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 15 毫秒
1.
Regeneration of bacteriorhodopsin from bacterioopsin and all-trans-retinal was studied in a mixed micelle system consisting of dodecyl sulfate, CHAPS and a water-soluble phospholipid dihexanoylphosphatidylcholine (hex2-PhosChol). Regeneration to approximately 40,000 M-1.cm-1 extinction at 550 nm (epsilon 550) was obtained with either 2.3 mM or 6.5 mM CHAPS along with 6.9 mM dodecyl sulfate and 4.5 mM hex2-PhosChol in 0.16 M NaCl and 40 mM phosphate (pH 6.0). Without CHAPS, the regeneration in 4.5 mM Hex2-PhosChol gave epsilon 555 = 27,800; without PhosChol, the 1:3 CHAPS/dodecyl sulfate mixture gave epsilon 550 approximately 20,000; and without PhosChol the nearly equimolar CHAPS/dodecyl sulfate mixture gave epsilon 550 approximately 10,000. The composition of the mixed micelles was estimated from fluorescence spectroscopy using pyrene butyryl hydrazine. The molecular weight was estimated by molecular seive chromatography to be 87,100 for 2.3 mM CHAPS, 6.9 mM dodecyl sulfate and 0.67 mM hex2-PhosChol; and 83,200 for 7.0 mM CHAPS, 6.9 mM dodecyl sulfate, and 1.1 mM hex2-PhosChol. These results are consistent with the idea that at low concentrations of CHAPS and dodecyl sulfate, CHAPS organizes the dodecyl sulfate into disk shaped bilayer micelles that are favorable for bacterioopsin refolding. However, a high concentration of either detergent inhibits regeneration. Added hex2-PhosChol can overcome the inhibitory effects of high concentrations of either CHAPS or dodecyl sulfate.  相似文献   

2.
Earlier neutron small-angle scattering experiments had revealed the low resolution structure of the complex between sodium dodecyl sulfate (SDS) and the single polypeptide (452 amino acid residues) of a water-soluble enzyme. The saturated complex consists of three globular micelles which are connected by short flexible polypeptide segments. New experiments, described here, were performed at subsaturating concentrations of free SDS in equilibrium with the complex. The data show a decrease in stoichiometry from one bound dodecyl sulfate (DS) anion per two amino acid residues near the critical micelle concentration (CMC) to one per four residues at half the CMC. At 0.3 CMC, a two-micelle complex is formed by the recombination of the small amino-terminal micelle with the middle one; and the center-to-center distance between the carboxyl-terminal micelle and the middle one decreases from 7.5 to 6.2 nm. These structural data allow us to better understand earlier results obtained with high-performance agarose gel chromatography of the same SDS-protein complexes.  相似文献   

3.
The interactions between the diblock copolymer S(15)E(63) and the surfactants sodium dodecyl sulfate (SDS), sodium decyl sulfate (SDeS), and sodium octyl sulfate (SOS) have been investigated by dynamic light scattering (DLS), transmission electron microscopy (TEM), and isothermal titration calorimetry (ITC). The surfactants with the same headgroup differentiate in their chain length. At 20 degrees C, the block copolymer is associated into micelles with a hydrodynamic radius of 11.6 nm, which is composed of a hydrophobic styrene oxide (S) core and a water-swollen oxypolyethylene (PEO or E) corona. The different copolymer/surfactant systems have been studied at a constant copolymer concentration of 2.5 g dm(-3) and in a vast range of surfactant concentrations, from 7.5 x 10(-6) up to 0.75 M. When SDS and SDeS are added to the block copolymer solution, different regions are observed in the DLS data: at low surfactant concentrations (c < 1.0 x 10(-4) M), single surfactant molecules associate with the copolymer micelle, probably the former being solubilized in the micelle core, leading to a certain disruption of the mixed micelle due to repulsive electrostatic interactions between surfactant headgroups followed by a stabilization of the mixed micelle. At higher concentrations (1.0 x 10(-4) < c < 0.1 M), two types of copolymer-surfactant complexes coexist: one large copolymer-rich/surfactant complex and one small complex consisting of one or a few copolymer chains and rich in surfactants. At higher SDS and SDeS concentrations, complete disintegration of mixed micelles takes place. In contrast, SOS-S(15)E(63) interactions are less important up to surfactant concentrations of 0.05 M due to its higher hydrophilicity, reducing the hydrophobic interactions between surfactant alkyl chains and copolymer micelles. At concentration larger than the critical aggregation concentration (cac) of the system, 0.05 M, disruption of copolymer micelles occurs. These regions have been confirmed by transmission electron microscopy. On the other hand, the titration calorimetric data for SDS and SDeS present an endothermic increase indicating the formation of mixed copolymer-rich-surfactant micelles. From that point, important differences in the ITC plot for both surfactants are present. However, the ITC curve obtained after titration of a SOS solution in the copolymer solution is quite similar to that of its titration in water.  相似文献   

