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 7 毫秒
1.
The stereoselectivity of the model active site formed by the adsorption of Ti2Cl7 on the (1 0 0) surface of MgCl2 was investigated by density functional calculations. The analysis of energetics for successive propylene insertions into the model active site reveals that the insertion of propylene into the model active site is energetically more favorable when a growing chain and one chlorine atom (that makes the active site chiral) are on the opposite side rather than on the same side. From this result, it is realized that the model active site is highly stereoselective. It is also observed that the Cl atoms near the growing chain significantly affect the activation energy barrier through the interaction with the growing chain.  相似文献   

2.
The kinetics of the ethylene/1-hexene copolymerization reaction with a Ti-based Ziegler-Natta catalyst has been studied. Kinetic analysis established the existence of several populations of active centers in the catalyst. The centers differ in two aspects: their ability to incorporate α-olefin units into copolymer chains (i.e., their reactivity ratios) and the average molecular weights of the polymer chains they produce. The centers of different populations are formed at different rates and have different kinetic stabilities. As a consequence, both the molecular weight distributions of the copolymers and their compositional distributions are relatively broad and change with in time. Two kinds of catalyst poisons were found. The poisons of the first type, arylalkoxysilanes, preferentially deactivate the centers which have the highest ability to copolymerize α-olefins with ethylene. These poisons decrease the average α-olefin content in the copolymers and the fraction of their olefin-rich components. The poisons of the second type, conjugated dienes, preferentially deactivate the centers which have the lowest ability to copolymerize α-olefins with ethylene. These poisons significantly increase the content of the olefin-rich components in the copolymers.  相似文献   

3.
Sequential polymerizations were carried out using a high-activity MgCl2/Ziegler-Natta catalyst to evaluate whether the in situ preparation of polypropylene/1-butene alloys was possible inside the reaction vessel and analyze the effects of 1-butene on the final material properties. Propylene/1-butene alloy resins were synthesized in a sequential two-stage process. In the first stage, liquid pool propylene polymerizations were carried out in batch. In the second stage, 1-butene was polymerized inside the polypropylene matrix in gas-phase in semibatch mode. According to the obtained results, it is possible to incorporate 1-butene upon the polypropylene matrix inside the reactor at very low pressures, without affecting the properties of the continuous polypropylene matrix significantly.  相似文献   

4.
5.
A soluble ethylene catalyst were obtained by mixing a methylene dichloride solution of dichlorobis(γ-cyclopentadienyl) titanium (Cp2TiCl2) with a heptane solution of ethylaluminium sesquichloride (Al2Et3Cl3) or of diethylaluminium chloride (AlEt2Cl). Ethylene was polymerized using these catalysts; the solution was examined by electron spin resonance technique before the polymerization and during the reaction. The catalyst activity remained constant for a long period, and the polymerization went on at the same rate for 6–8 hr. The mechanism of the reaction is discussed.  相似文献   

6.
7.
This article reveals the effects of mixed activators on ethylene polymerization and ethylene/1-hexene copolymerization over MgCl?/SiO?-supported Ziegler-Natta (ZN) catalysts. First, the conventional ZN catalyst was prepared with SiO? addition. Then, the catalyst was tested for ethylene polymerization and ethylene/1-hexene (E/H) co-polymerization using different activators. Triethylaluminum (TEA), tri-n-hexyl aluminum (TnHA) and diethyl aluminum chloride (DEAC), TEA+DEAC, TEA+TnHA, TnHA+ DEAC, TEA+DEAC+TnHA mixtures, were used as activators in this study. It was found that in the case of ethylene polymerization with a sole activator, TnHA exhibited the highest activity among other activators due to increased size of the alkyl group. Further investigation was focused on the use of mixed activators. The activity can be enhanced by a factor of three when the mixed activators were employed and the activity of ethylene polymerization apparently increased in the order of TEA+ DEAC+TnHA > TEA+DEAC > TEA+TnHA. Both the copolymerization activity and crystallinity of the synthesized copolymers were strongly changed when the activators were changed from TEA to TEA+DEAC+TnHA mixtures or pure TnHA and pure DEAC. As for ethylene/1-hexene copolymerization the activity apparently increased in the order of TEA+DEAC+TnHA > TEA+TnHA > TEA+DEAC > TnHA+DEAC > TEA > TnHA > DEAC. Considering the properties of the copolymer obtained with the mixed TEA+DEAC+TnHA, its crystallinity decreased due to the presence of TnHA in the mixed activator. The activators thus exerted a strong influence on copolymer structure. An increased molecular weight distribution (MWD) was observed, without significant change in polymer morphology.  相似文献   

