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 687 毫秒
1.
The metal-polyyne polymers consisting of transition metals and conjugated tetrayne systems where M represents ? Pt(PBu3)2? or ? Pd(PBu3)2? moiety were prepared by the oxidative coupling method and characterized by spectral analysis, associated with novel depolymerization giving binuclear transition metal complexes, and   相似文献   

2.
The 13C-NMR spectra of hydrogenated polyisoprenes were investigated. Polyisoprenes containing various ratios of 1,4 and 3,4 units were prepared with n-BuLi/Et2O and were hydrogenated by using p-toluenesulfonylhydrazide. 13C-NMR signals of the hydrogenated polymers were assigned for the triad sequences arising respectively from the 1, 4 and 3, 4 units of polyisoprenes. On the basis of these assignments it was confirmed that 1,4 and 3,4 units were randomly distributed in the polyisoprenes prepared with n-BuLi and that these polymers did not contain appreciable amounts of head-to-head or tail-to-tail 1,4 linkages.  相似文献   

3.
Details of the thermal behavior, including transition temperatures, transition heats and transition entropies, are reported for three series of nematic liquid-crystalline “backbone” polyesters having the general structure All polysters described above were examined by differential scanning calorimetry and were found to have reproducible thermal behavior once the polymer had been annealed by heating to the isotropic phase and then subsequently cooled. Enantiotropic nematic phases were found for all 30 polysters studied. Many polymers showed multiple endotherms on melting to the nematic phase. Plots of solid–nematic and nematic–isotropic transition temperatures versus number of carbon atoms in the diacid segment (x) for each (y) reveal an even–odd alternation reminiscent of trends in homologous series of small-molecule liquid crystals. Enthalpies for the solid → nematic and nematic → isotropic transitions do not show such a precise alternation.  相似文献   

4.
Certain polymer precursors of the general formula: where A is an aromatic structure bearing ester, amide, azomethine, or imide linkages were synthesized. Particularly, 4-hydroxybenzaldehyde was condensed with malonitrile to afford 4-hydroxy-β,β′-dicyanostyrene which reacted with a half molar amount of terephthaloyl dichloride in the presence of an acid acceptor. In addition, 3-nitrobenzaldehyde was condensed with malonitrile to yield 3-nitro-β,β′-dicyanostyrene that was catalytically hydrogenated to the corresponding amine. The latter reacted with a half molar amount of terephthaloyl dichloride, terephthaldehyde, pyromellitic dianhydride, or benzophenone tetracarboxylic dianhydride. The polymer precursors were characterized by elemental analyses as well as IR and 1H-NMR spectroscopy. Their curing behavior was investigated by DTA. Crosslinked polymers were obtained by curing the monomers at 300°C for 24 h. They were stable up to 407–437°C in N2 and afforded an anaerobic char yield of 65–50% at 800°C. The thermal stability of resins was correlated with their chemical structures. © 1993 John Wiley & Sons, Inc.  相似文献   

5.
Ab initio molecular orbital calculations with large, polarization basis sets and incorporating valence electron correlation have been employed to examine the [C2H2O] potential energy surface. Four [C2H2O] isomers have been identified as potentially stable, observable ions. These are the experimentally well-known ketene radical cation, [CH2?C?O] (a), and the presently unknown ethynol radical cation, [CH2?C? OH] (b), the oxirene radical cation (c) and an ion resembling a complex of CO with [CH2], (d). The calculated energies of b, c and d relative to a are 189, 257 and 259 kJ mol?1, respectively. Dissociation of ions a and d is found to occur without reverse activation energy.  相似文献   

6.
7.
8.
The reactions of fluorosulfonylurea derivatives with phosphorus pentachloride in carbon tetrachloride leads to and . These compounds are easily fluroinated by arsenic trifluoride to the corresponding fluorosulfonyl derivatives. It was shown that PCl5 is reliable for the fluorine-chlorine exchange. Isocyanate esterchlorides as well as carbonyl-fluoride derivatives react with compounds which have an active hydrogen atom. By this way were prepared: . By hydrolysis of , is probably formed as the first reaction product. The acid could be trapped with tetraphenylphosphoniumchloride as a well-cristalline salt. Fluorosulfonyl isocyanate reacts with aldehydes to the following compounds: FSO2N ? CHCH3, FSO2N ? CHC2H5, FSO2N ? CH? CH(CH3)2 and FSO2N ? CHC6H5. The physical and chemical properties as well as IR and NMR spectra of the compounds are described and, in part, compared and discussed.  相似文献   

