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 812 毫秒
1.
It was found that structurally isomeric polymers were formed by the ring-opening polymerization of β-(2-acetoxy ethyl)-β-propiolactone with (EtAlO)n and Et(ZnO)2ZnEt catalysts; that is, the Al catalyst catalyzed normal polymerization which led to poly-β-ester and the Zn catalyst formed isomerized poly-β-ester as the main product. The polymer structure was determined by nuclear magnetic resonance (NMR), T1-value, thermal decomposition product, and (Tg). The NMR studies for the monomer–catalyst systems indicated that the Al catalyst interacted predominately with the lactone group, whereas the Zn catalyst interacted with the side-chain ester group. These site-selective interactions could be related to the difference in the stereoregulation by the two catalysts during the poly(β-ester)-forming polymerization process.  相似文献   

2.
4-Vinylpyridinium trifluoromethanesulfonate monomers substituted at nitrogen with H, O, CH3, C2H5, C6H13, and C12H25 were synthesized and characterized spectroscopically. Thermal analyses (DSC and TGA) were carried out on all the compounds. The solid monomers (N? H, N? CH3, N? C6H13, and N? C12H25) exhibited endothermic melting followed by exothermic polymerization and exothermic decomposition (>400°C). Liquid N? C2H5 monomer revealed only exothermic polymerization and decomposition. The N? O polymer underwent thermal decomposition below 300°C. The N–C12H25 homopolymer, prepared from monomer in the DSC or in bulk, displayed an unusual thermal transition at 250°C, which has been attributed to a polymer backbone reorientation leading to side-chain ordering of the dodecyl groups.  相似文献   

3.
The isomerization and polymerization of propenylbenzene (PB) with various Ziegler–Natta catalyst systems have been investigated. With the TiCl3–(C2H5)3Al (Al/Ti > 2.0) catalyst at 80°C, PB polymerized to give a polymer exclusively consisting of allylbenzene (AB) unit. During the polymerization, AB, which polymerized readily with the catalyst, was produced through isomerization of PB, indicating that PB underwent monomer-isomerization polymerization. PB also polymerized with isomerization to AB in the presence of TiCl3?(C2H5)2AlCl?NiCl2 catalyst system, and a copolymer with PB and AB units was obtained. With TiCl3?C2H5AlCl2 catalyst, poly(PB) was formed via ordinary vinylene polymerization without isomerization. From these facts, it was concluded that the structure of the polymers produced from PB widely changed, depending on the catalyst systems used, which determine the rate of isomerization to AB and the polymerization reactivity of the PB and AB isomers formed.  相似文献   

4.
Bulk polymerization of alkyl N,N-dialkylfumaramates (FAE) and maleamates (MAE) was performed in the presence of a radical initiator. It has been found that FAE is more reactive than MAE when the reactivity of the two geometrical isomers was compared for their homo- and copolymerizations. From investigation on the effect of ester and N-substituents of these monomers, it has been found that the isopropyl ester shows a higher reactivity than the methyl ester and that N-ethyl and n-butyl substitution gives polymers with high molecular weight of more than several thousands. The resulting substituted polymethylenes from FAE and MAE were characterized and compared with each other. The isomerization of MAE to FAE with morpholine as an isomerization catalyst and monomer-isomerization radical polymerization were also investigated.  相似文献   

5.
Protonated aromatic aldehydes and methyl ketones 1a–10a, carrying initially the proton at the carbonyl group, are prepared by electron impact-induced loss of a methyl radical from 1?arylethanols and 2?aryl?2?propanols, respectively. The aryl moiety of the ions corresponds to a benzene group, a naphthalene group, a phenanthrene group, a biphenyl group, and a terphenyl group. respectively, each substituted by a CH3OCH2 side-chain as remote from the acyl substituent as possible. The characteristic reactions of the metastable ions, studied by mass-analyzed ion kinetic energy spectrometry, are the elimination of methanol, the formation of CH3OCH 2 + ions, and the elimination of an ester RCOOCH3 (R = H and CH3) . The mechanisms of these fragmentations were studied by using D-labeled derivatives. Confirming earlier results, it is shown that the ester elimination, at least from the protonated aryl methyl ketones, has to proceed by an intermediate [acyl cation/arylmethyl methyl ether]-complex. The relative abundances of the elimination of methanol and of the ester decrease and increase, respectively, with the size of the aromatic system. Clearly, the fragmentation via intermediate ion-neutral complexes is favored for the larger ions. Furthermore, the acyl cation of these complexes can move unrestricted over quite large molecular distances to react with the remote CH3OCH2-side-chain, contrasting the restricted migration of a proton by 1,2-shifts (“ring walk”) in these systems.  相似文献   

