首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
R B Vallee  R C Williams 《Biochemistry》1975,14(12):2574-2580
Upon exposure to conditions of low pH, beef B4 lactate dehydrogenase rapidly loses enzymatic activity, but this process can be completely reversed yielding 100% of the original activity if the enzyme is immediately returned to neutral conditions. As the time of exposure to low pH is increased, the fraction of activity recovered declines to a values of 50--60% and remains nearly constant over a period of many hours. Correlated with this behavior is a change in the kinetics of the recovery of activity. Recovery of activity has been shown to be a second-order process for enzyme exposed to low pH for brief periods of time (Anderson, S., and Weber, G. (1966), Arch. Biochem. Biophys. 116, 207). After several minutes at low pH recovery of activity is found to become first order and to occur at a considerably slower rate. Gel filtration chromatography at low pH separates the protein into two fractions. The lower molecular weight fraction is found to be primarily monomeric, as indicated by equilibrium ultracentrifugation, and is capable of recovering enzymatic activity. The higher molecular weight fraction is generated from the lower molecular weight fraction, and is incapable of recovering activity. These results are interpreted to indicate that the enzyme exists sequentially in three denatured forms at low pH, the first two capable of being restored to the native state, and the third irreversibly denatured.  相似文献   

2.
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   

3.
The action of 5 m urea on bovine serum albumin has been studied at pH 9.0 and 25°C. Analysis by the acrylamide gel electrophoresis revealed the presence of a few components 1, 1′, 2, 3, 4 and 5. The components 1 and 1′ are monomers, component 2 is a dimer, and components 3, 4 and 5 are aggregates. In presence of SH blocking reagent, bovine serum albumin gave only the zone 1, indicating that the components 1′-5 were formed by the SH to S-S exchange reactions. Component 1′ was formed by the intramolecular SH to S-S exchange reaction, and components 2–5 were formed by the intermolecular exchange reaction. Addition of cysteine either to bovine serum albumin or to the SH-blocked bovine serum albumin increased the percent of zone 1′, indicating that a complex bovine serum albumin-cysteine was formed or that the SH-catalyzed structural alteration occurred in bovine serum albumin. Components 1, 1′, 2 and 3 were isolated separately by the preparative disc gel electrophoresis. The sedimentation coefficients 1 and 1′ differed slightly indicating that they were different monomers, and values were slightly smaller than the normal value of bovine serum albumin, indicating that these components were in slightly expanded state. Isolated component 1 was exposed to 5 m urea again, but no further change occurred. This supports the concept of microheterogeneity of bovine serum albumin.  相似文献   

4.
Tetrameric rabbit muscle aldolase is dissociated to the inactive monomer at strongly alkaline pH (pH greater than or equal to 12). As shown by sedimentation velocity, fluorescence emission, and specific activity, the final profiles of dissociation, denaturation, and deactivation run parallel. Increasing incubation time proves the enzyme to be metastable in the pH range of deactivation. At 10 less than pH less than 12 "hysteresis" of the deactivation-reactivation reaction is observed. Short incubation at pH greater than or equal to 12 leads to high yields of reactivation (greater than or equal to 60%), while irreversibly denatured enzyme protein is the final product after long incubation. The kinetics of reconstitution under essentially irreversible conditions (pH 7.6) can be described by a sequential uni-bimolecular mechanism, assuming partial activity of the isolated subunits. The kinetic constants correspond to those observed for the reactivation after denaturation at acid pH or in 6M guanidine. HCl. Obviously the pH-dependent deactivation and reactivation of aldolase at alkaline pH obeys the general transconformation/association model which has been previously reported to hold for the reconstitution of numerous oligomeric enzymes after denaturation in various denaturants.  相似文献   

5.
Thermal denaturation of bacteriophage T4 lysozyme at neutral pH   总被引:7,自引:0,他引:7  
W J Becktel  W A Baase 《Biopolymers》1987,26(5):619-623
  相似文献   

6.
The change in apparent molal volume ? of DNA on thermal denaturation in carbonate buffer at pH 11.0 has been determined by the dilatometric method. It was found that ? increases sigmoidally during the helix–coil transition. Several methods, including a colorimetric technique that closely simulates the conditions used in the dilatometric experiments, were employed to estimate the protons lost by the DNA during the transition. These measurements indicated that the extent of the proton loss depends on the counterion present, increasing in the order Li+ < Na+ < K+ < Cs+. The major part of the volume changes observed during the denaturation is due to the volume changes expected to accompany the transfer of protons from the bases guanine and thym ne to carbonate ions. As has been previously reported for the denaturation of DNA at neutral pH, the volume change directly due to the change in shape of the polymer molecules is so small as to be experimentally undetectable.  相似文献   

