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 171 毫秒
1.
丁齐  邢晓东  李丽霞 《化工进展》2014,33(4):971-976
以N-异丙基丙烯酰胺(NIPAM)为温敏单体,引入线性聚甲基丙烯酸-β-羟乙酯(PHEMA)或乙二胺胺化的聚甲基丙烯酸-β-羟乙酯(PAEMA),制备多孔半互穿水凝胶,并引入马来酰亚胺基团修饰,制备了两种新型水凝胶酶固定化载体Ⅰ和Ⅱ,凝胶的平均孔洞大小均约在10 ?m以上。以Ⅰ和Ⅱ为载体,马来酰亚胺与巯基的点击反应为基础,进行脂肪酶与糖化酶的固定化研究。结果表明,通过点击反应固定化得到的固定化脂肪酶最高活力回收率达6.03%,是相同条件下戊二醛固定化的5倍左右,且稳定性较高。  相似文献   

2.
本论文采用一步法,以季戊四醇作为B4核单体、2,2-二羟甲基丙酸作AB2型单体合成了超支化聚酯内核(HBPE),并对产物进行表征,计算超支化聚酯的支化度为0.41;探讨并建立了超支化聚酯的缩聚反应动力学方程-d[COOH]/dt=K[COOH][OH],该缩聚反应属于二级反应,反应速率由羧基浓度和羟基浓度共同决定。采用月桂酸改性超支化聚酯合成了新的长链烷基化的超支化聚合物(LHBP),并用红外和核磁共振谱图进行结构表征。溶解性质表明月桂酸改性超支化聚酯溶解于极性较弱的溶剂中,较未改性的超支化聚酯具有更好的有机溶剂溶解性。采用改性后的超支化聚酯内核(LHBP)和2-羟乙胺基蒽醌,通过IPDI偶联法合成了超支化聚合染料(R-LHBP)。  相似文献   

3.
酯端基超支化聚(胺-酯)的合成与表征   总被引:1,自引:0,他引:1  
赵辉  罗运军  杨树  孙瑞敏 《化学世界》2007,48(10):629-632
以乙醇胺和丙烯酸甲酯为原料,通过Michael加成反应合成了N-羟乙基-3-胺基-N,N-二丙酸甲酯AB2单体,并采用FT IR1、H NMR和元素分析对单体的结构进行了表征;以丁二酸酐为反应中心核,N-羟乙基-3-胺基-N,N-二丙酸甲酯为单体,制得了1~5代酯端基超支化聚(胺-酯)。采用核磁共振法对1~5代酯端基超支化聚(胺-酯)的支化度进行了表征,结果表明,由于采取的是加入中心核并逐步加入单体的有核准一步缩聚法,所合成的超支化聚(胺-酯)具有较高的支化度;凝胶渗透色谱测定的结果表明,超支化聚(胺-酯)的相对分子质量分布较窄,具有单分散性;超支化聚(胺-酯)的粘度较低且在多种溶剂中均有较好的溶解性,分解温度均高于170°C,具有良好的热稳定性。  相似文献   

4.
通过n(二乙醇胺):n(甲基丙烯酸甲酯)=1:1.05进行Michael加成反应,制得N,N'一二羟乙基-3-氨基甲基丙酸甲酯(MMB).以季戊四醇为核,MMB为支化单体,采用准一步法合成第2代端羟基的超支化聚合物(PM-2).通过甲苯-2,4一二异氰酸酯(TDI)分别与甲基丙烯酸-β-羟乙酯(HEMA)、2-羟基-2-甲基-1-苯基丙酮(HMPP)反应,制得含双键的功能单体TDI-HEMA和含光引发基团的功能单体TDI-HMPP,通过这两种单体对PM-2进行端基改性反应,得到超支化UV自引发聚合物(PM-UV).最后通过调节PM-UV中双键和光引发基团的物质的量之比制得各种PM-UV,并研究各种PM-UV的涂膜性能.结果表明改性单体物质的量之比不同的PM-UV均具有耐热性好、凝胶含量高、成膜硬度好、耐冲击性高等优点.  相似文献   

