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.
Interactions of anionic polyelectrolyte (PE) with cationic monomeric (MS) and dimeric surfactants (DS) have been investigated by coarse-grained molecular dynamics (MD) simulation. A PE/surfactant mixture is observed to evolve over time into micellar complex of increasing size. The critical aggregation concentration (CAC) is qualitatively found to be much lower than the critical micellization concentration (CMC) of the free surfactant. Compared to the monomeric analog, a DS interacts more strongly with the oppositely charged polyion chain. The equilibrium complex size becomes larger with increasing surfactant concentration. Simulation results are consistent with experimental observations and reveal that the electrostatic and hydrophobic interactions play an important role in the formation of micellar complex.  相似文献   

2.
The interactions between the polysaccharide alginate with charged ionic surfactants (anionic and cationic) in aqueous solution have been investigated using pyrene as a photophysical probe. Static fluorescence determinations have been used to obtain information about the new microenvironments arising by these interactions. Micropolarity studies using the I(1)/I(3) ratio of the vibronic bands and I(E)/I(M) ratio between the excimer and monomer emissions of pyrene shows the formation of hydrophobic domains. The interactions between the natural polyelectrolytes and the oppositely charged surfactants lead to the formation of pre-micelles at surfactant concentrations lower than the CMC of the surfactants. The aggregation process is assumed to be due to electrostatic attraction. On the other side, systems containing an anionic surfactant do not show the same behaviour at low concentrations.  相似文献   

3.
Equilibrium measurements of the binding of central nervous system myelin basic protein to sodium dodecyl sulphate, sodium deoxycholate and lysophosphatidylcholine have been obtained by gel permeation chromatography and dialysis. This protein associates with large amounts of each of these surfactants: the apparent saturation weight ratios (surfactant/protein) being 3.58 +/- 0.12 and 2.30 +/- 0.15 for dodecyl sulphate at ionic strengths 0.30 and 0.10, respectively 1.34 +/- 0.10 for deoxycholate (at 0.12 ionic strength) and 4.0 +/- 0.5 for lysophosphatidylcholine. Binding to the ionic surfactants increases markedly close to their critical micelle concentrations. Sedimentation analysis shows that at 0.30 ionic strenght in excess dodecyl sulphate the protein is monomeric. It becomes dimeric when the binding ratio falls below 1 at a free detergent concentration of approximately 0.25 mM: below this concentration much of the protein and deterent forms an insoluble complex. The amount of dodecyl sulphate bound at high concentrations and at both above-mentioned ionic strengths corresponds closely to that expected for interaction of a single poly-peptide with two micelles. Variability of deoxycholate micelle size on interaction with other molecules precludes a similar analysis for this surfactant. Association was observed only with single micelles of lysophosphatidylcholine. The results provide strong evidence for dual lipid-binding sites on basic protein and indicate that lipid bilayer cross-linking by this protein may be effected by single molecules.  相似文献   

4.
Because in vitro studies indicate that pulmonary alveolar macrophages (PAM's) filled with phospholipid vesicles have depressed microbicidal capacity, we tested the intrapulmonary bactericidal activity of newborn PAM's after surfactant treatment. Term newborn rabbits received intratracheally either homologous surfactant or one of two artificial phospholipid vesicle preparations followed by pulmonary aerosol infection with group B streptococci (GBS). Four hours after lung infection, phagocytic killing of GBS was reduced by 70-90% in animals treated with the homologous and one of the artificial surfactants compared with untreated animals or animals that received intrapulmonary injections of the surfactant vehicle (P less than 0.02). The other artificial phospholipid preparation decreased intrapulmonary inactivation of GBS by 30-40% compared with the controls. The phospholipid vesicles in the three preparations were avidly ingested and processed by newborn PAM's. The diminished in vivo killing of GBS was not attributed to decreased viability or phagocytic behavior of the PAM's toward GBS. The bactericidal defect that was evident in the newborn PAM's appeared related to the uptake of large phospholipid vesicles in the preparations rather than to the phospholipid content of the surfactants themselves. When in vitro conditions that stimulated the alveolar environment were used, the natural surfactant preparation promoted GBS proliferation, whereas the artificial preparations did not. Our findings indicate that surfactant administration reduces the bactericidal activity of neonatal PAM's. We conclude that additional investigations are needed to ascertain the effect of surfactant replacement therapy on lost defenses of the lung.  相似文献   

