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 734 毫秒
1.
Urea and guanidinium hydrochloride dissociate the 12S protein of mustard and rapeseed to 1.8 S protein and the extent of dissociation depends on the concentration of the denaturant. Mustard (Brassica juncea) protein is more readily dissociated than the rapeseed (Brassica campestris) protein. The reagents denature the protein as evidenced by increase in viscosity, appearance of difference spectra and quenching of fluorescence. Rapeseed protein is denatured more readily than the mustard protein. Analysis of visctosity, spectral and fluoresence data suggests that the first event in the denaturation reaction is the perturbation of the aromatic amino acid residues followed by their exposure to the solvent medium and unfolding of the protein molecule.  相似文献   

2.
The effect of sodium dodecyl sulphate, urea or guanidinium hydrochloride on the sedimentation velocity, viscosity, ultra violet spectra and fluorescence spectra of the 11S protein of guar seed has been determined. Sodium dodecyl sulphate dissociates the protein directly to the 2S protein, whereas urea or guanidinium hydrochloride produces an intermediate 7S protein. These reagents denature the protein also. Both the dissociative and the denaturation effect follow the order, sodium dodecyl sulphate > guanidinium hydrochloride > urea when the concentration are expressed as mols per litre. The denatured states in the three cases probably differ.  相似文献   

3.
Studies of the denaturation and partial renaturation of ovalbumin   总被引:6,自引:4,他引:2       下载免费PDF全文
1. The denaturation of ovalbumin by the reagents sodium dodecyl sulphate and guanidinium chloride was investigated, by following the changes in sedimentation velocity, optical rotatory dispersion and viscosity as a function of denaturant concentration. 2. With sodium dodecyl sulphate both the optical-rotatory-dispersion parameters a(0) and b(0) become more negative, the sedimentation coefficient decreases and the viscosity increases; significant differences in the denaturation profiles are observed. The change in each parameter is indicative of only limited denaturation. 3. With guanidinium chloride the transition occurs over the concentration range 1-4m: more extensive changes occur in all the physical parameters than with sodium dodecyl sulphate. The values of a(0) and b(0) are indicative of complete denaturation. Reduction by mercaptoethanol produces only minor further changes. 4. Renaturation was attempted from both denaturants, the removal of reagent being accomplished reversibly by controlled slow dialysis. Partial renaturation was observed, but aggregated or insoluble material was produced in both cases at relatively low concentrations of denaturant. Similar behaviour was observed with fully reduced protein in guanidinium chloride-mercaptoethanol; complete renaturation could not be brought about even at very low protein concentrations.  相似文献   

4.
The effect of sodium dodecyl sulphate, urea, guanidinium hydrochloride and heat on the oligomeric structure of the 11 S protein of sunflower has been determined. Sodium dodecyl sulphate directly dissociates the protein to 2 S subunits, whereas urea and guanidinium hydrochloride dissociate it through an intermediate 7 S protein. Heating the protein at 90‡C for 20 min caused dissociation of the 11 S protein, without any precipitation.  相似文献   

5.
The effect of anionic detergent, sodium dodecyl sulphate, on the major protein, alpha-globulin of sesame seed (Sesamum indicum L.) has been investigated by gel filtration, sedimentation velocity, viscosity, optical rotation, difference spectra and fluorescence measurements. The detergent causes dissociation of the protein first and then denaturation. In the detergent concentration range of .175-4.0 X 10(-"3) M four components are observed in the ultracentrifuge. The specific rotation of the protein increases with the detergent concentration above 2.5 x 10 (-3) M detergent suggesting conformational change; above 8 X 10(-"3) M detergent the value of -[alpha] does not change. The reduced viscosity etared however, increases above .25 X 10(-3) M detergent and does not attain a plateau value. The difference spectrum of the protein indicates that both tryptophan and tyrosine groups have been affected by the detergent. The fluorescence intensity decreases and the maxima shifts towards red in the detergent solution resulting in an "isoemissive point" at 355 nm. The double difference spectra in sucrose-detergent protein system show that below 5-0 X 10(-3) M detergent, the difference absorption and fluorescence spectrum result from the binding of the detergent near the chromophoric groups and are not due to conformational change. Binding studies by equilibrium dialysis indicate the presence of 50 binding sites in the protein and binding constant of 3-0 X 10(3).  相似文献   

