首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The partition equilibria of sodium dodecyl sulfate (SDS) and lithium dodecyl sulfate between water and bilayer membranes were investigated with isothermal titration calorimetry and spectroscopic methods (light scattering, (31)P-nuclear magnetic resonance) in the temperature range of 28 degrees C to 56 degrees C. The partitioning of the dodecyl sulfate anion (DS(-)) into the bilayer membrane is energetically favored by an exothermic partition enthalpy of Delta H(O)(D) = -6.0 kcal/mol at 28 degrees C. This is in contrast to nonionic detergents where Delta H(O)(D) is usually positive. The partition enthalpy decreases linearly with increasing temperature and the molar heat capacity is Delta C(O)(P) = -50 +/- 3 cal mol(-1) K(-1). The partition isotherm is nonlinear if the bound detergent is plotted versus the free detergent concentration in bulk solution. This is caused by the electrostatic repulsion between the DS(-) ions inserted into the membrane and those free in solution near the membrane surface. The surface concentration of DS(-) immediately above the plane of binding was hence calculated with the Gouy-Chapman theory, and a strictly linear relationship was obtained between the surface concentration and the extent of DS(-) partitioning. The surface partition constant K describes the chemical equilibrium in the absence of electrostatic effects. For the SDS-membrane equilibrium K was found to be 1.2 x 10(4) M(-1) to 6 x 10(4) M(-1) for the various systems and conditions investigated, very similar to data available for nonionic detergents of the same chain length. The membrane-micelle phase diagram was also studied. Complete membrane solubilization requires a ratio of 2.2 mol SDS bound per mole of total lipid at 56 degrees C. The corresponding equilibrium concentration of SDS free in solution is C (sat)(D,F) approximately 1.7 mM and is slightly below the critical micelles concentration (CMC) = 2.1 mM (at 56 degrees C and 0.11 M buffer). Membrane saturation occurs at approximately 0.3 mol SDS per mol lipid and the equilibrium SDS concentration is C (sat)(D,F)approximately equal 2.2 mM +/- 0.6 mM. SDS translocation across the bilayer is slow at ambient temperature but increases at high temperatures.  相似文献   

2.
R Smith 《Biochemistry》1982,21(11):2697-2701
Self-association of basic protein has been proposed to be of functional significance in central nervous system myelin. In aqueous solution this protein self-associates, previous data being consistent with the formation of dimers, which then undergo an indefinite isodesmic self-association [Smith, R. (1980) Biochemistry 19, 1826-1831]. As this protein is membrane bound in vivo, we have now examined the effects of amphiphiles on the self-association equilibria. Contrary to the expected effects, at low molar ratios dodecyl sulfate, deoxycholate, Triton X-100, and lysophosphatidylcholine increased protein intermolecular attraction. The anionic detergents led to partial precipitation even at 1:1 protein:detergent molar ratios whereas the zwitterionic lipid and the nonionic detergent exerted less pronounced effects. Sedimentation velocity and equilibrium measurements have been used to define quantitatively the effects of lysophosphatidylcholine. The sedimentation coefficient increases up to a lipid:protein ratio of 4:1 and then remains constant up to a ratio of 12:1. The sedimentation equilibrium data suggest that the mode of protein-protein interaction is the same as in the absence of lipid but with substantially increased association constants. The dimerization constant is increased from 1.20 X 10(2) M-1 to 1.0 X 10(3) M-1 and the isodesmic association constant from 3.4 X 10(4) M-1 to 1.2 X 10(5) M-1. The effects of detergents on myelin basic protein are compared with the effects on other proteins, and the implications for the state of the protein with myelin are discussed.  相似文献   