5.
The biodegradation of polycyclic aromatic hydrocarbons (PAH) often is limited by low water solubility and dissolution rate. Nonionic surfactants and sodium dodecyl sulfate increased the concentration of PAH in the water phase because of solubilization. The degradation of PAH was inhibited by sodium dodecyl sulfate because this surfactant was preferred as a growth substrate. Growth of mixed cultures with phenanthrene and fluoranthene solubilized by a nonionic surfactant prior to inoculation was exponential, indicating a high bioavailability of the solubilized hydrocarbons. Nonionic surfactants of the alkylethoxylate type and the alkylphenolethoxylate type with an average ethoxylate chain length of 9 to 12 monomers were toxic to a PAH-degrading Mycobacterium sp. and to several PAH-degrading mixed cultures. Toxicity of the surfactants decreased with increasing hydrophilicity, i.e., with increasing ethoxylate chain length. Nontoxic surfactants enhanced the degradation of fluorene, phenanthrene, anthracene, fluoranthene, and pyrene.  相似文献   

6.
Kinetic and steady-state measurements of pyrene fluorescence in a variety of model membranes are evaluated in terms of the theory of collisional excimer formation. In the region of 10(-3)-0.1 M pyrene, molecular fluorescence decay in membranes is biphasic and the two component lifetimes do not depend on the pyrene concentration. The lifetime data are consistent with the rate constant for collisional excimer formation being of the order 10(6) M-1 X s-1 or less. The concentration dependence of the component amplitudes is inconsistent with the theory of collisional excimer formation and suggests that pyrene exists in two forms in membranes: a slowly diffusing monomeric form and an aggregated form. The component of molecular fluorescence decay associated with aggregated pyrene is highly correlated with steady-state excimer fluorescence, suggesting that excimer fluorescence in membranes arises from aggregated pyrene in which excimers are formed by a static rather than a collisional mechanism. It is suggested that the concentration dependence of excimer to molecular fluorescence intensity ratios in membranes is related to the equilibrium constant for exchange between monomeric and aggregated pyrene forms rather than to the collisional excimer formation rate constant.  相似文献   

7.
The hydrophobicity of the bacterial cell surface was determined by using nonionic surfactants. The method is based on the adsorption of nonionic surfactants at the hydrophobic sites of the cell surface. Among many nonionic surfactants, C18H37O(CH2CH2O)13H was preferred. The surfactant was added in excess to a bacterial suspension, and the suspension was mixed by sonication or mechanical stirring. The amount of surfactant remaining in the supernatant after centrifugation was determined spectrophotometrically by measuring the absorbance of tetrabromophenolphthalein ethylester. Effective dispersion of bacterial cells such as Staphylococcus aureus and Mycobacterium smegmatis was achieved by sonication in the presence of the nonionic surfactant. Adsorption measurements coincided with Langmuir's equation, indicative of monolayer adsorption. The method is useful for the determination of the hydrophobicity of various bacterial cell surfaces.  相似文献   

8.
Phospholipid flippases translocate phospholipids from the exoplasmic to the cytoplasmic leaflet of cell membranes to generate and maintain phospholipid asymmetry. The genome of budding yeast encodes four heteromeric flippases (Drs2p, Dnf1p, Dnf2p, and Dnf3p), which associate with the Cdc50 family noncatalytic subunit, and one monomeric flippase Neo1p. Flippases have been implicated in the formation of transport vesicles, but the underlying mechanisms are largely unknown. We show here that overexpression of the phosphatidylserine synthase gene CHO1 suppresses defects in the endocytic recycling pathway in flippase mutants. This suppression seems to be mediated by increased cellular phosphatidylserine. Two models can be envisioned for the suppression mechanism: (i) phosphatidylserine in the cytoplasmic leaflet recruits proteins for vesicle formation with its negative charge, and (ii) phosphatidylserine flipping to the cytoplasmic leaflet induces membrane curvature that supports vesicle formation. In a mutant depleted for flippases, a phosphatidylserine probe GFP-Lact-C2 was still localized to endosomal membranes, suggesting that the mere presence of phosphatidylserine in the cytoplasmic leaflet is not enough for vesicle formation. The CHO1 overexpression did not suppress the growth defect in a mutant depleted or mutated for all flippases, suggesting that the suppression was dependent on flippase-mediated phospholipid flipping. Endocytic recycling was not blocked in a mutant lacking phosphatidylserine or depleted in phosphatidylethanolamine, suggesting that a specific phospholipid is not required for vesicle formation. These results suggest that flippase-dependent vesicle formation is mediated by phospholipid flipping, not by flipped phospholipids.  相似文献   

