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.
Formation, deactivation and transformation of stereospecific active sites on TiCl4/dibutylphthalate (DBP)/Mg(OEt)2 Ziegler–Natta catalyst induced by short time reaction with triethylaluminum (TEA) cocatalyst (with TEA pretreatment time from 0 to 600 s) were investigated by stopped-flow propene polymerization combined with temperature rising elution fractionation (TREF) and GPC methods. It was demonstrated that both formation and deactivation of active sites with broad multiplicity in isospecificity on the catalyst are slow reactions with an induction period of ca. 0.2 s. It was most important to find that the formation of active sites with the highest isospecificity strongly depends on the interaction between the catalyst and cocatalyst (up to 60 s of pretreatment) even in the presence of internal donor. This newly observed phenomenon (according to our knowledge) suggested that the transformation of monometallic active sites (aspecific or less isospecific) into bimetallic active sites (highly isospecific) through reversible complexing with TEA cocatalyst (or its reaction product diethylaluminum chloride (DEAC)) in Ziegler–Natta catalysts cannot be overlooked even in the presence of internal electron donor. The existence of –OC2H5 ligand in the catalyst most probably gave birth to a new group of active titanium species. The stability of active sites increases with increasing isospecificity in the early stage of pretreatment (up to 60 s of pretreatment). While all the active sites became relatively stable in the later stage of pretreatment (from 60 to 600 s of pretreatment). The extraction of internal donor DBP by TEA from the catalyst within the pretreatment procedure is found to initiate from 60 s of pretreatment resulting in slight transformation of isospecific active sites into aspecific sites.  相似文献   

2.
Monocyclopentadienyltitanium trichloride (CpTiCl3) was directly immobilized on silica surface. The resulting CpTiCl3/SiO2-MAO was investigated in view of its suitability for syndiospecific polymerization of styrene using ESR spectroscopy. Polymerization results show that both the syndiotacticity and the activity decrease upon immobilization. A possible explanation is that the active sites may become aspecific after immobilization and part of the immobilized active sites can be leached into the solution in the presence of MAO.  相似文献   

3.
A novel layered calcosilicate (CAS-1) was employed to immobilize an iron-based diimine catalyst 2,6-bis[1-(2,6-diisopropylphenylimino)ethyl]pyridine iron chloride (I) onto it to form a supported catalyst (CC) for the first time. The crystal structure of CAS-1 was determined by X-ray crystallographic analysis in addition to SEM characterization. The CC-catalyzed ethylene polymerization exhibited good catalytic activities with either co-catalyst methylaluminoxane (MAO) or triethylaluminum (TEA). The resulting polyethylenes possessed not only higher molecular weight, melting temperature (Tm), and decomposition onset temperature (Tonset) than those obtained with its homogeneous counterpart, but also a unique morphology.  相似文献   

4.
Phillips catalyst has been contributing to about 40% of world high‐density polyethylene production because of its ability to give products with unique microstructures like broad molecular weight distribution as well as short and long chain branches. Even after 50 years' effort, some crucial problems concerning the nature of active sites, polymerization, and branching mechanisms are still kept mysterious. In this work, ethylene and 1‐hexene copolymerization with Phillips catalyst prereduced by CO was carried out in the presence of triethyl aluminum (TEA) cocatalyst. The microstructures of polymers were investigated by 13C NMR and gel permeation chromatography (GPC) methods. A hybrid‐type kinetics was found for both homo‐ and copolymerization kinetics, which indicated that there existed two types of active sites namely site A and site B. Site A with instant activation, high activity, and fast decay was transformed from a metathesis site, namely Cr(II) site, coordinated with CO or CO2 through desorption of CO or CO2 by TEA, which contributed to the formation of short chain branches, especially methyl branches. Site B with slow activation, low activity, and slow decay was generated from reduction of residual chromate (VI) by TEA. Both 1‐hexene and TEA can decrease the molecular weight of polyethylene as well as enhance short chain branching. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 4632–4641, 2005  相似文献   

