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.
The denaturation behavior of phaseolin in urea, guanidine hydrochloride, and sodium dodecyl sulfate solutions was examined by monitoring changes in the intrinsic fluorescence of tryptophan and tyrosyl residues. Changes in various fluorescence parameters, such as quantum yield, emission maximum, spectral half-width, fluorescence depolarization, and fluorescence quenching by acrylamide, have indicated that while phaseolin is relatively stable up to 8 M urea, it is completely destabilized in 6 M guanidine hydrochloride and 6 mM sodium dodecyl sulfate. Furthermore, while the denaturation of phaseolin in urea solutions followed a two-step process, that in guanidine hydrochloride and sodium dodecyl sulfate followed a single-step process. While the accessibility of tryptophan residues to the nonionic acrylamide quencher is almost 100% in 6 M guanidine hydrochloride and 6 mM sodium dodecyl sulfate, only about 72% was accessible in 8 M urea compared to 52% in native phaseolin. The results presented here suggest that the protomeric structure of phaseolin is quite stable to changes in the environment. This structural stability may be partly responsible for its resistance to proteolysis by various proteinases.  相似文献   

2.
Ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) is a key enzyme in photosynthesis and photorespiration. The inactivation and subsequent conformational changes and dissociation of rice Rubisco by SDS have been studied. At low SDS concentrations (0.4 mM), Rubisco completely lost its carboxylase activity and most of its sulfhydryl groups became exposed. Dissociation of small subunits and significant conformational changes occurred at higher SDS concentrations. Increasing SDS concentrations caused only slight changes in CD spectrum, indicating no significant effect of SDS on the secondary structure of the enzyme. The results prove that the active site of Rubisco is more fragile to denaturants than the protein as a whole. The results also suggest that small subunits are more liable to SDS denaturation and thus dissociate first, while the more hydrophobic large subunits remain complexed. The naturally existing hydrophobic surface of Rubisco may be an important factor in the interaction of Rubisco with other macromolecules.  相似文献   

3.
Although considerable information is available regarding protein–sodium dodecyl sulfate (SDS) interactions, it is still unclear as to how much SDS is needed to denature proteins. The role of protein charge and micellar surfactant concentration on amyloid fibrillation is also unclear. This study reports on equilibrium measurements of SDS interaction with six model proteins and analyzes the results to obtain a general understanding of conformational breakdown, reorganization and restructuring of secondary structure, and entry into the amyloid fibrillar state. Significantly, all of these responses are entirely resolved at much lower than the critical micellar concentration (CMC) of SDS. Electrostatic interaction of the dodecyl sulfate anion (DS) with positive surface potential on the protein can completely unfold both secondary and tertiary structures, which is followed by protein chain restructuration to α-helices. All SDS-denatured proteins contain more α-helices than the corresponding native state. SDS interaction stochastically drives proteins to the aggregated fibrillar state.  相似文献   

4.
5.
Sodium dodecyl sulfate (SDS) is used to denature and solubilize proteins, especially membrane and other hydrophobic proteins. A quantitative method to determine the concentration of SDS using the dye Stains-All is known. However, this method lacks the accuracy and reproducibility necessary for use with protein solutions where SDS concentration is a critical factor, so we modified this method after examining multiple parameters (solvent, pH, buffers, and light exposure). The improved method is simple to implement, robust, accurate, and (most important) precise.  相似文献   

6.
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   

7.
Annelid erythrocruorins are respiratory proteins with high cooperativity and low autoxidation rates. The giant extracellular hemoglobin of the earthworm, Glossoscolex paulistus (HbGp), has a molecular mass of 3.6 MDa. In this work, isothermal titration calorimetry (ITC), together with DLS and fluorescence emission have been used to investigate the interaction of SDS with the HbGp in the oxy‐form, at pH 7.0. Our ITC and DLS results show that addition of SDS induces oxy‐HbGp oligomeric dissociation, while a small amount of protein aggregation is observed only by DLS. Moreover, the oligomeric dissociation process is favored at lower protein concentrations. The temperature effect does not influence significantly the interaction of SDS with the hemoglobin, due to the similarities presented by the critical aggregation concentration (cac) and critical micelle concentration (cmc′) for the mixtures. The increase of oxy‐HbGp concentration leads to a slight variation of the cac values for the SDS‐oxy‐HbGp mixture, attributed mainly to the noncooperative electrostatic binding of surfactant to protein. However, the cmc′ values increase considerably, associated to a more cooperative hydrophobic binding. Complementary pyrene fluorescence emission studies show formation of pre‐micellar structures of the mixture already at lower SDS concentrations. This study opens the possibility of the evaluation of the surfactant effect on the hemoglobin stability by ITC, which is made for the first time with this extracellular hemoglobin. © 2014 Wiley Periodicals, Inc. Biopolymers 101: 1065–1076, 2014.  相似文献   

