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 46 毫秒
1.
Polymeric lipid nanoparticles were prepared in 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid buffer (pH 8.0, 10 mM) by taking advantage of salt bridges formed between poly(N-isopropylacrylamide-co-methacrylic acid) (P(NIPAM-co-MAA)) and N-[3-(dimethylamino)propyl]-octadecanamide (DMAPODA). The homogeneous nanoparticles of 200–250 nm were obtained when the copolymer was included in the preparation so that the relative mass of copolymer to lipid was more than 5. However, when the relative amount of copolymer was less than 5, large agglomerates more than 10 μm were observed together with nanoparticles. The protonated amino groups of DMAPODA will attach to the ionized carboxyl groups of P(NIPAM-co-MAA), and they would act as polymeric amphiphiles. It is believed that the hydrophilic copolymer can stabilize the hydrophobic core of the lipid. The critical association concentrations were determined to be 32, 112, and 158 mg/l, when the lipid/copolymer ratios were 1:5, 1:23, and 1:50, respectively.  相似文献   

2.
Vesicles composed of N-[3-(dimethylamino) propyl]-octadecanamide (DMAPODA) and 2-(hexadecyloxy) cinnamic acid (HOCA) in an equimolar ratio were prepared by taking advantage of salt bridge formed between the amino group and the carboxylic group. The structure of vesicle was observed on a Transmission electron microscopy (TEM), and the size was determined on a dynamic light scattering equipment. The phase transition of the vesicular membrane was found to be around 35 °C on a differential scanning calorimeter. HOCA of the vesicular membrane was readily dimerized under the irradiation of a UV light (λ?=?254 nm, 6 W). The release degree of rhodamine B from vesicle suspended in distilled water (pH 6.8) was about 70 % in 1 h at 25 °C, and the UV irradiation during the release experiment had little effect on the release degree. However, it had a significant effect when the temperature of release medium was 40 °C. Upon the photodimerization, HOCA would readily change its orientation in the vesicular membrane vesicle at 40 °C, possibly because the vesicular membrane is in a liquid crystalline state at the temperature, which is higher than the phase transition temperature (around 35 °C). In addition, the vesicle released rhodamine B in a pH-dependent manner. The release degrees were the highest at pH 3.0 and the lowest at pH 9.0 among the pH values tested. The salt bridge formed between DMAPODA and HOCA is labile at a strong acidic condition so it would be responsible for the extensive release at pH 3.0.  相似文献   

3.
Reduction-responsive vesicle was prepared by salt-bridging N-[3-(dimethylamino)propyl]-octadecanamide (DMAPODA, a cationic amphiphile) using 3,3′-dithiodipropionic acid (DTPA, a disulfide diacid compound). According to the transmission electron micrograph and the fluorescence quenching degree (53.2%), it could be said that vesicles were formed when the DMAPODA to DTPA molar ratio was 2:2. The DMAPODA/DTPA associate was considered to be a building block for vesicle formation because DTPA could electrostatically associate with DMAPODA and help the cationic amphiphile assemble into the vesicle. On a differential scanning calorimetric thermogram, the DMAPODA/DTPA vesicle showed two endothermic peaks at 50.6°C and 63.2°C. The peak found at the lower temperature was possibly due to the solid gel-to-liquid crystal phase transition of the vesicular membrane and the peak found at the higher temperature was considered to be due to the melting of DMAPODA, indicating that unassociated DMAPODA coexisted with DMAPODA/DTPA vesicles. The release of calcein enveloped in the vesicle was promoted by DL-dithiothreitol, possibly because DTPA can be broken by the reducing agent to form mercaptopropionic acids and the vesicle could be disintegrated and/or the vesicular membrane would become defective.  相似文献   