5.
肖舒  曹辉波  戴林  申越  何静 《精细化工》2012,29(7):712-716
在离子液体介质中利用原子转移自由基聚合方法合成了纤维素-甲基丙烯酸羟乙酯的接枝共聚物,采用甲基丙烯酸羟乙酯单体合成纤维素聚合物,并对聚合物进行结构测定与性质分析。通过FTIR、1HNMR和GPC对聚合物结构及相对分子质量(简称分子量,下同)进行分析,研究了不同溶剂对聚合条件、接枝效率的影响,以及在不同溶剂中聚合物的结构、分子量分布、热稳定性及形态有何变化。结果表明,在溶剂N,N-二甲基甲酰胺(DMF)中进行的反应聚合速率及终止速率均高于在丁酮中进行的反应。纤维素接枝甲基丙烯酸羟乙酯聚合物分子量分布约为1.81,该聚合物热稳定性较强;在DMF中呈球状颗粒,平均直径50 nm;在选择性溶剂丙酮中颗粒直径100 nm左右。  相似文献   

6.
超支化聚合物的研究及应用   总被引:1,自引:0,他引:1  
以二异丙醇胺(DIPA)及酸酐为原料合成了含有一个羧基、两个羟基的AB2型单体,采用AB2型单体自缩聚的方法合成了超支化聚酰胺酯,然后用甲基丙烯酸对超支化聚酰胺酯的端基功能改性。然后将超支化单体与马来酸酐、甲基丙烯酸共聚合成阻垢分散剂。  相似文献   

7.
以丁二酸酐和二乙醇胺为起始原料经加成反应合成新型AB2型单体,在溶液中采用“准 一步法”自缩聚反应合成超支化聚(酰胺-酯),使用傅里叶变换红外光谱、乌式粘度计、差热-热重联 用分析仪等测试分析超支化酰胺类聚合物的结构、粘度及热性能,红外光谱图分析进一步证实了聚合 物的超支化结构。实验表明封端剂对超支化聚合物的粘度有较大影响,封端产品粘度在0.019- 0.057 dL/g范围之间,而未封端产品粘度为0.146 dL/g。热分析中,超支化聚(酰胺-酯)的Td(热分 解温度)为316.3℃,Tg(玻璃态转化温度)为270℃;结果表明,超支化聚(酰胺-酯)具有较低的比浓 对数粘度、良好的溶解性和热稳定性等特点。  相似文献   

8.
马来酸酯烯类支化单体合成及其与甲基丙烯酸的共聚   总被引:1,自引:1,他引:0  
采用三羟甲基丙烷(TMP)和马来酸酐(MA)制备多官能度烯类支化单体单三羟甲基丙烷三马来酸单酯(MTPTM)。MTPTM与甲基丙烯酸(MAA)在水溶液中通过自由基聚合,制备超支化聚合物。采用NMR表征了MTPTM和超支化聚合物的结构。考察了MAA和MTPTM单体摩尔比对超支化聚合物相对分子质量、水溶液黏度、热稳定性以及玻璃化转变温度的影响。结果表明,随MAA单体用量增加,聚合物相对分子质量及其分布先降后升,n(MTPTM)∶n(MAA)=1∶9时出现最小值(Mw=2.09×104,Mw/Mn=1.66),产物水溶液黏度也呈现先降后升的趋势,并在n(MTPTM)∶n(MAA)=1∶6时出现最小值。而玻璃化转变温度则先升后降,n(MTPTM)∶n(MAA)=1∶9时出现最大值(Tg=274.5℃),MAA单体比例增加,有助于超支化聚合物热稳定性提高。  相似文献   

9.
介绍了原子转移自由基聚合(ATRP)制备超支化聚合物的原理以及近年来采用ATRP方法制备的各种支化/超支化聚合物,展望了ATRP的发展趋势.ATRP是目前可控,活性聚合最成功的方法之一,它以过渡金属配合物为催化剂,通过有机卤化物引发乙烯基单体的自由基聚合,合成相对分子质量可控、相对分子质量分布窄的多种聚合物.  相似文献   

