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.
Manning M  Colón W 《Biochemistry》2004,43(35):11248-11254
The term kinetic stability is used to describe proteins that are trapped in a specific conformation because of an unusually high-unfolding barrier that results in very slow unfolding rates. Motivated by the observation that some proteins are resistant to sodium dodecyl sulfate (SDS)-induced denaturation, an attempt was made to determine whether this property is a result of kinetic stability. We studied many proteins, including a few kinetically stable proteins known to be resistant to SDS. The resistance to SDS-induced denaturation was investigated by comparing the migration on polyacrylamide gels of identical boiled and unboiled protein samples containing SDS. On the basis of the different migration of these samples, eight proteins emerged as being resistant to SDS. The kinetic stability of these proteins was confirmed by their slow unfolding rate upon incubation in guanidine hydrochloride. Further studies showed that these proteins were also extremely resistant to proteolysis by proteinase K, suggesting that a common mechanism may account for their resistance to SDS and proteolytic cleavage. Together, these observations suggest that a rigid protein structure may be the physical basis for kinetic stability and that resistance to SDS may serve as a simple assay for identifying proteins whose native conformations are kinetically trapped. Remarkably, most of the kinetically stable SDS-resistant proteins in this study are oligomeric beta-sheet proteins, suggesting a bias of these types of structures toward kinetic stability.  相似文献   

2.
Unlike most proteins, which are in equilibrium with partially and globally unfolded conformations, kinetically stable proteins (KSPs) are trapped in their native conformations and are often resistant to harsh environment. Based on a previous correlation between kinetic stability (KS) and a protein's resistance to sodium dodecyl sulfate (SDS), we show here a simple method to identify KSPs by SDS‐capillary electrophoresis (CE). Control non‐KSPs were fully denatured by SDS and formed protein:SDS complexes that exhibited similar mobility in CE. In contrast, KSPs bound fewer SDS molecules, and showed a very different migration time and peak pattern in CE, thereby providing some insight about the structural heterogeneity of SDS:protein complexes and the relative KS of the various proteins.  相似文献   

3.
Xia K  Zhang S  Bathrick B  Liu S  Garcia Y  Colón W 《Biochemistry》2012,51(1):100-107
Globular proteins are usually in equilibrium with unfolded conformations, whereas kinetically stable proteins (KSPs) are conformationally trapped by their high unfolding transition state energy. Kinetic stability (KS) could allow proteins to maintain their activity under harsh conditions, increase a protein's half-life, or protect against misfolding-aggregation. Here we show the development of a simple method for quantifying a protein's KS that involves incubating a protein in SDS at high temperature as a function of time, running the unheated samples on SDS-PAGE, and quantifying the bands to determine the time-dependent loss of a protein's SDS resistance. Six diverse proteins, including two monomer, two dimers, and two tetramers, were studied by this method, and the kinetics of the loss of SDS resistance correlated linearly with their unfolding rate determined by circular dichroism. These results imply that the mechanism by which SDS denatures proteins involves conformational trapping, with a trapping rate that is determined and limited by the rate of protein unfolding. We applied the SDS trapping of proteins (S-TraP) method to superoxide dismutase (SOD) and transthyretin (TTR), which are highly KSPs with native unfolding rates that are difficult to measure by conventional spectroscopic methods. A combination of S-TraP experiments between 75 and 90 °C combined with Eyring plot analysis yielded an unfolding half-life of 70 ± 37 and 18 ± 6 days at 37 °C for SOD and TTR, respectively. The S-TraP method shown here is extremely accessible, sample-efficient, cost-effective, compatible with impure or complex samples, and will be useful for exploring the biological and pathological roles of kinetic stability.  相似文献   