4.
The structure of the complex between sodium dodecyl sulfate (SDS) and a deuterated bifunctional enzyme, N-5'-phosphoribosylanthranilate isomerase/indole-3-glycerol-phosphate synthase (Mr 49,484), has been studied in dilute solution by small-angle neutron scattering. The complex nearly acquired its final size, as shown by molecular-sieve chromatography, at the chosen SDS concentration of 1.6 mM, which is slightly below the critical micelle concentration of 1.8 mM (at the ionic strength of 0.1 M). The 452 amino-acid residues of the bifunctional enzyme were combined with 216 detergent molecules. The complex was found to be composed of three protein-decorated SDS micelles of unequal size, connected by short flexible polypeptide segments. The largest of the three micelles was the middle one. The SDS-protein complex contained the dodecyl hydrocarbon moieties in three globular cores. Each core was surrounded by a hydrophilic shell, formed by the hydrophilic and amphiphilic stretches of the polypeptide chain, and by the sulfate head groups of the detergent. The average thickness of these shells was 0.7-0.8 nm. The three-micelle complex was cleaved with trypsin at a single site, possibly in a micelle-connecting segment, into a single-micelle fragment at the carboxyl-terminal which comprised 73 SDS molecules and 163 amino-acid residues, and a dual-micelle fragment. One of the micelles within this larger fragment contained 42 SDS molecules and about 90 amino-acid residues; the other micelle contained 101 SDS molecules and about 190 amino-acid residues. The individual micelle sizes seemed to be determined by the amino-acid sequence.  相似文献   

5.
The vesicle-to-micelle transition of egg phosphatidylcholine LUVs induced by octylglucoside was studied in buffers with 0-4 M sodium chloride, sucrose or urea. We used both light scattering and fluorescent probes to follow the lipid-detergent complexes in these buffers. The vesicle-to-micelle transition process was fundamentally the same in each solute. However, the detergent-to-lipid ratio required for micelle formation shifted in ways that depended on the aqueous solute. The partitioning of octylglucoside between the vesicles and the aqueous phase was primarily determined by the change in its critical micelle concentration (cmc) induced by each solute. Specifically, the cmc decreased in high salt and sucrose buffers but increased in high concentrations of urea. Cmc for two additional nonionic detergents, decyl- and dodecyl-maltoside, and three zwittergents (3-12, 3-14 and 3-16) were determined as a function of concentration for each of the solutes. In all cases NaCl and sucrose decreased the solubility of the detergents, whereas urea increased their solubilities. The effects clearly depended on acyl chain length in urea-containing solutions, but this dependence was less clear with increasing NaCl and sucrose concentrations. The contributions of these solutes to solubility and to interfacial interactions in the bilayers, pure and mixed micelles are considered.  相似文献   

6.
We examined the ability of NaCl (at 0.15 to 3 M) to release non-histone proteins from chromatin of cultured rat hepatoma cells. The percentage of the non-histones released increased with increasing NaCl concentrations up to 0.75 M; 1 and 3 M NaCl were not significantly more effective. A maximum of 50% of the non-histone protein was recovered free of DNA. The release of non-histones from sheared and unsheared chromatin was similar. The electrophoretic patterns of the non-histone proteins released by NaCl resembled that of the non-histones released by sodium dodecyl sulfate, which indicates that many of the detectable components were at least partially released by NaCl. Some non-histones (especially low molecular weight polypeptides) were fully released by NaCl and other proteins were relatively resistant to NaCl release. Higher recoveries of NaCl-dissociated non-histones were obtained with sucrose gradient centrifugation than with centrifugation in the absence of sucrose.  相似文献   

