首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到19条相似文献,搜索用时 578 毫秒
1.
选用甲基丙烯酸异丙酯(iPMA)与N-异丙基丙烯酰胺(NIPAM)共聚,制备了一系列疏水改性、相转变温度可调的温敏性P(NIPAM-co-iPMA)微凝胶.利用透射电子显微技术(TEM)、浊度法、动态光散射(DLS)技术及示差扫描量热(DSC)技术对所得微凝胶的形态及去溶胀行为进行了表征.TEM与DLS结果表明,所制备的微凝胶具有规则的球型形态.浊度、DLS及DSC结果表明,疏水性单体iPMA的引入能有效调节共聚物微凝胶的相转变温度;在所考察的范围内,微凝胶的相转变温度随iPMA投料比的增加几乎呈线性降低.  相似文献   

2.
采用两步合成路线合成了二缩三乙二醇单甲基丙烯酸酯 (TREGMA) ,并对其结构进行了表征 ;利用无皂乳液聚合法使N 异丙基丙烯酰胺 (NIPAM)、TREGMA和N ,N 亚甲基双丙烯酰胺 (BA)交联共聚 ,制备了含有功能性羟基的温敏性微凝胶 .微凝胶的去溶胀性能研究表明 ,TREGMA的引入使得微凝胶的体积相转变温度得到提高 ,同时所制得微凝胶具有较好的温敏性 .该类微凝胶有望成为良好的生物医用材料 .  相似文献   

3.
利用预乳化乳液法制备了不同单体配比的聚(甲基丙烯酸甲酯-co-甲基丙烯酸-co-甲基丙烯酸羟乙酯)(P(MMA-co-MAA-co-HEMA))微凝胶分散液;采用透射电子显微镜、动态光散射仪研究了微凝胶的微观形态、粒径大小及其溶胀率;利用试管倒转法对微凝胶分散液的凝胶化相转变行为进行了研究,借助椎板流变仪考察了所形成胶态凝胶的储能模量与单体配比、微凝胶分散液浓度和温度的关系.结果表明,所制备的微凝胶的数均粒径为90 nm左右,当MMA与MAA的投料质量不变时,随着HEMA含量的增加,分散液凝胶化所需的临界最小浓度增大,临界最大pH值减小,胶态凝胶的储能模量增加.当保持单体MMA与HEMA的投料质量不变时,随着单体MAA投料质量的增多,微凝胶的数均粒径和溶胀率增大,胶态凝胶的储能模量先升高后降低;当MAA占单体总摩尔数的25%时,浓度为15 wt%的微凝胶分散液在扫描频率为100 rad/s时,胶态凝胶的储能模量最高可达2×104Pa.这类微凝胶分散液在组织工程支架材料方面有潜在的应用价值.  相似文献   

4.
温敏水凝胶的新进展   总被引:1,自引:0,他引:1  
从改变亲水/疏水单体比值、与离子单体共聚、改变水凝胶内部结构、调整凝胶溶胀剂四个方面介绍了合成具有不同临界相转变温度、温度响应速度及溶胀比的温敏水凝胶及此类凝胶在药物控释、分离萃取等方面的应用。  相似文献   

5.
N-异丙基丙烯酰胺-丙烯酰胺热敏凝胶的溶胀特性   总被引:1,自引:0,他引:1  
制备并表征了N-异丙基丙烯酰胺-丙烯酰胺热敏凝胶(NIPAm-Am),研究了单体配比、引发剂、交联剂用量和温度对其溶胀特性的影响。结果表明:NIPAm-Am热敏凝胶是由亲水和疏水基团组成的非晶高聚物。mAm/mNIPAm越大,凝胶的平衡溶胀率越大;增加交联剂的用量,凝胶的溶胀率减小,当引发剂的质量分数为0.008时溶胀率达最大值;温度的增加会使凝胶的溶胀率减小,在相转变温度时,溶胀率的变化最大。  相似文献   