6.
The effects of two ionic surfactants on the oligomeric structure of the giant extracellular hemoglobin of Glossoscolex paulistus (HbGp) in the oxy - form have been studied through the use of several spectroscopic techniques such as electronic optical absorption, fluorescence emission, light scattering, and circular dichroism. The use of anionic sodium dodecyl sulphate (SDS) and cationic cethyltrimethyl ammonium chloride (CTAC) has allowed to differentiate the effects of opposite headgroup charges on the oligomeric structure dissociation and hemoglobin autoxidation. At pH 7.0, both surfactants induce the protein dissociation and a significant oxidation. Spectral changes occur at very low CTAC concentrations suggesting a significant electrostatic contribution to the protein–surfactant interaction. At low protein concentration, 0.08 mg/ml, some light scattering within a narrow CTAC concentration range occurs due to protein–surfactant precipitation. Light scattering experiments showed the dissociation of the oligomeric structure by SDS and CTAC, and the effect of precipitation induced by CTAC. At higher protein concentrations, 3.0 mg/ml, a precipitation was observed due to the intense charge neutralization upon formation of ion pair in the protein–surfactant precipitate. The spectral changes are spread over a much wider SDS concentration range, implying a smaller electrostatic contribution to the protein–surfactant interactions. The observed effects are consistent with the acid isoelectric point (pI) of this class of hemoglobins, which favors the intense interaction of HbGp with the cationic surfactant due to the existence of excess acid anionic residues at the protein surface. Protein secondary structure changes are significant for CTAC at low concentrations while they occur at significantly higher concentrations for SDS. In summary, the cationic surfactant seems to interact more strongly with the protein producing more dramatic spectral changes as compared to the anionic one. This is opposite as observed for several other hemoproteins. The surfactants at low concentrations produce the oligomeric dissociation, which facilitates the iron oxidation, an important factor modulating further oligomeric protein dissociation.  相似文献   

7.
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   

8.
The effects of two ionic surfactants on the oligomeric structure of the giant extracellular hemoglobin of Glossoscolex paulistus (HbGp) in the oxy - form have been studied through the use of several spectroscopic techniques such as electronic optical absorption, fluorescence emission, light scattering, and circular dichroism. The use of anionic sodium dodecyl sulphate (SDS) and cationic cethyltrimethyl ammonium chloride (CTAC) has allowed to differentiate the effects of opposite headgroup charges on the oligomeric structure dissociation and hemoglobin autoxidation. At pH 7.0, both surfactants induce the protein dissociation and a significant oxidation. Spectral changes occur at very low CTAC concentrations suggesting a significant electrostatic contribution to the protein-surfactant interaction. At low protein concentration, 0.08 mg/ml, some light scattering within a narrow CTAC concentration range occurs due to protein-surfactant precipitation. Light scattering experiments showed the dissociation of the oligomeric structure by SDS and CTAC, and the effect of precipitation induced by CTAC. At higher protein concentrations, 3.0 mg/ml, a precipitation was observed due to the intense charge neutralization upon formation of ion pair in the protein-surfactant precipitate. The spectral changes are spread over a much wider SDS concentration range, implying a smaller electrostatic contribution to the protein-surfactant interactions. The observed effects are consistent with the acid isoelectric point (pI) of this class of hemoglobins, which favors the intense interaction of HbGp with the cationic surfactant due to the existence of excess acid anionic residues at the protein surface. Protein secondary structure changes are significant for CTAC at low concentrations while they occur at significantly higher concentrations for SDS. In summary, the cationic surfactant seems to interact more strongly with the protein producing more dramatic spectral changes as compared to the anionic one. This is opposite as observed for several other hemoproteins. The surfactants at low concentrations produce the oligomeric dissociation, which facilitates the iron oxidation, an important factor modulating further oligomeric protein dissociation.  相似文献   

