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 187 毫秒
1.
TiCl_4/MgCl_2催化丙烯/1-辛烯共聚合研究   总被引:3,自引:2,他引:3  
本文用TiCl_4/MgCl_2-Al(i-Bu)_3催化剂进行丙烯/1-辛烯共聚合,研究发现引入少量共聚单体1-辛烯时,能提高丙烯的聚合活性。30℃时,测得共聚合竟聚率为r_丙=5.63,r_辛=0.32。共聚物的结晶度和己烷不溶物含量随其1-辛烯含量的增加而迅速下降。X射线衍射及~(13)C-NMR测定结果表明,共聚物的己烷可溶部分为非结晶的无规共聚物,己烷不溶部分是具有镶嵌着半个1-辛烯单体单元的长嵌段聚丙烯链结构的结晶性共聚物。  相似文献   

2.
本文探索了乙烯/丙烯/极性单体三元共聚物的合成方法.乙烯/丙烯/ω-Cl-α-乙烯基单体三元共聚物由于分子中引入了ω-Cl-α-乙烯基极性单体,改变了乙烯丙烯共聚物的化学惰性.我们采用催化剂Cat.L-Pd配位催化乙烯/丙烯/ω-Cl-α-乙烯基单体三元共聚合,合成了极性三元无规共聚物.探讨了催化剂结构、聚合条件对三元共聚合行为的影响,并优化了聚合条件.采用红外光谱(FTIR)、核磁共振碳谱(氢谱)(~(13)C(~1H)NMR)、示差扫描量热(DSC)和高温凝胶渗透色谱(GPC)等方法研究了共聚物的结构与性能.FTIR与~(13)C(~1H)NMR结果表明,催化剂Cat.L-Pd能够有效催化乙烯/丙烯/ω-Cl-α-乙烯基单体三元共聚合,共聚物中ω-氯代极性单体的插入量达3.6 mol%.极性单体不发生均聚合反应,但能够有效参与乙烯和丙烯的共聚合反应,形成三元无规共聚物.丙烯能够发生均聚合反应,但是不能形成聚丙烯长链段,主要发生乙烯与丙烯共聚合反应.乙烯最易发生聚合反应,并能够形成较长链段的聚乙烯.共聚物的Mw高于2×10~5g/mol.分子量分布在1.6~3.0,说明该类催化剂催化乙烯/丙烯/ω-Cl-α-乙烯基单体三元共聚合行为遵循单中心聚合机理.  相似文献   

3.
用13C NMR测定了由单茂基钛化合物 /mMAO催化体系制备的乙烯 /丙烯共聚物大分子链的立体结构和单体序列分布 ,计算了单体的竞聚率r1=7 91± 0 0 6 ,r2 =0 135± 0 0 3 ,其乘积r1r2 ≈ 1.Fineman Ross计算得到的单体竞聚率与13C NMR测定值相近 ,即r1=7.94,r2 =0 .134,其乘积r1r2 =1 0 4.这表明共聚物是立构无规的 .共聚物经溶剂萃取后 ,乙醚可溶和己烷可溶两个级分中单体的序列分布和竞聚率略有不同 .乙醚可溶级分中丙烯链段稍长 ,而己烷可溶级分中乙烯链段稍长  相似文献   

4.
以球形高效负载的TiCl4/MgCl2/邻苯二甲酸二异丁酯(DIBP)为催化剂, 采用本体聚合方法进行丙烯与1-丁烯共聚合研究. 考察了共单体效应对共聚活性及聚合物立构规整性的影响; 表征了共聚物的结构. 结果表明, 随着1-丁烯/丙烯投料比的增加, 聚合活性呈先升高后降低的趋势, 在1-丁烯/丙烯摩尔投料比为0.26条件下聚合活性达到最高, 并随着共聚物中1-丁烯含量的增加, 共聚物的熔点明显下降, 分子量降低, 分子量分布变窄, 同时共聚物力学性能有明显提高, 透明度逐渐增加.  相似文献   

5.
使用Et(Ind)2ZrCl2/MAO催化剂催化乙烯和3种ω-对甲苯基-α-烯烃(对甲苯基-1-丙烯,4-对甲苯基-1-丁烯,6-对甲苯基-1-己烯)共聚,主要研究了共单体加入量对催化剂活性和所得共聚物性能的影响.4-对甲苯基-1-丁烯表现出最好的共聚性能.使用1H-NMR、13C-NMR、GPC和DSC对共聚物进行了表征.  相似文献   

