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.
Sodium dodecyl sulfate (SDS) is a highly effective and widely used protein denaturant. We show that certain amphipathic cosolvents such as 2-methyl-2,4-pentanediol (MPD) can protect proteins from SDS denaturation, and in several cases can refold proteins from the SDS-denatured state. This cosolvent effect is observed with integral membrane proteins and soluble proteins from either the α-helical or the β-sheet structural classes. The SDS/MPD system can be used to study processes involving native protein states, and we demonstrate the reversible thermal denaturation of the outer membrane protein PagP in an SDS/MPD buffer. MPD and related cosolvents can modulate the denaturing properties of SDS, and we describe a simple and effective method to recover refolded, active protein from the SDS-denatured state.  相似文献   

2.
A simple method for determination of proteins, initially solubilized in Tris buffer containing sodium dodecyl sulfate (SDS), mercaptoethanol, and sucrose, is described. This method is based on protein and Coomassie brilliant blue G-250 binding but it involves the removal of excess SDS by precipitation with 100 mm potassium phosphate buffer, pH 7.4 to 7.5, prior to protein determination. It has been established that the precipitation of excess SDS does not lead to the removal of the solubilized proteins. Therefore the method is applicable to both water-soluble and water-insoluble protein samples.  相似文献   

3.
We have investigated the utilization of a previously published dye-binding protein quantitation method for the analysis of proteins solubilized in SDS, mercaptoethanol-tris electrophoresis sample buffer. Although there is a direct relationship between the amount of protein applied and the absorbance of Coomassie blue G for BSA standard curves, complicated protein mixtures such as whole cell homogenates of tissue culture cells show more complicated functions. Apparently, components present in whole cell homogenates altered the dye-binding phenomenon to the extent that absolute values of the total amount of protein of such a mixture cannot be obtained by comparison to standard curve. However, over a relatively wide range of protein concentrations, the dye-binding method can provide a relative value for the amount of total protein present in a given complex protein mixture solubilized in SDS-mercaptoethanol electrophoresis buffer.  相似文献   

4.
Otzen DE 《Biophysical journal》2002,83(4):2219-2230
The 101-residue monomeric protein S6 unfolds in the anionic detergent sodium dodecyl sulfate (SDS) above the critical micelle concentration, with unfolding rates varying according to two different modes. Our group has proposed that spherical micelles lead to saturation kinetics in unfolding (mode 1), while cylindrical micelles prevalent at higher SDS concentrations induce a power-law dependent increase in the unfolding rate (mode 2). Here I investigate in more detail how micellar properties affect protein unfolding. High NaCl concentrations, which induce cylindrical micelles, favor mode 2. This is consistent with our model, though other effects such as electrostatic screening cannot be discounted. Furthermore, unfolding does not occur in mode 2 in the cationic detergent LTAB, which is unable to form cylindrical micelles. A strong retardation of unfolding occurs at higher LTAB concentrations, possibly due to the formation of dead-end protein-detergent complexes. A similar, albeit much weaker, effect is seen in SDS in the absence of salt. Chymotrypsin inhibitor 2 exhibits the same modes of unfolding in SDS as S6, indicating that this type of protein unfolding is not specific for S6. The unfolding process in mode 1 has an activation barrier similar in magnitude to that in water, while the activation barrier in mode 2 is strongly concentration-dependent. The strong pH-dependence of unfolding in SDS and LTAB suggests that the rate of unfolding in anionic detergent is modulated by repulsion between detergent headgroups and anionic side chains, while cationic side chains modulate unfolding rates in cationic detergents.  相似文献   