9.
The suggestion that the high molecular weight erythrocyte membrane protein, spectrin, consists of subunits resistant to dissociation by both sodium dodecyl sulfate and 6 m guanidine hydrochloride has been reevaluated. By gel electrophoresis in dodecyl sulfate and thin-layer gel filtration in 6 m guanidine hydrochloride as well as in the much more powerful denaturant guanidine thiocyanate, and by sedimentation velocity in 6 m guanidine hydrochloride, the molecular weight emerges in the range 2–2.5 × 105. Denaturation profiles as a function of guanidine hydrochloride concentration, observed by circular dichroism, reveal that the spectrin conformation is unusually labile, with a mid-point for the unfolding process at a denaturant concentration near 1 m. Complete acylation with succinic anhydride, as well as reaction with citraconic anhydride, leaves the molecular weight unchanged even in 6 m guanidine hydrochloride. The possibility of measuring molecular weights of proteins by viscosity determination in trifluoroacetic acid was explored. A calibration with a series of proteins gave a Mark-Houwink plot with high scatter, which did not result from low precision of viscosity determination or protein degradation. Evidence is adduced from infrared spectra that the scatter is due to a variable degree of protonation of the polypeptide backbone in the acid, leading to altered hydrodynamic characteristics. Within the semiquantitive limits of the method, spectrin is not further disaggregated in trifluoroacetic acid. The presence of refractory noncovalent interactions and of covalent cross-links has been variously invoked to explain an apparent microheterogeneity in spectrin preparations. The results here described appear to render the former explanation untenable.  相似文献   

10.
The activity and conformational change of human placental cystatin (HPC), a low molecular weight thiol proteinase inhibitor (12,500) has been investigated in presence of guanidine hydrochloride (GdnHCl) and urea. The denaturation of HPC was followed by activity measurements, fluorescence spectroscopy and Circular Dichroism (CD) studies. Increasing the denaturant concentration significantly enhanced the inactivation and unfolding of HPC. The enzyme was 50% inactivated at 1.5 M GdnHCl or 3 M urea. Up to 1.5 M GdnHCl concentration there was quenching of fluorescence intensity compared to native form however at 2 M concentration intensity increased and emission maxima had 5 nm red shift with complete unfolding in 4–6 M range. The mid point of transition was in the region of 1.5–2 M. In case of urea denaturation, the fluorescence intensity increased gradually with increase in the concentration of denaturant. The protein unfolded completely in 6–8 M concentration of urea with a mid-point of transition at 3 M. CD spectroscopy shows that the ellipticity of HPC has increased compared to that of native up to 1.5 M GdnHCl and then there is gradual decrease in ellipticity from 2 to 5 M concentration. At 6 M GdnHCl the protein had random coil conformation. For urea the ellipticity decreases with increase in concentration showing a sigmoidal shaped transition curve with little change up to 1 M urea. The protein greatly loses its structure at 6 M urea and at 8 M it is a random coil. The urea induced denaturation follows two-state rule in which Native→Denatured state transition occurs in a single step whereas in case of GdnHCl, intermediates or non-native states are observed at lower concentrations of denaturant. These intermediate states are possibly due to stabilizing properties of guanidine cation (Gdn+) at lower concentrations, whereas at higher concentrations it acts as a classical denaturant.  相似文献   

