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 0 毫秒
1.
The interactions between fluorocarbon‐modified poly (sodium acrylate) and various kinds of added surfactants have been studied by means of viscometric measurement. Association behavior was found in both hydrogenated and fluorinated anionic, nonionic and cationic surfactants. Among them, the interactions between fluorocarbon‐modified poly (sodium acrylate) and cationic surfactants are the strongest, owing to the cooperation of both electrostatic attractions and hydrophobic associations. The anionic surfactants have the weakest effects on the solution properties because of the existence of unfavorable electrostatic repulsion. The hydrophobic interactions between copolymers and fluorinated surfactants are much stronger than those between copolymers and hydrogenated surfactants.  相似文献   

2.
粘度法研究疏水改性聚丙烯酸与Np7.5的相互作用   总被引:4,自引:0,他引:4  
研究了氟基团改性和氟碳、碳氢基团同时改性的聚丙烯酸(HM-PAA)以及参与聚合物(PAA)与非离子表面活性剂Np7.5分别在稀溶液和亚浓溶液下的相互作用对特性粘数和Brookfield表观粘度的影响。结果表明,PAA与Np7.5无明显作用,而M-PAA在Np7.5达到一定浓度后,由于Np7.5参与了疏水缔合,特性粘数和表观粘度发生了明显变化。稀溶液中,氟碳基团较多的聚合物出现了链构象的伸展。亚浓溶液中,疏水改性聚合物的粘度都有先上升后下降的变化,但氟碳含量较多的变化更强烈。  相似文献   

3.
Hydrophobically modified poly(vinyl alcohol) (HMPVA) polymers were synthesized by potassium t‐butoxide‐catalyzed reaction of PVA with methyl 3,4,5‐tris(n‐octyloxy) benzoate (MGC8)/3,4,5‐tris(n‐dodecyloxy) benzoate (MGC12) and 1,3‐propane sultone. The concentration of 1,3‐propane sultone was kept constant at 10 mol % and that of MGC8 (2, 3, and 4 mol %)/MGC12 (2 and 3 mol %) was varied to obtain HMPVAs with different hydrophobic contents. The incorporation of MGC8/MGC12 and 1,3‐propane sultone onto HMPVA was confirmed by NMR spectroscopy. Rheological properties of aqueous solutions also confirmed the presence of hydrophobic and charged functional groups on HMPVAs. In the semidilute regime, the specific viscosity of HMPVAs followed concentration scaling that is typical of polyelectrolytes. At higher concentrations, the HMPVA solutions with 3 and 4 mol % of MGC8 exhibited large increase in specific viscosity. Oscillatory experiments on these solutions exhibited gel‐like behavior at polymer concentrations of 40–50 g/L. Confocal microscopy images of HMPVA with 4 mol % of MGC8 clearly indicated the existence of microgels. The tendency of formation of microgels further increased with increasing chain length of the hydrophobe, that is, with MGC12. These samples exhibited rheological behavior that is typical of soft solids and was therefore probed by the strain‐rate frequency superposition technique reported recently in the literature. HMPVAs with improved rheological properties show potential applications as thickeners in cosmetic creams, lotions and as drug carriers in pharmaceutical formulations. © 2010 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 48: 1054–1063, 2010  相似文献   

4.
The hydrophobic interaction between hydrophobically modified acrylamide copolymer (HMPAM) and poly(N‐isopropylacrylamide) (PNIPAM) in aqueous solutions was investigated. The results show that the solution properties of HMPAM are significantly influenced by the addition of PNIPAM. In dilute regime, the intrinsic viscosity of HMPAM in 0.025 wt % PNIPAM/0.1 M NaCl mixed solution is 17.52 dL g?1, about 2 times 8.66 dL g?1, that in 0.1 M NaCl solution, which is due to the attractive interaction between the hydrophobic parts of PNIPAM and HMPAM molecules. In semidilute regime, below the saturation concentration, the addition of PNIPAM can lead to both the apparent viscosity and the modulus of HMPAM solutions increasing, which is attributed to the number of aggregation junctions increasing, responsible for the increase of the contribution of the reversible network to the viscosity increase, the β value. In addition, a thermothickening behavior for the HMPAM/PNIPAM mixed solution is observed with increasing temperature over 15–30 °C, which is consistent with the large increase of the Huggins coefficient of HMPAM in the presence of PNIPAM from 1.95 to 7.59 as temperature increases from 25 to 30 °C. © 2005 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 43: 709–715, 2005  相似文献   

