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 185 毫秒
1.
基于互穿网络结构的pH/温度双重刺激响应性微凝胶的研究   总被引:1,自引:0,他引:1  
室温下采用氧化-还原引发体系,以低交联密度的聚(N-异丙基丙烯酰胺)(PNIPAM)微凝胶为种子,通过种子乳液聚合法合成由PNIPAM和聚丙烯酸(PAA)形成的具有互穿聚合物网络结构的微凝胶.傅立叶变换红外光谱分析结果表明微凝胶由PNIPAM和PAA两种聚合物组成,透射电镜表征结果证实微凝胶中PNIPAM和PAA两种聚合物形成了互穿网络结构.用动态激光光散射测试不同温度或pH值水介质中微凝胶的粒径,结果发现微凝胶具有良好的pH/温度双重刺激响应性.在水介质pH值大于5.5的情况下,PAA组分对微凝胶的体积相转变温度没有影响;而在水介质pH值为4.0的情况下,由于PAA与PNIPAM之间的氢键作用,微凝胶的体积相转变温度稍微降低.微凝胶中PAA组分含量越高,其pH刺激响应性越显著.  相似文献   

2.
CMC/PNIPAAm半互穿网络水凝胶的溶胀动力学研究   总被引:4,自引:3,他引:1  
以羧甲基纤维素钠(CMC)和N-异丙基丙烯酰胺(NIPAAm)为原料,制备了具有温度和pH敏感性的半互穿网络(CMC/PNIPAAmsemi-IPN)水凝胶,并研究了水凝胶在不同温度和pH值条件下的溶胀行为。结果表明:在弱碱性(pH-7.4)条件下,凝胶的溶胀速率和溶胀度都随着凝胶中CMC含量的增加而增大;而在酸性(pH-1.O)条件下则相反。在弱碱性条件下,水分子在凝胶中的扩散行为都可用non-Fickian扩散来描述,水分子在凝胶中的扩散系数D随着凝胶溶胀速率的增大而增大;在酸性条件下,20℃时凝胶的溶胀过程符合non-Fickian扩散规律,而37℃时凝胶的溶胀过程符合Fickian扩散规律,但水分子的扩散系数D相差不大。  相似文献   

3.
以N-异丙基丙烯酰胺(NIPAM)、甲基丙烯酸(MAA)为单体,N,N-亚甲基双丙烯酰胺(MBA)为交联剂,制备了温敏性聚(N-异丙基丙烯酰胺)(PNIPAM)和具有温度、pH敏感性的聚(N-异丙基丙烯酰胺-co-甲基丙烯酸)(PNIPAM-MAA)微凝胶。通过测定不同温度和pH条件下微凝胶浊度变化,表征微凝胶的温度及pH敏感性,描述了NaCl浓度和pH对微凝胶体积相转变温度的影响。同时,测定了微凝胶的临界聚沉浓度及临界絮凝温度,表征了微凝胶的稳定性,讨论了影响微凝胶的稳定性因素。  相似文献   

4.
以丙烯酰胺(AM)和丙烯酸(AA)单体的水溶液为分散相,失水山梨醇单油酸脂(Span80)/聚氧乙烯失水山梨醇脂肪酸脂(Tween80)/异辛烷为分散介质,分别以N,N′-亚甲基双丙烯酰胺(MBA)、过硫酸铵/亚硫酸氢钠((NH4)2S2O8/NaHSO3)为交联剂和氧化还原引发剂,在30℃进行反相微乳液聚合制备了一系列不同单体摩尔百分数的P(AM-co-AA)微凝胶.通过傅立叶红外光谱、浊度法、透射电镜(TEM)和动态光散射(DLS)等测试手段分别对微凝胶特征官能团的存在、pH敏感性、微观形态、粒径大小及粒径分布等进行表征分析.结果表明,共聚物中存在AM和AA结构单元;样品的TEM照片显示在原料中AA的摩尔百分数为60%时,P(AM-co-AA)微凝胶粒子的数均粒径为90 nm左右,呈现非规则球形;DLS结果表明,P(AM-co-AA)微凝胶与PAM微凝胶相比具有较宽的粒径分布,且随原料中AA摩尔百分数增加,粒径分布逐渐变宽;P(AM-co-AA)微凝胶具有良好的pH敏感性,敏感pH值与AA的解离常数有关,通过调节pH值可以迅速控制自身体积的溶胀与收缩.  相似文献   