11.
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   

12.
1. The fluorescence and circular dichroism of four homogeneous preparations of ATPase (adenosine triphosphatase) from Micrococcus lysodeikticus differing in molecular structure and enzymic properties were examined at pH 7.5 and 25 degrees. Emission was maximum at 325 and 335 nm and the relative intensities at these wavelengths may be used to characterize the different ATPase preparations. The circular-dichroism spectra exhibited negative extrema at 208 and 220 nm, and the relative value of the molar ellipticity at these wavelengths was also different for each molecular form of the enzyme. 2. The four preparations undergo two consecutive major unfolding transitions in guanidinium chloride (midpoints at 0.94 and 1.5 M denaturant), with concomitant destruction of the quaternary structure of the protein. A comparatively minor alteration in the ATPase structure also occurred in 0.05-0.2M-guanidine and led to complete inactivation of the enzyme. The inactivation and the first unfolding transition were reversible by dilution of the denaturant; the transition with midpoint at 1.5M-guanidine was irreversible. 3. Similar results were obtained in urea, except that the successive transitions had midpoints at concentrations of denaturant of 0.4, 2.0 and 4.5M. Low concentrations of urea caused a noticeable activation of the enzyme activity and alterations of the electrophoretic mobility of the ATPase. 4. A model is proposed in which one of the major subunits, alpha, is first dissociated and unfolded reversibly by the denaturants, followed by the irreversible unfolding and dissociation of the other major subunit, beta, from subunit delta and/or the components of relative mobility 1.0 in dodecyl sulphate/polyacrylamide-gel electrophoresis (rho).  相似文献   

13.
The mechanism for the solubilization of isolated central-nervous-system myelin by sodium dodecyl sulphate was studied in detail. The release of protein and phospholipid to the 100000 g x 1 h supernatant fraction is dependent on the total amount of detergent relative to the amount of membrane present and on the ionic strength of the solubilization system. Gel-filtration analysis of supernatant fractions indicate that at suboptimal concentrations of detergent these contain lipid-protein complexes. The complete dissociation of the individual protein components from lipid is dependent on the total amount of sodium dodecyl sulphate present in the system. The results indicate that for the analysis of membrane components in sodium dodecyl sulphate it is essential that sufficient detergent is present.  相似文献   

14.
The pH and denaturant stability of bovine milk galactosyl transferase was studied with particular reference both to aspects of published isolation procedures (acid casein precipitation) and to experiments probing the accessibility of reactive thiol groups. As monitored by catalytic activity or fluorescence spectroscopy, the enzyme undergoes an irreversible inactivation and concomitant structure change below pH 5.0, which is extremely rapid below pH 4.4. At catalytic pH (7.5) the enzyme inactivates (unfolds) at ~ 5 m urea, 2.0 m guanidine hydrochloride and 0.6 mm sodium dodecyl sulphate. The latter detergent apparently binds near Trp residues, as evidenced by a large (>10 nm) blue shift. Extreme caution should be taken in any (acid) casein precipitation steps.  相似文献   

15.
The effects of increasing concentrations of sodium deoxycholate on Semliki Forest have been studied. Sodium deoxycholate begins to bind to the virus at less than 0.1 mM free equilibrium concentration and causes lysis of the viral membrane at 0.9 +/- 0.1 mM free equilibrium concentration when 2.2 +/- 0.2 - 103 mol of sodium deoxycholate are bound per mol of virus. Liberation of proteins from the membrane begins at 1.5 +/- 0.1 mM sodium deoxycholate and the proteins released are virtually free from phospholipid above 2.0 mM sodium deoxycholate. The overall mechanism of sodium deoxycholate solubilization of the viral membrane resembles that of Triton X-100 and sodium dodecyl sulphate except that with sodium deoxycholate the various stages of membrane disruption occur at about 10-fold higher equilibrium free detergent concentrations. At sodium deoxycholate concentrations higher than 2.3 mM the viral spike glycoproteins can be separated by sucrose gradient centrifugation or gel filtration into constituent polypeptides E1, E2 and E3. E1 carries the haemagglutinating activity of the virus.  相似文献   

16.
The effect of urea, guanidine hydrochloride and sodium dodecyl sulphate on glycinin, the high molecular weight protein fraction from soybean has been investigated by analytical ultracentrifugation. Urea and guanidine hydrochloride dissociate the protein to a ‘2S’ protein through the intermediary 7S and 4S proteins. Howeαer, in sodium dodecyl sulphate the protein directly dissociates to a 2S protein. Analysis of the data by calculation of per cent fraction and S20,w value indicates that dissociation and denaturation of glycinin occur simultaneously in the presence of the aboαe reagents but to different extents.  相似文献   

