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.
The polymerization of styrene with two pyrazolylimine nickel (II) complexes of (2-(C3HN2Me2-3, 5)(C(Ph) = N(4-R2C6H2(R1)2-2, 6)NiBr2 (Complex 1 , R1 = iPr, R2 = H; Complex 2 , R1 = H, R2 = NO2)) activated by methylaluminoxane was studied. The influences of polymerization parameters such as polymerization temperature, Al/Ni molar ratio, and reaction time on catalytic activity and molecular weight of the polystyrene (PS) were investigated in detail. The electron-withdrawing of nitro group in Complex 2 could not enhance the catalytic activity for styrene polymerization; however, the molecular weights of polymers were increased. Both of the two catalytic systems exhibited high activity [up to 8.45 × 105 gPS/(mol Ni h)] for styrene polymerization and provide PS with moderate to low-molecular weights (Mw = 2.21 × 104∼ 5.71 × 103 g/mol) and narrower molecular weight distributions about 2.0. The obtained PS were characterized by means of IR, 1H NMR, and 13C NMR techniques. The results indicated that the PS was atactic polymer. © 2011 Wiley Periodicals, Inc. J Appl Polym Sci, 2012  相似文献   

2.
Styrene (St) was polymerized in toluene solution by using bis(β‐ketoamino)nickel(II) complex as the catalyst precursor and methylaluminoxane (MAO) as the cocatalyst. The polymerization conditions, such as Al : Ni ratio, monomer concentration, reaction temperature, and polymerization time, were studied in detail. Both of the bis(β‐ketoamino)nickel(II)/MAO catalytic systems exhibited higher activity for polymerization of styrene, and polymerization gave moderate molecular weight of polystyrene with relatively narrow molecular weight distribution (Mw/Mn < 1.6). The obtained polymer was confirmed to be atactic polystyrene by analyzing the stereo‐triad distributions mm, mr, and rr of aromatic carbon C1 in NMR spectrum of the polymer. The mechanism of the polymerization was also discussed and a metal–carbon coordination mechanism was proposed. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci 2007  相似文献   

3.
The polymerization of norbornene has been investigated in the presence of two novel bispyrazolylimine dinickel(II) complexes bis-2-(C3HN2Me2-3,5)(C(Ph) = N(4-RC6H4)Ni2Br4 (complex 1, R = H; complex 2, R = OCH3) activated by methylaluminoxane. The two catalytic systems show high activity (up to 1.83 × 106 gPNBE/(molNi·h)) for norbornene polymerization and provide polynorbornene (PNBE) with higher molecular weights (M w = 4.44 × 105–11.57 × 105 g/mol) and narrower molecular weight distributions about 2.0. The electron-donating of methoxyl group in complex 2 could enhance the catalytic activity for norbornene polymerization, however, the molecular weights of polymers were decreased. The influences of polymerization parameters such as polymerization temperature, Al/Ni molar ratio, reaction time and catalyst concentration on catalytic activity, and molecular weight of the PNBEs were investigated in detail. The obtained PNBEs were characterized by means of 1H NMR, FTIR, and thermogravimetric analyses. The analyses results of PNBE indicated that the norbornene polymerization is vinyl-type polymerization rather than ring-opening metathesis polymerization.  相似文献   

4.
Polymerization catalysts based on copper precursors appear particularly interesting due to the low metal cost, limited toxicity and modest sensitivity to deactivation by polar species. To date, α‐olefin and polar monomer coordination polymerization catalysed using copper catalysts has been scarcely investigated, and a good part of the literature is represented by patents. Here this research has been expanded to the study of the performances of bis(salicylaldiminate)copper(II)/methylaluminoxane (MAO) catalysts in the polymerization of n‐butyl methacrylate. The study of the catalytic activity of bis(salicylaldiminate)copper(II)/MAO systems in n‐butyl methacrylate polymerization was focused on the relationship between the catalytic behaviour and the main reaction conditions and ligand structures. The electronic and steric characteristics of the chelate ligands play an important role in the catalytic performances. The presence of electron‐withdrawing nitro groups on the chelate ligands increased the catalytic activity which reached 36 kgpolymer mol?1 h?1, the highest value up to now reported for copper systems in methacrylic or acrylic monomer polymerization. These performances were ascribed to copper catalysts activated by MAO: without copper precursor, working in the presence of MAO and free salicylaldimine ligand, complete inactivity was ascertained. Copyright © 2010 Society of Chemical Industry  相似文献   

