共查询到20条相似文献,搜索用时 15 毫秒
1.
Polycarbonates (PCs) were prepared by the oxidative carbonylation of bisphenol A and carbon monoxide with a hydrotalcite‐supported Pd–Co complex, a Pd–Co–poly(4‐vinylpyridine) complex [Pd–Co–(p‐4vpy)], a Pd–Co–polystyrene‐supported triphenylphosphine complex (Pd–Co–PS‐TPP), or a Pd–Co–polyvinylpyrrolidone complex (Pd–Co–pvp) as a heterogeneous Pd–Co bimetal catalyst to separate the PC solution and the Pd–Co bimetal catalyst after the reaction. Propylene carbonate was used as a halogen‐free solvent. Pd–Co–(p‐4vpy) and Pd–Co–PS‐TPP showed recycling potential, whereas Pd–Co–pvp, though not having recycling potential, yielded a high turnover number with a maximum of 1462. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2009 相似文献
2.
Polycarbonates (PCs) were prepared in a propylene carbonate solvent by the oxidative carbonylation of bisphenol A with Pd/bithienyl complexes, Pd/bipyridyl complexes, and Pd C σ-bonded complexes for comparison as homogeneous Pd catalysts. With the Pd/bipyridyl complexes, the 6,6′-disubstituted 2,2′-bipyridyl ligand showed a stronger substituent effect than the 2,2′-bipyridyl ligand, which lacked substituents at the 6,6′ positions. With the Pd/bithienyl complexes, however, the substituent effect was not seen. The Pd/bithienyl complexes, which lacked substituents at the 5,5′ positions, gave a PC yield that was the same as the yield of those that had substituents at the 5,5′ positions. The combination of the Pd C σ-bonded complexes and an inorganoredox cocatalyst showed a PC polymerization behavior that was different from the other two types of complexes. When Co(OAc)2·4H2O was used as the inorganoredox cocatalyst, all of the Pd C σ-bonded complexes gave a good PC yield. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008 相似文献
3.
Halogen‐free solvents for oxidative carbonylation of bisphenol A to polycarbonate using carbon monoxide were investigated. Tetrahydorofuran, γ‐butyrolacton, and acetophenone, which dissolve both bisphenol A and polycarbonate, gave a polycarbonate yield that was lower than using a halogenated solvent of dichloromethane. On the other hand, propylene carbonate that has a carbonate bond gave a polycarbonate yield the same as using dichloromethane. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci 2007 相似文献
4.
Bisphenol‐A polycarbonate (BAPC) was synthesized by solid‐state polymerization (SSP) using a semicrystalline prepolymer crystallized by antisolvent method. The antisolvent crystallization was investigated as a function of antisolvent types using X‐ray diffraction (XRD), different scanning calorimetry (DSC), and scanning electron microscopes (SEM). The results showed antisolvent types had a significant effect on the crystallization of BAPC. Prepolymer induced by acetone as an antisolvent gained a higher crystallinity of 37.0%, more uniform particle size, and mature crystal structure compared with the samples crystallized by methanol and ethanol. Then crystallization of BAPC by acetone was carried out at crystallization temperature in the range of 40–80 °C for 1–5 h. A high crystallinity of 42.0% was acquired with the crystallization conducted at 70 °C for 2 h. Prepolymer with appropriate crystallinity of 37.8% resulted in high‐molecular‐weight polymer of 57,411 via SSP due to the effect of crystallinity and plasticization of residual solvent. © 2016 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2016 , 133, 43636. 相似文献
5.
A series of linear aliphatic polycarbonate polyols were synthesized using dimethylcarbonate and a linear alkane diol or specific combinations of linear alkane diols. Polyol synthesis was carried out in a two‐stage process using dimethylcarbonate and a linear alkane diol to prepare a series of homopolymer polycarbonate polyols. Polyol grades were characterized using Fourier transform infrared spectroscopy, gel permeation chromatography, and differential scanning calorimetery techniques. Suitable reaction conditions were developed to yield polycarbonate polyols of number average molecular weight between 700 and 1700. The crystallinity of the polycarbonate polyols was shown to reduce as the molecular weight of the alkane diols used in the polycarbonate synthesis was increased. These polymers offer the potential for use in the synthesis of ether free polyurethane elastomers for biomedical applications. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2009 相似文献
6.