5.
温度、pH敏感性核壳结构微凝胶的制备及性质   总被引:8,自引:0,他引:8  
以无皂乳液分步聚合的方法, 将N-异丙基丙烯酰胺(NIPAM)与N,N-亚甲基双丙烯酰胺(MBA)交联反应3 h, 制得种子乳液, 再向种子乳液中加入甲基丙烯酸(MAA)功能性单体继续反应2 h, 制备了具有温度、pH敏感性的核壳结构微凝胶. 通过透射电子显微镜(TEM)、红外光谱(IR)等表征了微凝胶外貌形态及结构组成, 动态光散射(DLS)测定了微凝胶粒径响应热、pH的变化及微凝胶Zeta电位的变化. 结果表明凝胶形貌为异型核壳结构; Zeta电位与微凝胶粒径随温度、pH变化相关.  相似文献   

6.
利用溶剂热法通过控制反应时间和温度制得了分散性好和磁性强的Fe3O4,并利用溶胶凝胶法制备得到包覆SiO2的磁性微球(Fe3O4@SiO2)。以三聚氰胺为模板分子,甲基丙烯酸(MAA)为单体,采用本体聚合法制备了磁性分子印迹聚合物(MMIPs)。通过静态吸附实验表明,MMIPs对三聚氰胺的饱和吸附量高达10.22μg/mg,是磁性非印迹聚合物(MNIPs)的1.62倍。粒子扩散模型、Elovich模型和动态吸附实验表明所制得的MMIPs有较好的吸附性能。  相似文献   

7.
研究了在线型聚 (丙烯酸 ) (PAA)溶液中链长、pH、离子强度和水 /二甲亚砜混合溶剂组成对非离子聚 (N-乙烯基 - 2 -吡咯烷酮 ) (PVP)水凝胶溶胀特性的影响 .发现聚酸浓度的变化引起凝胶显著的体积相变 ,这是因为凝胶和聚合物通过氢键形成了大分子间凝胶 -聚合物复合物 .凝胶的溶胀特性取决于聚酸的链长而不是离子强度 .随着pH值和水 /二甲亚砜混合溶剂组成的变化 ,凝胶的溶胀率 (SR)发生变化  相似文献   

8.
设计合成了一类新型结构的聚环氧乙烷(PEO)大分子链转移剂,调控3-丙烯酰胺基苯硼酸(AAPBA)的可逆加成断裂-链转移(RAFT)自由基聚合,合成得到3种PAAPBA链段长度不同的PAAPBA-bPEO-b-PAABPA3嵌段共聚物.研究了3种聚合物在生理pH值下的凝胶化行为,证明凝胶的形成与PAABPA的长度有关,当该链段较长时,由于PAABPA链段疏水性太强,不能形成稳定的水凝胶.详细研究了聚合物浓度、温度、葡萄糖浓度对凝胶流变行为的影响,证明共聚物浓度越高,形成的凝胶的强度更大,性质上更接近于固体,浓度较高条件下形成的凝胶的转变温度较高.凝胶表现出葡萄糖敏感性,当高葡萄糖存在时,随时间延长,凝胶会发生崩解直至最后溶解.凝胶亲水微区能包载蛋白质FITC-BSA,加入葡萄糖后,FITC-BSA的释放加快.  相似文献   

9.
阳离子化热响应微凝胶的合成及在二氧化硅矿化中的应用   总被引:1,自引:0,他引:1  
采用无皂乳液聚合技术,在亚甲基双丙烯酰胺(MBA)为交联剂的情况下,N-异丙基丙烯酰胺(NIPAM)与甲基丙烯酰氧乙基三甲基氯化铵(DMC)发生共聚,生成具有阳离子功能化的热响应微凝胶poly-(NIPAM-co-DMC).TEM研究表明该微凝胶粒子的粒径约为200 nm左右,具有规则的球形形态.DLS和1H-NMR研究证实了微凝胶粒子的最低临界溶液温度(LCST)在34℃左右.进一步以此微凝胶为模板,在中性条件下,以四甲氧基硅烷(TMOS)为硅源,在此模板上仿生沉积S iO2,生成poly(NIPAM-co-DMC)/S iO2杂化纳米粒子.FTIR、TEM、1H-NMR及TGA等研究表明S iO2在聚合物模板上发生了沉积.能谱分析进一步证明了S iO2主要分布在杂化纳米粒子的壳层区域.另外,当矿化反应温度高于微凝胶的LCST值时,体系生成了具有明显核壳结构的异形杂化粒子.  相似文献   

