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.
To study the influence of the carbohydrate-moiety of ovalbumin on the formation of the heat-stable conformer S-ovalbumin, ovalbumin is deglycosylated with PNGase-F under native conditions. Although the enzymatic deglycosylation procedure resulted in a complete loss of the ability to bind to Concavalin A column-material, only in about 50% the proteins lost their complete carbohydrate moiety, as demonstrated by mass spectrometry and size exclusion chromatography. Thermal stability and conformational changes were determined using circular dichroism and differential scanning calorimetry and demonstrated at ambient temperature no conformational changes due to the deglycosylation. Also the denaturation temperature of the processed proteins remained the same (77.4 +/- 0.4 degrees C). After heat treatment of the processed protein at 55 degrees C and pH 9.9 for 72 h, the condition that converts native ovalbumin into the heat-stable conformer (S-ovalbumin), only the material with the intact carbohydrate moiety forms this heat-stable conformer. The material that effectively lost its carbohydrate moiety appeared fully denatured and aggregated due to these processing conditions. These results indicate that the PNGase-F treatment of ovalbumin prohibits the formation and stabilization of the heat-stable conformer S-ovalbumin. Since S-ovalbumin in egg protein samples is known to affect functional properties, this work illustrates a potential route to control the quality of egg protein ingredients.  相似文献   

2.
To provide a molecular explanation of the role of the disulfide (SS) bridge in the thermostability and structural integrity of ovalbumin (OVA), we prepared SS-mutated OVAs in which SS-forming residues were replaced by Ala or Ser (C73A, C73S, C120A, and C73/120A), and compared the conformation, thermostability, susceptibility to elastase, and formation of heat-stable OVA (S-OVA) with those of the wild-type. The circular dichroism (CD) and tryptophan fluorescence spectra revealed that the SS-mutated OVAs assumed a native-like conformation similar to the wild-type. The thermal denaturation temperature for the SS-mutated OVAs was significantly lower than that for the wild-type. C73S, C120A, and C73/120A mutants converted to S-OVA on alkaline treatment. Analyses for elastase digestion fragments showed that a non-native SS bridge was generated in all SS-mutated OVAs, but non-native SS-pairing did not contribute to thermostability. Hence, we concluded that the presence of the original SS bridge in OVA contributes to conformational stability but is not directly related to the conversion to S-OVA.  相似文献   

3.
To provide a molecular explanation of the role of the disulfide (SS) bridge in the thermostability and structural integrity of ovalbumin (OVA), we prepared SS-mutated OVAs in which SS-forming residues were replaced by Ala or Ser (C73A, C73S, C120A, and C73/120A), and compared the conformation, thermostability, susceptibility to elastase, and formation of heat-stable OVA (S-OVA) with those of the wild-type. The circular dichroism (CD) and tryptophan fluorescence spectra revealed that the SS-mutated OVAs assumed a native-like conformation similar to the wild-type. The thermal denaturation temperature for the SS-mutated OVAs was significantly lower than that for the wild-type. C73S, C120A, and C73/120A mutants converted to S-OVA on alkaline treatment. Analyses for elastase digestion fragments showed that a non-native SS bridge was generated in all SS-mutated OVAs, but non-native SS-pairing did not contribute to thermostability. Hence, we concluded that the presence of the original SS bridge in OVA contributes to conformational stability but is not directly related to the conversion to S-OVA.  相似文献   

