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 15 毫秒
1.
4-Toluenesulfonyl isocyanate cyclized with 1,2-epoxy-3-phenoxypropane and 2,3-epoxypropyl 4-methoxyphenyl ether, respectively, to give 3-(4-toluenesulfonyl)-5-phenoxymethylene-2-oxazolidone ( I ) and 3-(4-toluenesulfonyl)-5-(4-methoxyphenoxymethylene)-2-oxazolidone ( II ). Compounds I and II were hydrolyzed in 2 M sodium hydroxide solution to the corresponding uncyclized hydroxy amides, VII and VIII. Compound I was remarkably stable toward 6 M hydrochloric acid and amines. Styrene oxide, 1,2-epoxybutane, 3-chloro-1,2-epoxypropane, and 1-methoxy-2-methylpropylene oxide reacted with the isocyanate to afford 3-(4-toluene-sulfonyl)-4-phenyl-2-oxazolidone (III), 3-(4-toluenesulfonyl)-4-ethyl-2-oxazolidone ( IV ), 3-(4-toluenesulfonyl)-5-chloromethyl-2-oxazolidone ( V ), and 3-(4-toluenesulfonyl)-4,4-dimethyl-5-methoxy-2-oxazolidone ( VI ), respectively. The yield of VI was constant over a temperature range of 25–90°.  相似文献   

2.
Poly-2-oxazolidones of high molecular weight have been synthesized from diisocy anates and diepoxides. The synthetic method was first developed for the model compound 3-phenyl-5-phenoxymethyl-2-oxazolidone prepared from phenyl isocyanate and phenyl glycidyl ether. Conditions found most suitable for a high yield involve slow addition of isocyanate to a solution of epoxide and catalyst, temperatures of 160°C or higher, and a catalytic amount of n-butoxylithium in 1-butanol. The same procedure was used to prepare high molecular weight poly-2-oxazolidones. The polymer structure was confirmed by infrared spectroscopy and elemental analysis. All polymers prepared were of sufficient molecular weight to be compression molded. The bulk mechanical properties characterize poly-2-oxazolidones as potentially useful engineering thermoplastics. A mechanism for oxazolidone formation is proposed.  相似文献   

3.
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   

4.
Oxiranes, such as glycidyl phenyl ether, benzyl glycidate, glycidyl methyl ether, and styrene oxide, were copolymerized with dicarboxylic anhydrides, such as succinic anhydride, phthalic anhydride, and maleic anhydride, by the action of an enzyme in a stepwise reaction to produce the corresponding polyesters containing some ether linkages having a maximum M w of 13 500. Oxiranes, such as glycidol and glycidyl phenyl ether, were also homopolymerized and copolymerized with other oxiranes by the enzyme to produce the corresponding polyethers.

Enzymatic polymerization of oxiranes and dicarboxylic anhydrides.  相似文献   


5.
Crystalline and amorphous polymers have been obtained from the polymerization of phenyl glycidyl ether in the presence of tertiary amines. The crystalline fraction is high melting and insoluble at room temperature. The amorphous fractions are soluble at room temperature and their molecular weights are found to be ~950 in benzene at 30°C. The yields of the crystalline fraction and the amorphous paste fraction decreased considerably with increasing the catalyst concentration and reaction temperature above 50°C. The yield of the liquid fraction, however, increased with increasing concentration of the catalyst and the reaction temperature. The x-ray diffraction analysis of the crystalline fraction shows that the fraction has 47–50% crystallinity and that its diffraction pattern is similar to that of poly(phenyl glycidyl ether) obtained by Noshay and Price. The infrared spectra of these fractions have been obtained in the region of 650–4000 cm.?1. These data are compared with those of polystyrene and poly(styrene oxide) and are used to make an assignment of the normal modes of the poly(phenyl glycidyl ether) molecule. On the basis of analyses of polystyrene and poly(styrene oxide), and a study of the combination bands, it has been possible to make a fairly satisfactory assignment of all of the benzene ring fundamentals of the CH2, CH, and skeletal modes.  相似文献   

