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 453 毫秒
1.
采用耗散粒子动力学(Dissipative particle dynamics, DPD)模拟方法研究了三嵌段共聚物聚氧乙烯-聚氧丙烯-聚氧乙烯(PEO-PPO-PEO)的胶束化和凝胶化行为. 通过模拟得到了F127(EO99PO65EO99)水溶液的临界胶束浓度和临界凝胶浓度. 结果发现, 在298 K、 质量分数低于40%时, F127水溶液中形成的胶束形状均为球形. 此外,进一步研究了亲水嵌段长度对胶束结构及凝胶形成浓度的影响, 结果发现, 亲水嵌段越短, 越有利于长椭球状胶束的形成, 而临界凝胶浓度随着亲水嵌段PEO长度的增加而降低.  相似文献   

2.
采用原子转移自由基聚合伴随水解的方法合成了聚丙烯酸-聚醚嵌段共聚物(PAA-F108-PAA), 并通过氢核磁共振波谱和二维核Overhauser效应谱(2D NOE)研究了温度、 羧酸基团中和度(α)及盐浓度对PAA-F108-PAA嵌段共聚物在水溶液中胶束化行为的影响. 结果表明, PAA-F108-PAA分子的临界胶束化温度受α影响较小, 受盐的种类和浓度影响较大. 当α=0.14(0.01 mol/L KCl)时, 在6 ℃条件下, PAA-F108-PAA分子处于塌缩状态, 而在60 ℃条件下, 聚氧化丙烯(PPO)链段发生疏水聚集形成胶束的核, PAA链段与PEO链段相互作用形成胶束的壳; 当α=0.80(0.01 mol/L KCl)时, 在6 ℃条件下, PAA-F108-PAA分子处于相对伸展状态, 而在60 ℃条件下, PPO链段仍发生疏水聚集形成胶束的核, PEO与PAA彼此分离形成胶束的壳. 增加KCl的浓度至1 mol/L, PAA-F108-PAA分子的临界胶束化温度显著降低, KCl对PPO和PEO链段都表现出脱水作用. 但KI的浓度增加至1 mol/L时, PAA-F108-PAA分子的临界胶束化温度仅略微增加, KI对PPO链段表现出脱水作用, 而对PEO链段表现出增溶作用.  相似文献   

3.
采用TEM和AFM研究了PS(聚苯乙烯)43-b-PEO(聚氧乙烯)45-b-PS43和PS39-b-P4VP(聚4-乙烯基吡啶)98-b-PS39三嵌段共聚物在水介质中的球形胶束、囊泡和蠕虫状胶束之间的二次聚集行为.实验发现,初级聚集体的二次聚集具有不同的复杂程度.对称性的初级聚集体,如球形胶束和囊泡,其二次聚集表现出球形对称性;而非对称性初级聚集体(如蠕虫状胶束)二次聚集开始倾向于非对称性.BAB型的嵌段设计有利于初级聚集体的二次聚集发生.  相似文献   

4.
实验研究了剪切(搅拌)对ABC三嵌段共聚物PS720-b-P2VP200-b-PEO375在溶液中自组装形成的胶束形态的影响,研究结果表明剪切对多节状蠕虫胶束的生成和结构有着重要作用.在1500 r/min剪切速率时,嵌段共聚物自组装形成的球形胶束首先聚集形成蠕虫胶束的梭状轮廓,然后再经过不断地融合与调整形成蠕虫胶束节状部分的盘状结构,同时球的融合趋于沿着垂直于梭状结构的主轴方向(即流场方向).溶剂THF对PS嵌段充分的溶胀使得球形胶束进一步调整形成盘状结构,从而使梭状胶束聚集体顺利地向多节状蠕虫胶束过渡.通过透射电镜(TEM)和扫描电镜(SEM)对胶束形态进行表征,结果表明,多节状蠕虫胶束是剪切作用下球形胶束二次自组装的结果.  相似文献   

