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 46 毫秒
1.
6-dimethylamino-6-methylfulvene (7) was converted to the [(C5H4)–CMe2–NMe2] ligand system (8) by treatment with methyllithium. Its reaction with MCl4 (M = Zr, Ti) followed by treatment with CH3Li gave the respective [(C5H4)–CMe2–NMe2]2M(CH3)2 complexes (12). Their reaction with B(C6F5)3 led to reactive metallocene cation complexes that instantaneously underwent CH activation at a N–CH3 group to yield the metallacyclic cation complexes 15. (tert-butylaminomethyl)fluorene was prepared by the addition of tert-butylisocyanate to fluorenyllithium followed by hydride reduction. Deprotonation by a variety of bases gave rise to a series of competing and consecutive reactions to yield several unusually structured products, among them a fluorenyl-anellated η5-1-azapentadienyl anion equivalent (25) and [(flu)-CH2–NCMe3]Li2 (23). An improved way of generating synthetically useful C1-linked [Cp–C1(R) n –NR1]2- dianion equivalents was developed starting from 6-amino-6-methylfulvene (26). N-silylation followed by double deprotonation with, e.g., lithium diisopropylamide cleanly furnished the respective [(C5H4)–C(=CH2)–NSiMe3]2- dianion 33 (isolated as the dilithio derivative). Its reaction with Cl2Zr(NEt2)2 in THF gave [η5:κ-N-(C5H4)–C(=CH2)–NSiMe3]Zr(NEt2)2 36. Activation of 36 with methylalumoxane in toluene led to the formation of a C1-linked “constrained geometry” Ziegler catalyst that polymerized ethylene similarly as the [(C5Me4)SiMe2NCMe3]ZrCl2 derived literature system. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

2.
tert‐Butyl vinyl ether (tBVE) was polymerized with the catalyst dimethyl[rac‐ethylenebis(indenyl)] zirconium (ansa‐zirconocene) with tri(pentafluorophenyl) borane [B(C6F5)3] as a cocatalyst. The effects of various polymerization conditions, such as the polymerization time, type of polymerization solvent, polymerization temperature, and catalyst concentration, on the conversion of tBVE into poly(tBVE), its molecular weight and molecular weight distribution, and its stereoregularity were investigated. The maximum conversion of tBVE into poly(tBVE) was over 90% at a polymerization temperature of ?30°C with an ansa‐zirconocene and B(C6F5)3 concentration of 3.0 × 10?7 mol/mol of tBVE, respectively. The number‐average molecular weights of poly(tBVE) ranged from approximately 14,000 to 20,000, with a lower polydispersity index (weight‐average molecular weight/number‐average molecular weight) ranging from 1.48 to 1.77, at all polymerization temperatures. The number‐average molecular weight of poly(tBVE) increased with decreases in the polymerization temperature and catalyst concentration. The mm triad sequence fraction of poly(tBVE) polymerized with ansa‐zirconocene/B(C6F5)3 at ?30°C was much higher than that of poly(tBVE) polymerized with the B(C6F5)3 catalyst at ?30°C, and this indicated that the ansa‐zirconocene/B(C6F5)3 catalyst system affected the isospecific polymerization of tBVE. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

3.
The effects of B(C6F5)3 on the activation of the Ni(0) and Ni(II) complexes were studied in the polymerization of norbornene. The Ni(0) complex, such as bis(1,5-cyclooctadiene)nickel (Ni(COD)2 (1), biacetylbis(2,6-diisopropylphenylimmine)(1,3-butadiene)nickel (2), or tetrakis(triphenylphosphine)nickel (5), in combination with B(C6F5)3, was determined to have high activity in the polymerization of norbornene. On the other hand, the Ni(II) complex with B(C6F5)3 did not provide any activity at all under analogous conditions regardless of the structure of the Ni(II) complex. The use of other borane compounds, such as B(C6H5)3, BEt3, and BF3 etherate, with Ni(COD)2 (1) in place of B(C6F5)3 clearly showed the main functions of B(C6F5)3. The high Lewis acidity of B(C6F5)3 enabled it to activate catalytic complexes, thus inducing polymerization. The study of the 1H, 13C, and 19F NMR spectra of the polynorbornene produced with Ni(COD)2 (1) and B(C6F5)3, in the presence or absence of ethylene, showed that the initiation of addition polymerization occurred through the insertion of the exo face of the norbornene into the Ni-C bond of the C6F5 ligand. A new polymerization mechanism was proposed in norbornene polymerization, wherein the active complex formed from Ni(COD)2 (1) and B(C6F5)3 acts as a catalyst.  相似文献   

