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.
The cationic oligomerization of 2-ethyl-1,3-butadiene (2EBD) by a superacid (CF3SO3H) and a superacid derivative (CH3COClO4) accompanied monomer isomerization to 3-methyl-1,3-pentadiene (3MPD) before propagation to yield oligomers of the isomerized monomer as main products in benzene at 50°C. Detection of 3MPD in the reaction mixture and 1H-NMR structural analysis of the produced oligomers confirmed the occurrence of this “monomer-isomerization oligomerization.” On the other hand, in the presence of a metal halide catalyst (BF3OEt2) 2EBD reacted without isomerization and yielded oligomers that were different from those produced by the foregoing superacid catalysts. Monomer isomerization was suppressed in a polar solvent [(CH2Cl)2] or at lower temperatures. The mechanism of the oligomerization with monomer isomerization was discussed.  相似文献   

2.
MAO/CpTiCl3 is an active catalyst for the polymerization of various types of 1,3-dienes. Butadiene, (E) - and (Z) −1,3-pentadiene, (E) −2-methyl-1,3-pentadiene and 2,3-dimethylbutadiene yield, at room temperature, polymers with a cis-1,4 or a mixed cis/1,2 structure. 4-Methyl-1,3-pentadiene and (E,E) −2,4-hexadiene give, respectively, a 1,2 syndiotactic and a trans-1,4/1,2 polymer. MAO/CpTiCl2·2THF and MAO/(CpTiCl2)n are less active than the CpTiCl3 catalyst, but give the same type of polymers. A change of stereospecificity with temperature was observed in the polymerization of (Z)-1,3-pentadiene: a cis-1,4 isotactic polymer was obtained at +20°C, and a crystalline 1,2 syndiotactic polymer at −20°C. This effect was attributed to a different mode of coordination of the monomer, which is cis-η4 at +20°C and may be trans-η2 at −20°C. Results obtained with catalysts from CpTi(OBu)3 and Ti(OBu)4 are reported for comparison. An interpretation is given of the formation of cis-1,4 isotactic poly(2-methylpentadiene) and of 1,2 syndiotactic poly(4-methylpentadiene), as well as of syndiotactic polystyrene.  相似文献   

3.
Conclusions trans-Piperylene is selectively dimerized to 4-methyl-2,5,7-nonatriene on a catalytic system composed of chelated Co complexes and triethylaluminum.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 7, pp. 1663–1664, July, 1973.  相似文献   

4.
Following previous results showing that direct initiation was operating in the cationic polymerization of 1,3-pentadiene in the presence of AlCl3 in non-polar medium, it is shown on the same system that direct initiation also occurs in polar medium. In the case of 2-methylpropene the use of a proton trap (DtBP) allowed to show that at −30 °C, direct initiation mechanism was operating either in 64/36 or in 36/64 (v/v) CH2Cl2/pentane mixtures. These results show that direct initiation is a general mechanism with AlCl3. SEC studies showed that for 2-methylpropene transfer can be minimized.  相似文献   

5.
The hydrogenation of 1,3-pentadiene into pentenes over the commercial 0.5% Pd/Al2O3 catalyst and over a new catalyst containing 1.0% Pd and 3.7% Ag (μ-catalyst) has been investigated. The new catalyst has been prepared via the flameless wave conversion of cyclotrimethylenetrinitramine in a porous composite. The catalytic properties of the new composite in the hydrogenation reaction depend on the hydrogen/1,3-pentadiene ratio and on the catalyst activation temperature. The reaction conditions for selective 1,3-pentadiene hydrogenation have been optimized. The pentenes yield as a function of temperature passes through a maximum at any H2/C5H8 ratio between 1 and 2. The 2-pentene/1-pentene ratio in the reaction products increases as the temperature is raised.  相似文献   

6.
The homogeneous catalyst CpTiCl3-MAO is able to produce a random copolymer of 4-methyl-1,3-pentadiene with ethylene. 13C NMR analysis of the copolymers shows that after insertion of ethylene units, the next 4-ymethyl-1,3-pentadiene unit can be inserted in either 1,2 or 1,4 arrangement. The high chemoselectivity observed in the 1,2-syndiotactic homopolymerization of 4-methyl-1,3-pentadiene with respect to the lower one observed in the copolymerization with ethylene is attributed to a back-biting coordination to the Ti of the active species of the penultimate monomer unit of the growing chain.  相似文献   

7.
The cationic polymerization of 1,3-pentadiene in the presence of vanadium oxytrichloride is studied. 1,3-Pentadiene is shown to polymerize at a high rate to high monomer conversions in the absence of proton-donor compounds in the catalytic system. The initial rate of 1,3-pentadiene polymerization is proportional to the concentration of VOCl3 in the system and demonstrates an extremal dependence on the initial concentration of 1,3-pentadiene. The polymerization process is distinguished by an induction period whose duration increases with a decrease in the reaction temperature. Regardless of polymerization conditions, with an increase in the monomer conversion, the molecular-mass distribution of the polymer widens owing to formation of a high-molecular-mass fraction, which, depending on reaction conditions, can be consumed in formation of the gel fraction. It is shown that the degree of unsaturation and the microstructure of poly(1,3-pentadiene) are almost independent of the polymerization conditions.  相似文献   

