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 46 毫秒
1.
The factors determining the onset and extent of reconstructive denaturation of proteins were considered by comparing circular dichroism (CD) data of seven proteins and previously published findings. The effects of sodium dodecyl sulfate (SDS) on the conformation of the following proteins were tested: lysozyme, the mitogens fromPhytolacca americana (fractions Pa2 and Pa4), lectin fromWistaria floribunda, ovine lutropin, a Bence Jones protein, and histone H2B. While the helix content of lysozyme was raised by SDS slightly, in the Bence Jones protein andW. floribunda lectin it increased from near zero to about 25–30%. In histone H2B the helix content was raised by SDS even to about 48%. However, no clear indication of helix formation could be observed in the mitogens and lutropin, even at low pH or 2.0–2.5. The tertiary structure of the proteins was perturbed by SDS. It was concluded that the reorganization of secondary structure of the proteins was favored by the following factors: (1) presence of helicogenic amino acid sequences in the protein, (2) availability of positively charged sites of the basic amino acids for interactions with the dodecyl ion, (3) absence of a large surplus of negatively charged sites on the surface of protein, and (4) absence of extensive disulfide cross-linking within the macromolecule. Both hydrophobic and electrostatic interactions occur in reconstructive denaturation, and the newly formed helices are stabilized by hydrophobic shielding by the alkyl chains of the alkyl sulfate.  相似文献   

2.
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   

3.
The effect of sodium dodecyl sulfate (SDS) on the structure and activity of aprA-subtilisin, a secreted bacterial serine protease which is 85% homologous to subtilisin BPN', was examined. The addition of SDS resulted in the slow conversion of the subtilisin from the intact protein to the completely unfolded form of the enzyme. No intermediates between these two populations were detected. This conversion was accompanied by decreased activity, disruption of tertiary structure, a change in the mobility of the protein when subjected to SDS-polyacrylamide gel electrophoresis, and an increase in the apparent Stokes radius of the protein. After 2 h in 1% SDS at 20 degrees C, 25% of the subtilisin was still intact and active. The amount of protein existing in the unfolded form was increased by increasing the length of time in SDS, by increasing the concentration of SDS, and by increasing the temperature of the subtilisin-SDS solution. Analysis of the dependence of the rate of unfolding on SDS concentration indicated that one SDS micelle can destroy two protein molecules. The activation energy for the SDS-induced denaturation of aprA-subtilisin was 20 kcal mol-1, indicating that unfolding of the protein could be the rate-limiting step.  相似文献   

4.
Maestro B  Sanz JM 《FEBS letters》2007,581(3):375-381
We have investigated the stability of the choline-binding module C-LytA against sodium dodecyl sulphate (SDS)-induced unfolding at pH 7.0 and 20 degrees C. A major intermediate with an unfolded N-terminal region accumulates at around 0.75 mM SDS, whereas 2.0 mM SDS was sufficient for a complete unfolding. This might be the first report of a protein being extensively unfolded by submicellar concentrations of SDS, occurring through formation of detergent clusters on the protein surface. All transitions were reversible upon SDS complexation with beta-cyclodextrin, allowing the calculation of thermodynamic parameters. A model for the unfolding of C-LytA by SDS is presented and compared to a previous denaturation scheme by guanidine hydrochloride.  相似文献   

5.
The conformational transitions of calcium binding protein parvalbumin III from carp muscle were studied by scanning calorimetry, potentiometric titration and isothermal calorimetric titration. Changes of Gibbs energy, enthalpy and partial heat capacity were determined. The removal of calcium ions by EDTA is accompanied by 1) a heat absorption of 75 +/- 10 kJ per mole of the protein, 2) a decrease in the Gibbs energy of protein structure stabilisation of about 42 kJ mol-1 and 3) a decrease in thermostability by more than 50 K. The protonation of the acidic groups leads to a loss of calcium followed by denaturation, while the pH of the transition strongly depends on calcium activity. The enthalpy and heat capacity changes at denaturation are comparable with the values observed for other compact globular proteins.  相似文献   

6.
Rv3868 (EccA1) is an essential CbxX/CfqX-family ATPase of the Mycobacterium tuberculosis ESX-1 secretion system. Previously, we demonstrated that Rv3868 is composed of two domains; a regulatory N-terminal domain (NT-Rv3868) and an ATP binding C-terminal domain (CT-Rv3868). In the present report, chemical denaturation studies show that electrostatic interactions stabilize the Rv3868. Interestingly, Rv3868 has notable heat stability and retains about 50% of ATPase activity even at 60 °C. The C-terminal domain was found to be important for the heat stability as demonstrated by both enzymatic activity assays and thermal denaturation experiments. Furthermore a structure-sequence analysis based on the content of charged and aliphatic amino acids rationalizes the higher propensity of Rv3868 for thermophilic characteristics.  相似文献   