8.
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   

9.
10.
High enzyme activity of mitochondrial creatine kinase (creatine-N-phosphotransferase, mCK, EC 2.7.3.2) was detected in serum from a patient with advanced carcinoma of the rectum and its isoforms were characterized by means of isoelectric focusing (IEF). Three forms of mCK, membrane-bound (pI 6.9–7.0), octameric (pI 7.0–7.9) and dimeric (pI 6.7, 6.8, 6.9 and 7.0), were detected in the fresh serum. These three forms of mCK were converted to five dimeric isoforms, and these were characterized as one reduced form (pI 7.0) and four oxidized (pI 6.6, 6.7, 6.8 and 6.9) forms upon treatment with urea, hydrogen peroxide or 2-mercaptoethanol (2-ME). The C-terminal of the mCKs was concluded to be a lysine residue because the mCKs treated with carboxypeptidase B migrated to positions closer to the anode than did those not treated with carboxypeptidase B. Therefore, four bands were concluded to represent one reduced-delysined isoform (pI 6.4) and three oxidized-delysined isoforms (pI 6.1, 6.2 and 6.3). The broad octameric mCK band disappeared and a narrow band focused at pI 6.8–6.9 appeared upon probable delysination of the mCKs. Thus, the number of lysine residues at the C-terminal of the octamer was concluded to be variable due to variable catalysis by carboxypeptidase N in the plasma. mCKs seemed to be inactivated during conversion from a membrane-bound form to dimeric oxidized-delysined forms via the octameric, dimeric reduced and oxidized forms.  相似文献   

11.
Fatty acid synthase is an important enzyme participating in energy metabolism in vivo. The inactivation and conformational changes of the multifunctional fatty acid synthase from chicken liver in SDS solutions have been studied. The results show that the denaturation of this multifunctional enzyme by SDS occurred in three stages. At low concentrations of SDS (less than 0.15 mM) the enzyme was completely inactivated with regard to the overall reaction. For each component of the enzyme, the loss of activity occurred at higher concentrations of SDS. Significant conformational changes (as indicated by the changes of the intrinsic fluorescence emission and the ultraviolet difference spectra) occurred at higher concentrations of SDS. Increasing the SDS concentration caused only slight changes of the CD spectra, indicating that SDS had no significant effect on the secondary structure of the enzyme. The results suggest that the active sites of the multifunctional fatty acid synthase display more conformational flexibility than the enzyme molecule as a whole.  相似文献   

12.
Inactivation of prions by acidic sodium dodecyl sulfate   总被引:4,自引:0,他引:4       下载免费PDF全文
Prompted by the discovery that prions become protease-sensitive after exposure to branched polyamine dendrimers in acetic acid (AcOH) (S. Supattapone, H. Wille, L. Uyechi, J. Safar, P. Tremblay, F. C. Szoka, F. E. Cohen, S. B. Prusiner, and M. R. Scott, J. Virol. 75:3453-3461, 2001), we investigated the inactivation of prions by sodium dodecyl sulfate (SDS) in weak acid. As judged by sensitivity to proteolytic digestion, the disease-causing prion protein (PrPSc) was denatured at room temperature by SDS at pH values of < or =4.5 or > or =10. Exposure of Sc237 prions in Syrian hamster brain homogenates to 1% SDS and 0.5% AcOH at room temperature resulted in a reduction of prion titer by a factor of ca. 10(7); however, all of the bioassay hamsters eventually developed prion disease. When various concentrations of SDS and AcOH were tested, the duration and temperature of exposure acted synergistically to inactivate both hamster Sc237 prions and human sporadic Creutzfeldt-Jakob disease (sCJD) prions. The inactivation of prions in brain homogenates and those bound to stainless steel wires was evaluated by using bioassays in transgenic mice. sCJD prions were more than 100,000 times more resistant to inactivation than Sc237 prions, demonstrating that inactivation procedures validated on rodent prions cannot be extrapolated to inactivation of human prions. Some procedures that significantly reduced prion titers in brain homogenates had a limited effect on prions bound to the surface of stainless steel wires. Using acidic SDS combined with autoclaving for 15 min, human sCJD prions bound to stainless steel wires were eliminated. Our findings form the basis for a noncorrosive system that is suitable for inactivating prions on surgical instruments, as well as on other medical and dental equipment.  相似文献   

