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.
Cycloolefin copolymer, poly[(ethylene)‐co‐(5‐vinyl‐2‐norbornene)], was synthetized using the MAO activated ansa‐metallocene Ph2C(Flu)(Cp)ZrCl2 as a catalyst. Formation of mostly isolated 5‐vinyl‐2‐norbornene sequences was confirmed by NMR analysis. Comonomer incorporated selectively via cyclic double bond. No selectivity towards the reactivity of the particular isomer of 5‐vinyl‐2‐norbornene (5‐endo : 5‐exo) was observed. The molar mass was found to decrease when the comonomer concentration in product increased. At low comonomer incorporation levels copolymers were insoluble, cross‐linked elastomers. With high comonomer concentration in reaction medium copolymers were amorphous determined by means of DSC and X‐ray diffraction and the vinyl group of the comonomer remained unreacted for further modification.  相似文献   

2.
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   

3.
Styrene was copolymerized with ethylene using Me2Si(Me4Cp)(N-tert-butyl)TiCl2/methylaluminoxane (Me = methyl, Cp = cyclopentadienyl) varying monomer concentration, styrene/ethylene molar ratio and polymerization temperature. Increasing styrene or decreasing ethylene concentration, respectively, in the monomer feed lowered both activity of the catalyst system and molecular weight of the copolymer. Only at low styrene content, a linear correlation of styrene concentration in the monomer feed and styrene incorporation in the copolymer was found. From copolymerizations at various temperatures the activation energy of the insertion step was calculated to be 56 kJ/mol. According to Fineman-Ross plot, the copolymerization parameters are rE = 23,4 for ethylene and rS = 0,015 for styrene. Investigation of the thermal properties by means of differential scanning calorimetry and dynamic mechanical analysis revealed pronounced decrease of melting temperature and increase of glass transition temperature with increasing styrene content.  相似文献   

4.
At –25°C, the polymerizations of hydroxystyrene, the phenolic —OH of which was protected with trialkylsilyl compounds, were investigated by using a syndiospecific living polymerization catalyst system composed of (trimethyl)pentamethylcyclopentadienyltitanium (Cp*TiMe3), trioctylaluminium (AlOct3) and tris(pentafluorophenyl)borane (B(C6F5)3). The use of bulky trialkylsilyl protective groups was effective to control a stereoregularity and a molecular weight distribution (MWD) of polymer. In the case of 4‐(tert‐butyldimethylsilyloxy)styrene (TBDMSS) monomer, the number‐average molecular weights (n's) of polymer produced increased proportionally with increasing of monomer conversion. The MWD of polymer stayed narrow (w /n = 1.05–1.15). It was concluded, thus, the polymerizations of TBDMSS with Cp*TiMe3/B(C6F5)3/AlOct3 catalytic system proceeded under living fashion. The 13C NMR analysis clarified that the polymers obtained in this work had highly syndiotactic structure. By the deprotection reaction of silyl group with conc. hydrochloric acid (HCl), syndiotactic poly(4‐hydroxystyrene) (PHOST) with narrow MWD was prepared. The obtained syndiotactic PHOST had a good solubility for polar solvents and a high glass transition temperature (Tg) of 194°C.  相似文献   

5.
At –25°C, the sequential block copolymerizations of 4‐(tert‐butyldimethylsilyloxy)styrene (TBDMSS) and 4‐methylstyrene (4MS) were investigated by using a syndiospecific living polymerization catalyst system composed of (trimethyl)pentamethylcyclopentadienyltitanium (Cp*TiMe3), trioctylaluminum (AlOct3) and tris(pentafluorophenyl)borane (B(C6F5)3). The number‐average molecular weight (n) of the poly(TBDMSS)s increased linearly with increasing the polymer yield up to almost 100 wt.‐% consumption of TBDMSS used as 1st monomer. The n value of the polymer after the second monomer (4MS) addition continued to increase proportionally to the polymer yield. The molecular weight distributions (MWDs) of the polymers remained constant at around 1.05–1.18 over the entire course of block copolymerization. It was concluded that the block copolymerizations of TBDMSS and 4MS with the Cp*TiMe3 /B(C6F5)3 /AlOct3 catalytic system proceeded with a high block efficiency. The 13C NMR analysis clarified that the block copolymers obtained in this work had highly syndiotactic structure. By the deprotection reaction of silyl group with conc. hydrochloric acid (HCl), syndiotactic poly{(4‐hydroxystyrene)‐block‐[(4‐methylstyrene)‐co‐(4‐hydroxystyrene)]} (poly[HOST‐b‐(4MS‐co‐HOST)]) was successfully prepared.  相似文献   