4.
The immobilization and positioning of ultra small reaction vessels on solid supports open new pathways in applications such as lab-on-a-chip, sensors, microanalyses and microreactors. In our work block copolymer vesicles made from polystyrene-block-polyacrylic acid (PS-b-PAA) were immobilized from aqueous medium onto 3-amino propyl trimethoxysilane functionalized silicon surfaces exploiting electrostatic interactions. The immobilization of the vesicles was investigated by Fourier transform infrared (FTIR) spectroscopy, as well as fluorescence optical and atomic force microscopy (AFM). In addition, the influence of pH and ionic strength on the surface coverage of vesicles bound to the surface was elucidated. Finally micro-molding in capillaries (MIMIC) was utilized to create line patterns of the vesicles containing the enzyme trypsin and the fluorogenic substrate rhodamine 110 bisamide. The selective positioning of vesicle nanoreactors in conjunction with electrostatic immobilization serves as a proof of principle for potential applications in real-time observation of confined chemical reaction inside vesicles as nanocontainers and for the fabrication of integrated microarray systems.  相似文献   

5.
The biosorption of nickel(II) ions from aqueous solution by Acacia leucocephala bark was studied in a batch adsorption system as a function of pH, initial metal ion concentration, adsorbent dosage, contact time and temperature. The maximum Ni(II) adsorption was obtained at pH 5. Further, the biosorbents were characterized by Fourier Transformer Infrared Spectroscopy (FTIR). The experimental data were analysed using three sorption kinetic models viz., the pseudo-first- and second-order equations and the intraparticle diffusion model. Results show that the pseudo-second-order equation provides the best correlation for the biosorption process. The equilibrium nature of Ni(II) adsorption at different temperatures of 30, 40 and 50 °C have been described by the Langmuir and Freundlich isotherm models. The equilibrium data fit well Langmuir isotherm. The monolayer adsorption capacity of A. leucocephala bark as obtained from Langmuir isotherm at 30 °C was found to be 294.1 mg/g. The Chi-square (χ2) and Sum of the square errors (SSE) tests were also carried out to find the best fit adsorption isotherm and kinetic model. Isotherms have been used to determine thermodynamic parameters of the process, viz., free energy change (ΔG°), enthalpy change (ΔH°), and entropy change (ΔS°) were calculated indicating that this system was a spontaneous and endothermic process. Present investigation emphasized that A. leucocephala bark may be utilized as a low cost adsorbent for nickel removal.  相似文献   

6.
Molecular interactions between a terminally alkylated pH-sensitive N-isopropylacrylamide copolymer DODA-poly(NIPAM-co-MAA) and a monolayer of distearoylphosphatidylcholine (DSPC) at the air/water interface are investigated using the Langmuir balance technique. The compression isotherms ofthe copolymer monolayer at the air-water interface confirm that the copolymer undergoes a structural transition with a change in pH ranging from an extended coil state at neutral pH to a collapsed globular state at a pH corresponding to the pH of the polymer phase transition. Adsorption kinetics of DODA-poly(NIPAM-co-MAA) in the DSPC monolayer is analyzed using a first-order kinetics model allowing an effective interaction area Ax between DSPC and DODA-poly(NIPAM-co-MAA) molecules to be evaluated. The results clearly indicate that the interaction area increases with a decrease in pH. The results also suggest that the penetration of the DODA-poly(NIPAM-co-MAA) within the phospholipid monolayer is enhanced by a decrease in pH which causes a change in the copolymer structure and an increase in specific attractive interactions between the copolymer and the phospholipid. Therefore, the copolymer can trigger the destabilization or rupture of the phospholipidic layer through a simple variation in its structure associated with a variation in molecular interactions when coupled or inserted within the membrane. This study greatly supports the prospects of the copolymer-functionalized liposomes as stable and tunable carrier systems for in vivo applications in drug delivery.  相似文献   