5.
Syndiotactic polystyrene (sPS) was synthesized with a novel monotitanocene complex of η5‐pentamethylcyclopentadienyltri‐4‐methoxyphenoxy titanium [Cp*Ti(OC6H4OCH3)3] activated by methylaluminoxane (MAO) in different polymerization media, including heptane, toluene, chlorobenzene, and neat styrene. In all cases bulk polymerization produced sPS with the highest activity and molecular weight. Solution polymerization produced much better activity in heptane than in the other solvents. Using a solvent with a higher dipole moment, such as chlorobenzene resulted in lower activity and syndiotacticity because of the stronger coordination of solvent with the Ti(III) active species, which controlled syndiospecific polymerization of styrene. With bulk polymerization at a higher polymerization temperature the Cp*Ti(OC6H4OCH3)3–MAO catalyst produced sPS with high catalytic activity and molecular weight. The external addition of triisobutylaluminum (TIBA) to the Cp*Ti(OC6H4OCH3)3–MAO system catalyzing styrene polymerization led to significant improvement of activity at a lower Al:Ti molar ratio, while the syndiotacticity and molecular weight of the yields were little affected. © 2000 John Wiley & Sons, Inc. J Appl Polym Sci 79: 1243–1248, 2001  相似文献   

6.
Two types of salicylaldiminato-based nickel complexes, mono-ligated Ni(II) complexes ([O-C6H4-o- C(H)=N-Ar]Ni(PPh3)(Ph) (5), [O-(3,5-Br2)C6H2-o-C(H)=N-Ar]Ni(PPh3)(Ph) (6), [O-(3-t-Bu)C6H3-o-C(H)=N-Ar]Ni(PPh3)(Ph) (7)) and bis-ligated Ni(II) complexes ([O-(3,5-Br2)C6H2-o-C(H)=N-Ar]2Ni (8), [O-(3,5-Br2)C6H2-o-C(H)=N-2-C6H4(PhO)]2Ni (9), Ar=2,6-C6H3(i-Pr)2) were synthesized and characterized by Fourier transform infrared spectroscopy (FT-IR), nuclear magnetic resonance (NMR), mass spectrography (MS) and elemental analysis (EA). In the presence of methylaluminoxane (MAO) as cocatalyst, all the nickel complexes exhibited high activities for the polymerization of methyl methacrylate (MMA) and syndiotactic-rich poly(methyl methacrylate) (PMMA) was obtained. The complexes with less bulky substituents on salicylaldiminato framework possessed higher activities, while with the same salicylaldiminato, the mono-ligated nickel complexes showed higher catalytic activity than bis-ligated ones.  相似文献   

7.
Homopolymerization of disubstituted styrenes (2,4‐ and 2,5‐dimethylstyrene) and trisubstituted styrene (2,4,6‐trimethylstyrene) and their copolymerization with styrene were carried out using diphenylzinc–metallocene–methylaluminoxane initiator systems for metallocene (n‐BuCp)2TiCl2 and for half‐metallocene CpTiCl3. The studied comonomers were found to be less reactive than p‐tertbutylstyrene, p‐methylstyrene and styrene. The results indicate that, even though the methyl group has I+ inductive effect, di‐ and tri‐methylstyrenes are reluctant to undergo either homopolymerization or copolymerization. This behavior suggests that the reactivity is regulated not only by the inductive effect of the alkyl group but also by the steric impediment caused by the crowding of the substituents on the benzene ring. Copyright © 2006 Society of Chemical Industry  相似文献   

