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.
The effects of various concentrations of urea and guanidine hydrochloride on enzyme activity and on subunit association were determined. Incubation of thymidylate synthetase with buffered solutions of 3M to 3.5M guanidine hydrochloride or 5 M to 6 M urea resulted in the loss of about 90% of the enzyme activity. Under these denaturing conditions a red shift of the fluorescence emission maximum from 340 nm to 351 nm was observed together with a significant decrease in the relative fluorescence intensity of the protein. Studies at both 4 degrees C and 25 degrees C indicated that the enzyme was in the dimer form in 2 M guanidine hydrochloride but was dissociated into monomers in concentrations of this denaturant of 3 M and above. Although only monomeric species were evident at 4 degrees C in 6 M urea, at 25 25 degrees C this denaturant caused protein aggregation which increased with decreasing phosphate buffer concentration. Enzyme (5 mg/ml) in 0.5 M potassium phosphate buffer, pH 6.8, containing 4 M guanidine hydrochloride gave a minimum S20, w value of 1.22S at 25 degrees C. Sedimentation behavior of the native enzyme in the range of 5 to 20 mg/ml was only slightly concentration-dependent (4.28 S to 4.86 S) but extensive aggregation occurred above 20 mg/ml.  相似文献   

2.
Ornithine carbamoyltransferase (OTC; subunit, 36,000 Da) [EC 2.1.3.3] is initially synthesized as a precursor (pOTC) with a transient NH2-terminal presequence of 32 amino acid residues, then is imported posttranslationally nto the mitochondrial matrix. We expressed rat pOTC in Escherichia coli, purified it in a denatured form, and showed that could be transported into isolated mitochondria in the presence of rabbit reticulocyte lysate [Murakami et al. (1988) J. Biol. Chem. 263, 18437-18442]. In order to compare the properties of the precursor and mature form of OTC, the rat mature OTC was synthesized in E. coli and purified. The recombinant OTC represented about 5% of the total bacterial protein and was present in both the supernatant and precipitate of the disrupted bacteria. The OTC, extracted from the precipitate with 8 M urea or 6 M guanidine.HCl, was essentially homogeneous, as judged by SDS-PAGE. When guanidine.HCl-denatured mature OTC was diluted and incubated at 0 degrees C for 40-60 h, it was reactivated to a specific activity of 170 mumol/min/mg protein at 37 degrees C (18% of that of the purified mature enzyme). Guanidine.HCl-denatured pOTC was activated to a specific activity of 125 mumol/min/mg protein under similar conditions. The native and reactivated OTC sedimented with an s20.w value of 6.2S, whereas the activated pOTC sedimented with an s20.w of 5.2S. The activated pOTC was more unstable than the reactivated OTC at 50 degrees C. These observations indicate that the presequence does not prevent pOTC from folding into an enzymatically active trimeric form, although the pOTC trimer appears to be less compact than the mature trimer.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Thermal unfolding of dodecameric manganese glutamine synthetase (622,000 M(r)) at pH 7 and approximately 0.02 ionic strength occurs in two observable steps: a small reversible transition (Tm approximately 42 degrees C; delta H approximately equal to 0.9 J/g) followed by a large irreversible transition (Tm approximately 81 degrees C; delta H approximately equal to 23.4 J/g) in which secondary structure is lost and soluble aggregates form. Secondary structure, hydrophobicity, and oligomeric structure of the equilibrium intermediate are the same as for the native protein, whereas some aromatic residues are more exposed. Urea (3 M) destabilizes the dodecamer (with a tertiary structure similar to that without urea at 55 degrees C) and inhibits aggregation accompanying unfolding at < or = 0.2 mg protein/mL. With increasing temperature (30-70 degrees C) or incubation times at 25 degrees C (5-35 h) in 3 M urea, only dodecamer and unfolded monomer are detected. In addition, the loss in enzyme secondary structure is pseudo-first-order (t1/2 = 1,030 s at 20.0 degrees C in 4.5 M urea). Differential scanning calorimetry of the enzyme in 3 M urea shows one endotherm (Tmax approximately 64 degrees C; delta H = 17 +/- 2 J/g). The enthalpy change for dissociation and unfolding agrees with that determined by urea titrations by isothermal calorimetry (delta H = 57 +/- 15 J/g; Zolkiewski M, Nosworthy NJ, Ginsburg A, 1995, Protein Sci 4: 1544-1552), after correcting for the binding of urea to protein sites exposed during unfolding (-42 J/g). Refolding and assembly to active enzyme occurs upon dilution of urea after thermal unfolding.  相似文献   

