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 156 毫秒
1.
季铵盐型双子表面活性剂16-4-16聚集状态的NMR研究   总被引:1,自引:0,他引:1  
核磁共振弛豫,自扩散以及2D NOESY谱研究结果表明:双子表面活性剂16-4-16溶液在形成胶束的过程中,联结基团及其邻近的碳氢链质子形成胶束的壳层,而距离离子头较远的疏水质子位于胶束的内部. 与对应的单链的表面活性剂CTAB相比,其分子运动更受限制. 2D NOESY谱显示联接基团及临近的碳氢链的质子间有较强的交叉峰,表明形成胶束时,分子在联结基团附近堆积的较为紧密. 由2D NOESY谱计算得到的质子间距与HYPERCHEM模拟值有偏差,表明这些强交叉峰是分子间相互作用的结果,并且对应质子对在双子表面活性剂16-4-16分子中位于邻近的区域. 因此我们推测,双子表面活性剂16-4-16分子在球形胶束中形成特殊的排列方式.  相似文献   

2.
双子表面活性剂的NMR研究现状与展望   总被引:1,自引:1,他引:0  
双子表面活性剂是一类新型的表面活性剂, 其分子由化学键通过联接基团将两个单链表面活性剂的离子头联接构成. 由于其结构的特殊性, 使其具有特殊的性质, 如高表面活性, 特殊的聚集形态等. 多年来的研究结果表明, 联接基团的结构和长度对双子表面活性剂的活性影响最大. 与其它研究手段相比, NMR在从分子水平上研究双子表面活性剂溶液中胶束的形成以及自聚集状态, 有着独特的优势.  相似文献   

3.
应用核磁共振技术,对几种典型的表面活性剂在水溶液中的聚集行为、结构特征、动力学特性和相互作用等进行了研究. 
利用1D 1H NMR方法测得4-癸基萘磺酸钠(SDNS)在313 K温度时的临界胶束浓度(CMC)在0.82~0.92 mmol/L之间,与报道的298 K时的CMC范围相同. 弛豫时间和2D NOESY实验结果表明,与298 K时的SDNS胶束相比,313 K温度时,SDNS胶束中烷烃链排列得更紧密,其中与萘环相连的第一和第二个亚甲基参与了胶束紧密层的形成,更紧密地堆积在萘环之间. SDNS质子T2值随温度的变化表明,在单体和胶束两种状态下,质子运动对温度的敏感性明显不同. 由自扩散系数分析得到,SDNS胶束的水合半径约为其单体水合半径的5.3倍. 而在十二烷基磺酸钠(SDSN)胶束中,由于静电排斥力的作用,在同样温度下SDSN的胶束紧密层排列比SDNS更疏松. 
NMR实验表明,在SDNS/Triton X-100 (TX-100)和SDNS/SDSN体系中形成了混合胶束. 在SDNS/TX-100混合胶束中,TX-100的苯环靠近SDNS的烷烃链,而它的聚烷氧链除与苯环相连的第一个乙氧基基团以外都被限制在SDNS的萘环附近. 在SDNS/SDSN混合胶束中,SDSN的磺酸基比SDNS分子更靠近胶束内部. 而SDNS的萘环将SDSN的磺酸基分隔开,在降低带负电荷的磺酸基极性头之间的静电排斥力中起到了积极的作用,有助于混合胶束的形成.
从自扩散系数、横向弛豫和质子距离等NMR测定参数推测,在浓度为0.26 mmol/L(318 K)的N,N′-双(十六烷基二甲基)-α,ω-丙烷溴化铵(16-3-16)溶液中形成了近似球形的胶束,胶束表面的带正电荷的铵基极性头呈锯齿状排列以减弱分子间静电排斥力的影响. 弛豫时间测定表明,与N,N′-双(十六烷基二甲基)-α,ω-丁烷溴化铵(16-4-16)相比,16-3-16在胶束表面的spacer链段更僵硬, 在胶束核区的烷烃侧链排列的更紧密. NMR共振峰的线形分析表明,16-3-16和16-4-16侧链末端的甲基在胶束中位于两个不同的位置.   相似文献   

4.
本文应用DNMR的方法测量了双子表面活性剂12-s-12和14-s-14(s=2, 3, 4)单体在溶液和胶束间的分子交换动力学参数,根据DNMR理论的简化公式,变温条件下测量了表面活性剂分子的逸出速率常数(k~-),根据阿伦尼乌斯公式拟合得到了逸出活化能(E_a~-).结果显示14-s-14和12-s-12的逸出速率常数分别在10~1和10~3 s~(-1)数量级,逸出活化能分别为54.04~73.64 kJ/mol和33.42~47.09 kJ/mol.随着联结基团长度的增大,14-s-14和12-s-12的逸出速率常数增大,逸出活化能减小.结合荧光探针实验测量微极性的结果,揭示了双子表面活性剂分子从胶束中逸出时经历了构象变化,支持了双子表面活性剂的两步分子交换机制.  相似文献   