6.
Propylene polymerization was conducted with Ph2C(R-Cp)(Flu)ZrCl2 [R = Me, i-Pr, PhCH2, Me3Si] catalysts in combination with methylaluminoxane (MAO) and dimethylanilinium tetrakis(pentafluorophenyl)borate (Me2PhNH·B(C6F5)4) as cocatalyst; the dependence of the stereoregularity of poly(propylene) on cocatalysts and bulkiness of the substituents in β-position of the cyclopentadienyl ligand was studied. Methyl and i-propyl substituted metallocene catalysts produce hemi-isotactic poly(propylene). These results are in good agreement with the results of the isopropylidene bridged metallocene analogue. The benzyl substituted metallocene catalyst produces syndiotactic poly(propylene) regardless of the cocatalyst. This means that this substituent group does not affect migration insertion of propylene. Stereoregularity of poly(propylene) obtained with diphenylmethylidene(3-trimethylsilylcyclopentadienyl)(fluorenyl)zirconium dichloride (Ph2C(Me3SiCp)(Flu)ZrCl2) as a catalyst was significantly influenced by the cocatalyst. Me2PhNH· B(C6F5)4/triisobutylaluminium(i-Bu3Al) produces poly(propylene) with 65% racemic and 23% meso pentads at 40°C, whereas the MAO activated catalyst produces isotactic rich poly(propylene). Fractionation experiments indicated that Me2PhNH·B(C6F5)4/i-Bu3Al forms two active sites, one of them being the same as that of the MAO activated catalyst, the other one producing syndiotactic rich poly(propylene).  相似文献   

7.
Ethylene and butylene‐1 were homo‐ and co‐ polymerized using monotitanocene catalysts composed of η5‐pentamethylcyclopentadienyl tribenzyloxytitanium [Cp*Ti(OBz)3] and two kinds of solid modified methylaluminoxane (mMAO) containing different amounts of residual trimethylaluminium (TMA). The mMAO1 with lower TMA content to activate Cp*Ti(OBz)3 for homo‐ and co‐ polymerization of ethylene and butylene‐1 showed higher activity and more butylene‐1 incorporation in their copolymers than mMAO2 with higher TMA content. Influences of copolymerization conditions including ethylene/butylene‐1 ratio in the feed, polymerization temperature and Al/Ti molar ratio on activity, butylene‐1 content in copolymers, molecular weight and molecular weight distribution, the triad and tetrad sequence distributions and butylene‐1 content in the copolymers. The monomer reactivity ratios rE and rB (E = ethylene, B = butylene‐1) for ethylene/butylene‐1 copolymerizations were estimated from the 13C NMR spectra. The Cp*Ti(OBz)3/mMAO catalyst gave rise to rErB lower than 1. A correlation between the reactivity ratios and mMAO composition was also made. The thermal properties of copolymers determined by DSC are mainly dependent on the butylene‐1 content in the copolymers.  相似文献   

8.
Blends of syndiotactic polystyrene (sPS) and isotactic polypropene (iPP) have been prepared using TiCl4/MgCl2/β‐diketone activated with methylaluminoxane (MAO). The influences of the Al/Ti ratio and the polymerization temperature on the catalyst activity and the blend composition have been investigated. It is shown that the polymerization temperature is an important factor in determining catalyst activity and the blend composition. Through controlling the polymerization conditions, blends containing a wide composition range of styrene/propene molar fractions can be prepared. The supported catalyst shows a good activity of about 1.5×105 (g blend)·(mol Ti)–1·h–1. 13C NMR analysis allows the composition of the blends to be calculated. Dynamic mechanical analysis (DMA) reveals that the blends of sPS/iPP prepared this way are partially compatible; the two glass transition temperatures (Tg's) of the components shift towards each other. The values of Tg, as well as the peaks of sPS intensity, increase with an increasing molar fraction of styrene in the blend. When the sPS content is in the range of 59–80 wt.‐%, dual phase continuity occurs in the blends. The morphology found by microscopic observation is consistent with the DMA results.  相似文献   

