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.
The results of a thermodynamic and kinetic investigation on the homolytic reactivity of 3,3'-di-tert-butyl-5,5'-dimethyl(1,1'-biphenyl)-2,2'-diol (1) are reported. EPR studies of the equilibration between 1, 2,4,6-trimethylphenol, and the corresponding radicals obtained by abstraction of a hydroxylic hydrogen allowed us to determine the OH bond dissociation energy (BDE) of investigated bisphenol as 83.10 kcal/mol. This value is considerably larger than that reported for the structurally related 2,6-di-tert-butyl-4-methylphenol (BHT), i.e., 81.02 kcal/mol. Absolute rate constants for the reaction of 1 with alkyl, alkoxyl, and peroxyl radicals, at or nearly to room temperature, were also determined by competition kinetics in the first two cases and by autoxidation studies under controlled conditions in the last one. The experimental data indicate that this bisphenol is a moderately efficient antioxidant and polymerization inhibitor.  相似文献   

2.
Summary Specifically labelled [1-13C ]salicylic acid has been made by a logical route involving the thermolysis of basic copper(II) benzoates, and the position of labelling established from the spin-spin coupling pattern of the product, p-hydroxybenzoic acid, of isomerisation. The results reinforce findings on the origin of the carbon atoms of chelated oxalate derived from chelated salicylate.Part X, see ref. 1.  相似文献   

3.
Thermochemical studies on the thioproline   总被引:3,自引:0,他引:3  
The combustion energy of thioproline was determined by the precision rotating-bomb calorimeter at 298.15 K to be Δc U= –2469.30±1.44 kJ mol–1. From the results and other auxiliary quantities, the standard molar enthalpy of combustion and the standard molar enthalpy of formation of thioproline were calculated to be Δc H m θC4H7NO2S, (s), 298.15 K= –2469.92±1.44 kJ mol–1 and Δf H m θC4H7NO2S, (s), 298.15K= –401.33±1.54 kJ mol–1.  相似文献   

4.
The oxidation of polymers is characterized by the generation of free radicals, and the kinetic parameters depicting the progress of oxidation are specific for each material structure and formulation. The durability of materials depends intrinsically on the molecular structure, and degradation mechanism influences the long-term stability of products. The intimate transformations of macromolecules can be reliably characterized by chemiluminescence (CL), which depicts the evolution of oxidation state by the reaction of free radicals with molecular oxygen. The isothermal and nonisothermal CL spectra are complementary proofs for the interpretation of oxidation behavior involving hydroperoxide as the initiators of oxidation. The degradation of PLA takes place by decomposition and fragmentation to lactide. The values of five kinetic parameters (initial CL intensity, CL intensity at the first peak, temperature corresponding to the first peak, oxidation induction time and oxidation rate) obtained for oxidative degradation of PLA lead to the activation energies ranged between 49 and 99 kJ mol?1. The evolution of thermal degradation for poly(lactic acid) is an excellent example for the explanation of the decay fate of radical intermediates.  相似文献   

5.
The thermal degradation of polyacrylic carboxylic and polystyrene sulfonic cationites was investigated using thermal analysis (TG) combined with Scanning Electron Microscopy (SEM). Fourier Transform Infrared Spectroscopy (FTIR) was used to characterize the resins degradation steps. The carboxylic cationite undergoes degradation through dehydration forming polyanhydrides, decomposition of polyanhydrides through decarboxylation with elimination of CO2 and CO. The sulfonic cationite undergoes degradation through dehydration, followed by decomposition of sulfonic acid functional groups liberating SO2. It was observed that strong acid (−SO3H+) cationite shows small mass loss of 55%, as against 88% mass loss shown by low-acidity carboxylic cationite. The possible reason for small mass loss of sulfonic cationite is discussed. The text was submitted by the authors in English.  相似文献   

6.
The bromination and iodination of tetraphenyllead and the corresponding triphenyllead halides have been studied in a solution calorimeter and the enthalpies of formation of the crystalline solids Ph3PbBr, Ph2PbBr2, Ph3PbI, and Ph2PbI2 have been derived. Values for the gas phase have been calculated using enthalpies of sublimation derived from vapourpressure measurements. The PbBr and PbI bond energy terms have been derived.  相似文献   

