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 78 毫秒
1.
The heterogeneous polymerization of acrylonitrile photoinitiated by naphthalene is influenced by the polarity of the reaction medium. The rate of initiation increases with the increasing dielectric strength of the reaction medium. A similar trend is observable for Stern–Volmer constants of naphthalene fluorescence quenching by acrylonitrile. The ratio kp/kt1/2 of the rate constant for propagation and termination reactions is not influenced by a change in the polarity of the reaction medium. The effect of viscosity on the value of kp/kt1/2 known for polymerization in a homogeneous medium was not observed in the reaction systems studied.  相似文献   

2.
Kinetic studies on the Hinsberg condensation were performed trying to improve yields and achieve regio-selectivity in the attainment of benzene-substituted 3-methylquinoxalin-2(1H)-ones. The course of the reactions between o-phenylenediamine (o-PDA) and substituted o-PDA with pyruvic acid ( 2a ) or ethyl pyruvate ( 2b ) were followed by uv spectrophotometry at different pH values. The formation of 3-methylquinoxalin-2(1H)-one ( 6a ) was improved using sulphuric acid-water mixtures, in which the reaction proceeded by a different mechanism. 3-Methyl-7-methoxyquinoxalin-2(1H)-one ( 7b ) was regioselectively synthesized independently of the pH of the reaction media. Reaction of 2-amino-4-methylamine ( 1c ) with 2a or 2b led to a mixture of 6 and 7-quinoxalinone isomers, 6c and 7c , while 2-amino-4-nitroaniline ( 1d ) and 2,4-diaminoaniline ( 1e ) with 2a or 2b did not afford the heterocycle. In every case reactions with 2a were 100–1000 times faster than those with 2b . Mechanisms are proposed trying to account for the experimental results.  相似文献   

3.
The kinetics of the radiation-induced polymerization of vinyl chloride in the presence of precipitants has been successfully described by a one-parameter equation as follows, where ?0 is initial monomer volume fraction, X is conversion, t is time, and k is reaction constant. The equation was confirmed for extensive conditions of temperatures and monomer concentrations in the case of polymerization in methanol. The degree of polymerization was related with the reaction constant k, initial monomer volume fraction ?0, monomer chain transfer constant Cm, conversion X, and the initiation rate I as follows, The factors which determine the value of the reaction constant k were elucidated through measurements of the reaction constant k and the degree of polymerization DP n.  相似文献   

4.
The kinetic feature of the anionic polymerization of N-PMI was investigated in THF. The polymerization system initiated with lithium tert-butoxide was revealed to be so-called “slow-initiation” system. The rate constant of the initiation reaction, ki, was obtained to be 4.2 × 10?3 (L mol?1 s?1) at ?72°C. The apparent rate constants of the propagation reaction, k, at ?72°C were individually obtained from each slope of the first-order plots in the later stages of the polymerizations for four different initiator concentrations. Each k is fairly close to that of initiation rate around 10?3. The propagation reaction was concluded to be dominated by ion-pair mechanism from the analysis of the kinetic data and the results of the addition effects of crown ether and common salt.  相似文献   

5.
The formation of 1-substituted-3-hydroxy-1H-indazole and 1-substituted-benzimidazolin-2-one derivatives by thermal reaction of N-substituted-N-arylcarbamoyl azides was examined.  相似文献   

6.
The preparation of aminopterin analogs via 6-halomethylpteridine intermediates either were unsuccessful or gave low yields of pteridines by reaction of 2,4,5,6-tetraaminopyrimidine (1) with 2,3-dibromopropionaldehyde (2) , 1,1,3-trichloroacetone (11) , and 1,1,3-tribromoacetone (12) , respectively. Similarly, the preparation of a 6-formylpteridine was unsuccessful by the alkylation of 1 with bromomalonaldehyde and by the addition of the 5 -acetyl derivative of 1 to 2-bromopropenal (17). In contrast, the oxidation of 2,4-diaminopteridine-6-methanol (23) with N,N′-dicyclohexylcarbodiimide gave the corresponding 6-formylpteridine 20. In related work, a mechanism for the Waller reaction was suggested by the identification of some of the products resulting from the reaction of 2-bromopropenal with p-aminobenzoyl derivatives.  相似文献   