9.
The scientific study of protein surfactant interactions goes back more than a century, and has been put to practical uses in everything from the estimation of protein molecular weights to efficient washing powder enzymes and products for personal hygiene. After a burst of activity in the late 1960s and early 1970s that established the general principles of how charged surfactants bind to and denature proteins, the field has kept a relatively low profile until the last decade. Within this period there has been a maturation of techniques for more accurate and sophisticated analyses of protein-surfactant complexes such as calorimetry and small angle scattering techniques. In this review I provide an overview of different useful approaches to study these complexes and identify eight different issues which define central concepts in the field. (1) Are proteins denatured by monomeric surfactant molecules, micelles or both? (2) How does unfolding of proteins in surfactant compare with "proper" unfolding in chemical denaturants? Recent work has highlighted the role of shared micelles, rather than monomers, below the critical micelle concentration (cmc) in promoting both protein denaturation and formation of higher order structures. Kinetic studies have extended the experimentally accessible range of surfactant concentrations to far above the cmc, revealing numerous different modes of denaturation by ionic surfactants below and above the cmc which reflect micellar properties as much as protein unfolding pathways. Uncharged surfactants follow a completely different denaturation strategy involving synergy between monomers and micelles. The high affinity of charged surfactants for proteins means that unfolding pathways are generally different in surfactants versus chemical denaturants, although there are common traits. Other issues are as follows: (3) Are there non-denaturing roles for SDS? (4) How reversible is unfolding in SDS? (5) How do solvent conditions affect the way in which surfactants denature proteins? The last three issues compare SDS with "proper" membranes. (6) Do anionic surfactants such as SDS mimic biological membranes? (7) How do mixed micelles interact with globular proteins? (8) How can mixed micelles be used to measure the stability of membrane proteins? The growing efforts to understand the unique features of membrane proteins have encouraged the development of mixed micelles to study the equilibria and kinetics of this class of proteins, and traits which unite globular and membrane proteins have also emerged. These issues emphasise the amazing power of surfactants to both extend the protein conformational landscape and at the same time provide convenient and reversible short-cuts between the native and denatured state for otherwise obdurate membrane proteins.  相似文献   

10.
Equilibrium measurements of the binding of central nervous system myelin basic protein to sodium dodecyl sulphate, sodium deoxycholate and lysophosphatidylcholine have been obtained by gel permeation chromatography and dialysis. This protein associates with large amounts of each of these surfactants: the apparent saturation weight ratios (surfactant/protein) being 3.58 ± 0.12 and 2.30 ± 0.15 for dodecyl sulphate at ionic strengths 0.30 and 0.10, respectively, 1.34 ± 0.10 for deoxycholate (at 0.12 ionic strength) and 4.0 ± 0.5 for lysophosphatidylcholine. Binding to the ionic surfactants increases markedly close to their critical micelle concentrations. Sedimentation analysis shows that at 0.30 ionic strength in excess dodecyl sulphate the protein is monomeric. It becomes dimeric when the binding ratio falls below 1 at a free detergent concentration of approximately 0.25 mM: below this concentration much of the protein and detergent forms an insoluble complex. The amount of dodecyl sulphate bound at high concentrations and at both above-mentioned ionic strengths corresponds closely to that expected for interaction of a single polypeptide with two micelles. Variability of deoxycholate micelle size on interaction with other molecules precludes a similar analysis for this surfactant. Association was observed only with single micelles of lysophosphatidylcholine. The results provide strong evidence for dual lipid-binding sites on basic protein and indicate that lipid bilayer cross-linking by this protein may be effected by single molecules.  相似文献   

11.
The recombination of dimers of immunoglobulin peptide chains   总被引:13,自引:3,他引:10       下载免费PDF全文
1. Both the gamma and light peptide chains of human pooled and myeloma immunoglobulin G can be prepared as non-aggregating dimers at pH5.4 in 4mm-sodium acetate buffer. The dimeric state is maintained by non-covalent bonds, since the formation of interchain disulphide bonds was prevented by alkylation of the thiol groups. In the case of the light chains there is some evidence that the dimers are in equilibrium with a small amount of monomer. 2. When such dimers of the gamma and light chains are mixed at pH5.4 in 4mm-sodium acetate buffer they combine rapidly, yielding a product that resembles the original immunoglobulin G in its physicochemical and antigenic properties. However, the original optical rotatory dispersion spectrum was regained only with the homogeneous myeloma protein. The recombined pooled immunoglobulin G had a spectrum slightly different from the original, suggesting that at least some of the recombinant molecules had not regained native conformations. 3. Dimers of gamma chains stabilized by interchain disulphide bonds were able to recombine with light chains. However, light chains stabilized in the dimeric state by interchain disulphide bonds would not combine with gamma chains. 4. The chains of rabbit immunoglobulin G behave similarly to the human chains in this system, apart from the alkylated light chains showing clearer evidence of monomeric components.  相似文献   