7.
The absorbance maximum, lambda max, of a local anesthetic, benzyl alcohol, is shifted to longer wavelengths when solvent polarity is decreased. The shift was approximately a linear function of the dielectric constant of the solvent. This transition in electronic spectra according to the microenvironmental polarity is used to analyze benzyl alcohol binding to surfactant micelles. A facile method is devised to estimate the micelle/water partition coefficient from the dependence of lambda max of benzyl alcohol on surfactant concentrations. The effective dielectric constants of the sodium decyl sulfate, dodecyl sulfate and tetradecyl sulfate micelles were 29, 31 and 33, respectively. The partition coefficient of benzyl alcohol between the micelles and the aqueous phase was 417, 610 and 1089, respectively, in the mole fraction unit. The pressure dependence of the partition coefficient was estimated from the depression of the critical micelle concentration of sodium dodecyl sulfate by benzyl alcohol under high pressure up to 200 MPa. High pressure squeezed out benzyl alcohol molecules from the micelle until about 120 MPa, then started to squeeze in when the pressure was further increased. The volume change of benzyl alcohol by transfer from the aqueous to the micellar phase was calculated from the pressure dependence of the partition coefficient. The volume change, estimated from the thermodynamic argument, was 3.5 +/- 1.1 cm3.mol-1 at 298.15 K, which was in reasonable agreement with the partial molal volume change determined directly from the solution density measurements, 3.1 +/- 0.2 cm3.mol-1. Benzyl alcohol apparently solvates into the micelles close to surface without losing contact with the aqueous phase.  相似文献   

8.
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   

9.
The moderate halophile Vibrio costicola, growing on a chemically-defined medium, transformed choline into glycine betaine (betaine) by the membrane-bound enzyme choline dehydrogenase and the cytoplasmic enzyme betainal (betaine aldehyde) dehydrogenase. Choline dehydrogenase was strongly induced and betainal dehydrogenase less strongly induced by choline. The formation of these enzymes was also regulated by the NaCl concentration of the growth medium, increasing with increasing NaCl concentrations. Intracellular betaine concentrations also increased with increasing choline and NaCl concentrations in the medium. This increase was almost completely blocked by chloramphenicol, which does not block the increase in salt-tolerant active transport on transfer from a low to a high salt concentration.Choline dehydrogenase was inhibited by chloride salts of Na+, K+, and NH inf4 su+ , the inhibition being due to the Cl- ions. Betainal dehydrogenase was stimulated by 0.5 M salts and could function in up to 2.0 M salts.Cells grew as well in the presence as in the absence of choline in 0.5 M and 1.0 M NaCl, but formed no intracellular betaine. Choline stimulated growth in 2.0 M NaCl and was essential for growth in 3.0 M NaCl. Thus, while betaine is important for some of the adaptations to high salt concentration by V. costicola, it by no means accounts for all of them.Abbreviations CDMM chemically-defined minimal medium - PPT proteose-peptone tryptone medium - SDS sodium dodecyl sulfate Deceased, 1987  相似文献   

10.
Previously characterized as being non-luminescent in room-temperature fluid solution, the coordination compound chloro(2,2′:6′,2″-terpyridine)platinum(II) chloride can display two types of luminescence in certain microenvironments. In aqueous solutions of anionic and neutral surfactants having concentrations near or above their critical micelle concentration, [Pt(terpy)Cl]Cl (10-50 μM) displays broad emission centered at ∼610 nm that is characterized as metal-to-ligand charge transfer phosphorescence (3MLCT). In high concentration (10-100 mM) solutions having no surfactant, [Pt(terpy)Cl]Cl aggregates form. Excitation in the 470-540 nm region results in a long-wavelength emission centered at ∼720 nm that is characterized as metal-metal-to-ligand charge transfer phosphorescence (3MMLCT). This emission can also be detected in lower concentration solutions (10-50 μM) with surfactant concentration below its critical micelle concentration. Enhancement of 3MLCT luminescence is also found for the related phenylacetylide complex cation [Pt(terpy)(CCPh)]+ in micelles of the anionic surfactant sodium dodecyl sulfate.  相似文献   

11.
Absorption, fluorescence and excitation fluorescence spectra of pheophytin a have been measured in aqueous solutions of nonionic (Triton X-100), anionic (sodium lauryl sulphate) and cationic (Cetyl pyridinium chloride) detergents at different concentrations and pH after system relaxation. By measuring the second derivative and differential spectra, it has been shown, that if detergent concentrations are lower than critical micelles concentration, or if the detergent is completely absent, the pigment forms conglomerates containing both dimeric and monomeric forms with an efficient energy transfer between them. If detergent concentrations are higher than critical micelles concentration, pheophytin a localizes in detergent micelle in monomeric form at neutral and acidic pH, and allomerizes at alkaline pH. The spectral characteristics of pheophytin a dimers in the conglomerate and its monomers in micelles poorly (if at all) depend on the sign of the detergent molecule charge.  相似文献   