10.
采用三羟甲基丙烷三马来酸单酯(TMPTM)与甲基丙烯酸(MAA)在水中进行自由基共聚,合成线形-超支化嵌均聚合物(LHBPs)。单体竞聚率数值表明TMPTM与MAA之间为典型的嵌均共聚,NMR、 lg[η]-lgMw关系规律以及支化因子g’研究结果表明,当单体摩尔比n (MAA)/n(MTPTM)小于48时,产物显示超支化聚合物特征,当n (MAA)/n(MTPTM)≥48时,聚合物呈现线形聚合物特征。DSC分析表明,LHBP-3聚合物玻璃化转变温度约为209.98℃,TGA表明在50-600℃升温过程中,LHBP聚合物有两次分解,分解温度分别为206.38℃(脱羧)和429.57℃(骨架分解)。  相似文献   

11.
A versatile method is described to synthesize a new family of solvent‐responsive membranes whose response states can be not only tunable but also fixable via ultraviolet (UV) irradiation induced crosslinking. The atom transfer radical polymerization (ATRP) initiator 2‐bromoisobutyryl bromide was first immobilized on the poly(ethylene terephthalate) (PET) track‐etched membrane followed by room‐temperature ATRP grafting of poly(2‐hydroxyethyl methacrylate) (PHEMA) and poly(2‐hydroxyethyl methacrylate‐co‐2‐(dimethylamino)ethyl methacrylate) (P(HEMA‐co‐DMAEMA)) respectively. The hydroxyl groups of PHEMA were further reacted with cinnamoyl chloride (a photosensitive monomer) to obtain photo‐crosslinkable PET‐g‐PHEMA/CA membrane and PET‐g‐P(HEMA/CA‐co‐DMAEMA) membrane. The length of grafted polymer chains was controllable by varying the polymerization time. X‐ray photoelectron spectroscopy, Fourier transform infrared spectroscopy in attenuated total reflection and thermogravimetric analysis were employed to characterize the resulting membranes. The various membrane surface morphologies resulting from different states of the grafted chains in water and dimethylformamide were characterized by scanning electron microscopy. It was demonstrated that the grafted P(HEMA/CA‐co‐DMAEMA) chains had more pronounced solvent responsivity than the grafted PHEMA/CA chains. The surface morphologies of the grafted membranes could be adjusted using different solvents and fixed by UV irradiation crosslinking. © 2014 Society of Chemical Industry  相似文献   

12.
In this work, interpenetrating polymer networks (IPNs) of polydimethylsiloxane (PDMS) and poly(acrylic acid) or poly(2‐hydroxyethyl methacrylate) (PHEMA) have been synthesized employing a sequential method. Monomeric AAc or HEMA was introduced into the PDMS network by swelling the polymer in solutions of monomer. The polymerization of monomers was then conducted in the swollen network. The swelling properties of the IPNs were investigated by varying the monomer concentrations in the polymerization and more swelling was observed with low monomer concentrations due to the prevalence of cyclization reactions. Multi‐step polymerization used to achieve IPNs with high hydrogel contents, did not improve their water uptake. The kinetics of acrylic acid polymerization was studied under various conditions. Specifically, in the presence of confinement effects imposed by the PDMS network a considerable drop in the rate of reaction was observed. The cross‐linking density of the PDMS network was also studied how to affect the reaction rate. © 2011 Wiley Periodicals, Inc. J Appl Polym Sci, 2012  相似文献   