7.
Mechanisms of the electron-induced degradation of three polymers utilized in aerospace applications (polyethylene (PE), polytetrafluoroethylene (PTFE), and polystyrene (PS)) were examined over a temperature range of 10 K to 300 K at ultra high vacuum conditions (~10(-11) Torr). These processes simulate the interaction of secondary electrons generated in the track of galactic cosmic ray particles in the near-Earth space environment with polymer material. The chemical alterations at the macromolecular level were monitored on-line and in situ by Fourier-transform infrared spectroscopy and mass spectrometry. These data yielded important information on the temperature dependent kinetics on the formation of, for instance, trans-vinylene groups (-CH=CH-) in PE, benzene (C(6)H(6)) production in PS, fluorinated trans-vinylene (-CF=CF-) and terminal vinyl (-CF=CF(2)) groups in PTFE together with molecular hydrogen release in PE and PS. Additional data on the radiation-induced development of unsaturated, conjugated bonds were collected via UV-vis spectroscopy. Temperature dependent G-values for trans-vinylene formation (G(-CH=CH-) ≈ 25-2.5 × 10(-4) units (100 eV)(-1) from 10-300 K) and molecular hydrogen evolution (G(H(2)) ≈ 8-80 × 10(-5) molecules (100 eV)(-1) from 10-300 K) for irradiated PE were calculated to quantify the degree of polymer degradation following electron irradiation. These values are typically two to three orders of magnitude less than G-values previously published for the irradiation of polymers with energetic particles of higher mass.  相似文献   

8.
9.
10.
The 351.1 nm photoelectron spectrum of 1-pyrazolide anion has been measured. The 1-pyrazolide ion is produced by hydroxide (HO(-)) deprotonation of pyrazole in a flowing afterglow ion source. The electron affinity (EA) of the 1-pyrazolyl radical has been determined to be 2.938 +/- 0.005 eV. The angular dependence of the photoelectrons indicates near-degeneracy of low-lying states of 1-pyrazolyl. The vibronic feature of the spectrum suggests significant nonadiabatic effects in these electronic states. The gas phase acidity of pyrazole has been determined using a flowing afterglow-selected ion flow tube; Delta(acid)G(298) = 346.4 +/- 0.3 kcal mol(-1) and Delta(acid)H(298) = 353.6 +/- 0.4 kcal mol(-1). The N-H bond dissociation energy (BDE) of pyrazole is derived to be D(0)(pyrazole, N-H) = 106.4 +/- 0.4 kcal mol(-1) from the EA and the acidity using a thermochemical cycle. In addition to 1-pyrazolide, the photoelectron spectrum demonstrates that HO(-) deprotonates pyrazole at the C5 position to generate a minor amount of 5-pyrazolide anion. The photoelectron spectrum of 5-pyrazolide has been successfully reproduced by a Franck-Condon (FC) simulation based on the optimized geometries and the normal modes obtained from B3LYP/6-311++G(d,p) electronic structure calculations. The EA of the 5-pyrazolyl radical is 2.104 +/- 0.005 eV. The spectrum exhibits an extensive vibrational progression for an in-plane CCN bending mode, which indicates a substantial difference in the CCN angle between the electronic ground states of 5-pyrazolide and 5-pyrazolyl. Fundamental vibrational frequencies of 890 +/- 15, 1110 +/- 35, and 1345 +/- 30 cm(-1) have been assigned for the in-plane CCN bending mode and two in-plane bond-stretching modes, respectively, of X (2)A' 5-pyrazolyl. The physical properties of the pyrazole system are compared to the isoelectronic systems, pyrrole and imidazole.  相似文献   

11.
The calcination is one of the important steps in the preparation of nanostructured WC and WC–Co powders from chemically co-precipitated W–Co precursors. In the present paper, the processing of the precursor prepared by co-precipitation from cobalt(II) hydroxide and ammonium paratungstate is reported. The precursor was calcined at different temperatures under various atmospheres. The resulting powders were investigated using scanning electron microscopy, X-ray diffraction, thermogravimetric analysis (TGA), differential TGA, BET, and mass spectrographic analysis techniques. The results showed that the precursor decomposes differently by calcination under different atmospheres, i.e. nitrogen, air, and helium–5% oxygen. Different decomposition behaviour was observed when the calcination was carried out at medium temperature in flowing and non-flowing air. Nanostructured powders with particle size of around 90 nm can be obtained after calcination. The precursor decomposes into CoWO4 and WO3 oxides at 520 °C in air after a weight loss of 10.3%.  相似文献   