6.
采用无皂乳液聚合法使N-异丙基丙烯酰胺(NIPAM)、ε-丙烯酰基-L-赖氨酸(εACRLLY)和N,N-亚甲基双丙烯酰胺(BA)交联共聚,制备了含有自由氨基酸侧链的温敏性微凝胶.利用透射电子显微技术(TEM)、动态光散射技术(DLS)及浊度法对所制备的微凝胶的形态及相转变进行了表征.TEM结果表明,所得的微凝胶具有规则的球型形态,微凝胶的粒径随εACRLLY含量的增加而减小.DLS及浊度结果表明,微凝胶粒径呈单分散的窄分布,随着温度的升高,微凝胶粒径减小,有着明显的体积相转变温度(VPTT);亲水单体εACRLLY的引入能够有效地调节共聚物微凝胶的VPTT,并且VPTT随εACRLLY含量的增加几乎呈线性上升.微凝胶在对盐酸阿霉素的药物释放研究表明,所制备的微凝胶在20℃,12 h内释放了56%,37℃下释放了73%;37℃,pH=7.4下13 h内释放73%,pH=4.5下基本释放完毕,该微凝胶表现出良好的药物缓释性能.  相似文献   

7.
由于改变亲水/疏水单体比值、与离子单体共聚心、改变凝胶内部结构等均可不同程度地调整温敏水凝胶的溶胀性能,本研究选择一种既含疏水烷基又含季铵盐正离子型亲水基团的两亲性单体——甲基丙烯酰氧乙基二甲基辛基溴化铵(ADMOAB),结构如示意图1所示.与N-异丙基丙烯酰胺(NIPAM)聚合,制备了P(NIPAM-co-ADMOAB)共聚水凝胶,以便在引入离子型结构单元的同时,改变凝胶体系中亲水/疏水单体比值,避免单纯增加疏水单体引起的水凝胶溶胀性降低问题,并考察了ADMOAB对水凝胶溶胀性能的影响,对该类水凝胶迄今鲜见相关文献报道.该研究对进一步了解水凝胶的构效关系、探索有效控制溶胀性能的途径具有积极意义.  相似文献   

8.
合成了含金刚烷基的甲基丙烯酸金刚烷酯(AdMA)疏水单体,并通过与N-异丙基丙烯酰胺(NIPAM)共聚,制备了温敏性的(P(NIPAM-co-AdMA))共聚物水凝胶.用傅里叶变换红外光谱仪(FTIR)表征了凝胶的化学结构,用环境扫描电镜(ESEM)对凝胶断层结构的形貌进行了观察,用DSC测试了凝胶的体积相转变温度(LCST),并研究了共聚水凝胶的溶胀性能.结果表明,共聚物水凝胶的LCST能够高效地通过改变疏水单体的含量来调节,在实验所考察的范围内,LCST随AdMA含量的增加而线性降低;疏水单体的含量对凝胶的孔洞结构和溶胀性能存在一最优值,在最优的单体配比下,水凝胶具有均匀规整的大孔结构和超快的响应速率.如疏水单体含量为3%(AdMA∶NIPAM=3%)的共聚物水凝胶具有如渔网般均匀的多孔结构,当发生去溶胀时,在5min内就可以失去92%的水,不到10min的时间就可以完全达到去溶胀平衡,水保留率在4%以下.  相似文献   

9.
通过分子结构设计, 合成了疏水性单体4-乙酰基丙烯酰乙酸乙酯(AAEA), 并以该单体与丙烯酸(AA)进行自由基溶液共聚, 制备了P(AAEA-co-AA)新型温度敏感性水凝胶. AAEA的1H NMR及FT-IR分析表明, 该单体主要以烯醇式结构存在; P(AAEA-co-AA)的FT-IR分析发现, PAAEA与PAA之间存在较强烈的氢键作用, 使得AAEA烯醇异构体中的C—O伸缩振动吸收峰移向了低波数处. 对冷冻干燥后凝胶的电镜分析发现, 当AAEA用量较高时, 由于凝胶内部分子链段的疏水聚集, 各部分溶胀度以及溶胀速度不均一而使得凝胶表面粗糙不平. 采用DSC对凝胶的体积相转变进行了研究, 结果表明, 该水凝胶的体积相转变温度(VPTT)在48.2至61.8 ℃之间, 并且随着AAEA用量的减小, 凝胶的VPTT逐渐增加. 对该新型温度敏感性水凝胶在去离子水中的溶胀动力学研究发现, 当AAEA用量高于4.6 g时, 凝胶属于Fick凝胶; 反之凝胶则属于非Fick凝胶. 该水凝胶在去离子水中具有良好的温度敏感性, 当外界温度低于VPTT时, 凝胶能保持溶胀状态; 而当外界温度高于VPTT时, 凝胶的平衡溶胀度迅速下降, 表现为温度敏感性. 进一步研究发现, 凝胶组成不仅会影响凝胶的VPTT, 而且会影响凝胶温度敏感性的强弱.  相似文献   