6.
The isomerization of CH3S(OH)CH2 to CH3S(O)CH3 in the absence and presence of water has been investigated at the G3XMP2//B3LYP/6-311 + G(2df, p) level. The naked isomerization, the reaction without water, gives the high barrier height (21.56 kcal.mol^-1). Three models are constructed to describe the water influence on the isomerization, that is, water molecules are the catalyst and the microsolvation, and water molecules act as the catalyst and microsolvation simultaneously. Our results show that the isomerization barrier heights of CH3S(OH)CH2 to CH3S(O)CH3 are reduced by 12.32, 11.04, and 7.80 kcal.mol^-1, respectively, when one, two, and three water molecules are performed as catalyst, in contrast to the naked isomerization. Moreover, the rate constants of the isomerization are calculated using the transition state theory with the Wigner tunneling correction over the temperature range of 240-425 K. We find that the rate constant of a single water molecule as the catalyst is 1.58 times larger than the naked isomerization at 325 K, whereas it is slower by 6 orders of magnitude when water molecule serves as the microsolvation at 325 K, compared to naked reaction. So the water-catalyzed isomerization of CH3S(OH)CH2 to CH3S(O)CH3 is predicted to be the key role in lowering the activation energy. The isomerization involving water molecules acting as mierosolvation is unfavorable under atmospheric conditions.  相似文献   

7.
N,N‐dialkylaminoethyl methacrylate (DAEA) monomers are extensively used to prepare multi‐responsive polymers. However, these monomers face high risk of hydrolysis in their ester groups when being polymerized in water‐containing medias. Here, NMR spectroscopy was employed to continuously track the hydrolysis and solubility of four widely used DAEA monomers [CH2CH2R1COO(CH2)2N(R2)2; R1 = H or CH3; R2 = CH3, CH2CH3 or CH(CH3)2] under typical polymerization conditions. With this technique, the hydrolysis reactivity and absolute hydrolysis amount of these monomers are separately examined, and then their kinetic correlations with solubility, molecular structure, pH, and temperature are established, so that the hydrolysis of DAEA monomers and even other esters with similar cyclic structure can be predicted. The present efforts are expected to provide a general understanding for the hydrolysis of all the DAEA monomers, benefitting to the optimization of polymerization toward well‐defined DAEA copolymers, as well as the design of smart soft matter for specific applications. © 2018 Wiley Periodicals, Inc. J. Polym. Sci., Part B: Polym. Phys. 2018 , 56, 914–923  相似文献   

8.
Polymerization of 2‐pentene with [ArN?C(An)C(An)·NAr)NiBr2 (Ar?2,6‐iPr2C6H3)] ( 1‐Ni) /M‐MAO catalyst was investigated. A reactivity between trans‐2‐pentene and cis‐2‐pentene on the polymerization was quite different, and trans‐2‐pentene polymerized with 1‐Ni /M‐MAO catalyst to give a high molecular weight polymer. On the other hand, the polymerization of cis‐2‐butene with 1‐Ni /M‐MAO catalyst did not give any polymeric products. In the polymerization of mixture of trans‐ and cis‐2‐pentene with 1‐Ni /M‐MAO catalyst, the Mn of the polymer increased with an increase of the polymer yields. However, the relationship between polymer yield and the Mn of the polymer did not give a strict straight line, and the Mw/Mn also increased with increasing polymer yield. This suggests that side reactions were induced during the polymerization. The structures of the polymer obtained from the polymerization of 2‐ pentene with 1‐Ni /M‐MAO catalyst consists of ? CH2? CH2? CH(CH2CH3)? , ? CH2? CH2? CH2? CH(CH3)? , ? CH2? CH(CH2CH2CH3)? , and methylene sequence ? (CH2)n? (n ≥ 5) units, which is related to the chain walking mechanism. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 2858–2863, 2008  相似文献   