4.
The precipitating of effect of sodium dodecyl sulphate (SDS) on the egg white proteins ovalbumin, conalbumin and lysozyme was studied at 25 degrees C and at different pH values. The proteins precipitated below their respective isolectric points, provided the detergent to protein ratio was appropriate. The pH profile of precipitation was different for the three proteins reflecting net charge differences. The binding of SDS to the proteins was studied with [35S]-labelled SDS and for ovalbumin a ratio of 21--28 SDS molecules/protein molecule, dependent on the concentration of SDS initially used, seem to be required for precipitation at pH 4.5. This number, however, is dependent on pH and increases with an increased positive net charge of the protein. The precipitating effect of SDS was identical for ovalbumin, conalbumin and lysozyme when compared on a gram to gram basis (0.1--0.15 g SDS/g precipitated protein). The precipitated protein was denatured as measured by differential scanning calorimetry, but was also completely redissolved if pH was increased to above the isoelectric point. The precipitating effecto f SDS was also examined at elevated temperatures. The two-phase systems of the proteins induced by SDS at 25 degrees C were heated from 25 degrees C to 90 degrees C at a rate of 1.25 degrees C/min. The precipitation behaviour was similar for the three proteins upon heating. When the SDS concentration was increased the precipitation curves were transferred towards lower temperatures and the courses of precipitation became less sharp. The synergistic effect of SDS and heat on protein precipitation was differentiated by denaturation measurements and radioactive labelling. The ratio SDS to precipitated protein gradually diminished towards higher temperatures but no purely thermal precipitation was found.  相似文献   

5.
Freezing denaturation of ovalbumin at acid pH   总被引:1,自引:0,他引:1  
The effects of rapid freezing and thawing at acid pH on the physiochemical properties of ovalbumin were examined. At low pH (around 2), UV difference spectra showed microenvironmental changes around the aromatic amino acid residues; elution curves by gel permeation chromatography showed decreasing numbers of monomers after neutralization. These changes depended on the incubation temperature (between -196 and -10 degrees C) and the protein concentration (0.5-10 mg/ml), and a low concentration of ovalbumin incubated at around -40 degrees C suffered the most damage to its conformation. With freezing and then incubation at -40 degrees C, three of the four sulfhydryl groups in the ovalbumin molecule reacted with 2,2'-dithiodipyridine. The CD spectra showed these changes in the secondary structure, but they were smaller than those when guanidine hydrochloride was used for denaturation. Supercooling at -15 degrees C or freezing at -196 degrees C had little or no effect on the conformation of the ovalbumin molecule. Thus, irreversible conformational changes of ovalbumin were caused under the critical freezing condition at an acid pH. These changes arose from partial denaturation and resembled those with thermal denaturation of ovalbumin at neutral pH.  相似文献   

6.
The thermal denaturation, aggregation, and degradation of hen egg white ovalbumin dissolved in distilled and deionized water (60 mg/ml, pH 7.5) was investigated by differential scanning calorimetry (DSC), polyacrylamide gel electrophoresis (PAGE), and viscosity measurement. Two independent endothermic peaks were observed up to 180 degrees C by the DSC analysis. The first peak appeared at around 80 degrees C, corresponding to the denaturation temperature of ovalbumin. The second peak occurred around 140 degrees C due to the degradation of protein molecules as judged from the analysis by SDS-PAGE. The viscosity of the ovalbumin solution increased dramatically above 88 degrees C and maintained almost the same value up until heating to 140 degrees C. The increase in viscosity after heating to 88 degrees C was due to the denaturation and subsequent aggregation of ovalbumin molecules as observed by SDS-PAGE. The decrease in viscosity of the samples heated above 150 degrees C appears to have been the result of degradation of the ovalbumin molecules.  相似文献   

7.
The pH dependence of the reversible guanidine hydrochloride denaturation of the major fraction of ovalbumin (ovalbumin A1) was studied by a viscometric method in the pH range 1-7, at 25 degrees C and at six different denaturant concentrations (1.5-2.6 M). At any denaturant concentrationa reduction in pH favoured the transition from the native to the denatured state. The latter was essentially 'structureless', as revealed by the fact that the reduced viscosity of the acid and guanidine hydrochloride denatured state of ovalbumin A1 (obtained at different denaturant concentrations in acidic solutions) was measured (at a protein concentration of 3.8 mg/ml) to be 29.2 ml/g which is identical to that found in 6 M guanidine hydrochloride wherein the protein behaves as a cross-linked random coil. A quantitative analysis of the results on the pH dependence of the equilibrium constant for the denaturation process showed that on denaturation the intrinsic pK of two carboxyl groups in ovalbumin A1 went up from 3.1 in the native state to 4.4 in the denatured state of the protein.  相似文献   

