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.
Chaozhan Wang  Sa Zhao  Yinmao Wei 《中国化学》2012,30(10):2473-2482
Poly(glycidylmethacrylate) (PGMA) brushes were grafted from chloromethylated polysulfone (CMPSF) membrane surface by surface‐initiated atom transfer radical polymerization (SI‐ATRP), and the grafting was followed by hydrolysis of epoxy groups in the grafting chains to improve the membrane's hydrophilic property. Fourier transform infrared spectroscopy (FT‐IR) and X‐ray photoelectron spectroscopy (XPS) measurements confirmed the successful grafting and hydrolysis of PGMA. The grafting degree of the monomer, measured by periodic acid titration and gravimetric analysis, increased linearly with the polymerization time, while the static water contact angle of the membrane grafted with PGMA or hydrolyzed PGMA linearly decreased. In comparison with the PGMA‐grafted membranes, the hydrolyzed PGMA‐grafted membranes possess stronger hydrophilicity as indicated by their contact angle and hydration capacity, and as a result they have an improved antifouling property. Therefore, the control of the hydrophilicity of PSF membrane could be realized through adjusting the polymerization time and transforming the functional groups in the grafting chain.  相似文献   

2.
Glycidyl‐functional polymer nanoparticles [poly(glycidyl methacrylate) (PGMA)] were fabricated with microemulsion polymerization. The successful fabrication of PGMA nanoparticles was confirmed by Fourier transform infrared spectroscopy and transmission electron microscopy (TEM). A TEM image showed that the average diameter of the PGMA nanoparticles was approximately 10–28 nm and was fairly monodisperse. As the surfactant concentration increased, the average size of the nanoparticles decreased and approached an asymptotic value. A significant reduction of the nanoparticle size to the nanometer scale led to an enhanced number of surface functionalities, which played an important role in the curing reaction. The PGMA nanoparticles were cured with a low‐temperature curing agent, diethylene triamine, to produce ultrafine thermoset nanoparticles. The low‐temperature curing process was performed below the glass‐transition temperature of PGMA to prevent the coagulation and deformation of the nanoparticles. A TEM image indicated that the cured PGMA nanoparticles did not exhibit interparticle aggregation and morphological transformation during curing. The average size of the cured PGMA nanoparticles was consistent with that of the pristine PGMA nanoparticles © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 2258–2265, 2005  相似文献   

3.
Poly(glyceryl methacrylate) (PGMA) was reacted with meso‐tetra(4‐hydroxylphenyl)porphyrin (THPP) in a homogeneous system via the ring opening reaction between epoxy and hydroxyl groups, which exist on the side chain of PGMA and the outside ring of THPP, respectively. Porphyrin‐functionalized PGMA with line‐type (denoted as HPP‐PGMA), on whose side chains hydroxylphenyl porphyrin (HPP) was bonded, was obtained. Grafting particles PGMA/SiO2 were also reacted with THPP, and the porphyrin‐immobilized particles (denoted as HPP‐PGMA/SiO2), on which HPP was supported, were produced. The above two target products were characterized using spectroscopy methods, such as infrared (IR), nuclear magnetic resonance (1H‐NMR), electronic absorption, and fluorescence emission. The effects of various factors on the bonding and immobilization reactions of HPP were studied in detail. The experimental results show that the soluble HPP‐PGMA has all the spectral characteristics of porphyrins and the absorption or emission intensity is increased with the increase in the bonding degree of HPP. In the preparation process of HPP‐PGMA, in order to avoid the occurrence of the crosslinking reaction and to obtain HPP‐PGMA with complete line‐type, the catalyst should be selected and the reaction time should be controlled. NaHCO3 is an appropriate catalyst. In the immobilization process of HPP on the grafting particle PGMA/SiO2, the greater the used amount of the catalyst triethylamine (TEA), the more rapid is the rate of ring opening reaction, resulting in higher immobilization amount of HPP. Besides, the immobilization amount of HPP is increased with the enhancement of the reaction temperature. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