9.
Stereoregular polymerization involving asymmetry at phosphorus has been obtained from ethylene methyl or phenyl phosphorothioate with R2Mg? NH3 catalysts, or, in some cases, with R2Mg alone. The methyl ester gave two types of polymer: an amorphous rubber and a low-melting (75°C) crystalline polymer. The phenyl ester gave mainly a low-melting (68°C) crystalline polymer of 2.2 inherent viscosity. Proton and 31P NMR and infrared spectra of these polymers are in accord with the expected chain unit, ? CH2CH2? O? P(S)(OR)? O? . The polymerization mechanism probably involves an anionic ring-opening step with P? O cleavage. Ring opening with C? O cleavage appears to be largely excluded. This conclusion is based on the expectation that anionic ring opening with C? O cleavage should lead to a rearranged chain unit, ? CH2CH2? O? P(O)? (OR)? S? , because of the high nucleophilicity of sulfur as compared with oxygen. Proton and 31P NMR spectra give no evidence for the rearranged unit within the limit of detection (ca. 3%). However, on aging, the methyl ester polymer changes drastically to form up to 40% CH2SP groups. Presumably, the polymer undergoes the well-known thiono-thiolo rearrangement characteristic of simple phosphorothioate esters to form ? CH2CH2? O? P(O)(SCH3)? O? chain units. The phenyl ester polymer is stable under the same aging conditions.  相似文献   

10.
1,4-Cyclohexadiene underwent monomer-isomerization polymerization to yield poly(1,3-cyclohexadiene) with a Ziegler-Natta catalyst comprising TiCl4–Al(C2H5)3 catalyst with Al/Ti molar ratios of 0.5–3.0 at 60°C for 96 hr. Good yields of polymer were obtained (49.5% yield at Al/Ti = 3.0; [η] = 0.04 dl/g). The infrared and NMR spectra of the polymer were identical to those of poly-(1,3-cyclohexadiene), confirming that 1,4-cyclohexadiene first isomerizes to 1,3-cyclohexadiene and then homopolymerizes to give poly-1,3-cyclohexadiene. 1,3-Cyclohexadiene polymerized without isomerization easily in the presence of TiCl3–Al(C2H5)3 catalyst at Al/Ti molar ratios of 0.5–3.0 at 60°C for 3 hr (76.3% yield at Al/Ti = 3.0; [η] = 0.06 dl/g).  相似文献   

11.
Thiophene monomers displaying a dimethylenecarboxylate (CH2CH2COOR) substituent on the 3‐position of the aromatic ring can be easily obtained and in one step from the electrochemically induced reaction of 3‐bromothiophene with the corresponding acrylate (CH2?CHCOOR). The synthesis of the ethyl ester monomer, of related 2,5‐dihalogenothiophenes, and their polymerization are reported. Despite the surprisingly low solubilities displayed by the polymers, a full spectroscopic characterization could be performed and the data fully analyzed. Oxidative polymerizations (FeCl3 or electropolymerization) yield a regioirregular polythiophene, with 60–70% of head‐to‐tail diads. Both experimental and theoretical results suggest that the nonconjugated ester plays a very minor role—if any—in the polymerizations under oxidative conditions, but has a significant influence on the polymer properties. Preliminary attempts to polymerize the dihalogenothiophenes under reductive conditions were hampered by the even lower solubilities exhibited by the regioregular oligomers. © 2011 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2012  相似文献   

12.
The vinylic polymerization of norbornene and its copolymerization with norbornene carboxylic acid methyl esters were investigated. Norbornene was polymerized by us using di-μ-chloro-bis-(6-methoxybicyclo[2.2.1]hept-2-ene-endo-5σ,2π)-palladium(II) as catalyst. The polymerization time can be decreased by a factor of 100000 by activation of the catalyst with methylaluminoxane (MAO). With this palladium catalyst activated by MAO, 140 t of norbornene can be polymerized per mol palladium per h. This catalyst system was much more active than [Pd(CH3CN)4](BF4)2 ( I ). The polymerization of norbornene by (6-methoxybicyclo[2.2.1]hept-2-ene-endo-5σ,2π)-palladium(II) tetrafluoroborate was also possible but it was not as fast as the polymerization by Pd catalysts activated with MAO. We were also able to obtain copolymers of norbornene and 5-norbornene-2-carboxylic acid methyl ester (exo/endo = 1/4 or 2/3) containing between 15 and 20 mol-% ester units. The copolymerization of norbornene and 2-methyl-5-norbornene-2-carboxylic acid methyl ester (exo/endo = 7/3) was faster than the copolymerization mentioned before. In contrast the homopolymerization of 2-methyl-5-norbornene-2-carboxylic acid methyl ester was 10 times slower than that of 5-norbornene-2-carboxylic acid methyl ester (exo/endo = 1/4).  相似文献   

