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.
Oleoyl‐chitosans (O‐chitosans), with three degrees of substitution (DS), were synthesized by reacting chitosan with oleoyl chloride. The chemical structures of these polymers were characterized by 1H NMR and FTIR. The results suggested the formation of an amide linkage between amino groups of chitosan and carboxyl groups of oleic acid. These O‐chitosans exhibited poor solubility in aqueous acidic solution. The solubility of O‐chitosans decreased as the DS values increased. The transmittance of O‐chitosans (2 g/L) with DS 5%, 11%, 27% in 1% (v/v) HCl solution were 69.5%, 62.7%, 48.6%, respectively. These O‐chitosans were not soluble at neutral or alkali pH. Formation of self‐aggregation was observed using pyrene as a fluorescent probe in the O‐chitosans aqueous solution. The increase of DS of O‐chitosans resulted in significant decrease of critical aggregation concentration (CAC). The CAC of the O‐chitosans with DS 5%, 11%, 27% were 79.43, 31.6, 10 mg/L, respectively. The surface tension of solution could be reduced slightly by all of the O‐chitosans. The surface tension of O‐chitosans solution decreased with the increase of DS values. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 102:1968–1973, 2006  相似文献   

2.
Monolayer formation by poly[(maleic anhydride)‐alt‐(stearyl methacrylate)] (MA‐alt‐StM) on aqueous subphases, with and without chitosan, was studied by the Langmuir technique. Chitosan (CS) modified considerably the shape of the MA‐alt‐StM isotherms on water. To explain this behavior, the surface activity properties of chitosan at the air–solution interface were studied. The variations of the interfacial tension, γint, with chitosan concentration and temperature, were also determined. The results were discussed in terms of the modification of the air–water interface owing to the presence of chitosan in the subphase and the surface activity. It was found that the standard free energy of adsorption, , values were dependent on the degree of acetylation (DA) over the DA range being studied. Copyright © 2004 Society of Chemical Industry  相似文献   

3.
Kinetics of the solid-solid II-I phase transition of isotactic polybutene-1 was investigated. The fraction WI of phase I as a function of time ttr during the phase transition was measured by X-ray diffraction at various temperatures Ttr. The Avrami indices n of the WI-ttr plots are approximately unity for Ttr > 288 K. A bell-shaped temperature dependence of the transition rate V with the maximum transition rate at 285 K was obtained. The V-Ttr curve and the Avrami index n = 1 suggest that the rate-determining process is primary nucleation. The dependence of V on Ttr for Ttr < 283 K is described by the William-Landel-Ferry (WLF) equation, which shows that the glass transition affects the transition rate. The Avrami index decreases to n < 1 for Ttr < 283 K, indicating a broadened distribution of the transition rate caused by the spatial heterogeneity of the amorphous state at low temperatures near the glass transition. Those evidences at low temperature clearly suggest that the solid-solid phase transition is influenced by the mobility of chain folding, tie chains and cilia in the amorphous between the stacks of lamellar crystals.  相似文献   

4.
The analysis of phase equilibrium between methanol and glycerides during methyl esters of fatty acids (FAME or biodiesel) synthesis at high pressure and temperature is very important for describing the kinetic and process design. It was studied at pressure between 1.1 and 28.0 MPa and temperature from 150 to 270 °C. The transition of phases and composition of identified phases was calculated using RK-Aspen EOS and obtained values were also compared to experimentally determined data at subcritical condition (1.1-4.5 MPa and 150-210 °C).Results of experimental investigation, as well as performed simulation of some specified composition of reaction mixture, showed that system of triglycerides and methanol, at the beginning of reaction (at all analysed conditions except for supercritical state of mixture) is in equilibrium between two liquid phases. During the methanolysis of triglycerides, the phase's distribution was changed accordingly and it highly depends on actual composition of reaction mixture, temperature and pressure. Calculated and measured values indicated that distribution of methanol between the oil phase, the methyl esters, and the glycerol rich phase exists and depends of working condition. As a consequence of fact, that the methanolysis of triglycerides (oil) is mainly realized in the oil-rich phase, at the end of reaction, after all triglycerides are converted into FAME and glycerol, the oil phase disappears. Furthermore, according to the results of phase composition calculation, it was shown that from the beginning to the end of reaction one phase only exists, for methanolysis performed at 270 °C and 20.0 MPa.  相似文献   