7.
Marine endosymbiotic fungi Aspergillus ustus (MSF3) which produce high yield of biosurfactant was isolated from the marine sponge Fasciospongia cavernosa collected from the peninsular coast of India. Maximum production of biosurfactant was obtained in Sabouraud dextrose broth. The optimized bioprocess conditions for the maximum production was pH 7.0, temperature 20 °C, salt concentration 3%, glucose and yeast extract as carbon source and nitrogen sources respectively. The response surface methodology based analysis of carbon and nitrogen ratio revealed that the carbon source can increase the biosurfactant yield. The biosurfactant produced by MSF3 was partially characterized as glycolipoprotein based on the estimation of macromolecules and TLC analysis. The partially purified biosurfactant showed broad spectrum of antimicrobial activity. The strain MSF3 can be used for the microbially enhanced oil recovery process.  相似文献   

8.
A simple copolymer, poly(NIPAM-co-RD), consisting of N-isopropylacrylamide (NIPAM) and rhodamine (RD) units, behaves as a fluorescent temperature sensor exhibiting selective fluorescence enhancement at a specific temperature range (25–40 °C) in water. This is driven by a heat-induced phase transition of the polymer from coil to globule. At low temperature, the polymer exists as a polar coil state and shows very weak fluorescence. At >25 °C, the polymer weakly aggregates and forms a less polar domain within the polymer, leading to fluorescence enhancement. However, at >33 °C, strong polymer aggregation leads to a formation of huge polymer particles, which suppresses the incident light absorption by the RD units and shows very weak fluorescence. In the present work, effects of polymer concentration and type of acrylamide unit in the polymer have been investigated. The increase in the polymer concentration in water leads to a formation of less polar domain even at low temperature and, hence, widens the detectable temperature range to lower temperature. Addition of N-n-propylacrylamide (NNPAM) or N-isopropylmethacrylamide (NIPMAM) component to the polymer, which has lower or higher phase transition temperature than that of NIPAM, enables the aggregation temperature of the polymer to shift. This then shifts the detectable temperature region to lower or higher temperature.  相似文献   

9.
In order to design liposomes which release their contents in a glucose-sensitive manner, the surfaces of egg phosphatidylcholine (egg PC) liposomes or dioleoylphosphatidylethanolamine (DOPE) liposomes were modified with the copolymer of N-isopropylacrylamide/methacrylic acid/octadecylacrylate and hydrophobically modified glucose oxidase (EC 1.1.3.4.). Whichever the liposomes were prepared with egg PC or DOPE, an extensive release of calcein was observed at acidic conditions. And DOPE liposomes were more pH sensitive than egg PC liposomes in terms of the release. In glucose-dependent calcein release experiment, there was no release for 180 min when the suspension of liposome was free of glucose. When the glucose concentration was 50 mg/dl, no appreciable amount of calcein was released for the first 50 min, but a significant release was observed for the last 130 min. At glucose concentration of 200 mg/dl, calcein release became more extensive and the releases for 180 min from egg PC and DOPE liposome were 84% and 98%, respectively.  相似文献   

10.
Temperature- and pH-sensitive copolymer gels were synthesized by the simultaneously occurring radiation-induced polymerization and self-bridging of acryloyl- -proline methyl ester (A-ProOMe) with acrylic acid (AAc) in aqueous solutions. The gel swelling behavior and the metal permeation characteristic of its gel membrane were investigated with regard to very slight changes of temperature and pH. The pH threshold of the swelling of a copoly(A-ProOMe/AAc, 70/30 mol%) gel in the range of 5–30°C lay in the region between pH 4.0 and 5.0. The permeability results of metal ions showed that at 40°C the gel membrane blocks the permeation of lithium (Li) and cesium (Cs) ions at pH values lower than 4.75 and 4.60, respectively. The permselectivity (PLi/Cs value) of the two metal ions at 30°C was also studied and, as a result, its value was obtained to be 1.33 at pH 4.65 and 30°C. This permeation study indicates that the selective metal separation of copoly(A-ProOMe/AAc) gel membranes can be controlled by changing temperature and pH values.  相似文献   