8.
Rh4(CO)12 anchored on γ-Al2O3 (Rh4(CO)12/Al2O3) has been studied as a catalyst for the hydrogenation of 1,3-trans-pentadiene. Under mild conditions (1 atm H2 and temperatures between 60°C and 80°C) hydrogenation occurs at only one of the double bonds of the diene, and analysis of the products shows that the terminal double bond is preferentially hydrogenated. Hydrogenation of the second double bond of the conjugated diene occurring only after all the 1,3-trans-pentadiene has been consumed. In this respect Rh4(CO)12/Al2O3 behaves like toluene solutions of Rh4(CO)12. Anchoring of Rh4(CO)12 on the solid support gives a catalyst which is less active but more stable than toluene solutions of Rh4(CO)12. The effects of CO and of triphenylphosphine on catalytic activity and on specificity of Rh4(CO)12/Al2O3 have also been investigated and both shown to cause a reduction of the rate of hydrogenation of 1,3-trans-pentadiene.  相似文献   

9.
Isomeric C5H8 compounds are distinguished by monitoring the products of their reactions with mass-selected ions generated from the individual isomers. This procedure, done by selecting appropriate reaction times in a quadrupole ion trap, yields data for the compounds which are more structure-selective than those obtained by collision-induced dissociation or dissociative charge stripping, both procedures in which isomer distinction is based on the behavior of the molecular ions rather than the neutral molecules themselves. All isomers except cis and trans 1,3-pentadiene can be distinguished by their ion/molecule reactions. The conjugated dienes, 1,3-pentadiene and isoprene, form the deprotonated C10H15+ dimer which is not generated by 1,4-pentadiene, cyclopentene, or by the allenes, 2,3-pentadiene and 3-methyl-1,2-butadiene. This clear, qualitative difference enables the isomers 1,4- and 1,3-pentadiene to be distinguished, which is otherwise difficult.  相似文献   

10.
A novel cycloadduct of 1-boryl-3,4-dimethylphosphole was prepared by reaction of 3,4-dimethylphospholyl anion with monobromoborane-methylsulfide complex (CH3)2S · BH2Br at −60 °C. It was characterized as a six-membered trimer by spectroscopic means, and its structure confirmed by an X-ray crystal analysis and quantum chemical calculations. Density functional theory calculations (B3LYP) showed that the cyclic trimer is by far more stable than the monomer, dimers or open-chain forms. Various molecular and spectroscopic properties of the borylphosphole monomer and trimer were evaluated. In particular, the changes of the 31P NMR chemical shifts upon oligomerization were examined. The six-membered ring was demonstrated to exist preferentially in a chair-like conformation. Computed NMR chemical shifts (1H, 13C and a lesser extent 31P) appear to be a highly sensitive analytical tool for distinguishing ring conformations having only small energy differences.  相似文献   

11.
Isoprene, 1,3-butadiene and 2,3-dimethyl-1,3-butadiene react with HFe(CO)4SiCl3 by addition of the Fe---H function to the diene. Isoprene appears to add predominantly 1,4 and 2,3-dimethyl-1,3-butadiene appears to add 1,2, while 1,3-butadiene may add both ways. In the case of isoprene and 1,3-butadiene loss of CO from the addition compound gives a stable π-allyl- Fe(Co)3SiCl3 product. Either cis- or trans-1,3-pentadiene is reduced to pentene by HFe(CO)4SiCl3.  相似文献   

12.
The mode of formation of isotactic and syndiotactic polymers from 1,3-dienes is examined in the light of the most recent results. An interpretation is given for the formation of trans-1,4 isotactic polymers from CH2=CH-CH=CHR (R = Me, Et, Pr, etc.) type monomers with heterogeneous VCl3-based catalysts. Evidence is reported showing that stereoregular 1,2 or cis-1,4 polymers derive from a growing polymer chain anti-η3-bonded to the transition metal and a cis-η4 coordinated monomer. The influence on stereoselectivity of the substituents at the central carbon atoms of the monomer is discussed. The peculiar behavior of (Z)-1,3-pentadiene and 4-methyl-1,3-pentadiene, which give 1,2 polymers with catalysts that give 1,4 polymers from other monomers, is attributable to the fact that they can coordinate trans-η2, in addition to cis-η4.  相似文献   

13.
Conclusions The codimerization of either 1,3-butadiene or 1,3-pentadiene with ethylene in the presence of the catalyst system C5H5Ti[OSi(CH3)3]3-CH3MgI assures the selective synthesis of vinylcyclobutanes.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 12, pp. 2781–2783, December, 1978.  相似文献   

