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 31 毫秒
1.
13C NMR spectra of propylene-1-butene copolymers have been studied at 90.55 MHz. At this high field, many lines previously not resolvable were found to be sensitive to comonomer sequence distribution; these microstructures have been fully assigned. Each NMR spectrum was analyzed by a reaction probability model. Information available includes polymer composition, comonomer sequence distribution, and Markovian reaction probabilities. The use of this model reduces errors in the determination of sequence distribution and furthermore enables the product r1r2 of the reactivity ratios to be determined for the copolymer/catalyst system.  相似文献   

2.
The 22.6-MHz Fourier-transform noise-decoupled 13C (carbon-13) NMR spectra of several ethylene–vinyl acetate (E–VA) copolymers were obtained. We found that triad information on monomer placement can be deduced from carbonyl resonances, triad and pentad information can be deduced from methine carbon resonances, and triad information is available from the methylene carbon resonances. The random comonomer distributions in E–VA polymerizations were demonstrated up to pentad placements. In addition, the use of model-compound data in the analysis of copolymer spectra was shown.  相似文献   

3.
In this work, two series of ethylene-dimethylaminoethylmethacrylate (EDAM) and ethylene-methyl acrylate copolymers (EMA) with different comonomer content were studied by high-resolution solid-state 13C NMR spectroscopy. Biexponential 13C spin-lattice relaxation behaviors of the crystalline region were observed for all copolymer samples either melt-quenched or isothermally crystallized. The relative content of the shorter 13C T1 component to that of the longer 13C T1 component was found to increase with comonomer content. By employing a new pulse sequence which can be considered as a combination of Goldman-Shen's and Torchia's pulse sequences, it was demonstrated that the shorter 13C T1 component is corresponding to the intermediate part of the crystalline region. The thickness of the intermediate part is estimated to be about 0.85 nm.  相似文献   

4.
We report the complete 13C NMR characterization of a series of ethylene–propylene–1‐hexene terpolymers obtained with the metallocenic system rac‐ethylene bis‐indenyl zirconium dichloride, with different comonomer ratios. A detailed study of 13C NMR chemical shifts, triad sequence distributions, monomer‐average sequence lengths, and reactivity ratios for these terpolymers is presented. © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 2474–2482, 2004  相似文献   

5.
Acrylonitrile-methyl acrylate (A/M) copolymers of different monomer compositions were prepared by bulk polymerization using free radical initiator (benzoyl peroxide). Copolymer compositions were determined by elemental analyses and comonomer reactivity ratios were determined by the nonlinear least squares errors-in-variables methods (EVM). Terminal and penultimate reactivity ratios have been calculated using the observed monomer triad sequence distribution determined from 13C{1H}-NMR spectra. The triad sequence distribution was used to calculate diad concentrations, conditional probability parameters, number-average sequence lengths, and run number in the copolymers. The observed triad sequence concentrations determined from 13C{1H}-NMR spectrum agreed well with those calculated from reactivity ratios. Glass transition temperatures (Tg) of various copolymers determined from DSC gave good agreement with those obtained from NMR. © 1992 John Wiley & Sons, Inc.  相似文献   

6.
Poly(isobutylene‐co‐p‐methylstyrene) is an important precursor to Exxpro™ elastomers. A previous report detailed the characterization of both the proton and the carbon NMR spectra of the copolymer. 1 However, several resonances in the proton NMR spectrum of the copolymer were not assigned. Specifically, the proton methine resonance of the BSB triad sequence is now identified and used to calculate BSB triad contribution to the copolymer microstructure. This report describes the assignment of this resonance and other resonances associated with microstructural sequence distribution around p‐methylstyrene. The proton NMR signals of interest resonate at 2.8 ppm and 2.5 ppm in a typical spectrum for poly(isobutylene‐co‐p‐methylstyrene). The nature of these resonances were determined by preparation and characterization of specifically deuterated poly(isobutylene‐co‐p‐methylstyrene)s employing both one and two dimensional NMR techniques. The 2.8 ppm signal is assigned as the methine proton of a p‐methylstyrene incorporated between two isobutylene units (the BSB triad). The signal at 2.5 ppm is assigned to the meso‐BSS triad. Determination of these resonances allows for rapid evaluation of isolated p‐methylstyrene units (BSB triads) present in the copolymer using only 1H NMR. The utility of this technique is demonstrated by comparing BSB triad values determined by 1H and 13C NMR analysis. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 1680–1686, 2000  相似文献   

7.
The microstructure of vinylacetate-maleic anhydride copolymer is investigated by UV, IR, 1H, and 13C NMR methods alongside with quantum-chemical calculations. The macromolecules of the equimolar copolymer are found to consist of the sequence of comonomer units. As soon as the copolymer interacts with atmospheric water and solvents, 2,5-dihydroxyfuran derivatives stabilized by hydrogen bonding are formed in the copolymer macromolecules owing to cycloanhydride-enolic tautomerism. These derivatives are quasi-aromatic by their nature.  相似文献   