4.
We have examined the influence of urea on the properties of the rat liver glucocorticoid receptor (GR). A 1-h incubation of hepatic cytosol with 1-3 M urea at 0 or at 23 degrees C caused a progressive decrease in the steroid binding efficiency of GR. Urea treatment of cytosol incubated with 20 nM [3H]triamcinolone acetonide caused transformation of glucocorticoid-receptor complexes (GRc) and resulted in an increase in the binding of GRc to DNA-cellulose and ATP-Sepharose. The transforming effect was maximal with 2.5 M urea at 0 degrees C for 1 h, and it caused a shift in the rate of sedimentation of the 9 S untransformed GRc to a 4 S form, similar to that observed upon incubation of the cytosol GRc at 23 degrees C. This 9 to 4 S transformation could also be observed in the presence of Na2MoO4. The Stokes radii of the GRc eluted from a Bio-Gel-A-0.5m agarose column were determined to be 5.9 and 4.9 nm in the absence and presence of 2.5 M urea. The aqueous two-phase partitioning analysis revealed a significant change in surface properties of GR following urea treatment; the observed partition coefficient values (cpm upper phase/bottom phase) were 0.022, 0.208, and 0.60 for GRc, GRc + 23 degrees C, and GRc + 2.5 M urea, respectively. Furthermore, the urea treatment rendered the GRc less negatively charged, forcing their appearance in the flow-through fractions of a DEAE-Sephacel column. These results suggest that urea is a potent in vitro modulator of the physicochemical behavior of GR, influencing both the steroid binding and the process of receptor transformation.  相似文献   

5.
6.
Dihydrofolate reductase (DHFR) (5,6,7,8-THF: NADDP+ oxidoreductase, EC 1.5.1.3) was purified 205-fold to apparent homogeneity from the crude extracts of Lactobacillus leichmannii. It has UV absorption maxima at 280 nm, M(r) of 20,000, Stokes radius of 0.34 nm and a S20.w value of 0.12 S. The preparation showed the presence of 168 amino acid residues with threonine and lysine as the NH2- and COOH- terminal end-groups respectively and a single reactive sulfhydryl group. pCMB inhibited the enzyme activity (IC50 = 2 microM). The enzyme has a pH optimum of 7.4 and is thermally inactivated at > 35 degrees C. It is activated by 0.1 M KCl and KI and 2 M urea. 3-4 M urea completely inactivated the enzyme. Enzyme has Km values of 3.5 microM and 6.2 microM for NADPH and DHF respectively, and a Ki value of 7 nM for MTX, the inhibition being competitive.  相似文献   

7.
Effect of lower concentrations of urea on papain was monitored by optical spectroscopy, calorimetry and partial specific volume measurements. At lower concentrations of urea, papain exhibits a different structure and showed an increase in the intensity of circular dichroic (CD) spectra as compared to the native molecule. At lower concentrations (0.2-1.5 M) of urea, binding of 8-anilino-naphthalene sulfonic acid (ANS) to the papain molecule was higher; at 0.5 M, there was about 50% increase in ANS binding. Both calorimetric and spectroscopic studies indicated an increased thermal stability of the molecule at lower concentrations. At 0.5 M urea concentration, the apparent thermal denaturation temperature increased from a control value of 83 +/- 1 degrees C to 86 +/- 1 degrees C. At isopotential conditions, the partial specific volume of papain was found to be higher in presence of lower concentrations of urea, than the native protein or unfolded molecule. The preferential interaction parameter (deltag3/deltag2)(T,mu1,mu3) showed a negative value in the presence of lower concentrations of urea (0.2-2 M), which was maximum at 1 M urea with a value of -0.019 g/g. Above 3 M urea, the preferential interaction parameter was positive.  相似文献   