10.
PNI PAM/CS微凝胶的性质测定   总被引:1,自引:0,他引:1  
以N-异丙基丙烯酰胺(NIPAM)、壳聚糖(CS)为单体,N,N-亚甲基双丙烯酰胺(MBA)为交联剂,制备了PNIPAM/CS微凝胶.测定了不同单体配比对微凝胶体积相转变温度(VPTT)的影响和25℃不同pH条件下微凝胶液浊度及粒径的变化.研究表明,PNIPAM/CS微凝胶具有温敏性;并且随着pH的增大,微凝胶粒径先变小后变大,显示pH敏感性;浊度法测定结果与粒径测定一致.  相似文献   

11.
Two monomers containing functional ? OH groups with different hydrophilic long side chains (viz., triethyleneglycol methacrylate (TREGMA) and polyethyleneglycol methacrylate (PEGMA)) were selected to modify the swelling/deswelling behavior of poly(N‐isopropylacrylamide) (pNIPAM) microgels. Dynamic scattering technique, turbidimetric method, and differential scanning calorimetry (DSC) were employed to investigate the deswelling behavior of the microgels. Experimental results show that the two series of microgels are identical in that incorporation of hydrophilic chains containing ? OH groups causes the volume‐phase transition temperature (VPTT) of pNIPAM microgels to shift to higher temperature; the more hydrophilic the side chains, the more the VPTTs shift. Although PEGMA are more effective in elevating the VPTTs of pNIPAM microgels than TREGMA, p(NIPAM‐co‐TREGMA) microgels show better deswelling properties than p(NIPAM‐co‐PEGMA) microgels, i.e., they have much larger deswelling ratios (α) and display less continuous volume‐phase transition. The VPTTs of the modified microgels can be modulated to well close to the normal body temperature of human beings. These characteristics along with the functional ? OH groups they contain make the microgels competitive candidates for biomaterials. © 2005 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 43: 3575–3583, 2005  相似文献   

12.
The synthesis and properties of thermal/pH-sensitive core-shell copolymer nano/microgels were investigated. The crosslinked core consisted of N-isopropylacrylamide (NIPAAm) while the shell was stabilized by poly(ethylene glycol) methyl ether methacrylate (PEGMA) and 2-methacryloyloxybenzoic acid (2MBA) using a "one pot" soapless emulsion polymerization method. Monodisperse particles were produced with average hydrodynamic diameters ranging from 40 to 880 nm, as determined by dynamic light scattering (DLS) in water at 25°C, depending on the synthetic recipe used. The influence of PEGMA and 2MBA content on size and temperature transition at different pH values was studied. Zeta potential measurements and acid-base titration studies demonstrated almost complete incorporation of acid comonomer (2MBA) into the nano/microgels. Two different crosslinkers, a stable and an acid labile, were compared. The crosslinker used has a major influence on the size and charge density of the nano/microgels produced. Microscopic studies confirmed the core-shell morphology of the nano/microgels.  相似文献   

13.
New divinyl-functionalized acetal-based crosslinkers were synthesized as building elements to form acid-labile microgel particles for controlled-release applications. The synthesized crosslinkers underwent hydrolysis at slightly acidic pHs in less than 1 h while they were stable at neutral pHs for longer times. HEMA was copolymerized with the crosslinkers via an inverse emulsion polymerization technique using a redox initiator system at room temperature to form crosslinked, colloidal p(HEMA) microgels. Microgels in diameters ranging from 150 to 475 nm with narrow distribution could be produced. The crosslinking density and the diameter of the microgels were found to be controlled by monomer/crosslinker feed ratio. The microgels demonstrated a pH-dependent cleavage behavior that mimicked the pH-dependent hydrolysis profile of the acid-labile crosslinkers. Model biomacromolecules, i.e., Rhodamine B-labeled dextran and BSA were efficiently loaded into the microgels. The release of the biomolecules from p(HEMA) microgels was also found to be controllable by the pH of the environment similar to the particle degradation. The protein released from the microgels was observed to retain its structural stability.  相似文献   