3.
Polyacrylamide gel electrophoresis of highly purified Clostridium perfringens enterotoxin revealed electrophoretic microheterogeneity of the enterotoxin, apparently because of slight charge differences in the peptides. Detergent gel electrophoresis showed that purified enterotoxin formed high molecular weight aggregates in the presence of both sodium dodecyl sulfate (SDS) and cetyltrimethylammonium bromide. No conditions capable of inhibiting this phenomenon were found. Although a molecular weight of 35 000 daltons has been reported in the literature, the experimentally determined molecular weight values in the presence of detergents corresponded to multiples of a theoretical subunit molecular weight of 17 500 daltons. Binding studies performed by equilibrium dialysis and ultracentrifugation methods revealed that the enterotoxin bound very small amounts of SDS per gram of protein. The evidence presented indicates possible detergent induced structural alterations of the protein.  相似文献   

4.
Blue native electrophoresis (BN-PAGE) is presently considered as one of effective methods for the identification of membrane protein complexes. The choice of a nonionic detergent and the detergent to protein ratio are critically important. Our experiments with plasma membranes of etiolated pea (Pisum sativum L.) seedlings showed that various nonionic detergents—digitonin, dodecyl maltoside, and Triton X-100—solubilized similar assortments of protein complexes. Irrespective of the detergent type, PIP aquaporins were always observed in the 440-kD protein complex. Only in the case of dodecyl maltoside, the PIP aquaporins were also revealed in the complexes with the lower and higher molecular weights when the detergent/protein ratio increased.  相似文献   

5.
The interaction of free and immobilized myelin basic protein (MBP) with sodium deoxycholate (DOC) and sodium dodecyl sulfate (NaDodSO4) was studied under a variety of conditions. Free MBP formed insoluble complexes with both detergents. Analysis of the insoluble complexes revealed that the molar ratio of detergent/MBP in the precipitate increased in a systematic fashion with increasing detergent concentration until the complex became soluble. At pH 4.8, equilibrium dialysis studies indicated that approximately 15 mol of NaDodSO4 could bind to the protein without precipitation occurring. Regardless of the surfactant, however, minimum protein solubility occurred when the net charge on the protein-detergent complex was between +18 and -9. Complete equilibrium binding isotherms of both detergents to the protein were obtained by using MBP immobilized on agarose. The bulk of the binding of both detergents was highly cooperative and occurred at or above the critical micelle concentration. At I = 0.1, saturation levels of 2.09 +/- 0.15 g of NaDodSO4/g of protein and 1.03 /+- 0.40 g of DOC/g of protein were obtained. Below pH 7.0 the NaDodSO4 binding isotherms revealed an additional cooperative transition corresponding to the binding of 15-20 mol of NaDodSO4/mol of protein. Affinity chromatography studies indicated that, in the presence of NaDodSO4 (but not in its absence), [125I]MBP interacted with agarose-immobilized histone, lysozyme, and MBP but did not interact with ovalbumin-agarose. These data support a model in which the detergent cross-links and causes precipitation of MBP-anionic detergent complexes. Cross-linking may occur through hydrophobic interaction between detergents electrostatically bound to different MBP molecules.  相似文献   

6.
Wastewater sludge reduced the heat required to inactivate rotavirus SA-11, and ionic detergents were identified as the sludge components responsible for this effect. A similar result was found previously with reovirus (R. L. Ward and C. S. Ashley, Appl. Environ. Microbiol 36:889-897, 1978). The quantitative effects of individual ionic detergents on rotavirus and reovirus were very different, and rotavirus was found to be extremely sensitive to several of these detergents. However, neither virus was destabilized by nonionic detergents. On the contrary, rotavirus was stabilized by a nonionic detergent against the potent destabilizing effects of the ionic detergent sodium dodecyl sulfate. The destabilizing effects of both cationic and anionic detergents on rotavirus were greatly altered by changes in the pH of the medium.  相似文献   