8.
The molecular weight of pyruvate carboxylase isolated from pigeon and rat liver mitochondria was examined using analytical ultracentrifugation and electron microscopy. The enzyme molecule appeared as a tetramer with the four subunits arranged at the corners of a square. Sedimentation studies in the analytical ultracentrifuge, extrapolated to infinite dilution, showed the tetramer to have a molecular weight Mc=0r of 280 000 and an So20,w of 12.7 S. The tetramer could be dissociated into trimers and dimers of lower specific enzymic activity by storage at 4 degrees C or incubation at -- 20 degrees C at low protein concentrations. The isolated trimers and dimers had a molecular weight Mc=0r of 210 000 and 140 000, respectively, and an So20,w of 10.85 S and 7.55 S, respectively. Incubation with 2 M urea at 20 degrees C yielded enzymically inactive subunits (Mc=0r = 70 000; So20,w = 4.95 S). The molecular weights (for pyruvate carboxylase and its subunits), as calculated from the subunit diameter observed in the electron microscope, were consistent with the values obtained from sedimentation studies.  相似文献   

9.
The major protein component of bovine high density lipoprotein was investigated in solution by fluorescence polarization and ultracentrifugal techniques. A fluorescent derivative of this protein with 1-dimethylaminonaphthalene-5-sulfonyl chloride was employed in the fluorescence experiments. Over the concentration range from 5-10(-7) M to 5-10(-4) M of the protein monomer at pH values from 2 to 11 and ionic strengths from 0.03 to 2.0, at 23 degrees C, the major protein of bovine high density lipoproteinapoprotein (Apo-HOL-I) was found to exist in a stable aggregated form. The aggregate was not affected by dioxane additions of up to 20% nor by Triton X-100 to 0.2%, but dissociated readily in the presence of 0.07% sodium dodecylsulfate or 6 M urea. At concentrations below 5-10(-7) M, dissociation of the protein aggregate started spontaneously and continued down to 10(-8) M, the lowest measurable concentration. Several physiocochemical properties of the major protein of bovine high density lipoprotein were determined in the stable aggregate form. Molecular weight was 104 000 from ultracentrifugal analysis and 80 000 from gel-filtration. Rotational relaxation time was 115 ns at 25 degrees C, and s-0 20,w was 4.78 s. The results suggest very strong protein-protein interactions (Kd less than 10(-7) M) that are not electrostatic in nature. Hydrophobic interactions of a magnitude that could be affected by 20% dioxane or 0.2% Triton X-100 detergent are also excluded. There is saturation of the interaction sites by the aggregation of a few protein monomer units possibly to form a tetramer which is moderately asymmetric (1:4 axial ratio, assuming an ellipsoid of revolution) and relatively rigid. The strong protein-protein interactions in this pure apolipoprotein suggest the possibility of competition of inter-protein associations with protein-lipid interactions in in vitro lipid binding or lipoprotein reconstitution experiments.  相似文献   