17.
Three S100 protein species (S100a, S100b, S100a') have been purified from bovine brain using a modification of standard preparative methods. A higher yield for each protein was obtained at the last separation step. Characterization by urea/sodium dodecyl sulfate/polyacrylamide gel electrophoresis, UV absorption spectra, and fluorescence parameters provided evidence of a new tryptophan-containing S100 protein called S100a', which exhibits, as S100a and S100b, the properties of a Ca2+ binding protein.  相似文献   

18.
The extent of binding of sodium dodecyl sulphate to bovine serum albumin at high binding ratios was investigated by gel filtration. The weight ratio of bound sodium dodecyl sulphate to bovine serum albumin increases with the NaCl concentration, and, except at low salt concentrations, with the concentration of sodium dodecyl sulphate. In the presence of 1.0g of sodium dodecyl sulphate/l, the binding ratio varied from 1.0 (at 0.04m-Na(+)) to 2.2 (at 0.44m-Na(+)). In the presence of 0.24m-Na(+), the binding ratio increased with sodium dodecyl sulphate concentration, from 0.9 (0.2g of sodium dodecyl sulphate/l) to 2.0 (5g of sodium dodecyl sulphate/l), at 26 degrees C, in a dilute sodium phosphate buffer. No significant dependence of the binding ratio upon temperature in the range 26-45 degrees C was observed. These results differ from those of Reynolds & Tanford (1970a) obtained by equilibrium dialysis.  相似文献   

19.
Studies on the alpha-subunit of bovine brain S-100 protein.   总被引:3,自引:0,他引:3       下载免费PDF全文
A method is described for the rapid purification of both S-100 protein and calmodulin from crude bovine brain extracts by the use of a fluphenazine-Sepharose affinity column eluted stepwise with decreasing concentrations of free Ca2+. Protein containing only alpha-subunit was purified from preparations of S-100 protein by anion-exchange chromatography. This protein co-migrated with the alpha-subunit of S-100 protein on sodium dodecyl sulphate/urea/polyacrylamide-gel electrophoresis and had an amino acid composition identical with that previously reported for this subunit. The results of u.v.-absorption and fluorescence-emission spectroscopy indicate that the tryptophan residue of the purified alpha-subunit of S-100 protein undergoes a Ca2+-induced change in environment. Measurements of changes in tryptophan fluorescence with increasing Ca2+ concentrations suggest an apparent dissociation constant of the alpha-subunit for Ca2+ of 7 X 10(-5)M in the absence of K+. In the presence of 90mM-K+ this value is increased to 3.4 X 10(-4)M.  相似文献   

20.
During denaturation by sodium dodecyl sulfate (SDS), aminoacylase shows a rapid decrease in activity with increasing concentration of the detergent to reach complete inactivation at 1.0 mM SDS. The denatured minus native-enzyme difference spectrum showed two negative peaks at 287 and 295 nm. With the increase of concentration of SDS, both negative peaks increased in magnitude to reach maximal values at 5.0 mM SDS. The fluorescence emission intensity of the enzyme decreased, whereas there was no red shift of emission maximum in SDS solutions of increasing concentration. In the SDS concentration regions employed in the present study, no marked changes of secondary structure of the enzyme have been observed by following the changes in far-ultraviolet CD spectra. The inactivation of this enzyme has been followed and compared with the unfolding observed during denaturation in SDS solutions. A marked inactivation is already evident at low SDS concentration before significant conformational changes can be detected by ultraviolet absorbance and fluorescence changes. The inactivation rate constants of free enzyme and substrate-enzyme complex were determined by the kinetics method of the substrate reaction in the presence of inactivator previously described by Tsou [Tsou (1988),Adv. Enzymol. Related Areas Mol. Biol. 61, 381–436]. It was found that substrate protects against inactivation and at the same SDS concentrations, the inactivation rate of the free enzyme is much higher than the unfolding rate. The above results show that the active sites of metal enzyme containing Zn2+ are also situated in a limited and flexible region of the enzyme molecule that is more fragile to denaturants than the protein as a whole.  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