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 31 毫秒
1.
The behaviour of the homogeneous catalyst system Ti(Oi-C3H7)4/Al(C2H5)2F in ethylene/propene copolymerization was studied. The copolymer structure is discussed in view of the stereoregulation mechanism previously proposed for vanadium-based homogeneous catalyst systems.  相似文献   

2.
The influence of ethyl benzoate (E.B.) on the copolymerization of styrene with 1-hexene, initiated by isospecific Ziegler-Natta catalysts α-TiCl3(H)-AlEt3 and TiCl3 · 1/3 AlCl3-AlEt3 (mole ratio Al/Ti = 3), was examined. Ethyl benzoate was found to reduce the activity of the catalysts. In addition it leads, depending on the Ti catalyst, to opposite effects on the apparent reactivity order of the two monomers. Incorporation of styrene into the copolymers is reduced when E. B. is added to TiCl3 · 1/3 AlCl3-AlEt3. On the contrary, a much higher incorporation of styrene is observed with α-TiCl3(H)-AlEt3 in the presence of E. B. For this system the calculated reactivity ratio varies strongly with increasing proportion of E. B.: for a mole ratio AlEt3/E. B. = 3, rS = 0,94 and rH = 1,46 and for AlEt3/E. B. = 2, rS = 8 and rH = 0,1. Changes in the stereoregularity of copolymers suggest that E. B. leads to an inhibition of the less stereospecific sites for TiCl3 · 1/3 AlCl3-AlEt3, whereas its addition suppresses the stereospecificity of the α-TiCl3(H)-AlEt3 catalyst. Contributions of conventional cationic and/or radical processes to the copolymerization reaction were examined and may be ruled out.  相似文献   

3.
The chain propagation rate constants for the polymerization of ethylene and propene in the presence of δ-TiCl3/Al(CH3)3 at 22 °C are determined by means of a 13C NMR analysis of suitable block copolymers. The numerical values of the rate constants are compared with those previously reported.  相似文献   

4.
The allylneodymium chloride complexes Nd(C3H5)2Cl·1.5 THF and Nd(C3H5)Cl2·2 THF can be activated by adding hexaisobutylaluminoxane (HIBAO) or methylaluminoxane (MAO) in a ratio of Al/Nd = 30 for the catalysis of butadiene 1,4‐cis‐polymerization. A turnover frequency (TOF) of about 20 000 mol butadiene/(mol Nd·h) and cis‐selectivity of 95–97% are achieved under standard conditions ([BD]0 = 2 m, 35°C, toluene). Molecular weight determinations indicate a low polydispersity (w (LS)/n (LS) = 1–1.5), the formation of only one polymer chain per neodymium and the linear increase of the degree of polymerization (DP) with the butadiene conversion, as observed for living polymerizations. First indications of chain‐transfer reaction occur only at the highest conversion or degree of polymerization. The rate law rP = kP[Nd][C4H6]1.8 is derived for the catalyst system Nd(C3H5)2Cl·1.5 THF/HIBAO and for the system Nd(C3H5)Cl2·2 THF/MAO the rate law rP = kP[Nd] [C4H6]2 with kP = 3.24 L2/(mol2·s) (at 35°C). Taking into account the Lewis acidity of the alkylaluminoxanes and the characteristic coordination number of 8 for Nd(III) in allyl complexes the formation of an η3‐butenyl‐bis(η4‐butadiene)neodymium(III) complex of the composition [Nd(η3‐RC3H4)(η4‐C4H6)2(X‐{AlOR}n)2] is assumed to be a single‐site catalyst for the chain propagation by reaction of the coordinated butadiene via the π‐allyl insertion mechanism and the anticis and syntrans correlation to explain the experimental results.  相似文献   

5.
The catalytic activity of the system Al(C2H5)3/H2O was investigated. Instead of pure water aqueous solutions of inorganic bases were reacted with Al(C2H5)3 in toluene/diethyl ether media. A uniform soft gel-like product was readily obtained in this reaction (H2O/Al(C2H5)3 ~ 1.0–1.5) differing from the heterogeneous nature of the product obtained with pure water. The polymerization of acetaldehyde proceeded with higher stereo-regularity in this system Al(C2H5)3/aqueous inorganic base. The polymerization behavior and the stoichiometric reaction between the catalyst and the monomer (ethyl acetate formation) indicate that there seem not to be significant differences in the active species of Al(C2H5)3/aqueous inorganic base and Al(C2H5)3/H2O systems. Improved stereospecific polymerization in the presence of inorganic bases is discussed with respect to the formation of a cross-linked product in the Al(C2H5)3/H2O system in the preparation of the catalyst.  相似文献   