14.
Photodissociation permits the distinction of four isomeric [C5H8] ions (ionized 2-pentyne, 1,2-pentadiene, 1,3-pentadiene and cyclopentene) which cannot be identified via collisional activation spectrometry. Both the relative cross-section for photodissociation and the relative abundance of the photodissociated fragments can be used to characterize the ion structure. Furthermore, upper and lower limits for the barrier for interconversion between 1,3-pentadiene and the other isomers can be determined.  相似文献   

15.
1,3-Butadiene, 4-methyl-1,3-pentadiene and styrene were polymerized with dicyclopentadienyltitanium dichloride/methylaluminoxane (Cp2TiCl2/MAO) and dicyclopentadienyltitanium chloride/MAO (Cp2TiCl/MAO). These systems are less active than cyclopentadienyltitanium trichloride/MAO (CpTiCl3/MAO), but show the same stereospecificity as the latter; they give predominantly cis-1,4-polybutadiene, 1,2-syndiotactic poly(4-methyl-1,3-pentadiene) and syndiotactic polystyrene. Cp2TiCl/MAO is much more active than Cp2TiCl2/MAO; this is probably due to the fact that in the reaction of Cp2TiCl2 with MAO, only a small amount of Ti(IV) is reduced to Ti(III), which is the active species in the polymerization of styrene and 1,3-dienes. An interpretation of the structure of the active species in Cp2TiCl/MAO is reported.  相似文献   

16.
Polymerization of 1,3-pentadiene was performed using the Gustavson complex (AlCl2.xylene.0.5HCl) at a nearly full conversion of the monomer. The possibility of control over molecular characteristics of the polymer was demonstrated. Recommendations for application of the poly-1,3-pentadiene synthesized are given.  相似文献   

17.
Jeremy M. Carr 《Tetrahedron》2008,64(13):2897-2905
Reductive desymmetrization of 2-methyl-2-substituted-cycloalkane-1,3-diones can be effected using either NaBH4 in DME or lithium tri-tert-butoxyaluminum hydride (LTBA) in THF at −60 °C. The former is a new approach that offers slightly greater diastereoselectivity in the reduction of 2,2-disubstituted-cyclopentane-1,3-diones while LTBA is superior with 2,2-disubstituted-cyclohexane-1,3-diones. Both conditions minimize subsequent reduction to diols thereby furnishing high yields of 1,3-ketols. Particularly rapid monoreductions are observed with 2-methyl-2-nitroethylcyclopentane-1,3-dione and 2-cyanoethyl-2-methylcyclopentane-1,3-dione when treated with NaBH4 in DME at −60 °C. As expected, diastereoselectivity varies considerably with the substitution at C-2.  相似文献   

18.
When the bulk oligomerization of 1,3‐dioxolan‐2‐one (ethylene carbonate, EC) and 4‐methyl‐1,3‐dioxolan‐2‐one (propylene carbonate, PC) with the 2,2‐bis(4‐hydroxyphenyl)propane (bisphenol‐A, BPA)/base system (bases such as KHCO3, K2CO3, KOH, Li2CO3, and t‐BuOK) was investigated at elevated temperature, significant differences were observed. Oligomerization of EC initiated by BPA/base readily takes place, but the oligomerization of PC is inhibited. The very first propylene carbonate/propylene oxide unit readily forms a phenolic ether bond with the functional groups of BPA phenolate, but the addition of the second monomer unit is rather slow. The oligomerization of EC yields symmetrical oligo(ethylene oxide) side chains. According to IR studies the oligomeric chains formed from PC with BPA contain not only ether but also carbonate bonds. The in situ step oligomerization of the BPA dipropoxylate was also identified by SEC, and a possible reaction mechanism is proposed. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 545–550, 1999  相似文献   

19.
1.  Addition of mercaptans of 3,5,5-trichloro-1,3-pentadiene proceeds with the formation of products of 1,2 and 1,4 addition.
2.  Arguments are presented in favor of rearrangement of the CHCl2HCCl=CHCH2SC4H9 radical to the ClHCHC1CCI=CHCH2SC4H9 radical during addition of butyl mercaptan to 3,5,5, -trichloro-1,3-pentadiene.
  相似文献   

20.
Summary Hydrogenations of individual C5 dienic substrates (1,3-cyclopentadiene, isoprene, 1,4-pentadiene, cis-1,3-pentadiene, trans-1,3-pentadiene) and their binary mixtures on a heterogeneous palladium catalyst were studied in cyclohexane at 25°C and 2 MPa. Selectivities of the competitive hydrogenations were determined and the substrates relative adsorption coefficients calculated. Effects of the diene structure on their reactivity and the stability of the surface complex are discussed. It was found that differences in selectivity of the competitive hydrogenations of C5 dienes are caused by the difference both in adsorptivity values and in reactivity of adsorbed molecules. The presence of a five-membered ring in the C5 dienes leads, , in significant reduction of surface complex stability as compared with acyclic structures of C5 dienes. On the other hand, it has a very positive effect on the rate of surface reaction.  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