8.
Summary The polymerization of butadiene(Bd), isoprene(Ip) and styrene(St) has been examined using the six catalyst systems composed of lanthanocene, (C5H9Cp)2NdCl(I), (C5H9Cp)2SmCl(II), (MeCp)2Sm OAr'(III), (Ind)2NdCl(IV), Me2Si(Ind)2NdCl(V) and (Flu)2NdCl(VI), and methylaluminoxane(MAO) respectively. All of them can be used to form the polyisoprene with molecular weights of 1 to 10 thousand and cis-1,4-unit contents of 41 to 47%. (I), (II) and (III) of them can be also used to form the polybutadiene with molecular weights of 10 to 20 thousand and cis-1,4-unit contents of 62 to 78%. In addition, the catalysts from (II) to (V) are still active for St polymerization and (II) of them gives a syndio -rich random polystyrene. It is noteworthy that (II) and (III) are active for homopolymerization of Bd, Ip and St in the same polymerization condition. Received: 16 December 1997/Revised version: 17 March 1998/Accepted: 24 March 1998  相似文献   

9.
The polymerization of butadiene was performed with catalysts based on the complex Nd{N(SiMe3)2}3 (1). This amide complex in combination with methyaluminoxane or with a boron compound ([HNMe2Ph][B(C6F5)4], [CPh3][B(C6F5)4] or B(C6F5)3) and Al(iBu)3 showed high activity and stereospecificity in polymerization of butadiene. The cationic complex [Nd{N(SiMe3)2}2(THF)2][B(C6F5)4] (2) was prepared by reaction of 1 and [HNMe2Ph][B(C6F5)4]. The catalyst 2/Al(iBu)3 (ratio Al/Nd: 10/1) was highly active for butadiene polymerization. Copolymerization of butadiene and styrene was performed with the catalytic system Nd{N(SiMe3)2}3/[HNMe2Ph][B(C6F5)4]/Al(iBu)3 (3). Copyright © 2004 Society of Chemical Industry  相似文献   

10.
[Bis(N,N′‐dimesitylimino)acenaphthene]dibromonickel ( 1 ) when activated with diethylaluminium chloride (DEAC) is a very active catalyst for ethylene homopolymerization. The activity (AE) of 1 /100 DEAC is twenty times greater than that of 1 /100 MAO and of the same order of magnitude as 1 /2000 MAO. In the case of homopolymerization of propylene the highest activity (AP) was obtained at a ratio of 25/15 for AlDEAC/Ni. Trialkylaluminium compounds were also found to act as cocatalysts for 1 . The PE synthesized with four different cocatalysts was found by 13C NMR to have dissimilar branching distributions. 1 /DEAC shows no activity for the polymerization of proximately substituted polar monomers. The introduction of dibutylmagnesium, (DBM) activates the 1 /DEAC system to copolymerize ethylene and a number of proximately substituted polar monomers. Compared with the 1 /MAO/monomer.AlR3 catalyst system the former is three times more active for copolymerization of 5‐hexene‐1‐ol or 10‐undecen‐1‐oic acid with ethylene. The activity of copolymerization is 1 /24, 1 /5 and 1 /2 as active as homopolymerization, respectively, in the case of methyl vinyl ketone, vinyl acetate and ?‐caprolactam. In the case of tetrahydrofuran/ethylene, the 1 /MAO catalyst produced copolymers using AlR3 pretreated THF whereas the 1 /DEAC/DBM catalyst produces homopolyethylene only. No polymerization occurred with an acrylonitrile/ethylene mixture in the presence of 1 /DBM/DEAC catalyst. © 2002 Society of Chemical Industry  相似文献   

11.
Polymerization of methyl methacrylate (MMA) with nickel(II) acetylacetonate [Ni(acac)2] in combination with methylaluminoxane (MAO) was investigated. Ni(acac)2 was found to be an effective catalyst for the polymerization of MMA. From a kinetic study of the polymerization of MMA with the Ni(acac)2–MAO catalyst, the overall activation energy was estimated to be 15 kJmol−1. The polymerization rate (Rp) was expressed as follows: Rp = k [MMA]1.0[Ni(acac)2–MAO]0.6 (the MAO/Ni mole ratio was kept constant). The mechanism for the polymerization of vinyl monomers with the Ni(acac)2–MAO catalyst is discussed. © 2000 Society of Chemical Industry  相似文献   