6.
用13C-NMR测定了由单茂基钛化合物/mMAO催化体系制备的乙烯/丙烯共聚物大分子链的立体结构和单体序列分布,计算了单体的竞聚率r1=7.91±0.06,r2=0.135±0.03,其乘积r1r2≈1.Fineman-Ross计算得到的单体竞聚率与13C-NMR测定值相近,即r1=7.94,r2=0.134,其乘积r1r2=1.04.这表明共聚物是立构无规的.共聚物经溶剂萃取后,乙醚可溶和己烷可溶两个级分中单体的序列分布和竞聚率略有不同.乙醚可溶级分中丙烯链段稍长,而己烷可溶级分中乙烯链段稍长.  相似文献   

7.
应用扫描电子显微镜(SEM)研究乙烯-丙烯嵌段共聚物和聚乙烯/聚丙烯共混物的冲击断裂表面形态。结果表明,这二种材料力学性能的差别,明显地反映在它们的断面形态上。前者显示出一种典型的韧性断裂特征;后者则呈现出一种剥离层状结构。~(13)C-NMR分析结果表明,在上述乙烯-丙烯嵌段共聚物中,确实存在着乙烯-丙烯共聚链段。这是该材料具有良好的抗冲击性能的主要原因。  相似文献   

8.
利用具有"颗粒反应器技术(RGT)"特征的Ziegler-Natta催化剂进行丙烯多相共聚(丙烯均聚+乙烯/丙烯无规共聚),通过在乙丙共聚阶段引入双烯烃单体1,9-癸二烯,使乙丙共聚物在聚合的同时实现交联,制备了新型抗冲聚丙烯合金.聚合反应结果表明,1,9-癸二烯可参与乙丙共聚,同时对聚合反应速率和共聚物组成影响较小;1,9-癸二烯使乙丙共聚物发生支化/部分交联,合金聚合物的熔体流动速率在引入1,9-癸二烯后显著降低,且凝胶含量随1,9-癸二烯用量的增加而增大.形态研究结果表明,乙丙共聚物的交联显著降低了其在聚丙烯基体中的分散尺度,提高了分散均匀性,分散相粒径随支化/交联程度提高而减小.力学性能测试结果表明,乙丙共聚物的交联使合金聚合物在保持较高韧性的同时显著提升了刚性,有利于实现抗冲聚丙烯合金的刚韧平衡.  相似文献   

9.
采用合成的催化剂五甲基环戊二烯基三烯丙氧基钛 [Cp Ti(OAllyl) 3]与改性甲基铝氧烷 (mMAO)组成催化体系制备乙烯 /丙烯共聚物 .红外分析显示 ,乙醚可溶和己烷可溶两个级分的化学结构几乎相同 .GPC测试结果表明共聚物分子量高 ,分子量分布窄 .X 射线衍射分析 ,大多数样品的图谱为宽的弥散峰 ,表明它们是无规共聚物 ;只有当乙烯含量很高时 ,样品的谱图才有较为尖锐的结晶峰 ,结晶度不高 .经热分析(DSC、TG) ,大多数样品没有出现明显的熔点 ,只有当乙烯含量很高时才显示出熔点 ;共聚物的热稳定性较高 .DMA分析表明 ,共聚物样品中乙烯含量多的 ,其储能模量 (E′)大一些 ;共聚物的玻璃化转变温度随着丙烯链节的增多而升高 .  相似文献   

10.
改性甲基铝氧烷(mMAO)激活五甲基茂基三氯化钛(Cp TiCl3)催化乙烯 丙烯共聚合,控制两种单体的进料配比,得到单元序列分布不同的共聚物.混合单体中含有少量丙烯,共聚合活性高于相同聚合条件下乙烯均聚合的活性.用1 3C -NMR测定共聚物分子链的微观结构和单元序列分布,计算出单体的竞聚率;结果表明共聚物分子链中两种单体的序列分布均匀.混合单体中丙烯含量较大时,共聚物为完全无规共聚物;而当丙烯含量少时,丙烯链节或短的聚丙烯链段均匀分布于聚乙烯链段之间.共聚物经DSC分析,也证明不存在长序列的聚乙烯链段;因此,即便在进料气体中丙烯含量很少的情况下,共聚物仍然没有明显的熔融温度和结晶性.  相似文献   

11.
锆茂均相催化剂对乙烯/丙烯和乙烯/1-丁烯共聚合研究姚晖,肖士镜,陆宏兰(中国科学院化学研究所北京100080)关键词共聚合,乙烯,丙烯,丁烯,锆茂,甲基铝氧烷金属茂均相催化剂中过渡金属的性质,及其周围配位体的结构对催化剂性能有很大影响[1,2].金...  相似文献   