13.
In this work, two novel monomers, a cyclooctene and a norbornene derivative, including the trifluoromethylphenyl ester group, are prepared by chemical modification of 1,5-cyclooctadiene and 5-norbornene-2-carboxylic acid, respectively. The ring-opening metathesis polymerization of each one of these monomers, as well as the commercial carboxylic acid, is confirmed in the presence of Grubbs’ second generation catalyst in CH2Cl2. The three functionalized monomers are subsequently copolymerized with cis-cyclooctene in order to obtain modified polycyclooctenes. Solubility is enhanced by the incorporation of the functional group. The thermal properties of the obtained homopolymers and copolymers are studied by DSC and TGA, showing that crystalline behavior strongly depends on the cyclooctene content.  相似文献   

14.
Four α‐diimine nickel complexes [(Ar? N?C(R)? C(R)?N? Ar)NiBr2; R?H, CH3, cyclohexane‐1,2‐diyl, naphthalene‐1,8‐diyl, Ar?2,6‐i‐Pr2‐C6H3‐) were investigated in propene and hex‐1‐ene polymerization to identify the limits of backbone substituent R size needed to provide living/controlled α‐olefins polymerization by the sufficient suppression of βH elimination transfer. Propagation kinetics measurements, molar mass on monomer conversion dependencies and reinitiation tests were used to evaluate the livingness of hex‐1‐ene polymerization. Interestingly, living/controlled hex‐1‐ene polymerization was observed in the case of all diimine derivatives including the one bearing only hydrogen atom in backbone positions. Unexpectedly, in the case of catalysts bearing H and CH3 backbone substituents, we observed the unusual isomerization of hex‐1‐ene into internal hexenes in parallel with its polymerization. Nevertheless, by subtracting the amount of monomer consumed in isomerization side reaction, polymerization still keeps the features of living/controlled process. © 2016 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2016 , 54, 3193–3202  相似文献   

15.
The potential of several alkylcobalt complexes as catalysts for hydrogenation and isomerization of alkenes has been investigated. The complexes CH3Co(CO)2(Pom-Pom) (Pom-Pom = 1,2 bis(dimethoxyphosphino)ethane), CH3Co(CO)3P(OMe)3 and C6H5CH2Co(CO)3PPh 3 are compared to CH3Co(CO)2(P(OMe)3)2, for their ability to function in catalytic cycles. Each is active for hydrogenation and isomerization of alkenes under conditions where the carbonylation-decarbonylation equilibrium is readily established. The lifetime for the complexes is much shorter than for CH3Co(CO)2 (P(OMe)3)2 suggesting that two phosphorus donors in trans positions in an intermediate is a requirement for catalyst stability in these alkylcobalt complexes.  相似文献   

16.
Several α-olefins containing the trifluoromethyl group were prepared and characterized. 4,4,4-Trifluoro-1-butene, 3-trifluoromethyl-1-butene, 5,5,5-trifluoro-1-pentene, and 4-trifluoromethyl-1-pentene were homopolymerized with VCl3–Al(i-Bu)3 catalyst. The trifluorobutenes gave low-melting polymers with low fluorine contents. Polymers obtained from the trifluoropentenes were soluble having moderately high intrinsic viscosities. Copolymerizations of these monomers with their nonfluorinated homologs by the same catalyst system indicated low reactivities of the fluoromonomers. Nuclear magnetic resonance spectra of the fluorinated and nonfluorinated monomers and their respective spectroscopic studies with the catalyst (C5H5)2TiCl2–Al(CH3)3 indicated an electron deficiency of the vinyl group of the fluorobutenes. This was related to the inductive effect of the trifluoromethyl group. The inductive effect of this group was absent in the fluoropentenes and the nonfluorinated monomers. The electron-deficient vinyl group of the fluorobutenes apparently did not allow these monomers to coordinate with the active sites of the catalyst. Polymerization studies of the nonfluorinated monomers, 1-butene, 3-methyl-1-butene, 1-pentane, and 4-methyl-1-pentene, with the catalyst VCl3–Al(Bu)3, were performed in the presence of compounds containing the trifluoromethyl group. Results indicated that this group did not retard the rate of polymerization of these monomers. Evidence is presented to show that a catalytic amount of benzotrifluoride enhanced the rate of polymerization of α-olefins, particularly that of sterically hindered monomers such as 3-methyl-1-butene.  相似文献   

