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 0 毫秒
1.
Compositional heterogeneity of copolymers of butadiene with isoprene in copolymerization with supported titanium catalyst was investigated. The causes of compositional heterogeneity were found based on solution of an inverse problem of formation of the molecular weight distribution. It was shown that the pre-hydrodynamic effects in a turbulent reactor at the stage of the reaction mixture formation allowed the synthesis of more uniform composition of the copolymer.  相似文献   

2.
Titanates are versatile in the 1,4 polymerization of isoprene. The (R′O)4Ti/RAlCl2 catalyst gives either cis- or trans-1,4-polyisoprene, depending on the nature of both the titanate and the solvent. Primary titanates give cis-1,4-polyisoprene in both aliphatic and aromatic solvents. Secondary titanates give cis-polyisoprene in aliphatic solvents, and trans-1,4-polyisoprene in aromatic solvents. Tertiary titanates give trans-polyisoprene in both aliphatic and aromatic solvents. A mechanism is postulated which takes into consideration the role of the solvent. ESR studies of the various titanate–RAlCl2 catalysts were made; the paramagnetic structures are related to polymerization mechanisms.  相似文献   

3.
Commercial microheterogeneous titanium catalyst in the step of its preliminary formation was subjected to hydrodynamic action by circulation in a tubular diffuser-confuser apparatus. As a result of such treatment, the Mooney viscosity of the high-molecular-weight polyisoprene prepared in the presence of this catalyst becomes more stable.  相似文献   

4.
1,4-cis-polymerizations of 1,3-dienes of TiCl4-AIR3 catalyst systems for isoprene and on TiI2Cl2-AIR3 caltalyst systems for butadiene have been investigated. AIR3 denotes herein an organoaluminum compound (OAC) containing hydrocarbon radicals, linear of different lengths or cyclic, both saturated and unsaturated, as well as those containing O, N, or Si atoms. Neither the OAC radical structure nor the Al/Ti ratios in the range investigated have been found to influence the stereospecificities and reactivities (chain propagation constant) of the active centers. Differing activities of the catalyst systems have been established to result from the variations in the concentrations of the active centers involving <1–5% of all the Ti species. The nature of the OAC radical does not change the mode of the molecular weight distribution in the polymers obtained. An increase in the molecular weight (MW) of the polydienes produced with the use of higher-MW trialkylaluminums was provoked by an increase in the macromolecules propagation period owing to the lower constant of chain transfer to the higher-MW OAC.  相似文献   

5.
The possibility of using a turbulent prereactor of diffuser-confuser design for directional synthesis of butadiene and isoprene polymers and copolymers on a multisite titanium catalyst was examined. Preliminary forming of the reaction mixture under turbulent conditions decreases the contribution of low- and high-molecularmass modes with smoothing of nonmonotonic dependences of the intermediate-molecular-mass modes on the monomer ratio, which is due to disintegration of catalyst particles.  相似文献   

6.
A polymer-supported Ziegler–Natta catalyst, polystyrene-TiCl4AlEt2Cl (PS–TiCl4AlEt2Cl), was synthesized by reaction of polystyrene–TiCl4 complex (PS–TiCl4) with AlEt2Cl. This catalyst showed the same, or lightly greater catalytic activity to the unsupported Ziegler–Natta catalyst for polymerization of isoprene. It also has much greater storability, and can be reused and regenerated. Its overall catalytic yield for isoprene polymerization is ca. 20 kg polyisoprene/gTi. The polymerization rate depends on catalyst titanium concentration, mole ratio of Al/Ti, monomer concentration, and temperature. The kinetic equation of this polymerization is: Rp = k[M]0.30[Ti]0.41[Al]1.28, and the apparent activation energy ΔEact = 14.5 kJ/Mol, and the frequency factor Ap = 33 L/(mol s). The mechanism of the isoprene polymerization catalyzed by the polymer-supported catalyst is also described. © 1993 John Wiley & Sons, Inc.  相似文献   