9.
The catalytic activity of imidazole-containing polymers in the hydrolyses of substrates with poor leaving groups was examined. Hydrolyses of p-methoxyphenyl esters (Sn) catalyzed by poly[4(5)-vinylimidazole] (pvIm) in relation to imidazole (Im) indicates that both cooperative and hydrophobic interactions are operative. Hydrolyses of 3-methoxy-4-acyloxybenzoic acid substrates (S) catalyzed by pvIm and a water-soluble copolymer, copoly[1-methyl-4-vinylimidazole/4(5)-vinylimidazole], exhibit many characteristics of enzyme-catalyzed reactions like saturation kinetics, bellshaped pH-rate profiles, and nonproductive binding. The importance of general-base, cooperative interactions for substrates with poor leaving groups and hydrophobic interactions in the formation of a stable catalyst-substrate complex in the case of long-chain esters is demonstrated. The interesting similarity between the Michaelis constants Km for the two polymer catalysts, pvIm and the copolymer, suggests that the common underlying principle involved in binding is hydrophobic interactions and the presence of N? CH3 group in the copolymer does not increase the hydrophobicity significantly. The pH-rate profiles for the hydrolyses of S by pvIm show that the optimum pH is around neutrality, which indicated that the presence of neutral Im units to serve as nucleophiles and protonated Im units to serve as electrostatic binding sites is essential for maximum catalytic efficiency. The rate enhancement as a function of acyl chain length at different pH leads to the conclusion that a cooperative effect between electrostatic and hydrophobic interactions would reduce the nonspecificity of hydrophobic interactions and result in their better realization. An approximate calculation shows that the binding of S, in relation to S, with the copolymer in aqueous solution corresponds to the additional interactions of two methylene units. Our recent studies1–10 in the field of catalysis by polymeric imidazoles are oriented mainly toward the recognition of the significance of hydrophobic interactions. The role of hydrophobic interactions in enzymatic catalysis11–13 and synthetic macromolecular catalysts1–10,14–29 has been understood only recently. Hydrophobic interactions describe the tendency of nonpolar groups to associate themselves in aqueous solution30,31. Because the catalytic mechanism generally involves the prior complexation of the catalyst with the substrate and the catalyst-substrate complex is apolar, the influence of hydrophobic interactions in catalysis is conceivable. The favorable free energy of formation of hydrophobic interaction leads to better binding and better catalysis. In the present study we examined the hydrolytic reactions of the following substrates with poor leaving groups, catalyzed by imidazole-containing vinyl polymers, poly[4(5)-vinylimidazole] (pvIm), and a water-soluble copolymer, copoly[1-methyl-4-vinylimidazole/4(5)-vinylimidazole], ~(1:1)M, Esters of varying acyl chainlength are chosen to determine the influence of hydrophobic interactions.  相似文献   

10.
Sulphamoyl chlorides and chlorosulphonyl isocyanate react with monosubstituted hydrazones and alkylhydrazonates to sulphamoyl hydrazones and sulphamoyl hydrazonates respectively. Reaction of benzil monoalkylhydrazones with chlorosulphonyl isocyanate results in formation of 2‐alkyl‐4,5‐aryl‐2H‐ [1λ6,2,3,6]‐thiatriazine‐1,1‐dioxides.  相似文献   

11.
Protonated acetophenones, substituted with a methoxymethyl group in the para and meta positions, have been generated by electron impact induced fragmentation of the correspondingly substituted 2-phenylpropan-2-ols. The metastable ions, formed in the second field-free region of a VG ZAb 2F mass spectrometer, react unimolecularly by elimination of CH3OH, formation of CH3CO+ and ions, loss of CH3COOCH3, and loss of CH2O. The mechanisms of these fragmentations have been elucidated with the aid of deuterated analogues of the protonated acetophenones. It is shown that these reactions are initiated by an endothermic transfer of the proton at the carbonyl group of the protonated acetophenones to the benzene ring. A further migration of the proton to the ether O atom of the methoxymethyl side-chain leads eventually to the elimination of CH3OH. Protolytic bond cleavages of either side-chain gives rise to the CH3CO+ and ions. At low internal energies both these ions may be trapped by the aromatic neutral fragment in ion-molecule complexes. Reactions within these complexes result in the energetically favourable losses of CH3COOCH3 and CH2O, respectively. With respect to these reactions, the protonated acetophenones behave analogously to the correspondingly substituted and protonated benzaldehydes.  相似文献   

12.
NMR spectra of twelve newly synthesised mercaptoacid amides are being reported and discussed. The syntheses and structure of more than thirty mercaptoacid amides have been reported(1-3) from our laboratories. The study has revealed that N -phenyl-2-mercaptoacetamide and related compounds have intramolecular hydrogen bonded structure (I).   相似文献   

13.
Quinones (Q) do not affect the liquid-phase oxidation of hydrocarbons and retard the solid-phase oxidation of polymers terminating chains in reactions with alkyl macroradicals. This difference is the result of specific influence of a polymer matrix on the kinetics of free radical reactions. Quinones were found to terminate chains in oxidizing polypropylene containing hydroperoxide groups by the reaction with hydroperoxyl radicals. This is the result of peroxide group's decomposition with hydrogen peroxide production and chain termination in the following reactions: The equilibrium reaction between hydroquinone (QH2) and quinonimine (QI) was evidenced to proceed as a chain reaction with chain propagation in reactions of •QH with QI and HQI• with QH2. Analysis of the reactions of quinone with phenols (ArOH) proved that it can be important as an additional way of chain termination in oxidizing polymers and hydrocarbons at elevated temperatures.  相似文献   