5.
6.
《Ceramics International》2020,46(8):12275-12281
Alumina and zirconia ceramic particles exhibit high hardness and excellent wear resistance at high temperature, and hence are used as ceramic reinforcement phases in some plasma sprayed coatings. In this study, the interface evolution of a zirconia/alumina eutectic ceramic and the phase transition of zirconia in a plasma-sprayed coating were investigated. Scanning electron microscopy and transmission electron microscopy combined with focused-ion beam and energy dispersive X-ray were used to analyze the microstructure and composition of the ceramic interface. The results showed that the eutectic ceramic particles consisted of alumina (outer) and columnar zirconia (inner) before and after the plasma spraying process. The inner zirconia part showed the martensitic transformation of t-type zirconia to stripe-like m-type zirconia. After the plasma spraying, the interface between alumina and zirconia changed significantly, which formed a new oxide layer. The phase transition mechanism in the ceramic particle and oxide layer formation mechanism at the alumina/zirconia interface were investigated.  相似文献   

7.
The microstructure and macroscopic properties of ferroelectric materials at high pressure are of great interest in both the engineering and scientific arenas. The effect ofthe pressure value, loading time (the time taken for the pressure to increase from atmospheric pressure to the highest pressure) and loading direction on the evolution of domains and the ferroelectric phase transition for a BaTiO3 single crystal was investigated using a phase field approach. It was found that under symmetrical compression loading the pressure loading time affected the phase transition path and rate but did not affect the phase transition pressure or the ultimate stable phase. For example, at room temperature, even when the loading time increased from 1 ns to 10 μs, the phase transition pressure remained stable at 2.1 GPa, but the phase transition time was prolonged. At −70 °C the orthorhombic–cubic phase transition was induced when the loading pressure was 5 GPa and the loading time was 1 ns, whereas the orthorhombic–tetragonal–cubic phase transition occurred when the loading time increased to 10 μs. In addition, it was found that the application of symmetrical pressure tended to reduce the degree of ferroelectricity, while one-dimensional compression favored the ferroelectric phase.  相似文献   

8.
The behavior of amphiphilic diblock copolymers containing 80–89% of poly(N‐hexylisocyanate) (PHIC) with different hydrophobic segments spread at the air–water interface has been studied. Surface pressure‐area isotherms (π‐A) at the air–water interface were determined. It was found that these diblock copolymers form stable monolayers and the isotherms present a pseudoplateau region at low surface pressure, irrespective of the nature of the partner block: poly(styrene) (PS) or poly(isoprene). Surface pressure variation at the semidilute region of the monolayer was expressed in terms of the scaling laws as power function of the surface concentration. The critical exponents of the excluded volume ν obtained for copolymers with PHIC and PS blocks are 0.58 for the copolymer with 85% of PHIC and 15% of PS, and 0.63 for the copolymer with 89% of PHIC and 11% of PS. The hydrophobicity degree of the diblock copolymers was estimated from the determination of the surface energy values by wettability measurements. The morphology of the monolayers was determined by means of Brewster angle microscopy. Molecular dynamic simulation was performed to explain the experimental behavior of diblock copolymers at the air–water interface. © 2011 Wiley Periodicals, Inc. J Appl Polym Sci, 2011  相似文献   

9.
Silver‐nanoparticle‐doped poly(9‐vinylcarbazole) (PVK) nanocomposites were prepared via the reduction of Ag+ ions and the self‐assembly of PVK on AgNO3 aqueous solution surfaces. The formed composite nanostructures depended strongly on the experimental temperature. Thick round disks of PVK surrounded by discrete Ag nanoparticles and/or with irregular holes formed at room temperature; nanotubes and micronetworks doped with Ag nanoparticles formed at about 30–40°C, and networks formed at higher temperature. Further investigation revealed that the nanotubes were transformed from thin round disks. The length of the PVK/Ag composite nanotubes were longer than 10 μm, and the average size of the embedded Ag nanoparticles was found to be about 3.5 nm. The composite networks were composed of round pores with diameters of several hundred nanometers and fine silver nanoparticles embedded in the thin polymer films that covered the pores. The formation of the nanotubes was a very interesting self‐assembly phenomenon of the polymer at the air–water interface that has not been reported before. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2010  相似文献   