7.
Wastewater sludge reduced the heat required to inactivate rotavirus SA-11, and ionic detergents were identified as the sludge components responsible for this effect. A similar result was found previously with reovirus (R. L. Ward and C. S. Ashley, Appl. Environ. Microbiol 36:889-897, 1978). The quantitative effects of individual ionic detergents on rotavirus and reovirus were very different, and rotavirus was found to be extremely sensitive to several of these detergents. However, neither virus was destabilized by nonionic detergents. On the contrary, rotavirus was stabilized by a nonionic detergent against the potent destabilizing effects of the ionic detergent sodium dodecyl sulfate. The destabilizing effects of both cationic and anionic detergents on rotavirus were greatly altered by changes in the pH of the medium.  相似文献   

8.
We have examined the role of the environment on the interactions between transmembrane helices using, as a model system, the dimerization of the glycophorin A transmembrane helix. In this study we have focused on micellar environments and have examined a series of detergents that include a range of alkyl chain lengths, combined with ionic, zwitterionic, and nonionic headgroups. For each we have measured how the apparent equilibrium constant depends on the detergent concentration. In two detergents we also measured the thermal sensitivity of the equilibrium constant, from which we derive the van't Hoff enthalpy and entropy. We show that several simple models are inadequate for explaining our results; however, models that include the effect of detergent concentration on detergent binding are able to account for our measurements. Our analysis suggests that the effects of detergents on helix association are due to a pair of opposing effects: an enthalpic effect, which drives association as the detergent concentration is increased and which is sensitive to the chemical nature of the detergent headgroup, opposed by an entropic effect, which drives peptide dissociation as the detergent concentration is raised. Our results also indicate that the monomer-monomer interface is relatively hydrophilic and that association within detergent micelles is driven by the enthalpy change. The wide variations in glycophorin a dimmer, stability with the detergent used, together with the realization that this results from the balance between two opposing effects, suggests that detergents might be selected that drive association rather than dissociation of peptide dimers.  相似文献   

9.
P K Werner  R A Reithmeier 《Biochemistry》1985,24(23):6375-6381
Band 3 protein, the anion transport protein of the human erythrocyte membrane, was purified in the presence of the nonionic detergent octyl glucoside. A molecular characterization was carried out to investigate whether the native structure of the protein was retained in the presence of this detergent. Band 3 bound octyl glucoside below the critical micelle concentration (cmc) of the detergent, approaching saturation above the cmc. At 40 mM octyl glucoside, close to saturating concentrations, 0.64 g of octyl glucoside is bound per gram of band 3 protein, corresponding to 208 molecules of detergent bound per monomer of band 3. Sedimentation velocity and gel filtration studies, performed at 40 mM octyl glucoside, indicated that the band 3-octyl glucoside complex had an average molecular weight of 1.98 X 10(6), which corresponds to a dodecamer. Sedimentation equilibrium experiments confirmed that band 3 in octyl glucoside exists in a heterogeneous and high oligomeric state. This high oligomeric state did not change dramatically over octyl glucoside concentrations ranging from 6 to 60 mM. The circular dichroism spectrum of band 3 changed only slightly over this range of octyl glucoside concentrations. The alpha-helical and beta-sheet contents of band 3 in 2 mM octyl glucoside were calculated to be 40% and 27%, respectively, indicating that no gross alteration in the secondary structure of the protein had occurred in octyl glucoside. The ability of band 3 to bind 4-benzamido-4'-aminostilbene-2,2'-disulfonate (BADS), a potent inhibitor (Ki = 1 microM) of anion transport, was measured to assess the integrity of the inhibitor binding site of the protein in octyl glucoside.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
The preparation of a pure and homogeneous protein sample at proper concentration is a prerequisite for success when attempting their crystallization for structural determination. The detergents suitable for solubilization particularly of membrane proteins are not always the best for crystallization. Myelin of the peripheral nervous system of vertebrates is the example of a membrane for which neutral or "gentle" detergents are not even strong enough to solubilize its proteins. In contrast, sodium- or lithium-dodecyl sulfate is very effective. We solubilized myelin membrane in 2%(w/v) sodium dodecyl sulfate, followed by chromatographic purification of the hydrophobic myelin glycoproteins P0 and PASII/PMP22, and finally, we have exchanged the sodium dodecyl sulfate bound to protein for other neutral detergents using ceramic hydroxyapatite column. Theoretically, we should easily exchange sodium dodecyl sulfate for any neutral detergent, but for some of them, the solubility of myelin glycoproteins is low. To monitor the potential variability in the secondary structure of glycoproteins, we have used circular dichroism. Sodium dodecyl sulfate seems to be the appropriate detergent for the purpose of purification of very hydrophobic glycoproteins, since it can be easily exchanged for another neutral detergent.  相似文献   