12.
锆茂均相催化剂对乙烯/丙烯和乙烯/1-丁烯共聚合研究姚晖,肖士镜,陆宏兰(中国科学院化学研究所北京100080)关键词共聚合,乙烯,丙烯,丁烯,锆茂,甲基铝氧烷金属茂均相催化剂中过渡金属的性质,及其周围配位体的结构对催化剂性能有很大影响[1,2].金...  相似文献   

13.
Nascent form of random copolymers of propylene with ethylene, 1-butene, 1-hexene, 1-octene, and 4-methyl-1-pentene was studied by Raman spectroscopy. The most significant spectral alterations with a change in propylene content were observed in two lines at 809 and 841 cm−1. The first line corresponds to vibrations of polypropylene helical chains in the crystalline phase, while the second one is associated with vibrations of polypropylene helical chains having isomeric defects. Raman data confirm that conformational composition and phase state of copolymer macromolecules strongly depend on the comonomer content as well as on the size of the comonomer units.  相似文献   

14.
Copolymerization of propylene and 1-butene with highly stereospecific three-component coordination catalysts produced multiblock crystalline copolymers having stereo-regular sequences of both propylene and 1-butene. Copolymers containing from 3 to about 80% 1-butene had two DTA melting points which were attributable to polypropylene and poly-1-butene crystallinity. Those containing from 18 to about 70% 1-butene had x-ray diffraction patterns showing peaks characteristic of polypropylene and form I poly-1-butene, but form II poly-1-butene crystallinity was never observed. The multiblock copolymer structure observed is also supported by the fact that the product of the reactivity ratios is greater than unity. The composition distributions of low-conversion and continuously prepared copolymers were similar and relatively broad. For example, copolymers containing an average of 12% 1-butene had species containing from 5–30% 1-butene. High-conversion copolymers had an even broader composition distribution due to the gradual increase of the 1-butene concentration in the comonomer mixture as the copolymerization proceeded. The absence of homopolymers was demonstrated by fractionation. The ability to detect homopolymers was proved by the fact that a mixture of stereoregular polypropylene and poly-1-butene were readily separated. Increasing the amount of 1-butene tended to decrease those properties dependent upon crystallinity such as hardness, tensile strength, stiffness, density, and melting point, but tended to improve significantly the impact strength, low temperature properties, and clarity of molded objects. These duocrystalline copolymers retained a much higher level of properties than that observed for random copolymers prepared with less stereospecific coordination catalysts.  相似文献   

15.
Batch liquid pool and semibatch gas-phase polymerizations were performed with high-activity Ziegler-Natta catalysts to evaluate the effect of 1-butene on the crystallinity, the melt temperature and the average molecular weights of the final 1-butene/propylene copolymers and alloys. According to the obtained results, 1-butene can be significantly incorporated into the polymer chain over the whole range of copolymer compositions in both gas and liquid-phases, leading to the decrease of the melting temperature of the copolymer resins. On the other hand, the properties of the polymer alloys seem to be less sensitive to 1-butene incorporation, indicating the development of a distinct 1-butene phase. The average molecular weights, the polydispersities and the reactivity ratios are quite different in the liquid pool and gas-phase processes, indicating that sorption/ diffusion effects may exert an important role during the copolymerization. The obtained reactivity ratios in the gas-phase are close to 1, while the reactivity ratios of propylene are systematically higher than the reactivity ratios of 1-butene in the liquid pool process. Polymer materials with large molecular weights and good particle morphology can be obtained in all analyzed cases, indicating that development of propylene/1-butene copolymer grades is indeed possible in both liquid pool and gas-phase processes.  相似文献   

16.
Studies devoted to the homo-and copolymerization of propylene with ethylene and higher olefins (1-butene, 1-hexene, 4-methyl-1-pentene, 1-octene) in liquid propylene under the action of homogeneous metallocene catalysts of various types are surveyed in brief. The main kinetic features of the processes and the properties of the polymers are discussed. The optimal conditions for the highly efficient syntheses of isotactic, syndiotactic, hemiisotactic, and stereoblock PPs are described. It is shown that the combined cocatalyst—polymethylaluminoxane coupled with (i-Bu)3Al—shows promise for the processes under consideration. Depending on the type of catalyst used, the copolymerization of propylene with ethylene yields copolymers with a block, random, or close to alternating distribution of comonomer units in a polymer chain. The copolymerization of propylene with higher olefins in the monomer bulk initiated by highly active sterically hindered isospecific catalytic systems shows an ideal character, and the reactivity ratios are r 1r 2 ≈ 1; that is, the composition of the copolymer is equal to the composition of the monomer mixture at all comonomer ratios. It is demonstrated that the synthesis of homo-and copolymers of propylene in the monomer bulk in the presence of modern homogeneous catalysts is promising for highly efficient production of both traditional and new polymer materials with a unique combination of mechanical and thermal properties.  相似文献   