4.
Soluble amyloid beta-protein (Abeta) is normally present in the cerebrospinal fluid (CSF) and plasma. However, it is fibrillized and deposited as plaques in the brains of patients with Alzheimer's disease. Cerebrospinal fluid (CSF) contains several circulating proteins (apolipoprotein E, apolipoprotein J, and transthyretin) that bind to Abeta. We report here that gelsolin, a secretory protein, also binds to Abeta in a concentration-dependent manner. Under similar conditions, other proteins such as G-actin, protein kinase C, polyglutamic acid, and gelatin did not bind to Abeta. Solid phase binding assays showed two Abeta binding sites on gelsolin that have dissociation constants (Kd) of 1.38 and 2.55 microM. Abeta was found to co-immunoprecipitate along with gelsolin from the plasma, suggesting that gelsolin-Abeta complex exists under physiological conditions. The gelsolin-Abeta complex was sodium dodecyl sulfate (SDS)stable in the absence of reducing agent, but was dissociated when the SDS stop solution contained dithiothreitol (reducing agent). This study suggests that the function of secretory gelsolin in the CSF and plasma is to bind and sequester Abeta.  相似文献   

5.
The most effective protein purification method of low picomole amounts for sequence analysis involves polyacrylamide gel electrophoresis followed by electroblotting to polyvinylidene difluoride (PVDF) membranes. Since a critical factor in this procedure is the protein recovery at the blotting step, different types of PVDF membranes were systematically evaluated for their ability to bind proteins during electrotransfer. Differences in electroblotting recoveries occurred between types of PVDF membranes for some proteins. Some variability persisted even when optimized electroblotting procedures were used which reduce the sodium dodecyl sulfate (SDS) concentration in the gel and improve protein-PVDF binding. The membranes which were evaluated could be grouped as either "high retention" membranes (ProBlott, Trans-Blot, and Immobilon-PSQ) or "low retention" membranes (Immobilon-P and Westran). The high retention membranes showed higher protein recoveries under most conditions tested, especially for small proteins or peptides. These high retention membranes were also less sensitive to the exact electroblotting conditions, especially those factors which affect the amount of SDS present during either electrotransfer or direct adsorption from protein solutions. High retention PVDF membranes are therefore preferred in most cases for optimal protein or peptide recovery prior to direct sequence analysis. In contrast, low retention membranes are preferred for procedures where subsequent extraction of the proteins from the membranes is required. Even under identical conditions, substantial protein-to-protein variation for both adsorption and subsequent extraction is routinely observed for both groups of membranes, indicating that the nature of protein-PVDF interactions is more complex than simple hydrophobic interactions.  相似文献   

6.
Allostery, the modulation of function of a protein at one site by the binding of a ligand at a different site, is a property of many proteins. Two kinetically distinct models have been proposed: i) The induced fit model in which the ligand binds to the protein and then induces the conformational change. ii) The population selection model, in which the protein spontaneously undergoes a conformational change, which is then ‘captured’ by the ligand. Using measured kinetic constants for the lac repressor the contribution of population selection vs. induced dissociation is quantified by simulating the kinetics of allostery. At very low inducer concentration, both mechanisms contribute significantly. Total induction, though, is small under these conditions. At increasing levels of induction the induced dissociation mechanism soon dominates, first due to binding of one inducer, and then from two inducers binding.  相似文献   

7.
The possibility of occurrence of laminin binding proteins in cardiac tissue under different stages of growth was examined by affinity chromatography of the soluble fraction of human fetal myocardial plasma membrane over Ln-Sepharose. A 67 kDa protein was isolated by elution with glycine/HCl buffer containing 1 M NaCl and visualized as a coomassie stainable band on SDS gel electrophoresis under reducing conditions. Dot blot assays of the radioiodinated protein revealed the binding of 67 kDa protein with high affinity to laminin in a cation independent manner. This protein appears to be present in relatively higher amounts in tissues from early stage fetus. The occurrence of cation dependent laminin binding proteins was also examined by affinity chromatography. Electrophoresis of the EDTA eluate under reducing conditions followed by silver staining showed two prominent bands with average molecular size 130 and 174 kDa which under non-reducing conditions appeared as two bands with average molecular weight of 115 and 135 kDa. Using radioiodinated protein in dot blot assays, its binding to Ln was found to be maximum in the presence of Mn++ ions. Immunoblotting using anti-β1 integrin antibodies showed that 115 kDa protein is a β1 integrin suggesting the possibility of this protein belonging to the integrin group of receptors. The occurrence of multiple laminin binding proteins and the relative abundance of one of these proteins viz. the 67 kDa protein during early stages than in late stage tussue suggest a possible role for these proteins in cellular interactions with laminin during myocardial tissue development.  相似文献   