6.
The effect of varying 14CO contact time upon the concentration of active centres C* in ethylene homopolymerization using the TiCl4/MgH2 · Al(C2H5)3 catalytic system shows that the polymer radioactivity, and hence C*, increase sharply in the first sixty minutes of 14CO contact with the polymerization centres. For contact times longer than one hour, the polymer radioactivity continues to increase, but very slowly. Studies on the effect of the mole ratio Al(C2H5)3/Ti on ethylene homopolymerization show that both the catalytic activity and C* increase sharply when increasing the mole ratio Al(C2H5)3/Ti in the range from 5 to 20. When increasing the mole ratio Al(C2H5)3/Ti above 50, C* tends to decrease very slightly. In ethylene/1-hexene copolymerization a similar effect of the mole ratio Al(C2H5)3/Ti on the polymerization is observed. However, even though the catalytic activity in copolymerization is observed to be higher than in homopolymerization, at the same mole ratio Al(C2H5)3/Ti, yet C* in both cases is found to be more or less the same.  相似文献   

7.
The reaction of (C5(CH3)5)Ti(CH2C6H5)3 with B(C6F5)3 in chlorobenzene at 25°C produces a mixture of [(C5(CH3)5)Ti(CH2C6H5)2]+ [B(CH2C6H5)(C6F5)3]? ( 1 ) and Ti(III) complexes which have been characterized by NMR and electron spin resonance (ESR) spectroscopy, respectively. Spectroscopic data combined with polymerisation activity measurements suggest that a Ti(III) complex is the active species in the styrene syndiospecific polymerisation.  相似文献   

8.
Polymerization of ethylene was carried out in toluene using the homogeneous catalytic system Cr[OC(CH3)3]4/Al(C2H5)2Cl, combined with various metal chlorides (MtClx). The polymerization activity was found to be strongly dependent upon the MtClx used. A clear correlation was found between the activity and the elctronegativity X(Mtx+) of the metal ion in MtClx. MtClx containing metal ions with X(Mtx+) < 8,5 (electronegativity of the active Cr2+) show a markedly increased activity, whereas those with X(Mtx+) > 8,5 have a rather decreased activity. Thus, for example, the catalyst combined with MgCl2 shows very high activity to give a linear polyethylene with a remarkably high molecular weight and narrow molecular weight distribution. A plausible mechanism for the enhancement of the activity by suitable MtClx is proposed on the basis of these results.  相似文献   

9.
The η3, η2, η2-dodeca-2(E), 6(E), 10(Z)-trien-1-yl-nickel(II) complexes [Ni(C12H19)]X (X = SbF6, O3SCF3) were treated in toluene with amorphous aluminium trifluoride (which was prepared from AlEt3 and BF3 · OEt2) in a mole ratio 1 : 10 to 20, forming a highly active catalyst for the 1,4-cis polymerization of butadiene. This catalyst is comparable in its activity and selectivity, and in the molar mass distribution of the polybutadiene, with the technical nickel catalyst Ni(O2CR)2/BF3?OEt2/AlE3 developed by Bridgestone Tire Company thirty years ago. The existence of the C12-allynickel(II) cation [Ni(C12H19)]+ on the AlF3 support could be proved by FAB mass spectroscopic measurements. In agreement with our reaction model for the allyl nickel complex catalyzed butadiene polymerization, it is concluded that the technical nickel catalyst in its effective structure can be described as a polybutadienylnickel(II) complex co-ordinated to a polymeric fluoroaluminate anion via a fluoride bridge.  相似文献   

10.
Copolymers sPS‐B consisting of blocks of syndiotactic polystyrene (sPS) and polybutadiene (B) have been prepared using CpTiX3 (Cp = C5H5, X = Cl, F; Cp = C5Me5, X = Me) and TiXn (n = 3, X = acetylacetonate (acac); n = 4, X = O‐tert‐Bu) activated with methylaluminoxane (MAO). If proper conditions are used, copolymers containing a range of styrene and butadiene molar fractions can be prepared. Structural analysis of these copolymers by means of 13C NMR spectroscopy allowed the assignment of different monomer diads (SS, SB, BB; S = styrene, B = butadiene) and the calculation of reactivity ratio products r1·r2. Differential scanning calorimetry (DSC) analysis further confirmed the block‐like structure of these copolymers. The melting points (Tm) of syndiotactic styrene sequences decrease as the styrene molar fraction decreases, whereas the glass transition temperature (Tg) increases with decreasing butadiene molar fraction in the copolymer. The polydispersity values (Mw/Mn) determined by GPC suggest that these copolymers are produced by a single site catalyst.  相似文献   