12.
Analogues of chorismate and isochorismate were designed and tested as potential inhibitors in the first inhibition study against a salicylate synthase.  相似文献   

13.
镨配合物的热化学及其对酵母菌作用的热动力学研究   总被引:2,自引:0,他引:2  
用六水合氯化镨、硫代脯氨酸(C4H7NO2S)和水杨酸(C7H6O3)合成了三元固体配合物[Pr(C7H5O3)2(C4H6NO2S)]-2H2O.根据盖斯定律设计一个热化学循环,用溶解-反应量热法研究得到合成反应的标准摩尔焓变为(133.70±1.02)kJ/mol,配合物298.15K时的标准摩尔生成焓为-(2909.3±3.2)kJ/mol.用TAMair微量热仪测定其在28.00℃时对粟酒裂殖酵母作用的产热曲线,进而算出在配合物作用下,酵母菌生长代谢的最大发热功率Pmax、速率常数κ、传代时间tG、抑制率I和半抑制浓度cI,50等热动力学参数.结果表明:稀土水杨酸硫代脯氨酸配合物在低浓度下对酵母菌有刺激作用,高浓度下为抑制作用,即稀土配合物对微生物的生长具有双向生物效应,也称为Hormesis效应.  相似文献   

14.
Nuclear magnetic resonance, infrared, and ultraviolet spectroscopy were used to elucidate the structure of oxidized polystyrene. To identify the nuclear magnetic resonance peaks of the degraded polystyrene, deuterated polystyrenes were synthesized and degraded. The primary structure present in the degraded polystyrene was found to be an aromatic carbonyl group. It was shown that this structure was formed regardless of the presence of ultraviolet light and was also present when polystyrene was degraded in carbon tetrachloride solution. When polystyrene was degraded in carbon tetrachloride solution, stable polymeric peroxides were formed in a concentration of 4 peroxide groups for every 1000 styrene units.  相似文献   

15.
Summary The equilibrium and kinetics of the complexation of CoII by folic acid have been studied at 25 °C, the ionic strength being regulated with KNO3 in the 5.6–7.3 pH range. Kinetic studies were performed using the stopped-flow method. Under our experimental conditions one process was observed; a reaction which took place within a few seconds. The results are consistent with the formation of a 11 complex between the reactants and a mechanism is proposed to account for the observed behaviour. Equilibrium constants for the CoII and folic acid system, as well as activation parameters, are reported.Member of CICPBA, Argentina.  相似文献   

16.
Pyrolysis-gas-liquid-chromatography (“thermocouple feedback” technique) has been used to study the thermal degradation kinetics of ionically-initiated and free-radical-initiated samples of polystyrene. Although mass-spectrometric measurements confirm that the pyrolysis products from large samples (1 mg) contain oligomers up to at least hexamer in addition to monomer, only monomer is detected when small thin samples (0.1 μg, 102–105 Å) are used. This effect is not due to a sensitivity problem in detecting oligomers, nor to the incapacity of such compounds of limited volatility to elute from the GLC apparatus. In studying the kinetics of monomer evolution from thin films, initial work was concerned with the effect of film thickness and the limits of first-order behaviour. Then the specific rate of monomer evolution (kobs) was measured as a function of molecular weight for both types of sample at 723 K and 753 K; the results indicate that the pyrolysis mechanism involves both initiation at the chain-ends and initiation by random scission. Kinetic schemes involving mixed initiation have been proposed, and on this basis the results have been analysed to yield activation energies for scission and end-initiation for both types of sample. Comparison of the activation energies obtained with the quoted value for scission of a CC bond has shown that the depolymerization chain termination process cannot be second order and must be first order in the concentration of long chain radicals. The experimental results also indicate that the ionically-initiated polystyrenes are more stable than free-radical-initiated samples of comparable molecular weight. Possible initiation sites have been discussed with reference to the samples examined and to previous published studies. Several mechanisms leading to first order termination have been proposed; it is suggested that the most probable process involves intramolecular transfer with subsequent scission to give an oligomer radical which is small enough to diffuse readily from the system without further reaction.  相似文献   