17.
Rates of oxygen absorption and formation of oxidation products were determined in γ-initiated oxidations of thin films of high- and low-density polyethylene, atactic and isotactic polypropylene, and of three ethylene–propylene copolymers. Radiation yields G for O2 absorbed and formation of hydroperoxides depend on dose rates and decrease sharply with increasing ethylene content of the copolymers and moderately with increasing crystallinity of any base polymer. G values for dialkyl peroxide and carbonyl formation, and therefore for chain initiation and termination, do not change much with polymer composition and crystallinity and not at all with dose rates. A few experiments with atactic polypropylene and an amorphous ethylene–propylene copolymer, initiated by di-tert-butylperoxy oxalate, indicate that 37 mole-% of ethylene in the polymer increases the efficiency of initiation and the tendency toward crosslinking.  相似文献   

18.
Intrinsic viscosities have been measured at 25° on five ethylene–propylene copolymer samples ranging in composition from 33 to 75 mole-% ethylene. The solvents used were n-C8 and n-C16 linear alkanes and two branched alkanes, 2,2,4-trimethylpentane and 2,2,4,4,6,8,8-heptamethylnonane (br-C16). This choice was based on the supposition that the branched solvent would prefer the propylene segments and the linear solvent the ethylene segments, due to similarity in shape and possibly in orientational order. It was found that [η]n ? [η]br ≡ Δ[η] is indeed negative for propylene-rich copolymers, zero for a 56% ethylene copolymer, and positive for ethylene-rich copolymers. The Stockmayer–Fixman relation was used to obtain from Δ[η] a molecular-weight independent function of composition. The quantities (Δ[η]/[η])(1 + aM?1/2) and Δ[η]/M are linear with the mole percent ethylene in the range investigated with 200 ≤ a ≤ 2000. The possibility of using these results for composition determination in ethylene–propylene copolymers is discussed. Intrinsic viscosities in the same solvents are reported for two samples of a terpolymer with ethylidene norbornene.  相似文献   

19.
Fast scanning chip calorimetry has been employed to study the effect of the type and concentration of co-units on the rate of mesophase formation and crystallization in random isotactic copolymers of propylene and 1-alkenes, including ethylene, 1-butene, 1-hexene, and 1-octene. The dependence of the rate of ordering on temperature of the propylene homopolymer shows two distinct maxima around 300 and 340–350 K which are related to mesophase formation and crystallization, respectively. Addition of 1-alkene co-units leads to a decrease of the maximum rate of both crystallization and mesophase formation. At comparable temperature and molar percentage of co-units in the propylene chain, ethylene, and 1-butene co-units cause less reduction of the maximum rate of ordering than 1-hexene or 1-octene co-units. The experimental observations are discussed in the context of possible incorporation of these chain defects into the ordered structures.  相似文献   

20.
The effect of a direct current discharge on the films of polypropylene and copolymers of propylene and hexene-1 synthesized with an isospecific catalytic system, rac-Me2SiInd2ZrCl2-polymethylaluminoxane, was investigated. The treatment of isotactic polypropylene films by the discharge did not affect the ratio of crystalline phases in the polymer to a measurable extent. However, for the plasma treated films of copolymers of propylene and hexene-1 (the hexene-1 content of 1-2 mol%), a structural transformation of γ-modification into α-modification has been noticed. The observed phase transition has no apparent relation to any changes in microstructure of the copolymer chain because melting temperature values and the stereoregularity parameters of the samples remained practically unchanged. An experimental investigation of the specific influence exerted by individual components of a direct current discharge on the crystalline structure of copolymers has been undertaken. The exposure to a quantum component of the discharge did not induce any changes in the phase composition of the irradiated samples. The heating of the samples led to a negligible change of their phase composition. It has been determined that the surface of polypropylene and propylene/hexene-1 copolymer films facing the cathode in the course of the direct current discharge treatment had an accumulated negative charge Q > 10 nC/cm2 which persisted for a long time afterwards. It has been suggested that the electrical field of a negative discharge may be the main cause of the γ-into α-phase transition in propylene/hexene-1 copolymers under the plasma effect. To verify this assumption, a propylene/hexene-1 copolymer film was charged under electron beam with energy of 4 keV. The electron beam treatment of the film resulted to the negative charge value of 11 nC/cm2. The electron beam irradiation has induced the phase transition which was quite similar to the transition observed as the result of plasma treatment. So, it may be concluded that the phase transition from crystal γ-modification to α-modification under the effect of direct current discharge which has been investigated for copolymers of propylene and hexene-1 is induced by electric field of the negative charge accumulated at the surface layers of the films of the copolymers.  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