10.
For the study of DNA conformations, conformational transitions, and DNA-protein interactions, covalently closed supercoiled ColE1-plasmid DNA has been purified from cultures of Escherichia coli harboring this plasmid and grown in the presence of chloramphenicol according to the method of D.B. Clewell [J. Bact. 110 (1972)667]. The open circular and linear forms of the plasmid were prepared by digestion of the covalently closed, supercoiled form with pancreatic deoxyribonuclease and EcoRI-restriction endonuclease, respectively. The linear form was found to be very homogeneous by electron microscopy and sedimenting boundary analysis. Its physical properties (s0 20,w=16.3 S,D0 20,W=1.98 X 10(-8) cm2 s-1 and [eta]=2605 ml g-1) have been carefully determined in 0.2 M NaCl, 0.002 M NaPO4 pH 7.0,0.002 M EDTA, at 23 degrees C. Combination of s0 20, w (obtained by quasielastic laser light scattering) gave Ms,D=4.39 x 10(6). This value is in reasonable agreement with the molecular weight from total intensity laser light scattering M=4.30 x 10(6). The covalently closed and open circular forms of the ColE1-plasmid are less homogeneous due to slight cross-contamination and the presence of small amounts of dimers in these preparations. The weight fractions of the various components as determined by boundary analysis or electron microscopy are given together with the average quantities obtained in the same solvent for the supercoiled form ((s0 20,w)w=25.4 S, (D0 20,w)z=2.89 x 10(-8) cm2 s-1, [eta]= 788 ML G-1,Ms,D=4.69 x 10(6) and Mw=4.59 x 10(6)) and the open circular form (s0 20, w)w=20.1 S, (D0 20,w)z=2.45 x 10(-8) cm2 s-1, [eta]=1421 ml g-1,Ms,D=4.37 x 10(6) and Mw=4.15 x 10(6)). Midpoint analysis of the sedimenting boundaries allows unambiguous determination of the sedimentation coefficients of these two forms: s0 20,w=24.5 S and s0 20,w=18.8 S, respectively. Also deduced from total intensity light scattering were radii of gyration Rg (103.5, 134.2 and 186 nm) and second virial coefficients A2 (0.7, 4.8 AND 5.4 x 10(-4) mole ml/g2) for the supercoiled, the open circular and linear forms, respectively. The data presented are discussed in relation to the conformational parameters for the three forms in solution.  相似文献   

11.
Isolated, intact rat liver nuclei have high-affinity (Kd = 10(-9) M) binding sites that are highly specific for nonsteroidal antiestrogens, especially for compounds of the triphenylethylene series. Nuclear [3H]tamoxifen binding capacity is thermolabile, being most stable at 4 degrees C and rapidly lost at 37 degrees C. More [3H]tamoxifen, however, is specifically bound at incubation temperatures of 25 degrees C and 37 degrees C than at 4 degrees C although prewarming nuclei has no effect, suggesting exchange of [3H]tamoxifen for an unidentified endogeneous ligand. Nuclear antiestrogen binding sites are destroyed by trypsin but not by deoxyribonuclease I or ribonuclease A. The nuclear antiestrogen binding protein is not solubilized by 0.6 M potassium chloride, 2 M sodium chloride, 0.6 M sodium thiocyanate, 3 M urea, 20 mM pyridoxal phosphate, 1% (w/v) digitonin or 2% (w/v) sodium cholate but is extractable by sonication, indicating that it is tightly bound within the nucleus. Rat liver nuclear matrix contains high-affinity (Kd = 10(-9) M) [3H]tamoxifen binding sites present in 5-fold higher concentrations (4.18 pmol/mg DNA) than in intact nuclei (0.78 +/- 0.10 (S.D.) pmol/mg DNA). Low-speed rat liver cytosol (20 000 X g, 30 min) contains high-capacity (955 +/- 405 (S.D.) fmol/mg protein), low-affinity (Kd = 10.9 +/- 4.5 (S.D.) nM) antiestrogen binding sites. In contrast, high-speed cytosol (100 000 X g, 60 min) contains low-capacity (46 +/- 15 (S.D.) fmol/mg protein), high-affinity (Kd = 0.61 +/- 0.20 (S.D.) nM) binding sites. Low-affinity cytosolic sites constitute more than 90% of total liver binding sites, high-affinity cytosolic sites 0.3%-3.2%, and nuclear sites less than 0.5% of total sites.  相似文献   

12.
The effects of metal ions on the thermal denaturation and Mg2+ binding of native spinach ferredoxin and its acetylated derivative were investigated. The denaturation of ferredoxin in a metal-free solution at 40 degrees C was quickly prevented by the addition of Mg2+ or Na+ at appropriate concentrations. The metal concentrations required for 50% protection from thermal denaturation were 1.54 x 10(-4) M Mg2+ or 8.0 x 10(-3) M Na+ for native ferredoxin and 1.05 x 10(-3) M Mg2+ or 6.0 x 10(-2) M Na+ for acetylated ferredoxin. It was also found that native ferredoxin in the presence of over 20 mM Mg2+ was almost completely protected from thermal denaturation at 40 degrees C. The D-form which has been observed in acetylated ferredoxin by Masaki et al. (1977) (J. Biochem. 81, 1-9) was confirmed to be present in native ferredoxin at high temperature (49 degrees C) and is suggested to be an important form in the denaturation processes of the ferredoxin molecule.  相似文献   