4.
Poly(3‐hydroxybutyrate) (PHB)/poly(glycidyl methacrylate) (PGMA) blends were prepared by a solution‐precipitation procedure. The compatibility and thermal decomposition behavior of the PHB/PGMA blends was studied with differential scanning calorimetry, thermogravimetric analysis, and differential thermal analysis (DTA). The blends were immiscible in the as‐blended state, but for the blends with PGMA contents of 50 wt % or more, the compatibility was dramatically changed after 1 min of annealing at 200 °C. In addition, PHB/PGMA blends showed higher thermal stability, as measured by maximum decomposition temperatures and residual weight during thermal degradation. This was probably due to crosslinking reactions of the epoxide groups in the PGMA component with the carboxyl chain ends of PHB fragments during the degradation process, and the occurrence of such reactions can be assigned to the exothermic peaks in the DTA thermograms. © 2001 John Wiley & Sons, Inc. J Polym Sci Part A: Polym Chem 40: 351–358, 2002  相似文献   

5.
Controlled grafting of well-defined epoxide polymer brushes on the hydrogen-terminated Si(100) substrates (Si-H substrates) was carried out via the surface-initiated atom-transfer radical polymerization (ATRP) at room temperature. Thus, glycidyl methacrylate (GMA) polymer brushes were prepared by ATRP from the alpha-bromoester functionalized Si-H surface. Kinetic studies revealed a linear increase in GMA polymer (PGMA) film thickness with reaction time, indicating that chain growth from the surface was a controlled "living" process. The graft polymerization proceeded more rapidly in the dimethylformamide/water (DMF/H(2)O) mixed solvent medium than in DMF, leading to much thicker PGMA growth on the silicon surface in the former medium. The chemical composition of the GMA graft-polymerized silicon (Si-g-PGMA) surfaces were characterized by X-ray photoelectron spectroscopy (XPS). The fact that the epoxide functional groups of the grafted PGMA were preserved quantitatively was revealed in the reaction with ethylenediamine. The "living" character of the PGMA chain end was further ascertained by the subsequent growth of a poly(pentafluorostyrene) (PFS) block from the Si-g-PGMA surface, using the PGMA brushes as the macroinitiators.  相似文献   

6.
Amphiphilic block copolymers composed of a hydrophilic poly(ethylene glycol) (PEG) block and a hydrophobic poly(glycidyl methacrylate) (PGMA) block were synthesized through cationic ring‐opening polymerization with PEG as the precursor. The model reactions indicated that the reactivity of the epoxy groups was higher than that of the double bonds in the bifunctional monomer glycidyl methacrylate (GMA) under the cationic polymerization conditions. Through the control of the reaction time in the synthesis of block copolymer PEG‐b‐PGMA, a linear GMA block was obtained through the ring‐opening polymerization of epoxy groups, whereas the double bond in GMA remained unreacted. The results showed that the molecular weight of the PEG precursor had little influence on the grafting of GMA, and the PGMA blocks almost kept the same length, despite the difference of the PEG blocks. In addition, the PGMA blocks only consisted of several GMA units. The obtained amphiphilic PEG‐b‐PGMA block copolymers could form polymeric core–shell micelles by direct molecular self‐assembly in water. The crosslinking of the PGMA core of the PEG‐b‐PGMA micelles, induced by ultraviolet radiation and heat instead of crosslinking agents, greatly increased the stability of the micelles. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 2038–2047, 2005  相似文献   

7.
采用Adler法合成了Meso-四(对羟基苯基)卟啉(THPP), 在均相体系中使聚甲基丙烯酸缩水甘油酯(PGMA)与THPP发生开环反应, 得到侧链键合有羟基苯基卟啉的线型PGMA(HPP-PGMA); 进一步使HPP-PGMA与锰离子发生配合作用, 得到侧链键合有锰卟啉(MnP)的线型PGMA(MnP-PGMA), 测定了HPP-PGMA的1H NMR谱, 表征了其化学结构; 测定了HPP-PGMA与MnP-PGMA的UV-Vis光谱及荧光发射光谱, 考察了其光物理行为. 实验结果表明, 通过THPP外环上羟基与PGMA侧基环氧键的开环成醚反应, 可以顺利地将羟基苯基卟啉及其锰配合物键合在PGMA的侧链上. HPP-PGMA具有THPP的特征电子吸收光谱与荧光发射光谱, 且随着THPP键合度的增加, 光谱的强度增强. MnP-PGMA具有小分子锰卟啉的特征电子吸收光谱与荧光特征, 其Soret吸收带发生显著红移(红移58 nm), Q吸收带的数量减为3个吸收峰; 实验发现, MnP-PGMA与小分子锰卟啉类似, 在Q发射带没有荧光发射.  相似文献   