7.
An activity which releases free uracil from bacteriophage PBS1 DNA has been purified over 10,000 fold from extracts of Saccharomyces cerevisiae. The enzyme is active on both native and denatured PBS1 DNA and is active in the absence of divalent cation, and in the presence of 1 mM EDTA. The enzyme has a negative molecular weight of 27,800 as estimated by glycerol gradient centrifugation and gel filtration. Enzyme activity has been recovered after denaturation in SDS and electrophoresis in an SDS polyacrylamide gel. This analysis suggests that the enzyme consists of a single polypeptide chain of about 27,000 daltons. Normal levels of uracil-DNA glycosylase activity were found in partially purified extracts of the nitrous-acid sensitive rad18-2 mutant of yeast.  相似文献   

8.
Cell extracts from Pyrococcus furiosus were found to contain five proteases, two of which (S66 and S102) are resistant to sodium dodecyl sulfate (SDS) denaturation. Cell extracts incubated at 98 degrees C in the presence of 1% SDS for 24 h exhibited substantial cellular proteolysis such that only four proteins could be visualized by amido black-Coomassie brilliant blue staining of SDS-polyacrylamide gels. The SDS-treated extract retained 19% of the initial proteolytic activity as represented by two proteases, S66 (66 kilodaltons [kDa]) and S102 (102 kDa). Immunoblot analysis with guinea pig sera containing antibodies against protease S66 indicated that S66 is related neither to S102 nor to the other proteases. The results of this analysis also suggest that S66 might be the hydrolysis product of a 200-kDa precursor which does not have proteolytic activity. The 24-h SDS-treated extract showed unusually thermostable proteolytic activity; the measured half-life at 98 degrees C was found to be 33 h. Proteases S66 and S102 were also resistant to denaturation by 8 M urea, 80 mM dithiothreitol, and 5% beta-mercaptoethanol. Purified protease S66 was inhibited by phenylmethylsulfonyl fluoride and diisopropyl fluorophosphate but not by EDTA, ethylene glycol-bis(beta-aminoethyl ether)-N,N,N',N'-tetraacetic acid, or iodoacetic acid. These results indicate that S66 is a serine protease. Amino acid ester hydrolysis studies showed that protease S66 was hydrolytically active towards N-benzoyl-L-arginine ethyl ester.  相似文献   

9.
Cell extracts from Pyrococcus furiosus were found to contain five proteases, two of which (S66 and S102) are resistant to sodium dodecyl sulfate (SDS) denaturation. Cell extracts incubated at 98 degrees C in the presence of 1% SDS for 24 h exhibited substantial cellular proteolysis such that only four proteins could be visualized by amido black-Coomassie brilliant blue staining of SDS-polyacrylamide gels. The SDS-treated extract retained 19% of the initial proteolytic activity as represented by two proteases, S66 (66 kilodaltons [kDa]) and S102 (102 kDa). Immunoblot analysis with guinea pig sera containing antibodies against protease S66 indicated that S66 is related neither to S102 nor to the other proteases. The results of this analysis also suggest that S66 might be the hydrolysis product of a 200-kDa precursor which does not have proteolytic activity. The 24-h SDS-treated extract showed unusually thermostable proteolytic activity; the measured half-life at 98 degrees C was found to be 33 h. Proteases S66 and S102 were also resistant to denaturation by 8 M urea, 80 mM dithiothreitol, and 5% beta-mercaptoethanol. Purified protease S66 was inhibited by phenylmethylsulfonyl fluoride and diisopropyl fluorophosphate but not by EDTA, ethylene glycol-bis(beta-aminoethyl ether)-N,N,N',N'-tetraacetic acid, or iodoacetic acid. These results indicate that S66 is a serine protease. Amino acid ester hydrolysis studies showed that protease S66 was hydrolytically active towards N-benzoyl-L-arginine ethyl ester.  相似文献   

10.
The helicities in various fragments of bovine serum albumin (BSA) were examined in the thermal denaturation and in sodium docecyl sulfate (SDS) denaturation. The thermal denaturation was examined in a temperature range between 2 and 65°C. The helicity decreased with a rise of temperature and it recovered to some degree upon cooling temperature. A rather high reversibility was observed in the BSA fragments, which were located in the N-terminal of the parent protein and then contained the first large loop with no disulfide bridge. The high reversibility was available also for the helicity in the first large loop of the fragment, disulfide bridges of which were reduced. The fragments, which were smaller than one domain, became unstable in the SDS denaturation. The helicities of such fragments decreased in lower SDS concentrations compared with those of the intact BSA and the large fragments, which contained one or more domains. A resistance to the SDS denaturation appeared in the helices of every large loop even after the fragmentation. On the other hand, helicities of the fragments decreased to 20–25% upon the reduction of disulfide bridges. However, the helicities of these fragments increased to 35–40% in the SDS denaturation.  相似文献   