7.
Part I: Kinetic data for the static system silane pyrolysis (from 640–703 K, 60–400 torr) are presented. For conversion from 3–30%, first-order kinetics are obtained, with silane loss rates equal to half the hydrogen formation rates. At conversions greater than 40%, rate inhibition attributable to the back reaction of hydrogen with silylene occurs. Overall reaction rates are not surface sensitive, but disilane and trisilane yield maxima under some conditions are. A nonchain mechanism capable of describing quantitatively all stages of the silane pyrolysis is proposed. Post 1.0% initiation is both homogeneous (gas phase) and heterogeneous (on the walls), and reaction intermediates are silylenes and disilenes. Free radicals are not involved at any stage of the reaction. Rate data at high conversions and with added hydrogen provide kinetics for the addition of silylene to hydrogen [reaction (?1)1] relative to its addition to silane [reaction (2)]: k?1,/k2 = 10?0.65 × e?3200 cal/RT. With E2 = 1300 cal, this gives a high pressure activation energy for silylene insertion into hydrogen of E?1 = 8200 cal. Part II: An analysis is made of each rate constant of the silane mechanism and the modeling results are compared with experimental results. Agreement is excellent. It is concluded that the dominant sink reaction for silylene intermediates is 1,2—H2 elimination from disilane (followed by Si2H4 polymerization and wall deposition). The model is in accord with slow isomerization between disilene and silylsilylene and near exclusive 1,2—H2 elimination from Si2H6. It is also concluded that disilene is about 10 kcal/mol more stable than silylsilylene and that the activation energy for isomerization of silylsilylene to disilene is greater than 26 kcal/mol.  相似文献   

8.
The radiation-initiated graft polymerization of 4-vinylpyridine to high-density polyethylene was studied over a wide range of reaction conditions of radiation intensity I, monomer concentration M1, and polymer film thickness L. The conditions included both diffusion-free and diffusion-controlled graft polymerizations. The results corroborate our previous theoretical predictions on the effect of I, M1, and L on the experimental grafting rate. The grafting rate is inverse first order in L for diffusion-controlled reaction and independent of L for diffusion-free reaction. The dependence of grafting rate on radiation intensity decreases from 1/2 to 1/4 order for diffusion-controlled reaction. Diffusion control results in a decrease in the dependence of rate on monomer concentration. The observed decrease is somewhat greater than theoretically predicted.  相似文献   

9.
The reaction of cyclohexyl isocyanate with phenylglycidyl ether was selected as model reaction for the synthesis of cycloaliphatic isocyanate-based poly(2-oxazolidone)s. The selectivity of AlCl3 and AlCl3-triphenylphosphine oxide (AlCl3–TPPO) and AlCl3-hexamethylphosphoric triamide (AlCl3–HMPA) complexes was studied for 2-oxazolidone formation. The reaction products were identified by means of the melting point, 1H-NMR, and IR spectroscopy. The kinetics of the model reaction was studied using AlCl3-TPPO in o-dichlorobenzene at 120 and 140°C.  相似文献   

10.
This paper reports our results for the direct experimental determination of the equilibrium constant for the hydrogen-isotope-exchange reaction, 1/2D2(g)+HCl(hexOH)=1/2H2(g)+DCl(hexOD), where hexOH isn-hexanol and hexOD isn-hexanol with deuterium substitution in the alcohol function. The reaction was studied in electrochemical double cells without liquid junction for which the net cell reaction is the above isotope-exchange reaction. The experimentally determined value of ε° (296.0°K) for this cell is 4.03±0.95 mV (strong electrolyte standard states, mole-fraction composition scale); the value of the equilibrium constant for the reaction is 1.17±0.05. The contributions of isotope-exchange and transfer effects to the magnitude of the standard Gibbs energy change for the above reaction and for the analogous reaction 1/2D2(g)+HCl(aq)=DCl(daq)+1/2H2(g) are considered. Our results support the conclusion of Heinzinger and Weston that the formulation of the solvated proton in water as H3O+, as opposed to H9O4 +, is sufficient for the interpretation of the thermodynamics of hydrogen-isotope-exchange reactions in water. We also find that the formulation of the solvated proton inn-hexanol as ROH 2 + is sufficient for the interpretation of our results on the thermodynamics of hydrogen-isotope-exchange inn-hexanol.  相似文献   

11.
By pyrolyzing di-t-butyl peroxide over the temperature range of 405–450 K in the presence of hexafluoroacetone the kinetics of the addition reaction (1), CH3 + (CF3)2CO→; (CF3)2C(?)CH3, have been studied. Detailed analyses have shown that the principal product of the adduct radical, (CF3)2C(?)CH3, is CF3COCH3 from reaction (2), (CF3)2C(?)CH3 → CF3COCH3 + CF3. The rate constant of the addition reaction was determined to be k1(dm3/mol·s) = (1.1 ± 4.0) + 109 exp(-(3680 ± 480)/T) over the temperature range 405–450 K, based on the value k3 = 2.2 × 1010 dm3/mol·s for reaction (3), 2CH3 → C2H6. The results are discussed in relation to existing data for radical additions to carbonyl groups.  相似文献   