8.
Addition reactions of pendant epoxide groups in poly(glycidyl methacrylate) (PGMA) with various active esters such as 1-benzotriazolyl benzoate, S-(2-benzoxazolyl) thiobenzoate, S-(2-benzothiazolyl) thiobenzoate, 4-nitrophenyl benzoate (4NPB), and S-phenyl thiobenzoate (PTB) were carried out using quaternary salts as catalysts. The reactions of PGMA with those active esters proceeded in diglyme at 100°C for 24 h quantitatively without the formation of 2-hydroxyl pendant groups in the polymer when 10 mol % of tetraethylammonium bromide was used as a catalyst. Furthermore, it was found that the respective quaternary salts have higher catalytic activity than tertiary amines in the reaction of PGMA with the active esters, and the reaction of PGMA with 4NPB gave the corresponding polymer with the highest conversion by addition of tetrabutylammonium bromide as a catalyst, while tetraethylammonium chloride showed the highest activity for the reaction of PGMA with PTB. In addition, the rate of reaction of PGMA with 4NPB was proportional to third order kinetics of the epoxide concentration, the ester concentration and the catalyst concentration as follows: ?d[Epoxide]/dt = ?[Ester]/dt = k3[Epoxide] [Ester] [Catalyst].  相似文献   

9.
Macromolecular anchoring layer approach was used for preparation of an effective macroinitiator for the synthesis of grafted polymer layers by atom transfer radical polymerization (ATRP) initiated from the surface. For the initial surface modification, a thin layer of poly(glycidyl methacrylate) (PGMA) was deposited on the surface of a silicon wafer. The ATRP macroinitiator was synthesized on the substrate surface by the reaction between epoxy groups of PGMA and carboxy functionality of bromoacetic acid (BAA). Variation of the time and temperature of the BAA deposition as well as PGMA layer thickness allowed control over the amount of BAA attached to the surface. The PGMA anchoring layer allowed the achievement of initiator surface density significantly higher than that reported for a self-assembled monolayer of ATRP initiators. Polymer brushes were synthesized on the PGMA/BAA-modified substrates by ATRP. Different surface concentrations of BAA were used in our grafting experiments to acquire knowledge about the relationship between the amount of initiator anchored to the surface through PGMA and the rate of the grafted layer formation. The increase in the surface density of the initiating moieties led to the increase in the grafting rate. However, a cutoff initiator concentration beyond which no increase of the thickness of the grafted layer was observed. From comparison between the surface densities of the initiator and the attached polymer it was determined that the efficiency of the initiation from the surface was on the level of 5-15%.  相似文献   

10.
11.
Poly(glycidyl methacrylate) (PGMA) was synthesized by the RAFT method in the presence of 2‐cyanoprop‐2‐yl dithiobenzoate (CPDB) chain transfer agent using different [GMA]/[CPDB] molar ratios. The living radical polymerization resulted in controlled molecular weights and narrow polydispersity indices (PDI) of ≈1.1. The polymerization of pentafluorostyrene (PFS) with PGMA as the macro‐RAFT agent yielded narrow PDIs of ≤1.2 at 60 °C and ≤1.5 at 80 °C. The epoxy groups of the PGMA block were hydrolyzed to obtain novel amphiphilic copolymer, poly(glyceryl methacrylate)‐block‐poly(pentafluorostyrene) [PGMA(OH)‐b‐PPFS]. The PGMA epoxy group hydrolysis was confirmed by 1H NMR and FTIR spectroscopy. DSC investigation revealed that the PGMA‐b‐PPFS polymer was amorphous while the PGMA(OH)‐b‐PPFS displayed a high degree of crystallinity.