6.
The insertion reaction of various epoxy compounds such as phenyl glycidyl ether (PGE), methyl glycidyl ether, butyl glycidyl ether (BGE), and styrene oxide into the phenyl ester linkage in the polymer chain was investigated using quaternary ammonium salts as catalysts in diglyme, anisole, sulfolane, o-dichlorobenzene, or DMSO at 100–150°C. The reaction of PGE with poly[4-(4-nitrobenzoyloxy)styrene] (polymer 1a ) proceeded almost quantitatively to give the corresponding polymer using tetrabutylammonium bromide (TBAB) as catalyst in diglyme at 100°C for 24 h. The reactions of BGE with poly(4-nitrophenyl methacrylate), and copolyarylate derived from 2,2-bis(4-hydroxyphenyl)-1,1,1,3,3,3-hexafluoropropane and iso- and terephthaloyl chlorides also produced the corresponding polymers with 86 and 89 mol% new structural units, respectively, using TBAB in sulfolane at 150°C for 24 h. Furthermore, it was found that the degree of insertion of the epoxy compound into the ester linkage in the polymer chain was affected by the kind of epoxy compound, reaction solvent, catalyst concentration, substituent group on the phenyl ester, and structure of the polymer. Chiral polymers were also synthesized with high degrees of insertion by the reaction of chiral epoxides such as (R)-1,2-epoxyhexane, (R)-1,2-epoxyheptane, and (R)-1,2-epoxydecane with polymer 1a and poly(2,4-dichlorophenyl methacrylate) using TBAB in diglyme at 120°C for 24 h.  相似文献   

7.
Copolymerizations of tritiated phenyl glycidyl ether with p- and m-methoxy, p-methyl, and p-chloro analogs, as well as with propylene oxide and epichlorohydrin are reported. The relative reactivities indicate greater Lewis acid character for triethylaluminum–water than for diethyl-zinc–water or ferric chloride–propylene oxide catalyst systems. The copolymerization of the PGE analogs is promoted by electron-donor groups and retarded by electron-withdrawing groups.  相似文献   

8.
High molecular weight, linear polyethers were prepared by polymerizing a series of ring-substituted phenyl glycidyl ethers by using the ferric chloride–propylene oxide and dibutylzinc–water catalyst systems. The α-naphthyl, β-naphthyl, p-phenylphenyl, the o-, m-, and p-methyl, and the o- and p-chlorophenyl polymers resemble the parent polymer in that they are readily crystallizable polyethers which have melting points above 170°C. The other substituted poly(phenyl glycidyl ethers), including the o- and p-isopropyl, p-tert-butyl, p-octyl, and 2,4,6-trichloro derivatives show much less tendency to crystallize and are lower melting. The x-ray and electron diffraction data established that poly(o-chlorophenyl glycidyl ether) crystallizes in an orthorhombic unit cell; data obtained in a parallel study of unsubstituted poly(phenyl glycidyl ether) did not allow assignment of a specific structure.  相似文献   

9.
The processes occurring during the modification of epoxy polymers by various polymorphic aluminum oxide modifications (γ-AlO(OH), γ-Al2O3, α-Al2O3) with epoxy groups were studied by the methods of IR Fourier spectroscopy, chemical analysis, and differential scanning calorimetry (DSC) by an example of a model compound (phenyl glycidyl ether). Two types of interactions were revealed: a direct chemical reaction of phenyl glycidyl ether with the surface hydroxy groups of alyminum oxide, and phenyl glycidyl ether homopolymerization. By processing by graphical method the data of chemical analysis on the diminishing in amount of epoxy groups in the course of the polycondensation reaction the value of activation energy 106–110 kJ mol−1 of the process of phenyl glycidyl ether interaction with aluminum γ-oxide was determined.  相似文献   

10.
A small quantity of 10-methylphenothiazine cation radical (MPT.+), electrochemically prepared and stocked in acetonitrile solution, initiated cationic polymerizations of n-butyl, t-butyl, and 2-methoxyethyl vinyl ethers and p-methoxystyrene, while no initiation occurred for phenyl vinyl ether, styrene, methyl methacrylate, and phenyl glycidyl ether. 1H-NMR studies of oligomers and low molecular weight compounds isolated from the reaction mixture for the polymerization of t-butyl vinyl ether in the presence of a small amount of D2O indicated that electron transfer from the monomer to MPT.+ was involved in the initiation step. 1H- and 13C-NMR and MO calculation implied that monomers with higher electron densities on the vinyl groups and with lower ionization potentials were more susceptible to the initiation of MPT.+. © 1994 John Wiley & Sons, Inc.  相似文献   

11.
Conclusions 3-Phenyl-5-phenoxymethyl-2-N-phenyliminooxazolidine, formed by the reaction of diphenyl-carbodiimide with phenyl glycidyl ether, reacts with an excess of the latter to give 3-phenyl-5-phenoxymethyl-2-oxazolidone.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 9, pp. 2152–2153, September, 1983.  相似文献   