8.
Differential scanning calorimetry (DSC) was used to assay thermal transitions that might be responsible for cell death and other responses to hyperthermia or heat shock, such as induction of heat shock proteins (HSP), in whole Chinese hamster lung V79 cells. Seven distinct peaks, six of which are irreversible, with transition temperatures from 49.5 degrees C to 98.9 degrees C are detectable. These primarily represent protein denaturation with minor contributions from DNA and RNA melting. The onset temperature of denaturation, 38.7 degrees C, is shifted to higher temperatures by prior heat shock at 43 degrees and 45 degrees C, indicative of irreversible denaturation occurring at these temperatures. Thus, using DSC it is possible to demonstrate significant denaturation in a mammalian cell line at temperatures and times of exposure sufficient to induce hyperthermic damage and HSP synthesis. A model was developed based on the assumption that the rate limiting step of hyperthermic cell killing is the denaturation of a critical target. A transition temperature of 46.3 degrees C is predicted for the critical target in V79 cells. No distinct transition is detectable by DSC at this temperature, implying that the critical target comprises a small fraction of total denaturable material. The short chain alcohols methanol, ethanol, isopropanol, and t-butanol are known hyperthermic sensitizers and ethanol is an inducer of HSP synthesis. These compounds non-specifically lower the denaturation temperature of cellular protein. Glycerol, a hyperthermic protector, non-specifically raises the denaturation temperature for proteins denaturing below 60 degrees C. Thus, there is a correlation between the effect of these compounds on protein denaturation in vivo and their effect on cellular sensitivity to hyperthermia.  相似文献   

9.
Circular dichroism was used to monitor the thermal unfolding of ribonuclease A in 50% aqueous methanol. The spectrum of the protein at temperatures below -10 degrees C (pH* 3.0) was essentially identical to that of native ribonuclease A in aqueous solution. The spectrum of the thermally denatured material above 70 degrees C revealed some residual secondary structure in comparison to protein unfolded by 5 M Gdn.HCl at 70 degrees C in the presence or absence of methanol. The spectra as a function of temperature were deconvoluted to determine the contributions of different types of secondary structure. The position of the thermal unfolding transition as monitored by alpha-helix, with a midpoint at 38 degrees C, was at a much higher temperature than that monitored by beta-sheet, 26 degrees C, which also corresponded to that observed by delta A286, tyrosine fluorescence and hydrodynamic radius (from light scattering measurements). Thus, the loss of beta-sheet structure is decoupled from that of alpha-helix, suggesting a step-wise unfolding of the protein. The transition observed for loss of alpha-helix coincides with the previously measured transition for His-12 by NMR from a partially folded state to the unfolded state, suggesting that the unfolding of the N-terminal helix in RNase A is lost after unfolding of the core beta-sheet during thermal denaturation. The thermally denatured protein was relatively compact, as measured by dynamic light scattering.  相似文献   

10.
The phosphotriesterase (PTE) from Pseudomonas diminuta, a metalloenzyme that catalyses the hydrolysis of organophosphorus pesticides and nerve agents, has been described as a remarkably heat-stable protein [Grimsley et al., Biochemistry 36 (1997), 14366-14374]. Because substitution of the naturally occurring zinc ions by cobalt ions was found to enhance the enzyme catalytic activity, we investigated the thermal stability of the Co(2+)/Co(2+)-PTE. This study, carried out using capillary electrophoresis under optimised conditions in the pH range 9-10 compatible with optimal enzyme activity, provided evidence for irreversible denaturation according to the Lumry-Eyring model. A temperature-induced conformational transition (T(m) approximately equal to 58 degrees C) and an early growing of aggregates were observed. Comparison of UV spectra with heat-induced inactivation data clearly demonstrated that the PTE state populated above T(m) was neither native nor active. Differential scanning calorimetry showed only an exothermic trace due to aggregation of the denatured protein at T=76 degrees C. Accordingly, the temperature-induced denaturation process of the PTE could be described by a consecutive reaction model, including formation of an intermediate with enhanced activity at T approximately equal to 45 degrees C and an inactive unfolded state populated at T approximately equal to 58 degrees C, which leads to denatured aggregates. Thus, the wild type Co(2+)/Co(2+)-PTE displays a middle-range thermostability. Hence, for decontamination purposes under extreme Earth temperatures, wild type and engineered mutants of PTE substituted with other metal cations should be evaluated.  相似文献   