5.
A method which facilitates the rapid and quantitative electrophoretic transfer of proteins from gels not containing sodium dodecyl sulfate (SDS) to nitrocellulose membranes is described. The equilibration of non-SDS-polyacrylamide gel electrophoretic gels in a buffer containing SDS confers a net negative charge to the proteins present, presumably as a result of the formation of SDS-protein complexes. Proteins from gels equilibrated in the SDS buffer and then electroblotted in a Tris-glycine buffer at pH 8.3 are transferred with much greater efficiency than are proteins from untreated gels. The method has been shown to significantly enhance the electrophoretic transfer of polyoma viral proteins resolved in either acetic acid-urea or isoelectric-focusing gels to nitrocellulose membranes, and it is suggested that the method should have universal applicability to all gel electrophoresis systems currently employed. The proteins from isoelectric-focusing gels treated with SDS and transferred to nitrocellulose membranes were found to retain antigenicity to antisera prepared against either denatured or native viral proteins.  相似文献   

6.
Actin-binding proteins have been identified in human platelets with a gel-overlay technique that uses 125I-G-actin. Platelet proteins were separated on SDS polyacrylamide gels using the buffer system of Laemmli (1970, Nature [Lond.] 227:680-685). The proteins were fixed in the gels with methanol-acetic acid, the SDS was washed out, and the proteins were renatured. The gels were incubated with 125I-G-actin from rabbit skeletal muscle that was radiolabeled with 125I according to the method of Bolton and Hunter (1973, Biochem. J. 133:529-538) and has been shown to retain biological activity. After nonspecifically bound radioactivity was washed out, gels were dried and processed for autoradiography. The 125I-G-actin binds to several proteins in human platelets, platelet extracts, and the particulate fraction. Control experiments demonstrate that the 125I-G-actin can be displaced by use of increasing amounts of unlabeled actin, that the binding is stable to 0.6 M NaCl, and that preheating the 125I-G-actin to 90 degrees C for 3 min eliminates all binding. Prominent 125I-G-actin-binding activities were present at Mr 90,000 and 40,000. The binding to the 90,000 Mr protein appears to be at least partially Ca++ sensitive, whereas the binding to the 40,000 Mr protein does not. 125I-G-actin bound to proteins in the SDS gels can be fixed in situ and compared directly with the stained gel. This technique should prove generally useful in identification and purification of some actin-binding proteins from cells and tissues.  相似文献   

7.
Consistent results have not been obtained yet on the presence of antibody to the M protein of measles virus in the sera of patients with subacute sclerosing panencephalitis (SSPE). We performed a comparative study on various immunoprecipitation systems which appeared in the literature and found that the difference in the composition of the solubilizing buffer produced a large variety of results on the immunoprecipitation. [35S]Methionine-labeled Vero cells infected with the Edmonston strain of measles virus were solubilized by 10 different buffers and reacted with hyperimmune rabbit serum to whole virus, monospecific antisera to H, NP, and M proteins of the virus, normal adults' sera, and the sera from 16 SSPE patients. The immune complex was absorbed by protein A and both solubilization and precipitation rates were compared with each viral protein. Although viral proteins were solubilized by all buffers, the solubilization rate varied considerably. M protein was solubilized and was not coprecipitated nonspecifically with any of the other viral proteins. Purified protein A conjugated to Sepharose was preferable to Staphylococcus aureus for absorption of the immune complex since the latter absorbed both viral and host proteins nonspecifically. The precipitation rates of the viral proteins also varied according to the buffers. Better solubilization of the viral proteins seemed to reduce their rate of precipitation for which the presence of SDS may be responsible, and the presence of the protease inhibitors may also affect the results of immunoprecipitation. Detection of M protein in the immunoprecipitates was largely influenced by the kind of buffer used: some buffers could detect it clearly, but others could not defect it at all. Among the solubilizing buffers tested, Saleh's buffer (Virology 93: 369-376 (1979)),, which contains 0.5% DOC and 0.5% Triton X-100, was most reliable for detection of the anti-M antibody in the rabbit serum, because it showed a high solubilization and high precipitation rates of viral proteins without nonspecific absorption by protein A or coprecipitation of M proteins with any of the other proteins. Using this buffer, we could definitely detect M proteins in the immunoprecipitates from the sera of all six healthy adults and 15 out of 16 patients with SSPE. It was found, however, that the amount of M proteins in SSPE patients was lower than that in healthy adults and varied considerably.  相似文献   