12.
Ozone is shown to react with lysozyme in reverse micelles formed by 0.1 M sodium di-2-ethylhexylsulfosuccinate and 1.2-3 M water (pH 7.4) in isooctane solvent. The reaction of ozone is assessed by the oxidation of tryptophan residues in the protein to N-formylkynurenine. Cosolubilization of oleate in lysozyme-containing reverse micellar solutions at concentrations of 0.5-10 mM results in a progressive inhibition (19% to 82%) of the oxidation of tryptophan residues with a concentration for 50% inhibition around 2 mM. At this concentration of oleate, the magnitude of inhibition is independent of the micelle size and concentration, the overall interfacial area of reverse micelles, and the amount of ozone employed. These findings are discussed in terms of competitive reactions of ozone with unsaturated fatty acids and proteins in the lung lining fluid and in biological membranes.  相似文献   

13.
Summary Hydrophobie zeolite Y was used to adsorb detergents from protein solutions and within one minute the commonly used detergents sodium dodecyl sulfate, cetyl trimethyl ammonium bromide, and Triton X-100 at concentrations of 10 mg/ml were adsorbed to a level below their critical micelle concentrations. From the detergent depleted solutions 77 to 85 % of the proteins were recovered; the lower value was obtained with protein concentration below one mg/ml.  相似文献   

14.
Meniscus depletion sedimentation equilibrium ultracentrifuge experiments were performed on purified MoFe and Fe proteins of Azotobacter vinelandii. The MoFe protein was found to have a molecular weight of 245,000, using an experimentally confirmed partial specific volume of 0.73. The MoFe protein formed one band on sodium dodecyl sulfate gel electrophoresis and had a subunit molecular weight of 56,000. The subunit molecular weight from ultracentrifuge experiments in 8 M urea was 61,000. The molecular weight of the Fe protein was calculated to be 60,500 in meniscus depletion experiments. Similar experiments in 8 M urea solvent indicated a subunit molecular weight of 30,000. A subunit molecular weight of 33,000 was obtained from sodium dodecyl sulfate gel electrophoresis experiments.  相似文献   

15.
In view of the low solubility of calcium deoxycholate and the possible induction of cholesterol precipitation in the gallbladder by calcium insoluble salts, we find it of interest to study the precipitation of calcium deoxycholate and its dependence on other bile components. The findings of these studies were as follows: (i) Precipitation of calcium deoxycholate from mixtures of calcium chloride and monomeric deoxycholate (at concentrations below the critical micelle concentration (CMC] is very slow even at relatively high CaCl2 concentrations (more than 20 days at 50 mM CaCl2). (ii) At higher deoxycholic acid (DOC) concentrations, precipitation of micellar DOC is faster and requires much lower calcium chloride concentrations. For any given calcium concentration, the rate of precipitation is maximal at an optimal DOC concentration. In solutions containing 150 mM NaCl, the maximal rate of precipitation occurs at about 10 mM DOC, almost independent of Ca2+ concentration. At lower ionic strength (10 mM NaCl), the optimal DOC concentration is 30 mM. These observations suggest that the most important factors in determining the rate of Ca(DOC)2 precipitation are (a) the ratio between calcium ions bound to the surface of a DOC micelle, and the [DOC] (the Ca2+/DOC binding ratio) and (b) the concentration of DOC micelles. (iii) In the presence of conjugated deoxycholates, the crystallization of calcium deoxycholate is inhibited. Phosphatidylcholine has a similar, although smaller, inhibitory effect. Upon precipitation of calcium deoxycholate from a mixed micellar system containing sodium deoxycholate, phosphatidylcholine and cholesterol, the latter two components spontaneously form vesicles. The anti-nucleating effect of PC and conjugated bile salts is explained in terms of "poisoning" of the crystallization process. In view of the latter results we conclude that under normal conditions calcium deoxycholate is not likely to precipitate in the gallbladder.  相似文献   

16.
A discontinuous sodium dodecyl sulfate-polyacrylamide slab gel electrophoresis system is described which provides superior resolution of polypeptides with molecular weights from approximately 2500 to 90,000. The system utilizes a relatively low-mobility acetate ion in the stacking gel and high-mobility strong anions, sulfate and chloride, as leading and trailing ions in the separating gel. The entire system is run at pH 7.8. The separating gel contains 8 M urea, and can be used at acrylamide concentrations from 5 to 18%, all with 5% crosslinker concentrations. Using a number of protein standards, the calibration curves obtained with this system are linear over the molecular weight range from 2500 to 90,000, regardless of acrylamide concentration. These studies indicate that by providing good resolution of small peptides, this system greatly extends the utility of one-dimensional peptide mapping techniques.  相似文献   

