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 31 毫秒
1.
Interaction between amphiphiles and water molecules in micelle or bilayer structure has been investigated using aqueous colloids of various amphiphiles through the rheological data and the spin-lattice relaxation timeT 1 of the proton of water molecule.T 1 of the water proton has been measured by the inversion recovery method and determined as a single exponential relaxation process.The chemical shift of the water proton is almost independent of the amphiphilic concentration; however, it shifts toward a higher magnetic field with increasing temperature in a way similar to that in pure water and in the amphiphilic aqueous systems. These facts mean that there is no significant difference in the magnetic field environment of the water protons in these systems.The water molecule is not necessarily bound in the fully developed micelle or bilayer (rod-like or lamella) structure which induces the high viscosity or high rigidity of the colloidal system. On the other hand, the water molecule is bound in the micelle colloids of amphoteric amphiphiles or amphiphiles whose molecular assembly creates a relatively strong electrostatic field. The activation entropy of the bound water is negative and this suggests that water molecules assume some ordered structure in the bound state.  相似文献   

2.
Strips of gelatin have been prepared by extrusion at different water contents varying from 20 to 50% H2O (dry weight basis, d.w.b.). The processes of subsequent hydration or dehydration of these strips were followed by dynamic mechanical thermal analysis (DMTA), wide-angle X-ray diffraction and NMR relaxation measurements. A comparison of the calculated dependence of theT g of gelatin (T g anhydrous, 200?C) on water content (using the Ten Brinke and Karasz equation) with experimental results derived from DMTA showed that freshly extruded material followed the theoretical plot below 25% H2O (d.w.b.), but at higher water contents, the7 g deviated positively, probably due in part to the effect of delayed re-equilibration of water content after thawing of separated ice crystals. The experimental results determined after storage for one week fell on a different line, with aT g of 145?C for anhydrous gelatin Possibly, theT g is elevated by crystallization — a view supported by the WAXS spectra. The NMR relaxation results also showed a profound mobilization of the gelatin protons at water contents greater than 25% d.w.b.  相似文献   

3.
Differential-scanning-calorimetry was applied to study the lyotropic and thermotropic properties of the two ternary systems dimyristoylcephaline (di-(C14:0)-PE)/Palmitic acid (C15COOH)/water (H2O) and dimyristoylcephaline (di-(C14:0)-PE)/palmitic acid methyl ester (C15COOMe)/water (H2O) in dispersions with excess water (50 wt.%).The phase diagrams of both systems showed that the two systems differ in their miscibility behavior. The system di-(C14:0)-PE/C15COOH/H2O is completely miscible in its high-temperature phase. In the low-temperature phase the mixing gap was found within the concentration range of C15COOH and was also indicated by a maximum value of the transition enthalpy of the pseudo-binary mixtures.In the pseudo-binary system di-(C14:0)-PC/C15COOMe/H2O, the tendency towards demixing is much more pronounced. It was observed that the incorporated C15COOMe melted above its normal melting point, but below the transition temperature of di(C14:0)-PE/H2O system; therefore, the phase transition started at lower temperature. In the low-temperature phase, both lipids are partially miscible. The demixing range of the phase diagram lies within the concentration region of C15COOMe. Up to the mole fraction ofXC15COOMe=0.43, C15COOMe can be incorporated into theL-phase of the system di-(C14:0)-PE/H2O.  相似文献   

4.
The variation of amorphous orientation and crystalline regularity of hard elastic polypropylene (HEPP) films during cyclic deformation and stress relaxation processes were studied using a FTIR spectrometer. The result proves entropic elasticity and shows the orientational hysteresis in the amorphous region or within the microfibrils, and also shows that the amorphous orientation increases, but that the crystalline regularity decreases with the increase of extension rate.Three spin-spin relaxation timesT 2f,T 2m, andT 2s and associated mass fractionsF f,F m, andF s of HEPP fibers were measured with a solid echo of NMR method at different elongations and after relaxation or recovery for a long time A new possible interpretation was proposed that, while the microfibrils are formed in HEPP, the medium decay component should be ascribed to inner molecules of the microfibrils, and the slow decay component to the surface molecules of the microfibrils. According to this interpretation, the results implied that subfibrillation is the main process when HEPP is stretched up to 15% strain, and that at above 15% strain thinning and lengthening of the microfibrils become the main process. Thickening of the microfibrils was found in the recovery and relaxation processes.  相似文献   