8.
Treating proteins with low concentrations of sodium dodecyl sulfate (SDS) and boiling for 2-3 min increased the linear range and total amount of protein that could be bound to nitrocellulose. Human serum albumin (HSA) and cathepsin G (Cat G) were both optimally bound at an SDS concentration of 10 micrograms/ml, while bronchial leukocyte proteinase inhibitor (BLPI) required 50 micrograms/ml SDS for optimum binding, corresponding to SDS-to-protein weight ratios of 0.5 and 2.5, respectively. Ionic strength and pH of the blotting buffers had a greater effect on the binding of SDS-treated proteins than on native proteins, with the linear binding range and total capacity for SDS-treated proteins being increased. Boiling SDS-treated human leukocyte extracts inactivated endogenous peroxidases, eliminating their interference with peroxidase-linked secondary antibodies in immunoassays. The nonionic detergents, Tween 20 and Nonidet P-40, were shown to rapidly wash both native and SDS-treated HSA off the filters, but these HSA samples were stable to washing with SDS. Although SDS-treated Cat G was more stable with nonionic detergents than was native Cat G, it was less resistant to washing with SDS. The substitution of SDS for nonionic detergents improved the response of immunoassays with native and SDS-treated proteins. Affinity-purified antibodies to human mast cell tryptase cross-reacted with native Cat G, but not with SDS-treated Cat G, indicating that SDS treatment can improve the specificity of immunoassays employing polyclonal antisera. These effects appear to be the result of partial denaturation and increases in the hydrophobicity of SDS-treated relative to native proteins.  相似文献   

9.
Different folding states of the small, globular milk protein bovine alpha-lactalbumin (BLA) induced by the anionic surfactant sodium dodecylsulphate (SDS) have been examined by fluorescence spectroscopy, CD and NMR. The solution structure of the protein in the absence of SDS was also determined, indicating fluidity even under native conditions. BLA is partly denatured to a molten globule (MG)-like state by micromolar concentrations of SDS, and the transitions from native to MG-like state are dependent on pH, the protein being more sensitive to the surfactant at pH 6.5. As indicated by measurements of the intrinsic emission fluorescence, the tertiary structure disappears at lower concentrations of SDS than most of the secondary structure, as estimated from CD data. The MG-like state induced by low concentrations of SDS is not observable by NMR, and is probably fluctuating and/or aggregating. At higher concentrations of SDS above the critic concentration of micelles, an NMR-observable state reappears. This micelle-associated conformer was partially assigned, and found to bear strong resemblance to the acid-tri-fluoroethanol state, retaining weakened versions of the A and C helix of native BLA. We discuss the results in terms of the inherent flexibility of the protein, and its ability to form multiple folding states and to bind to membranes. Also, we propose that proteins with stable MG-like conformers can have these states stabilized by low levels of compounds with surfactant properties in vivo.  相似文献   