5.
Using triethylaluminum (TEA), triisobutylaluminum (TIBA) or TEA/TIBA mixtures of molar ratio 75/25, 50/50 and 25/75 as the cocatalyst, five different ethylene-propylene copolymer samples were synthesized by a MgCl2/SiO2/TiCl4/diester type Ziegler-Natta catalyst in a slurry polymerization process. The synthesized copolymers are strongly heterogeneous in chain structure and were fractionated into part of nearly random copolymer and part of segmented copolymer. Both polymerization activity and copolymer structure were found to be markedly changed when the cocatalyst was changed from TEA to TEA/TIBA mixtures or pure TIBA. As the content of TEA in cocatalyst increases, yield of the random part of product increases and the yield of the crystalline segmented copolymer part decreases. There is also a decrease in ethylene content of the whole product with increasing TEA amount. Copolymerization behaviors of the TEA/TIBA mixture activated catalysis systems are not simple superposition of those activated by pure TEA and TIBA. When a 50/50 TEA/TIBA mixture was used as cocatalyst, the copolymerization activity became the highest, and yields of both the random copolymer part and the segmented copolymer part are close to the highest level. On the other hand, both parts of the copolymer produced with a 50/50 TEA/TIBA mixture are relatively more blocky than the products of TEA or TIBA systems, and difference in ethylene content between the random part and the segmented part was the smallest. The segmented copolymer part of three typical samples was further fractionated by temperature-gradient extraction fractionation into fractions of different ethylene content and sequence distribution. Changing TEA content in the cocatalyst exerted strong influences also on the fraction distribution of the segmented part of copolymer.  相似文献   

6.
Phenylacetylene and its homologues bearing methoxy or chloro groups in the para position of the phenyl ring were stereoregularly polymerized by Rh complex catalysts, e.g., [Rh (norbornadiene) Cl]2, in the presence of various solvents. The 1H-NMR and UV study of the Rh complex showed that the activity of the Rh catalysts greatly depends upon not only the ligand in the Rh catalyst but also on the solvent used, e.g., polyphenylacetylene with a highly cis-transoidal content was quantitatively obtained when norbornadiene was coordinated as the ligand, and triethylamine (TEA) was used as solvent, whereas tetrahydrofuran (THF) and benzene reduced the cis content. It was also found that the Rh catalyst as a dimer complex dissociates into the monomeric species not only when TEA was used but also when alcohol, benzene, and THF were used. The resulting monomeric species was considered as the important propagation species in this polymerization. The sharp line width observed in the 1H-NMR spectrum of polyphenylacetylene and its homologues which were prepared in ethanol or TEA was interpreted by the narrow chemical shift distribution arising from the high cis content. © 1994 John Wiley & Sons, Inc.  相似文献   

7.
The copolymerization of ethylene and butadiene with a supported titanium catalyst (TiCl4/MgCl2/EB/Φ2SiCl2/AlEt3) is described. The resulting products were characterized by IR, 13C-NMR, x-ray diffraction, differential thermal analysis, electron microscopy, and solvent extraction. It was found that the butadiene units are substantially in trans-1,4 configuration and blocked sequences. Both ethylene and butadiene blocks form crystalline phases. The presence of unsaturated bonds made it possible to graft MMA and maleic anhydride. The influences of monomer composition, temperature, Al/Ti ratio, catalyst concentration, and solvents on the copolymerization were investigated.  相似文献   

8.
Fourier transform infrared (FTIR) spectra were obtained for a typical MgCl2-supported, high-mileage catalyst for propylene polymerization. When ball-milling MgCl2 with ethyl benzoate (EB), the latter is incorporated into the support (I) by Lewis acid-base complexation involving both oxygen atoms of the ester. Reaction of (I) with p-cresol (PC) resulted in a material (II) that contains all the characteristic IR bands of PC. The reaction of (II) that contains all the characteristic IR bands of PC. The reaction of (II) with AlEt3 (TEA) resulted in (III) whose spectrum supports the reaction observed by product analysis and NMR spectroscopy. There was no evidence of any reaction between TEA and EB. Further reaction of (III) with an excess of TiCl4 caused substantial removal of the p-cresol moiety as shown by the diminution of its characteristic bands. Finally, activation with 3TEA-1MT (methyl-p-toluate) complexes gave spectra that revealed the presence of MT in the activated catalyst without any visage of p-cresol moiety. The nondestructive FTIR method, however, is not quantitative. Quantitative analysis of the organic components in the support materials (I), (II), and (III) and the catalysts was accomplished by hydrolysis of the inorganic components, extraction with ether, and analysis by gas chromatography. The results are in good agreement with composition deducted from elemental analysis and substantiate the FTIR conclusions.  相似文献   