11.
The major proteins of myelin have classically been extracted in organic solvents. Here we investigated some of the characteristics of brain myelin solubilization in aqueous detergent solutions. At comparable molar concentrations, two nonionic detergents, i.e., octyl glucoside and Lubrol PX, proved relatively better myelin solubilizers than the detergents related to the bile salts, i.e., cholate and CHAPS. The two former detergents solubilized more protein than lipid and the two latter ones more lipid than protein from myelin membranes. All four detergents solubilized the phospholipid more efficiently than the cholesterol component of myelin. The detergent concentrations required for myelin solubilization were reduced substantially if the temperature and the salt concentration of the media were increased. As much as 3 mg of lyophilized myelin (about 1 mg of protein) were solubilized readily per milliliter of a solution containing 30 mM octyl glucoside and 0.1 M sodium sulfate in 0.1 M sodium phosphate buffer, pH 6.7. Each of the detergents studied, including the above four, sodium dodecyl sulfate (SDS). Triton X-100, and Zwittergent 3-14, had its own advantages and drawbacks as myelin protein extractors. The nonionic amphiphiles and CHAPS left a small residue mainly composed of proteins of the Wolfgram fraction, as revealed by SDS-polyacrylamide gel electrophoresis. Octyl glucoside was preferred, given its versatility as solubilizer, ultraviolet transparency, and high critical micellar concentration. Observations on possible difficulties that may be encountered are also included.  相似文献   

12.
The effect of the different detergent mixtures on immunodiffusion and immunoprecipitation was studied. The anionic detergent sodium dodecyl sulfate at concentrations above 0.2% (wv) inhibits the reaction between antigen and antibody by more than 90%. Nonionic detergents at a concentration of 1% (wv) have little or no detectable effect. In contrast, when we used mixtures of various concentrations of ionic and nonionic detergents the inhibitory effect of the ionic detergent decreased.  相似文献   

13.
The binding of deoxycholate and dodecyl sulfate to the mouse and bovine myelin basic proteins and two peptide fragments, obtained by cleavage of the bovine basic protein at its single tryptophan residue, was examined. Complete equilibrium binding isotherms for both detergents were obtained by examining their binding to each of the polypeptides immobilized on agarose. The bulk of the binding of dodecyl sulfate was found to be highly cooperative, and at saturation all four polypeptides bound far more detergent than globular, water-soluble proteins. The sum of the dodecyl sulfate bound by each of the two bovine basic protein cleavage fragments was almost twice that bound by the intact protein at saturation, suggesting that cleavage of the bovine basic protein exposes sites for additional binding of dodecyl sulfate. At pH values below pH 8.0, an additional cooperative transition was observed below the critical micelle concentration of sodium dodecyl sulfate in the binding isotherms of all four polypeptides. The midpoint of this transition corresponded to an apparent pK of approximately 5.5; however, the destruction of 90% of the histidine residues in the bovine basic protein had no effect on this transition. At pH 9.2 and moderate ionic strength (I = 0.1), the bulk of the binding of deoxycholate to the mouse and bovine basic proteins occurred at and above the critical micelle concentration of the detergent; and saturation values of deoxycholate binding to these two proteins were considerably higher than that reported for globular, water-soluble proteins. In marked contrast to the results with dodecyl sulfate, neither cleavage fragment was observed to bind deoxycholate. The results suggest that the higher ordered structure of the bovine basic protein may play an important role in the binding of anionic amphiphiles to the protein.  相似文献   