17.
The laser-induced pH jump (Gutman, M. and Huppert, D.J. (1979) Biochem. Biophys. Methods 1, 9–19) has a time resolution capable of measuring the diffusion-controlled rate constant of proton binding. In the present study we employed this technique for measuring the kinetics of protonation-deprotonation of surface groups of macromolecules.The heterogeneous surface of proteins excludes them from serving as a simple model, therefore we used micelles of a neutral detergent (Brij 58) as a high molecular weight structure. The charge was varied by the addition of a low concentration of sodium dodecyl sulfate and the surface group with which the protons react was an adsorbed pH indicator (bromocresol green or neutral red).The dissociation of a proton from adsorbed bromocresol green is slower than that from free indicator. This effect is attributed to the enhanced stabilization of the acid form of the indicator in the pallisade region of the micelle. The pK shift of bromocresol green adsorbed on neutral micelles is thus quantitatively accounted for by the decreased rate of proton dissociation. Indicators such as neutral red, which are more lipid soluble in their alkaline form, do not exhibit such decelerated proton dissociation in their adsorbed state nor a pK shift on adsorption to neutral micelles.The protonation of an indicator is a diffusion-controlled reaction, whether it is free in solution or adsorbed on micelles. By varying the electric charge of the micelle this rate can be accelerated or decelerated depending on the total charge of the micelle. The micellar charge calculated from this method was corroborated by other measurements which rely only on equilibrium parameters.The high time resulation of the pH jump is exemplified by the ability to estimate the diffusion coefficient of protons through the hydrated shell of the micelle.  相似文献   

18.
The effects of sodium dodecyl sulfate on the fluorescent labeling of proteins were studied. Of 57 primary amine groups in bovine serum albumin, no more than 7 are titrable by fluorescamine. Fluorescamine labeling does not cause appreciable conformational changes of proteins. The extent of labeling of proteins decreases as the concentrations of sodium dodecyl sulfate increases. The fluorescence properties of labeled primary amine are only slightly affected by the polarities of the solvents. The inhibitory effects of sodium dodecyl sulfate upon labelings are interpreted as the low permeability of fluorescamine toward the highly charged envelopes of sodium dodecyl sulfate-protein micelles.  相似文献   

19.
The interaction of the Alzheimer's amyloid beta peptide, Aβ40, with sodium dodecyl sulfate (SDS) micelles, together with the self-assembly of SDS molecules around the peptide from an initial random distribution were studied using atomistic and coarse-grained (CG) molecular dynamics simulations. In atomistic simulations, the peptide structure in the micelle was characterized by two helical regions connected through a short hinge. The initial structure of the system was shown to affect the simulation results. The atomistic self-assembly of SDS molecules resulted in a 38-molecule micelle around the peptide, along with some globules and individual molecules. Coarse-grained simulation results, however, did not show such a difference, and at the end of all CG simulations, a complete 60-molecule micelle was obtained, with the peptide located at the interface of the micelle with water. The obtained CG radial density profiles and SDS micelle size and shape properties were identical for all CG simulations.  相似文献   

20.
In this work, osmotic second virial coefficients (B(22)) were determined and correlated with the measured solubilities for the proteins, α-amylase, ovalbumin, and lysozyme. The B(22) values and solubilities were determined in similar solution conditions using two salts, sodium chloride and ammonium sulfate in an acidic pH range. An overall decrease in the solubility of the proteins (salting out) was observed at high concentrations of ammonium sulfate and sodium chloride solutions. However, for α-amylase, salting-in behavior was also observed in low concentration sodium chloride solutions. In ammonium sulfate solutions, the B(22) are small and close to zero below 2.4 M. As the ammonium sulfate concentrations were further increased, B(22) values decreased for all systems studied. The effect of sodium chloride on B(22) varies with concentration, solution pH, and the type of protein studied. Theoretical models show a reasonable fit to the experimental derived data of B(22) and solubility. B(22) is also directly proportional to the logarithm of the solubility values for individual proteins in salt solutions, so the log-linear empirical models developed in this work can also be used to rapidly predict solubility and B(22) values for given protein-salt systems.  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