11.
The stabilities of trypsin and soybean trypsin inhibitor in sodium dodecylsulfate (SDS) were examined by SDS-polyacrylamide gel electrophoresis (PAGE). Both samples contained several bands, all of which migrated to positions corresponding to the appropriate molecular weight or less, even when the samples were unheated, suggesting that both the trypsin and trypsin inhibitor are susceptible to SDS-induced denaturation. When they were mixed together prior to addition of SDS-PAGE sample buffer (1% SDS), a new smearing band appeared which corresponded to a molecular weight of around 46,000, suggesting that these proteins form a stable complex in SDS. This was confirmed by electroblotting and sequence analysis, which indicated that this band contains both the trypsin and inhibitor sequences. At a fixed concentration of the inhibitor, increasing concentrations of the trypsin resulted in an increase in the intensity of the complex band. When the mixture was heated for 10 min in 1% SDS, the complex band disappeared in a temperature-dependent manner. The melting temperature determined under the experimental conditions used was about 35|MoC. Similar results were obtained with Bowman-Birk trypsin inhibitor, except that the complex with the above inhibitor had a higher melting temperature, around 41|MoC, suggesting that the Bowman-Birk inhibitor/trypsin complex is more stable than the soybean inhibitor/trypsin complex.  相似文献   

12.
Accumulation of aggregated amyloid-β protein (Aβ) is an important feature of Alzheimer's disease. There is significant interest in understanding the initial steps of Aβ aggregation due to the recent focus on soluble Aβ oligomers. In vitro studies of Aβ aggregation have been aided by the use of conformation-specific antibodies which recognize shape rather than sequence. One of these, OC antiserum, recognizes certain elements of fibrillar Aβ across a broad range of sizes. We have observed the presence of these fibrillar elements at very early stages of Aβ incubation. Using a dot blot assay, OC-reactivity was found in size exclusion chromatography (SEC)-purified Aβ(1–42) monomer fractions immediately after isolation (early-stage). The OC-reactivity was not initially observed in the same fractions for Aβ(1–40) or the aggregation-restricted Aβ(1–42) L34P but was detected within 1–2 weeks of incubation. Stability studies demonstrated that early-stage OC-positive Aβ(1–42) aggregates were resistant to 4 M urea or guanidine hydrochloride but sensitive to 1% sodium dodecyl sulfate (SDS). Interestingly, the sensitivity to SDS diminished over time upon incubation of the SEC-purified Aβ(1–42) solution at 4 °C. Within 6–8 days the OC-positive Aβ42 aggregates were resistant to SDS denaturation. The progression to, and development of, SDS resistance for Aβ(1–42) occurred prior to thioflavin T fluorescence. In contrast, Aβ(1–40) aggregates formed after 6 days of incubation were sensitive to both urea and SDS. These findings reveal information on some of the earliest events in Aβ aggregation and suggest that it may be possible to target early-stage aggregates before they develop significant stability.  相似文献   

13.
Native bovine liver catalase [EC 1.11.1.6] and catalase acetylated with N-acetylimidazole (AI) both combined with sodium dodecyl sulfate (SDS) to form catalase-SDS complexes. The differences between native and acetylated catalase bound to SDS were investigated as regards enzymatic activity, absorption spectra, ORD and CD, sedimentation velocity and fluorescence spectra. It was found that the binding of SDS with both catalases depended on incubation time and SDS concentration, and that the acetylation of catalase had some protective effect on the denaturation of the molecule by SDS, which may be ascribed to a reduction of ionic interaction between SDS and the protein on acetylation. The native catalase was found to split into three smaller components on incubation with 1% SDS for 96 hr, whereas the acetylated catalase split into two smaller components. These smaller components were isolated by gel filtration through Sephadex G-100. The isolated components has estimated molecular weights of 60,000, 30,000, aide. It seemed likely that the modification occurred stepwise. Approximately 26% of the carboxyl groups of fibrinogen was modified finally. The modified fibrinogen had no interaction with cationic detergent, and did not form any complex with the detergent. In dilute acid, fibrinogen was observed to show only a slight interaction with cationic detergent. It is probable that the exposed and ionized carboxyl groups are essential for the formation of a complex between fibrinogen and cationic detergent.  相似文献   

