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.
A free-radical substitution for the chlorine atoms with the 1,4-dioxan-2-yl fragment in the tris-dichloro-substituted dioximate iron(II) clathrochelate is accompanied by the oxidative dioxane ring opening as a side process. The compound obtained was characterized by elemental analysis data, 1D and 2D 1H and 13C{1H} NMR spectroscopy (in solution), MALDI TOF mass spectrometry, as well as by single-crystal X-ray diffraction.  相似文献   

2.
The effect of 2-thio-6-aminouracil on the oxygen uptake kinetics has been studied in 1,4-dioxane free-radical chain oxidation as a model system. The presence of a thiocarbonyl group in the 2-position of the uracil ring makes 6-aminouracil highly reactive towards 1,4-dioxane peroxy radicals. The rate constant of the 1,4-dioxane peroxy radical interaction with 2-thio-6-aminouracil has been measured to be k7 = (3.0 ± 0.5) × 105 L mol–1 s–1 (333 K). The stoichiometric inhibition factor f = 1.1 ± 0.1 has been determined.  相似文献   

3.
The antioxidation activity of 5-substituted 6-methyluracils was quantitatively estimated in the model system of initiated radical-chain oxidation of 1,4-dioxane. The rate constants of the reactions of 1,4-dioxane peroxide radicals with 6-methyluracil (1), 6-methyl-5-piperidinouracil (2), 6-methyl-5-morpholinomethyluracil (3), 6-methyl-5-morpholinouracil (4), 6-methyl-5-methylaminouracil (5), 5-ethylamino-6-methyluracil (6), and 5-hydroxy-6-methyluracil (7) were measured. Among compounds 1–7, derivative 7 is most efficient with an inhibition rate constant of (5.2±0.1) · 104 L mol-1 s-1 (60 °C).  相似文献   

4.
The conformation of several trans-2,3-diaryloxy-1,4-dioxanes has been studied using 1H NMR techniques. Trans-2,3-bis(4-nitrophenoxy)-1,4-dioxane and trans-2,3-bis(4-methoxyphenoxy)-1,4-dioxane have been found to be predominantly ( ≈98%) in diaxial conformation in CDCl3). On the other hand, trans-2,3-bis(2,6-dimethylphenoxy)-1,4-dioxane exists in the same conditions as a 66:33 mixture of diaxial and diequatorial conformers. An explanation based on the fulfilment of the exo-anomeric effect is provided.  相似文献   

5.
The complexation of a macrocycle containing thiopyrimidine and uracil moieties (M) with amino acids and some dicarboxylic acids was studied by pH-metric, UV-VIS, 1H NMR spectroscopy methods in chloroform, methanol, aqueous 1,4-dioxane, and biphasic water–chloroform media. The complexation of M with acids is too weak to solubilize them from the solid state into chloroform solutions containing M. The 1H NMR spectra and pH-metric data of aqueous 1,4-dioxane (80 vol.%) reveal the pH-dependent 1:1 binding between M and the acids studied. The protonation of M is not a prerequisite for binding of fumaric, succinic, o-phtalic acids and the series of amino acids, whereas binding of maleic acid requires the protonation of both thiopyrimidine moieties of M. Therefore,M·(H+)2 exhibits strong selectivity towards maleic acid in aqueous 1,4-dioxane and in biphasic water–chloroform media.  相似文献   

6.
The dichlorine-containing tris-dioximate iron(II) clathrochelate undergoes a radical addition-elimination reaction with the radical derivatives of 1,4-dioxane to yield ribbed-functionalized C-alkylated macrobicyclic complexes. The clathrochelates obtained have been characterized using elemental analysis, MALDI-TOF mass, IR, UV-Vis and multinuclear NMR spectra, as well as X-ray crystallography. This unexpected and previously unknown radical reaction is a novel approach to the ribbed functionalization of tris-dioximate d-metal clathrochelates, and it can be extended to other cyclic ethers.  相似文献   

7.
Novel polyamides of 2,5-bis(amino methyl) 1,4-dioxane (cis/trans-BAMD) with adipic/sebacic acids and of 2,5-bis(carboxy methyl)/2,5-bis(carboxy) 1,4-dioxane with 1,6-hexane diamine have been prepared. Because of the slightly higher conformational flexibility of the 1,4-dioxane ring in comparison with that of the cyclohexane ring, the BAMD polyamides have lower Tg and Tm than the corresponding 1,4-bis(amino methyl) cyclohexane (BAMC) polyamides. The symmetry of the trans-dioxane moiety permits high crystallinity in the trans-BAMD polymers. The crystallinity Tg and Tm of the trans polymers are decreased with the incorporation of the cis-dioxane moiety which lacks a plane of symmetry. Because of the hydrophilic nature of the dioxane ring, t-BAMD-6 has good moisture-regain properties, yet is melt processable (Tm = 292°C).  相似文献   