12.
Zusammenfassung Die Reaktion von 1-Nitroso-2-naphthol mit 1-und 2-Naphthol sowie die Reaktion von 2-Nitroso-1-naphthol mit 2-Naphthol in Äthanol und in Äther bei Anwesenheit von HNO3 gibt 5H-Dibenzo[a,j]phenoxazon-(5) (I), 5H-Dibenzo[a,j]phenoxazon-(5)-14-oxid (II), 5H-Dibenzo[a,h]phenoxazon-(5) (III) sowie 5H-Dibenzo[a,h]phenoxazon-(5)-14-oxid (IV). Es wurde ein Reaktionsmechanismus vorgeschlagen und die Konstitution der hergestellten Verbindungen spektrophotometrisch und potentiometrisch bestimmt.
The reaction of 1-nitroso-2-naphthol with 2-and 2-naphthol and the reaction of 2-nitroso-1-naphthol with 2-naphthol in ethanol or ether in the presence of nitric acid have been studied. The main reaction products isolated were the dibenzophenoxazones I–IV. The reaction mechanism for their formation is proposed.


Mit 4 Abbildungen  相似文献   

13.
ABSTRACT

The Horner-Wadsworth-Emmons (HWE) reaction was performed on methyl 3,6-di-O-benzoyl-2-deoxy-α-D-glycero-hex-2-enopyranosid-4-ulose (1) with the potassium enolates of dimethyl [(methoxycarbonyl)methyl]phosphonate (2) or diethyl [(ethoxycarbonyl) methyl]phosphonate (3) under different conditions (metallic cation and solvent) in order to study regio- and stereochemical aspects of the reaction. In the presence of lithium ions, no reaction took place. When sodium enolates were employed, 1,2-addition was the main reaction in chelating solvents, whereas the 1,4-adduct is favoured in the less polar, non chelating toluene. Only 1,2-addition was observed with potassium enolates. Evidence of phosphonate-phosphate rearrangements through five membered cyclic intermediates is described.  相似文献   

14.
The rate of the thermal cycloaddition of ethylene to cis and trans butene-2 has been measured at 693°K and at pressures of about 12 atmospheres. The ratio of trans- to cis-1,2-dimethylcyclobutane from the reaction of trans-butene-2 with ethylene was 5.1, obtained from the initial rates of formation of the products. Similarly, the ratio of cis- to trans-1,2-dimethyl-cyclobutane from the reaction of cis-butene-2 with ethylene was 2.8. The results show that the cycloaddition reactions are the reverse of the decomposition reactions of the dimethyl-cyclobutanes and may be interpreted in terms of a biradical intermediate. Several ratios of rate constants have been measured as well as the rate constants for the reaction of the olefins to form the intermediate biradical.  相似文献   

15.
First ever non-deformylating transdiazotization of acylacetaldehydes was achieved: the reactions of 2-azido-l-ethylpyridinium tetrafluoroborate ( 4 ) with acylacetaldehydes 3 proceeded partially without deformylation to yield 16 new α-diazo-β-oxoaldehydes 1 along with diazomethyl ketones 2 , especially in the presence of NaOAc (Scheme 1, Tables 1 and 2). The product distribution was substituent-dependent and could be correlated quantitatively. This new diazotization reaction appears as an alternative, direct, and more general method for the synthesis of these diazooxoaldehydes. α-Oxocycloalkanecarbaldehydes 5 gave only traces (if any) of α-diazocycloalkanones 7 , and rearrangement products 6 were isolated (Scheme 2). Mechanisms of the reactions are discussed (Schemes 4 and 5).  相似文献   

16.
The rate of polymerization of thiophene, at concentrations of catalyst (SnCl4), and thiophene of the same order as was subsequently used in studying the reaction between thiophene and di(chloromethyl)benzene, is of the order of 10-2%/hr at 30°C. There is no significant self-condensation of DCMB under the same conditions. Since the reaction between thiophene and DCMB is complete at 30°C in minutes rather than hours, it is assumed that self-condensation of thiophene or DCMB during the reaction between them will be negligible and should not influence the course of the reaction or the structure of the resulting polymer. Reaction at 30°C is much too fast for convenient study. A temperature of 0°C is more appropriate and was used in subsequent kinetic work. The first two products of the condensation of p-di(chloromethyl)benzene (DCMB) with thiophene have been identified by a combination of mass, infrared, and nuclear magnetic resonance spectroscopy as thenylchloromethylbenzene (TCMB) and dithenylbenzene (DTB). DCMB, TCMB, and DTB have been estimated quantitatively during the course of the reaction by gas-liquid chromatography (GLC), and it has been established that the rates of each of the two reaction steps is first-order with respect to the chloro compound (DCMB and TCMB respectively), thiophene, and SnCl4. Rate constants for these two consecutive reactions were calculated to be k1 = 2.79 × 10-4l.2/mole2-sec, k2 = 6.37 × 10-3l.2/mole2-sec; the corresponding energies of activation are E1 = 7.93 kcal/mole, E2 = 7°67 kcal/mole. These rate constants are appreciably higher than values previously obtained for the corresponding DCMB–benzene reactions.  相似文献   