This article describes a new, nonphosgene method for the synthesis of poly(bisphenol A carbonate) (PC). The method involves three steps: the reaction of an aliphatic diol with phenyl chloroformate to form an alkylene diphenyl dicarbonate, the reaction of the alkylene diphenyl dicarbonate with bisphenol A to produce an aromatic–aliphatic polycarbonate, and the thermal treatment of the polycarbonate at 180–210°C under a stream of nitrogen with Ti(OBu)4 to give PC and a cyclic alkylene carbonate. The method furnished low to moderate molecular masses of PC upon the complete elimination of the aliphatic moieties. The approach may be considered a new method, based on polycarbonate thermochemical degradation, for the synthesis of cyclic aliphatic carbonates. The obtained polymers were characterized by intrinsic viscosity and IR, 1H‐NMR, and 13C‐NMR spectroscopy. The thermal treatment step was conducted in a glass reaction tube at 180–210°C under a stream of nitrogen, and the reaction was completed by heating to 250°C. In the thermal treatment step, semisolid effluents composed of cyclic alkylene carbonates were formed and subsequently eliminated from the reaction mixture. Heating to 250°C under nitrogen or under a dynamic vacuum furnished the pure aromatic PC residue. This intrachange reaction provides a flexible method for the synthesis of polycarbonates with alkylene diols containing two or three methylene groups, from which the pure PC homopolymer can be prepared. The potential of this approach was demonstrated by the successful synthesis of PC homopolymer from five different polycarbonates with a bisphenol A unit linked to 1,2‐propylene, 1,3‐propylene, 2‐methyl‐1,3‐propylene, 2,2‐dimethyl‐1,3‐propylene, and 1,3‐butylene as the alkane chains. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008 相似文献
7.
Heterogeneous catalytic synthesis of poly(butylene succinate) by attapulgite‐supported Sn catalyst 下载免费PDF全文
An attapulgite‐supported Sn catalyst (Sn‐palgorskite) was successfully prepared by ion‐exchange co‐impregnation method. The catalytic effect of Sn‐attapulgite on the melt polycondensation reaction of poly(butylene succinate) (PBS) was examined by comparing with that of anhydrous SnCl2 and a blank experiment. The structures and properties of the PBS products synthesized with the three catalytic systems were characterized by means of Fourier transform‐infrared spectra (FT‐IR), viscosity measurements, wide angle X‐ray diffraction (WARD), differential scanning calorimetry (DSC), and thermogravimetric analysis (TGA). Meanwhile, the acidolysis degradation behaviors of the PBS samples were also studied. It was found that the intrinsic viscosities and the number‐average molecular weights of the PBS were remarkably increased under the catalysis of Sn‐attapulgite, further leading to an increase in the decomposition temperature, the crystallization temperature, and the viscous flow activation energy, while a decrease in the melting temperature, the relative degree of crystallinity, and the rate of degradation. Additionally, no significant differences were observed in the crystal structures of all the samples, regardless of the reaction system with or without catalyst. The experiment results confirmed that Sn‐attapulgite was an effective heterogeneous catalyst for the synthesis of PBS, and the recycled Sn‐attapulgite still exhibited higher activity than anhydrous SnCl2 under identical reaction conditions. © 2014 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2015 , 132, 41729. 相似文献
8.
Phase inversion is a very flexible technique to obtain membranes with a large sort of morphologies. Membrane properties can vary greatly depending on the kind of polymer system used. Bisphenol A polycarbonate (PC) could be used as a phase inversion membrane base polymer, and presents very good properties. Nevertheless, very little information on membrane preparation using PC and the phase inversion process can be found in the literature. In this work flat‐sheet microporous membranes were obtained by the phase inversion process using the immersion precipitation technique. A new polymer system was studied, consisting of polycarbonate, N‐methyl‐2‐pyrrolidone as solvent, water as the nonsolvent, and an additive. The influence of some parameters on membrane morphology, such as polymer solution composition, exposition time before immersion into the precipitation bath, and the kind of additive was investigated. Precipitation was followed using light transmission experiments and membrane morphology was observed through Scanning Electron Microscopy (SEM). The viscosity and cloud points of all polymer solutions were also determined. The results were related to the studied synthesis parameters, using the basic principles of membrane formation by the phase inversion technique, looking forward to establishing criteria to control the morphology of flat‐sheet membranes using polycarbonate as the base polymer. The results showed that both additives were able to increase pore interconnectivity and even suppress macrovoid formation. The decrease in the miscibility region of the polymer system and increase in mass transfer resistance are found to be the determining factors during polymer solution precipitation. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 86: 3085–3096, 2002 相似文献
9.