13.
A monoclonal IgG 2a antibody directed against the activated rat liver glucocorticoid receptor (GR) was used to prepare an immunoaffinity matrix of high capacity. The molybdate-stabilized GR from rat liver cytosol was immunoadsorbed on this gel. A non-hormone-binding protein of Mr approximately 90,000, as determined after denaturing gel electrophoresis, was eluted from this matrix following removal of molybdate and exposure to heat (25 degrees C) and salt (0.15 M NaCl). Subsequently, the Mr approximately 90,000 protein was purified to homogeneity using high-performance ion-exchange chromatography, covalently radiolabelled, and analyzed by high-performance size-exclusion chromatography and sucrose gradient ultracentrifugation. Hydrodynamic characterization indicates that, under our experimental conditions, the molybdate-stabilized rat liver GR (Rs approximately 7.4 nm, s20,w approximately 9.1 S, calculated mol. wt Mr approximately 285,000) includes one steroid-binding unit (Rs approximately 5.5 nm, S20,w approximately 4.3 S, calculated Mr approximately 100,000) and a dimer of Mr approximately 90,000 non-hormone-binding protein (Rs approximately 6.9 nm, S20,w approximately 6.1 S, calculated native Mr approximately 180,000).  相似文献   

14.
Recently, protease 2A of human rhinovirus 2 (HRV2 2A) was shown to require a zinc ion for the formation of an active enzyme although zinc is not involved mechanistically. The data presented clearly show that the zinc ion bound to a picornaviral-specific motif represents an essential component of the native structure, probably representing a new Zn-binding motif. This structure, containing mostly beta-strand elements as shown by CD spectroscopy, changes drastically upon removal of zinc. The zinc-depleted form does represent an intermediate with mostly unchanged secondary structure, but not a fully denatured random coil as obtained by guanidinium hydrochloride. This is indicated by the blue-shifted fluorescence spectra and by CD. The native protein exhibited a cooperative phase transition at 53 degrees C. In contrast, the zinc-depleted form did not show any transition at all, again demonstrating the stabilizing role of the zinc ion. A structural intermediate was observed during thermal and pH denaturation that may represent a molten globule, as suggested by its ANS binding.  相似文献   

15.
Conformational changes of apo A-1, the principal apoprotein of human plasma high density lipoprotein, have been studied by differential scanning calorimetry and ultraviolet difference spectroscopy as a function of temperature, pH, concentration of apoprotein, and urea concentration. Calorimetry shows that apo A-1 (5 to 40 mg/ml, pH 9.2) undergoes a two-state, reversible denaturation (enthalpy = 64 +/- 8.9 kcal/mole), between 43--71 degrees (midpoint temperature, Tm = 54 degrees), associated with a rise in heat capacity (deltaCvd) of 2.4 +/- 0.5 kcal/mole/degrees C. Apo A-1 (0.2 to 0.4 mg/ml, pH 9.2) develops a negative difference spectrum between 42--70 degrees, with Tm = 53 degrees. The enthalpy (deltaH = 59 +/- 5.7 kcal/mole at Tm) and heat capacity change (2.7 +/- 0.9 kcal/mole/degrees C) in the spectroscopic experiments were not significantly different from the calorimetric values. Below pH 9 and above pH 11, the calorimetric Tm and deltaH of denaturation are decreased. In the pH range of reversible denaturation (6.5 to 11.8), delatH and Tm are linearly related, showing that the heat capacity change (ddeltaH/dT) associated with denaturation is independent of Tm. In urea solutions, the calorimetric Tm and deltaH of denaturation are decreased. At 25 degrees, apo A-1 develops a negative difference spectrum between 1.4 and 3 M urea. Fifty per cent of the spectral change occurs in 2.4 M urea, which corresponds to the urea concentration obtained by extrapolation of the calorimetric Tm to 25 degrees. In urea solution of less than 0.75 M there is hyperchromicity at 285 nm (delta epsilon = 264 in 0.75 M urea), indicating strong interaction of aromatic amino acid residues in the native molecule with the solvent. Spectrophotometric titration of apo A-1 shows that 6.6 of the 7 tyrosine groups of apo A-1 titrate at pH less than 11.9, with similar titration curves obtained in aqueous solutions and in 6 M urea. The free energy of stabilization (deltaG) of the native conformation of apo A-1 was estimated, (a) at 37 degrees, using the calorimetric deltaA and deltaCvd, and (b) at 25 degrees, by extrapolation of spectroscopic data to zero urea concentration. The values (deltaG (37 degrees) = 2.4 and deltaG (25 degrees) = 2.7 kcal/mole) are small compared to typical globular proteins, indicating that native apo A-1 has a loosely folded tertiary structure. The low values of deltaG reflect the high degree of exposure of hydrophobic areas in the native protein molecule. The loosely folded conformation of apo A-1 allows extensive binding of lipid, since this can involve both surface hydrophobic sites and hydrophobic areas exposed by a cooperative, low energy unfolding process.  相似文献   