7.
T R Cech  F Reisler  J E Hearst 《Biochemistry》1976,15(9):1865-1873
A new technique--partial denaturation of DNA in equilibrium CsCl density gradients at pH 11.4--is used to determine the distribution of intermediate states in the melting of mouse DNA. When the technique is applied in the preparative ultracentrifuge, the DNA is fractionated according to stability. Neutralization of the partially denatured fractions results in the recovery of most of the DNA in its native form. The individual fractions are more homogeneous than the total DNA: they have decreased density heterogeneity (smaller band widths), neutral CsCl buoyant densities that differ from the average, and more homogeneous melting profiles with melting temperatures that differ from the average.  相似文献   

8.
Dilatometric measurements were made to determine the change in apparent specific volume φ of DNA resulting from thermal denaturation in neutral solution, φ increased continuously with temperature in the range 10–85°C. No deviations from a monotonically rising curve were observed in the φ versus temperature profile in the region of the melting temperature. The results are interpreted in terms of a partial loss of the preferentially bound DNA hydration shell. The nature of the well known buoyant density difference between native and denatured DNA was investigated by evaluating the densities in a series of cesium salt gradients at constant temperature. Extrapolation of the results to zero water activity indicates that the partial specific volumes of anhydrous native and denatured DNA are equal. The density difference at nonzero water activities is attributed to decreased hydration in the denatured state. The absence of a related change in φ accompanying the denaturation in the dilatometric experiments suggests that the probable volume change associated with loss of bound water during denaturation is accompanied by other compensatory volume effects. The possible nature of these volume effects is discussed.  相似文献   

9.
10.
The conformational stability and the folding process of alpha, beta and Psi bovine trypsin at pH 3.0 followed by circular dichroism (CD) and size exclusion in HPLC have been analyzed as a function of urea concentration. The thermodynamic stability for a and b are deltaG = 15.91 +/- 0.28 kcal/mol, deltaG = 15.54 +/- 2.39 kcal/mol. respectively, and y trypsin is deltaG = 16.10 +/- 2.51 kcal/mol. The transition curves for alpha, beta and Psi forms suggest a molten globule state.  相似文献   

11.
The heat-induced denaturation kinetics of two different sources of ovalbumin at pH 7 was studied by chromatography and differential scanning calorimetry. The kinetics was found to be independent of protein concentration and salt concentration, but was strongly dependent on temperature. For highly pure ovalbumin, the decrease in nondenatured native protein showed first-order dependence. The activation energy obtained with different techniques varied between 430 and 490 kJ*mole(-1). First-order behavior was studied in detail using differential scanning calorimetry. The calorimetric traces were irreversible and highly scan rate-dependent. The shape of the thermograms as well as the scan rate dependence can be explained by assuming that the thermal denaturation takes place according to a simplified kinetic process where N is the native state, D is denatured (or another final state) and k a first-order kinetic constant that changes with temperature, according to the Arrhenius equation. A kinetic model for the temperature-induced denaturation and aggregation of ovalbumin is presented. Commercially obtained ovalbumin was found to contain an intermediate-stable fraction (IS) of about 20% that was unable to form aggregates. The denaturation of this fraction did not satisfy first-order kinetics.  相似文献   

12.
DSC analysis has been used to quantify the reversibility of unfolding following thermal denaturation of lysozyme. Since the temperature at which protein unfolding occurs, Tm, varies with different solution conditions, the effect on the melting temperature and the degree of refolding after thermal denaturation in low ionic strength sodium phosphate buffers (5–1000 mM) over a range of pH (5–9) in the presence/absence of disaccharides is examined. This study compares the enthalpies of unfolding during successive heating cycles to quantify reversibility following thermal denaturation. The disaccharides, trehalose and maltose were used to assess if the disaccharide induced increase in Tm is reflected in the reversibility of thermally induced denaturation. There was extensive overlap between the Tm values where non-reversible and reversible thermal denaturation occurred. Indeed, for pH 6, at the highest and lowest Tm, no refolding was observed whereas refolding was observed for intermediate values, but with similar Tm values having different proportions of refolded protein. We established a method to measure the degree of reversible unfolding following thermal denaturation and hence indirectly, the degree to which protein is lost to irreversible aggregation, and show that solution conditions which increase melt transition temperatures do not automatically confer an increase in reversibility. This type of analysis may prove useful in assessing the stability of proteins in both the biopharmaceutical and food industries.  相似文献   