10.
制备了在修复受损组织方面有应用潜能的纳米级聚(甲基丙烯酸羟乙酯/甲基丙烯酸) (P(HEMA/MAA))微凝胶; 采用试管倒转法对不同pH值和浓度的P(HEMA/MAA)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低浓度和高浓度微凝胶分散液的流变性能, 并对pH触发物理凝胶化相转变机理进行了推测. 结果表明: 在生理pH值环境下, 一定浓度的P(HEMA/MAA)微凝胶分散液可以发生凝胶化相转变形成凝胶态, pH=7时, HEMA/MAA进料摩尔比为8/2的微凝胶分散液凝胶化后得到的凝胶力学性能最佳, 最大弹性模量(G')可达7.58×103 Pa; P(HEMA/MAA)微凝胶颗粒在不同条件下具有不同的溶胀效果, 导致低浓度分散液的表观粘度发生相应的变化, 并由此推测出微凝胶颗粒的溶胀过程由外及内, 分为三个阶段; 高浓度微凝胶分散液发生凝胶化相转变主要是由颗粒间或颗粒与分散介质间形成的空间静电稳定作用和氢键共同作用引起的.  相似文献   

11.
聚酰胺酸结构及其亚胺化的红外光谱分析   总被引:2,自引:0,他引:2  
利用变温透射红外光谱方法,通过跟踪聚酰胺酸(PAA)的亚胺化过程,对由均苯四酸二酐和4,4′-二氨基二苯醚合成的聚酰胺酸及经过加热亚胺化后生成的聚酰亚胺(PI)的红外吸收光谱进行分析,对聚酰胺酸和聚酰亚胺的红外谱峰进行合理的归属,发现聚酰胺酸在亚胺化过程中有-COO-和-NH+2存在,-COO-中羰基的对称与反对称伸缩振动分别位于1607和1406 cm-1,NH+2的伸缩振动则有3200、3133、2938、2880、2820和2610 cm-1等多个精细谱带。 并根据对-COO-和-NH+2谱峰的归属,提出聚酰胺酸生成聚酰亚胺的机理为聚酰胺酸中COOH的H+转移到聚酰胺酸中的NH上,形成NH+2,然后脱水环化生成聚酰亚胺。  相似文献   

12.
pH‐sensitive poly (vinylidene fluoride) (PVDF)/poly (acrylic acid) (PAA) microgels membranes are prepared by phase inversion of the N, N‐dimethylformamide solution containing PAA microgels and PVDF in aqueous solution. The composition and structure of the blend membrane are investigated by Fourier transform infrared spectra, X‐ray photoelectron spectroscopy measurements, thermo gravimetric analysis, field‐emission scanning electron microscope and atomic force microscope. The results indicate the surface and cross section of the blend membranes have a porous structure with PAA microgels immobilized inside the pore and on the membrane surface. The blend PVDF membranes exhibit pH‐sensitive water flux, with the most drastic change in permeability observed between pH 3.7 and 6.3. The blend membranes are fouled by bovine serum albumin, and their antifouling property is enhanced by increasing PAA microgels, mainly derived from the improved hydrophilic property. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