10.
The chaperone calreticulin is a highly conserved eukaryotic protein mainly located in the endoplasmic reticulum. It contains a free cysteine SH group but does not form disulfide-bridged dimers under physiological conditions, indicating that the SH group may not be fully accessible in the native protein. Using PAGE, urea gradient gel electrophoresis, capillary electrophoresis and MS, we show that dimerization through the SH group can be induced by lowering the pH to 5-6, heating, or under conditions that favour partial unfolding such as urea concentrations above 2.6 m or SDS concentrations above 0.025%. Moreover, we show that calreticulin also has the ability to self-oligomerize through noncovalent interactions at urea concentrations above 2.6 m at pH below 4.6 or above pH 10, at temperatures above 40 degrees C, or in the presence of high concentrations of organic solvents (25%), conditions that favour partial unfolding or an intramolecular local conformational change that allows oligomerization, resulting in a heterogeneous mixture of oligomers consisting of up to 10 calreticulin monomers. The oligomeric calreticulin was very stable, but oligomerization was partially reversed by addition of 8 m urea or 1% SDS, and heat-induced oligomerization could be inhibited by 8 m urea or 1% SDS when present during heating. Comparison of the binding properties of monomeric and oligomeric calreticulin in solid-phase assays showed increased binding to peptides and denatured proteins when calreticulin was oligomerized. Thus, calreticulin shares the ability to self-oligomerize with other important chaperones such as GRP94 and HSP90, a property possibly associated with their chaperone activity.  相似文献   

11.
Affinity purification of Strep-tagged fusion proteins on resins carrying an engineered streptavidin (Strep-Tactin) has become a widely used method for isolation of protein complexes under physiological conditions. Fusion proteins containing two copies of Strep-tag II, designated twin-Strep-tag or SIII-tag, have the advantage of higher affinity for Strep-Tactin compared to those containing only a single Strep-tag, thus allowing more efficient protein purification. However, this advantage is offset by the fact that elution of twin-Strep-tagged proteins with biotin may be incomplete, leading to low protein recovery. The recovery can be dramatically improved by using denaturing elution with sodium dodecyl sulfate (SDS), but this leads to sample contamination with Strep-Tactin released from the resin, making the assay incompatible with downstream proteomic analysis. To overcome this limitation, we have developed a method whereby resin-coupled tetramer of Strep-Tactin is first stabilized by covalent cross-linking with Bis(sulfosuccinimidyl) suberate (BS3) and the resulting cross-linked resin is then used to purify target protein complexes in a single batch purification step. Efficient elution with SDS ensures good protein recovery, while the absence of contaminating Strep-Tactin allows downstream protein analysis by mass spectrometry. As a proof of concept, we describe here a protocol for purification of SIII-tagged viral protein VPg-Pro from nuclei of virus-infected N. benthamiana plants using the Strep-Tactin polymethacrylate resin cross-linked with BS3. The same protocol can be used to purify any twin-Strep-tagged protein of interest and characterize its physiological binding partners.  相似文献   

12.
Despite detailed knowledge of the overall structural changes and stoichiometries of surfactant binding, little is known about which protein regions constitute the preferred sites of attack for initial unfolding. Here we have exposed three proteins to limited proteolysis at anionic (SDS) and cationic (DTAC) surfactant concentrations corresponding to specific conformational transitions, using the surfactant‐robust broad‐specificity proteases Savinase and Alcalase. Cleavage sites are identified by SDS‐PAGE and N‐terminal sequencing. We observe well‐defined cleavage fragments, which suggest that flexibility is limited to certain regions of the protein. Cleavage sites for α‐lactalbumin and myoglobin correspond to regions identified in other studies as partially unfolded at low pH or in the presence of organic solvents. For Tnfn3, which does not form partially folded structures under other conditions, cleavage sites can be rationalized from the structure of the protein's folding transition state and the position of loops in the native state. Nevertheless, they are more sensitive to choice of surfactant and protease, probably reflecting a heterogeneous and fluctuating ensemble of partially unfolded structures. Thus, for proteins accumulating stable intermediates on the folding pathway, surfactants encourage the formation of these states, while the situation is more complex for proteins that do not form these intermediates. © 2008 Wiley Periodicals, Inc. Biopolymers 91: 221–231, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

13.
When sodium dodecyl sulfate (SDS) is added to a high-performance gel chromatographic column equilibrated with a buffer solution containing SDS at a level above the critical micelle concentration, the surplus SDS migrates as micelles giving a sharp peak. The presence of an unfolded protein in the sample solution gives a polypeptide peak in advance of the SDS micelle peak. As the result of SDS binding to the polypeptide, the SDS micelle peak is attenuated in comparison to that in the absence of protein. Thus the amount of SDS bound to the polypeptide can be determined accurately and simply from the decrease in the area of the SDS micelle peak. This approach is particularly useful for precise determination of bound SDS, which is pertinent to understanding the state of the protein polypeptide-SDS complex under the conditions of SDS-polyacrylamide gel electrophoresis.  相似文献   

