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 997 毫秒
1.
A new class of catalysts for polymerization of cycloalkenes was developed, based on the interaction of these hydrocarbons with π-allylic transition metal complexes. The reaction route (decyclization or double-bond opening) and the microstructure of polyalkenamers obtained is determined primarily by the nature of the transition metal and also by the nature and the number of ligands bound to the metal. π-Allylic complexes of zirconium and chromium catalyze the polymerization of cycloalkenes in both directions. For complexes of the group VI metals the substitution of molybdenum or tungsten for chromium results in complete suppression of the double-bond opening. In the presence of Group VIII metal compounds the polymerization of cycloolefins occurs exclusively at the double bond. The introduction of acidic ligands into the internal sphere of the π-allylic compound does not affect the reaction route but results in a significant rise of the yield of polyalkenamers and enriches the content of trans-1,4-units. Lewis acids act as activators for complexes of all the metals, the halides of molybdenum and tungsten affecting favorably the decyclization. It is assumed that the ring-opening polymerization of cycloolefins proceeds via π-allylic complex formation.  相似文献   

2.
Various π-allylic nickel complexes were synthesized by reacting bis(1,5-cyclooctadiene)-nickel with an allylic ester of trifluoroacetic acid. A study of the stereospecific polymerization of 1,3-butadiene initiated by these catalysts showed the strong influence of the initial substitution in the π-allylic ligand as well as of the solvent used for synthesis of the catalysts on the course of the reaction, i.e., on both the polymerization rate and the stereospecificity. Moreover, new catalytic species were isolated by adding electron-accepting or electron-donating ligands to some of the π-allylic complexes. In that way, we were able to modify the behavior of the catalysts and, in some cases, to enhance catalytic activity considerably.  相似文献   

3.
In the course of investigations of polymerization of diolefins by transition metal derivatives, we have synthesized various monometallic nickel coordination catalysts. The complexes were prepared by reacting 2,6,10-dodecatriene-1,12-diyl nickel with protonic acids; they were shown to initiate the stereospecific polymerization of 1,3-butadiene. The study of these catalysts showed the strong influence of the nature of the counteranion used on the stereospecificity and the polymerization rate. Moreover, by adding various ligands, we were able to modify the behavior of the catalytic systems and to prepare either pure cis-1,4 or pure trans-1,4 or cistrans equibinary polybutadienes, starting from the same complex and keeping a high 1,4 specificity. Some of these modifications were shown to be reversible.  相似文献   

4.
Turnover frequencies of catalytic systems based on nickel complexes with 1,4-diaza-1,3-butadiene (α-diimine) ligands in the reactions of styrene hydrogenation and ethylene polymerization were determined. Results are presented of a study by the electron paramagnetic resonance method and IR spectroscopy of 1,4-diaza-1,3-butadiene complexes of nickel(II) and anion radicals formed in the interaction of the starting components under the conditions of catalysis. It was shown that the paramagnetic Ni(I) complexes are precursors of complexes catalytically active in the hydrogenation and polymerization reactions.  相似文献   

5.
This article provides an overview on recent progress in the polymerization of 1,3-dienes catalyzed by transition metal complexes with phosphorus and nitrogen ligands. Polymers having different microstructures (cis-1,4; 1,2; mixed cis-1,4/1,2) and tacticity (iso- or syndiotactic) were obtained from various 1,3-dienes (1,3-butadiene, isoprene, 1,3-pentadienes, 1,3-hexadienes) depending on the catalyst used, clearly suggesting that the catalyst structure (i.e. metal nature, type of ligand) strongly affects the polymerization chemo- and stereoselectivity. However, as indicated by the results obtained in the polymerization of substituted butadienes, a fundamental role in determining the selectivity is also played by the type of monomer: polymers with different structure, some of them completely new, were obtained from different monomers with the same catalyst. All these observations permitted to confirm, and in some cases to improve, the knowledge on the diene polymerization mechanism.  相似文献   

6.
The turnover frequency and number have been determined for eighteen catalytic systems based on triphenylphosphine and 1,4-diazo-1,3-butadiene complexes of nickel in the formal oxidation states 0, +1, and +2 in the oligoand polymerization of lower alkenes. The main catalytic characteristics are almost independent of the oxidation state of nickel in the precursor and depend on the nature and concentration of the cocatalyst (Lewis acid). The catalytic systems have been studied by ESR. The ESR spectral parameters are presented for nickel(I) 1,4-diazo-1,3-butadiene complexes and radical anions resulting from the reactions of the cocatalyst with nickel α-diimine complexes. Reactions describing the formation, functioning, decomposition, and regeneration of the catalytically active nickel hydride complexes are proposed.  相似文献   