11.
Propylene polymerization was carried out using the [ArN(CH2)3NAr]TiCl2 (Ar = 2,6‐iPr2C6H3)/Al(iBu)3/Ph3CB(C6F5)4 catalyst system in the presence of cyclohexene. It was found that isospecific polymerization is promoted by adding cyclohexene even at low propylene concentration. It was also indicated that a considerable number of isospecific active species retain the metal‐polymer bond. Based on this fact, isotactic poly(propylene)‐block‐poly(1‐hexene) could be obtained.  相似文献   

12.
Polymerization of 3-(o-formylphenyl)propionaldehyde ( 4 ) and o-phenylenediacetaldehyde ( 5 ) were performed with several ionic catalysts. The cationic polymerization of 4 yielded polyacetals which contained a seven-membered cyclic unit ( 7 ) in mole fractions of 0,3–0,5. The uncyclized unit ( 6 ) contained the pendant aromatic aldehyde. Apparently, the aromatic aldehyde group reacted during the course of the intramolecular cyclization. Polymerizations of 5 with BF3·O(C2H5)2, lithium tert-butoxide, or Al(C2H5)3-TiCl4 as catalysts, similarly gave polyacetals containing a seven-membered cyclic unit ( 10 ) in mole fractions of 0,5–0,8. The extent of cyclization decreased with increase in temperature in the cationic polymerization, resulting in a difference of ?7,5 kJ mol?1 (?1,8 kcal mol?1) between the activation energy of intramolecular cyclization and that of intermolecular propagation. The polymerization with Al(C2H5)3 as catalyst gave a polyester via propagation involving a hydride shift.  相似文献   

13.
The effect of transition metal chlorides (MtClx) as isomerization catalysts was examined in the monomer-isomerization polymerization of cis-2-butene with TiCl3/Al(C2H5)3 as a catalyst. The isomerization and polymerization depend on both, MtClx and MtClx/TiCl3 mole ratio. The rate of polymerization of 2-butene with TiCl3/Al(C2H5)3/MtClx (mole ratio Al/Ti = 3,0, Mt/Ti = 1,0) as catalysts decreases in the following order: NiCl2 > CoCl2 > FeCl3 > None > MnCl2 > CrCl3. This order was found to be in a good agreement with the electronegativity of the metal atom in the chloride.  相似文献   

14.
By X-ray powder diffraction methods the crystal structures of α-TiCl3 and of the products obtained by dry ball-milling activation of this form were studied. TiCl3 microcrystals show initially a structural disorder in the positions of titanium atoms and the successive mechanical activation introduces into the structure some stacking faults, mainly associated to ±60° rotations of the triple layers Cl? Ti? Cl. The structural investigation has been carried out by an accurate fit of observed X-ray patterns to those calculated in correspondence of structures containing well-defined disordered sequences and of likewise well-defined crystallite sizes. Like γ-TiCl3 and MgCl2, α-TiCl3 is very sensitive to mechanical activation, which improves the performances of these three compounds when they are employed as catalysts, or supports for catalysts, in Ziegler-Natta polymerization processes.  相似文献   

15.
The activation of the tris(allyl)neodymium complex Nd(η3-C3H5)3 · dioxane with alkylaluminoxanes (MAO or HIBAO) results in highly selective catalysts for the 1,4-cis-polymerization of butadiene (cis-selectivity up to 80%). Under standard conditions (50°C, toluene), the turnover frequency (TOF) of the catalyst/MAO system amounts to 10–15000 mol butadiene/(mol Nd · h). Molecular weight determinations indicate the formation of only one polymer chain per neodymium center as in a living polymerization reaction, and for the catalyst/HIBAO system the rate law rp = kp [Nd][C4H6] with kp = 8,7 · 10?2 mol/(L · s) (at 25°C) has been derived. As the catalytically active species, a cationic monobutenyl neodymium(III) complex is discussed, which is stabilized through coordinative interaction with the counter anion as well as the growing polybutadiene chain. This cationic complex reacts under insertion with butadiene in a bimolecular fashion.  相似文献   