11.
Polyprenyl phosphates, as well as polyprenyl alcohols bearing different isopentenyl C(5) units, have been synthesized. The pH range of spontaneous vesicle formation of polyprenyl phosphates with or without polyprenyl alcohols was defined by fluorescence microscopy. A variety of the acyclic or monocyclic polyprenyl phosphates studied formed stable vesicles in water over a wide range of pHs, and the addition of polyprenyl alcohols allowed the vesicle formation of polyprenyl phosphates at higher pHs. Osmotic swelling of a suspension of unilamellar vesicles using the stopped-flow/light-scattering method enabled us to evaluate the water permeability of polyprenyl phosphate vesicles with or without 10 mol% of free polyprenyl alcohol. The addition of many polyprenyl alcohols to polyprenyl phosphate vesicles decreased the water permeability, and some reduced it even more efficiently than cholesterol.  相似文献   

12.
Polytetrafluoroethylene (PTFE) was grafted (g) with acrylic acid (AAc) by γ-ray pre-irradiation method to get PTFE-g-AAc films, then N-isopropylacrylamide (NIPAAm) was grafted onto PTFE-g-AAc films with γ-ray to get (PTFE-g-AAc)-g-NIPAAm. PTFE films were irradiated in air at a dose rate of 3.0 kGy h–1 and different radiation dose. The irradiated films were placed in glass ampoules, which contained aqueous solutions with different monomer concentration (AAc), and then they were heated at different temperatures and reaction time. NIPAAm onto PTFE-g-AAc was carried out with the same procedure with monomer concentration of 1 mol L−1. The thermosensitivity of the samples was defined and calculated as the ratio of the grafted samples swelling at 28 and 35 °C, and pH sensitivity defined as the ratio of the grafted samples swelling at pH 2 and 8.  相似文献   

13.
A range of carbonated apatites was prepared by aqueous precipitation at 37, 60, and 85°C and at controlled pH values varying from 6.00–9.50 in 0.25 increments. The products were analyzed for Ca, P, Sr, Mg, Na, F, and carbonate. Their initial dissolution rates were measured in a pH 4.5, 0.01 mol · liter−1 acetate buffer. Information about crystal morphologies and crystal defects was obtained by x-ray diffraction and high-resolution transmission electron microscopy (TEM). The molar ratios of the products, together with their Sr and Mg contents, increased with increasing pH. Initial dissolution rates of the products, when adjusted for carbonate content, were in the order 37 > 60 > 85°C whereas apparent particle sizes determined by TEM and x-ray diffraction were ordered 37 < 60 < 85°C. Carbonated apatites precipitated at pHs of 7.0 or less were observed to have planar defects parallel to (100) that were identified as unit-cell-thick intergrowths of octacalcium phosphate. Carbonated apatites precipitated at higher pHs and noncarbonated apatites did not have these defects. A crystal growth mechanism is proposed to account for the presence of the (100) defects.  相似文献   

14.
A series of novel hydrogels based on poly(N-isopropylacrylamide) (PNIPAAm) with pendant poly(N-(2-(dimethylamino) ethyl)-methacrylamide) (PDMAEMA) grafts were designed and synthesized. The influence of the pendant PDMAEMA grafts on the properties of the resulted hydrogels was examined in terms of morphology observed by scanning electron microscopy (SEM), thermal property characterized by differential scanning calorimetry (DSC) and shrinking/swelling kinetics upon external temperature changes. In comparison with the conventional PNIPAAm hydrogels, resulting hydrogels presented favorable pH sensitivity as well as improved thermosensitive properties, including enlarged water containing capability at room temperature and faster shrinking/swelling rate upon heating. In addition, fish DNA, used as a model drug, was loaded into the hydrogels, and the controlled release behavior of the drug-loaded hydrogels at different temperatures (22 and 37 °C) was further studied.  相似文献   