7.
This work analyzes stereochemical aspects of olefin polymerization processes mediated by the C1-symmetric constrained geometry catalyst H2Si(ind)(tBuN)TiCH3+ (ind = indenyl), including the role of the cocatalyst/counteranion. The energetics of catalyst activation are first analyzed and shown to compare favorably with experiment. The energetics of heterolytic ion pair separation are next scrutinized, and the effects of solvation environment are assessed. Computed thermodynamic profiles for ethylene insertion at H2Si(ind)(tBuN)TiCH3+ indicate that the kinetics of insertion processes at the H2Si(ind)(tBuN)TiR+ cation can be analyzed in terms of SCF potential energies. We next compare the energetic profile for ethylene insertion at the naked H2Si(ind)(tBuN)TiCH3+ cation with that at the related H2Si(ind)(tBuN)TiCH3+H3CB(C6F5)3- ion pair to understand counterion effects. It is seen that the counterion, although affecting overall catalytic activity, does not significantly influence enchainment stereochemistry or polymer microtacticity. Next, the second ethylene insertion at H2Si(ind)(tBuN)Ti(nC3H7)+H3CB(C6F5)3- is analyzed to evaluate counteranion influence on the propagation barrier. It is found that the ethylene uptake transition state is energetically comparable to the first insertion transition state and that solvation has negligible effects on the energetic profile. These findings justify analysis of the propylene insertion process within the less computationally demanding "naked cation" model. Thus, monomer enchainment at H2Si(ind)(tBuN)TiR+ is analyzed for H2Si(ind)(tBuN)TiCH3+ + propylene (first insertion) and for H2Si(ind)(tBuN)Ti(iC4H6)+ + propylene (second insertion). Data describing the first insertion highlight the sterically dominated regioselection properties of the system with activation energies indicating that olefin insertion regiochemistry is predominantly 1,2 (primary), while the second insertion similarly reflects the catalyst stereoinduction properties, with steric effects introduced by the growing chain (mimicked by an isobutyl group) preferentially favoring insertion pathways that afford isotactic enrichment, in agreement with experiment.  相似文献   

8.
Copolymerization of butadiene and isoprene on the titanium catalyst proceeds on three types of active sites, and this correspond to the standard set of homopolymerization of isoprene. Sites responsible for the preparation of low-molecular-mass fractions of macromolecules are more active in the copolymerization as compared with homopolymerization of butadiene and isoprene. Change in the hydrodynamic regime in the reaction region is accompanied by an increase in the copolymerization rate due to an increase in the over-all concentration of active sites but it does not affect the typical set of growth sites of macromolecules. Distribution curve of active sites with respect to their reactivity is shifted to the high-molecular-mass region with increasing average weight molecular mass with decreasing average number molecular mass when the corresponding molecular-mass distribution becomes broader.  相似文献   

9.
10.
The kinetics of butadiene polymerization in hexane initiated by a catalyst prepared through the interaction of cobalt 2-ethylhexanoate with ethylaluminum sesquichloride and isoprene in the absence of water at Co: Al: isoprene = 1: 20: 20 (mol/mol) has been studied. The reaction orders with respect to the monomer and cobalt ethylhexanoate are estimated as 1.45 ± 0.04 and 1.5 ± 0.3, respectively. The concentration of active centers of the catalyst turns out to be 60 mol % of the cobalt concentration, as determined by the fractional inhibition of polymerization with cyclopentadiene. The rate constant of chain propagation, as calculated from the kinetic data, appears to be 1700 l/(mol min). The effective activation energy is 25.1 ± 7.5 kJ/mol.  相似文献   

11.
1.  The kinetics of the selective hydrogenation of isoprene to isoamylenes in a circulation-flow system in the vapor phase on a palladium-lead catalyst was studied; the results indicate a sequential mechanism of selectivity. An equation was derived describing the kinetics of the process.
2.  A kinetic model of the process was proposed, satisfying the aggregate of data obtained.
  相似文献   

12.
Lithium diethylamide has been found to be an active initiator for the polymerization of isoprene both in hydrocarbon media and in a variety of polar solvents, such as diethyl ether and tetrahydrofuran. The successful initiation of styrene polymerization is, however, strongly dependent upon the type of solvent employed. Thus no polymerization is observed in hydrocarbon media or in diethyl ether solution, but polymerization occurs rapidly in either tetrahydrofuran or 1,2-dimethoxyethane solution. These polymerization processes are anionic in nature and are characterized by sigmoidal conversion–time plots, indicating that the initiation reactions are relatively slow compared to chain propagation.  相似文献   

13.
The propagation rate constant Kp was used as a measure of reactivity of propagation centers in ethylene polymerization with oxide catalysts. This constant was determined by a radiotracer quenching technique for oxide catalysts of different compositions and activation conditions. For catalysts based on various transition metal oxides, an increase of Kp was observed in the series W < Mo < Cr and V < Cr. In the case of chromium oxide catalyst it was shown that Kp value does not depend on the content of the transition metal in a catalyst. A change of propagation center reactivity was found when oxides of different composition (SiO2, Al2O3, ZrO2, TiO2) were used as supports. An increase of the vacuum activation temperature of a catalyst results in increasing Kp. Pretreatment of catalyst with different reducing agents (SO2, CO2, NH3, HCN) results in the change of Kp value in comparison in comparison with the catalyst activated by the vacuum treatment only. The data obtained on the variation of the reactivity of the propagation centers permit one to draw a conclusion about the composition of surface compounds as acitve centers of the oxide polymerization catalysts.  相似文献   

14.
Cationic polyisoprene characteristic of different molecular parameters can be obtained with high yields in the presence of the TiCl4-trichloroacetic acid catalyst system. At low monomer concentrations, polyisoprene has a unimodal molecular-mass distribution. At elevated monomer concentrations, the polydispersity of polyisoprene increases significantly with conversion because of chain transfer to the polymer and branching. As the TiCl4/trichloroacetic acid ratio in the catalyst and the total concentration of the catalytic complex increase, the average molecular masses of the produced cationic polyisoprene decrease as a result of chain transfer to trichloroacetic acid.  相似文献   