7.
The ruthenium-catalysed double addition of trimethylsilyldiazomethane to alkynes developed by Dixneuf and co-workers was applied to the synthesis of 2-alkyl- or 2-aryl-3-(4,4,5,5-tetramethyl-1,3,2-dioxaborolan-2-yl)-1,4-bis(trimethylsilyl)-1,3-butadienes by use of alkynylboronates instead of alkynes. Di- and tetrasubstituted 1,3-butadienes were prepared from a 2-boryl-1,4-disilyl-1,3-butadiene, using the Suzuki-Miyaura coupling reaction, iodolysis of the alkenylsilane moieties with N-iodosuccinimide and hydrolysis of the carbon-silicon bonds with trifluoroacetic acid. The same compound was converted also to a bicyclic compound, a trisubstituted 1,3-butadiene and a dienone through the Diels-Alder reaction, oxidation of the alkenylboronate moiety and the Mukaiyama aldol reaction.  相似文献   

8.
The main path of the interaction of chromium and nickel η3-allyl complexes and titanium tetrachloride is the transfer of allyl ligands to titanium with the concomitant reduction of TiCl4 to η-TiCl3 and the formation of the products initiating isoprene polymerization. The stereospecific effect of the system is because of the formation of bimetallic complexes with bridged metal—carbon bonds which are the most probable centers of stereospecific cis-1,4-polyisoprene chain propagation.  相似文献   

9.
胡雁鸣  张学全 《高分子科学》2016,34(9):1060-1069
Cobalt and nickel complexes (1a-1d and 2a-2d, respectively) supported by 2-imidate-pyridine ligands were synthesized and used for 1,3-butadiene polymerization. The complexes were characterized by IR and element analysis, and complex 1a was further characterized by single-crystal X-ray diffraction. The solid state structure of complex 1a displayed a distorted tetrahedral geometry. Upon activation with ethylaluminum sesquichloride (EASC), all the complexes showed high activities toward 1,3-butadiene polymerization. The cobalt complexes produced polymers with high cis-1,4 contents and high molecular weights, while the nickel complexes displayed low cis-1,4 selectivity and the resulting polymers had low molecular weights. The catalytic activities of the complexes highly depended on the ligand structure. With the increment of polymerization temperature, the cis-1,4 content and the molecular weight of the resulting polymer decreased.  相似文献   

10.
On the base of structural and thermodynamic data using modern methods of quantum chemistry, a comparative theoretical study of the elementary acts of initiation and growth of a polymer chain during the polymerization of 1,3-butadiene was carried out. Ziegler–Natta catalysts based on Nd(III) and Gd(III) were used as polymerization initiators. It was shown that the higher stereospecificity of the action of gadolinium catalytic complexes in comparison with neodymium complexes is due to the increased stability of the anti-form of the π-allylic terminal structure of the growing polymer chain, and the reduced activity is due to the higher activation energy of the processes of initiation and chain growth.  相似文献   

11.
2-Arylhydrazono-1,3-diketones react with ethylenediamine to give, depending on the substituents, dihydro-1,4-diazepine derivative or N,N'-ethylenebis(1,3-aminovinyl ketones). Treatment of the latter with nickel(II) or copper(II) acetate results in formation of the corresponding metal chelates. Nickel complexes of N,N'-ethylenebis(1,3-aminovinyl ketones) can also be synthesized from 2-arylhydrazono-1,3-diketones and ethylenediamine on a metal template. Reactions of 2-arylhydrazono-3-oxo esters with ethylenediamine yield N,N'-ethylenebis(3-alkyl-2-arylhydrazono-3-oxopropionamides). Ethyl 2-arylhydrazonoacetoacetate reacts with ethylenediamine under mild conditions, affording ethyl 2-p-tolylazo-3-[2-(2-p-tolylhydrazono-1,3-dioxobutylamino)ethylamino]-2-butenoate.  相似文献   

12.
By 13C NMR spectroscopy, the structures of (C4D7Nil)2 adducts with isoprene, 2-ethyl-, 2-isopropyl- and 2-tert. butylbutadienes have been investigated. The nature of the 2-alkyl substituent in the anti- and syn-1,2-disubstituted π-allylic complexes influences the character of electron charge densities distribution on the carbon atoms of a π-allylic ligand. Mutual arrangement of substituents in the π-allylic active sites has been shown to remain in the double bonds of elementary units of a polymer chain.  相似文献   

13.
Synthesis of alkoxyolefins from alcohols and dienes can be achieved in the presence of hydrated rhodium trichloride. Because of the probable intermediate formation of π-allylic rhodium complexes, the regioselectivity of the addition varies according to the reaction conditions. Without any additives, the greatest selectivity in 1,4-addition is 53% for a 78% isoprene conversion. In the presence of zinc chloride, 1,4-addition is significantly increased (68% for a 70% conversion).  相似文献   

14.
π-Allyl complexes of nickel induce stereospecific homopolymerization of cyclohexadiene-1,3 and 2,3-dimethylbutadiene with the formation of crystalline polymers. These polymers consist of 1,4-monomer units in cis-configuration for polycyclohexadiene and in trans-configuration for poly-2,3-dimethylbutadiene. Peculiarities of copolymerization of cyclohexadiene-1,3 with butadiene and isoprene and of 2,3-dimethylbutadiene with butadiene under the influence of various π-allyl complexes of nickel are studied. By i.r.- and NMR-spectroscopy and radiochemical methods, the composition of copolymers for the above pairs of monomers are determined and the reactivity ratios are found. Influence of the monomers on the microstructure of the chain in copolymerization is established; the mechanism of this phenomenon is discussed.  相似文献   