10.
We examine the dynamics and morphology of graphitic films at an air‐water interface in a Langmuir trough by varying interfacial surface coverage, by observing in situ interfacial structure, and by characterizing interfacial structure of depositions on mica substrates. In situ interfacial structure is visualized with Brewster angle microscopy and depositions of the interface are characterized with atomic force microscopy and field‐emission scanning electron microscopy. Compression/expansion curves exhibit a monotonically decreasing surface pressure between consecutive compressions, but demonstrate a “rebound” of hysteretic behavior when the interface is allowed to relax between consecutive compressions. This dynamic results from a competition between consolidation of the interface via agglomeration of particles or the stacking of graphene sheets, and a thermally‐driven relaxation where nanometer‐thick particles are able to overcome capillary interactions. These results are especially relevant to applications where functional films with controlled conductivity and transparency may be produced via liquid‐phase deposition methods. © 2018 American Institute of Chemical Engineers AIChE J, 64: 3177–3187, 2018  相似文献   

11.
12.
A simulation model for adsorption kinetics of hen egg white lysozyme (HEW) adsorption to hydrophilic silica is proposed. The adsorption kinetic data were monitored by usingin-situ ellipsometry. The model is based on an irreversible adsorption mechanism allowing two different adsorbed states. The adsorbed states were differentiated based on binding strengths resistant to the concentration gradient exerted by rinse. Molecules desorbing and remaining upon rinse were identified as loosely bound (state 1) and tightly bound (state 2) states, respectively. The adsorption rate constants were assumed to be a time-dependent nonlinear function in order to account for the change in surface properties originating from the protein layer formed on the surface. The parameters of adsorption rate constants were evaluated by using adsorption kinetic data at different protein concentrations, and the relationships between the adsorption parameters and protein concentration were established which eventually demonstrated a linear relationship. The established relations between the adsorption parameters and concentration elucidated the effect of protein concentration on adsorption to hydrophilic silica.  相似文献   

13.
1,2-Dipalmitoyllecithine forms adsorbed films at electrodes in two distinct potential regions. At temperatures below the phase transition temperature the film at negative potentials is highly ordered and blocks the surface for electrode reactions. Potential steps between both regions enable us to monitor the random nucleation during the film transition by its influence upon the charge transfer process. The time distribution of current pulses obeys the Poisson distribution. The current power spectrum is given.  相似文献   

14.
《Ceramics International》2022,48(18):25996-26002
NaNbO3 (NN) is a well-known perovskite-type dielectric material. However, its phase transition is a complex process and the phase transition mechanism is insufficiently investigated yet. Therefore, in the work, the NN superfine particles were synthesized by a hydrothermal method with the assistance of K+ and the microstructures of NN were carefully investigated to understand the strain-driven phase transition and morphology evolution of NN. The results suggest that K+ leads to strong planar bending of NbO6 by influencing the Na + vibration, which accounts for the O-R phase transition as the K+ increases. Once the R phase is triggered, the O-R phase transition proceeds spontaneously based on phonon calculations of orthorhombic (O) NN and R NN. The antiferroelectric nature of rhombohedral (R) NN with the space group R–3H is confirmed through charge density distribution. Additionally, the morphology evolution is deduced on the TEM analysis basis.  相似文献   

15.
Vapor-liquid-equilibria (VLE) and vapor-liquid-liquid equilibria (VLLE) data for the carbon dioxide + 1-nonanol system were measured at 303.15, 308.15, 313.15, 333.15, and 353.15 K. Phase behavior measurements were made in a high-pressure visual cell with variable volume, based on the static-analytic method. The pressure range under investigation was between 1.15 and 103.3 bar. The Soave-Redlich-Kwong (SRK) equation of state (EOS) coupled with both classical van der Waals and a Gibbs excess energy (GE) mixing rules was used in semi-predictive approaches, in order to represent the complex phase behavior (critical curve, liquid-liquid-vapor (LLV) line, isothermal VLE, LLE, and VLLE) of the system. The topology of phase behavior is correctly predicted.  相似文献   

16.
Two‐phase flow pattern and friction characteristics for an air–water system in a 3.17 mm smooth tube are reported in this study. The range of mass flux is between 50 and 700 kg/m2s. The experimental data show that the two‐phase friction multipliers are strongly related to the flow pattern. For a stratified‐wavy flow pattern, a mass‐flux dependence of the two‐phase multipliers is seen. For a non‐stratified flow pattern, the two‐phase frictional multipliers are comparatively independent of mass flux. Correlations of the frictional multipliers are developed for stratified and non‐stratified flow. To use the appropriate correlation in different regime, a simple criterion is proposed.  相似文献   