13.
Copolymers of N-isopropylacrylamide (IPA) and alkyl acrylates [methyl acrylate (MA), ethyl acrylate, and butyl acrylate] or vinyl acetate have been prepared and their phase transitions in water have been observed by means of IR spectroscopy. The incorporation of these alkyl acrylates into a poly(IPA) (PIPA) chain induces a decrease in the phase-transition temperatures, Tp, and the magnitude increases with increasing size of the alkyl chains. The profiles of the C=O stretching absorption bands of the ester groups [9(C=O)ester] and the IR bands due to IPA units exhibit critical changes at the Tp of these copolymers. The 9(C=O)ester bands shift slightly toward higher wavenumbers (blueshift) upon phase transition, while the amide I and amide II bands of the IPA units undergo a blueshift and a redshift, respectively. Analysis of the 9(C=O)ester band of PIPA-MA by using a curve-fitting method shows that it consists of three components, at 1,703, 1,720, and 1,738 cm-1. The relative peak area of the largest component (1,720 cm-1) is almost constant, and those of the 1,703-cm-1 and 1,738-cm-1 components increase and decrease with increasing temperature during the phase transition, respectively. However, the changes are rather small, suggesting that changes in hydrogen bonding of the C=O groups of MA units upon phase transition are not significant. The 9(C=O)ester bands of other comonomers examined here also exhibit similar changes. The situation is consistent with the change in the hydration states of the amide groups of IPA units, most of which associate with water molecules through hydrogen bonds even after the phase separation.  相似文献   

14.
The miscibility of poly(N-isopropylacrylamide) (PNIPA) with poly(vinyl pyrrolidone) (PVP) and a cross-linked poly(acrylic acid) (Carbopol® 971P) was evaluated from the rheological data of aqueous dispersions and the temperature of glass transitions of films made of binary mixtures. PNIPA has a low critical solubility temperature (LCST) of about 33°C, below which 1% dispersion behaves as a viscous system. At temperatures above LCST, the hydrophobic interactions among the isopropyl groups initially provide transient networks of greater elasticity. The LCST of PNIPA as well as its T g (144°C, estimated by DSC and MTDSC of films) were not modified by the presence of PVP. The immiscibility of PNIPA and PVP was confirmed by the absence of interaction between both polymers as shown by FTIR analysis of the films. In contrast, PNIPA and carbopol were miscible and the behaviour of their mixtures differed significantly from that of the parent polymers; i.e. a strong synergistic effect on the viscoelasticity of the dispersions was observed below the LCST. As temperature increased, the blends showed a decrease in the loss and storage moduli, especially those with greater PNIPA proportions. The fall was smoother as the PNIPA proportion decreased. This behaviour may be explained as the result of the balance between PNIPA/carbopol hydrogen bonding interactions (as shown in the shift of C=O stretch in FTIR spectra) and PNIPA/PNIPA hydrophobic interactions. The T g values of the films of the blends showed a positive deviation from the additivity rule; the mixtures containing more than 1:1 amide:carboxylic acid groups have a notably high Tg (up to 181°C). This increase is related to the stiffness induced in the films by the PNIPA/carbopol interactions.  相似文献   

15.
A novel pH- and temperature-sensitive nanocomposite microgel based on linear Poly(acrylic acid) (PAAc) and Poly(N-isopropylacrylamide) (PNIPA) crosslinked by inorganic clay was synthesized by a two-step method. First, PNIPA microgel was prepared via surfactant-free emulsion polymerization by using inorganic clay as a crosslinker, and then AAc monomer was polymerized within the PNIPA microgel. The structure and morphology of the microgel were confirmed by FTIR, WXRD and TEM. The results indicated that the exfoliated clay platelets were dispersed homogeneously in the PNIPA microgels and acted as a multifunctional crosslinker, while the linear PAAc polymer chains incorporated in the PNIPA microgel network to form a semi-interpenetrating polymer network (semi-IPN) structure. The hydrodynamic diameters of the semi-IPN microgels ranged from 360 to 400 nm, which was much smaller than that of the conventional microgel prepared by using N,N′-methylenebis(acrylamide) (MBA) as a chemical crosslinker, the later was about 740 nm. The semi-IPN microgels exhibited good pH- and temperature-sensitivity, which could respond independently to both pH and temperature changes.  相似文献   

16.
The phenomenon of self-assembly of aggregates formed by relatively short chains of poly(vinyl alcohol) (PVA) on the long macromolecules of polyacrylamide (PAA) in aqueous medium are discussed. PVA and PAA form intermolecular polycomplexes (InterPC) of a constant composition independently on a ratio of polymer components. The complex formation between high-molecular-weight PAA and relatively low-molecular-weight poly(ethylene oxide) (PEO) are considered also. PEO with M ⩽ 4·104 g.mol−1 weakly interacts with PAA. The polymer-polymer interaction can be intensified when the part of amide groups (∼20 mol %) on PAA chain to transform into the carboxylic groups. InterPCs formed by PEO and initial or modified PAA have associative structure with friable packing of the polymer segments. They are stabilized by the hydrogen bond system.  相似文献   