  相似文献   


12.
We have demonstrated the successful deposition of poly(glycidyl methacrylate) (PGMA) thin films using hot filament chemical vapor deposition (HFCVD) with tert-butyl peroxide as the initiator. The introduction of the initiator allows for film deposition at low filament temperatures (<200 degrees C) and greatly improves the film deposition rates. The retention of the pendant epoxide chemical functionality and the linear polymeric structure in the deposited films were confirmed by infrared spectroscopy and X-ray photoelectron spectroscopy. The number-average molecular weight of the PGMA films can be systematically varied from 16,000 to 33,000 by adjusting the filament temperature and flow ratio of the initiator to the precursor. The apparent activation energies observed from PGMA deposition kinetics (100.9+/-9.6 kJ/mol) and from molecular weight measurements (-54.8+/-2.0 kJ/mol) are close to the calculated overall activation energies for the polymerization rate (104.4 kJ/mol) and number-average molecular weight (-59.2 kJ/mol), which supports the hypothesis of the free radical polymerization mechanism in the HFCVD PGMA deposition.  相似文献   

13.
Poly(glycidyl methacrylate), PGMA, chains in linear and arborescent structures were incorporated onto surfaces of poly(tetrafluoroethylene), PTFE, films by hydrogen plasma and ozone treatment and atom transfer radical polymerization. The epoxide groups of the PGMA chains were further reacted with acetic acid (AAc), oxalic acid (XAc), allyl amine (AA), and ethylenediamine (EDN) to introduce hydroxyl and amine groups to the surfaces of the PTFE films. Surface characterizations performed by Fourier Transform infrared attenuated total reflectance (FTIR-ATR) spectroscopy and X-ray photoelectron spectroscopy (XPS) confirmed the surface modification and the chemical structure. The PGMA chains in arborescent structures show a high effectiveness for the enhancement of the adhesion of PTFE films. The adhesion of PTFE films was also significantly enhanced by ring-opening reactions of the PGMA epoxide groups with acetic acid and amine compounds. A high value of 9.5 N cm(-1) in the optimum 180 degrees peel strength test was observed with PTFE/copper assemblies.  相似文献   

14.
A series of five near-monodisperse sterically stabilized polystyrene (PS) latexes were synthesized using three well-defined poly(glycerol monomethacrylate) (PGMA) macromonomers with mean degrees of polymerization (DP) of 30, 50, or 70. The surface coverage and grafting density of the PGMA chains on the particle surface were determined using XPS and (1)H NMR spectroscopy, respectively. The wettability of individual latex particles adsorbed at the air-water and n-dodecane-water interfaces was studied using both the gel trapping technique and the film calliper method. The particle equilibrium contact angle at both interfaces is relatively insensitive to the mean DP of the PGMA stabilizer chains. For a fixed stabilizer DP of 30, particle contact angles were only weakly dependent on the particle size. The results are consistent with a model of compact hydrated layers of PGMA stabilizer chains at the particle surface over a wide range of grafting densities. Our approach could be utilized for studying the adsorption behavior of a broader range of sterically stabilized inorganic and polymeric particles of practical importance.  相似文献   

15.
Core‐shell structured barium titanate‐poly(glycidyl methacrylate) (BaTiO3‐PGMA) nanocomposites were prepared by surface‐initiated atom transfer radical polymerization of GMA from the surface of BaTiO3 nanoparticles. Fourier transform infrared spectroscopy confirmed the grafting of the PGMA shell on the surface of the BaTiO3 nanoparticles cores. Transmission Electron Microscopy results revealed that BaTiO3 nanoparticles are covered by thin brushes (~20 nm) of PGMA forming a core‐shell structure and thermogravimetric analysis results showed that the grafted BaTiO3‐PGMA nanoparticles consist of ~13.7% PGMA by weight. Upon incorporating these grafted nanoparticles into 20 μm‐thick films, the resultant BaTiO3‐PGMA nanocomposites have shown an improved dielectric constant (ε = 54), a high breakdown field strength (~3 MV/cm) and high‐energy storage density ~21.51 J/cm3. AC conductivity measurements were in good agreement with Jonscher's universal power law and low leakage current behavior was observed before the electrical breakdown field of the films. Improved dielectric and electrical properties of core‐shell structured BaTiO3‐PGMA nanocomposite were attributed to good nanoparticle dispersion and enhanced interfacial polarization. Furthermore, only the surface grafted BaTiO3 yielded homogenous films that were mechanically stable. © 2014 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2015 , 53, 719–728  相似文献   