5.
采用耗散粒子动力学方法,研究了两亲性嵌段共聚物和双疏性嵌段共聚物共混体系的自组装行为,探讨了双疏性嵌段共聚物的浓度以及双疏性嵌段共聚物的嵌段体积分数对聚集体结构的影响.结果表明,随着双疏性嵌段共聚物浓度的增加,聚集体发生自囊泡到棒状胶束再到同心圆多舱胶束的转变,且当浓度较高时,同心圆多舱胶束的同心圆层数量与浓度密切相关.当双疏性嵌段共聚物中的嵌段体积分数降低时,球形胶束由同心圆结构转变为非同心圆结构.此外,利用Minkowski泛函方法表征了多舱胶束的形成过程,发现这是一个先形成大尺度球形结构、再形成小尺度内核结构的分级组装过程.  相似文献   

6.
两亲嵌段共聚物溶液内胶束形成的温度效应   总被引:2,自引:0,他引:2  
合成了一系列具有两亲嵌段结构的聚(乙二醇)(PEO)一聚(丙二醇)(PPO)共聚物.利用荧光探针及示差量热法测定了共聚物水溶液的临界胶束形成温度(CMT)值.发现二嵌段共聚物(PEO-PPO)和三嵌段共聚物(PEO-PPO-PEO)有着类似的变化规律,即随共聚物分子中疏水链(PPO)长度的增大,其CMT值降低.但三嵌段共聚(PPO-PEO-PPO)则因疏水链段处于共聚物分子的两端,因而在溶液中有可能形成立体网状交联结构.此外,利用探针分子在不同极性溶剂中荧光峰值波长发生位移的现象可以对形成胶束内核的组织程度、极性大小进行估测.  相似文献   

7.
聚合诱导自组装(PISA)是一种在高浓度溶液中可连续大量制备纳米材料的新技术,结合计算模拟方法,研究其动力学过程可强化对PISA的认识和调控.通过耗散粒子动力学(DPD)模拟,研究了ABC三嵌段共聚物的聚合诱导自组装过程.先利用亲溶剂A链段引发B单体聚合,随着疏溶剂B链段的增长,AB二嵌段共聚物可组装并发生聚集体结构的连续转变,由球形胶束→蠕虫状胶束→层状结构→囊泡.再将C单体逐步聚合到AB共聚物上,调控C链段的亲疏溶剂性,可聚合诱导组装或解组装形成不同的ABC三嵌段共聚物聚集体.  相似文献   

8.
采用透射电子显微镜(TEM)观察由结晶性均聚物聚氧乙烯(h-PEO)与半结晶性嵌段共聚物聚苯乙烯-嵌-聚氧乙烯-嵌-聚苯乙烯(SEOS)组成的干、湿刷共混物薄膜结构.结果表明:4种共混系统的薄膜结构均由嵌段PEO和PEO均聚物组成球形PEO分散相;随着均聚物含量的增加,PEO分散相尺寸逐渐增大;当均聚物质量分数增大到33.3%时,在聚苯乙烯(PS)连续相中形成了类似胶束的"核-壳"结构.  相似文献   

9.
采用耗散粒子动力学方法(DPD),模拟了聚氧乙烯-聚氧丙烯-聚氧乙烯(PEO—PPO—PEO)三嵌段共聚物在乙醇溶液中的自组装行为,考察了该共聚物的体积分数和聚氧乙烯(PEO)嵌段链长对介观形貌的影响。当F88(PEO104-PPO39-PEO104)体积分数为20%时,胶柬由初始的均衡分散态逐渐聚合,最终形成PPO为核、PEO为壳的平衡态柱状团聚体。改变共聚物的体积分数和PEO链的长度,会形成不同的介观结构,如:球状、柱状、立体网络、层状和穿孔状结构等。结果表明,DPD方法是研究三嵌段共聚物自组装行为和介观结构形成机理的有效工具,对合成具有特定结构性能的材料有一定的指导意义。  相似文献   