4.
A detailed kinetic analysis of ethylene homopolymerization reactions and its copolymerization reactions with 1-hexene with a supported Ti-based Ziegler–Natta catalyst (reactions in the absence and the presence of hydrogen) shows a number of distinct kinetic features which are interpreted as a manifestation of multi-site catalysis; the catalyst contains several types of polymerization centers which differ in stability and formation rates, the molecular weight of polymers they produce, and in their response to the presence of α-olefins and hydrogen. All these effects require introduction of a special kinetic mechanism which postulates an unusually low activity of growing polymer chains containing one ethylene unit, the Ti–C2H5 group, in the ethylene insertion reaction into the Ti–C bond. This peculiarity of the Ti–C2H5 group, which is probably caused by its β-agostic stabilization, predicts two kinetic/chemical features of ethylene polymerization reactions which have not been described yet, the deuterium effect on the homopolymer structure and the activation effect of α-olefins on chain initiation. Both effects were confirmed experimentally. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

5.
Piers  Warren E.  Sun  Yimin  Lee  Lawrence W.M. 《Topics in Catalysis》1999,7(1-4):133-143
The highly electrophilic boranes HB(C6F5)2 and B(C6F5)3 are effective reagents for generating a variety of zwitterionic olefin polymerization catalysts. The former borane can be used to incorporate Lewis acid activators into ancillary ligand structures via hydroboration of pendant olefinic functions; alternatively, direct reaction with simple organozirconocenes can lead to a family of hydridoborate stabilized girdle-type zwitterions where charge separation is minimal. The ethylene polymerization activity of these compounds is, in general, poor by virtue of the tight intramolecular ion pairing. More active zwitterionic catalysts can be generated through reaction of B(C6F5)3 with suitable organozirconium pre-catalysts. The solution and solid state structures of several of these compounds are discussed, highlighting the various mechanisms of stabilization found in the catalyst structures. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

6.
Treatment of a series of 2-alkylfurans ( 1 ) and related systems with n-butyllithium, followed by addition of the resulting lithiofurans to 2-indanone, H3O+-catalyzed elimination, deprotonation, and transmetallation to zirconium tetra-chloride, gave a series of differently substituted bis [(5-alkyl-2-furyl) indenyl] ZrCl2 complexes [alkyl = n-butyl ( 5d ), n-hexyl ( 5e ), n-octyl ( 5f ), tert-butyl ( 5g )], as well as the corresponding benzofuryl ( 5h ) and 4,5-dimethylfuryl ( 5i ) complexes. Their treatment with methylalumoxane (>1000-fold excess) generated active homogeneous Ziegler-Natta polymerization catalysts. Propene polymerization at temperatures between −20 °C and +20 °C gave elastomeric polypropylenes. The elastomeric features were characterized by stress/strain curves that revealed a pronounced dependence on the actual alkyl substituent attached at the furyl rings at the periphery of the bent metallocene complexes. Increasing steric bulk of these substituents at the catalyst framework seems to result in increased isotacticities and a corresponding improvement in elastomeric properties of the polymers obtained by these systems.  相似文献   

7.
Four titanium complexes having β-ketoiminate chelate ligands with fluorine or alkyl groups [(Ar)NC(CH3)C(H)C(CH3)O]2TiCl2 (3a: Ar = 2.6-F2C6H3; 3b: Ar = C6F5; 3c: Ar = 2.6-Me2C6H3; 3d: Ar = 2.6-iPr2C6H3) have been synthesized and characterized by 1H NMR and EA. Complexes 3a, 3c and 3d were further characterized by X-ray diffraction analysis and demonstrated distorted octahedral coordination structure around the titanium center. The substituents in ligands greatly affect the coordination mode, resulting in three different isomeric structures. These complexes are active catalysts for polymerization of ethylene with MMAO as cocatalyst. The substituents in ligands have also great influences on catalytic activity. The complexes with alkyl groups have lower activity, while the complexes with fluorine atoms have middle or high catalytic activity.  相似文献   