12.
6-Aminopyrazolo[1,2-a][1,2,4]triazine-4,8-dione derivative 3 was obtained upon the reaction of the acid hydrazide derivative 2a with ethyl cyanoacetate. The reactions of 3 with several electrophiles such as aldehydes, isatin, acetic anhydride, phenyl isocyanate, benzoyl isothiocyanate, and p-toluenesulfonyl chloride were studied. The structures of the newly synthesized compounds were established on the basis of IR, 1H NMR, mass spectra, and elemental analyses. The antitumor activities of some selective compounds were examined against two cell lines as liver carcinoma cell line (HEPG-2) and human breast cancer cell line (MCF7).  相似文献   

13.
A number of derivatives of 5-(2-aminoethyl)bicyclo[2.2.1]hept-2-ene (85-90% of the endo isomer) were synthesized by reactions with p-nitrobenzenesulfonyl chloride, p-toluenesulfonyl chloride, benzoyl chloride, p-nitrobenzoyl chloride, benzyl isocyanate, m-tolyl isocyanate, phenyl isothiocyanate, and endic anhydride. By reactions of the resulting sulfon- and carboxamides with peroxyphthalic acid generated in situ from phthalic anhydride and 40-45% hydrogen peroxide the corresponding epoxy derivatives were obtained. These reactions were not accompanied by heterocyclization into azabrendane derivatives, which is typical of homologous N-(p-nitrophenylsulfonyl)-endo-5-aminomethylbicyclo[2.2.1]hept-2-ene.  相似文献   

14.
The preparation of 3-aminopropylcellulose from cyanoethylcellulose is readily achieved. Reduction of the cyano groups with borane-dimethyl sulfide in tetrahydrofuran or a borane-tetrahy-drofuran complex proceeds quantitatively in 3 h to a corresponding 3-aminopropylcellulose. The presence of primary amine functions is confirmed by spectroscopy and a positive ninhydrin test; the concentration of amino substituents, as ascertained by titration, ranged from 1.2 to 6.4 meq/g. Because the derivatives are neither soluble nor excessively swollen in water, applications as ion-exchange resins or chromatographic supports can be envisioned. Treatment of 3-aminopropyl-cellulose with acetyl chloride, phenyl isocyanate, or p-toluenesulfonyl isocyanate produced 3-acetamido-, 3-(N′-phenyluredo)-, or 3-(N′-p-toluenesulfonyluredo)-N-propylcellulose. Alkylation with methyl chloride yielded a water-soluble quaternary ammonium salt.  相似文献   

15.
《Tetrahedron: Asymmetry》1998,9(3):459-466
The recombinant epoxide hydrolase from Agrobacterium radiobacter AD1 was used to obtain enantiomerically pure epoxides by means of a kinetic resolution. Epoxides such as styrene oxide and various derivatives thereof and phenyl glycidyl ether were obtained in high enantiomeric excess and in reasonable yield. The enantioselectivity (E-value) of the resolution was calculated from progress curves for styrene oxide (E=16.2) and para-chlorostyrene oxide (E=32.2).  相似文献   

16.
Oxazolidone group-containing vinyl monomers, 4-(2-oxo-3-oxazolidinyl)methylstyrene (OS) and 4-[2-(2-oxo-3-oxazolidinyl)ethoxy]methylstyrene (OES), were synthesized and their polymerization and copolymerization behaviors with styrene (St), p-methoxystyrene (PMS), and m-hydroxystyrene (MHS) were investigated. OS was prepared in 70% yield by the reaction of 2-oxazolidone with p-chloromethylstyrene in the presence of sodium hydride. OES was obtained by the similar reaction of p-chloromethylstyrene with N-hydroxyethyl-2-oxazolidone which was prepared by the reaction of 2-oxazolidone with ethylenecarbonate. Homopolymerization of OS and OES afforded mainly gelled polymers, but also soluble polymers on high dilution. In the copolymerization with styrene derivatives, an alternating nature was suggested from the copolymerization parameters obtained by either the nonlinear least-squares analysis method or the Fineman–Ross method. The alternating copolymerizability decreased in the following order: MHS > PMS > St. Q?e values of OS and OES were calculated and demonstrated that OS and OES behaved as stronger electron-accepting monomers in the copolymerization with MHS than in those with St and PMS. The copolymerization behavior of OS (OES) with MHS was compared with those of 4-(2-oxo-1-pyrrolidinyl)methylstyrene (PS) and 4-[2-(2-oxo-1-pyrrolidinyl)ethoxy]methylstyrene (PES). From an IR study examining the shift of carbonyl absorption by addition of MHS, the interaction which contributed to the increase of the alternating copolymerizability in the copolymerization of OS (OES) with MHS was concluded to be based on hydrogen bonding. © 1993 John Wiley & Sons, Inc.  相似文献   