14.
1,5‐Hydrogen transfer reactions in methyl acrylate and butyl acrylate free‐radical polymerization are studied using quantum chemistry and transition state theory to estimate the kinetic parameters (ktr, Ea, and A) with tetrameric radicals, requiring a number of atoms that ranks among the largest polymeric mimics to date. A two‐step transformation accounted for the overall reaction: rotation from an extended conformation to a coiled conformation and abstraction of the fifth hydrogen atom by the end‐chain radical. UB3LYP/6‐31G(d) was used for geometry optimization, validation of the transition states, and calculation of frequencies that were used to obtain thermodynamic properties. The more computationally demanding level of theory, MPWB1K/6‐31G(d,p), was used for calculation of the electronic energy.

  相似文献   


15.
This article concerns the synthesis and characterization of new epoxy and aldehyde telechelic polyisobutylenes, that is, and . The synthesis of the epoxy derivative was achieved by quantitative epoxidation of α,ω-di(isobutenyl) polyisobutylene with m-chloroperoxybenzoic acid and that of the dialdehyde by quantitative isomerization of the epoxy termini with zinc bromide. Infrared (IR) and 1H-NMR analysis of these new telechelic polymers and ultraviolet (UV) analysis of the 2,4-dinitrophenylhydrazine derivative of the dialdehyde indicate quantitative conversions and yields, that is, essentially theoretical functionalization (F n = 1.95 ± 0.05).  相似文献   

16.
The structure of polymers obtained by polymerization of methoxyallene and ethoxyallene with transition-metal catalysts depends on the catalyst employed. Rapid polymerization at 0°C through the unsubstituted double bond occurred with π-allylnickel halides, NiCl2/AlEt3 and CoCl2/AlEt3, yielding polymers with the structure Typical Ziegler–Natta catalysts (TiCl4, VOCl3 or FeCl3 with AIEt3) gave polymers mainly with the structure although some of the structural units were probably present as well. Polymers having conjugated double bonds were prepared with PdCl2, [(π-allyl)PdCl]2, and PdCl2(C6H5CN)2 as catalysts. Palladium iodide produced polymers with all three of the above structural units present. Polymerization occurred more slowly with these palladium catalysts. A preliminary examination of the effect of variation of solvent, ligand, co-catalyst, and temperature on the rate and structure of the polymers obtained with the palladium catalysts is reported.  相似文献   

17.
3‐Bromopyridine‐4‐carbaldehyde is cyclized with carboxylic acids in acetonitrile at 100° under carbon monoxide pressure in the presence of a catalytic amount of a palladium catalyst along with a base to afford the corresponding 3‐oxo‐1,3‐dihydrofuro[3,4‐c]pyridine‐1‐yl alkanoates in moderate to good yields.  相似文献   

18.
A study of the isomorphous relationships existing among five crystalline polyaromatic polymers containing O, S, CH2 bridges, i.e., where Ph is phenyl, is reported. Appropriate series of oligomers corresponding to these polymers have also been prepared in order to investigate how the solid-state properties of the oligomers correlate with those of the polymers. A preliminary x-ray analysis is reported showing that the polymers are actually isostructural.  相似文献   

19.
Collisional activation demonstrates that the stable ions from o-nitrobenzaldehydedimethylacetale possess the structure of ionized o-nitroso benzoic acid methyl ester. Contrary to previous conclusions it is demonstrated that the structure of the stable ions (m/e 135) from different precursors [i.e. o-nitrobenzyl alcohol o-nitrobenzyl cyanide and o-nitrobenzaldoxime is best represented by 2,1-benzisoxazoline-3-one. Ionized o-nitrosobenzaldehtde rearranged to 2,1-benzisoxazoline-3-one prior to collision induced decomposition, whereas 2-benzoxazolinone and 3-hydroxy-1,2-benzisoxazole do not rearrange within 10?5 s.  相似文献   

20.
Diphenylphosphorous chloride and methyl iodide add readily to the N-bonded P(III)-atom of (C6H5)2P? P(C6H5)3?N? P(C6H5)2 forming the salts [(C6H5)2P? P(C6H5)2 N P(C6H5)2? P(C6H5)2]Cl and [(C6H5)2P? P(C6H5)2 N P(C6H5)2. CH3]I, respectively. A similar behaviour is observed with sulfur: Under mild conditions (C6H5)2P? P(C6H5)2?N? P(C6H5)2 = S is formed but forcing conditions are required to produce S = P(C6H5)2? P(C6H5)2?N? P(C6H5)2?S. The monosulfide is also obtained by treating (C6H5)2P(S)N[Si(CH3)3]2 with diphenylphosphorous chloride, indicating the favoured formation of the phosphazene system as compared with the phosphazane system Confirmation of the structures comes from 31P nmr and IR data, and for the sulfides also from their degradation with bromine.  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