8.
A new technique associating the detergent Sodium Dodecyl Sulphate (SDS) and an alcohol‐type co‐solvent has been set up, showing an unexpected efficiency to refold several types of soluble or membrane proteins. The present contribution deepens the fundamental knowledge on the phenomena underlying this process, considering the refolding of two model peptides featuring the main protein secondary structures: α‐helix and β‐sheet. Their refolding was monitored by fluorescence and circular dichroism, and it turns out that: (i) 100% recovery of the folded structure is observed for both peptides, (ii) the highest the SDS concentration, the more co‐solvent to be added to recover the peptides' native structures, (iii) a high alcohol concentration is required to alter the SDS denaturing properties, (iv) the co‐solvent performance relies on its specific lipophilic–hydrophilic balanced character, (v) the size of the micelle formed by the detergent does not enter the process critical parameters, and (vi) increasing the salt concentration up to 1 M NaCl has a beneficial impact on the process efficiency. These mechanistic aspects will help us to improve the method and extend its application. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
A simple method for detection of DNA-binding proteins is offered. These proteins can be revealed, following their electrophoretic separation in sodium dodecyl sulfate (SDS)-polyacrylamide gel containing labeled DNA, by washing the gel in buffer to remove SDS and to allow protein renaturation. Protein-free DNA is washed out, remaining in the DNA-binding proteins that restored their original characteristic. After autoradiography these proteins are seen as black bands (by one-dimensional gel electrophoresis) or spots (by two-dimensional gel electrophoresis) on a grey background. High sensitivity of the method is shown by using protein fractions of rat liver and a standard method.  相似文献   

10.
Although GroE chaperonins and osmolytes had been used separately as protein folding aids, combining these two methods provides a considerable advantage for folding proteins that cannot fold with either osmolytes or chaperonins alone. This technique rapidly identifies superior folding solution conditions for a broad array of proteins that are difficult or impossible to fold by other methods. While testing the broad applicability of this technique, we have discovered that osmolytes greatly simplify the chaperonin reaction by eliminating the requirement for the co-chaperonin GroES which is normally involved in encapsulating folding proteins within the GroEL–GroES cavity. Therefore, combinations of soluble or immobilized GroEL, osmolytes and ATP or even ADP are sufficient to refold the test proteins. The first step in the chaperonin/osmolyte process is to form a stable long-lived chaperonin–substrate protein complex in the absence of nucleotide. In the second step, different osmolyte solutions are added along with nucleotides, thus forming a ‘folding array’ to identify superior folding conditions. The stable chaperonin–substrate protein complex can be concentrated or immobilized prior to osmolyte addition. This procedure prevents-off pathway aggregation during folding/refolding reactions and more importantly allows one to refold proteins at concentrations (~mg/ml) that are substantially higher than the critical aggregation concentration for given protein. This technique can be used for successful refolding of proteins from purified inclusion bodies. Recently, other investigators have used our chaperonin/osmolyte method to demonstrate that a mutant protein that misfolds in human disease can be rescued by GroEL/osmolyte system. Soluble or immobilized GroEL can be easily removed from the released folded protein using simple separation techniques. The method allows for isolation of folded monomeric or oligomeric proteins in quantities sufficient for X-ray crystallography or NMR structural determinations.  相似文献   