14.
Glucosamine-carrying temperature- and pH-sensitive microgels with an average diameter of about 100 nm were successfully prepared by free radical precipitation polymerization. The thermo- and pH-responsive properties of the microgels were designed by the incorporation of N-isopropylacrylamide (NIPAM) and acrylic acid (AAc) to copolymerize with acrylamido-2-deoxyglucose (AADG). The stimuli sensitivity of the microgels was studied by the measurement of their sizes and volume phase transition temperature (VPTT) under different surrounding conditions. The results showed that the microgels were responsive to temperature, pH, and ionic strength, and could have a desired VPTT by modifying AADG and AAc contents. The effect of temperature and pH on insulin release from the microgels was also investigated. The release of drug at the tumor-surrounding environment is faster than that under normal physiological conditions. A preliminary in vitro cell study showed that the glucosamine-carrying microgels are more biocompatible to mouse fibroblast cells, compared to the microgels without glucosamine. These glucosamine-carrying dual-sensitive microgels may be promising carriers for targeted drug delivery to tumors.  相似文献   

15.
The ability to manipulate and control the surface properties of nylons is of crucial importance to their widespread applications. In this work, surface-initiated atom-transfer radical polymerization (ATRP) is employed to tailor the functionality of the nylon membrane and pore surfaces in a well-controlled manner. A simple two-step method, involving the activation of surface amide groups with formaldehyde and the reaction of the resulting N-methylol polyamide with 2-bromoisobutyryl bromide, was first developed for the covalent immobilization of ATRP initiators on the nylon membrane and its pore surfaces. Functional polymer brushes of 2-hydroxyethyl methacrylate (HEMA) and poly(ethylene glycol)monomethacrylate (PEGMA) were prepared via surface-initiated ATRP from the nylon membranes. A kinetics study revealed that the chain growth from the membranes was consistent with a "controlled" process. The dormant chain ends of the grafted HEMA polymer (P(HEMA)) and PEGMA polymer (P(PEGMA)) on the nylon membranes could be reactivated for the consecutive surface-initiated ATRP to produce the corresponding nylon membranes functionalized by P(HEMA)-b-P(PEGMA) and P(PEGMA)-b-P(HEMA) diblock copolymer brushes. In addition, membranes with grafted P(HEMA) and P(PEGMA) brushes exhibited good resistance to protein adsorption and fouling under continuous-flow conditions.  相似文献   

16.
Combining hydrophobic materials such as polydimethylsiloxane (PDMS), a natural hydrophobic material with typical hydrophilic monomers without using organic solvent remains a big challenge due to their extreme different properties; hence, fabricating a class of silicone hydrogels with two extremes without use of organic solvents could bring us a novel class of silicone hydrogels. Herein, a range of PDMS‐HEMA‐PEGMA hydrogels was prepared from hydroxyl‐terminated PDMS, 2‐hydroxyethylmethacrylate (HEMA), poly (ethylene glycol) methacrylate (PEGMA), and isophorone diisocyanate via condensation and radical copolymerization reactions. The infrared results confirmed the PDMS‐HEMA‐PEGMA network formation, while the hydrophilicity of the as‐prepared block copolymer was dependent on (PDMS‐HEMA)/PEGMA ratio. Increasing the PEGMA content resulted in increased equilibrium water content, phase separation, surface roughness, and tensile strength, while the tensile modulus, elongation at break, optical transmittance, water contact angle, and oxygen permeability (Dk) were decreasing. At PEGMA content of 28.3%, the relative protein adsorption ratio decreased to 20% and 36% for bovine serum albumin and lysozyme, respectively, compared with that of the control (PDMS‐HEMA), suggesting antiprotein adsorption ability. In overall, the results showed that the PDMS‐HEMA‐PEGMA hydrogels not only exhibited remarkable hydrophilicity and suppressed protein adsorption but also maintained higher optical transparency and oxygen permeability (Dk).  相似文献   