15.
The effect of the buffer composition and the preparation protocol on the dynamic surface tension (DST) and vesicle sizes of aqueous dipalmitoylphosphatidylcholine (DPPC) dispersions was studied. Four isotonic buffers were used in preparing DPPC dispersions at physiological conditions for possible biological applications: (1) a standard PBS solution; (2) the above PBS with 1 mM CaCl2; (3) PBS with one tenth the previous standard phosphate salt concentrations and 2.5 mM CaCl2; and (4) 150 mM NaCl with 2.5 mM CaCl2 and 10 mM HEPES (4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid). Two protocols, with a new method and an old method (Bangham method), were used in preparing the DPPC dispersions. The DPPC dispersions prepared with the new method contained mostly vesicles and were quite stable at 25 or 37 °C. Dynamic light scattering (DLS) and spectroturbidimetry (ST) results showed that the DPPC vesicle sizes in buffer (4) were much smaller than those in the other buffers. When the DPPC dispersions were prepared with the new method, the diameter of the DPPC particles was smaller than those with the old method. The DPPC vesicles with the new method were more stable than those with the other method. The DPPC dispersions of 1000 ppm at 37 °C with the new method produced, at pulsating area conditions at 20 cycles per minute, low tension minima (γmin), lower than 10 mN/m, in buffers (1), (2), and (4). With buffer (4) the DSTs were lower and were achieved faster than with the other buffers. A minimum concentration of 1000 or 250 ppm DPPC was needed to produce DSTs lower than 10 mN/m within 10 min or less, with buffer (2) or (4), respectively. IRRAS results suggest that DPPC in buffer (2) or (4) forms a close-packed monolayer at the interface. These results have implications for designing efficient protocols of lipid dispersion preparation and lung surfactant replacement formulations in treating respiratory disease.  相似文献   

16.
The differential microemulsion polymerization technique was used to synthesize the nanoparticles of glycidyl-functionalized poly(methyl methacrylate) or PMMA via a two-step process, by which the amount of sodium dodecyl sulfate (SDS) surfactant required was 1/217 of the monomer amount by weight and the surfactant/water ratio could be as low as 1/600. These surfactant levels are extremely low in comparison with those used in a conventional microemulsion polymerization system. The glycidyl-functionalized PMMA nanoparticles are composed of nanosized cores of high molecular weight PMMA and nano-thin shells of the random copolymer poly[(methyl methacrylate)-ran-(glycidyl methacrylate)]. The particle sizes were about 50 nm. The ratios of the glycidyl methacrylate in the glycidyl-functionalized PMMA were achieved at about 5–26 wt.%, depending on the reaction conditions. The molecular weight of glycidyl-functionalized PMMA was in the range of about 1 × 106 to 3 × 106 g mol−1. The solid content of glycidyl-functionalized PMMA increased when the amount of added glycidyl methacrylate was increased. The glycidyl-functionalized polymer on the surface of nano-seed PMMA nanoparticles was a random copolymer which was confirmed by 1H-NMR spectroscopy. The amounts of functionalization were investigated by the titration of the glycidyl functional group. The structure of the glycidyl-functionalized PMMA nanoparticles was investigated by means of TEM. The glycidyl-functionalized PMMA has two regions of Tg which are at around 90 °C and 125 °C, respectively, of which the first one was attributed to the poly[(methyl methacrylate)-ran-(glycidyl methacrylate)] and the second one was due to the PMMA. A core/shell structure of the glycidyl-functionalized PMMA latex nanoparticles was observed.  相似文献   

17.
We have studied the pH dependence of the light-induced proton release and uptake by bacteriorhodopsin. The quantum efficiency of proton release in cell envelopes and proton uptake in phospholipid vesicles is high in the low pH range and begins to decline between pH 6 and 7 in cell envelopes and between pH 7–8 in phospholipid vesicles. In the cell envelope vesicles the proton release increases again above pH 8–8.5; in phospholipid vesicles a proton release is observed before proton uptake at pHs greater than 9. We suggest that the light-induced proton release observed at high pHs are due to protons released and rebound on the carboxyl terminal side of bacteriorhodopsin.  相似文献   