5.
The complex system of hydrophobically modified poly (acrylic acid) (HMPA) and wormlike micelles formed by sodium oleate in the presence of sodium acetate is investigated by apparent viscosity and rheological measurements. Addition of small amount of HMPA markedly enhances the viscosity of wormlike micelles, with further increase of HMPA concentration, a viscosity drop is observed. At high salinity, the complex system still remains strong viscoelastic and eliminates the limitation of susceptibility to salt for HMPA. DPD simulation results are in good agreement with experimental data, which provide theoretical confirmation for a synergistic mechanism between HMPA and wormlike micelles.   相似文献   

6.
A hyperbranched polyester was fractionated by precipitation to produce 10 fractions with molecular weights between 20 × 103 and 520 × 103 g mol?1. Each of these fractions was examined by size exclusion chromatography, dilute‐solution viscometry, intensity, and quasi‐elastic light scattering in chloroform solution at 298 K. High‐resolution solution‐state 13C NMR was used to determine the degree of branching; for all fractions this factor was 0.5 ± 0.1. Viscometric contraction factors, g′, decreased with increasing molecular weight, and the relation of this parameter to the configurational contraction factor, g, calculated from a theoretical relation suggested a very strong dependence on the universal viscosity constant, Φ, on the contraction factor. A modified Stockmayer–Fixman plot was used to determine the value of (〈r2o/Mw)1/2, which was much larger than the value for the analogous linear polymer. The scaling relations of the various characteristic radii (Rg, Rh, RT, and Rη) with molecular weight all had exponents less than 0.5 that agreed with the theoretical predictions for hyperbranched polymers. The exponent for Rg was interpreted as fractal dimension and had a value of 2.38 ± 0.25, a value that is of the same order as that anticipated by theory for branched polymers in theta conditions and certainly not approaching the value of 3 that would be associated with the spherical morphology and uniform segment density distribution of dendrimers. Second virial coefficients from light scattering are positive, but the variation of the interpenetration function, ψ, with molecular weight and the friction coefficient, ko, obtained from the concentration dependence of the diffusion coefficient suggests that chloroform is not a particularly good solvent for the hyperbranched polyester and that the molecules are soft and penetrable with little spherical nature. © 2003 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 41: 1339–1351, 2003  相似文献   

7.
The interactions between sodium dodecyl sulphate (SDS) and three autoassociative polymers (hydrophobically end-capped poly(ethylene oxide) urethanes) are studied by conductimetry, fluorescence, and viscometry. By the first two techniques the interactions are found to increase with the hydrophobicity of the polymer and to be stronger than those observed for a polyethylene oxide (PEO) of the same order of molecular weight. The results seem to be consistent with a model where SDS micelles are formed around hydrophobic chain ends in a first stage, and along the main chain in a second stage. For the more hydrophobic sample already self-aggregated in water and of relatively large reduced viscosity, the addition of SDS induces a loss in viscosity contrary to the less hydrophobic ones and PEO. Such effect may be attributed to the destruction of the chain end association by the formation of SDS micelles around one end in the first stage of the interaction.  相似文献   