16.
RNase Rs showed an approx. 2-fold increase in its activity when incubated in the presence of 2 M urea at 37 degrees C. The increase in its activity, in the presence of urea, was comparable to the activity at its optimum temperature, i.e. 45 degrees C. Compared to the native enzyme at 37 degrees C, the K(m) and V(max) of RNase Rs at 45 degrees C and in the presence of 2 M urea at 37 degrees C showed an increase while k(cat)/K(m) decreased. Arrhenius plots in the presence and absence of urea showed a decrease in the activation energy in the presence of urea. Though there was no change in the secondary structure of the protein in the presence of urea, minor changes were observed in the tertiary structure. Hence, the increase in the activity of RNase Rs, in the presence of 2 M urea at 37 degrees C, is due to the lowering of the activation energy as a result of changes in the microenvironment of the active site.  相似文献   

17.
Huang Q  Zhang L 《Biopolymers》2005,79(1):28-38
From Poria cocos mycelia yielded via a pilot scale facility-fermentation tank, a water-insoluble (1-->3)-alpha-D-glucan coded as Pi-PCM3-I was isolated by extraction with 0.5 M NaOH/0.01 M NaBH(4) aqueous solution. Nine fractions from F1 to F9 with a weight-average molecular mass (M(w)) range from 7.75 x 10(4) to 57.3 x 10(4) were prepared from the Pi-PCM3-I sample by a nonsolvent addition method. The fractions were reacted with chlorosulfonic acid-pyridine complex to product water-soluble sulfated derivatives coded as S1 to S8 with M(w) from 2.36 x 10(4) to 14.5 x 10(4) and degree of substitution (DS) of 0.86-1.38. M(w), z-average radius of gyration (s(2) (z) (1/2)), the second virial coefficient (A(2)), and the intrinsic viscosity ([eta]) of the native and sulfated Pi-PCM3-I were measured by laser light scattering (LLS), size-exclusion chromatography combined with LLS (SEC-LLS), and viscometry at 25 degrees C. The Mark-Houwink equations for Pi-PCM3-I in 0.25 M LiCl/dimethylsulfoxide (DMSO) (Me(2)SO) and for its sulfated derivative in 0.15 M NaCl aqueous solution at 25 degrees C were established to be [eta] = 1.33 x 10(-2) M(w) (0.75+/-0.01) (mL g(-1)) and [eta] = 1.46 x 10(-4) M(w) (1.13+/-0.01) (mL g(-1)), respectively. On the basis of theories for a wormlike cylinder model, the conformational parameters of the native and sulfated Pi-PCM3-I were calculated to be 760 +/- 50 and 1060 +/- 30 nm(-1) for the molar mass per unit contour length (M(L)), 6.3 +/- 0.5 and 13.1 +/- 1 nm for the persistence length (q), and 14.9 +/- 0.2 and 31.8 +/- 1 for the characteristic ratio (C( proportional, variant)), respectively. The results revealed that Pi-PCM3-I existed as an extended flexible chain in 0.25 M LiCl/Me(2)SO, and its sulfated derivative existed as a semistiff chain in 0.15 M NaCl aqueous solution. Furthermore, Pi-PCM3-I possessed similar structure and molecular parameters to wc-PCM3-I from a rotary shaker; this suggests promising industrialization of Poria cocos polysaccharides.  相似文献   