14.
15.
Triglyceride lipase from Thermomyces lanuginosus (TlL) has been reported to be resistant to denaturation by sodium dodecyl sulfate (SDS). We have found that at neutral pH, structural integrity is strongly dependent on ionic strength. In 10 mM phosphate buffer and SDS, the lipase exhibits a far-UV CD spectrum similar to other proteins denatured in this surfactant while the near-UV CD spectrum shows a complete loss of tertiary structure, observations supported by steady state fluorescence spectroscopy. However, when increasing the ionic strength by the addition of NaCl, the lipase was rendered resistant towards SDS denaturation, as observed by all techniques employed. The effect of salt on the critical micelle concentration (CMC) of SDS was observed to correlate with the effect on the degree of SDS-induced denaturation. This finding is compatible with the notion that the concentration of SDS monomers is a crucial factor for SDS–lipase interactions. The presented results are important for the understanding and improvement of protein stability in surfactant systems.  相似文献   

16.
Green fluorescent protein (GFP) is an unusually stable fluorescent protein that belongs to a family of related auto-fluorescent proteins (AFPs). These AFPs have been generated from jellyfish GFP by mutating the amino acids in the chromophore or its vicinity. Variants that emit light in the blue region (Blue Fluorescent Protein, BFP), red region, or yellow region are readily available and are widely used in diverse applications. Previously, we had used fluorescence spectroscopy to study the effect of pH on the denaturation of GFP with SDS, urea, and heat. Surprisingly, we found that SDS, urea or heat, did not have any significant effect on the fluorescence of GFP at pH 7.5 or 8.5, however, at pH 6.5, the protein lost all fluorescence within a very short period of time. These results suggested that GFP undergoes a structural/stability shift between pH 6.5 and 7.5, with the GFP structure at pH 6.5 being very sensitive to denaturation by SDS, urea, and heat. In the present study, we wanted to explore whether the stability or structure of the closely related BFP is also pH dependent. As expected, we found heat-induced denaturation and renaturation of BFP to be pH dependent, very much like GFP. However, when exposed to other denaturants like urea/heat or SDS we found BFP to behave very differently than GFP. Unlike GFP, which at pH 8.5 and 7.5 is very resistant to SDS-induced denaturation, BFP readily lost about 20% of its fluorescence at pH 8.5 and about 60% fluorescence at pH 7.5. Also, our denaturation and renaturation studies show that under certain conditions, BFP is more stable than GFP, such that under conditions where GFP is completely denatured, BFP still retained significant fluorescence. Taken together, our preliminary results show that despite being very similar in both amino acid sequences and overall structures, there may be subtle and important structural/conformational differences between BFP and GFP.  相似文献   

17.
When urea is added to ribonuclease A that has already been denatured by salt (CaCl2, LiClO4 or LiCl were used), a second co-operative transition occurs, supporting the previous demonstration that these salts cause only partial denaturation. Also we have studied the effect of the salts on the urea denaturation, and the effect of urea on the salt denaturation. At low concentrations urea makes the salt transitions occur at lower concentrations, but at higher concentrations it changes the transition so that the completely disordered protein found in urea is produced by the salt. At low concentrations the salts actually stabilize the protein against denaturation by urea, but at higher concentrations they destabilize it. The data are presented in “phase diagrams” which are found to be very useful for such three-component systems.  相似文献   

18.
The fungicide dodine combines the cooperative denaturation properties of guanidine with the mM denaturation activity of SDS. It was previously tested only on two small model proteins. Here we show that it can be used as a chemical denaturant for phosphoglycerate kinase (PGK), a much larger two‐domain enzyme. In addition to its properties as a chemical denaturant, dodine facilitates thermal denaturation of PGK, and we show for the first time that it also facilitates pressure denaturation of a protein. Much higher quality circular dichroism and amide I′ infrared spectra of PGK can be obtained in dodine than in guanidine, opening the possibility for use of dodine as a denaturant when UV or IR detection is desirable. One caution is that dodine denaturation, like other detergent‐based denaturants, is less reversible than guanidine denaturation.  相似文献   

19.
The strong specific binding of streptavidin (SA) to biotin is utilized in numerous biotechnological applications. The SA tetramer is also known to exhibit significant stability, even in the presence of sodium dodecylsulfate (SDS). Despite its importance, relatively little is known about the nature of the thermal denaturation pathway for SA. This work uses a homogeneous SA preparation to expand on the data of previous literature reports, leading to the proposal of a model for temperature induced structural changes in SA. Temperature dependent data were obtained by SDS and native polyacrylamide gel electrophoresis (PAGE), differential scanning calorimetry (DSC), and fluorescence and ultraviolet (UV)-visible spectroscopy in the presence and absence of SDS. In addition to the development of this model, it is found that the major thermal transition of SA in 1% SDS is reversible. Finally, although SA exhibits significant precipitation at elevated temperatures in aqueous solution, inclusion of SDS acts to prevent SA aggregation.  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