15.
Stereodefined multisubstituted 1-cyano- and 1,4-dicyano-1,3-butadiene derivatives were obtained in excellent yields of the isolated product from their corresponding monohalo- and dihalobutadienes and CuCN. This reaction proceeded with high stereoselectivity and retention of the stereochemistry of the starting halobutadienes. A study of the utility of the thus-obtained 1-cyano- and 1,4-dicyano-1,3-butadiene derivatives was demonstrated by their reactions with organolithium reagents. 2H-Pyrrole or iminocyclopentadiene derivatives were formed in high yields from 1-cyano-4-halo-1,3-butadienes and organolithium reagents. When 1,4-dicyano-1,3-butadienes were treated with organolithium reagents followed by trapping with electrophiles, a tandem process took place to afford 2H-pyrrolyl nitriles in excellent yields. Reduction of 1,4-dicyano-1,3-butadiene derivatives with LiAlH4 showed novel reaction patterns relative to normal nitriles.  相似文献   

16.
《Tetrahedron: Asymmetry》1998,9(13):2271-2277
The hetero-Diels–Alder reactions of 1-alkyl-3-(tert-butyldimethylsilyl)oxy-1,3-butadienes with ethyl glyoxylate catalyzed by chiral salen–metal complexes have been studied. With a cobalt(II) complex as the catalyst, the reaction of 1-(2-benzyloxyethyl)-3-(tert-butyldimethylsilyl)oxy-1,3-butadiene with ethyl glyoxylate gave the Diels–Alder product in 75% isolated yield with the endo:exo ratio >99:1 and the enantiomeric excess ≥52%. The effects of temperature, catalyst and solvent are also discussed.  相似文献   

17.
The cationic polymerizations of dimethyl-1,3-butadienes with various catalysts in methylene chloride and toluene have been investigated. The activity of catalysts decreased in the order WCl6 > AcClO4 > SnCl4·TCA > BF3OEt2. The homopolymerization rate of dimethyl-1,3-butadienes with WCl6, AcClO4, and SnCl4·TCA decreased in the order 1,3-dimethyl-1,3-butadiene > 2,3-dimethyl-1,3-butadiene > 1,2-dimethyl-1,3-butadiene > 2,4-hexadiene. The polymers prepared with WCl6, SnCl4.TCA, and BF3OEt2 were rubberlike polymers or white powders, whereas those prepared with AcClO4 were oily oligomers. The 1,4-propagation increased in the order 1,2-dimethyl-1,3-butadiene < 1,3-dimethyl-1,3-butadiene < 2,3-dimethyl-1,3-butadiene < 2,4-hexadiene. This order may indicate that the steric effect of methyl group determine primarily the microstructure of the polymer. The relative reactivity of dimethyl-1,3-butadienes toward a styryl cation decreased in the order 1,3-dimethyl-1,3-butadiene > 1,2-dimethyl-1,3-butadiene > 2,3-dimethyl-1,3-butadiene > 2,4-hexadiene. This order may be explained in terms of the stability of the resulting allylic cation.  相似文献   

18.
The homocoupling reaction between the conjugated n-(2-chloroethenyl)pyridine; n, 2-, 3- and 4- (or quinoline; n, 2- and 4-) mediated by zero-valent nickel complexes at room temperature affords to the corresponding 1,4-diaryl-1,3-butadiene, always as the 1E,3E stereoisomer. The yield in 1,4-diaryl-1,3-butadiene increases with the nickel catalyst and hence, the active zero-valent nickel catalyst is not regenerated during the homocoupling reaction.The stereospecific synthesis of (1Z,3Z)-1,4-di(4′-pyridyl)-1,3-butadiene stereoisomer was efficiently carried out by partial hydrogenation of the appropriate 1,4-di(4′-pyridyl)-1,3-butadiyne.  相似文献   

19.
The reaction of 2-t-butyl-1,3-butadiene with π-allylic palladium chlorides is described. The 2-chloroallyl complex gives a π-allylic complex with an open chain structure, in which the butenyl group in is the anti-position. Allyl-1-carbomethoxyallyl-, methallyl-, and 1-carbomethoxymethallyl-palladium chlorides give six-membered cyclic complexes. The product of the former two are π-allyllic and those of the latter are σ-bonded complexes.  相似文献   

20.
Y. Ohbe  M. Takagi  T. Matsuda 《Tetrahedron》1974,30(16):2669-2675
The reactions of allyl bromide and crotyl chloride with Grignard reagent catalyzed by π-allyl and crotyl metal complexes of nickel, cobalt, and iron, and the stoichiometric reaction of the complexes with the Grignard reagent have been examined. The similarity in catalytic behaviour of the complex and the corresponding metallic halide affords further evidence in support of the previous proposal that the π-allylic metal intermediate plays an important role in the catalytic reaction. The stoichiometric reaction suggests that the dependence of distribution of product in the catalytic process on the type of both allylic halide and metal is attributable to the facility of ligand exchange between the π-allylic complex and Grignard reagent.  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