18.
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   

19.
The secondary and tertiary structures of Xenopus oocyte and somatic 5S rRNAs were investigated using chemical and enzymatic probes. The accessibility of both RNAs towards single-strand specific nucleases (T1, T2, A and S1) and a helix-specific ribonuclease from cobra venom (RNase V1) was determined. The reactivity of nucleobase N7, N3 and N1 positions towards chemical probes was investigated under native (5 mM MgCl2, 100 mM KCl, 20 degrees C) and semi-denaturing (1 mM EDTA, 20 degrees C) conditions. Ethylnitrosourea was used to identify phosphates not reactive towards alkylation under native conditions. The results obtained confirm the presence of the five helical stems predicted by the consensus secondary structure model of 5S rRNA. The chemical reactivity data indicate that loops C and D are involved in a number of tertiary interactions, and loop E folds into an unusual secondary structure. A comparison of the data obtained for the two types of Xenopus 5S rRNA indicates that the conformations of the oocyte and somatic 5S rRNAs are very similar. However, the data obtained with nucleases under native conditions, and chemical probes under semi-denaturing conditions, reveal that helices III and IV in the somatic 5S rRNA are less stable than the same structures in oocyte 5S rRNA. Using chimeric 5S rRNAs, it was possible to demonstrate that the relative resistance of oocyte 5S rRNA to partial denaturation in 4 M urea is conferred by the five oocyte-specific nucleotide substitutions in loop B/helix III. In contrast, the superior stability of oocyte 5S rRNA in the presence of EDTA is related to a single C substitution at position 79.  相似文献   

20.
A majority of the untransformed glucocorticoid-receptor complexes (GRc) from rat liver cytosol sedimented in the 9S region in 5-20% sucrose gradients containing 0.15 M KCl and 20 mM Na2MoO4. Incubation of the cytosol at 23 degrees C, or at 0 degree C with 10 mM ATP or 0.3 M KCl caused appearance of a slower migrating (4S) form which exhibited an increased affinity toward DNA-cellulose and ATP-Sepharose. Presence of 20 mM Na2MoO4 blocked this 9S to 4S transformation of GRc. A complete conversion of the 9S to the 4S form occurred upon a 2 h incubation of GRc with 10 mM ATP at 0 degree C. Other nucleoside triphosphates (GTP, CTP, and UTP), ADP and PPi (but not AMP or cAMP) were also effective in transforming the 9S form. The heat transformation occurred in a time-dependent manner and was complete within 1 h at 23 degrees C; presence of 10 mM ATP during this 23 degrees C incubation period allowed a complete 9S to 4S alteration in 10-20 min. Addition of ATP also accelerated the rate of salt activation of the GRc; a 50% conversion to the 4S form occurred in 20 min or 3 min in the absence or the presence of 10 mM ATP during the 0 degree C incubation of GRc with 0.15 M KCl. An absolute requirement of the hormone for 9S to 4S transformation of glucocorticoid receptor (GR) was evident, as no conversion of the 9S form to the 4S form could be achieved with the ligand-free GR under any of the above conditions. Incubation of cytosol preparations at 23 degrees C or at 0 degree C with KCl or ATP caused dissociation of the GRc and reduced the steroid binding capacity of GR. Although aurintricarboxylic acid, pyridoxal 5'-phosphate, Na2MoO4, Na2WO4, o-phenanthroline, Rifamycin AF/013 and heparin inhibited the ATP-Sepharose and DNA binding of the GRc, only Na2MoO4 and Na2WO4 selectively blocked the 9S to 4S conversion. We suggest that the 9S to 4S transformation in vitro of rat liver GRc represents an acquisition of DNA and ATP-Sepharose binding ability and may involve a separation of subunits from an oligomeric receptor structure.  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