13.
In order to broaden the scope and increase the utility of differential scanning calorimetry, a theoretical model of calorimetric thermograms is presently proposed which facilitates their biophysical interpretation and accounts explicitly for their modifications induced by denaturing agents and/or pH. The model rests mainly on statistical-physical considerations, the denaturation-linked increase of the number of binding sites for denaturants (including H+) serving as the conceptual basis for thermogram modelling. Denaturants were envisioned as contributing indirectly to thermal denaturation by forming complexes preferentially with unfolded protein molecules, shifting thus the equilibrium towards the denatured phase. After postulating the probability of complex formation, mean numbers of the relevant molecular species were computed by ensemble averaging. Finally, an eight-parameter expression has been derived defining protein heat capacity as a function of both temperature and denaturant concentration (or pH), each of the eight parameters having a distinct biophysical meaning. The model has been tested by applying it to the prediction of the pH-dependence of thermograms. Four proteins have been considered (lysozyme, myoglobin, apomyoglobin, and ribonuclease A), each represented by a series of three to four published thermograms recorded under different pH conditions. Model equations, fitted simultaneously to all thermograms in a pH series, reproduced correctly experimental tracings. Parameter values obtained as best-fit requirements (particularly those representing the number of binding sites unmasked by denaturation and the free energy of ion binding) were in close agreement with empirical, mainly potentiometric, data from literature. The empirically established pH-independence of the total enthalpy of denaturation, the phenomenon of cold denaturation, the pH-dependence of the Gibbs free energy of denaturation, of the melting temperature and of the temperature of cold denaturation, were all correctly predicted by the model. Combined effects of multiple denaturants, including the effects of pH in the presence of denaturants other than protons, are also predictable by the model.  相似文献   

14.
We have previously shown that the denaturation of TK with urea follows a non-aggregating though irreversible denaturation pathway in which the cofactor binding appears to become altered but without dissociating, then followed at higher urea by partial denaturation of the homodimer prior to any further unfolding or dissociation of the two monomers. Urea is not typically present during biocatalysis, whereas access to TK enzymes that retain activity at increased temperature and extreme pH would be useful for operation under conditions that increase substrate and product stability or solubility. To provide further insight into the underlying causes of its deactivation in process conditions, we have characterised the effects of temperature and pH on the structure, stability, aggregation and activity of Escherichia coli transketolase. The activity of TK was initially found to progressively improve after pre-incubation at increasing temperatures. Loss of activity at higher temperature and low pH resulted primarily from protein denaturation and subsequent irreversible aggregation. By contrast, high pH resulted in the formation of a native-like state that was only partially inactive. The apo-TK enzyme structure content also increased at pH 9 to converge on that of the holo-TK. While cofactor dissociation was previously proposed for high pH deactivation, the observed structural changes in apo-TK but not holo-TK indicate a more complex mechanism.  相似文献   

15.
16.
Equilibrium unfolding of barstar with guanidine hydrochloride (GdnHCl) and urea as denaturants as well as thermal unfolding have been carried out as a function of pH using fluorescence, far-UV and near-UV CD, and absorbance as probes. Both GdnHCl-induced and urea-induced denaturation studies at pH 7 show that barstar unfolds through a two-state F<->U mechanism and yields identical values for delta GU, the free energy difference between the fully folded (F) and unfolded (U) forms, of 5.0 +/- 0.5 kcal.mol-1 at 25 degrees C. Thermal denaturation of barstar also follows a two-state F<->U unfolding transition at pH 7, and the value of delta GU at 25 degrees C is similar to that obtained from chemical denaturation. The pH dependence of denaturation by GdnHCl is complex. The Cm value (midpoint of the unfolding transition) has been used as an index for stability in the pH range 2-10, because barstar does not unfold through a two-state transition on denaturation by GdnHCl at all pH values studied. Stability is maximum at pH 2-3, where barstar exists in a molten globule-like form that forms a large soluble oligomer. The stability decreases with an increase in pH to 5, the isoelectric pH of the protein. Above pH 5, the stability increases as the pH is raised to 7. Above pH 8, it again decreases as the pH is raised to 10. The decrease in stability from pH 7 to 5 in wild-type (wt) barstar, which is shown to be characterized by an apparent pKa of 6.2 +/- 0.2, is not observed in H17Q, a His 17-->Gln 17 mutant form of barstar. This decrease in stability has therefore been correlated with the protonation of His 17 in barstar. The decrease in stability beyond pH 8 in wt barstar, which is characterized by an apparent pKa of 9.2 +/- 0.2, is not detected in BSCCAA, the Cys 40 Cys 82-->Ala 40 Ala 82 double mutant form of barstar. Thus, this decrease in stability has been correlated with the deprotonation of at least one of the two cysteines present in wt barstar. The increase in stability from pH 5 to 3 is characterized by an apparent pKa of 4.6 +/- 0.2 for wt barstar and BSCCAA, which is similar to the apparent pKa that characterizes the structural transition leading to the formation of the A form. The use of Cm as an index of stability has been supported by thermal denaturation studies.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