5.
Two polystyrenes terminated withp-cyanobenzyl andp-cyanobiphenyl groups (=labels) were prepared and their complex dielectric constants were measured in the glass transition region in the frequency range 10–2–106 Hz. The glass temperaturesT g (DSC) were considerably different, 92 and 97.5°C, resp., although their molecular weights were very similar (11 000 and 10 000 g/mol, resp.). Their relaxation behavior showed that the cyano groups relax cooperatively with the polymer segments. The cyanophenyl groups were found to relax with shorter relaxation times than the cyanobiphenyl groups. The measured relaxation strengths showed that there was no association between the dipoles. The relaxation mechanisms of the cyano groups in both labels seemed to be different although the only difference between them was an additional phenyl group in the case of the second label.  相似文献   

6.
An isotactic polypropylene film was stretched at 120 °C in poly(ethylene glycol) and thermally shrunk at various temperatures. Proton spin-lattice,T 1, and spin-spin,T 2, relaxation times were measured using a broad line pulse spectrometer operating at 19.8 MHz in the temperature range 40 °C–100 °C. The temperature ofT 1 minimum shifts to higher temperatures and the value ofT 1 minimum increases in magnitude as the stretching ratio is increased. In contrast the temperature ofT 1 minimum shifts to lower temperatures as shrinkage is increased, whereas the value ofT 1 minimum increases in magnitude because of the increase in crystallimty during shrinkage. T2a, the longestT 1 associated with the mobile amorphous regions, increases during shrinkage, indicating that chain mobility in the amorphous regions increases substantially during shrinkage. It was found that an orientation function of the amorphous regions,f a, correlates well withT 2a .Presented in part at the 52nd Annual Meeting of the Japan Chemical Society, Kyoto, April 1986.  相似文献   

7.
Polymers having phosphoric acid groups were prepared as a model binder for magnetic coatings, and the correlation among the adsorption behavior of the polymers onto-Fe2O3 particles and the dispersibility, the orientation, and the packing density of-Fe2O3 particles in the magnetic coatings was investigated.PMMA homopolymer molecules hardly adsorbed on-Fe2O3, and the interfacial tension at a water/polymer solution (toluene) interface ( W/T) was scarcely changed compared with a water/toluene interface. Increasing with the content of polymeric phosphoric acid group, the adsorbance of polymer increased and the interfacial tension ( W/T) decreased. When the content of polymeric phosphoric acid groups was over 0.4 mol%, the adsorbance of polymer and interfacial tension ( W/T ) remained constant. When these polymers were used as a binder for magnetic tapes, the dispersibility of-Fe2O3 in the magnetic coatings was improved, increasing with the content of polymeric phosphoric acid group; however, when the content of phosphoric acid group was over 0.2 mol%, its dispersibility decreased abruptly.Studies on Recording Magnetic Materials and Magnetic Composite. XVIII.  相似文献   

8.
The phase transitions and proton dynamics of Cs5H3(SO4)4·0.5H2O single crystals were studied by measuring the NMR line shape, the spin-lattice relaxation time, T1, and the spin-spin relaxation time, T2, of the 1H and 133Cs nuclei. The “acid” protons and the “water” protons in Cs5H3(SO4)4·0.5H2O were distinguished. The loss of water protons was observed above TC1, whereas the content of water protons was found to recover above TC2. Therefore, the water protons play a special role in the stability of the superprotonic phase at high temperatures. The mechanism of fast proton conduction was found to consist of hydrogen-bond proton transfer involving the breakage of the weak part of the hydrogen bond and the formation of a new hydrogen bond. Thus, these structural phase transitions probably involve significant reorientation of the SO4 tetrahedra and dynamical disorder of the hydrogen bonds between them.  相似文献   