13.
Reversible denaturation of enzymes by sodium dodecyl sulfate   总被引:58,自引:0,他引:58  
  相似文献   

14.
The precise diagnosis for bovine spongiform encephalopathy (BSE) is crucial for preventing new transmission to humans. Several testing procedures are reported for determining protease-resistant prion protein in various tissues as a major hallmark of prion diseases such as BSE, scrapie, and Creutzfeldt-Jakob disease. However, contamination of materials from tissues or degradation of the specimens sometimes disturbs the accuracy of the assay. Here, we have developed a novel method for solid-phase immunoassay of the disease-specific conformational isoform, PrP(Sc), using filtration blotting of protein in the presence of sodium dodecyl sulfate (SDS) followed by a filtration-based immunoassay with a single anti-prion protein antibody, together with the improved fractionation procedure involving high concentrations of surfactant/detergent. The SDS/heat treatment renders unfolded PrP(Sc) quantitative retention on a polyvinylidene difluoride filter and allows enhancement of the analyte signal with immunodetection; thus, all of the tested specimens are determined with 100% accuracy. In addition, the immunoassay is completed in approximately 1h, indicating its usefulness not only for the screening of BSE specimens but probably also for the postmortem BSE diagnosis of fallen stock as the antibody recognizes the core part of PrP(Sc). The solid-phase immunoassay method, including the filtration blotting with SDS, would be applicable to determining even more sensitively proteins other than PrP(Sc), especially those having rigid conformations.  相似文献   

15.
The extent of helical structure of 19 intact proteins and of 15 proteins with no disulfide bridges in the absence and presence of 10 mM sodium dodecyl sulfate (SDS) was determined using the curve-fitting method of circular dichroic spectra. The change in helicity caused by the addition of SDS was examined as a function of each amino acid fraction. An increase in the helicity upon the addition of SDS occurred in most of the proteins with no disulfide bridges (C proteins) and containing more than 0.06 Lys fraction. In most of the intact proteins (B proteins), most of which contained disulfide bridges, helicity in SDS decreased with an increase in Lys fraction. The helicity of the C proteins in SDS also tended to increase with an increase in the Leu and Phe fractions, while it decreased with an increase in the Gly fraction. For the helicity of the B proteins in SDS, there was a tendency to increase with increased Asn fraction and decrease with increased His fraction. On the other hand, amino acids were divided into eight groups according to their side-chain properties and the conformational preference for each of the amino acid groups of C proteins was calculated using a simple assumption.  相似文献   

16.
In addition to our previously reported versatile methods for sodium dodecyl sulfate (SDS)-polyacrylamide gel electrophoresis [1] and isoelectric focusing [IEF]-gel [2], I have achieved molecular weight gradient flattening of the SDS-polyacrylamide gel and pH gradient flattening of the IEF gel at any segment using the same electrophoresis system. Any crowded gel segment where congregated components are not separated well can easily be widened for good separation and any dispersed gel segment where components are too far can easily be narrowed. Therefore, every gel segment can be used effectively and meaningfully because the gradient curve can be ajusted to any distribution of the components. In the crowded area, any small spots of components which could not be detected previously because of nearby heavy staining or strong radioactivity of an abundant component can be sufficiently separated from the nearby spots in a small gel without sacrificing other areas.  相似文献   