13.
Core–shell hydrogel latexes, composed of a poly(2‐hydroxyethyl methacrylate) (PHEMA) core chemically coated with chitosan (CS) shell, were synthesized via an emulsifier‐free emulsion polymerization, free radically initiated by a redox couple of tert‐butyl hydroperoxide and amine groups on CS itself. The variation of some polymerization parameters [e.g., polymerization time, CS/2‐hydroxyethyl methacrylate (HEMA) weight ratio, and content of crosslinker] was systematically investigated in this study. We found that the weight ratios between CS and the HEMA monomer influenced the course of polymerization, which was traced by the change in percentage monomer conversions, and the colloidal stability of the PHEMA–CS hydrogel latexes obtained. Moreover, the polymerization time affected their particle sizes and surface charges. For the colloidally stable PHEMA–CS hydrogel latexes, their sizes and charges ranged from 600 to 689 nm and from 32 to 51 mV, respectively. N,N′‐Methylene bisacrylamide was used as a crosslinking agent for the core component; this was found to be able to enhance the hydrogels' thermal stability and water uptake. Moreover, the 3‐[4,5‐dimethylthiazol‐2‐yl]‐2,5‐diphenyltetrazolium bromide assay showed that 100% cell viability was achieved during the treatment of the PHEMA–CS latex (0.2–2.5 mg/mL) with Caco‐2 cells. © 2013 Wiley Periodicals, Inc. J. Appl. Polym. Sci., 2014 , 131, 40003.  相似文献   

14.
A series of microemulsions have been formulated, with 2-hydroxyethyl methacrylate (HEMA) or HEMA/water/propanol mixtures as the continuous phase and methylcyclohexane as the discontinuous phase. The effect of surfactant type was investigated with the utilization of both anionic and nonionic surfactants. The microemulsion continuous phase was polymerized by UV radiation and a thermal post-cure. The resultant polymers were extracted to remove the discontinuous phase and the surfactant. On swelling, the majority of the polymers became opaque, although transparent PHEMA hydrogels were synthesized with an improved equilibrium water content (EWC). The cause of opacity was shown by field emission scanning electron microscopy (FESEM). The breakdown in the microemulsion on polymerization is caused by unfavourable interactions between the PHEMA and the stabilizing surfactants causing agglomerization of the discontinuous phase. All the hydrogels were found to have higher water retention than PHEMA, with EWCs of up to 70%. The modified polymers also demonstrated an increased rate of water diffusion into the matrix. A preliminary study of oxygen permability revealed that a significant improvement had been made over standard PHEMA membranes. The porous structure of the PHEMA gels has been shown to be dependent on the type of surfactant used during synthesis.  相似文献   

15.
A new low-toxicity gelcasting system with a 2-hydroxyethyl methacrylate (HEMA) monomer was applied for casting of alumina ceramics. Polyvinyl pyrrolidone (PVP) was adopted for modifying the homogeneity of the PHEMA (poly-HEMA) gel. The rheological properties of alumina suspension in the HEMA–PVP premix solution were studied. After preparation of a concentrated alumina suspension, homogenous alumina green body with a relatively high strength of about 19 MPa could be formed through the PVP-modified HEMA system. Dense complex-shaped ceramic parts can be successfully produced through the system. Besides, the surface exfoliation phenomenon that seems inherent to the acrylamide gelcasting system could also be eliminated by using the PVP-modified HEMA system. Analysis of the interaction of HEMA and PVP suggested that the improved microstructure and strength homogeneity, as well as the elimination of surface exfoliation in the new system, should be attributed to the intermolecular hydrogen bonding formed between PHEMA and PVP molecules.  相似文献   

16.
The hydrogen‐terminated Si (100) (Si? H surface) was functionalized by coupling with 4‐vinylbenzyl chloride (VBC) to form a Si? VBC surface, which serves as macroinitiators for the surface‐initiated aqueous atom transfer radical polymerization (ATRP) of 2‐hydroxyethyl methacrylate (HEMA) and poly(ethylene glycol)methacrylate (PEGMA) to prepare Si? VBC? g? PHEMA and Si? VBC? g? PPEGMA substrates, respectively. The ellipsometric results revealed that the surface‐initiated ATRP of both PHEMA and PPEGMA brushes proceeded in a controlled fashion. By adjusting the monomer concentration, an eccentric polymer thickness dependence on the initial monomer concentration [M]0 was observed for both HEMA and PEGMA, i.e., in the dilute regime, the thickness of the polymer film increases with the increase in [M]0; however, beyond critical [M]0, the thickness deceases gradually with the further increase. Such an eccentricity was tentatively correlated to the counteractive combination of the increase in [M]0 and decrease in the apparent polymerization rate constant. Both Si? VBC? g? PHEMA and Si? VBC? g? PPEGMA substrates were esterified for the subsequent surface‐initiated ATRP, resulting in corresponding comb‐like brushes. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 102: 2590–2599, 2006  相似文献   