9.
A basic requirement for that type ofL 2-phase which exists in the system sodium octanoate-octanoic acid-water is the formation of acid-soaps. In order for the phase to be formed at all, the temperature must lie above the melting point of the fatty acid so that a reaction in non-aqueous milieu between neutral soap and fatty acid is possible. In order to obtain the characteristic shape and complete extension of the phase in direction of high water content the temperature must be so high that also the hydrated acid-soaps occur in fluid state. On the other hand the temperature cannot be so high that the acid-soaps become unstable.At temperatures at which the phase has obtained its full extension those circumstances differs which in different regions regulate the location of the phase borders; they depend on the composition of the acid soaps and on their amounts. In that part of the phase where the molar ratio between octanoic acid and sodium octanoate lies between 2 and 3 and where one has a continuous transition from reversed to normal structure only the two acid octanoates 1 NaC8 2 HC8 x H2O and 1 NaC8 3 HC8 x H2O occur and both are at 20 °C in fluid state.At water contents from about 22 % to 40 % the hydrate-water molecules belonging to the first mentioned soap are capable of contributing actively to the formation of large aggregates of acid-soap, a process which however is counteracted by the inmixing of the latter acid-soap. This mixture of the two acid-soaps decides in this region where the border of the phase will lie in direction towards an increased content of sodium octanoate; the result is that in spite of the fact that the hydration is increased, the border is only slowly displaced towards a higher content of fatty acid. As soon as the hydration of the acid octanoates has been completed and the additional water occurs as unbound bulkwater, the location of the phase boundary will no longer be influenced by the water content — now it will be the amphiphilic composition of the acid-soaps that determines the location of the border and it remains at the molar ratio 2.5 between octanoic acid and sodium octanoate at water contents from about 40% and up to 82%.In the direction of decreasing content of neutral sodium octanoate and increased content of water theL 2-phase both at the highest content of fatty acid and the highest contents of water will be in equilibrium with the water-richL 1-phase; in the first mentioned region with theL 1-phase below the lac where at the border it is saturated with octanoic acid and in the latter region with theL 1-phase just above the lac, where the dilute sodium octanoate solution contains dissolved 1 NaC8 1HC8 x H2O. In the large central part of theL 2-phase, from about 20 % to about 86 % of water, the location of the border is dominated by the acid octanoate 1 NaC8 3 HC8 x H2O and that makes an equilibrium with theL 1-phase impossible; instead one has an equilibrium via a two-phase zone between the amphiphile-rich region of theL 2-phase and its water-rich region. In the first region the location of the border is regulated by the decreasing capability of the hydrated acid octanoate 1 NaC8 3 HC8 x H2O to dissolve octanoic acid; in the latter it is regulated by the fact that 1 NaC8 3 HC8 x H2O is the most fatty acid-rich acid-soap that is formed and that the octanoic acid is very little soluble in water and in the aqueous solution of this acidsoap.The middle part of theL 2-phase, especially the region between about 55 % and 82 % of water, constitutes a direct continuation of the liquid crystalline lamellarD-phase. The liquid crystalline character of theD-phase is lost at the transition, but the lamellar organization is retained. That the molecules at least up to a water content of about 40 % are of the original reversed type and have an elongated shape with a central part of hydrated polar groups, from which core the hydrocarbon chains extend in two opposite directions, is the reason to that they, at crowding, form transient layer-like agglomerates of tightly packed more or less parallel molecules; this facilitates the transformation to coherent double amphiphilic layers, in which all molecules lie with the hydrated polar groups outwards toward coherent domains of bulk-water, without another liquid phase occurs.  相似文献   

10.
Dynamic and static light scattering techniques were used to study the droplet size and the interdroplet interaction of w/o microemulsions consisting of cetyltrimethyl-ammonium bromide (CTAB), hexyl carbitol, toluene, water and poly(ethylene glycol). The results were analyzed in terms of a hard-sphere model with a perturbation. For the microemulsions without polymer, their droplet sizes increased only slightly (R=10.1 to 11.0 nm) and the perturbation became more attractive as the molar ratio of H2O/CTAB was raised from 50 to 82. In contrast, an increase in polymer concentration or polymer molecular weight not only increased the droplet sizes but also changed the perturbation to become more repulsive. In addition, it is envisaged that the interactions between the cationic groups of CTAB and the ether linkages of the poly(ethylene glycol) may also enhance the rigidity of the interfaces, hence the stability of the microemulsions.  相似文献   

11.
Crystalline silicic acids are prepared from alkali layer silicates by exchanging protons for the alkali ions. The acid H2Si20O41 · xH2O (parent material K2Si20O41 · xH2O) exhibits some outstanding gas adsorption properties which are related to the layer structure and the interlamellar microporosity. The external surface, about 20 m2 g–1, is estimated from nitrogen adsorption data after blocking the micropores. Slit-shaped ultramicropores (with diameters similiar to that of the nitrogen molecule) between the layers are widened to supermicropores near the crystal edges. During an adsorption run the nitrogen molecules penetrate more deeply into the ultramicropores. Nitrogen molecules strongly adsorbed in the ultramicropores are not desorbed at 77 K. Additional amounts of nitrogen are adsorbed by widening of the slit-shaped micropores at the crystal edges when pressure increases. This process proceeds slowly and is reversible.  相似文献   