9.
The bulk polymerization of styrene has been investigated at 105°C in the presence of exclusively dialkylmagnesium or combination of chlorolanthanocene and dialkylmagnesium. In the presence of butylethylmagnesium or n,s‐dibutylmagnesium, styrene polymerization proceeds via thermal self‐initiation, but is accompanied by a reversible transfer to dialkylmagnesiums to yield in turn oligostyrylmagnesium species; the latter are finally hydrolysed to oligostyrenes with Mn = 500–1 500 and Mw/Mn = 2.0–2.8. The analysis of the oligostyrenes by MALDI‐TOF mass spectrometry establishes the presence of ethyl and butyl headgroups, consistent with the transfer process. When the dialkylmagnesium is combined with a lanthanocene such as (C5Me5)2NdCl2Li(OEt2)2 ( 1 ), an increase in activity is obtained which is ascribed to additional styrene polymerization initiated by in situ generated alkyl(hydride)lanthanocene species. The influence of various reaction parameters on the performance of this system has been investigated. The oligostyrenes (Mn = 500–9 000) produced under optimum conditions have a relatively narrow molar mass distribution (Mw/Mn = 1.20–1.40) which can be explained in terms of an efficient transfer between the chain‐growing lanthanide and the oligostyrylmagnesium species. The MALDI‐TOF mass spectra of the oligostyrenes produced with various dialkylmagnesium‐lanthanocene combinations gives an insight into the initiation mechanism. Finally, the combination of butylethylmagnesium and Cp*2NdCl2Li(OEt2)2 has been used to achieve (styrene‐co‐ethylene) block copolymers.  相似文献   

10.
Two kinds of syndiotactic AB type block copolymers were prepared, which were (1) poly(4‐methylstyrene)‐block‐polystyrene {Poly(4MS‐b‐S), (A: poly(4MS), B: polystyrene (S))}, (2) poly(4‐methylstyrene)‐block‐poly(styrene‐co‐3‐methylstyrene) {poly[4MS‐b‐(S‐co‐3MS)] (A: poly(4MS), B: styrene/3‐methylstyrene (3MS) copolymer)}. For the syntheses of these diblock copolymers, the living polymerization catalytic system composed of (trimethyl)pentamethylcyclopentadienyltitanium (Cp*TiMe3) premixed with trioctylaluminium (AlOct3), and tris(pentafluorophenyl)borane (B(C6F5)3) was used at –25°C. Chlorination of the methyl groups of poly[4MS‐b‐(S‐co‐3MS)] was conducted by aqueous sodium hypochlorite (NaOCl) and phase‐transfer catalyst such as tetrabutylammonium hydrogensulfate (TBAHS). The novel tapered densely grafted diblock copolymer was synthesized with by coupling reaction of living poly(2‐vinyl pyridine)lithium (Poly(2VP)Li) with the partly chloromethylated poly[4MS‐b‐(S‐co‐3MS)].  相似文献   

11.
Ethylene‐co‐norbornene polymers [P(E‐co‐N)s] have been synthesized with the dicarbollide catalysts (η5‐C2B9H11)M(NEt2)2(NHEt2) [M = Ti ( 1 ), Zr ( 2 )] activated with methylaluminoxane (MAO) or alkylaluminium compounds, as AlMe3 (TMA), Al(iBu)3 (TIBA), AlH(iBu)2 (DIBALH), AlEt2Cl (DEAC). Polymerization results by 1 ‐TIBA, 2 ‐TIBA, 1 ‐MAO and 2 ‐MAO catalysts have been preliminary compared with those of cyclopentadienyl derivatives of the group 4 metals, (η5‐C5R5)TiX3 [R = H, X = Cl ( 3 ); R = Me, X = Cl ( 4 ), Me( 5 )], (η5‐C5H5)ZrCl3 ( 6 ), ethylenebis(1‐indenyl) zirconium dichloride ( 7 ) and “Cp‐free” compounds M(NEt2)4 [M = Ti ( 8 ), Zr ( 9 )] precursors of the dicarbollide compounds 1 – 2 , under the same conditions (Tp = 50°C, [M] = 1×10–3 M ; PE = 1 atm; [N] = 1.75 M ). The 1 ‐TIBA and 2 ‐TIBA catalysts exhibit productivity values greater than the corresponding MAO activated system and incorporate high concentration of the cyclic monomer in the copolymer products (N mol‐% = 38–45). Crystalline blocks of isotactic alternating NENE sequences were identified in the P(E‐co‐N)s copolymers produced by these catalysts by means of 13C NMR and DSC analysis. The dicarbollide derivatives 1 and 2 were also efficiently activated with the alkyl aluminium compounds TMA, DIBALH and DEAC at very low Al/M molar ratio (Al/M = 10); the titanium and zirconium cyclopentadienyl derivatives 3 , 4 , 6 , 7 are inactive under these conditions.  相似文献   