18.
We report the variation of the molecular assembly and crystallinity of polyethylene-block-poly(ethylene oxide) (EEO, 1400 g mol−1) in a non initiated liquid resin of unsaturated polyester (UP). We particularly focus on the driving force that governs the variation of the molecular assembly of the block copolymer in the UP resin upon heating. For this purpose, we performed a set of experiments combining time resolved in situ SAXS and WAXS measurements upon heating and cooling. Upon heating, SAXS shows that the inter-distance of the EEO domains decreases dramatically for temperature above 75 °C (between 90 and 30 nm for the initial system at room temperature vs 9 nm at 110 °C) suggesting that the initially homogeneously dispersed EEO domains undergo aggregation. This is consistent with the macroscopic phase separation observed in such temperature range. In situ WAXS shows that the onset of the aggregation coincides with the melting of the PE domains suggesting that the flexibility of the platelets plays an important role in their aggregation. Upon cooling, peculiar behaviour is observed for the systems with higher EEO content, with an irreversible structure formation leading to macroscopic EEO rich domains dispersed in continuous UP rich phase.  相似文献   

19.
The kinetics of pH-induced formation and dissociation of vesicles self-assembled from a biocompatible zwitterionic diblock copolymer, poly(2-(methacryloyloxy)ethyl phosphorylcholine)-b-poly(2-(diisopropylamino)ethyl methacrylate) (PMPC- b-PDPA), was investigated in detail via a combination of stopped-flow light scattering and laser light scattering (LLS). Upon jumping from pH 2 to 10, stopped-flow light scattering reveals three distinct relaxation processes for the early stages of vesicle self-assembly (0-40 s). Kinetic sequences associated with the obtained three characteristic relaxation times have been tentatively proposed. Moreover, the kinetics of vesicle formation in the later stage (from 3 min onward) was investigated by dynamic LLS. It was found that both the intensity-averaged hydrodynamic radius, R h, and the polydispersity, mu2/Gamma (2), decrease exponentially, yielding a characteristic relaxation time of approximately 350 s. To our knowledge, this is the first report on the kinetics of the unimer-to-vesicle transition of a stimulus-responsive diblock copolymer. The kinetics of vesicle dissociation for a pH jump from 12 to 2 was also investigated. The breakdown of polymeric vesicles is extremely fast and is independent of polymer concentration; it is complete within approximately 5 ms and is in marked contrast to the much slower rate of vesicle formation.  相似文献   

20.
Equilibrium, kinetics and thermodynamic aspects of sorption of Promethazine hydrochloride (PHCl) onto iron rich smectite (IRS) from aqueous solution were investigated. The effect of pH on sorption of PHCl onto IRS was also found out. Experimental data were evaluated by using Langmuir, Freundlich and Dubinin–Raduschkevich (DR) isotherm equations. Freundlich and DR equations provided better compatibility than Langmuir equation. Besides, it was determined that the maximum sorption of PHCl takes place at about pH 5. From kinetic studies, it was obtained that sorption kinetics follow pseudo-second-order kinetic model for PHCl sorption onto IRS. When thermodynamic studies are concerned, the values of activation energy (Ea), ΔG°, ΔH° and ΔS° were obtained. ΔG° values are in the range of −8.84 and −9.45 kJ mol−1 indicating spontaneous nature of physisorption. The negative value of the ΔH° (−3.20 kJ mol−1) indicates exothermic nature of adsorption. FTIR analysis and SEM observations of IRS and PHCl adsorbed IRS were also carried out. Sorption experiments indicate that IRS may be used effectively for the adsorption of PHCl.  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