10.
庄莹  王立权  林嘉平 《高分子学报》2011,(11):1320-1328
采用实空间求解的自洽场理论,研究了两亲性二嵌段共聚物(AB)/均聚物(C)超分子体系在溶液中的自组装行为,其中B疏水嵌段的自由末端与C均聚物的一个末端形成可逆的非共价键.在稀溶液中,AB/C超分子聚合物体系通过自组装形成了一系列不同形貌的胶束,如核-壳-冠的三层胶束和蠕虫状胶束等.研究发现,胶束形貌受到非共价键强度和初...  相似文献   

11.
Dissipative particle dynamics (DPD) was used to simulate the formation and stabilization of gold nanoparticles in poly(ethylene oxide)–poly(propylene oxide)–poly(ethylene oxide) (PEO–PPO–PEO) block copolymer micelles. Primary gold clusters that were experimentally observed in the early stage of gold nanoparticle formation were modeled as gold bead in DPD simulation. It showed that gold beads were wrapped by the block copolymer and aggregated into spherical particles inside the micelles and forming stable Pluronic–gold colloids with two-layer structures. Increasing Pluronic concentration, molecular weight, and PPO block length led to the formation of more uniform and more stable gold nanoparticles. Density profiles of water beads suggested that the micelles, especially the hydrophobicity of the micellar cores, played an important role in stabilizing gold nanoparticles. Dynamic process indicated that the formation of gold nanoparticles was controlled by the competition between aggregation of primary gold clusters and the stabilization by micelles of block copolymers.. The DPD simulation results of gold–copolymer–water system agree well with previous experiments, while more structure information on microscopic level could be provided.  相似文献   

12.
J. Jin  F. Pan 《Thermochimica Acta》2007,456(1):25-31
Crystallisation behaviour of poly(ethylene oxide) (PEO)/multi-walled carbon nanotubes (MWCNT) and PEO/chemically modified MWCNT nanocomposites were investigated by means of differential scanning calorimetry. Non-isothermal crystallisation experiments showed that incorporation of MWCNT and chemically modified MWCNT reduced the crystallinity and restricted the spherical crystal growth of PEO. The nucleation sites decrease and spherical crystal size increased compared to the neat PEO. Change of crystal structure from spherical to disk-like was revealed by Avrami equation when MWCNT was added up to 1 wt.%.  相似文献   

13.
Small-angle neutron scattering has been used to investigate the associative structures formed by triblock copolymers of poly(ethylene oxide) (PEO)-polypropylene oxide (PPO)-poly(ethylene oxide) (PEO) (also known as Pluronics) and to monitor the structural changes occurring upon complexation with heptakis(2,6-di-O-methyl)-beta-cyclodextrin (hbeta-CD) over the temperature range from 5 to 70 degrees C. At low temperature, the Pluronics are dispersed as unimers. Close to ambient temperature, the hydrophobicity of PPO causes the aggregation of the polymers into spherical micelles with core sizes between 40 and 50 A and a high inclusion of solvent. The aggregation number increases with temperature as the hydrophobicity of the core is gradually enhanced. hbeta-CD spontaneously forms pseudopolyrotaxanes with the triblock copolymers either when in their unimer form or micellized. The complexation results in an increase in the effective critical micellar concentration. It is suggested that the cyclodextrins thread onto the polymer backbone to localize preferentially on the central PPO block, therefore improving its water solubility. At temperatures where the polymers exist in micellar form, complexation with hbeta-CD gives rise to a complete disruption of the aggregates. These processes are highly temperature-dependent. Above 50 degrees C, the break-up of the aggregates is inhibited, and large-scale aggregation is observed.  相似文献   