8.
Unsteady diffusion kinetics, recently advanced by this laboratory, is applied to the examination of some polymerization and molecular chain structure problems. Hitherto deemed “anomalous” phenomena, such as the faster rate of copolymerization of ethylene/α-olefin than the homopolymerization of ethylene and the enrichment in the incorporation of a higher α-olefin in its copolymerization with ethylene by a lower α-olefin, are reasonably explained by unsteady diffusion of monomers. Molecular chain structure of copolymers, such as compositional heterogeneity and its dependence on comonomer incorporation originates from the difference in diffusion coefficients of the monomers. A copolymer composition equation taking into consideration the unsteady diffusion was developed. In cases where simulated curves were compared with experimental curves, good agreements were found.  相似文献   

9.
10.
A modified chromium oxide supported catalyst has been developed and applied in industry for the manufacture of LPPE via the gas-phase (co)polymerization of ethylene. The catalyst contains surface chromium oxide in the oxidation number Cr2+, two modifiers (aluminum oxide and fluorine surface compounds), and silicon dioxide as a support. The activity of the new chromium oxide catalyst in the gas-phase copolymerization of ethylene with 1-butene is higher by a factor of 4–5 than that of the traditional commercial catalytic system based on the supported bis(triphenylsilylchromate) catalyst. An increased reactivity of 1-butene in its copolymerization with ethylene in the presence of the chromium oxide catalyst makes it possible to reduce the consumption of 1-butene in the synthesis of a linear medium-density PE (0.937–0.938 g/cm3). Gas pipes made of PE prepared with the new catalyst are characterized by improved resistance to crack propagation.  相似文献   

11.
The copolymerization of 3-methyl-1-butene (3M1B), 2-methyl-2-butene (2M2B), or 2-methyl-1-butene (2M1B) with trans-2-butene (2B) was attempted in the presence of a Ziegler-Natta catalyst. It was found the 3M1B underwent monomer-isomerization copolymerization with 2B to give a copolymer consisting of both 3M1B and 1-butene (1B) units, with an infrared (IR) spectrum in good agreement with that obtained from the copolymerization of 3M1B with 1B under similar conditions. When the apparent copolymerization parameters obtained by a TiCl3–(C2H5)3Al catalyst were compared, the apparent reactivity of 3M1B observed in the 3M1B-2B system was much higher than that in the 3M1B-1B system. However, 2M2B and 2M1B did not undergo monomer-isomerization copolymerization with 2B, and only the homopolymer of 1B was obtained under similar conditions.  相似文献   

12.
Poly(propylene) samples produced by heterogeneous Ziegler-Natta catalysts containing different internal electron donors were fractionated by using temperature rising elution fractionation; some key fractions were analyzed with 13C NMR. It was found that internal electron donors with different structures can convert aspecific active sites into different isospectrific ones. The employment of bis(2-ethylhexyl) phthalate as internal electron donor leads to chemically inverted structures in the atactic fraction. This suggests that an internal electron donor may also exist in the environment of aspecific active sites.  相似文献   

13.
The effect of photodegradation in isotactic poly(1-butene) (PB-1) have been investigated using rheology, differential scanning calorimetry and infrared spectroscopy. Two commercially available grades of PB-1 with different average molecular weight were chosen. Specimens prepared by compression moulding were UV irradiated in the interval from 0 to 70 h. UV-induced changes in molecular structure have been followed by evolution of rheological properties, thermal properties and degradation by-products. Thermal analysis showed significant changes in crystallization behaviour influencing morphology and resulting thermal properties. Moreover it has been confirmed that the degradation significantly retards the phase transformation. Rheological measurement has been found as an effective method for determination of early stages of photodegradation of PB-1.  相似文献   

