution Es = EsOH + Esα, where: EsOH = 0 or 0.15 for an axial or equatorial hydroxyl respectively and Esα = ?0.070 (or ?0.115) for substituting an α-hydrogen by a methyl (or benzyl) group. An equatorial hydroxylic function is therefore 40% more reactive than its axial homologue. These kinetic data are fairly consistent with structural information resulting from IR spectroscopy (vco and vOH vibrations) and from NMR (hydroxylic chemical shifts and coupling constants).  相似文献   

11.
The stretched film technique has been used to resolve the two overlapping 1La and 1Lb transitions of indole. 5-Methyl indole was studied as a model compound in two polymer matrices, polyethylene (PE) and polyvinylalcohol (PVA). The directions of the transition moments with respect to the long molecular axis are found to be in PE − 10° and +22° for 1La and 1Lb, respectively. These angles are found to be −25° and +29° in PVA. 1La has two prominent vibrations in PE at 2040 and 2895 cm−1 from the origin. 1Lb has four prominent vibrations in PE at 761, 1308, 1937 and 3196 cm−1 from the origin. In PVA, 1La has two prominent vibrations at 2149 and 2874 cm−1 from the origin and 1Lb has four prominent vibrations at 1049, 1352, 2229 and 3196 cm−1 from the origin.  相似文献   

12.
The density, viscosity and conductivity of binary mixtures of the front line antitubercular drug isoniazid (INH), in aqueous solution and dimethylsulfoxide (DMSO) solution, were determined at various temperatures (25, 37 and 55?°C) up to 0.3 mol?L?1 of INH. The apparent molar volumes were calculated from the density data. In the INH + water system the apparent molar volume of INH changed smoothly, whereas in the INH + DMSO system it passes through a maximum. Also, both systems showed pronounced maxima in their viscosity and conductivity isotherms. In addition, UV?CVis, FT-IR and 1H NMR spectroscopy were performed on the solutions. On the basis of this data, the predominant molecular interactions occurring between INH and water and between INH and DMSO were found to be hydrogen bonds. Furthermore, the susceptibility profile of DMSO, INH and its combination was studied against M. tuberculosis H37Rv and the minimum inhibitory concentration (MIC) determined. The results suggest a synergistic effect of INH at sub-MIC concentrations and DMSO.  相似文献   

13.
The effect of heat treatment at temperatures above 300°C on the low temperature relaxation of poly(4,4′-oxydiphenylenepyromellitimide) (Kapton H-film) was studied by wide-line nuclear magnetic resonance (NMR), mechanical, and dielectric measurements. In the NMR line spectrum of the as-received film, a narrow component above ?60°C and a broad component which begins to narrow at about ?100°C were observed. It is proposed that the narrow component is associated with absorbed water, because it disappeared in the heat-treated film at 300°C in N2. On the other hand, the behavior of the broad component was not influenced by heating to 300°C in N2. Mechanical and dielectric loss peaks were observed at ?75°C (11 Hz) and ?65°C (1 kHz), respectively, in the as-received film. These loss peaks did not change in intensity with heating to 300°C in N2. It is proposed that the mechanical and dielectric loss peaks corresponding to the narrowing of the NMR broad component are associated with the local-mode motion of the diphenylether portion of the polypyromellitimide chain. It was found that crosslinks are formed by heating to 374°C in air through coupling of the diphenylether portions of the molecular chains. With the formation of crosslinks the dielectric loss peak shifted toward higher temperature and the intensity decreased through restriction of the local-mode motion of the diphenylether portion of the molecular chain.  相似文献   

14.
The ionization enthalpy of benzoic acid has been measured calorimetrically at 25°C in H2ODMSO mixtures ranging from pure water to a maximum DMSO molar ratio XDMSO = 0.80. With the increase of DMSO content, the ionization becomes more and more endothermic, and for XDMSO = 0.8 the ionization enthalpy is about 6 kcal mol?1 higher than in water. By also measuring the solution enthalpy of crystalline benzoic acid in the mixtures, it has been shown that the solvation of the undissociated molecule is the main cause for the increase of the dissociation enthalpy. A comparison has been made between the relative enthalpies of benzoic and hydroxide ions in H2ODMSO mixtures.  相似文献   

15.
The reaction of tetraphenylphosphonium chloride with an equimolar amount of potassium tetrachloroplatinate or hexachloroplatinic acid in dimethyl sulfoxide gave the complexes [Ph4P]+[PtCl3(DMSO)]? (I) and [Ph4P]+[PtCl5(DMSO)]? (II), respectively. The phosphorus atoms in the cations have tetrahedral environment, the CPC angles and P-C distances 105.63(13)°–112.13(14)°, 1.795(3)–1.797(3) Å I) and 105.7(3)°–112.9(3)°, 1.783(7)–1.791(6) Å II). The platinum coordination polyhedra in the anions [PtCl3(DMSO)]? and [PtCl5(DMSO)]? are distorted square (Pt-S, 2.1937(8); Pt-Cl, 2.2894(10)–2.3024(10) Å; trans-angles: SPtCl, 177.38(4)°; ClPtCl, 175.40(4)°) and octahedron (Pt-S 2.291(2) Å; Pt-Cl, 2.312(2)–2.334(2) Å, trans-angles: SPtCl, 178.28(9)°; ClPtCl, 178.80(9)° and 178.88(8)°).  相似文献   