8.
The dilute solution behavior of several alternating copolymers of maleic acid has been characterized by static and dynamic light scattering, intrinsic viscosity, and pulsed-gradient spin-echo NMR spectroscopy. The copolymer of maleic acid–sodium salt and isobutylene (IBMA-Na, Mw ∼350 kg/mol) dissolves readily in concentrated aqueous salt solutions. Changes in chain dimensions with ionic strength and pH are similar to those of the lesser salt solution-soluble poly(acrylic acid-sodium salt). The hydrophobically modified (with n-butyl, n-hexyl, n-octyl, and phenethyl amines) copolymers of maleic acid–sodium salts and isobutylene (IBMA-NHR-Na) show no sign of large intermolecular aggregation in 0.1 N sodium acetate (NaAc). However, the sizes of the copolymers are relatively small compared to that of the ionized parent copolymer (IBMA-Na, Mw ∼350 kg/mol), suggesting intramolecular aggregation of the alkyl side-chain groups along the polymer backbone. The copolymer modified with the longer chain n-decyl, on the other hand, forms stable large intermolecular aggregates containing 33 chains/aggregate. The copolymers of maleic acid–sodium salt and styrene (SMA-Na) appear to have no signs of aggregation, despite being a hydrophobic polyelectrolyte. The copolymer of maleic acid–sodium salt and di-isobutylene (DIBMA-Na) has a similar salting-out concentration as SMA-Na. The radius of gyration measurements by static light scattering suggest that at least some fraction of the DIBMA-Na chains form large intermolecular aggregates. The copolymers of maleic acid–sodium salt with n-alkenes (n-CmMA-Na) in 0.1 N NaAc form small intermolecular aggregates (three to five chains/aggregate). In contrast to these static light scattering results, PGSE NMR diffusion measurements for the above aggregated systems indicate only one diffusion coefficient consistent with the motion of single isolated chains. A plausible explanation for this discrepancy is that the population of the aggregates is too small to be sufficiently detected in the PGSE NMR experiment. Furthermore, it is likely that the aggregate has a larger relaxation rate than the nonaggregate, and therefore has a comparatively reduced signal in the PGSE NMR experiment. © 2004 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 42: 3584–3597, 2004  相似文献   

9.
The intrinsic viscosity [η] of dilute solutions of poly(1,4-phenylene terephthalamide) (PPPT) is found to depend strongly on sulfuric acid strength, exhibiting a maximum at about 100% H2SO4. This behavior instigated measurements of [η] and light scattering from dilute solutions of unfractionated PPPT in concentrated (≈96%) and 100% H2SO4. From [η] and weight-average molecular weight M w relationships, Mark-Houwink exponents a were determined to be 1.36 in 96.6% and 1.62 in 100.2 ± 0.2% H2SO4, indicating that the PPPT molecule can undergo considerable expansion in 100% H2SO4. For the case of 100% H2SO4, a noticeable polyelectrolyte effect is observed in the reduced viscosity versus concentration curves. This result suggests that the repulsive charges generated along the PPPT backbone may be responsible for the change in configuration of PPPT upon increasing the acid strength from 96.6% to 100% H2SO4. It is pointed out that there is considerable experimental difficulty in measuring consistent values of M w, and this may be the reason for the variation among published data.  相似文献   

10.
pt‐Butyl calix[4]arene diol (distal cone) (1) was grafted with poly (acrylic acid) (PAA) to obtain hydrophobically modified PAA (PAA‐C) bearing calixarene moieties. The grafting method includes the direct esterification reaction of PAA with calixarene diol 1 which was carried out in a system of tosyl chloride (TsCl), pyridine (Py), and N,N‐dimethylformamide (DMF). The grafting yield was studied using different molar ratios of PAA to calix[4]arene diol 1, temperature, and reaction time. The chemical composition of the PAA‐C was studied by IR and 1H NMR spectroscopy. Also, the morphology of PAA‐C was evaluated by scanning electron microscopy. The PAA‐C had different solubility and thermal properties. The extraction ability measurements of modified PAA toward alkali metal cations (Na+, K+, Cs+) and Ag+ showed a remarkable efficiency and selectivity of PAA‐C toward Na+. The main goal of this work was to design hydrophobically modified PAA with binding ability that is suitable for ion selective membranes and chemical sensor devices such as ion‐specific electrodes, semipermeable membranes, and quartz microbalances. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

11.
本文通过沉淀聚合制备了具有不同主链结构的疏水改性聚丙烯酸(HMPA),由Huggins方程确定了HMPA在溶液(水,盐/水,醇/水)中的特性粘数[η]和Huggins常数KH,采用流变学法研究了HMPA溶液的疏水缔合行为和流变特性.研究表明,HMPA溶液具有典型的剪切变稀行为,主链结构对HMPA溶液的缔合行为和流变特性有显著影响.主链含有疏水链段的HMPA在乙二醇/水溶液中形成类似弹性体的凝胶网络结构,具有较大的活化能,其表观粘度具有明显的温度敏感性.  相似文献   