14.
2-Pentene and 2-hexene were found to undergo monomer-isomerization copolymerizations with 2-butene by Al(C2H5)3–VCl3 and Al(C2H5)3–TiCl3 catalysts in the presence of nickel dimethylglyoxime or transition metal acetylacetonates to yield copolymers consisting of the respective 1-olefin units. For comparison, the copolymerizations of 1-pentene with 1-butene and 1-hexene with 1-butene by Al(C2H5)3–VCl3 catalyst were also attempted. The compositions of the copolymers obtained from these copolymerizations were determined by using the calibration curves between the compositions of the respective homopolymer mixtures and the values of D766/D1380 in the infrared spectra. The monomer reactivity ratios for the monomer-isomerization copolymerizations of 2-butene (M1) with 2-pentene and 2-hexene, in which the concentrations of both 1-olefins calculated from the observed isomer distribution were used as those in the monomer feed mixture, and for the ordinary copolymerizations of 1-butene (M1) with 1-pentene and 1-hexene by Al(C2H5)3-VCl3 catalyst were determined as follows: 2-butene (M1)/2-pentene (M2): r1 = 0.14, r2 = 0.99; 1-butene (M1)/1-pentene (M2): r1 = 0.30, r2 = 0.74; 2-butene (M1)/2-hexene (M2): r1 = 0.11, r2 = 0.62; 1-butene (M1)/1-hexene (M2): r1 = 0.13, r2 = 0.90.  相似文献   

15.
The kinetics of isoprene polymerization catalyzed by VCl3 and Et3Al were studied by measuring fractional conversions, polymer composition, and molecular weight distributions at a series of reaction times and temperatures. The rate of polymerization plotted against temperature shows an inflection point with a minimum and maximum in the 60–90°C range. The isomeric composition of the polymer changes with temperature but not with reaction time, while the molecular weight distribution undergoes substantial change with both of these variables. The rate of polymerization at sites producing low molecular weight polymers was measured, and the activation energy calculated to be about 10 kcal/mole. The active sites were found to deactivate at different rates. The results support the hypothesis that several species of active sites are present in the system and that these exhibit characteristic polymerization behavior.  相似文献   

16.
New models for the steric environment of Ti isospecific polymerization sites for poly(propylene) on MgCl2 microcrystals are proposed. They directly involve a donor molecule in order to obtain isospecific activable Ti atoms otherwise belonging to isolated adsorbed TiCl4 molecules or Ti2Cl8 dimers which are lacking of the required chirality for stereocontrol. The donor molecules able to attain at best this effect keep to some peculiar conformational rules settled by the authors in a previous theoretical-correlative study on highly active Lewis bases. The new 1,3-dimethoxypropane series suggested by the authors and recently patented by Montell has been examined in detail. Essentially three different types of closeness between Ti atoms and donor molecules can take place, in which different moieties of the diether compound help to build the ‘right’ steric environment in the site's neighbouring. In the three proposed models S1, S2, S3 the stereocontrol is attained through, respectively, one of the methoxy moieties, one of the methyls, and one of the central carbon atom substituents. New hypotheses on the role of Lewis bases in the preparation of isospecific heterogeneous Ziegler-Natta catalysts are discussed.  相似文献   

17.
Poly-(4-phenyl-1-butene) was prepared by using a titanium tetrachloride–triethyl-aluminum catalyst. The crystalline polymer melts at 158°C. Double orientation could not be obtained but all the reflections on the x-ray fiber diagram can be indexed on the basis of an hexagonal cell (a = 20.8 A., c (fiber repeat) = 6.61 A.). However the crystal structure proposed does not belong to an hexagonal space group but to the monoclinic (pseudo-orthorhombic) space group Pa (a = 10.4 A., b = 18.0 A., c = 6.61 A.). The 31 helix, common to systems of this type, is consistent with an isotactic (not syndiotactic) configuration for the polymer chain.  相似文献   

18.
19.
Polymerization of styrene with the neodymium phosphonate Nd(P507)/H2O/Al(i-Bu)3 catalytic system has been examined. The polymer obtained was separated into a soluble and an insoluble fraction by 2-butanone extraction. 13C-NMR spectra indicate that the insoluble fraction is isotactic polystyrene and the soluble one is syndiotactic-rich atactic polystyrene. The polymerization features are described and discussed. The optimum conditions for the polymerization are as follows: [Nd] = (3.5–5.0) × 10−2 mol/L; [styrene] = 5 mol/L; [Al]/[Nd] = 6–8 mol/mol; [H2O]/[Al] = 0.05–0.08 mol/mol; polymerization temperature around 70°C. The percent yield of isotactic polystyrene (IY) is markedly affected by catalyst aging temperature. With increase of the aging temperature from 40 to 70°C, IY increases from 9% to 48%. Using AlEt3 and Al(i-Bu)2H instead of Al(i-Bu)3 decreases the yield of isotactic polystyrene. Different neodymium compounds give the following activity order: Nd(P507)3 > Nd(P204)3 > Nd(OPri)3 > NdCl3 + C2H5OH > Nd(naph)3. With Nd(naph)3 as catalyst, only atactic polystyrene is obtained. © 1998 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 36: 1773–1778, 1998  相似文献   

20.
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