17.
Creatine kinase (ATP:creatine N-phosphotransferase, EC 2.7.3.2) is a good model for studying dissociation and reassociation during unfolding and refolding. This study compares self-reassociated CK dimers and CK dimers that contain hybrid dimers under proper conditions. Creatine kinase forms a monomer when denatured in 6 M urea for 1 h which will very quickly form a dimer when the denaturant is diluted under suitable conditions. After modification by DTNB, CK was denatured in 6 M urea to form a modified CK monomer. Dimerization of this modified subunit of CK occurred upon dilution into a suitable buffer containing DTT. Therefore, three different types of reassociated CK dimers including a hybrid dimer can be made from two different CK monomers in the proper conditions. The CK monomers are from a urea-denatured monomer of DTNB-modified CK and from an unmodified urea dissociated monomer. Equal enzyme concentration ratios of these two monomers were mixed in the presence of urea, then diluted into the proper buffer to form the three types of reassociated CK dimers including the hybrid dimer. Reassociated CK dimers including all three different types recover about 75% activity following a two-phase course (k 1 = 4.88 × 10–3 s–1, k 2 = 0.68 × 10–3 s–1). Intrinsic fluorescence spectra of the three different CK monomers which were dissociated in 6 M urea, dissociated in 6 M urea after DTNB modification, and a mixture of the first two dissociated enzymes were studied in the presence of the denaturant urea. The three monomers had different fluorescence intensities and emission maxima. The intrinsic fluorescence maximum intensity changes of the reassociated CK dimers were also studied. The refolding processes also follow biphasic kinetics (k 1 = 3.28 × 10–3 s–1, k 2 = 0.11 × 10–3 s –1) after dilution in the proper solutions. Tsou's method [Tsou (1988), Adv. Enzymol. Rel. Areas Mol. Biol. 61, 381–436] was also used to measure the kinetic reactivation rate constants for the different three types of reassociated CK dimers, with different kinetic reactivation rate constants observed for each type. CK dissociation and reassociation schemes are suggested based on the results.  相似文献   

18.
Methods are presented for the complete removal of dodecyl sulfate from proteins. Themethods utilize the extraction of dodecyl sulfate anions as ion pairs with triethylammonium or tributylammonium cations into an organic solvent. The protein is insoluble in the organic solvent and is recovered as a precipitate. The methods are applicable to microgram as well as milligram amounts of protein. In all cases studied, the recovery of protein ranges from 70 to 100%. The recovered protein is suitable for N-terminal Edman degradation, tryptic peptide mapping, and amino acid analysis and can be renatured to regain enzymatic activity and antigenicity.  相似文献   

19.
The number of lithium dodecylsulfate (LDS) molecules tightly bound to creatine kinase has been found by isotachophoresis to be 87 at 25°C upon saturation of the enzyme dimer with LDS. The binding shows positive cooperativity by both the Hill and Scatchard plots. The enzyme is completely inactivated when its high-affinity sites are fully occupied with LDS. However, at partial LDS saturation the activity remaining is definitely higher than can be accounted for by the amount of free enzyme left, showing the presence of species of active molecules with the tight LDS sites only partly saturated. The presence of ATP leads to a decrease in detergent bound at the high-affinity sites with partial restoration of activity.  相似文献   

20.
Sodium dodecyl sulfate (SDS) bound to proteins in solution could be estimated by passing through Extracti-Gel that removes free SDS followed by specific interaction of the fluorophore Rhodamine B with protein-bound SDS. The resulting fluorescence intensity is compared with a calibration curve. Whereas globular proteins respond to binding of 1.4 mg SDS/mg protein under native conditions, “kinetically stable” proteins that are otherwise resistant to denaturation due to structural integrity show a low level of SDS binding. Analysis of the circular dichroism spectrum shows that in spite of the low level of SDS binding to kinetically stable proteins under nondenaturing conditions, the detergent generates considerable secondary structure in these proteins. Because the low level of SDS binding is a general feature of kinetically stable proteins, the protocol may fulfill one of the criteria to classify a protein as kinetically stable.  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