12.
Copolymerization of ethylene and styrene was carried out with CpTiCl3/MgCl2‐PMAO as a catalyst at various temperatures and comonomer concentrations. The present catalyst system produces a pseudorandom copolymer of ethylene and styrene beside syndiotactic poly(styrene) (sPS) and poly(ethylene) (PE). The copolymers were obtained at temperature ⪈60°C, indicating the active species promoting the copolymerization being formed at elevated temperatures. On the other hand, styrene incorporation in the copolymer increases progressively with the increase of styrene concentration.  相似文献   

13.
Radical copolymerization of 2,2‐dimethyl‐1,3‐dioxolan‐4‐one (DMDO) consisting of a hybrid structure of acrylate and vinyl ether moieties with styrene, methyl methacrylate, and vinyl acetate was examined. The radical copolymerization was carried out without a solvent or in chlorobenzene in the presence of 3 mol‐% of 2,2′‐azoisobutyronitrile at 60°C for 20 h to obtain the copolymers with number‐average molecular weights of 1 400‐700 000 in 27–86% yields. No ring‐opening occurred but vinyl polymerization of DMDO selectively proceeded in the copolymerization. The monomer reactivity ratios were evaluated as r1 = 6.42, r2 = 0.08 (M1, DMDO; M2, styrene) by the Fineman‐Ross method. The Alfrey‐Price Q‐ and e‐values of DMDO were calculated as 5.97 and –0.13, respectively. Ab initio molecular orbital calculations were carried out to compare the reactivity of DMDO with methyl α‐methoxyacrylate and vinyl ether.  相似文献   

14.
The utilization of the Ziegler catalyst Cp*TiMe3/B(C6F5)3 for the terpolymerization of ethylene, propylene and 5‐ethylidene‐2‐norbornene to give EPDM is investigated, and the major factors affecting yields, molecular weights, molecular weight distributions and compositional distributions of the EPDM polymers are assessed. High molecular weight polymers with narrow molecular weight distributions are obtained at ˜18°C.  相似文献   

15.
A series of five fluorinated dithioesters PhC(S)SRCH2CnF2n+1 (where R represents an activating spacer and n = 6 or 8) was obtained in fair to high yields (57–88%). These transfer agents were successfully used in reversible addition‐fragmentation transfer (RAFT) of styrene (S), methyl methacrylate (MMA), ethyl acrylate (EA) and 1,3‐butadiene. Well‐chosen fluorinated dithioesters were able to lead to a good control of the radical polymerization of these monomers (i.e., molar masses of the produced polymers increased linearly with the monomer conversion and the polydispersity indexes ranging between 1.1 and 1.6 remained low). The relationship between the structures of the dithioesters and the living behavior of the radical polymerization of these above monomers is discussed and it is shown that the nature of the R group influences the living behavior from different contributions to radical stabilization. Furthermore, the RAFT process also yielded PMMA‐b‐PS and PEA‐b‐PS block copolymers bearing a fluorinated moiety.  相似文献   