8.
Summary: The kinetics of ethene polymerization with metallocene/[Me2PhNH]+[B(C6F5)4]?/AliBu3 (ternary systems) and metallocene/methylaluminoxan (MAO) systems respectively has been investigated at 100–140 °C and 7 MPa. Overall, eight different unbridged and ansa‐metallocenes were tested. The effect of ligand structure, cocatalyst and catalyst concentration on the thermostability and the activation energy of deactivation were studied. Deactivation with respect to the catalyst concentration followed a first order reaction. The half‐life as well as the activation energy of the ternary systems depended strongly on ligand structure while the ligand structure of MAO‐activated metallocenes merely influenced the half‐life. The half‐life associated with MAO activation is approximately twice as high as that in ternary activation. Based on these results, conclusions about the deactivation reactions have been drawn.

Influence of catalyst concentration on deactivation.  相似文献   


9.
Synthesis of new ferrocenylenesilylene polymers was effected by direct hydrosilylation/polymerization of 1,1′-methylsilyl-ferrocenophane [FC–SiMeH] (FC = (η5-C5H4)Fe(η5-C5H4) with acetylenes and organometallic olefins using a Pt0 catalyst. The reaction of [FC–SiMeH] with HC2Ph, HC2SiMe3, CH2=CHSiMe2Fp, Fp = (η5-C5H5)Fe(CO)2 and CH2=CHCH2SiMe2Fp in the presence of a platinum catalyst resulted in high yields of the corresponding hydrosilylated polymers. In the case of the acetylenes only β products were obtained, both E and Z isomers, whereas for the olefins both α and β-isomers were noted. Only in the case of the more bulky allylsilicon material was any unreacted SiH functionality retained in the polymer, an effect also noted when using [FCSiPhH] as the starting ferrocenophane. Cyclic voltammetric studies on these polymers revealed metal–metal interaction with two redox processes associated with the ferrocenylene Fe center along with an irreversible oxidation at higher potential for the Fp Fe atom. Ian: A pleasure to contribute; keep up the good work-aptp, cheers, Keith.  相似文献   

10.
Propylene polymerization was carried out using an ansa‐zirconocene pyrrolidide based catalytic system of racemic ethylene‐1,2‐bis(1‐indenyl)zirconium dipyrrolidide [rac‐(EBI)Zr(NC4H8)2 or (rac‐1)] and methylaluminoxane (MAO) or a noncoordinating anion. In situ generation of cationic alkylzirconium species was also investigated by NMR‐scale reactions of rac‐1 and MAO, and rac‐1, AlMe3, and [Ph3C] [B(C6F5)4]. In the NMR‐scale reaction using CD2Cl2 as a solvent, a small amount of MAO ([Al]/[Zr] = 30) was enough to completely activate rac‐1 to give cationic methylzirconium cations that can polymerize propylene. The resulting isotactic polypropylene (iPP) isolated in this reaction showed a meso pentad value of 91.3%. In a similar NMR‐scale reaction rac‐1 was stoichiometrically methylated by AlMe3 to give rac‐(EBI)ZrMe2, and the introduction of [Ph3C] [B(C6F5)4] into the reaction mixture containing rac‐(EBI)ZrMe2 led to in situ generation of cationic [rac‐(EBI)Zr(μ‐Me)2AlMe2]+ species that can polymerize propylene to give iPP showing a meso pentad value of 94.7%. The catalyst system rac‐1/MAO exhibited an increase of activity as the [Al]/[Zr] ratio increased within an experimental range ([Al]/[Zr] = 930–6511). The meso pentad values of the resulting iPPs were in the range of 83.2–87.5%. The catalytic activity showed a maximum (R p = 6.66 × 106 g PP/mol Zr h atm) when [Zr] was 84.9 × 10−6 mol/L in the propylene polymerization according to the concentration of catalyst. MAO‐free polymerization of propylene was performed by a rac‐1/AlR3/noncoordinating anion catalytic system. The efficiency of AlR3 was decreased in the order of AlMe3 (R p = 13.0 × 106 g PP/mol Zr h atm) > Al(i‐Bu)3 (8.9 × 106) > AlPr3 (8.8 × 106) > Al(i‐Bu)2H (8.4 × 106) > AlEt3 (8.4 × 106). The performance of the noncoordinating anion as a cocatalyst was on the order of [HNMePh2][B(C6F5)4] (R p = 13.0 × 106 g PP/mol Zr h atm) > [HNMe2Ph][B(C6F5)4] (10.8 × 106) > [Ph3C][B(C6F5)4] (8.4 × 106) > [HNEt2Ph][B(C6F5)4] (7.8 × 106). The properties of iPP were characterized by 13C‐NMR, FTIR, DSC, GPC, and viscometry. © 1999 John Wiley & Sons, Inc. J Appl Polym Sci 71: 875–885, 1999  相似文献   