11.
Two-dimensional gel electrophoresis with immobilized pH gradients in the first dimension, initially applied for the separation of soluble and total cellular proteins, has been extended to the analysis of membrane proteins. We show that the usual procedures lead to artifacts and irreproducible results due to aggregation and precipitation of proteins and protein-phospholipid complexes during isoelectric focusing (first dimension) and sodium dodecyl sulfate (SDS) gel electrophoresis (second dimension). Optimized solubilization procedures for hydrophobic membrane proteins are presented and the use of dilute samples is shown to be essential to overcome the major problems in isoelectric focusing. Increased volumes of samples dissolved in rehydration buffer are applied by direct rehydration of dry immobilized pH gradient (IPG) gels. Isoelectric focusing in 2% 3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonate (CHAPS) without urea gives good results as does 2% Nonidet-P40 with 8 M urea. Heat denaturation should be avoided. An optimized equilibration procedure for IPG gel strips in SDS sample buffer prior to separation in the second dimension was developed that minimizes loss of proteins and results in high-resolution two-dimensional electropherographic maps with a minimum of streaking. The gel strips are partially dehydrated at 40 degrees C and shortly reswollen in situ on the SDS slab gel in SDS-sample buffer containing agarose.  相似文献   

12.
The HPLC-type hydroxyapatite chromatography in the presence of sodium dodecyl sulfate (SDS) was assessed with special attention to the behavior of the surfactant. A significant amount of SDS was found to be adsorbed to the hydroxyapatite packed in the column from the starting buffer, 50 mM sodium phosphate buffer, pH 7.0, only when the buffer contained SDS in a concentration at or above its critical micelle concentration. When the phosphate buffer concentration was increased while the SDS concentration was kept at 1 mg/ml, the adsorbed surfactant was desorbed in advance of the release of proteins. Polypeptides derived from proteins could be successfully separated only when the column had been thoroughly equilibrated with the above-mentioned starting buffer solution. When a protein polypeptide complexed with SDS, which had been similarly equilibrated, was applied to the column, an amount of SDS corresponding to 75-90% (w/w) of the surfactant originally bound to the polypeptide was released upon its binding to the hydroxyapatite. On the other hand, porin, an Escherichia coli outer membrane protein, retaining its trimeric native structure in the presence of SDS, released a significantly smaller amount of SDS. When the membrane protein was denatured to give a single polypeptide, it behaved in a manner similar to that of the other protein polypeptides. The mechanism of binding of the protein polypeptides was discussed on the basis of these results. The native and denatured entities of porin could be efficiently separated as the result of the difference in their mode of interaction with the hydroxyapatite.  相似文献   

13.
Sodium dodecyl sulfate (SDS) preparations from different sources were investigated with regard to their effect on the separation of the proteins of foot and mouth disease virus (FMDV) types 01K, A2S and C-Obb as well as of Sindbis virus during gel electrophoresis in continuous and discontinuous buffer systems in the presence of 8 m urea. In the continuous system, the different SDS preparations did not markedly alter the separations of any of the virus proteins, but the distances between the FMDV protein bands were small. In the discontinuous system, the source of SDS and the kind of virus protein strongly affected the separations obtained. In this system, adequate resolutions were obtained only with A2S and C-Obb proteins in the presence of MCB or Pierce SDS preparations. An appreciable percentage of the carbon chains in these SDS preparations were found to be longer than C12. The separations of Sindbis virus proteins which occurred in the continuous buffer PAGE system using any of the SDS preparations were preferable to those which were obtained in the discontinuous system.  相似文献   

14.
Arginine suppresses the aggregation of proteins. However, little is known about its mechanism. Here we have used HsNDK (Halobacterium salinarum nucleoside diphosphate kinase) to examine the solvent property of arginine. After exposure to 2 M arginine, HsNDK was diluted to a low salt buffer, resulting in fully active protein. Since unfolded HsNDK cannot refold in such low salt buffer, the observed activity indicates that HsNDK was in the native state in 2 M arginine. Enzyme activity was also examined directly in the presence of arginine, showing that it was active in the presence of 1 M arginine and, to less extent, 2 M arginine. Arginine, however, could not support refolding of heat-denatured HsNDK. HsNDK was stable at 40 degrees C for 19 h incubation in the presence of 1M arginine.  相似文献   