5.
采用水溶液均聚合方法,制备了阳离子型表面活性单体(2-丙烯酰胺基)乙基十四烷基二甲基溴化铵(AMC14AB)的均聚物,使用荧光探针法、表面张力测定及电导测定法,重点考察了均聚物P(AMC14AB)在水溶液中的胶束化行为与表面吸附现象. 在水溶液中,均聚物P(AMC14AB)呈现单分子链胶束的聚集形态,具有零临界胶束浓度,从开始加入P(AMC14AB)起,水溶液中随即产生单分子链胶束,不存在Krafft温度. P(AMC14AB)在溶液表面也发生表面吸附,使水的表面张力下降,即P(AMC14AB)也具有表面活性;随着浓度增大,表面吸附量增大,水的表面张力持续下降;当表面吸附达饱和时,表面张力~浓度曲线上出现突变点,该点定义为饱和的表面吸附浓度,而不应该再称为临界胶束浓度. P(AMC14AB)单分子链胶束溶液对疏水有机物(甲苯)的增溶情况,明显不同于普通小分子表面活性剂十六烷基溴化铵(CTAB)的多分子胶束溶液,甲苯增溶量~P(AMC14AB)浓度的关系曲线上无突变点,而且对甲苯的增溶能力高于CTAB的多分子胶束溶液.  相似文献   

6.
林元  李干佐 《波谱学杂志》1990,7(3):297-304
本文用高分辨~1H.NMR谱研究了多种短链芳烃及其衍生物在阳离子表面活性剂—十六烷基三甲基溴化铵(CTAB)水溶液胶束和非离子表面活性剂—月桂基聚氧乙烯(4)醚(Brij30)非水溶液胶束中的增溶作用。得到这些不同种类的增溶物在两种不同类型的胶束中增溶位置的详细结果。并讨论了增溶物与表面活性剂分子之间的相互作用。这一研究对深入探讨增溶机理十分有益。  相似文献   

7.
表面活性离子液体(SAILs)兼备离子液体和表面活性剂的性质~([1])。将其特性引入到传统的有序分子聚集体如PEO-PPO-PEO中,不但有助于改善聚集体的性质,还有望形成形态新颖的聚集体结构。因此,离子液体参与构筑的有序分子聚集体及其与两亲性分子的相互作用引起人们的日益关注~([2-3])。本工作C_8mimBr为SAILs代表,采用NMR和DLS研究了其与F127的相互作用。研究发现,F127的CMT随C_8mimBr的添加而降低,且CMT的下降程度紧密依赖于C_8mimBr的添加量。但是,当溶液中C_8mimBr浓度超过其自身CMC时,F127胶团解聚并分散于以C_8mimBr为主的胶团中。NOESY研究发现F127的EO链段的醚氧原子与C_8mimBr的咪唑环间存在氢键作用,PO链段与C_8mimBr的烷基链间存在疏水相互作用。  相似文献   

8.
通过紫外-可见光谱法研究了阴离子偶氮染料甲基橙(MO)与3种季铵盐表面活性剂之间的相互作用.当表面活性剂浓度远低于临界胶束浓度(CMC)时,表面活性剂与染料形成聚集体,反映为吸收带蓝移并伴随着吸收强度下降.继续加入表面活性剂,引起染料在阳离子胶束中的特征吸收.当表面活性剂浓度达到CMC及以上时,混合溶液的λ<,max>...  相似文献   

9.
为探究不同促进剂在甲烷水合物生成过程的微观作用机理,选取动力学促进剂十二烷基硫酸钠(SDS)和热力学促进剂十六烷基三甲基溴化铵(CTAB)作为添加剂,采用分子动力学方法研究其对甲烷水合物生成速率的影响.通过分析势能变化、均方位移、径向分布函数、分子簇生长速率,发现质量分数为0.9%SDS、1.2%SDS、1.2%CTAB、1.6%CTAB的溶液均可促进水合物生成.质量分数为1.2%的SDS溶液水合物生长速率最快,且SDS促进效果优于CTAB.通过分析甲烷分子密度分布云图,发现呈阴性的SDS分子头部基团吸附了大量甲烷分子,水分子受挤压向中间聚集;CTAB含氮的头部基团朝向均相溶液,包含在不稳定的水合物笼中,形成半笼型水合物.相比之下,CTAB溶液中水合物含气率更高.  相似文献   