9.
Pt-based catalysts cannot be used permanently for the diesel after-treatment system because the catalytic activity is decreased due to coarsening of Pt particles at high temperature of the exhaust gas. In this study, to prevent Pt-based catalyst from deactivation, Nd was added to the Pt/SiO2 catalyst, and the effect of the Nd addition on the catalytic activity was investigated. The Pt/SiO2 catalyst showed a high catalytic activity for the oxidation of NO but was severely deactivated after the fast thermal aging process. Pt crystallite size was increased and some Pt particles were buried in the SiO2 pore during the fast thermal aging process, which led to the decrease of catalytic activity. Nd-added Pt/SiO2 catalyst showed lower activity than Pt/SiO2 catalyst, but Pt–Nd/SiO2 catalyst maintained its catalytic activity after fast thermal aging process. It can be postulated that a stable Nd silicate, on which Pt particle is placed, protects SiO2 pores from destruction and so the number of the catalytically active sites remains nearly unchanged. As a result the Pt–Nd/SiO2 catalyst maintained its catalytic activity after fast thermal aging process.  相似文献   

10.
The relation between stereoregularity of active centers on a MgCl2/TiCl4 catalyst and functions of inside donor (ID) and outside donor (OD) was investigated in the case of ethyl benzoate (EB)/methyl p-toluate (MPT) as an ID/OD pair. The results indicate that stereregularity depends merely on the amount of MPT supported on the catalyst. On the other hand, regioregularity of active centers was investigated with a MgCl2/TiCl4/dioctyl phthalate(DOP)-Et3Al/diphenyldimethoxysilane(DPDMS) catalyst system. Regio-irregular inserted units were detected only in end groups of PP. It indicates that regio-irregular insertion leads to dormant centers with respect to propene insertion, though such centers are active for hydrogen transfer.  相似文献   

11.
The kinetics of propylene polymerization by superactive CH-catalyst prepared from toluene solution of MgCl2 · EH/PA/TiCl4–TEA/PES was investigated. The results are compared with CW-catalyst prepared from crystalline MgCl2/EB/PC/TEA/TiCl4–TEA/MPT (abbreviations given in the text). The former is four times more active than the latter and produces more isotactic polypropylene. The CH-catalyst has 25% of the Ti as isospecific sites as compared to 6.7% for the CW-catalysts. These sites have the same rate constant of propagation so that the higher polymerization activity of the CH-catalyst is attributable simply to a greater number of active sites. Differences in the kinetics of deactivation and of chain transfer for the two catalysts are described.  相似文献   

12.
Hydrogen effects for propylene polymerization were investigated with ultra low TiCl3 loading MgCl2-supported catalysts in which the electric states of Ti species can be almost uniform. Hydrogen did not affect the catalyst activity, while the efficiency of hydrogen as a chain transfer agent was found to depend on the Ti content of the catalyst and the stereospecificity of the polymerization sites: Hydrogen was effective for isospecific sites independent of Ti contents, but inert for aspecific sites only at the extremely low Ti content. These results were explained within the island model, where isospecific sites may be located in the islands with other Ti species in their surroundings acting, as a steric hindrance for isospecific polymerization and as hydrogen dissociation sites after deactivation. Most of the aspecific sites should be isolated only at the extremely low Ti content. These isolated sites have no other Ti species in their surroundings, i.e. no hydrogen dissociation sites, and are inert to hydrogen.  相似文献   