11.
Margl  Peter  Deng  Liqun  Ziegler  Tom 《Topics in Catalysis》1999,7(1-4):187-208
We present a generalized view of d0and d0f n metal complexes as olefin polymerization catalysts from computational studies of the {L}M–C2H5 (0,+,2+)-fragments (M = Sc(III), Y(III), La(III), Lu(III), Ti(IV), Zr(IV), Hf(IV), Ce(IV), Th(IV) and V(V); L = NH–(CH)2–NH2- {1}, N(BH2)–(CH)2–(BH2)N2- {2}, O–(CH)3–O- {3}, Cp2 2- {4}, NH–Si(H2)–C5H4 2- {5}, {(oxo)(O–(CH)3–O)}3- {6}, (NH2)2 2- {7}, (OH)2 2- {8}, (CH3)2 2- {9}, NH–(CH2)3–NH2- {10} and O–(CH2)3–O2- {11}). This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

12.
Summary The catalytic performance (activity and polymer properties) of metallocenes with different symmetries in combination with methylaluminoxane (MAO) in the polymerization of propylene has been investigated at different temperatures, under standardized reaction conditions. The zirconocene rac-ethylene (5-1-indenyl) zirconium (IV) dichloride, with C2 symmetry, produces isotatic polypropylene and isopropylidene(5-cyclopentadienyl (5-9-fluorenyl) zirconium (IV) dichloride, with C S symmetry, syndiotactic polypropylene. The degree of the tacticity of these polymers increases with decreasing polymerization temperature. Only atactic polypropylene was formed with the unbridged zirconocenes bis(5-cyclopentadienyl) zirconium (IV) dichloride and bis(5-indenyl zirconium (IV) dichloride at any temperature investigated (10–60°C).  相似文献   

13.
Two novel late transition metals complexes with bidentate O?N chelate ligand, Mt(benzocyclohexan‐ketonaphthylimino)2 {Mt(bchkni)2: bchkni ?C10H8(O)C[N(naphthyl)CH3]; Mt ? Ni, Pd}, were synthesized. In the presence of B(C6F5)3, both complexes exhibited high activity toward the homo‐polymerization of norbornene (NB) (as high as 2.7 × 105 gpolymer/molNi·h for Ni(bchkni)2/B(C6F5)3 and 2.3 × 105 gpolymer/molPd·h for Pd(bchkni)2/B(C6F5)3, respectively). Additionally, both catalytic systems showed high activity toward the copolymerization of NB with 1‐octene under various polymerization conditions and produced the addition‐type copolymer with relatively high molecular weights (0.1–1.4 × 105g/mol) as well as narrow molecular weight distribution. The 1‐octene content in the copolymers can be controlled up to 8.9–14.0% for Ni(bchkni)2/B(C6F5)3 and 8.8–14.6% for Pd(bchkni)2/B(C6F5)3 catalytic system by varying comonomer feed ratios from 10 to 70 mol %. The reactivity ratios of two monomers were determined to be r1‐octene = 0.052, rNB = 8.45 for Ni(bchkni)/B(C6F5)3 system, and r1‐octene = 0.025, rNB = 7.17 for Pd(bchkni)/B(C6F5)3 system by the Kelen‐TÜdÕs method. The achieved NB/1‐octene copolymers were confirmed to be noncrystalline and exhibited good thermal stability (Td > 400°C, Tg = 244.1–272.2°C) and showed good solubility in common organic solvents. © 2012 Wiley Periodicals, Inc. J. Appl. Polym. Sci., 2013  相似文献   

14.