8.
An incremental method for characterizing triad and pentad distribution by 13C NMR spectroscopy was applied to the poly(methyl methacrylate-co-isobutyl acrylate) copolymer. Calculation of the intensity of individual lines was performed applying Bernoulli statistics, while the chemical shifts for each sequence were calculated by an incremental method. Based on these data, the carbonyl signal was simulated yielding good agreement at the triad and pentad level.  相似文献   

9.
To describe the detailed microstructure of homogeneous ethylene‐1‐alkene copolymer chains and to study the characteristics of single site catalysts, Markov statistics are used to fit peak intensities of all relevant 13C NMR signals of series of copolymers. In the case of the occurrence of inverted comonomer units, a first‐order Markov terpolymer is applied, otherwise a second‐order Markov copolymer model. Chain propagation probabilities are obtained via modeling of the entire NMR spectrum. This procedure results in an accurate reproduction of the chain microstructure, including ethylene, 1‐alkene, and methylene sequence length distributions. If the experimental (co)monomer feeds are known, the reactivity ratios and the theoretical (co)monomer feeds are also found providing information about the copolymerization kinetics and the characteristics of the catalyst. © 2006 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 44: 722–737, 2006  相似文献   

10.
乙烯基单体-N-苯基马来酰亚胺共聚物序列结构研究   总被引:4,自引:3,他引:1  
用NMR谱研究了甲基丙烯酸甲酯(MMA)-N-苯基马来酰亚胺(PMI)、苯乙烯(St)-PMI共聚物的序列结构.结果表明,MMA-PMI共聚物单元属无规序列分布,St-PMI共聚物单元属交替序列分布.由1HNMR结果可得MMA-PMI共聚物空间立构部分信息,由13CNMR三单元组实验结果算得的序列长度与末端基理论计算结果一致,且MMA-PMI共聚物链属一级Markov链.由St-PMI共聚物序列长度与末端基理论计算结果的偏差提出更为合理的增长基元反应.  相似文献   

11.
The copolymerization of L ,L -lactide and ε-caprolactone was carried out using antimony trioxide and stannous octoate as catalysts. The effect of polymerization catalysts on the physical and the chemical microstructures of this copolymer was investigated by 13C NMR and DSC analysis. Antimony trioxide causes more random sequence distribution within the copolymer chain due to its higher transesterification characteristic than stannous octoate. The copolymer samples made with the antimony trioxide catalyst seem to have more amorphous phase structure, than those prepared using stannous octoate which are semicrystalline for the entire compositional range due to blocky copolymer sequences. © 1994 John Wiley & Sons, Inc.  相似文献   

12.
Copolymerization of propene and 1-olefins including 1-octene, 1-decene, 1-dodecene, 1-tetradecene, and 1-hexadecene were studied with the catalyst system MgCl2/TiCl4-Al(i-Bu)3. It was found that the polymerization productivity and the consumption rate of propene are improved significantly in the presence of the comonomer. The total productivity of propene/1-olefin copolymerizations decreases as follows: 1-octene> 1-decene>1-dodecene>1-hexadecene>1-tetradecene. The reactivity ratios were estimated from the copolymerization results. 13C NMR was used to characterize the microstructures of propene/1-octene copolymer. Finally, the oxygen enrichment behavior of propene/1-octene copolymer was investigated.  相似文献   

13.
Acrylamide/vinyl acetate and acrylamide/vinyl propionate copolymers were prepared by solution polymerization using benzoyl peroxide as the initiator. The copolymer composition was determined from the percent nitrogen in the copolymers.

The stability constants and related thermodynamic parameters (e.g., ΔG°, ΔH°, and ΔS°,) of the interpolymer complexes with Poly(ethyleneimine) were determined by using Osada's method. These parameters have been correlated with the sequence distribution of monomer units in the copolymer chains which were obtained from 13C{1H} NMR spectroscopy. The sequence distribution of the comonomer units in the copolymer chains influence the association between the copolymers and the polyelectrolyte which is reflected on the stability of the interpolymer complexes.  相似文献   

14.
13C-NMR spectroscopy was used in a detailed study of vinyl chloride-vinyl acetate and vinyl chloride-vinyl propionate copolymers. The NMR spectra of the methylene carbon region showed three split peaks whose intensities changed with composition of the copolymers. These peaks were assigned to diad sequences and the observed diad concentrations were in good agreement with the calculated concentrations in terms of the random copolymerization theory. For the methine carbon spectra of vinyl acetate or vinyl propionate units in the copolymers the degree of splitting of the signal was improved by the addition of tris(1,1,1,2,2,3,3-heptafluro-7,7-diemthyl-1,4,6-octanedinata)-praseodymium as a shift reagent. Four peaks assigned to the methine carbon were interpreted in terms of triad sequence distribution and tacticity.  相似文献   