13.
The stopped‐flow polymerization of propylene was carried out using an internal donor‐free MgCl2‐supported Ziegler catalyst in the absence or presence of external electron donors. The variation in the isospecific active sites was investigated based on the isotacticity distribution of the poly(propylene) analyzed by the TREF method. Highly isospecific active sites derived from the highest isotactic fraction (elution temperature by TREF: >112 °C) exist in the electron donor‐free catalyst system. The addition of external electron donors converted parts of the aspecific into isospecific active sites, but showed no effects on the highest isospecific active sites. The external electron donor sterically affects a coordination vacancy of each aspecific titanium species and, consequently, transfers it into an isospecific active site of high, but not highest isospecificity.  相似文献   

14.
The effect of chemical structure of 2,2′-disubstituted 1,3-dimethoxypropane (so-called 1,3-diether) on the performance of Ziegler-Natta (ZN) catalyst was investigated by using density functional theory and molecular mechanics. Calculation of the energy barrier during insertion of propylene reveals that the isospecific active site created on the (1 0 0) surface of MgCl2 is more active than the aspecific active site created on the (1 1 0) surface of MgCl2 for propylene polymerization. When the adsorption energies of various 1,3-diethers are calculated and analyzed in terms of isotacticity, it is found that the isotacticity of polypropylene increases as 1,3-diether is adsorbed more preferentially on the (1 1 0) surface. Since analysis of energetics for insertion of propylene into the active site created on the (1 1 0) surface with 1,3-diether coordinated to Mg atom in the vicinity of the active site reveals that the coordination of 1,3-diether does not transform the aspecific active site on the (1 1 0) surface into isospecific one, it is concluded that the primary function of 1,3-diether is to prevent the formation of aspecific site on the (1 1 0) surface, without significant decrease in the number of the isospecific active site created on the (1 0 0) surface. A systematic analysis of various model compounds for 1,3-diether suggests that the substitution of highly branched hydrocarbon at the C2 position of 1,3-diether results in better performance of ZN catalyst.  相似文献   

15.
采用X射线衍射、扫描电镜、原子吸收、程序升温还原、X射线吸收近边吸收谱、X射线光电子能谱、氮吸附等手段对水热合成(HS)、等体积浸渍(PVI)与离子交换(IE)法制备的CuSAPO-34样品进行了表征,并评价了老化前后催化剂上C3H6-SCR与NH3-SCR脱除模拟柴油车尾气中NOx的反应活性.结果表明,IE法制得的催化剂活性最高,尤其在C3H6-SCR低温阶段;PVI法制得的催化剂活性最差.制备方法影响CuSAPO-34催化剂的比表面积、孔径分布和活性组分价态从而改变其催化活性.各催化剂均存在Cu+和Cu2+,但比例明显不同.HS样品以Cu2+为主,另两种样品则含较多的Cu+.老化不仅部分破坏了分子筛的形貌、降低了分子筛的比表面积,尤其是表面Cu含量,而且有部分Cu生成了CuSO4,使得老化后催化剂的脱硝活性降低.PVI法制得的催化剂老化后活性下降幅度最小,表明该分子筛抗老化能力较强.  相似文献   

16.
Intercalation of cetyltrimethylammonium (CTA+) cations within the nanolayers of montmorillonite (MMT) clay followed by reaction with Keggin-type phosphomolybdic acid (PMo) resulted in the synthesis of (CTA)3PMo-MMT nanocomposite catalyst. The prepared nanocomposite catalyst was characterized using different physicochemical methods such as Fourier-transform infrared and inductively coupled plasma–optical emission spectroscopies, X-ray diffraction, and nitrogen adsorption–desorption (Brunauer–Emmett–Teller method) analyses. Characterization techniques demonstrated the intercalation of (CTA)3PMo species into the nanolayers of MMT. The resulting (CTA)3PMo-MMT nanocomposite catalyst efficiently catalyzed the synthesis of 2,4,5-trisubstituted imidazoles under solvent-free conditions. The efficiency is due to the fact that the presence of CTA+ species makes the nanocomposite catalyst hydrophobic and facilitates the accessibility of hydrophobic reactants to active sites in the course of the reaction. High activity and selectivity were achieved in the presence of the prepared nanocomposite catalyst. The nanocomposite catalyst was readily isolated from the reaction mixture using simple filtration, washed with ethanol, and recycled five times without a major loss of activity.  相似文献   