12.
The pH dependence of dispersion of titanium dioxide (TiO2) particles has been examined in the presence of surfactant molecules in water. Whereas particles were dispersed in water at acid and alkaline regions rather than at neutral region, the dispersion was enhanced at neutral region in an aqueous sodium dodecyl sulfate (SDS) solution and at acid and alkaline regions in an aqueous dodecyldimethylamine oxide (C12DAO) solution. Considering the pH dependence of zeta potential, the adsorption models of surfactant molecules on a particle were estimated on the basis of the modes of hemimicelle and double-layer compression. While the particles that adsorbed Al3+ were remarkably dispersed around pH 6, their dispersion does not largely depend on pH in the addition of SDS, indicating the adsorption of SDS molecules to form double-layer compression in the whole pH region. Dynamic light-scattering measurement and electron microscopic observation suggested that the particles were dispersed in water as small flocs.  相似文献   

13.
Concentration fluctuations in polymer blends and their change after a temperature jump were studied by time-dependent small angle X-ray scattering experiments. Measurements were conducted on homogeneous mixtures of polystyrene and a partially brominated derivative. Structure factors in thermal equilibrium show the form given by the random phase approximation, thus enabling a direct determination of the-parameter and the mean radius of gyration. TheT-dependence of can be understood as the result of superposed enthalpic contributions and a free volume term. In theT-jump experiments, samples were quenched to temperatures near Tg. Relaxation occurs on the time scale of minutes and is nonexponential, becoming slower with time. Initial relaxation rates increase with increasing scattering vectorsq in accordance with theoretical predictions.  相似文献   

14.
By ene-reaction of 4-phenyl-1,2,4-triazoline-3,5-dione with double bonds, polar 4-phenyl-1,2,4-triazolidine-3,5-dione (phenyl urazole) groups are introduced into unpolar matrices. Hydrogen-bond complexes are formed between two amide-like units. The temperature dependence of complex formation in dilute hydrocarbon medium is obtained from remperature-dependent IR spectra in the region of the C=O stretching vibration. Results obtained for a low molecular-weight model system are compared with modified polybutadienes, where the groups are attached statistically along the polymer backbone. The enthalpy and the entropy of complex formation (ΔH f =−28.6 kJ/mol; ΔS f =−52 J/mol K−1) obtained for the low molecular-weight model prove a dimeric chelate like complex involving two hydrogen bonds. The lower equilibrium constants observed for the groups attached to the polymer are related to additional topological constraints emerging from the polymer backbone.  相似文献   

15.
Viscosity measurements were made in the temperature range of 10 °–40 °C. The equation= o exp(B/(T-T o )) was used with the parameterT 0 as structure indicator, which is called the limiting temperature. For instance, hydrocarbons, as liquids with quasifree molecules, haveT 0=O; water as a highly structured liquid hasT 0= 140–150 K.The polymer investigated was ovalbumin in aqueous solution in a concentration comparable to that of blood. Acetylsalicylic acid produces a protein conformation which breaks the water structure in solution at a pH of within the in vivo region.The question of whether only the acidity determines the water structure breaking properties of the protein is investigated by acidifying albumin-water solutions with hydrochloric acid, lysine chloride and ascorbic acid. All these acids exhibit similar effects. A stronger influence is observed for ammonium chloride. Its interaction with ovalbumin produces a strong structure-breaking effect. The most powerful water structure breaker in albumin-water solutions is dextrane. In a concentration of 10 % it changes the polymer conformation so that the water structure is broken to such an extent that the solution behaves as an almost quasifree liquid withT 0=O.  相似文献   

16.
Usingn-hexadecyl acrylate, surface pressure-area (F-A) curves and equilibrium spreading pressuresF e were measured at various temperatures (5.7°–46°C) by the Langmuir balance (F-A) and the Wilhelmy-plate method (F e). At low temperatures (T<13 °C) condensed films and the liquid-condensed/solid condensed transition can be observed. At high temperatures (T>30 °C) liquid-expanded films occur. In the intermediate range the compression curves have two transition points. The transition pressureF 1 between liquid-expanded and condensed film has a marked temperature dependence. The transition enthalpiesH 1 decrease with increasing temperature and become zero at 29.2 °C. The second transition is related to a transition between the condensed films (F 2). The slight temperature dependence of this transition is accompanied by an increasing change of area as well as by increasing transition enthalpiesH 2.TheF e-T curve has two distinct breaks, at the melting pointT m and atT=30 °C. The break atT m is due to the melting process and the break atT=30 °C is caused by a phase transition between a liquid-expanded film and a condensed film.The phase diagram was constructed from the transition pressures. It can be demonstrated that the highest pressures of the thermodynamic stable film occurs atT m. At temperaturesT>T m equilibrium spreading pressure and equilibrium collapse pressure are identical whereas atT m supercompression of the monolayer occurs. The film in this state behaves like a supercooled liquid. Obviously, rupture and collapse of such a film lead to a thermodynamically metastable bulk phase.  相似文献   