15.
Methyl acrylate–styrene copolymers of different copolymer compositions were free-radically prepared. The relative intensities of the carbonyl frequencies of the methyl acrylate units at v 1730 cm?1 were correlated with the copolymer composition. The positions and shapes of the carbonyl bands in the infrared absorption spectra of the copolymers-dissolved in chloroform, were shown to depend on the composition of the copolymers and upon the presence of different proportions of methyl acrylate centered triads. The results obtained by infrared spectroscopy were compared with those obtained by 13C-NMR. Infrared spectra may be used to yield information about both the copolymer composition and the triad sequence distribution.  相似文献   

16.
The copolymerization of acrylamide (AM) with sodium 3-(N-propyl)acrylamido-3-methlbutanoate (NaPAMB) has been studied in the range from 40 to 90% AM in the feed. The copolymer compositions have been determined from elemental analysis and 13C NMR. Reactivity ratio studies have been performed, and the value of r1r2 determined to be 0.36. The molecular weights, as determined by low-angle laser light scattering of the copolymers, were found to decrease sharply with increasing NaPAMB content, and were in the range 0.5–7.0 × 106. The copolymer microstructures, including mean sequence length distributions, were calculated from the reactivity ratios. Knowledge of polymer composition, micro-structure, and molecular weight is utilized for assessment of structure/dilute solution property relationships reported in a subsequent paper in this series.  相似文献   

17.
Glycolide (GL) and ?‐caprolactone (CL) were copolymerized in bulk at relatively high temperatures using stannous octoate as a catalyst. To investigate the relationship among microstructure, thermal properties, and crystallinity, three series of copolymers prepared at various reaction temperatures, times, and comonomer feed ratios were prepared and characterized by 1H and 13C NMR, DSC, and wide‐angle X‐ray diffraction (WAXD). The 600‐MHz 1H NMR spectra provided information about not only the copolymer compositions but also about the chain microstructure. The reactivity ratios (rG and rC) were calculated from the monomer sequences and were 6.84 and 0.13, respectively. In terms of overall feed compositions, the sequence lengths of the glycolyl units calculated from the reactivity ratios exceeded those measured from the polymeric products. Mechanistic considerations based on reactivity ratios, monomer consumption data, and average sequence lengths are discussed. The unusual phase diagram of GL/CL copolymers implies that the copolymer melting temperature does not depend on its composition alone but rather on the nature of the sequence distribution. The DSC and WAXD measurements show a close relationship between polymer crystallinity and the nature of the polymer sequence. © 2002 John Wiley & Sons, Inc. J Polym Sci Part A: Polym Chem 40: 544–554, 2002; DOI 10.1002/pola.10123  相似文献   

18.
13C NMR spectra of styrene/butyl methacrylate copolymers prepared in bulk at 125°C using 2,2,6,6‐tetramethyl‐1‐(1‐phenylethoxy)piperidine as the initiator were analyzed. After assignment of the carbon resonances of the aromatic C‐1 region, the Mayo‐Lewis terminal model (MLTM) was tested by comparison of experimental and predicted values of configurational triad sequences. It follows that MLTM provides an excellent prediction across a wide conversion range for all comonomer feeds.  相似文献   

19.
The Ziegler–Natta system Cp*TiMe3/B(C6F5)3 catalyzed the copolymerization of ethylene and 1‐hexene in toluene into materials that were characterized by 1H and 13C{1H} NMR spectroscopy, differential scanning calorimetry, and gel permeation chromatography. The effects of temperature and ethylene/1‐hexene and olefin/catalyst ratios on catalyst activities and copolymer molecular weights and molecular weight distributions were studied; the ethylene proportions varied from less than 5% to 85% or more. In addition, significant amounts of 1‐hexene were incorporated into the growing polymer chain in a 2,1‐fashion; consequently, conventional 13C NMR analytical methodologies for deducing monomer proportions and dispersions and polymer microstructures, based on a low 1,2‐incorporation of α‐olefin, did not work very well. A soluble (in toluene at ambient temperature) but very high molecular weight (weight‐average molecular weight ∼ 8 × 105, weight‐average molecular weight/number‐average molecular weight = 1.8) rubbery copolymer that formed at −78 °C exhibited a predominantly alternating microstructure. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 3966–3976, 2000  相似文献   

20.
The kinetics and reactivity ratios of styrene‐acrylonitrile (SA) copolymerization have been studied extensively in bulk and in a variety of solution media using conventional free radical polymerizations (FRPs). Due to the significant difference in the two reactivity ratios for this monomer pair, at certain feed ratios the copolymers display composition drift with conversion due to monomer depletion. In this study, the kinetics of SA copolymerization using Reversible Addition‐Fragmentation Chain Transfer (RAFT) has been studied in bulk at 80 °C. The reactivity ratios for the terminal model were calculated from the comonomer sequence distributions for the RAFT process at low conversion for nine different compositions and found to be in the same range as those reported for conventional FRP of SA. The changes in the composition and sequence distribution with conversion were studied for three feed compositions. The copolymers show compositional drift with conversion, except at the azeotropic composition, and match the predictions from the reactivity ratios obtained at low conversion. From quantitative 13C NMR the triad distributions of these copolymers were estimated and found to match the predicted triad distributions as conversion increased. © 2017 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2017 , 55, 919–927  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