16.
The influence of ethylaluminium compounds (AIEt3 and AICIEt2) modified by 2,6-di-tert-butyl-4-methylphenol (BHT), tributylamine (TBA) and triphenylphosphine (TPP) on the polymerization of propene was investigated using catalysts based on TiCI3 modified by internal Lewis bases. Two catalysts were employed in this study, one (Cat. A) was prepared by the reduction of TiCI4 complexed with dibutyl ether (DBE)—mole ratio DBE/TiCI4 = 0,67—with AlClEt2 and the other one (Cat. B) was prepared in the same way, but a second internal base (ethyl benzoate (EB)) was added in a mole ratio EB/TiCl3 = 0,04. Activity and stereospecificity of the catalyst systems were strongly affected by these modified cocatalysts. The role of the modifiers is discussed.  相似文献   

17.
[(η5‐C5Me4)SiMe2(NtertBu)]TiCl2 was used as catalyst in the presence of methylaluminoxane (MAO) for the copolymerisation of ethylene with 5,7‐dimethylocta‐1,6‐diene (5,7‐DMO) and with 3,7‐dimethylocta‐1,6‐diene (3,7‐DMO), two linear non‐conjugated dienes readily available from terpene feedstock. The effects of reaction temperature, Al/Ti mole ratio and diene concentration in feed on the polymerisation activity and on the incorporation rates of the diene were investigated. The structure of the polymer and the distribution of the comonomer along the chains were investigated through NMR and DSC.  相似文献   

18.
Propene was polymerized with conventional and Solvay-type TiCl3 combined with Al(C2H5)2Cl using Zn(C2H5)2 as a chain transfer reagent. The produced terminal Zn-carbon bonds of the polymer were reacted with carbon dioxide and chlorine in the presence of N-methylimidazole followed by hydrolysis to obtain terminally carboxylated and chlorinated poly(propylene) in approximately 50% yield.  相似文献   

19.
1H and 2H NMR spectroscopic monitoring of ferrous species formed via interaction of 2,6‐bis[1‐(2,6‐dimethylphenylimino)ethyl]pyridineiron(II) chloride ( 1 ) with AlMe3, MAO, AlMe3/B(C6F5)3 and AlMe3/CPh3 (C6F5)4 is reported. At interaction of 1 with MAO in toluene solution, the new stable heterodinuclear neutral complexes with proposed structures LFe(II)(Cl)(μ‐Me)2AlMe2 and LFe(II)(Me)(μ‐Me)2AlMe2 are formed (L is initial tridentate ligand). Complex LFe(II)(Cl)(μ‐Me)2AlMe2 predominates at low Al/Fe ratios (less than 50), while LFe(II)(Me)(μ‐Me)2AlMe2 at high Al/Fe ratios (more than 500). Complex assigned to LFe(II)(Me)(μ‐Me)2AlMe2 can be prepared via interaction of 1 with AlMe3. Activation of LFe(II)(Me)(μ‐Me)2AlMe2 by B(C6F5)3 and CPh3B(C6F5)4 gives rise to formation of new complexes with proposed structures [LFe(μ‐Me)2AlMe2]+[MeB(C6F5)3] and [LFe(μ‐Me)2AlMe2]+ [B(C6F5)4]. Unexpectedly, the activity at ethylene polymerization was even higher for 1 /AlMe3 than for 1 /MAO catalytic system. The co‐catalytic activity of MAO towards 1 dramatically decreased with the diminishing of AlMe3 content in the composition of MAO. Activity of the catalyst 1 /AlMe3 and the molecular structure of polyethylene produced do not change noticeably at the addition of B(C6F5)3 to 1 /AlMe3.These data allow to suggest, that active species of 1 /AlMe3 and 1 /MAO systems are neutral methylated ferrous complexes but not cationic intermediates. Probably, complex LFe(II)(Me)(μ‐Me)2 AlMe2 is the closest precursor of these active species.  相似文献   

20.
2-Chloroethyl propenyl ether (CEPE) was synthesized, and its cationic polymerizability by BF3·O(C2H5)2 and the steric structure of the resulting polymer were studied. In the polymerization in toluene at ?78°C, the rate of consumption of monomers decreased in the following order: cis-CEPE > 2-chloroethyl vinyl ether > trans-CEPE. The steric structure of the β-methyl group of the resulting polymer was determined quantitatively on the basis of the NMR spectrum of the β-methyl protons decoupled from β-methine protons. In the polymerization of CEPE in toluene at ?78°C, the polymer obtained from trans-CEPE was rich in threo-meso structure whilst cis-CEPE gave a polymer containing almost the same amount of threo-meso and erythro-meso structures. With methylene chloride as solvent, the amount of threo-meso structure increased and that of erythro-meso structure decreased for the polymer obtained from cis-CEPE. The steric structure of the polymer obtained from trans-CEPE was independent of the nature of the solvent used. The effect of polymerization temperature on the steric structure of the polymers was also investigated.  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