17.
The poisoning effect of CO2 on a HY catalyst for the dehydration of 2-(2-hydroxyethyl)-pyridine (HEP) to 2-vinylpyridine (VP) has been investigated by FT-IR analysis in the presence and absence of pyridine. CO2 was found to adsorb on sites not occupied by pyridine,i.e. on weak basic sites accompanying the strongly acidic sites, characteristic of Y zeolites in the protonated form. When the basic sites are occupied by preadsorbed CO2, HEP dehydration cannot take place any more through the minor mechanism, involving a couple of acid-base sites, and the reaction proceeds only through the major mechanism, involving a carbocation intermediate, on Br?nsted acid sites only.  相似文献   

18.
In recent years, the dramatic increase of the prices of Pd and Rh brings about increasingly heavier cost pressure for the exhaust gas after-treatment industry, thus it is of great significance to develop advanced Pd-Rh bimetallic catalysts with low metal loading. In this work, on the basis of conventional catalyst composed of Al2O3-supported Pd and CZ (ceria-zirconia-based oxide)-supported Rh, an advanced Pd-Rh catalyst was obtained by adjusting the synthesis procedure. That was, a small portion of Rh was co-impregnated with Pd on Al2O3, while the other portion was still supported on CZ. In this way, the co-impregnated Rh may act as an auxiliary for Pd/Al2O3 component. As revealed by the various characterization results, the agglomeration of Pd species is well suppressed, and larger amount of PdO species is maintained for the modified catalyst after hydrothermal aging treatment. On the other hand, the state of Rh species also appears to be somewhat modified. Consequently, after hydrothermal aging treatment, the modified catalyst behaves much better for the catalytic conversions of CO, HC and NO.  相似文献   

19.
刘柏平 《高分子科学》2013,31(4):591-600
The formations of defective MgCl2 surfaces, and subsequent adsorption of Ti species and electron donor, as well as propylene polymerization over the Ziegler-Natta catalyst have been investigated using density functional theory (DFT) method. Twelve possible support models of regular and defective MgCl2 (110) and (100) surfaces were built. The individual adsorptions of titanium chlorides as mononuclear or dinuclear, and ethyl benzoate (EB) as electron donor, on these models were evaluated. The analysis of energies presented the cases of EB adsorption were generally more stable than titanium chlorides on both surfaces. Thus, EB as internal electron donor mainly prevented TiCl4 from coordinating on the MgCl2 surfaces where mostly non-stereospecific active sites could be formed. Exceptionally, A5 the site model with terminal Cl-vacancy on the MgCl2 support, presented stronger adsorption of TiCl4 than that of EB on (110) surface. Since the TiCl4 and ethyl benzoate (EB) would compete to adsorb on the support surface, it seems reasonable to assume that TiCl4 might predominately occupy this site, which can act as the most plausible active site for propylene polymerization. The first insertion of propylene monomer into the A5 active site model showed that it exhibited good regioselectivity but poor stereospecificity in the absence of electron donor.  相似文献   

20.
在固定床高压微反装置上考察了预硫化型NiMoS/γ-Al2O3催化剂上二苯并噻吩(DBT)加氢脱硫(HDS)反应和吲哚加氢脱氮(HDN)反应之间的相互影响。结果表明,吲哚对DBT的加氢脱硫反应具有抑制作用,其中对加氢路径(HYD)比对氢解路径(DDS)的抑制作用强,温度升高后,吲哚的抑制作用减弱。吲哚对DBT加氢脱硫反应的抑制作用源于吲哚及其HDN反应的中间产物在活性位上的竞争吸附。DBT和原位生成的H2S促进了催化剂表面硫阴离子空穴(CUS)向B酸位的转化,从而提高1,2-二氢吲哚(HIN)分子中C(sp3)—N键的断裂能力,使得吲哚的转化率和产物中邻乙基苯胺(OEA)的相对含量增大。HDN活性相的形成虽然需要硫原子的参与,但是活性相的保持并不需要大量的硫原子,较高含量硫化物存在时加氢活性位减少,不利于脱氮反应。  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