15.
Kinetic measurement of protein folding is limited by the method used to trigger folding. Traditional methods, such as stopped flow, have a long mixing dead time and cannot be used to monitor fast folding processes. Here, we report a compound, 4‐(bromomethyl)‐6,7‐dimethoxycoumarin, that can be used as a “photolabile cage” to study the early stages of protein folding. The folding process of a protein, RD1, including kinetics, enthalpy, and volume change, was studied by the combined use of a phototriggered caging strategy and time‐resolved photoacoustic calorimetry. The cage caused unfolding of the photolabile protein, and then a pulse UV laser (~10?9 s) was used to break the cage, leaving the protein free to refold and allowing the resolving of two folding events on a nanosecond time scale. This strategy is especially good for monitoring fast folding proteins that cannot be studied by traditional methods. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

16.
A membrane protein insoluble in water was isolated by gel chromatography in the presence of 0.1% sodium dodecyl sulfate (SDS) from chromatophores of a photosynthetic bacterium, Rhodospirillum rubrum. This is one of the major membrane proteins of the chromatophore. The protein was found to bind about four grams of SDS per gram, a value which is more than twice the amount generally observed with protein polypeptides derived from water-soluble globular proteins. The electrophoretic behavior of the complex between the membrane protein and SDS is abnormal due to this high capacity for binding SDS. Estimation of the molecular weight of this protein by SDS-polyacrylamide gel electrophoresis was thus impossible. Such an anomaly in SDS binding is unlikely to be restricted to the particular membrane protein described in this paper. The possibility of such a deviation from standard behavior in the interaction with SDS should be taken into consideration in studies of other membrane proteins, since SDS is often used both in analytical and preparative procedures.  相似文献   

17.
Nucleotide excision repair (NER) is the major DNA repair process that removes diverse DNA lesions including UV-induced photoproducts. There are more than 20 proteins involved in NER. Among them, XPC is thought to be one of the first proteins to recognize DNA damage during global genomic repair (GGR), a sub-pathway of NER. In order to study the mechanism through which XPC participates in GGR, we investigated the possible modifications of XPC protein upon UV irradiation in mammalian cells. Western blot analysis of cell lysates from UV-irradiated normal human fibroblast, prepared by direct boiling in an SDS lysis buffer, showed several anti-XPC antibody-reactive bands with molecular weight higher than the original XPC protein. The reciprocal immunoprecipitation and siRNA transfection analysis demonstrated that XPC protein is modified by SUMO-1 and ubiquitin. By using several NER-deficient cell lines, we found that DDB2 and XPA are required for UV-induced XPC modifications. Interestingly, both the inactivation of ubiquitylation and the treatment of proteasome inhibitors quantitatively inhibited the UV-induced XPC modifications. Furthermore, XPC protein is degraded significantly following UV irradiation in XP-A cells in which sumoylation of XPC does not occur. Taken together, we conclude that XPC protein is modified by SUMO-1 and ubiquitin following UV irradiation and these modifications require the functions of DDB2 and XPA, as well as the ubiquitin–proteasome system. Our results also suggest that at least one function of UV-induced XPC sumoylation is related to the stabilization of XPC protein.  相似文献   

18.
A prerequisite for the purification of any protein to homogeneity is that the protein is not non-specifically associated with other proteins especially during the final stage(s) of the fractionation procedure. This requirement is not so often fulfilled when nonionic detergents (for instance Triton X-100) are used for solubilization of membrane proteins. The reason is that these detergents are not efficient enough to prevent the protein of interest from forming aggregates with other proteins upon contact with chromatographic or electrophoretic supporting media, which, due to their polymeric nature, have a tendency to induce aggregation of other polymers, for instance, hydrophobic proteins. The aggregation can be avoided if sodium dodecyl sulfate (SDS) is employed as detergent. We therefore suggest that membrane proteins should be purified by conventional methods in the presence of SDS and that the purified proteins, which are in a denatured state, are allowed to renature. There is good change to renature internal membrane proteins since they should not be so susceptible to denaturation by detergents as are water-soluble proteins because the natural milieu of the former proteins is lipids which in fact are detergents. In this paper we present a renaturation method based on the removal of SDS by addition of a large excess of G 3707, a nonionic detergent. By this technique we have renatured a 5'-nucleotidase from Acholeplasma laidlawii and a neuraminidase from influenza virus. The enzyme activities were higher (up to 6-fold) after the removal of SDS than prior to the addition of SDS.  相似文献   