17.
Radical polymerization studies on diallyl oxalate (DAO), diallyl malonate (DAM), diallyl succinate (DASu), diallyl adipate (DAA), and diallyl sebacate (DAS) have been conducted kinetically from the standpoint of cyclopolymerization. Benzoyl peroxide was employed as the initiator. The initial overall rate of polymerization, Rp was not proportional to the square root or the first power of the initiator concentration, [I]. But Rp/[I]1/2 and [I]1/2 bore a linear relationship, provided the monomer concentration was kept constant. The residual unsaturation of the polymers decreased with decreasing monomer concentration. The ratio of the rate constant of the unimolecular cyclization reaction to that of the bimolecular propagation reaction of the uncyclized radical, Kc, was evaluated from the above relationship between the residual unsaturation and the monomer concentration at 60°C. The Kc values obtained were 3.6, 3.2, 2.8, 2.5, and 1.2 mole/l. for DAO, DAM, DASu, DAA, and DAS, respectively. The overall activation energies of polymerization were found to be 21.1 (DAO), 24.2 (DAM), 21.7 (DASu), 22.0 (DAA), and 22.2 (DAS) kcal/mole.  相似文献   

18.
By the reaction of anthranilic hydrazide 1 with cis-2-(p-methylbenzoyl)-1-cyclohexanecarboxylic acid 2a or diendo-3-(p-methylbenzoyl)bicyclo[2.2.1]heptane-2-carboxylic acid 2b , fused tetra- and pentacyclic ring systems 3a, b were prepared, trans-2-Amino-1-cyclohexanecar-bohydrazide 4b was reacted with 3-(p-chlorobenzoyl)propionic acid 5 to yield the pyridazino[6,1-b]quinazolinone 6 . From the reaction of cis-2-amino-1-cyclohexanecarbohydrazide 4a with 2a , three isomeric partially saturated 8H-phthalazino[1,2-b]quinazolin-8-ones 7a-c were formed. The reaction of diexo-2-aminobicyclo[2.2.1]heptane-3-carbohydrazide 4c and 2a furnished the pentacyclic derivatives 8 and 9 containing a 3-aryl-4,5-dihydropyridazine or 3-arylhexahydropyridazine ring C with cis annelated C/D rings. The formation of 8 and 9 involving different ring systems can be rationalized by two reaction pathways: (i) in the bislactam 9 the carboxyl group acylates the hydrazide, while (ii) in 8 it forms a pyridazine ring with the cyclic amino group by cyclocondensation. The structures of the products were elucidated by 1H and 13C nmr methods, including DEPT, DNOE and 2D-HSC measurements.  相似文献   

19.
The reaction of N1-acyl-2-ethoxycarbonylacetamidrazones with ethyl ethoxymethylenecyanoacetate (EMCA) has been examined. The acetamidrazone 1a reacts with EMCA in boiling dimethyl sulphoxide to give the 1,2,4-triazolo[4,3-a]pyridin-3(2H)-one 2 in excellent yield. Similarly from the amidrazones 1b-h the 1,2,4-triazolo[4,3-a]pyridines 3b-h were obtained. When the reaction between the amidrazones and EMCA was performed in ethanolic solution, the 6-(2-acylhydrazino)-pyridines 4 were isolated. Ring closure of 4 afforded the 1,2,4-triazolo[4,3-a]pyridines.  相似文献   

20.
The Bernthsen reaction between N-1-naphthyl-2-naphthylamine and 2-methylbutanoic acid and its anhydride at 200–230° for seven hours gives a low yield of 12- or 13-s-butyldibenz[a,h]acridine, instead of the expected 14-isomer. The parent molecule dibenz[a,h]acridine results from the same reaction conducted at 270° for thirteen hours. It is suggested that alkyl migration may have occurred in some other cases where the Bernthsen reaction was reported to yield 14-alkyldibenz[a,h]acridines.  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