12.
The interaction of B18 peptide with surfactants has been studied by circular dichroism spectroscopy and fluorescence measurements. B18 is the fusogenic motif of the fertilization sea urchin protein. The peptide forms an alpha-helix structure when interacting with positively or negatively charged surfactants below and above the critical micellar concentration (CMC). The alpha-helix formation is due to binding of surfactant monomers rather than the formation of surfactant micelles on the peptide. Fluorescence measurements show that the CMC of the negatively charged surfactant increases in the presence of B18, supporting the fact that there is a strong interaction between the peptide and monomers. Nonionic surfactant monomers have no effect on the peptide structure, whereas the micelles induce an alpha-helical conformation. In this case the helix stabilization results from the formation of surfactant micelles on the peptide.  相似文献   

13.
表面活性剂对小麦吸收多环芳烃(PAHs)的影响   总被引:14,自引:1,他引:13  
李滢  区自清  孙铁珩 《生态学报》2000,20(1):99-102
通过研究施加两表面活性剂(Tween80和LAS)后小麦对多环芳烃的吸收情况得出,含有过量菲、芘和苯并(a)芘营养液中生长的小麦PAHs含量受表面活性剂影响显著。在培养40d后,CMC以上Tween80使小麦根中菲、芘和苯并(a)芘含量下降,即促进了小麦茎叶中菲和芘的含量。CMC和CMC以下LAS也使小麦中PAHs含量降低而茎叶中PAHs含量增加,但主要是LAS对植物毒害作用结果,与表面活性剂胶束  相似文献   

14.
Four surfactants were tested at five concentrations to determine their abilities to solubilize soil-adsorbed pyrene. Inoculation with pyrene degraders in the presence of the surfactant Witconol SN70 was the most effective treatment for pyrene mineralization (46 to 80%) under unsaturated conditions, but the surfactant inhibited the effectiveness of these inoculants in soil slurries.  相似文献   

15.
The complexes of horse myoglobin (Mb) with the anionic surfactant sodium dodecyl sulfate (SDS), and with the cationic surfactants cetyltrimethylammonium chloride (CTAC) and decyltrimethylammonium bromide (DeTAB), have been studied by a combination of surface tension measurements and optical spectroscopy, including heme absorption and aromatic amino acid fluorescence. SDS interacts in a monomeric form with Mb, which suggests the existence of a specific binding site for SDS, and induces the formation of a hexacoordinated Mb heme, possibly involving the distal histidine. Fluorescence spectra display an increase of tryptophan emission. Both effects point to an increased protein flexibility. SDS micelles induce both the appearance of two more heme species, one of which has the features of free heme, and protein unfolding. Mb/CTAC complexes display a very different behavior. CTAC monomers have no effect on the absorption spectra, and only a slight effect on the fluorescence spectra, whereas the formation of CTAC aggregates on the protein strongly affects both absorption and fluorescence. Mb/DeTAB complexes behave in a very similar way as Mb/CTAC complexes. The surface activity of the different Mb/surfactant complexes, as well as the interactions between the surfactants and Mb, are discussed on the basis of their structural properties.  相似文献   

16.
The gel properties of ovalbumin mixtures with three different surfactants (sodium perfluorooctanoate, sodium octanoate and sodium dodecanoate) have been studied by rheological techniques. The gel elasticities were determined as a function of surfactant concentration and surfactant type. The fractal dimension of the formed structures was evaluated from plots of storage modulus against surfactant concentration. The role of electrostatic, hydrophobic and disulfide SS interactions in these systems has been demonstrated to be the predominant. The viscosity of these structures tends to increase with surfactant concentration, except for the fluorinated one. Unfolded ovalbumin molecules tend to form fibrillar structures that tend to increase with surfactant concentration, except for the fluorinated one. This fact has been related to the particular nature of this molecule.  相似文献   