17.
Biocompatible polymers with specific shape and tailored hydrogel properties were obtained by polymerization of mixtures of 2‐hydroxyethyl methacrylate (HEMA) with 1–8 wt % ethylene glycol dimethacrylate (EGDMA) or tetra(ethylene glycol) diacrylate (TEGDA) as crosslinking agents, by using a redox initiator. Introduction of charged positive and negative groups was easily achieved by direct polymerization of appropriate monomer mixtures and by chemical transformation of preformed hydrogels. Investigation of the swelling behavior of the prepared hydrogels evidenced an appreciable dependence on both solvent type and polymer chemical structure. Additionally, the solvation process resulted in being controlled by solvent diffusion, according to a Fickian II mechanism. The presence of several types of water with different melting behavior was observed in fully swollen hydrogels. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 85: 2729–2741, 2002  相似文献   

18.
Narrow‐dispersion or monodisperse polymer microspheres with active hydroxyl groups were prepared by distillation–precipitation polymerization in the absence of any stabilizer. The monomer hydroxyethyl methacrylate (HEMA) was copolymerized with either commercial divinylbenzene (DVB) (containing 80 % of DVB isomers) or ethyleneglycol dimethacrylate (EGDMA) as crosslinker by distillation–precipitation polymerization technique with 2,2′‐azobisisobutyronitrile (AIBN) as initiator in neat acetonitrile. The effects of the crosslinker and the crosslinking degree on the morphology and the loading of the active hydroxyl group of the resultant microspheres were investigated. The agitation caused by distilling off a portion of the polymerization solvent during the polymerization avoided coagulation and resulted in the narrow‐dispersion or monodisperse polymer microspheres for the distillation precipitation technique. Copyright © 2005 Society of Chemical Industry  相似文献   

19.
A detailed kinetic Monte Carlo simulation was used to predict the characteristics of the batch miniemulsion polymerization of an isocyanate and an acrylic monomer mixture that contains a hydroxyl functional monomer (HEMA). The simulation takes into account the simultaneous polyaddition of the polyurethane prepolymer with the hydroxyl group of HEMA and the free radical polymerization of the acrylic monomers and all reactions in aqueous and polymer particle phases. The model has been assessed by batch miniemulsion polymerizations carried out using an aliphatic isocyanate prepolymer, n-butyl acrylate, 2-hydroxyethyl methacrylate monomers and potassium persulfate as an initiator. It was found that partitioning of water had a significant effect on both kinetics and microstructure of the resulting polymer. Evolution of different species of PU prepolymer produced in the reaction and the sol and gel fractions revealed that the terminal pendent double bond of the HEMA in polymer chains has significantly lower reactivity than that of the HEMA free monomer. Detailed information on gel microstructure has been derived in the model by both distribution of molecular weight between crosslinking points in acrylic chains and distribution of chain extension of PU prepolymers. These crosslinking density distributions can be related to mechanical and adhesive properties of the polymer.  相似文献   

20.
The thermal behavior of poly(2‐hydroxyethyl methacrylate) [PHEMA] homopolymer and poly(2‐hydroxyethyl methacrylate‐co‐itaconic acid) [P(HEMA/IA)] copolymeric networks synthesized using a radiation‐induced polymerization technique was investigated by differential scanning calorimetry, thermogravimetric analysis, and Fourier transform infrared spectroscopy. The glass‐transition temperature (Tg) of the PHEMA homopolymer was found to be 87°C. On the other hand, the Tg of the P(HEMA/IA) networks increased from 88°C to 117°C with an increasing amount of IA in the network system. The thermal degradation reaction mechanism of the P(HEMA/IA) networks was determined to be different from the PHEMA homopolymer, as confirmed by thermogravimetric analysis. It was observed that the initial thermal degradation temperature of these copolymeric networks increased from 271°C to 300°C with IA content. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 103: 1602–1607, 2007  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