10.
季铵盐型双子表面活性剂在空气-水界面的单分子膜   总被引:1,自引:0,他引:1  
利用Langmuir膜天平测定不对称双子表面活性剂12-2-16和对称双子表面活性剂12-2-12在空气-水界面上形成单分子膜的表面压,通过单分子膜的π-A等温线、微分等温线和静态弹性,分析其相转变和成膜特性.结果表明,与传统阳离子表面活性剂DTAB比较,12-2-16、12-2-12比DTAB单分子膜的分子极限面积大,崩溃压高,并有较高的静态弹性,因而具有较高的凝聚性和稳定性,其中12-2-16单分子膜的凝聚性和稳定性最高.  相似文献   

11.
The aqueous solubility enhancement of the polycyclic aromatic hydrocarbons (PAHs) naphthalene, anthracene and pyrene by micellar solutions of single gemini surfactant hexanediyl-1,6-bis(dimethylcetylammonium bromide) (G6) and its mixtures with cationic cetyltrimethylammonium bromide (CTAB), anionic sodium bis(2-ethylhexyl)sulfosuccinate (AOT) and nonioinic polyoxyethylene (20) cetyl ether (Brij 58) have been investigated. Above the cmc, maximum solubilization occurs in the Brij 58 surfactant micelles whereas the solubilization is least in presence of AOT. The PAHs are solubilized synergistically in mixed gemini-conventional surfactant solutions, which is attributed to the formation of mixed micelles, their lower cmc values, and the increase of the solvents' molar solubilization ratios or micellar partition coefficients because of the lower polarity of the mixed micelles.  相似文献   

12.
The viscosity and morphology of gemini surfactant in aqueous solution at various temperatures are investigated by rheometer and transmission electron microscopy with a freeze fracture replication technique. The viscosity of 18-3-18 or 16-3-16 reaches its maximum when the solution temperature is near the Krafft temperature of the surfactant. The morphology of a gemini surfactant in aqueous solution is strongly influenced by the temperature: vesicles or sphere-like aggregates are formed in the dilute solution of 18-3-18 under 25 °C environment; when the solution temperature goes up to 50 °C, such aggregates will transit to worm-like micelles with the length of each from 100 nm to 400 nm; meanwhile, the increase of concentration may lead to the formation of network; however, the worm-like micelles or network will transit to vesicles or sphere-like aggregates again when the temperature reaches 80 °C.  相似文献   

13.
Electrodeposition of cobalt on a copper electrode was successfully performed from aqueous and reverse micellar solutions of a cationic surfactant, cetyltrimethylammonium bromide (CTAB), using constant potential electrolysis method. The potential to be applied for electrodeposition was judged from the cyclic voltammetric behavior of cobalt(II) in aqueous and reverse micellar solutions of CTAB at different compositions. The morphology, dimension, and crystallinity of cobalt deposited onto a copper substrate were evaluated from scanning electron microscopy (SEM) images and X-ray diffraction technique. The cobalt deposited on copper from aqueous solution does not show any definite shape and size, while the deposition from reverse micellar solutions occurred with definite shapes such as star-, flower-, and nanorod-like structures depending on the composition. The slow kinetics governed by the reverse micelles associated with the deposition brings about oriented growth of cobalt onto the copper substrate and offers the potential to electrochemically tune cobalt deposit with desirable morphology.  相似文献   

14.
The aqueous solutions of sodium cumene sulfonate (NaCS) and its mixtures with cetyl trimethylammonium bromide (CTAB) and sodium dodecyl sulfate (SDS) are studied by Small Angle Neutron Scattering (SANS), Fourier Transform Infrared (FTIR) spectroscopy and Nuclear Magnetic Resonance (NMR) spectroscopy. The compositions of mixed micelles are determined using Rubingh's Regular Solution Theory. NaCS when added to CTAB solution leads to the formation of long rod shaped micelles with dramatic increase in the CTAB aggregation number. Its addition to SDS on the other hand results in the formation of smaller mixed micelles where parts of SDS molecules in the micelle are replaced by NaCS molecules. NaCS–SDS mixed micelles prefer elongated ellipsoidal geometry in order to accommodate short NaCS molecules. The FTIR spectroscopy results indicate enhanced ordering of CTAB tails inside the NaCS–CTAB mixed micelles with reduction in the gauche/trans conformer ratio. Addition of NaCS to SDS on the other hand results in decreased ordering of SDS tails, as compared to SDS micelles alone. The chemical shifts observed in 1H NMR spectra of NaCS–SDS and NaCS–CTAB mixture indicate that NaCS resides near the surface of the SDS micelle.  相似文献   