11.
The existence of a practical minimum pH for the dissolution of heat-induced whey gels in alkaline solutions has been studied using beta-lactoglobulin (betaLg) as a model protein. A sharp transition in solubility was observed between pH 11 and 12; this transition shifts to higher pHs for gels formed at higher temperatures and for longer gelling times. The breakdown reactions of heat-induced aggregates in alkali were monitored with size exclusion chromatography. The destruction of large aggregates was faster at higher pH and also showed a transition between pH 11 and 12. Using tryptophan fluorescence and near- and far-UV circular dichroism, this transition was assigned to the base-induced denaturation observed in solutions of aggregates (pK 11.53). It is suggested that the high protein repulsion caused by the large number of charges at pH > 11.5 drives the unfolding of the protein and the disruption of the intermolecular noncovalent bonds. Concentrated urea and GuHCl were found to be less effective than a pH 12 solution in destroying large aggregates. Aggregates formed for a long time (80 degrees C for 24 h) contained a larger number of intermolecular disulfide bonds that hinder the dissolution process. Gels formed at low temperatures (65 degrees C for 60 min), with fewer intermolecular noncovalent bonds, showed a similar solubility-pH profile to that observed for the base-induced denaturation of unheated beta-lactoglobulin (betaLg) (pK 10.63).  相似文献   

12.
The kinetics of thermal denaturation of a biliprotein, C-phycocyanin (C-PC) isolated from Spirulina platensis were studied at different pH values, ranging from 4.0 to 8.0. The denaturation of C-PC follows the first order kinetics and rate constant at pH 5.0 and temperature 55 degrees C is found to be 4.37 x 10(-5) s(-1), which increases to 5.46 x 10(-1) s(-1) at pH 7.0. The denaturation rate is much higher at 65 degrees C and pH 7.0 (7.96 x 10(-4)), as compared to at pH 5.0 (1.46 x 10(-4)). The thermal stability of C-PC is more at pH 5.0, as compared to other pH values. The observed differences in entropy values at pH 5.0, as compared to other pH values indicate a considerably close fit structure of the protein at pH 5.0, which increases the stability of native structure, even at higher temperature (65 degrees C).  相似文献   

13.
In order to elucidate the stabilization mechanism of CutA1 from Pyrococcus horikoshii (PhCutA1) with a denaturation temperature of nearly 150 degrees C, GuHCl denaturation and heat denaturation were examined at neutral and acidic pHs. As a comparison, CutA1 proteins from Thermus thermophilus (TtCutA1) and Oryza sativa (OsCutA1) were also examined, which have lower optimum growth temperatures of 75 and 28 degrees C, respectively, than that (98 degrees C) of P. horikoshii. GuHCl-induced unfolding and refolding curves of the three proteins showed hysteresis effects due to an unusually slow unfolding rate. The midpoints of refolding for PhCutA1, TtCutA1 and OsCutA1 were 5.7 M, 3.3 M, and 2.3 M GuHCl, respectively, at pH 8.0 and 37 degrees C. DSC experiments with TtCutA1 and OsCutA1 showed that the denaturation temperatures were remarkably high, 112.8 and 97.3 degrees C, respectively, at pH 7.0 and that the good heat reversibility was amenable to thermodynamic analyses. At acidic pH, TtCutA1 showed higher stability to both heat and denaturant than PhCutA1. Combined with the data for DSC and denaturant denaturation, the unfolding Gibbs energy of PhCutA1 could be depicted as a function of temperature. It was experimentally revealed that (1) the unusually high stability of PhCutA1 basically originates from a common trimer structure of the three proteins, (2) the stability of PhCutA1 is superior to those of the other two CutA1s over all temperatures above 0 degrees C at neutral pH, due to the decrease in both enthalpy and entropy, and (3) ion pairs of PhCutA1 contribute to the unusually high stability at neutral pH.  相似文献   