14.
In our initial attempts to solubilize and purify a chemically synthesized 22-amino acid, membrane-spanning peptide, we encountered numerous difficulties. The peptide was not soluble in dilute acids, organic solvents, or chaotropic agents (+/- detergent) following standard HF cleavage protocols. The insolubility was a direct result of the formation of peptide-(peptide)n aggregates that occurred during the initial phase of isolation, i.e., during the HF cleavage. Eliminating the ether precipitation and subsequent washes did not decrease the degree of aggregation of the product. Inclusion of nonionic detergents in the HF-cleavage reactions displayed little ability in preventing aggregation. Cleavage in the presence of sodium dodecyl sulfate, however, dramatically reduced the degree of aggregation, even after washing with organic solvents. The cleaved peptide was purified to homogeneity using a detergent-based HPLC protocol. This column procedure also permits the quantitative exchange of the sodium dodecyl sulfate for n-octyl-beta-D-glucopyranoside. Combined use of the two protocols results in high-yield isolations for a class of peptides that is generally difficult to handle.  相似文献   

15.
As a preliminary to a study of the biogenesis of individual plasma membrane glycoproteins, the marker enzyme nucleotide pyrophosphatase (NPPase) and a major rat liver plasma membrane sialoprotein, subsequently found to be identical with the enzyme dipeptidyl peptidase IV (DPP IV), were purified 10,000- and 2,000-fold, respectively, from rat liver. Both were amphipathic proteins which formed defined micellar complexes with detergents and aggregated in their absence. Gel filtration, sucrose density gradient centrifugation, and polyacrylamide gel electrophoresis in the presence of sodium dodecyl sulfate showed the Triton X-100 complex of NPPase to contain a single 150,000-dalton peptide, while that of DPP IV was composed of two 120,000-dalton subunits; each complex also contained about 150,000-dalton Triton X-100. Trypsin cleaved the detergent complexes with release of major hydrophilic fragments which no longer bound detergent micelles; the accompanying change in peptide size was small for NPPase and undetectable for DPP IV, which also retained the dimer structure of its native form. DPP IV was the only major glycoprotein in rat liver plasma membrane which bound strongly to wheat germ agglutinin. Monospecific rabbit antibodies against NPPase and DPP IV precipitated the antigens without affecting their enzymatic activities.  相似文献   

16.
A discontinuous gradient polyacrylamide gel electrophoresis under nondenaturing conditions has been used to demonstrate monodispersity of procaryotic and eucaryotic cytochrome c oxidase preparations. Alkaline treated bovine enzyme which contains nine subunits as analysed by subsequent discontinuous SDS-polyacrylamide gel electrophoresis is a monodisperse dimer in 0.1% Triton X-100 and a monomer in 0.1% dodecyl maltoside. The Mr-values corrected for bound detergent are 286,000 in Triton X-100 and 152,000 in dodecyl maltoside respectively. The two-subunit bacterial cytochrome c oxidase of Paracoccus denitrificans is proved to be a monomer with a corrected Mr of 76,000 in both nonionic detergents Triton X-100 and dodecyl maltoside.  相似文献   

17.
The amyloid beta peptide (Abeta) with 39-42 residues is the major component of amyloid plaques found in brains of Alzheimer's disease patients, and soluble oligomeric peptide aggregates mediate toxic effects on neurons. The Abeta aggregation involves a conformational change of the peptide structure to beta-sheet. In the present study, we report on the effect of detergents on the structure transitions of Abeta, to mimic the effects that biomembranes may have. In vitro, monomeric Abeta(1-40) in a dilute aqueous solution is weakly structured. By gradually adding small amounts of sodium dodecyl sulfate (SDS) or lithium dodecyl sulfate to a dilute aqueous solution, Abeta(1-40) is converted to beta-sheet, as observed by CD at 3 degrees C and 20 degrees C. The transition is mainly a two-state process, as revealed by approximately isodichroic points in the titrations. Abeta(1-40) loses almost all NMR signals at dodecyl sulfate concentrations giving rise to the optimal beta-sheet content (approximate detergent/peptide ratio = 20). Under these conditions, thioflavin T fluorescence measurements indicate a maximum of aggregated amyloid-like structures. The loss of NMR signals suggests that these are also involved in intermediate chemical exchange. Transverse relaxation optimized spectroscopy NMR spectra indicate that the C-terminal residues are more dynamic than the others. By further addition of SDS or lithium dodecyl sulfate reaching concentrations close to the critical micellar concentration, CD, NMR and FTIR spectra show that the peptide rearranges to form a micelle-bound structure with alpha-helical segments, similar to the secondary structures formed when a high concentration of detergent is added directly to the peptide solution.  相似文献   