16.
刘吉众  黄嫣嫣  杨博  常建华  刘国诠  赵睿 《色谱》2013,31(4):310-316
以具有双孔结构的聚甲基丙烯酸环氧丙酯(PGMA)微球为基质,以葡萄糖进行表面亲水改性,制备了强阳离子交换色谱填料,并将其用于复杂生命体系中生物大分子的快速而高效的分离、分析与纯化。葡萄糖亲水改性增进了填料的生物相容性,提高了蛋白质样品的回收率;双孔结构及较高的比表面积赋予填料良好的柱渗透性和样品负载量。以标准蛋白质为样品,考察了该填料对生物样品的分离性能。以100 mm×4.6 mm的色谱柱分离4种蛋白质,在6 min内实现了基线分离;以溶菌酶为样品,填料的吸附容量为39.5 g/L,在蛋白质快速分离纯化分析中显示了良好的应用前景。  相似文献   

17.
A multifunctional ferrocene‐modified poly(glycidyl methacrylate) (PGMA‐Fc) and a difunctional β‐cyclodextrin derivative (bis‐CD) has been prepared for the construction of an electrically driven removable and self‐healing polymeric materials based on the complexation reaction between ferrocene and β‐CD groups. The chemical structures of PGMA‐Fc and bis‐CD have been characterized with Fourier transform infrared, 1H nuclear magnetic resonance, and X‐ray photoelectron spectroscopy. The effects of electrical voltages and medium conductivity on the decrosslinking efficiency of the crosslinked PGMA‐Fc/CD polymer have been examined. The PGMA‐Fc/CD network has shown removable feature and properties for application as a reworkable crosslinked material. Moreover, the crosslinked PGMA‐Fc/CD sample has shown electrically driven self‐healing behavior. The self‐healing performance could be enhanced with wetting the sample to increase the electrical conductivity. As a result, the material could serve as a self‐healing agent for commercial painting products. Preparation and application of a novel and efficient self‐healing polymer have been demonstrated. © 2013 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2013 , 51, 3395–3403  相似文献   

18.
分散聚合制备粒度均匀的聚甲基丙烯酸环氧丙酯微球   总被引:13,自引:0,他引:13  
文中描述了粒度均匀的聚甲基丙烯酸环氧丙酯微球的制备,所采用的是分散聚合方法,系统地研究了溶剂体系、单体浓度、引发剂类型与浓度、稳定剂用量、反应温度等各种聚合参数,对聚合产物粒度及其分散性的影响.在优化反应条件的基础上,制备出了微米级(1~8μm)粒度均匀性基本呈现单分散的聚合物微球.  相似文献   

19.
Polymer with pendant cinnamic ester and chloromethyl groups was synthesized by the addition reaction of poly(glycidyl methacrylate–co–methyl methacrylate) (PGMA) with cinnamoyl chloride. Also, polymers with pendant benzoic esters and chloromethyl groups were synthesized by reaction of PGMA with the corresponding benzoyl chlorides. Furthermore, polymers with cinnamic or benzoic esters and alkylazide groups were prepared by the substitution reaction of the obtained polymers with sodium azide.  相似文献   

20.
通过可逆加成-断裂链转移(RAFT)聚合制备出聚甲基丙烯酸缩水甘油酯(PGMA),经过叠氮钠与PGMA环氧基团的反应引入叠氮基和羟基,然后依次通过端炔基聚乙二醇(PEG-alk)与叠氮基的点击反应,己内酯(CL)在羟基存在下的开环聚合反应,获得双亲支链梳形共聚物(PGMA-g-PEG/PCL)。利用该梳形共聚物的两亲性,在氯仿-水混合体系中,进行自乳化高效负载阿霉素(DOX),得到负载DOX的纳米粒子。利用核磁共振氢谱、红外光谱和凝胶渗透色谱确认了最终产物及其前体聚合物的结构。利用动态光散射、紫外可见分光光度计和扫描电镜研究该载药粒子在pH为7.0和5.0的水溶液中的释放。结果表明:该纳米粒子平均粒径约为100nm,该粒子能有效释放DOX,在酸性条件下释放速率加快,且伴随PCL的降解。  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