15.
Butene-1 was polymerized with a highly-active supported titanium catalyst which was developed in this laboratory. The influences of various conditions (e.g., catalyst composition, temperature, external ester, H2, triethylaluminum, and catalyst concentration) on the catalytic activity, decay of polymerization rate, molecular weight, and isotacticity of the products were studied in detail. The structural properties of the PB-1 were characterized by WAXD, DSC, and 13C-NMR. It was found that the catalyst TiCl4, Ti(OBu)4/MgCl2/ethyl benzoate (EB)/Ph2SiCl2–AlEt3 shows high activity, i.e., 3.2 × 104 g PB/g Ti h. Isotacticity of the product was increased by adding p-CH3C6H4COOEt into the catalytic system. Molecular weight of the product can be easily controlled by H2. The decay of polymerization rate with time fulfills the equation: Rt ? Rs = (Ro ? Rs)et. © 1993 John Wiley & Sons, Inc.  相似文献   

16.
The kinetic heterogeneity of the active sites of titanium-containing catalytic systems in the stereospecific polymerization of isoprene was studied based on solving inverse problems for the molecular-weight distribution of polyisoprene with the use of the Tikhonov regularization method. It was found that from two to four types of active sites can occur depending on the nature of the organoaluminum compound used in the catalytic system. The rate constants of elementary steps of the polymerization process for particular types of active sites were obtained for the first time by solving inverse kinetic problems.  相似文献   

17.
Pd black was exposed to ethylene alone or in its mixture with hydrogen at 300 and 573 K. The samples were investigated by X-ray photoelectron spectroscopy (XPS) and ultraviolet photoelectron spectroscopy (UPS). Room temperature introduction of C(2)H(4) (also in the presence of H(2)) induced a binding-energy (BE) shift in the Pd 3d doublet and changed its full width at half-maximum (fwhm). The UPS features indicate shifting of electrons from the Pd d-band to Pd-H, Pd-C, and even Pd-OH species. Vinylidene (BE approximately 284.1 eV) may be the most abundant individual surface species on disperse Pd black, along with carbon in various stages of polymerization: "disordered C" (BE approximately 284 eV), graphite (approximately 284.6 eV), and ethylene polymer (approximately 286 eV), and also some "atomic" C (BE approximately 283.5 eV). Introduction of H(2) followed by ethylene brought about stronger changes in the state of Pd than exposure in the reverse sequence. This may indicate that the presence of some surface C may hinder the decomposition of bulk PdH. Formation of Pd hydride was blocked when ethylene was introduced prior to H(2). The C 1s intensity increased, the low-binding-energy C components disappeared, and graphitic carbon (BE approximately equal to 284.6 eV) prevailed after ethylene treatment at 573 K. The loss of the Pd surface state and "PdH" signal were observed in the corresponding valence band and UPS spectra. Hydrogen treatment at 540 K was not able to decrease the concentration of surface carbon and re-establish the near-surface H-rich state. UPS showed overlayer-type C in these samples. The interaction of Pd with components from the feed gas modified its electronic structure that is consistent with lattice strain induced by dissolution of carbon and hydrogen into Pd, as indicated by the d-band shift and the dilution of the electron density at E(F).  相似文献   

18.
The short-time polymerization of butadiene induced by heterogeneous titanium catalysts was studied for the first time. In the 0.1?0.3 s time interval, polymerization is characterized by a considerable decrease in the chain propagation rate constant and 1,4-stereospecificity of the catalysts. A structure kinetic continuum model was proposed in which the initiation step and the first propagation steps are kinetically continuous. The violation of this continuity changes the catalyst stereospecificity and creates conditions for chain transfer reactions.  相似文献   

19.
对正辛酸一三氟化硼乙酸络合物(n-C_8H_(17)0H-B)预混、乙酸正丁酯一三氟化硼乙醚络合物(CH_3COOC_4H_9-B)预混和稀三氟化硼乙醚络合物(B)单加三种体系的丁二烯聚合行为进行了比较。结果表明;B组分中加入n-C_8H_(17)OH,能在不拓宽三异丁基铝(Al)/B摩尔比范围的情况下,有效地提高聚丁二烯的分子量;B组分中加入CH_3COOC_4H_9,能适当地拓宽Al/B比,但对聚丁二烯分子是不产生影响。对n-C_8H_(17)0H-B在聚合体系中的作用机理进行了探讨,提出了在n-C_8H_(17)0H-B预混体系中存在两种活性中心。  相似文献   

20.
The possibility of reducing the consumption of titanium catalyst in butadiene polymerization by using a tubular turbulent prereactor of the diffuser-confuser design in the step of mixing the reaction mixture components was examined.  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