15.
The aqueous solutions of sodium cumene sulphonate (NaCS) and its mixtures with each of cetyl trimethylammonium bromide (CTAB) and sodium dodecyl sulphate (SDS) are characterized by small angle neutron scattering (SANS). NaCS when added to CTAB solution leads to the formation of long rod-shaped micelles with a dramatic increase in the CTAB aggregation number. Its addition to SDS on the other hand results in the formation of smaller mixed micelles where part of SDS molecules in the micelle is replaced by NaCS molecules.   相似文献   

16.
In this paper we report the effect of dicationic ‘gemini’ surfactants (CH3)2C16H33N+? (CH2)m? N+C16H33(CH3)2, 2Br? (where m = 4, 5, 6) on the reaction of ninhydrin with DL ‐tryptophan. The gemini surfactant micellar media are comparatively more effective than their conventional monomeric counterpart cetyltrimethylammonium bromide (CTAB) micelles. Also, whereas typical rate constant (kψ) increase and leveling‐off regions, just like CTAB, are observed with geminis, the latter produce a third region of increasing kψ at higher concentrations. These subsequent increases are ascribed to changes in micellar morphologies, consistent with changes in 1H NMR line widths. Quantitative kinetic analysis of the rate constant–[surfactant] data has been performed on the basis of modified pseudophase model. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

17.
The use of mixed surfactants in the synthesis of mesoporous silica nanoparticles (MSNs) is of importance in the context of adjusting pore structures, sizes and morphologies. In the present study, the arrangement of molecules in micelles produced from a mixture of two surfactants, cetyltrimethylammonium bromide (CTAB) and cetylpyridinium bromide (CPB) was detailed by solid-state NMR spectroscopy. Proximities of methyl protons in the trimethylammonium headgroup of CTAB and protons in the pyridinium headgroup of CPB were observed under fast magic angle spinning (MAS) by (1)H-(1)H double quantum (DQ) MAS NMR and NOESY. This result suggested that CTAB and CPB co-exist in the pores without forming significant monocomponent domain structures. (1)H-(29)Si heteronuclear correlation (HETCOR) NMR showed that protons in the headgroups of CTAB are in closer proximity to the silica surface than those in the CPB headgroups. The structural information obtained in this investigation leads to better understanding of the mechanisms of self-assembly and their role in determining the structure and morphology of mesoporous materials.  相似文献   

18.
Adsorption of a series of cationic gemini surfactants 12-2-m (m = 8, 12, 16) on the surface of silica was investigated. The critical micelle concentrations, cmcs, of cationic gemini surfactants in the initial solutions and in the supernatants were measured by conductometry and tensiometer. The changes in cmc values indicate that the ion exchanges take place between polar groups of gemini surfactants adsorbed and ions bound on the surface of silica. The adsorption isotherms of cationic gemini surfactants were obtained by a solution depletion method. Based on the driving force, the adsorption includes two steps, one of which is ion exchange, and the other is hydrophobic interaction. In each step, the tendency of surfactant molecules in the solution to form aggregates or to be adsorbed on the silica varies with their structures. The maximum adsorption amount of gemini surfactants on the silica, τmax, decreases as increasing in the length of one alkyl chain, m, from 8, 12 to 16. So the results show that the adsorption behaviors of gemini surfactants are closely related to the dissymmetry of gemini molecules.  相似文献   

19.
Fluorescence quenching measurements were performed on aqueous solutions of the cationic surfactant cetyltrimethylammonium halide (CTAX) and two bile salts, sodium cholate (NaC) and sodium deoxycholate (NaDC), to study the state of aggregation in the mixtures. Pyrene was used as a photoluminiscence probe in the study, and dimethylbenzophenone (DMBP) as the quencher. Analysis of time-resolved decay data with and without quencher using a simple kinetic model gave information of the different aggregation characteristics in the above two cases. Mixed micelles of CTAX/NaC were small and spherical at all compositions, while those of CTAX/NaDC tended to grow from spherical micelles to larger rod-like mixed aggregates at equimolar and close-to-equimolar concentrations. In the latter case more complex kinetics ensues and the fluorescence decays were treated using a generatized model for diffusion-controlled quenching along one dimension for infinitely long rod-like micelles. The mutual diffusion coefficient for the probe-quencherpair was determined.  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