17.
Several reports have pointed out the existence of intermediate states (both kinetic and equilibrium intermediate) between the native and the denatured states. The molten globule state, a compact intermediate state in which the secondary structure is formed but the tertiary structure fluctuates considerably, is currently being studied intensively because of its possible implication in the folding process of several proteins. We have examined the thermal stability of horse cytochrome c at low pH between 2.0 and 3.2 and different potassium chloride concentrations by absorbance of the Soret band, far and near-ultraviolet circular dichroism (u.v. c.d.) and tryptophan fluorescence using a multidimensional spectrophotometer. The concentration of potassium chloride ranged from 0 M to 0.5 M. The experimental thermal denaturation curves show that: (1) the helical content of cytochrome c remains stable at higher temperature when the concentration of salt is increased; whereas (2) the extent of ordering of the tertiary structure is weakly dependent on salt concentration; and (3) for cytochrome c, the stabilization of the molten globule state is induced by the binding of anions. Other salts such as NaCl, LiCl, potassium ferricyanide (K3Fe(CN)6) and Na2SO4 may also be used to stabilize the molten globule state. The thermodynamic analysis of the denaturation curves of c.d. at 222 nm and c.d. at 282 nm shows that, whereas a two-state (native and denatured) transition is observed at low-salt concentration, the far and near-u.v. c.d. melting curves of cytochrome c do not coincide with each other at high-salt concentration, and a minimum of three different thermodynamic states (IIb, intermediate or IIc, and denatured) is necessary to achieve a sufficient analysis. The intermediate state (called IIc) is attributed to the molten globule state because of its high secondary structure content and the absence of tertiary structure. Therefore, at low pH, cytochrome c is present in at least four states (native, IIb, IIc and denatured) depending on the salt concentration and temperature. The thermodynamic parameters, i.e. the Gibbs free energy differences (delta G), the enthalpy differences (delta H), the midpoint temperatures (Tm) of the transition (IIb in equilibrium intermediate (IIc in equilibrium denatured) are determined. We also give estimates of the heat capacity differences (delta Cp) from the temperature dependence of the enthalpy differences. The enthalpy change and the heat capacity difference of the IIc in equilibrium denatured transition are non-zero. The number of charges (protons or chloride anions) released upon transitions are determined by analysing the pH and chloride anion concentration dependence of the Gibbs free energy.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

18.
19.
A Tamura  K Kimura  K Akasaka 《Biochemistry》1991,30(47):11313-11320
Structural transitions of the protein Streptomyces subtilisin inhibitor (SSI) from the native state to the cold-denatured and heat-denatured states were studied by 1H NMR spectroscopy in the temperature range from -10 to 60 degrees C in the acidic pH range. Assignments of some of the 1H NMR signals of SSI in the cold-denatured and heat-denatured states were performed by a combined use of selective deuteration and site-directed mutagenesis. Throughout the pH range from 2.1 to 3.1, both transitions were cooperative and basically only three distinct spectra corresponding to structures in the cold-denatured, native, and heat-denatured states were detected. In the cold-denatured state, the side-chain signals of Met73, His106, at least one Val, and two Leu were observed at distinctly shifted positions from those for a random-coiled structure, suggesting the formation of a tertiary structure, while those of Met70, His43, and Ala2 were observed at positions for a random-coiled structure. This tertiary structure in the cold-denatured state is entirely different from that in the native state, as some amino acid residues exposed to the solvent in the native state (e.g., Met73, His106) are buried while those sequestered in the native state (e.g., His43) are exposed. In the heat-denatured state, however, most 1H NMR signals were observed at random-coiled positions, indicating that there is much less tertiary structure in the heat-denatured state than in the cold-denatured state. At pH values below 2.09, a structural transition was observed from the cold-denatured state to the heat-denatured state without passing through the native state.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
Upon heating cytochrome c peroxidase (ferrocytochrome c: hydrogen-peroxide oxidoreductase, EC 1.11.1.5) at pH 4 and 5, the enzyme precipitates at 41 degrees C and 51 degrees C, respectively. Incubating the enzyme at lower temperatures causes a slow dissociation of the heme from the protein. The heme precipitates, while the apoprotein remains soluble. Between pH 6 and 8, the native enzyme is converted to a low-spin ferric form upon heating. The Soret maximum shifts from 408 to 414 nm. The midpoint of this transition is pH-dependent, with a value of 46 degrees C at pH 6 decreasing to 29 degrees C at pH 8. At high temperatures the 414 nm form is converted to a species which has a 'free heme' spectrum with low absorptivity and Soret maximum at 390 nm. The midpoint temperature of this latter transition is 62 degrees C and 57 degrees C at pH 7 and 8, respectively.  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