Structure and properties of ultra‐high molecular weight bisphenol a polycarbonate synthesized by solid‐state polymerization in amorphous microlayers 下载免费PDF全文
The structure and properties of ultrahigh molecular weight polycarbonate synthesized by solid‐state polymerization in micro‐layers (SSPm) are reported. A low molecular weight prepolymer derived from the melt transesterification of bisphenol A and diphenyl carbonate as a starting material was polymerized to highly amorphous and transparent polycarbonate of molecular weight larger than 300,000 g mol?1 in the micro‐layers of thickness from 50 nm to 20 µm. It was observed that when the polymerization time in micro‐layers was extended beyond conventional reaction time, insoluble polymer fraction increased up to 95%. Through the analysis of both soluble and insoluble polymer fractions of the high molecular weight polycarbonate by 1H NMR spectroscopy and pyrolysis‐gas chromatography mass spectrometry (Py‐GC/MS), branches and partially crosslinked structures have been identified. The thermal, mechanical and rheological properties of the ultra‐high molecular weight nonlinear polycarbonates synthesized in this study have been measured by differential scanning calorimetry (DSC), dynamic mechanical analysis (DMA), and rheometry. The nonlinear chain structures of the polymer have been found to affect the polymer's thermal stability, mechanical strength, shear thinning effect, and elastic properties. © 2014 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2015 , 132, 41609. 相似文献
10.
Small‐molecule‐induced miscibility of isosorbide‐based polycarbonate with bisphenol A polycarbonate 下载免费PDF全文
In this study, we investigated the influence of the small molecule 4,4′‐thiobis(6‐tert‐butyl‐m‐methyl phenol) (AO300) on the miscibility of poly(isosorbide‐co‐1,4‐cyclohexanedimethanol carbonate) (IcC–PC) with bisphenol A polycarbonate (BPA–PC) through the formation of hydrogen‐bonding networks. Differential scanning calorimetry and morphological observation revealed that the initially, immiscible BPA–PC/IcC–PC blends become miscible through the addition of small molecules. Fourier transform infrared spectroscopy confirmed that intermolecular hydrogen bonds formed between the hydroxyl groups of AO300 and the carbonyl groups of the studied polycarbonates. These polycarbonates exhibited different hydrogen‐bonding behaviors and various degrees of glass‐transition temperature composition dependence. Dynamic mechanical analysis demonstrated that AO300 played an antiplasticization role in the BPA–PC/IcC–PC blends with improved storage moduli. To our knowledge, this article is the first to describe the miscibility of isosorbide‐based polycarbonate with BPA–PC. © 2016 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2017 , 134, 44537. 相似文献
11.
Ce在负载Pd催化苯酚氧化羰基化合成碳酸二苯酯反应中的作用 总被引:2,自引:0,他引:2
采用微乳液法制备了Ce为助剂的Pd-Ce-O/SiO2催化剂,用于苯酚氧化羰基化合成碳酸二苯酯(DPC)反应。活性评价结果显示,催化剂性能随着Ce用量的增加而提高,当Ce/Pd摩尔比为10/1时,苯酚转化率为64.4%,碳酸二苯酯选择性为83.4%。利用XRD表征发现,部分Ce4+进入到PdO晶格中,使得失活Pd原子中的电子更容易向Ce转移,从而易于再生,表现出更好的催化性能。根据上述结果,设计制备了Pd-O/CeO2催化剂用于本反应,苯酚转化率和碳酸二苯酯选择性分别仅为24.0%和23.3%。表征发现,在Pd-Ce-O/SiO2催化剂表面,Pd物种主要是PdO,而Pd-O/CeO2表面的Pd物种则以PdO2为主。由于苯酚氧化羰基化反应的活性中心为Pd(Ⅱ),所以Pd-O/CeO2催化性能较差。并且,由于Pd与CeO2之间存在强相互作用,催化剂表面Pd含量较低,这也是Pd-O/CeO2催化活性较差的原因之一。 相似文献
12.