19.
Amphipols are short amphipathic polymers designed to stabilize membrane proteins in aqueous solutions in the absence of detergent. Bacteriorhodopsin (BR), a light-driven proton pump, has been denatured, either by direct solubilization of the purple membrane in sodium dodecylsulfate (SDS) solution or by a procedure that involves delipidation with organic solvent followed by transfer to SDS, and renatured in amphipol A8-35. The effect of different renaturation procedures and of the presence or absence of lipids and the cofactor retinal have been investigated. The resulting samples have been characterized by absorbance spectroscopy, size-exclusion chromatography, thermostability measurements, and determination of photocycle kinetics. Transfer to A8-35 can be achieved by SDS precipitation, dilution, or dialysis, the first route resulting in the highest yield of refolding. Functional BR can be refolded whether in the presence or absence of lipids, higher yields being achieved in their presence. Retinal is not required for the protein to refold, but it stabilizes the refolded form and, thereby, improves folding yields. Lipids are not required for BR to perform its complete photocycle, but their presence speeds up the return to the ground state. Taken together, these data indicate that a membrane or membrane-mimetic environment is not required for correct decoding of the chemical information contained in the sequence of BR; functional folding is possible even in the highly foreign environment of lipid-free amphipols. BR interactions with lipids, however, contribute to an effective photocycle.  相似文献   

20.
We report an improved shotgun method for analyzing proteomic samples containing sodium dodecyl sulfate (SDS). This method is based on the use of strong-cation exchange (SCX) liquid chromatography (LC) for SDS removal that can be integrated with peptide separation as the first dimension of the two-dimensional LC tandem mass spectrometry workflow. To optimize the performance of SDS removal, various experimental conditions, including the concentrations of chemical reagents and salts in the sample, the SDS concentration, and the SCX mobile phase composition, were investigated. It was found that a peptide recovery rate of about 90% could be achieved while removing SDS efficiently. One key finding was that, by increasing the SDS concentration to a certain level (0.5%) in the digested peptide sample, the sample recovery rate could be increased. The peptide recovery rate of BSA digests was found to be 90.6 ± 1.0% (n = 3), and SDS in the SCX fractions collected was not detectable by pyrolysis GC-MS, i.e., below the detection limit of 0.00006% for the undesalted SCX fractions. The peptide recovery rates were found to be 90.9% ± 2.7 (n = 3) and 89.5% ± 0.5% (n = 3) for the digests of the membrane-protein-enriched fractions of E. coli cell lysates and the MCF-7 breast cancer cell line, respectively. Compared to the methods that use acid-labile surfactants, such as RapiGest and PPS, for the MCF-7 membrane fraction sample, the SDS method identified, on average (n = 3), more peptides (~5%) and proteins (~16%) than the RapiGest method, while the RapiGest method identified more peptides (~21%) and proteins (~7%) from the E. coli membrane fraction than the SDS method. In both cases, the two methods identified more peptides and proteins than the PPS method. Since SCX is widely used as the first dimension of 2D-LC MS/MS, integration of SDS removal with peptide separation in SCX does not add any extra steps to the sample handling process. We demonstrated the application of this method for 2D-LC MS/MS profiling of the MCF-7 membrane protein fraction and identified 6889 unique peptides, corresponding to 2258 unique proteins or protein groups from two replicate experiments with a false peptide discovery rate of ~0.8%, compared to 5172 unique peptides and 1847 unique proteins identified by the RapiGest method.  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