Abstract  

Copolymerization of norbornene with n-butyl methacrylate (n-BMA) was carried out with catalytic systems of bis-(β-ketoamino)nickel(II) complexes Ni{RC(O)CHC[N(naphthyl)]CH3}2 (R = CH3, CF3) and B(C6F5)3 in toluene and exhibited high activity for both catalytic systems. Influence of the catalyst structure and comonomer feed content on the polymerization activity as well as on the incorporation rates were investigated. The catalysis was proposed to involve the insertion mechanism of norbornene and n-BMA catalyzed by bis-(β-ketoamino)nickel(II)/B(C6F5)3 catalytic systems, and the decreasing polymerization activity with an increasing content of n-BMA in the feedstock composition could be attributed to the competition of carbonyl group coordination onto the Ni(II) active center instead of the olefin double bond. The reactivity ratios were determined to be r n-BMA = 0.095 and r norbornene = 12.626 by the Kelen–Tüd?s method. The copolymer films prepared show good transparency in the visible region.  相似文献   

15.
We have analyzed the computational results for several elementary reactions of the ethylene polymerization process catalyzed by an alternative (to the existing metallocene catalysts) “non-cyclopentadienyl” catalysts such as diimine-M(II) (where M = Ni and Pd) and chelating bridged Ti- and Zr-complexes. The obtained data have been compared with those for the existing zirconocene-based catalysts. In general, it was shown that: (i) the resting stage of the process is a metal-olefin-alkyl complex, the olefin coordination energy of which is a few kcal/mol larger for diimine-M(II) systems than zirconocene or dialkoxide systems; (ii) the rate-determining barrier is a migratory insertion barrier calculated from the metal-olefin-alkyl complex, which is found to be a few kcal/mol larger for the diimine-M(II) system compared to the Cp2ZrCH 3 + catalyst. The presence of certain flexible bridging ligands X in the Ti-alkoxide complex, [Y(Ph)X(Ph)Y]TiCH 3 + , which are capable of donating electron density to the cationic metal center at various stages during the reaction makes this barrier a few kcal/mol smaller for the dialkoxide than the Cp2ZrCH 3 + catalyst. It was shown that an increase in the metal-bridge interaction decreases the migratory insertion barrier and, consequently, increases the catalytic activity of these complexes. Although the diimine-M(II) catalysts are less active than zirconocene-based ones, the microstructure of the polymers produced by the former catalyst, which is found to be a function of temperature, ethylene, steric bulkiness of the auxiliary ligands, and transition metal center, makes them attractive for practice. We also have studied the mechanisms of several chain termination/transfer reactions, as well as the role of steric effects in the studied elementary reactions. We have clearly demonstrated tremendous possibilities of the computational chemistry in solving complex problems of the homogenous catalyst, and its high capability of predicting new and more active catalysts for different commercially important processes including olefin polymerization reactions. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

16.
In the presence of methylaluminoxane (MAO), ethylene polymerization was successfully performed with homobinuclear zirconocene complexes {[(C5H5)ZrCl2](C5H4CH2 C6H4CH2C5H4)[(C5H5)ZrCl2]; 3o , 4m , and 5p }, which were prepared conveniently by the reaction of disodium(phenylenedimethylene)dicyclopentadienide [C6H4(CH2C5H4Na)2] with 2 equiv of (N5‐Cyclopentadienyl)trichlorozirconium dimethoxyethane (CpZrCl3(DME)) in tetrahydrofuran and characterized by 1H‐NMR and elemental analysis. The effects of the polymerization parameters, such as the temperature, time, concentration of the catalyst, MAO/catalyst molar ratio, and isomeric difference of the homobinuclear metallocene complexes 3o , 4m , and 5p were studied in detail. The results showed that all three catalytic systems had moderate activities in ethylene polymerization and afforded polyethylene with relatively broad polydispersities. The catalytic activity of 4m was somewhat higher than that of 3o and 5p but lower than that of 4,4′‐bis(methylene)biphenylene‐bridged zirconocene catalysts; this indicated that the distance between the two metal centers was too short in comparison with a 4,4′‐bis(methylene)biphenylene bridge to increase the catalytic activity. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci, 2006  相似文献   

17.
[Me2C(Cp) (Ind)]ZrCl2 metallocene catalyst has been prepared and employed in a study of ethene polymerization in the presence of the cocatalyst methylaluminoxane. C1 and C2 signals are detected in the 13C NMR spectra of the resultant polymers; this reveals that the resultant polymer is a branched polyethene (polyethylene). The influence of polymerization temperature, catalyst concentration and [Al]/[Zr] ratio on catalytic activities and polymerization kinetics is investigated. A plausible mechanism for forming branched polyethene is suggested. © 2000 Society of Chemical Industry  相似文献   