8.
The results from electrooptical absorption measurements (EOAM) on the ground and excited Franck–Condon state dipole moments of Prodan and Laurdan in 1,4-dioxane and cyclohexane are presented. The ground and excited Franck–Condon state electric dipole moments as well as the respective transition moment of both probes are parallel. The electric dipole moments of Prodan and Laurdan in the ground state in cyclohexane and 1,4-dioxane have values within the range (15.7–16.5) × 10−30 C m. On optical excitation the dipole moments increase by (42.1–49.5) × 10−30 C m. The obtained results are compared with the values of the dipole moments of Prodan and Laurdan determined by other methods.  相似文献   

9.
The antiradical activity of fullerene C60 was studied for the oxidation of 1,4-dioxane and styrene initiated by azobisisobutyronitrile and benzoyl peroxide as model reactions. The effective rate constants of the reaction of peroxyl radicals with fullerene C60 (k 7) and the stoichiometric inhibition factor (f eff) were determined in air ( $P_{O_2 }$ = 0.21 atm) and oxygen ( $P_{O_2 }$ = 1.0 atm). The rate of the liquid-phase oxidation of 1,4-dioxane does not depend on $P_{O_2 }$ , and the effective rate constant of inhibition is k 7 = (2.4 ± 0.2) × 104 L mol?1 s?1. Chain termination in the oxidation of styrene occurs when C60 reacts with both the peroxyl radicals (k 7 = (1.2 ± 0.1) × 103 L mol?1 s?1) and alkyl (k 8 = 1.07 × 107 L mol?1 s?1) radicals.  相似文献   

10.
The structure, vibrational force constants and infrared and Raman intensities have been calculated for 1,3,5-trioxane using a 3-21G basis set. These results have been used to identify some possible inaccuracies in experimental diffraction based structures and in vibrational assignments. It is demonstrated that there is a marked contrast between the trends in the vibrational force constants and in the King's effective atomic charges of the axial and equatorial CH bonds in the series cyclohexane, 1,4-dioxane and 1,3,5-trioxane. The axial CH stretching force constant decreases by 0.04 mdyne Å−1 for each adjacent oxygen atom, whereas that of the equatorial CH bond increases by 0.15 mdyne Å−1 per oxygen. In trioxane the effective atomic charge of the axial hydrogen is twice that of the equatorial. Atomic polar tensors are calculated in a bond oriented frame, and the effect of the oxygens on CH stretching and bending mode intensities discussed. Some properties are also calculated using the 4-31G basis.  相似文献   

11.
Heats of solution of cyclohexane and squalane in mixtures of 1-octanol with acetone or 1,4-dioxane within the entire range of compositions were measured calorimetrically. The enthalpies of solution of cyclohexane, squalane, and hexadecane in the mixtures of acetone or 1,4-dioxane with 1-octanol show linear correlation. Enthalpies of solvation of propane, decane, and eicosane in these mixtures were calculated.  相似文献   

12.
The equilibrium solubility and preferential solvation of triclocarban in {1,4-dioxane (1) + water (2)} mixtures at 298.15 K was reported. Mole fraction solubility varies continuously from 2.85 × 10–9 in neat water to 2.39 × 10–3 in neat 1,4-dioxane. Solubility behaviour was adequately correlated by means of the Jouyban-Acree model. Based on the inverse Kirkwood-Buff integrals, preferential solvation parameters were calculated. Triclocarban is preferentially solvated by water in water-rich mixtures (0.00 < x1 < 0.18) and also in 1,4-dioxane-rich mixtures (0.78 < x1 < 1.00) but preferentially solvated by 1,4-dioxane in mixtures with similar solvent compositions.  相似文献   

13.
Excess enthalpies, excess heat capacities, excess volumes and sound velocities of the mixture of dioxane isomers, 1,3-dioxane and 1,4-dioxane, were measured. One of the isomers, 1,4-dioxane is considered as non-polar liquid and the other as polar liquid. Excess enthalpies are positive and small, less than 55 J mol-1. Excess heat capacities are also very small and the curve is W-shaped, and the values are from 0.03 to -0.08 J mol-1 K-1. Excess volumes and excess isentropic compressibilities are small and positive, and less than 0.03 cm3 mol-1 and 0.8 TPa-1. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

14.
Abatract  The reaction of the dichloride iron(ii) tris-dioximate clathrochelate with triethyl orthoformate in the presence of metallic sodium afforded the diethoxy-substituted macrobicyclic complex. The first alkoxy-containing clathrochelate was characterized using elemental analysis, MALDI-TOF mass spectrometry, IR, UV-Vis, 1H and 13C{1H} NMR spectroscopy, and X-ray diffraction analysis. Dedicated to the 90th anniversary of the L. Ya. Karpov Institute of Physical Chemistry. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 6, pp. 1283–1288, June, 2008.  相似文献   