17.
A novel separation process based on the hydrophobic adsorption at the n‐hexadecane–water interface was developed for the recovery of Acinetobacter radioresistens lipase from a pre‐treated fermentation broth. In a mixture containing water, lipase and n‐hexadecane, a water‐in‐oil emulsion was formed when the n‐hexadecane‐to‐water ratio (o/w ratio) was larger than 3, and a large amount of lipase was found to be adsorbed at the interface. Compared with the oil‐in‐water emulsion (occurring when o/w ratio < 3), the water‐in‐oil emulsion generated smaller droplets and larger interfacial area, and was more stable. The harvested emulsion phase could be centrifuged to give an aqueous, concentrated lipase solution. Adsorption of lipase at the interface could be described by the Langmuir isotherm. For lipase concentrations ranging from 8.4 to 87.2 U cm?3, a single‐stage adsorption resulted in a six‐ to four‐fold concentration and 16–45% activity recovery, where lipase concentration was the dominant factor. A method using data from a single‐stage adsorption to predict multiple‐stage operation was described, and the agreement between the experimental and the predicted results was good. To improve the enzyme recovery, a multiple‐run adsorption process was proposed. The use of salts enhanced the hydrophobic interaction between lipase and n‐hexadecane. Advantages of the proposed process include simple operation, low operational cost, environmentally friendly, no requirement for pre‐concentration of the enzyme solution, and negligible enzyme denaturation. Copyright © 2003 Society of Chemical Industry  相似文献   

18.
It is known that a reflectance spectrum for an optimal colour takes on the value 0 or 1 at every wavelength, with at most two transitions between those values. This article shows that any non‐optimal colour can be produced by a reflectance spectrum that takes on the value 0 or 1 at every wavelength, with at most four transitions. While the two‐transition optimal spectrum is unique, the four‐transition non‐optimal spectrum is not unique.  相似文献   

19.
The monolayer behavior of phospholipids at the air/water interface and their transfer characteristics on the: solid substrates have been investigated with a constant-perimeter type Langmuir trough. From the surface pressure-area (Π-A) isotherm, evaluated were the cross-sectional area of an oriented molecule, phase transition, and the miscibility of mixed monolayers. The monolayer state depends on the chain length and its mobility, and by adding proper salts, the monolayers were stabilized. The miscibility of mixed monolayers was also discussed with collapse pressure and excess area. The monolayers of all lipids were transferred into only one layer on hydrophobic substrates, and up to two layers on hydrophobic substrates. The multilayer formation of dipalmitoylphosphatidic acid (I)PPA) was significantly affected by the subphase pH and the addition of multivalent salts. As a transfer promoter. DPPA or octadecylamine (ODA) was used as a component in a mixture with lipid materials, and their mixed monolayers resulted in good transfer characteristics. The transferred mass and film thickness of DPPA, determined by using a quartz crystal microbalance (QCM) and by an ellipsometry respectively, were proportional to the number of transfer. The lattice spacing of DPPA was 20.3 Å per layer by ellipsometry, and 22.78 Å per layer by X-ray diffraction. These indicated that the transferred multilayers had well-defined layered structures of Y-type.  相似文献   

20.
Silica nanoparticles with well‐defined, highly grafted dense poly(methyl methacrylate) (MMA) were prepared by surface‐initiated activators regenerated by electron transfer for atom transfer radical polymerization (ARGET ATRP) of methyl methacrylate with an initiator‐fixed silica particle in the presence of air. Two different polymerizations of MMA were carried out under the same conditions using tris[2‐(dimethylamino)ethyl]amine (Me6TREN) and N,N,N′,N′,N″‐pentamethyldiethylene‐triamine (PMDETA) as the ligand, respectively. In the CuCl2/PMDETA system, polymerization appeared to be more controlled with a lower polydisperisty compared with the CuCl2/Me6TREN system. The monolayer of these particles was formed at the air–water interface using Langmuir‐Blodgett (LB) technique. Multilayers of the particles were fabricated by repetition of LB depositing. A surface pressure–area (π–A) measurement and SEM observation were used to characterize the particle arrays. POLYM. ENG. SCI., 2010. © 2009 Society of Plastics Engineers  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