16.
The paper describes a kinetic study of 1‐hexene oligomerization reaction with the (n‐Bu‐Cp)2ZrCl2–MAO catalyst system at 70 °C. GC analysis of the monomer/oligomer mixtures at different reaction times in the course of a 5‐h reaction, provided information on the formation rates of individual oligomer molecules, from the 1‐hexene dimer, C12H24, to the tetradecamer, C84H168. Kinetic modeling of the oligomerization kinetics strongly supports the principal premise of polymerization kinetics, that the value of the propagation rate constant of the chain growth reaction does not depend on the number of monomer units in the growing polymer chain. The only exception from this rule is the insertion of a 1‐hexene molecule into the Zr+? C bond in the active center bearing a single monomer unit, (n‐Bu‐Cp)2Zr+? CH2? CH2? C4H9. The rate constant of this reaction in one order of magnitude is higher than the rate constant of 1‐hexene insertion into active centers which are β‐branched with respect to the Zr atom, (n‐Bu‐Cp)2Zr+? [CH2? CH(C4H9)]n? CH2? CH2? C4H9.

  相似文献   


17.
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   

18.
Well‐defined, narrow molecular weight distribution (Mw/Mn ≤ 1.1) poly[(styrene)‐block‐(propylene oxide)] block copolymers with relatively high molecular weight poly(propylene oxide) blocks [e. g. Mn (PPO) = 10 000–12 000 g/mol] have been prepared by anionic polymerization. The polystyrene block (Mn = 5 000; Mw/Mn = 1.1) was prepared by alkyllithium‐initiated polymerization of styrene followed by chain‐end functionalization with ethylene oxide and protonation with acidic methanol. The resulting ω‐hydroxyl‐functionalized polystyrene was converted to the corresponding alkali metal salts with alkali metals (Na/K alloy, Rb, Cs) and then used to initiate block polymerization of propylene oxide in tetrahydrofuran. The effects of crown ethers (18‐crown‐6 and dicyclohexano‐24‐crown‐8) and added dimethylsulfoxide were investigated. Chain transfer to the monomer resulted in significant amounts of poly(propylene oxide) formation (50%); however, the diblock molecular weight distributions were narrow. The highest molecular weight poly(propylene oxide) blocks (12 200 g/mol) were obtained in tetrahydrofuran with cesium as counterion without additives.  相似文献   

19.
Summary: Syndiotactic polymers of 1‐pentene, 1‐hexene and 1‐octene were obtained with the CS‐symmetric metallocene catalysts [Ph2C(Cp)(2,7‐tert‐Bu2Flu]ZrCl2 and [(4‐MePh)2C(Cp)(2,7‐tert‐Bu2Flu)]ZrCl2, which have already been proven to give high molar masses and excellent tacticities in propene polymerisation. The monomers were polymerised in bulk and in solution processes at temperatures between 0 and 60 °C using methylaluminoxane as a cocatalyst. The effects on catalyst activity as well as on polymer microstructure and molar mass were determined. Polymers with high syndiotacticities and molar masses (up to 550 000 g · mol?1 in the 1‐octene polymerisation) compared to other metallocenes were obtained.

  相似文献   


20.
Free radical copolymerization of water‐soluble N‐vinylamides such as N‐vinylacetamide (NVA) and N‐vinylformamide (NVF) with hydrophobic vinyl acetate (VAc) gave amphiphilic copolymers. The monomer reactivity ratios were determined as r1 = 5.8 and r2 = 0.68 (M1 = NVA, M2 = VAc) and r1 = 6.2 and r2 = 0.37 (M1 = NVF, M2 = VAc), respectively. The growing radical of the terminals of N‐vinylamides propagates more favorably for N‐vinylamide monomers than for VAc monomer, resulting in the possible formation of blocky copolymers. It is found that aqueous solutions of these amphiphilic copolymers exhibited a lower critical solution temperature (LCST), depending on their chemical composition, followed by coacervate formation above the LCST. Furthermore, thermosensitive hydrogels could be prepared by the free radical copolymerization of N‐vinylamide and VAc in the presence of the crosslinker butylenebis(N‐vinylacetamide) (Bis‐NVA). The swelling ratios of these hydrogels decreased with an immediate increase in temperature from 20 to 80 °C, and then reversibly increased with decreasing temperature. These hydrogels showed the same thermosensitive properties as linear copolymers of NVF and VAc.

Relationship between LCST and vinyl acetate content in poly(N‐vinylamide‐co‐VAc)s.  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