欧阳华勇 《化学工业与工程技术》2009,30(4):8-11
考察了Ni—Pd/C双金属催化剂上乙醇气相羰基化合成丙酸的反应。实验结果表明,活性炭负载Ni—Pd双金属催化剂相比普通非贵金属催化剂具有更高的活性及选择性,优化的反应工艺条件为压力0.15MPa,温度220℃,nCO:nEtOH,n:2.0:1,乙醇液体空速1.5h^-1。在上述条件和少量碘乙烷存在下,乙醇羰基化反应产物中丙酸及丙酸酯的选择性可高达96.7%,通过GC—MS对反应产物分析,分析了反应过程的机理。 相似文献
13.
酯交换缩聚法合成聚碳酸酯的研究进展 总被引:2,自引:0,他引:2
聚碳酸酯的酯交换缩聚工艺具有绿色环保的特性,符合当今世界可持续发展的主题,相对于传统的光气法工艺更具有发展前途。综述了国内外在酯交换缩聚工艺的机理以及催化剂选择和工艺条件上取得的进展。指出碳酸二苯酯与双酚A酯交换反应是四面体机理,正反应是二级反应,逆反应是三级反应;以La(ACAC)_3作催化剂,得到的产品不但粘均分子量较高,而且热稳定性非常好;利用酯交换熔融聚合工艺可以解决传统光气法的环境问题,而固相聚合可以得到超高分子量的产品。 相似文献
14.
This article deals with the development of kinetic parameters for bisphenol acetophenone–polycarbonate made by melt transesterification with diphenyl carbonate. The understanding of the influence of borosilicate glass of the reactor construction materials on the accuracy of the kinetic data is reported. During the development of analytical methods, the use of high performance liquid chromatography‐mass spectrometry (HPLC‐MS) was proven to be a valid tool to determine the oligomers existing in the reaction mixture. Accurate kinetics parameters were obtained by elimination of the interference of the construction materials. We provide the rate expressions, kinetic parameters [forward reaction frequency factor = 2.456 × 1013 ± 0.01 (cm3/mol)2/min, forward reaction activation energy = 45.69 ± 0.2 kJ/mol, reverse reaction frequency factor = 2.068 × 1014 ± 0.01 (cm3/mol)2/min, and reverse reaction activation energy = 56.37 ± 0.1 kJ/mol], and equilibrium constants at various temperatures. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 103: 4072–4079, 2007 相似文献
15.
Jong Hun Kim 《应用聚合物科学杂志》2009,111(2):883-889
A novel crystallization method for the production of high‐molecular‐weight bisphenol A polycarbonate by solid‐state polymerization is suggested. In this method, a low‐molecular‐weight polycarbonate prepolymer is dissolved in a solvent and then partially crystallized with a novel spray‐crystallizing method to prepare crystallized polycarbonate particles having a very uniform and porous structure with a narrow melting region. As a result, during solid‐state polymerization, the phenol byproduct can be easily removed from the polymerizing porous polycarbonate particles, and the polymerization rate is dramatically increased. In particular, the effects of the crystallization methods on secondary crystallization during solid‐state polymerization and the melting behavior have been investigated with differential scanning calorimetry studies. The final product, a high‐molecular‐weight polycarbonate, displays a very narrow molecular weight distribution and uniform physical properties. A simultaneous process and an adequate reactor design for spray crystallization and solid‐state polymerization are also suggested. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2009 相似文献
16.