14.
Deposition kinetics of beta-lactoglobulin at a solid-liquid interface was studied with optical waveguide lightmode spectroscopy (OWLS) over a range of temperatures between 61 and 83 degrees C. A new temperature-controlled cell for OWLS measurements allows fast, on-line monitoring of the deposit formation at elevated temperatures. Primary protein layers were deposited at 25 degrees C in order to precondition and stabilize the waveguide surface. Sustained deposition lasting from a few minutes (around 80 degrees C) to hours (below 70 degrees C) resulted in multilayer deposits up to several tens of nanometers thick. The measured deposition rates were strongly influenced by temperature, pH, and NaCl concentration. Deposition rates decreased with increasing pH from 5.5. to 7.4, in a trend similar to that for noncovalent aggregation of beta-lactoglobulin in solution. Activation energies for deposition rates decreased with increasing pH, from 340 kJ/mol at pH 5.5 to 230 kJ/mol at pH 7.4 and were similar to the activation energies for denaturation of beta-lactoglobulin in solution.  相似文献   

15.
The wild-type p53 protein can be driven into a conformation corresponding to that adopted by structural mutant forms by heterodimerization with a mutant subunit. To seek partially folded states of the wild-type p53 core domain (p53C) we used high hydrostatic pressure (HP) and subzero temperatures. Aggregation of the protein was observed in parallel with its pressure denaturation at 25 and 37 degrees C. However, when HP experiments were performed at 4 degrees C, the extent of denaturation and aggregation was significantly less pronounced. On the other hand, subzero temperatures under pressure led to cold denaturation and yielded a non-aggregated, alternative conformation of p53C. Nuclear magnetic resonance (1H15N-NMR) data showed that the alternative p53C conformation resembled that of the hot-spot oncogenic mutant R248Q. This alternative state was as susceptible to denaturation and aggregation as the mutant R248Q when subjected to HP at 25 degrees C. Together these data demonstrate that wild-type p53C adopts an alternative conformation with a mutant-like stability, consistent with the dominant-negative effect caused by many mutants. This alternative conformation is likely related to inactive forms that appear in vivo, usually driven by interaction with mutant proteins. Therefore, it can be a valuable target in the search for ways to interfere with protein misfolding and hence to prevent tumor development.  相似文献   

16.
The thermostable tailspike endorhamnosidase of bacteriophage P22 has been investigated by laser Raman spectroscopy to determine the protein's secondary structure and the basis of its thermostability. The conformation of the native tailspike, determined by Raman amide I and amide III band analyses, is 52 to 61% beta-sheet, 24 to 27% alpha-helix, 15 to 21% beta-turn and 0 to 10% other structure types. The secondary structure of the wild-type tailspike, as monitored by the conformation-sensitive Raman amide bands, was stable to 80 degrees C, denatured reversibly between 80 and 90 degrees C, and irreversibly above 90 degrees C. The purified native form of a temperature-sensitive folding mutant (tsU38) contains secondary structures virtually identical to those in the wild-type in aqueous solution at physiological conditions (0.05 M-Na+ (pH 7.5], at both permissive (20 degrees C) and restrictive (40 degrees C) temperatures. This supports previous results showing that the mutational defect at 40 degrees C affects intermediates in the folding pathway rather than the native structure. At temperatures above 60 degrees C the wild-type and mutant forms were distinguishable: the reversible and irreversible denaturation thresholds were approximately 15 to 20 degrees C lower in the mutant than in the wild-type protein. The irreversible denaturation of the mutant tailspikes led to different aggregation/polymerization products from the wild-type, indicating that the mutation altered the unfolding pathway. In both cases only a small percentage of the native secondary structure was altered by irreversible thermal denaturation, indicating that the aggregated states retain considerable native structure.  相似文献   