17.
X-ray crystallography of collagen model peptides has provided high-resolution structures of the basic triple-helical conformation and its water-mediated hydration network. Vibrational spectroscopy provides a useful bridge for transferring the structural information from X-ray diffraction to collagen in its native environment. The vibrational mode most useful for this purpose is the amide I mode (mostly peptide bond C=O stretch) near 1650 cm-1. The current study refines and extends the range of utility of a novel simulation method that accurately predicts the infrared (IR) amide I spectral contour from the three-dimensional structure of a protein or peptide. The approach is demonstrated through accurate simulation of the experimental amide I contour in solution for both a standard triple helix, (Pro-Pro-Gly)10, and a second peptide with a Gly --> Ala substitution in the middle of the chain that models the effect of a mutation in the native collagen sequence. Monitoring the major amide I peak as a function of temperature gives sharp thermal transitions for both peptides, similar to those obtained by circular dichroism spectroscopy, and the Fourier transform infrared (FTIR) spectra of the unfolded states were compared with polyproline II. The simulation studies were extended to model early stages of thermal denaturation of (Pro-Pro-Gly)10. Dihedral angle changes suggested by molecular dynamics simulations were made in a stepwise fashion to generate peptide unwinding from each end, which emulates the effect of increasing temperature. Simulated bands from these new structures were then compared to the experimental bands obtained as temperature was increased. The similarity between the simulated and experimental IR spectra lends credence to the simulation method and paves the way for a variety of applications.  相似文献   

18.
Temperature‐induced phase separation of poly(N‐isopropylacrylamide) in aqueous solutions was studied by attenuated total reflectance (ATR)/Fourier transform infrared spectroscopy. The main objectives of the study were to understand, on a molecular level, the role of hydrogen bonding and hydrophobic effects below and above the phase‐separation temperature and to derive the scenario leading to this process. Understanding the behavior of this particular system could be quite relevant to many biological phenomena, such as protein denaturation. The temperature‐induced phase transition was easily detected by the ATR method. A sharp increase in the peaks of both hydrophobic and hydrophilic groups of the polymer and a decrease in the water‐related signals could be explained in terms of the formation of a polymer‐enriched film near the ATR crystal. Deconvolution of the amide I and amide II peaks and the O? H stretch envelope of water revealed that the phase‐separation scenario could be divided, below the phase‐separation temperature, into two steps. The first step consisted of the breaking of intermolecular hydrogen bonds between the amide groups of the polymer and the solvent and the formation of free amide groups, and the second step consisted of an increase in intramolecular hydrogen bonding, which induced a coil–globule transition. No changes in the hydrophobic signals below the separation temperature could be observed, suggesting that hydrophobic interactions played a dominant role during the aggregation of the collapsed chains but not before. © 2001 John Wiley & Sons, Inc. J Polym Sci Part B: Polym Phys 39: 1665–1677, 2001  相似文献   

19.
This study aimed to investigate the effect of COOH group distribution within a polymer network having amide groups, with which the COOH could form hydrogen bonds. We employed here two polyelectrolyte gels composed of N-isopropylacrylamide (NIPA) networks, either copolymerized with acrylic acid (AA) or within which poly(acrylic acid) (PAA) was entrapped. Both gels (AA–NIPA ∼ 1:4 mol/mol) were prepared by aqueous red-ox polymerization with N,N’-methylenebisacrylamide as the cross-linker. Finely divided gels in NaCl solutions (0.025 and 0.1 M) were titrated with NaOH and back-titrated with HCl at 25 °C. The results of the copolymer gel (CG) agreed well with those of a linear copolymer and a nanoscale gel which had a similar AA content to CG. However, marked differences were observed in the titration behaviors of the AA-copolymerized and PAA-entrapped gels, mainly due to the hydrogen bonding between the entrapped PAA chain and its surrounding NIPA network.  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