The photo‐activated, acid catalyzed decomposition of polycarbonate was investigated in this study. The impact of the chemical and physical properties of the photoacid generators (PAG) and the ambient atmosphere effect on polycarbonate decomposition were discussed. The photo‐patterns resulted from the photoacid catalyzed decomposition of a polycarbonate can be used as a sacrificial placeholder for fabrication of microelectromechanical and microfluidic devices. The effects of acid strength, vapor pressure of the acid, PAG activation process, and ambient conditions (temperature, moisture, and oxygen concentrations) on polymer film decomposition were studied. Several superacids (e.g. triflic and nonaflic based PAGs) were not suitable for decomposition of the polycarbonate because of their high vapor pressures resulting in the high volatility properties. From the decomposition experiments it was found that the nonfluorinated sulfonic acid based PAGs do not posses the superacidity needed for decomposition. Perfluorinated methide and a tetrakis (pentafluoropheyl)borate PAG were effective in the decomposition of polycarbonate films. The combination of two PAGs, one which generates high vapor pressure acid (thus, highly volatile) and the other with a lower vapor pressure acid (thus, less volatile) showed very low residue levels. This is because of the volatility of the generated high vapor pressure acid (usually remaining acid in the film was the cause of the residue left behind) and the remaining nonvolatile low vapor pressure acid was sufficient to decompose the polycarbonate that was not decomposed by the generated high volatile acid. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci, 2007 相似文献
17.
Heterogeneous cocatalysts, catalysts, and catalyst mixtures for the polymerization of ethylene were prepared applying “fumed silica” and mesoporous MCM-41 support materials and zirconocene dichloride, titanocene dichloride, and a bis(arylimino)pyridine iron complex as catalyst precursors. The catalyst mixtures produced polyethylenes which exhibit the properties of two single polymers. Polyethylenes with the desired bimodal molecular weight distributions could be obtained with a series of ternary Zr/Ti/Fe catalysts. The ability of the zirconium and titanium species to copolymerize short-chain 1-olefins produced by the iron centers (“in situ” copolymerization) is useful for the production of copolymers from only one monomer (ethylene). 相似文献
18.
A study on lactic acid polycondensation under melt conditions was carried out and a preliminary assessment revealed tin powder as a very good catalyst for poly(lactic acid) (PLA) synthesis by melt polycondensation while confirming previous information on SnCl2 good performance. However, these catalysts also promoted side reactions leading to racemization and yellowing of the final polymer. The use of p‐toluenesulphonic acid (p‐TSA) or triphenylphosphine (PPh3) as co‐catalysts proved to be very effective hindering colour formation and allowing synthesizing PLA samples with enhanced properties. The addition of these compounds to neat tin powder increased the PLA optical purity, whereas their addition to SnCl2 speeded up the polymerization. A significant increase in molecular weight, from 32,500 to 52,000 g mol?1, was recorded, with the new catalytic system SnCl2/PPh3 showing catalytic activity comparable with the one reported in the literature for SnCl2/p‐TSA. Several characterization techniques were used for assessing polymer samples: the molecular weights were determined by SEC, thermal behavior measured by DSC, and racemization extent calculated from specific rotation measurements. UV/vis spectroscopy was confirmed as a powerful technique for evaluating yellowing of final polymers. © 2012 Wiley Periodicals, Inc. J. Appl. Polym. Sci., 2013 相似文献
19.
The microstructure of blends of bisphenol A polycarbonate (PC), and poly(ethylene terepthalate) (PETP) has been studied by solvent extraction, infrared spectrophotometry, differential scanning calorimetry and dynamic mechanical thermal analysis. The blends appear to contain two amorphous phases over the whole composition range. The tensile behaviour and the Charpy impact strength of some of the blends have been determined, before and after heat treatment at 125°C for 18 hours. Improved performance of the blends, compared with that of the homopolymers PC and PETP, has been demonstrated. 相似文献
20.
A novel nanocomposite catalyst was prepared from immobilization of aluminum oxide hydroxide onto the attapulgite. Characterizations with scanning electron microscopy (SEM) and wide angle X‐ray diffraction (XRD) of the as‐prepared catalyst revealed that AlO(OH) nanoparticles were distributed on the attapulgite. Thermogravimetric analysis‐infrared spectrometry (TGA‐IR) of the mixture prepared by mixing of bishydroxy ethylene terephthalate (BHET) and the catalyst indicated that attapulgite‐supported aluminum oxide hydroxide catalyst can catalyze BHET polycondensation under the applied conditions. A kinetic model for determining the activation energy has been applied to evaluate the catalyst activity. The catalyst activity was examined through comparative experiments, and the results showed that the new catalyst exhibited higher activity for BHET polycondensation under identical reaction conditions, and the viscosity‐average molecular weight of poly(ethylene terephthalate) (PET) product obtained was increased about 2000 g/mol. © 2013 Wiley Periodicals, Inc. J. Appl. Polym. Sci., 2013 相似文献