14.
In an attempt to understand the role of interfacial friction on solute rotation, fluorescence anisotropy decays of a cationic solute, rhodamine 110, have been measured in polymorphic environments of a triblock copolymer, (PEO)20-(PPO)70-(PEO)20 (P123) (PEO = poly(ethylene oxide), PPO = poly(propylene oxide)). It has been noticed that even though rhodamine 110 is located in the interfacial region of the micelles, sol-gel transition does not significantly influence its rotation. Micelle-micelle entanglement, which is responsible for gelation, persists even in the micellar solution phase, perhaps to a lesser degree, and this entanglement is responsible for the observed behavior. This hypothesis has been substantiated by undertaking concentration-dependent studies in which it is shown that the reorientation time of the solute increases with an increase in the micellar concentration. In the case of reverse micelles, it has been observed that an enhancement in the water content facilitates solute rotation, which has been rationalized on the basis of solute migration from the hydrated poly(ethylene oxide) region to the poly(ethylene oxide)-water interface within the core.  相似文献   

15.
Amphiphilic block copolymers,poly(ethylene oxide)-b-poly(N-acryloxysuccinimide) (PEO-b-PNAS) with various molecular weights have been successfully synthesized by atom transfer radical polymerization (ATRP) of NAS using functionalized PEO (PEO-Br) as ATRP macroinitiator.The self-assembling of the block copolymers in water,which is a good solvent for PEO and a non-solvent for PNAS.yielded spherical core-shell micelles with PNAS as core and PEO as shell.The cross-linked reaction of oxysuccinimide in PNAS ch...  相似文献   

16.
An out line and summary of literature studies on interactions between different types of amphiphilic copolymer micelles with surfactants has been given. This field of research is still emerging and it is difficult presently to make generalisations on the effects of surfactants on the copolymer association. The effects are found to be varied depending upon the nature and type of hydrophobic (hp) core and molecular architecture of the copolymers and the hydrocarbon chain length and head group of surfactants. The information available on limited studies shows that both anionic and cationic surfactants (in micellar or molecular form) equally interact strongly with the associated and unassociated forms of copolymers. The beginning of the interaction is typically displayed as critical aggregation concentration (CAC), which lies always below the critical micelle concentration of the respective surfactant. The surfactants first bind to the hydrophobic core of the copolymer micelles followed by their interaction with the hydrophilic (hl) corona parts. The extent of binding highly depends upon the nature, hydropobicity of the copolymer molecules, length of the hydrocarbon tail and nature of the head group of the surfactant. The micellization of poly(ethylene oxide) (PEO)–poly(propylene oxide) (PPO)–poly(ethylene oxide) was found to be suppressed by the added surfactants and at higher surfactant concentrations, the block copolymer micelles get completely demicellized. This effect was manifested itself in the melting of liquid crystalline phases in the high copolymer concentrations. However, no such destabilization was found for the micelles of polystyrene (PS)–poly(ethylene oxide) copolymers in water. On the contrary, the presence of micellar bound surfactant associates resulted in to large super micellar aggregates through induced intra micellar interactions. But with the change in the hydrophobic part from polystyrene to poly(butadiene) (PB) in the copolymer, the added surfactants not only reduced the micellar size but also transformed cylindrical micelles to spherical ones. The mixtures in general exhibited synergistic effects. So varied association responses were noted in the mixed solutions of surfactants and copolymers.  相似文献   

17.
We report our finding of an optimal length scale for toughening of epoxies using spherical micelles formed by block copolymers. The amphiphilic diblock copolymer poly(hexylene oxide)‐poly(ethylene oxide) (PHO‐PEO) with 30 wt % PEO self‐assembled to form spherical micelles in a bisphenol A epoxy resin with a phenol novolac hardener. We systematically increased the size of the spherical micelles from 20–30 nm to 0.5–10 μm by swelling their PHO core using PHO homopolymer. Although all the blends were tougher than the unmodified epoxy, the largest enhancement of fracture resistance was measured in blends containing 0.1–1 μm spherical inclusions. This enhanced toughness was correlated with plastic deformation by shear banding in tensile test and greater roughness of the fracture surface. Smaller micelles neither induced plastic deformation nor contributed to surface roughness significantly whereas larger micelles acted as local defects resulting in early failure. These findings provide a framework in assessing the toughening effects of blended block copolymers on epoxy resins. © 2009 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 47: 1125–1129, 2009  相似文献   