17.
Molecular weight change studies have shown that the thermal degradation of random copolymers of styrene — namely HIPS, SAN, and ABS-at low temperatures and in air involves random chain scission. The dominant process in the degradation of HIPS is random chain scission due to weak links, whereas in SAN it is intermolecular chain transfer. In ABS, the degradation is initially random scission due to weak links and then mainly intermolecular chain transfer. The infrared spectra show that during degradation the labile weak links are attacked by oxygen and peroxidic free radicals are produced. Via hydrogen abstraction or autoxidation of olefinic links, these free radicals are responsible for the formation of aliphatic ketonic or peroxyester structures, and for isomerization and cyclization. The activation energies of overall degradation of HIPS, SAN, and ABS are 134, 142, and 92 kJ.mol–1 respectively.Part of the PhD dissertation of Mrs. Jaya Nambiar, University of Gorakhpur, Gorakhpur-273001, 1980.  相似文献   

18.
The 351.1 nm photoelectron spectra of the N-methyl-5-pyrazolide anion and the N-methyl-5-imidazolide anion are reported. The photoelectron spectra of both isomers display extended vibrational progressions in the X2A' ground states of the corresponding radicals that are well reproduced by Franck-Condon simulations, based on the results of B3LYP/6-311++G(d,p) calculations. The electron affinities of the N-methyl-5-pyrazolyl radical and the N-methyl-5-imidazolyl radical are 2.054 +/- 0.006 eV and 1.987 +/- 0.008 eV, respectively. Broad vibronic features of the A(2)A' ' states are also observed in the spectra. The gas-phase acidities of N-methylpyrazole and N-methylimidazole are determined from measurements of proton-transfer rate constants using a flowing afterglow-selected ion flow tube instrument. The acidity of N-methylpyrazole is measured to be Delta(acid)G(298) = 376.9 +/- 0.7 kcal mol(-1) and Delta(acid)H(298) = 384.0 +/- 0.7 kcal mol(-1), whereas the acidity of N-methylimidazole is determined to be Delta(acid)G(298) = 380.2 +/- 1.0 kcal mol(-1) and Delta(acid)H(298)= 388.1 +/- 1.0 kcal mol(-1). The gas-phase acidities are combined with the electron affinities in a negative ion thermochemical cycle to determine the C5-H bond dissociation energies, D(0)(C5-H, N-methylpyrazole) = 116.4 +/- 0.7 kcal mol(-1) and D(0)(C5-H, N-methylimidazole) = 119.0 +/- 1.0 kcal mol(-1). The bond strengths reported here are consistent with previously reported bond strengths of pyrazole and imidazole; however, the error bars are significantly reduced.  相似文献   

19.
A kinetic method for the simultaneous determination of chlorpyrifos and carbaryl is described. The method is based on the degradation of these compounds in alkaline hydrogen peroxide medium by measuring the absorbance at two different reaction times and at two different wavelengths. A homemade stopped-flow system coupled with a diode-array detector and the proportional equation method was used. Resolution was achieved over chlorpyrifos: carbaryl ratios ranging from 61 to 14, with relative errors of 4 and 3%, respectively. The method was successfully applied to the analysis of a commercial formulation.  相似文献   

20.
The phase diagrams of the binary systems of picric acid with naphthalene, anthracene and phenanthrene, and of -naphthol withp-toluidine, determined by the thaw-melt method, show the formation of a molecular complex and two eutectics in each system. The heats of the pure components, eutectics and molecular complexes were determined by differential scanning calorimetry. Comparison of the experimental heats of fusion with the theoretical values calculated via the mixture law suggests cluster formation in the melts. The entropy of fusion, enthalpy of mixing and excess thermodynamic functions were also calculated from the heat of fusion data.
Zusammenfassung Die Phasendiagramme der binären Systeme von Pikrinsäure mit Naphthalin, Anthrazen und Phenanthren sowie von -Naphthol mit p-Aminotoluol wurden bestimmt. In jedem System zeigt sich die Bildung eines Molekülkomplexes sowie je zwei Eutektika. Mittels DSC wurden die Schmelzwärmen der reinen Komponenten, der eutektischen Mischungen und der Molekülkomplexe ermitteis. Der Vergleich der experimentell ermittelten Schmelzwärmen mit den mittels der Mischungsregel errechneten Werten läßt auf eine Clusterbildung in der Schmelze schließen. Weiterhin wurden aus den Schmelzwärmen auch Werte für Schmelzentropie und Mischungsenthalpie errechnet.

— , , - . , . , . , .


Thanks are due to Prof. I. S. Ahuja, Head, Chemistry Department, Banaras Hindu University, for providing research facilities. Thanks are also due to CSIR, New Delhi, for financial assistance.  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