17.
The effect of spiro cyclohexane substitution on the polymerizability of the 1,3,2-dioxathiolan-4-one-2-oxide ring in various solvents has been examined. The steric hindrance of the cyclohexane ring inhibits the bimolecular chain propagation reaction which involves direct attack by a terminal hydroxyl group on the ring and which has been shown to occur in simpler dioxathiolan systems. The conjoined cyclohexane ring does not, however, markedly affect the “thermal” polymerization which occurs in nonhydroxylic solvents and in which chain propagation is thought to involve a reactive α-lactone intermediate. The rate-determining step in the sequence of reaction leading to polymer formation is a ring-scission process in which sulfur dioxide is evolved and the α-lactone intermediate formed. The values of the activation energy (25–30 kcal/mole) and frequency factor (1011–1013sec?1) associated with this reaction are, therefore, those which govern the the overall polymerization, since the subsequent steps are sensibly instantaneous. In the presence of adventitious traces of water the resultant polymer, poly(1-hydroxycyclohexanecarboxylic acid) has one carboxyl and one hydroxyl endgroup per chain. Polymers having M?n ~ 15,000 are readily obtained; these are amorphous materials, in contrast to the analogous poly-β-ester and dialkyl-substituted poly-α-esters which are crystalline. At temperatures in excess of 120°C a competitive first-order fragmentation reaction leading to the formation of cyclohexanone, carbon monoxide, and sulfur dioxide was observed. Kinetic studies demonstrated that this reaction, which is characterized by an activation energy of ~40 kcal/mole is unimportant, in the sense that it does not interfere with polymer formation at temperatures below 100°C.  相似文献   

18.
The direct polymerization of deprotonated acidic monomers in aqueous solutions was achieved via surface‐confined atom transfer radical polymerization (SC‐ATRP) to produce surface‐tethered polyelectrolyte brushes. Layers of poly(itaconic acid), poly(methacrylic acid), and sodium poly(styrene sulfonate) were grown by SC‐ATRP from self‐assembled initiator monolayers of [BrC(CH3)2COO(CH2)11S]2 on gold substrates. The polymer layers were characterized with variable‐angle ellipsometry and external‐reflection Fourier transform infrared spectroscopy. Without intervention, atom transfer radical polymerization catalysts were deactivated by complexation with the deprotonated acidic monomers, disproportionation, and dissociation during the polymerization of these monomers in water; the result was the cessation of polymer growth. The addition of an alkali salt to the reaction media suppressed catalyst deactivation, allowing polymer layers to increase in thickness linearly for longer periods of time with respect to salt‐free conditions. This result suggested an improved degree of polymerization control. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 45: 566–575, 2007  相似文献   

19.
This study investigated the synthesis of the biocatalyst, magnesium 2,6‐di‐tert‐butyl‐4‐methylphenoxide (Mg(BHT)2) complex, and the ring‐opening polymerization (ROP) of ε‐caprolactone (CL). The complex demonstrates high catalytic activity and controllable of molecular weight for the ROP of CL in tetrahydrofuran at room temperature, even when polymerization was performed under air. Before this study, the polymerization of CL had never been performed using a magnesium catalyst under air at room temperature. Various forms of alcohols with different purposes were also used as initiators with Mg(BHT)2. The results show that the magnesium complex acts as a perfect catalyst because of its high catalytic activity and control ability without any cytotoxicity in the polymerization of CL, making it suitable for biomedical applications. In addition, nanoparticle formation, cytotoxicity, and phototoxicity of tri‐2‐hydroxyethyl ester [Ce6‐(CH2CH2OPCL)3] were also studied in this article and Ce6‐(CH2CH2OPCL)3 formed nanoparticle can act as a nanophotosensitizer for photodynamic therapy. © 2012 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2012  相似文献   

20.
The reactivity of n-butyllithium-CH3OCH2CH2OLi in toluene (System A) in styrene-butadiene copolymerization reaction was found to be similar to that of the catalyst system n-butyllithium-poly(propylene oxide), which formed copolymers containing more styrene units than the feed monomer ratio at early stage of polymerization. Metalation reaction of toluene to produce benzyllithium was observed to take place in System A. Relative reactivity of each of chemical species present in System A was measured in the reactiom with butadiene, styrene, and 1,1-diphenylethylene, in connection with initiation and propagation reactions of copolymerization. From this study a living end anion possessing larger conjugated system was concluded to be more reactive toward monomers possessing larger conjugated system.  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