12.
A series of hydrophobically modified polyacrylamide and polyacrylamide-co-poly(acrylic acid) gels with systematically varying hydrophobicity were prepared by free-radical polymerization of acrylamide, n-alkylacrylamides (n = 10, 12, and 14), and acrylic acid. The swelling of these gels was examined in water and in both anionic and cationic surfactant solutions. It was found that the gels which incorporated acrylic acid showed extremely high swelling in water. Maximum swelling was observed in gels which incorporated 10 mol% acrylic acid. The swelling of these gels was much less in solutions of both anionic and cationic surfactants than in water. The gels which did not incorporate acrylic acid demonstrated little swelling in water, but showed increased swelling in both anionic and cationic surfactant solutions with increased hydrophobicity of the gel. Received: 1 February 1999 Accepted in revised form: 5 March 1999  相似文献   

13.
The association of a series of hydrophobically modified poly(sodium acrylate) (HMPA) with lysozyme, a cationic globular protein, or with bovine serum albumin (BSA), an anionic globular protein, was investigated at pH=9 by rheology and to a lesser extent by steady-state fluorescence spectroscopy. Under suitable concentration conditions, this association leads to a drastic viscosity enhancement which is improved when the polymer hydrophobicity is increased. A mechanism is proposed: the hydrophobic regions of the globular proteins interact strongly with the alkyl groups of one or more polymer chains. In the later case, the macromolecules are crosslinked via the proteins, which leads to viscosity enhancement and even gelation. Analogies and differences between these systems and surfactant/HMPA systems previously studied in our laboratory are emphasized and discussed.  相似文献   

14.
We report on the synthesis and characterization of a novel hydrophobically modified end-capped poly(ethylene oxide)s. The end-capping agent of this polymer was designed and synthesised from a renewable resource material namely, gallic acid (i.e. 3,4,5-trihydroxybenzoic acid), the byproduct of tannin industry. The hydroxyl groups at 3, 4 and 5 positions of gallic acid provide an opportunity for varying the hydrophobicity of the compound. The hydrophobic end-capping compound, 3,4,5-tridodecyloxy bezoylazide was prepared from gallic acid and PEGs with different chain lengths (of number average molecular weights, 10000 and 35000 g/mol) were end-capped using 3,4,5-tridodecyloxybenzoyl azide. The quantitative analysis of end-capping in the polymers was demonstrated by 1H-NMR spectroscopy and the rheological studies were carried out in the surfactant solutions.  相似文献   

15.
Novel cationic polyelectrolytes with a brush-on-brush structure of poly(ethylene oxide) (PEO) side chains and a charge-containing polyacrylate backbone were synthesized. The PEO side chains were not directly attached to the backbone but via polymethacrylate spacers, thus locating the PEO chains a distance away from the charged units of the backbone. The cationic brush-on-brush polyelectrolytes with high density of PEO chains showed a strong affinity to silica surfaces, provided the backbone charge density was high enough. The adsorption of these polymers was studied by QCM-D giving very high sensed mass, 20 mg/m2. It was shown by direct force measurements that protective surface layers were formed by the novel polyelectrolytes, generating strongly repulsive steric forces, which provided an effective barrier against flocculation. The adsorbed layer was sufficiently robust to withstand sliding experiments under a pressure of up to 35 MPa. The friction force in water was very low, and the lubrication was characterized by a friction coefficient in the range of 0.02-0.06.  相似文献   

16.
A high molecular weight poly(N-acryloyl-m-aminobenzoic acid) (PMAB) was synthesized in a mixed solvent medium by radical polymerization. The polymer was separated into different fractions by fractional precipitation. All those eight fractions were characterized by viscometry, vapor pressure osmometry and gel permeation chromatography. Mark-Houwink-Kuhn-Sakurada relationships were established for three different solvents at 30°C. Unperturbed dimensions were estimated from the viscosity-molecular weight data using graphical methods based on approximate theories of Stockmayer and Fixman, and of Kurata, Stockmayer, and Roig, and on a first-order perturbation treatment of the excluded volume effect. The results indicate that the PMAB chain is of intermediate stiffness.  相似文献   