12.
Copolymerization of styrene (St) and butadiene (Bd) with nickel(II) compound (NiX2) in combination with methylaluminoxane (MAO) was investigated at the monomer feed ratio of 1:1. Copolymerization of St and Bd induced with NiX2–MAO catalysts (X = acac, OCOC6H5, OCOC18H35, Cl, Cp) gave copolymers with high cis‐1,4 contents of Bd units. The St and cis‐1,4 units of the Bd units in the copolymer were not significantly affected by the X group of NiX2, although the copolymer yields depended on the substituent. The copolymer yields depended on the solvent used for the copolymerization with the Ni(acac)2–MAO catalyst; an aromatic hydrocarbon was more favourable than a non‐aromatic hydrocarbon. The effects of triphenylphosphine (TPP) and trifluoroacetic acid (TFA) on copolymerization of St and Bd with the Ni(acac)2–MAO catalyst were seen the microstructure of Bd units in the copolymer. © 2001 Society of Chemical Industry  相似文献   

13.
The effect of different aluminum‐based cocatalysts (MAO, pMAO, and TMA) on butadiene (Bd) polymerization catalyzed by VO(P204)2 was investigated. The bimodal dependence of the polymer yield on the [MAO]/[V] molar ratio was revealed, and an highest polymer yield was achieved at a rather low [MAO]/[V] molar ratio ([MAO]/[V] = 13). The microstructures of the resulting poly(Bd)s were also significantly influenced by the ratio. In the TMA or pMAO system, the polymer yields were also very sensitive to the [Al]/[V] molar ratio. However, the microstructures of the resulting poly(Bd)s were almost independent of the ratio. In relation to the microstructures of poly(Bd)s obtained by the MAO and TMA systems at various temperatures, the 1,2‐unit contents were found to be the most abundant microstructure for both systems. In the pMAO system, the trans‐1,4‐units were the most abundant. The results of the additions of Lewis bases (THF and TPP) into Bd polyerization system comfirmed the existing of the two types of the reactions of VO(P204)2‐MAO catalyst and had the polymerization process controlled to some extent. The different thermal behaviors of these catalytic systems also show that multiple types of active centers were formed during the reaction between VO(P204)2 and MAO. © 2012 Wiley Periodicals, Inc. J. Appl. Polym. Sci., 2013  相似文献   

14.
The polymerizations of norbornene were investigated using a series of bis(β‐ketoamino)nickel(II) complexes( 1–6 ) in combination with methylaluminoxane (MAO) in toluene solution. The effects of catalyst structure, Al/Ni molar ratio, reaction temperature, and reaction time on catalytic activity and molecular weight of the polynorbornene were examined in detail. The electronic effect of the substituent around the imino group in the ligand is stronger than the steric bulk one on the polymerization activities, and the activities are in the order of 1 > 2 > 4 > 5 > 6 > 3 . The obtained polynorbornenes were characterized by means of 1H‐NMR, 13C‐NMR, FTIR, TG, and WAXD techniques. The analyses results of polymers' structures and properties indicate that the polymerization reaction of norbornene runs in vinyl‐addition polymerization mode. The obtained polynorbornene was confirmed to be vinyl‐type and atactic polymers and showed good thermostability (Tdec > 458°C) and were noncrystalline but had short‐range order. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 101: 4172–4180, 2006  相似文献   

15.
The catalytic activity in the polymerization of styrene has been examined using commercially available simple rare earth metal compounds such as Sm(OiPr)3, Sm(acac)3, Sm(OCOMe)3, SmI2(THF)2 or SmCl3 coupled with Et3Al or methylaluminoxane (MAO). Among these compounds, the Sm(OiPr)3/AlEt3 system shows the highest catalytic activity, especially in the presence of a minor amount of toluene at 60 °C. The random copolymerization of styrene with methyl methacrylate suggests that the present polymerization proceeds with a radical polymerization mechanism. (C5Me5)SmCl3Na(THF) and (C5Me5)SmCl3Li(THF) systems exhibit relatively low catalytic activity, even in the presence of AlEt3. © 2001 Society of Chemical Industry  相似文献   