14.
15.
Previously, a 2D gel electrophoresis approach was developed for the Escherichia coli inner membrane, which detects membrane protein complexes that are stable in sodium dodecyl sulfate (SDS) at room temperature, and dissociate under the influence of trifluoroethanol [R. E. Spelbrink et al., J. Biol. Chem. 280 (2005), 28742-8]. Here, the method was applied to the evolutionarily related mitochondrial inner membrane that was isolated from the yeast Saccharomyces cerevisiae. Surprisingly, only very few proteins were found to be dissociated by trifluoroethanol of which Lpd1p, a component of multiple protein complexes localized in the mitochondrial matrix, is the most prominent. Usage of either milder or more stringent conditions did not yield any additional proteins that were released by fluorinated alcohols. This strongly suggests that membrane protein complexes in yeast are less stable in SDS solution than their E. coli counterparts, which might be due to the overall reduced hydrophobicity of mitochondrial transmembrane proteins.  相似文献   

16.
Proteins of fresh, cold and dark-stored and illuminated tomato leaves were fractionated by SDS electrophoresis. The total soluble proteins extracted from fresh leaves were separated into 5 main fractions with MWs of 54,000, 45,000, 32,000, 23,000 and 14,000. The cold and dark storage of the leaves causes a marked reduction mainly in the fraction with MW of 45,000 which increased with the illumination of the cold and dark-storaged leaves. The polypeptides with MWs of 54,000 and 14,000 (probably large and small subunits of ribulose, bisphosphate carboxylase) were stable under these conditions. In contrast, the polypeptides with MWs of 54,000 and 14,000 are decreased following the storage of tomato leaves in the dark at room temperature. Chloroplast soluble proteins were seperated by SDS electrophoresis into fractions with MWs of 64,000, 54,000, 20,000 and 14,000. The same fractions in similar proportions were observed in soluble-chloroplast proteins from fresh as well as coold and dark-stored and illuminated leaves. No drastic changes in structural polypeptides were observed following cold and dark-storage and illumination of the leaves. The results indicated that the main protein fraction, which degradated following cold and dark storage of tomato leaves and synthetized during illumination, is the fraction of cytoplasmic protein which in SDS electrophoresis gives polypeptides of about 45,000 MW. The fractions of chloroplast proteins were stable under such conditions.Abbreviations DCIP 2,6-dichlorophenolindophenol - FFA free fatty acid - MW molecular weight - RuBP carboxylase ribulose 1,5-bisphosphate carboxylase - SDS sodium dodecyl sulphate - TCA trichloroacetic acid  相似文献   

17.
Z-DNA-binding proteins from bull testis.   总被引:3,自引:1,他引:2       下载免费PDF全文
Three Z-DNA-binding proteins of Mr 31, 33 and 58 kD were isolated from mature bull testis. They were obtained in a native state suitable for binding studies. These are the first examples of Z-DNA-binding proteins from a mammalian tissue. Purification involved tissue extraction with 0.35 M NaCl, cation exchange chromatography on CM-Trisacryl M and two consecutive anion exchange FPLC runs on Mono Q. The proteins appeared virtually homogeneous by anion exchange FPLC, SDS polyacrylamide gel electrophoresis and reverse phase HPLC (58 kD protein only). Yields from 50 g of testis tissue were: 31 kD protein, 40 micrograms; 33 kD protein, 100 micrograms; and 58 kD protein, 150 micrograms. Z-DNA binding was determined by Scatchard analysis of filter binding data using brominated poly(dG-dC).poly(dG-dC) as a conformation-specific ligand. Dissociation constants (Kz, in mol nucleotide/liter) were: 31 kD protein, 7 X 10(-7) M; 33 kD protein, 8 X 10(-7) M; 58 kD protein, 6 X 10(-8) M (primary binding site) and 6 X 10(-7) M (secondary binding site). B-DNA binding to poly(dG-dC).poly(dG-dG) was too low for reliable determination under the conditions of assay. This attested to a high degree of conformational specificity of the three proteins. The 58 kD protein bound Z-DNA at the primary site with an affinity almost equivalent to that of a polyclonal anti-Z-DNA antiserum raised in a rabbit (Kz, 4 X 10(-8) M).  相似文献   