17.
The special nature of the outer-most water-rich region of theL 2-phase in the ternary system sodium octanoate-octanoic acid-water is evidenced by its somewhat turbid appearance and by the character of its equilibria with adjacent phases. The phase contains aggregated acid sodium octanoate which is dispersed in a very dilute aqueous solution of sodium octanoate. The acid octanoate has the composition 1 NaC82 HC8x H2O and is composed of closely packed amphiphilic units, all with the polar groups in the same direction. This acid soap obviously forms double-layered aggregates with the lipophilic hydrocarbon chains pointing inwards and the polar groups pointing outwards towards the surrounding bulk-water. The phase is formed when octanoic acid is added to theL 1-phase of the system just above the l.a.c.; in this aqueous solution, the acid reacts with dissolved acid octanoate 1 NaC81 HC8x H2O and that results in the formation of the slightly soluble acid soap 1 NaC8 2 HC8x H2O that separates as a new phase, the turbidL 2 phase. On further addition of octanoic acid, the content of the mentioned acid soap increases until the solution phase is transformed into a liquid crystalline lamellarD-phase with the same acid soap composition. This formation of acid soap 1 NaA2 HA on addition of fatty acid to the dilute soap solution just above the l.a.c., has been known for a long time to occur in various systems containing a long-chain sodium soap. However, at suitably low temperatures, the reaction in these systems does not result in separation of the acid soap in the liquid crystalline, but in the solid crystalline state.  相似文献   

18.
Multinuclear solid‐state NMR and powder X‐ray diffraction data collected for phosphonate materials Zr(O3PC6H4PO3) · 3.6H2O and Sn(O3PC6H4PO3)0.85(O3POH)0.30 · 3.09H2O have resulted in the layered structure, where the phosphonic acids cross‐link the layers. The main structural motif (the 111 connectivity in the PO3 group) has been established by determination of chemical shift anisotropy parameters for phosphorus nuclei in the phosphonate groups. An analysis of the variable‐temperature 31P T1 measurements and the shapes of the phosphorus resonances in the 31P static NMR spectra have resulted in the dipolar mechanism of the phosphorus spin‐lattice relaxation, where the rotating phenylene rings reorient dipolar vectors PH as a driving force of the relaxation process. It has been found that water protons do not affect the 31P T1 times. The activation energy of the phenylene rotation in both compounds has been determined as low as 12.5 kJ/mol. The interpretation of the phosphorus relaxation data has been independently confirmed by the measurements of 1H T1 times for protons of the phenylene rings.  相似文献   

19.
Based on the phenomenon of freezing point depression of a solvent byT, experimental evidence is presented to show that the distance between the junction points can be calculated fromT. Direct measurements of the temperature-time-curve of the cooling network and the Differential Scanning Calorimetry offer the determination ofT. Except the mean distances ¯d c in dependence on cross-linking density, swelling degree, and other network parameters, the distribution of the distance between the junction pointsH(dc) can be determined, which allows conclusions on the course of cross-linking reaction. This paper attempts to give experimental evidence of influences of the breadth ofH(dc) on application-relevant properties.  相似文献   

20.
The dynamic-structural changes and polymer - solvent interactions during the thermotropic phase transition in poly(vinyl methyl ether) (PVME)/D2O solutions in a broad range of polymer concentrations (c = 0.1-60 wt.-%) were studied combining the measurements of 1H NMR spectra, spin-spin (T2) and spin-lattice (T1) relaxation times. Phase separation in solutions results in a marked line broadening of a major part of polymer segments, evidently due to the formation of compact globular-like structures. The minority (∼15%) mobile component, which does not participate in the phase separation, consists of low-molecular-weight fractions of PVME, as shown by GPC. Measurements of spin-spin relaxation times T2 of PVME methylene protons have shown that globular structures are more compact in dilute solutions in comparison with semidilute solutions where globules probably contain a certain amount of water. A certain portion of water molecules bound at elevated temperatures to (in) PVME globular structures in semidilute and concentrated solutions was revealed from measurements of spin-spin and spin-lattice relaxation times of residual HDO molecules.  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