17.
Kinetic analysis of formulations based on glycerol diglycidyl ether and phenyl glycidyl ether were carried out in the presence of sulfonium salt as initiator at 35 mW cm?2using photo differential scanning calorimeter and the final conversion was found to increase with an increase in phenyl glycidyl ether content. The effects of formulation monomer ratios at three different temperatures were studied. The variations in the observed kinetic parameters can be related to increase in mobility of reactive species with temperature, distance of counter ion from the propagating cationic center, as well as extent of crosslinking reaction which controlled the course and duration of the reaction. The applicability of autocatalytic kinetic model was also evaluated and the system underwent early gelation and the activation energy decreased with an increase in phenyl glycidyl ether content. Analysis of stable photocured films containing glycerol diglycidyl ether and phenyl glycidyl ether showed better thermal stability than rigid films obtained with glycerol diglycidyl ether.  相似文献   

18.
Bis-aminimide compounds [bis-N, N,-dimethyl-N,-(2-hydroxypropyl)-amine-N,′-adipimide ( 1 ) and bis-trimethylamine adipimide ( 2 )] were found to exhibit different thermal decomposition behavior and polymerization efficiency for an epoxide (phenyl glycidyl ether, PGE). The thermal decomposition rate of 1 was much higher than that of 2. It seemed that hydrogen bonding enhanced the decomposition rate. Compound 1 was thermolyzed to give a diisocyanate and a tertiary aminoalcohol, which subsequently reacted with each other to give a urethane. When 2 was heated, the isocyanate generated from 2 remained unreacted. PGE reacted with thoseaminimides to give different products, depending on their thermolyzed products. Mixtures of diisocyanate, tertiary amine, and PGE were used in the model reactions, and the thermal reaction between the expected decomposition products of aminimides was investigated in the presence and absence of PGE. The amount of high molecular weight fraction in PGE + 2 is greater than that of PGE + 1 . In the former, the free isocyanate groups may act as a chain extender to give higher molecular weight fractions.  相似文献   

19.
Mono-aminimide compounds [N,N-dimethyl-N-(2-hydroxypropyl)-amine-N′-propionimide ( 1 ), trimethylamine valerimide ( 2 ), and trimethylamine benzimide ( 3 )] were found to exhibit different thermal decomposition behaviors and polymerization efficiencies for an epoxide (phenyl glycidyl ether, PGE). The thermal decomposition rate of aminimides at 150°C decreased in the order 1 > 2 > 3 . It seemed that hydrogen bonding enchanced the decomposition rate, and the resonance effect induced by the phenyl group suppressed the decomposition. 1 was thermolyzed to give isocyanate and tertiary aminoalcohol, which subsequently reacted with each other to give isocyanate and tertiary aminoalcohol, which subsequently reacted with each other to give urethane. When 2 was heatted, the isocyanurate generated from 2 remained intact. On heating of 3 , we observed the formation of triphenyl isocyanurate. PGE reacted with those aminimides and gave different products depending on their thermolyzed products. Equimolar mixtures of isocyanate, tertiary amine, PGE were used in the model reactions, and the thermal reaction between the expected decompostion products of aminimides was investigated in the presence and absence of PGE. The rate of PGE consumption was in the order PGE + 2 > PGE + 1 > PGE + 3 . It is clear that the formation of urethanes and oxazolidone derivatives affects the polymerization process.  相似文献   

20.
1,1-Dimethyl-1-(2-hydroxypropyl)amine p-substituted benzimide (“aminimide”) derivatives were prepared by the reaction of p-substituted methyl benzoates with equimolar amounts of 1,1-dimethylhydrazine and propylene oxide. These ylide compounds are shown to be useful as thermally latent initiators for the polymerization of glycidyl phenyl ether (GPE). Bulk polymerization of GPE with 3 mol % of these aminimides was carried out at 40–150°C for 8 h, showing ≥ 100°C was required for an effective rate. No consumption of the monomer could be observed at temperatures lower than 80°C. p-Methoxy substituted 1 showed the largest thermal latency among four aminimides tested. The activities of the aminimides increased with an increase of electron-donating ability of the substituents on the benzene ring, according to the following order: 1 (p-MeO) > 2 (p-Me) > 3 (H) > 4 (p-NO2). © 1997 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 35: 689–694, 1997  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