18.
Solubilization and characterization of yeast signal peptidase   总被引:2,自引:0,他引:2  
An efficient post-translational assay for solubilized yeast signal peptidase has been developed. The enzyme can be solubilized in nonionic detergent (0.5% Nikkol) without added salt, but salt increased the efficiency of solubilization. Radiosequencing of the cleaved substrate revealed that the enzyme removed the signal peptide. The substrate (prepro-alpha-factor) must be pretreated with sodium dodecyl sulfate to be cleaved. The enzyme displays a broad, alkaline pH optimum, retaining activity at pH 12. Moderately high temperatures (35 degrees C), excess detergent (greater than 0.5% Nikkol), or high salt (greater than 300 mM KOAc) will inactivate the enzyme. Phosphatidylcholine is necessary for optimal activity. The optimal ratio of Nikkol:lipid:sodium dodecyl sulfate is 6.4:2.2:1. The membrane association of yeast signal peptidase is resistant to carbonate extraction, indicating that it is an integral membrane protein.  相似文献   

19.
The parallel-mode electron paramagnetic resonance (EPR) spectrum of the S(1) state of the oxygen-evolving complex (OEC) shows a multiline signal centered around g=12, indicating an integer spin system. The series of [Mn(2)(2-OHsalpn)(2)] complexes were structurally characterized in four oxidation levels (Mn(II)(2), Mn(II)Mn(III), Mn(III)(2), and Mn(III)Mn(IV)). By using bulk electrolysis, the [Mn(III)Mn(IV)(2-OHsalpn)(2)(OH)] is oxidized to a species that contains Mn(IV) oxidation state as detected by X-ray absorption near edge spectroscopy (XANES) and that can be formulated as Mn(IV)(4) tetramer. The parallel-mode EPR spectrum of this multinuclear Mn(IV)(4) complex shows 18 well-resolved hyperfine lines center around g=11 with an average hyperfine splitting of 36 G. This EPR spectrum is very similar to that found in the S(1) state of the OEC. This is the first synthetic manganese model complex that shows an S(1)-like multiline spectrum in parallel-mode EPR.  相似文献   

20.
G S Lukat  K R Rodgers  H M Goff 《Biochemistry》1987,26(22):6927-6932
Electron paramagnetic resonance (EPR) studies of the nitrosyl adduct of ferrous lactoperoxidase (LPO) confirm that the fifth axial ligand in LPO is bound to the iron via a nitrogen atom. Complete reduction of the ferric LPO sample is required in order to observe the nine-line hyperfine splitting in the ferrous LPO/NO EPR spectrum. The ferrous LPO/NO complex does not exhibit a pH or buffer system dependence when examined by EPR. Interconversion of the ferrous LPO/NO complex and the ferric LPO/NO2- complex is achieved by addition of the appropriate oxidizing or reducing agent. Characterization of the low-spin LPO/NO2- complex by EPR and visible spectroscopy is reported. The pH dependence of the EPR spectra of ferric LPO and ferric LPO/CN- suggests that a high-spin anisotropic LPO complex is formed at high pH and an acid-alkaline transition of the protein conformation near the heme site does occur in LPO/CN-. The effect of tris(hydroxymethyl)aminomethane buffer on the LPO EPR spectrum is also examined.  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