15.
The 1H spectrum of (15N)acetamide has been measured in dimethyl sulphoxide (DMSO), methyl propyl ketone (MPK), 1,3-dioxane, 1,4-dioxane, D2O, acetonitrile and pyridine-d5 at various temperature intervals within the range of 278–343 K. From the temperature dependence of the NMR spectra of the amide protons, the free energy of activation, ΔG, for hindered rotation about the central C? N bond was determined by means of total line shape analysis in the four solvents DMSO, MPK and the two dioxanes. Observed values of ΔG (298 K) (72.7 in DMSO, 70.1 in MPK, 70.0 in 1,3-dioxane and 70.1 kJ mol1 in 1,4-dioxane) were not very sensitive to the choice of solvent or concentration. The concentration dependence of the internal chemical shift between the amide protons was studied in MPK, D2O, acetonitrile and pyridine-d5. The free energy of activation and the internal chemical shift are discussed on the basis of solvent-amide and amide–amide specific hydrogen bonding interactions, and in comparison to the results of molecular orbital calculations.  相似文献   

16.
17.
Two stereoisomers of the title compound are observed by H NMR at 10°. Their spectra coalesce at higher temperature (10°-90°). The equilibrium and rate constants K and k, strongly dependent on the solvent used (1,4-dioxane, tetrahydrofuran, acetone, chloroform); typical values for these parameters and the related thermodynamic functions are: K(25°)= 0.170; k(25°)=23.2s?1; ΔHR and ΔH=4.94 and 17.9 kcal.mol.?1; ΔSR and ΔS =13.1 and 7.7 e.u, in a 0.2 molar solution in 1,4-dioxane. The two isomers are shown to result from a hindered rotation around the aryl-to-nitrogen bond, presumably due to a direct resonance effect between the amide and nitro groups. The more abundant isomer was assigned a planar molecular structure in which the O atom of the amide group is close to the S atom of the thiophen ring, presumably on account of an electrostatic interaction between these two atoms which bear partial electrical charges of opposite sign.  相似文献   

18.
Summary Headspace-gas chromatography was used to determine the contents of toxic 1,4-dioxane, ethylene oxide and ethylene glycol in ethoxylated alcohols and alkylamines, and in commercial cosmetics and washing products. A Permaphase PEG capillary column was used for the determination of 1,4-dioxane and ethylene oxide and a DB-17 column for ethylene glycol determination. Dimethylformamide was used as the solvent in the determination of 1,4-dioxane and ethylene oxide, and undecanol in the case of ethylene glycol. The detection limits for ethylene oxide, 1,4-dioxane and ethylene glycol are 1,2 and 10 μg·g−1, respectively.  相似文献   

19.
Hypervalent organic ammonium radicals were generated by collisional neutralization with dimethyl disulfide of protonated 1,4-diazabicyclo[2.2.2]octane (1H+), N,N′-dimethylpiperazine (2H+) and N-methylpiperazine (3H+). The radicals dissociated completely on the 5.1 μs time-scale. Radical 1H underwent competitive N−H and N−C bond dissociations producing 1,4-diazabicyclo[2.2.2]octane and small ring fragments. Dissociations of radical 2H proceeded by N−H bond dissociation and ring cleavage, whereas N−CH3 bond cleavage was less frequent. Radical 3H underwent N−H, N−CH3 and N−Cring bond cleavages followed by post-reionization dissociations of the formed cations. The pattern of bond dissociations in the hypervalent ammonium radicals derived from six-membered nitrogen heterocycles is similar to those of aliphatic ammonium radicals. © 1997 John Wiley & Sons, Ltd.  相似文献   

20.
The IR spectra of six monosubstituted and of four 2,6-disubstituted 1,4-dioxanes have been studied in the 650–1800cm–1 region. The assignment of the bands due to the vibrations of the 1,4-dioxane ring and to the deformation vibrations of the methylene groups of the ring is given. The appearance of a whole series of new absorption bands on passing from unsubstituted 1,4-dioxane to its derivatives is explained by the change in the symmetry of the molecule and the removal of the prohibition from the vibrations previously inactive in the IR spectra connected with this reduction in symmetry. It is proposed to use, in order to confirm the presence of a 1,4-dioxane ring in a molecule from the results of IR spectroscopy, not only the 1126-cm–1 band but the whole group of bands lying in the frequency ranges 800–950, 1000–1150, and 1200–1300 cm–1.For part I, see [3].  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