18.
A unique pH-dependent phase behavior from a copolymer micellar solution to a collapsed hydrogel with micelles ordered in a hexagonal phase was observed. Small-angle neutron scattering (SANS) was used to follow the pH-dependent structural evolution of micelles formed in a solution of a pentablock copolymer consisting of poly((diethylaminoethyl methacrylate)-b-(ethylene oxide)-b-(propylene oxide)-b-(ethylene oxide)-b-(diethylaminoethyl methacrylate)) (PDEAEM25-b-PEO100-b-PPO65-b-PEO100-b-PDEAEM25). Between pH 3.0 and pH 7.4, we observed the presence of charged spherical micelles. Increasing the pH of the micelle solution above pH 7.4 resulted in increasing the size of the micelles due to the increasing hydrophobicity of the PDEAEM blocks above their pKa of 7.6. The increase in size of the spherical micelles resulted in a transition to a cylindrical micelle morphology in the pH range 8.1-10.5, and at pH >11, the copolymer solution undergoes macroscopic phase separation. Indeed, the phase separated copolymer sediments and coalesces into a hydrogel structure that consists of 25-35 wt % water. Small-angle X-ray scattering (SAXS) clearly indicated that the hydrogel has a hexagonal ordered phase. Interestingly, the process is reversible, as lowering of the pH below 7.0 leads to rapid dissolution of the solid into homogeneous solution. We believe that the hexagonal structure in the hydrogel is a result of the organization of the cylindrical micelles due to the increased hydrophobic interactions between the micelles at 70 degrees C and pH 11. Thus we have developed a pH-/temperature-dependent, reversible hierarchically self-assembling block copolymer system with structures spanning nano- to microscale dimensions.  相似文献   

19.
Amphiphilic dendritic–linear–dendritic triblock copolymers based on hydrophilic linear poly(ethylene oxide) (PEO) and hydrophobic dendritic carbosilane were synthesized with a divergent approach at the allyl end groups of diallyl‐terminated PEO. Their micellar characteristics in an aqueous phase were investigated with dynamic light scattering, fluorescence techniques, and transmission electron microscopy. The block copolymer with the dendritic moiety of a third generation could not be dispersed in water. The block copolymers with the first (PEO–D ‐Si‐1G) and second (PEO–D ‐Si‐2G) generations of dendritic carbosilane blocks formed micelles in an aqueous phase. The critical micelle concentrations of PEO–D ‐Si‐1G and PEO–D ‐Si‐2G, determined by a fluorescence technique, were 27 and 16 mg/L, respectively. The mean diameters of the micelles of PEO–D ‐Si‐1G and PEO–D ‐Si‐2G, measured by dynamic light scattering, were 170 and 190 nm, respectively, which suggests that the micelles had a multicore‐type structure. The partition equilibrium constants of pyrene in the micellar solution increased with the increasing size of the dendritic block (e.g., 7.68 × 104 for PEO–D ‐Si‐1G and 9.57 × 104 for PEO–D ‐Si‐2G). The steady‐state fluorescence anisotropy values (r) of 1,6‐diphenyl‐1,3,5‐hexatriene were 0.06 for PEO–D ‐Si‐1G and 0.09 for PEO–D ‐Si‐2G. The r values were lower than those of the linear polymeric amphiphiles, suggesting that the microviscosity of the dendritic micellar core was lower than that of the linear polymeric analogues. © 2001 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 39: 918–926, 2001  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