16.
Bromo dimethyl sulfoxide osmium(II) complexes were synthesized: trans-[OsBr2(dmso-S)4] (1) was obtained by the reaction of K2[OsBr6] with DMSO in the presence of SnBr2 at 100°C and cis,fac-[OsBr2(dmso-S)3(dmso-O)] (2) was prepared by thermal isomerization of 1 in a DMSO solution at 150°C. The coordination mode of DMSO molecules was determined by IR and 1H and 13C NMR spectroscopy. X-ray diffraction analysis showed that compound 2 crystallizes in the monoclinic system, space group P21/n; a = 8.4711(5) Å, b = 27.7876(15) Å, c = 8.5569(5) Å, β = 115.7110(10)°; Z = 4. The coordination polyhedron of osmium is a distorted octahedron; the osmium environment is formed by two cis-arranged bromine atoms and three fac-S-coordinated and one O-coordinated DMSO molecules. The interconversion of complexes in solutions was studied by UV/Vis and 1H and 13C NMR spectroscopy. In chloroform and DMSO, complex 2 isomerizes to cis-[OsBr2(dmso-S)4] and (in the light) to 1. The complexes trans-[OsX2(dmso-d6)4], where X = Cl, Br, were isolated from DMSO-d 6 and characterized by the IR spectra.  相似文献   

17.
Nanocrystalline magnesium chromite spinel was synthesized through hydrothermal reaction of metal nitrate solutions in stoichiometric amount at different pH, temperature and time intervals. The synthesized products were characterized for crystallinity, phase identification, and surface morphology by X-ray diffraction (XRD) and scanning electron microscope (SEM). XRD patterns showed that as-synthesized product remained amorphous up to 250 °C. However, well-crystallized magnesium chromite spinel structure is formed after calcination at 850 °C. Rietveld refinement study confirms the formation of single-phase cubic structure MgCr2O4 with lattice parameter a = 8.3347 Å, and Fd3m space group. The as-processed MgCr2O4 products showed extensive XRD line broadening, and the mean crystallite size of such crystals was found to be mainly in size range of 85–124 nm. Surface SEM images of calcined specimens revealed that the matrix is uniform, and no separation of secondary phase was detected. Thermal stability was examined by thermogravimetry (TG), differential thermal analysis (DTA), and differential scanning calorimetry. TG/DTA reveals that MgCr2O4 is thermally stable above 700 °C. Fourier transform infrared (FTIR) spectra studies shows two strong bands, one around 600 cm?1 which is attributed to the intrinsic vibrations of tetrahedral and other at 400 cm?1 is due to octahedral one. FTIR confirms the formation of metal oxides. The bandgap energy was estimated by absorption spectroscopy in ultraviolet–visible range and was found to be 0.693 eV for MgCr2O4 specimen sintered at 1,000 °C. Isothermal shrinkage characteristic and coefficient of thermal expansion were determined by dilatometry. The powder specimens showed excellent densification at 1,250 °C temperature and uniformly fine grain sintered ceramics (>90 % relative density) with submicron grain size (2–5 μm) were obtained after sintering at 1,000–1,250 °C. Impedance studies were carried out at room temperature and equivalent circuit model (R 1 Q 1) (R 2 Q 2) (R 3 Q 3) is used to explain different relaxation processes. We report largest impedance values i.e., 6.74 × 108 Ω, reduced dielectric constant (≈1.0), and low tangent loss (0.8) for MgCr2O4 sintered at 1,250 °C.  相似文献   

18.
Thioacetamide has been studied by electron diffraction in the gas phase, utilizing a new nozzle construction and using a broad electron beam. The molecule has Cs symmetry, and one C-H bond eclipses the CS bond. The most important structural parameters are: rg(C-N) = 135.6(3) pm, rg(C—C) = 151.2(4) pm, rg(CS) = 164.7(3) pm,∠αCCS = 122.9(3)° and ∠αCCN = 114.8(4)°. Parenthesized values are one standard deviation where correlation among data and uncertainty in the electron wavelength have been included. The methyl barrier, V3, is found from the electron diffraction data to be 4.56 kJ mol?1. This corresponds to a torsional frequency of 131 cm?1.  相似文献   

19.
The Raman and infrared spectra of all-trans-astaxanthin (AXT) in dimethyl sulfoxide (DMSO) solvent were investigated experimentally and theoretically. Density functional cal-culations of the Raman spectra predict the splitting of the υ1 band into υ1-1 and υ1-2 compo-nents. The absence of splitting in Raman experimental spectra is ascribed to the competition between the two symmetric C=C stretching vibrations of the backbone chain. The υ1 band is very sensitive to the excitation wavelength: resonance excitation stimulates the higher-frequency υ1-2 mode, and off-resonance excitation corresponds to the lower-frequency υ1-1 mode. Analyses of the intramolecular hydrogen bonding between C=O and O-H in the AXT/DMSO system reveal that the C4=O1...H1-O3 and C4'=O2...H2-O4 bonds are strengthened and weakened, respectively, in the electronically excited state compared with those in the ground state. This result reveals significant variations of the AXT molecular structure in different electronic states.  相似文献   

20.
The molecular structure of carbonyl fluoride has been determined by electron diffraction. The results have been used in conjunction with the rotational constants reported by Carpenter in a combined structure analysis. The values so obtained are rz (C=O) = 1.1717 ± 0.0013 Å, rz (C-F) = 1.3157 ± 0.0005 Å, and ∠zF-C-F = 107.71 ± 0.08°. These agree with the corresponding parameters estimated by Carpenter from the rotational constants alone. The effective constants, α3, representing the cubic anharmonicity of bond stretching vibrations have been estimated.  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