17.
We have studied the stability and reassociation behaviour of native molecules of Rapana venosa hemocyanin and its two subunits, termed RvH1 and RvH2. In the presence of different concentrations of Ca(2+) and Mg(2+) ions and pH values, the subunits differ not only in their reassociation behaviour, but also in their formation of helical tubules and multidecamers. RvH1 revealed a greater stability at higher pH values compared to RvH2. Overall, the stability of reassociated RvH and its structural subunits was found to be pH-dependent. The increasing stability of native Hc and its subunits, shown by pH-induced CD transitions (acid and alkaline denaturation), can be explained with the formation of quaternary structure. The absence of a Cotton effect at temperatures 20-40 degrees C in the pH-transition curves of RvH2 indicates that this subunit is stabilized by additional "factors", e.g.: non-ionic/hydrophobic stabilization and interactions of carbohydrate moieties. A similar behaviour was observed for the T-transition curves in a wide pH interval for RvH and its structural subunits. At higher temperatures, many of the secondary structural elements are preserved especially at neutral pH, even at extreme high temperatures above 90 degrees C the protein structures resemble a "globule state".  相似文献   

18.
The soluble ATPase (adenosine triphosphatase) from Micrococcus lysodeikticus underwent a major unfolding transition when solutions of the enzyme at pH 7.5 were heated. The midpoint occurred at 46 degrees C when monitored by changes in enzymic activity and intrinsic fluorescence, and at 49 degrees C when monitored by circular dichroism. The products of thermal denaturation retained much secondary structure, and no evidence of subunit dissociation was detected after cooling at 20 degrees C. The thermal transition was irreversible, and thiol groups were not involved in the irreversibility. The presence of ATP, adenylyl imidodiphosphate, CaCl2 or higher concentrations of ATPase conferred stability against thermal denaturation, but did not prevent the irreversibility one denaturation had taken place. In the presence of guanidinium chloride, thermal denaturation occurred at lower temperatures. The midpoints of the transition were 45 degrees C in 0.25 M-, 38 degrees C in 0.5 M-and 30 degrees C in 0.75 M-denaturant. In the highest concentration of guanidinium chloride a similar unfolding transition induced by cooling was observed. Its midpoint was 9 degrees C, and the temperature of maximum stability of the protein was 20 degrees C. The discontinuities occurring the the Arrhenius plots of the activity of this enzyme had no counterpart in variations in the far-u.v. circular dichroism or intrinsic fluorescence of the protein at the same temperature.  相似文献   

19.
Thermal denaturation and aggregation of rabbit muscle glyceraldehyde-3-phosphate dehydrogenase (GAPDH) have been studied using differential scanning calorimetry (DSC), dynamic light scattering (DLS), and analytical ultracentrifugation. The maximum of the protein thermal transition (T(m)) increased with increasing the protein concentration, suggesting that the denaturation process involves the stage of reversible dissociation of the enzyme tetramer into the oligomeric forms of lesser size. The dissociation of the enzyme tetramer was shown by sedimentation velocity at 45 degrees C. The DLS data support the mechanism of protein aggregation that involves a stage of the formation of the start aggregates followed by their sticking together. The hydrodynamic radius of the start aggregates remained constant in the temperature interval from 37 to 55 degrees C and was independent of the protein concentration (R(h,0) approximately 21 nm; 10 mM sodium phosphate, pH 7.5). A strict correlation between thermal aggregation of GAPDH registered by the increase in the light scattering intensity and protein denaturation characterized by DSC has been proved.  相似文献   

20.
R H Hatley  F Franks 《FEBS letters》1989,257(1):171-173
The cold-induced denaturation of lactate dehydrogenase has been determined in an unfrozen, cryoprotectant free solution at sub-zero temperatures. The cold-induced denaturation temperature (TL) has been found to be -28 degrees C. The results for the first time clearly establish that temperature alone can induce denaturation in a cooled protein solution. The validity of earlier data, obtained in the presence of perturbants (particularly pH or guanidinium chloride), is discussed.  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