17.
Isotactic and syndiotactic poly(2-hydroxyethyl methacrylate) (PHEMA) have been prepared. Intrinsic viscosity–molecular weight relationships were established for the isotactic and syndiotactic PHEMA in N,N-dimethylformamide (DMF) at 25°C by solution viscometry and light scattering. The unperturbed dimensions and interaction parameters were examined in DMF, water, methanol, ethanol, and water–methanol (1:7 by volume) mixture for isotactic PHEMA and in DMF, methanol, and water–methanol (1:7 by volume) mixture for syndiotactic PHEMA using the Stockmayer–Fixman representation. The results suggest that the compact random coil structure for isotactic PHEMA occurs in water solvent and the isotactic PHEMA is more highly extended in polar solvents.  相似文献   

18.
Kuhn-Mark-Houwink-Sakurada relations were obtained in methyl ethyl ketone N,N-di-methylformamide and dichlorethane at 30 for (styrene acrylonitrile) alternating copolymer. The values of the unperturbed dimensions. [K0 or (ro?2/M)1/2] and conformational parameter σ have been determined, using several graphical and semiempirical methods, and the results were compared with the direct determinations in a θ solvent. The best values for Kθ were obtained using the methods of Stockmayer-Fixman and Inagaki-Suzuki-Kurata. By comparing the values of σ for polystyrene, polyacrylonitrile, random and alternating (styrene-acrylonitrile) copolymers, it is to be concluded that the short-range interactions do not markedly influence the chain dimensions in solutions for random and alternating (styrene-acrylonitrile) copolymers.  相似文献   

19.
A sequence-ordered, periodic copolymer of ethylene, ethylene, and acrylic acid, poly (ethylene-per-ethylene-per-acrylic acid) (PEEA), with M w=1.44×105 has been synthesized by alternating copolymerization of 1,3-butadiene and methyl acrylate, followed by hydrogenation and hydrolysis. Aqueous solution and dissociation properties of the alkali-soluble PEEA were explored by potentiometric titration and intrinsic viscosity at 25 °C. The pH values of PEEA were almost constant (pH = 6.48 ∼ 6.55) with an increasing degree of dissociation (α) from 0.3 to 0.8 at C s=50 mN NaCl. Correspondingly, the plots of negative logarithm of apparent dissociation constant (pK a) against α showed a reversed S-shape curve over the whole α, indicating an extensive precipitation and subsequent tran-sition from compact to coiled conformation. The intrinsic viscosity steeply increased with α above 0.4 up to 9.97 dl/g at α = 1.0. Good agreement between the observed electrostatic potential and that calculated from the rod model with a smeared charge density was observed in the region of α higher than 0.9. The dissociation and dissolution processes of PEEA with neutralization in water were described. Received: 14 April 1998 Accepted: 3 June 1998  相似文献   

20.
A four-branched cage-shaped poly(ethylene oxide) (4C-PEO) was prepared by a coupling reaction at very dilute condition between two kinds of end-functional four-armed star-shaped PEOs having amino and N-hydroxysuccinimide groups on their four ends, followed by purification using α-cyclodextrin (α-CD). The raw coupling reaction product shows multiple peaks including various high molecular weight multimeric products in size-exclusion chromatography measurements, while the product after α-CD purification shows a single peak eluted slightly earlier than the precursor star PEOs. Moreover, the final product obtained has about twice higher molecular weight than the star precursors. These results suggest that the targeted 4C-PEO polymer was successfully isolated through the α-CD purification. Small-angle neutron scattering (SANS) measurements of the final product in dilute solution were conducted, and its chain conformation was evaluated from the scattering profile in comparison with linear and star polymers. It has been found that the cage-shaped sample exhibits a distinct peak at lower q-region in the Kratky plot, which is in strong contrast with the linear counterpart. This result must be originated from the characteristic cage-shaped architecture, that is, having branched and closed-loop configurations, and hence having higher segmental density than simple linear and star molecules. In fact, the present experimental result is consistent with the recent Monte Carlo simulation reported by Uehara and Deguchi.  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