18.
Tianxu Sun  Zhiqiang Fan 《Polymer》2010,51(14):3091-2485
Heterobinuclear cobalt and nickel complex {2-[2,6-R2-C6H3NC(CH3)-(CH3)CN-(3,5-R2′)C6H2-CH2-(3′,5′-R2)C6H2NC(CH3)]-6-[2,6-R2-C6H3NC(CH3)] pyridine}CoCl2NiBr2 (R = isopropyl) (N5CoNi) was prepared by reaction of pentadentate nitrogen ligand containing 2,6-bis(imino)-pyridine and α-diimine moieties with CoCl2 and NiBr2(DME) in turn. The complex was applied as catalyst for ethylene polymerization activated by AlEt3, MMAO and AlEt3/[PhMe2NH] [B(C6F5)4] respectively. The performance of the heterobinuclear complex in ethylene polymerization was compared with corresponding mononuclear complexes (α-diimine nickel bromide and 2,6-bis(imino)-pyridine cobalt chloride) and their equivalent mixture (binary complexes). When the complex N5CoNi was activated by AlEt3 or MMAO, its ethylene polymerization activity was lower than its control, the binary complexes. Both heterobinuclear complex and binary complex produced PE with bimodal molecular weight distribution. The amount of high-molecular-weight polyethylene produced by nickel center of N5CoNi was less than the binary complexes, which reveals that productivity of nickel center of N5CoNi is selectively suppressed. When the heterobinuclear complex N5CoNi is activated by AlEt3/[PhMe2NH][B(C6F5)4], the relative productivity of nickel center increased, although the total activity of catalyst decreased compared with AlEt3 as cocatalyst. With respect to AlEt3, [PhMe2NH][B(C6F5)4] can preferably activate nickel center of heterobinuclear complex. The results suggest that metal site in the heterobinuclear complex is selectively activated by cocatalyst.  相似文献   

19.
The cycloaddition of carbon dioxide to epichlorohydrin was performed without any solvent in the presence of ionic liquid as catalyst. 1-Alkyl-3-methyl imidazolium salts of different alkyl group (C2, C4, C6, C8) and anions (Cl, BF4, Br, PF6) were used for this reaction carried out in a batch autoclave reactor. The conversion of epichlorohydrin was affected by the structure of the imidazolium salt ionic liquid; the one with the cation of longer alkyl chain length and with more nucleophilic anion showed better reactivity. The conversion of epichlorohydrin increased as the temperature increased from 60°C to 140°C. It also increased with increasing carbon dioxide pressure probably due to the increase of the absorption of carbon dioxide into the mixture of epichlorohydrin and the ionic liquid. Zinc bromide was also tested for its use as a cocatalyst in this reaction. This work was presented at the 6 th Korea-China Workshop on Clean Energy Technology held at Busan, Korea, July 4–7, 2006.  相似文献   

20.
The effects of aluminum alkyls on the gas‐phase ethylene homopolymerization and ethylene/1‐hexene copolymerization over polymer‐supported metallocene/methylaluminoxane [(n‐BuCp)2ZrCl2/MAO] catalysts were investigated. Results with triisobutyl aluminum (TIBA), triethyl aluminum (TEA), and tri‐n‐octyl aluminum (TNOA) showed that both the type and the amount of aluminum alkyl influenced the polymerization activity profiles and to a lesser extent the polymer molar masses. The response to aluminum alkyls depended on the morphology and the Al : Zr ratio of the catalyst. Addition of TIBA and TEA to supported catalysts with Al : Zr >200 reduced the initial activity but at times resulted in higher average activities due to broadening of the kinetic profiles, i.e., alkyls can be used to control the shape of the activity profiles. A catalyst with Al : Zr = 110 exhibited relatively low activity when the amount of TIBA added was <0.4 mmol, but the activity increased fivefold by increasing the TIBA amount to 0.6 mmol. The effectiveness of the aluminum alkyls in inhibiting the initial polymerization activity is in the following order: TEA > TIBA >> TNOA. A 2‐L semibatch reactor, typically run at 80°C and 1.4 MPa ethylene pressure for 1 to 5 h was used for the gas‐phase polymerization. © 2004 Wiley Periodicals, Inc. J Appl Polym Sci 92: 3549–3560, 2004  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