17.
Rotational diffusion of the neutral probe coumarin 153 has been examined in a mixed surfactant system containing n-octyl-β-D-thioglucoside (OTG) and Triton X-100 (TX100), two non-ionic surfactants belonging to the alkyl glucoside and polyoxyethylene alkyl ether families, respectively. Both steady-state and time-resolved fluorescence measurements indicate that the polarity of the dye decreases slightly as the amount of ethoxylated surfactant in the mixed micelle increases. This behaviour can be attributed to migration of the probe towards the inner region of the micellar palisade layer as a result of the increasing hydration induced by the presence of TX100. The anisotropy decay curves, r(t), were well fitted to a biexponential function, characterized by two reorientational times of the probe in the micellar pseudophase. The experimental data were analyzed on the basis of the two-step and wobbling-in-a-cone model, and the results obtained correlated with the changes that occur in the palisade layer of the mixed micelles due to the different structure and nature of the head groups of both surfactants. It was found that, although the average reorientation time of the probe molecule is almost unaffected with the participation of the ethoxylated surfactant, the observed increase in the generalized order parameter suggests a certain rise in the compactness of the mixed micelles.  相似文献   

18.
The gel-to-liquid-crystalline phase transition of dipalmitoylphosphatidylcholine (DPPC) vesicle membrane was measured in the presence of sodium octanoate (SO) (pH 3 and 10) and sodium perfluorooctanoate (SPFO) (pH uncontrolled) by monitoring the scattered light intensity of the vesicle suspension. The phase transition temperature, Tm, decreased linearly with the concentration of added SO within the measured concentration range; the uncharged form of SO (pH 3) was much more effective for the depression of Tm than the charged form (pH 10). On the other hand, with increasing SPFO concentration, levelling off of Tm was observed after depression at an initial stage. From the depression of Tm, the partition coefficients, K, of these surfactants between bulk solution and DPPC vesicle membrane were estimated and compared with those obtained previously for other surfactant systems. The value of K for charged SO fell on the straight line of log K vs. Nc plot for anionic surfactants, where Nc is the carbon number of the hydrocarbon chain of surfactants, whereas K for uncharged SO showed a large positive deviation from the straight line of the plot for non-ionic surfactants. The latter suggested that some specific interaction, presumably hydrogen bond formation, may act between the protonated carboxyl group of SO and the lipid head group. The K value estimated for SPFO was much larger than that for charged SO. This difference in the affinity for the lipid bilayer between fluorocarbon surfactant and hydrocarbon surfactant may be attributed to the difference in their hydrophobicity.  相似文献   

19.
The circular dichroism studies on calcium ionophore, A23187, incorporated in Dipalmitoyl phosphatidyl choline (DPPC) vesicle showed interesting time dependent changes in the CD spectra. Analysis of the data indicated the possible aggregation of the observed dimeric structure of this molecule in non-polar solvents into a stacked dimeric pore in the phospholipid vesicle.  相似文献   

20.
The state of aggregation and the steady-state size of mixed aggregates made of phospholipids and surfactants are both determined by the surfactant/lipid ratio in the mixed aggregates (Re). Water-soluble polymers, such as dextrans and polyethylene glycols (PEGs) of different molecular weights, induce reversible aggregation of phospholipid vesicles, mostly due to dehydration of the vesicle surface and depletion forces, and only at much higher concentrations, PEGs (but not dextran) also induce irreversible size growth of the vesicles. Here we show that the water-soluble polymers dextrans and PEGs do not affect the vesicle-micelle phase boundaries in mixtures of phosphatidylcholine and the anionic surfactant sodium cholate. By contrast, these polymers affect markedly the steady-state size of cholate-containing vesicles. As compared with pure phosphatidylcholine vesicles, the cholate-containing vesicles have a lower tendency to undergo polymer-induced aggregation, probably due to the electrostatic repulsion between the negatively charged vesicles, but a higher tendency to undergo irreversible size growth at relatively low polymer concentrations. Such irreversible size growth was observed not only for PEG but also for dextran, which in the absence of cholate is incapable of inducing vesicle size growth. These findings are consistent with the prevailing concept that the polymer-induced size growth is due to the effect of large structural fluctuations in the bilayers of deformed aggregated vesicles, the surface of which is dehydrated by the polymer. The presence of cholate in the bilayers at sufficiently high concentrations induces such fluctuations, yielding irreversible size growth within the clusters of dehydrated vesicles formed upon mixing with polymers.  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