17.
The flocculation behavior of poly(N-isopropylacrylamide) (pNIPAM) microgels containing polar -(OCH(2)CH(2))(3)OH chains, incorporated by the copolymeric components (triethyleneglycol methacrylate, TREGMA), in aqueous NaCl solution was investigated. Determination of the critical flocculation temperatures (CFTs) and the critical flocculation concentrations (CFCs) of the microgels at 45 degrees C shows that polar -(OCH(2)CH(2))(3)OH chains have different influence on the flocculation behavior of the microgels at temperatures below and above their volume phase transition temperatures (VPTTs). The flocculation of the microgels becomes more difficult with the increase of -(OCH(2)CH(2))(3)OH chains below the VPTT. In contrast, the microgels flocculate more easily with more -(OCH(2)CH(2))(3)OH chains above the VPTT. Preliminary investigation on the flocculation kinetics of the microgels further shows that -(OCH(2)CH(2))(3)OH chains have different effects on the flocculation rate at temperatures below and above the VPTT. The flocculating rate of the microgels at 25 degrees C decreases with the increase of -(OCH(2)CH(2))(3)OH chains. While the flocculation rate at 45 degrees C increases with the increase of -(OCH(2)CH(2))(3)OH chains due to their enrichment on the surface of the microgels as a result of the temperature-induced volume-phase transition, which was verified by variable temperature (1)H NMR spectroscopy. The polar -(OCH(2)CH(2))(3)OH chains rich in the surface increase the attractive force between the microgels, promoting the flocculation.  相似文献   

18.
The influence of butyl acrylate (BA) and methyl methacrylate (MMA) on hydroxyl functionalized latexes was investigated. The hydrophobicity of the monomer feed was varied via the BA/MMA ratio. In addition to monitoring the effect of hydrophobic monomer feed on secondary nucleation, the polymerization kinetics and final latex properties were also obtained for comparison. Five different BA to MMA molar ratios were combined with five 2‐hydroxyethyl methacrylate (HEMA) concentrations (0, 10, 20, 30 and 40 mol% in monomer composition). All latexes were synthesized through seeded semibatch emulsion polymerization process. Particle size distributions and average particle sizes of the latexes were determined by dynamic light scattering (DLS) and qualitatively compared with transmission electron microscope (TEM) images. The BA to MMA ratio significantly influences the boundary HEMA concentration at which homogeneous secondary nucleation occurs. © 2017 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2017 , 55, 2190–2202  相似文献   

19.
Emulsion polymerization of 2-(diethylamino)ethyl methacrylate (DEA) in the presence of a bifunctional cross-linker at pH 8-9 afforded novel pH-responsive microgels of 250-700 nm diameter. Both batch and semicontinuous syntheses were explored using thermal and redox initiators. Various strategies were evaluated for achieving colloidal stability, including charge stabilization, surfactant stabilization, and steric stabilization. The latter proved to be the most convenient and effective, and three types of well-defined reactive macromonomers were examined, namely, monomethoxy-capped poly(ethylene glycol) methacrylate (PEGMA), styrene-capped poly[2-(dimethylamino)ethyl methacrylate] (PDMA50-St), and partially quaternized styrene-capped poly[2-(dimethylamino)ethyl methacrylate] (10qPDMA50-St). The resulting microgels were pH-responsive, as expected. Dynamic light scattering and 1H NMR studies confirmed that reversible swelling occurred at low pH due to protonation of the tertiary amine groups on the DEA residues. The critical pH for this latex-to-microgel transition was around pH 6.5-7.0, which corresponds approximately to the known pKa of 7.0-7.3 for linear PDEA homopolymer. The microgel particles were further characterized by electron microscopy and aqueous electrophoresis studies. Their swelling and deswelling kinetics were investigated by turbidimetry. The PDEA-based microgels were compared to poly[2-(diisopropylamino)ethyl methacrylate] (PDPA) microgels prepared with identical macromonomer stabilizers. These PDPA-based microgels had a lower critical swelling pH of around pH 5.0-5.5, which correlates with the lower pKa of PDPA homopolymer. In addition, the kinetics of swelling for the PDPA microgels was somewhat slower than that observed for PDEA microgels; presumably this is related to the greater hydrophobic character of the former particles.  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