16.
Poly(acrylamide-maleic acid) resin P(AAm-MA) was prepared by template polymerization. Polyacrylamide PAAm was used as a template for the polymerization of MA in an aqueous solution using gamma rays as the initiator. The effects on the capacity of P(AAm-MA), such as concentration of maleic acid and amount of template polymer, were investigated. P(AAm-MA) has been utilized as an adsorbent for the removal of Co(II) and Ni(II) ions from an aqueous solution. The effects of time of equilibrium, pH, temperatures, and dosage of the adsorbent on the removal of Co(II) and Ni(II) ions have been studied. The equilibrium data were analyzed using the Langmuir and Freundlich isotherm models. The equilibrium process was described well by the Langmuir isotherm model.  相似文献   

17.
Polymerization reactions of styrene in toluene at 60°C have been carried out using Ziegler–Natta catalysts containing praseodymium trichloroacetate‐triethyl aluminum [Pr(OCOCCl3)3‐Et3Al] (system I) and praseodymium trichloroacetate‐diethyl aluminum bromide [Pr(OCOCCl3)3‐Et2AlBr] (system II). The reaction was of first order with respect to both monomer and catalyst concentrations. At a stoichiometric ratio of Al/Pr = 8, an optimum catalytic activity was observed. The activation energies for systems I and II ranged between 21.06 kJ/mol and 13.79 kJ/mol, respectively. System I was white in color and system II is waxy. Molar masses of systems I and II were 9.1 × 105 and 2.2 × 104, respectively. The polymers were characterized by viscometry and 1H NMR methods. © 2001 John Wiley & Sons, Inc. J Appl Polym Sci 80: 995–1002, 2001  相似文献   

18.
The hexa-coordinated chelate complexes of the type [Ru(CO)2Cl2(P-P)](1a,b) [where P-P = 9,9-dimethyl-4,5-bis(diphenylphosphino)xanthene(a) and [bis(2-diphenylphosphinophenyl)ether(b)] have been synthesized by reacting the polymeric precursor [Ru(CO)2Cl2]n with the ligands in 1:1 molar ratio. The complexes 1a,b are characterized by elemental analyses, Mass, IR and NMR spectroscopy together with the single crystal X-ray structure determination of 1a. The compound 1a crystallizes in a monoclinic system with space group C2/c showing a slightly distorted octahedral geometry around the Ru centre. The complexes 1a and 1b are thermally stable up to 300 °C and exhibit high catalytic activity in transfer hydrogenation of aldehyde and ketones to corresponding alcohols. The complexes 1a and 1b show much higher catalytic activity for the hydrogenation of aldehyde than ketones. In general, the catalytic efficiency of 1b is higher compared with 1a.  相似文献   

19.
用阴离子聚合的方法合成了星型结构的高苯乙烯橡胶,对高苯乙烯橡胶的聚合反应动力学过程进行了研究,求得不同聚合温度、不同引发剂浓度下假一级表观增长反应速度常数kp"。并根据不同温度下的kp",求取了高苯乙烯橡胶聚合表观增长活化能Ep’=56.7 kJ/mol,高苯乙烯橡胶聚合反应动力学方程表明,聚合反应速率与单体浓度一次方、引发剂浓度1/2次方成正比关系。  相似文献   

20.
Two new half‐sandwich titanocene catalysts for the stereospecific polymerization of styrene, RCpTi(o‐MeOPhCPh2O)Cl2 (R = H (1); R = Me (2)), were prepared by reaction of the corresponding RCpTiCl3 complexes with o‐methoxyphenyldiphenylmethanol (o‐MeOPhCPh2OH) in the presence of triethylamine. Upon activation with excess methylaluminoxane (MAO), they showed high activities and high thermal stabilities for the stereospecific polymerization of styrene. The influences of polymerization temperature, Al/Ti molar ratio, solvent (aliphatic or aromatic) and time on the activity and syndiotacticity of the styrene polymerization were investigated. Copyright © 2006 Society of Chemical Industry  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