18.
A rapid and simple method has been developed which allows detection and isolation of covalent DNA/protein adducts. The method is based upon the use of an ionic detergent, SDS, to neutralize cationic sites of weakly bound proteins thereby resulting in their dissociation off the helix. Proteins tightly or covalently bound to DNA that are not dissociable by SDS, result in the precipitation of the DNA fragment by the addition of KCl; however, free nucleic acid does not precipitate. The method is particularly useful as an analytical tool to titrate the binding of prototypic covalent binding proteins, topoisomerase I and II; thus, quantitation of topoisomerase activity is possible under defined conditions. As an analytical tool the method can be used as a general assay in the purification of as yet unidentified topoisomerases or other activities that bind DNA covalently. Moreover, the technology can be adapted for use in a preparative mode to separate covalent complexes from free DNA in a single step.  相似文献   

19.
In the microsome of scallop adductor striated muscle, 30K, 55K, 90K, and 360K proteins were detected as calcium binding proteins by 45Ca autoradiography on the transferred nitrocellulose membrane after sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS PAGE). The 360K protein was directly extracted with Triton X-100 from the whole homogenate of striated portion of scallop adductor muscle and purified through DEAE cellulose and hydroxyapatite column chromatography. This purified scallop high molecular weight calcium binding protein (SHCBP) showed a faster mobility in SDS PAGE in the presence of Ca2+ than in its absence. The decrease of tryptophan fluorescence had a half maximum near pCa 7 and was slightly co-operative with Mg2+. UV absorbance was slightly increased with Ca2+. The CD spectrum also changed with Mg2+ and Ca2+. These results reflect that this SHCBP binds calcium ions under near physiological conditions. SHCBP-like high molecular weight calcium binding proteins were also detected in the smooth muscle portion of adductor muscle and branchiae of scallop by 45Ca autoradiography, but not in liver. The adductor muscle of clam had a high molecular weight calcium binding protein whose molecular weight was a little smaller than that of SHCBP. The foot of turban shell had the same molecular weight calcium binding protein as SHCBP. Stains-all, a cationic carbocyanine dye, which has been reported to stain calcium binding proteins blue, stained SHCBP blue. The spectrum of SHCBP stained with Stains-all was very similar to that of calsequestrin. Although the function of SHCBP is still unknown, it might be expected to correspond to calsequestrin of vertebrate skeletal muscle, a calcium sequestering protein, in the sarcoplasmic reticulum.  相似文献   

20.
We describe a cyclic on‐column procedure for the sequential degradation of complex O‐glycans on proteins or peptides by periodate oxidation of sugars and cleavage of oxidation products by elimination. Desialylated glycoproteins were immobilized to alkali‐stable, reversed‐phase Poros 20 beads followed by two degradation cycles and the eluted apoproteins were either separated by SDS gel electrophoresis or digested with trypsin prior to LC/ESI‐MS. We demonstrate on the peptide and protein level that even complex glycan moieties are removed under mild conditions with only minimal effects on structural integrity of the peptide core by fragmentation, dehydration or by racemization of the Lys/Arg residues. The protocol is applicable on gel‐immobilized glycoproteins after SDS gel electrophoresis. Conversion of O‐glycoproteins into their corresponding apoproteins should result in facilitated accessibility of tryptic cleavage sites, increase the numbers of peptide fragments, and accordingly enhance protein coverage and identification rates within the subproteome of mucin‐type O‐glycoproteins.  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