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.
1. A unique caseinolytic activity was found in the crude extract from chicken and rat skeletal muscle. Hardly any activity was detected at physiological assay temperatures at pH 8.0 but did well at around 60 degrees C. 2. The activity partially purified from rat skeletal muscle showed optimum pH at around 8.0 at 60 degrees C. It hardly hydrolyzed casein below 50 degrees C, but in the presence of 5 M urea it showed relatively high activity at 30 degrees C. The activity was completely stable at 50 degrees C for 1 hr. 3. The activity seems to be contained in a high mol. wt (450,000) protein from the elution volume and is due to cysteine proteinase from the effect of inhibitors. 4. The above properties agreed with those of the heat-stable alkaline proteinase (HAP) of fish purified homogeneously by electrophoresis. This seems to suggest that HAP may also exist in rat skeletal muscle.  相似文献   

2.
K Lohner  A F Esser 《Biochemistry》1991,30(26):6620-6625
The thermotropic behavior of purified human complement protein C9 was investigated by high-sensitivity differential scanning calorimetry. When dissolved in physiological buffers (pH 7.2, 150 mM NaCl), C9 underwent three endothermic transitions with transition temperatures (Tm) centered at about 32, 48, and 53 degrees C, respectively, and one exothermic transition above 64 degrees C that correlated with protein aggregation. The associated calorimetric enthalpies of the three endothermic transitions were about 45, 60, and 161 kcal/mol with cooperative ratios (delta Hcal/delta HvH) close to unity. The total calorimetric enthalphy for the unfolding process was in the range of 260-280 kcal/mol under all conditions. The exothermic aggregation temperature was strongly pH dependent, changing from 60 degrees C at pH 6.6 to 81.4 degrees C at pH 8.0, whereas none of the three endothermic transitions was significantly affected by pH changes. They were, however, sensitive to addition of calcium ions; most affected was Tm1 which shifted from 32 to 35.8 degrees C in the presence of 3 mM calcium, i.e., the normal blood concentration. Kosmotropic ions stabilized the protein by shifting the endothermic transitions to slightly higher temperatures whereas inclusion of chaotropic ions (such as choline), removal of bound calcium by addition of EDTA, or proteolysis with thrombin lowered the transition temperatures. Previous studies had indicated the formation of at least three different forms of C9 during membrane insertion or during heat polymerization, and it is suggested that the three endothermic transitions reflect the formation of such C9 conformers. Choline, which is present at high concentrations on the surface of biological membranes, and calcium ions have the ability to shift the transition temperatures of the first two transitions to be either close to or below body temperature. Thus, it is very likely that C9 is present in vivo in a partially unfolded state when bound to a membrane surface, and we propose that this facilitates membrane insertion and refolding of the protein into an amphiphilic conformation.  相似文献   

3.
The critical role played by temperature in ligand-induced protein aggregation was investigated. Recombinant human interleukin-1 receptor antagonist (rhIL-1ra) and the ligands benzyl alcohol and 8-anilinonaphthalene-1-sulfonate (ANS) were used. We investigated aggregation kinetics and the conformation and cysteine reactivity of rhIL-1ra in buffer alone or in the presence of 0.9% (w/v) benzyl alcohol or 4.2 or 21 mM ANS at 25 and 37 degrees C. In buffer, protein aggregation was not detected at 25 degrees C but occurred at 37 degrees C. At 25 degrees C, neither benzyl alcohol nor 4.2 mM ANS enhanced aggregation. However, at 37 degrees C, both compounds greatly accelerated protein aggregation. With 21 mM ANS, rhIL-1ra aggregation was accelerated at both temperatures, but the effect was more pronounced at 37 degrees C than at 25 degrees C. Increasing the temperature from 25 to 37 degrees C caused a minor perturbation in the tertiary structure of rhIL-1ra in buffer but no detectable alteration in secondary structure. Benzyl alcohol enhanced the tertiary structural perturbation at 37 degrees C, but the secondary structure was not affected by the ligand. The reactivity of buried free cysteines of rhIL-1ra was enhanced by benzyl alcohol at 37 degrees C but not at 25 degrees C, consistent with the structural results. Isothermal titration calorimetry documented that the interaction of benzyl alcohol with rhIL-1ra was hydrophobic and that the degree of hydrophobic interactions increased with temperature. At 25 degrees C, the interaction of ANS with rhIL-1ra was electrostatic, but at 37 degrees C, both electrostatic and hydrophobic interactions were important. Taken together, our results support the conclusion that benzyl alcohol and ANS interact hydrophobically with partially unfolded aggregation-prone protein molecules, resulting in temperature-dependent increases in their levels and acceleration of protein aggregation.  相似文献   

4.
The temperature-induced structural transitions of the full length Alzheimer amyloid beta-peptide [A(beta)(1-40) peptide] and fragments of it were studied using CD and 1H NMR spectroscopy. The full length peptide undergoes an overall transition from a state with a prominent population of left-handed 3(1) (polyproline II; PII)-helix at 0 degrees C to a random coil state at 60 degrees C, with an average DeltaH of 6.8 +/- 1.4 kJ.mol(-1) per residue, obtained by fitting a Zimm-Bragg model to the CD data. The transition is noncooperative for the shortest N-terminal fragment A(beta)(1-9) and weakly cooperative for A(beta)(1-40) and the longer fragments. By analysing the temperature-dependent 3J(HNH(alpha)) couplings and hydrodynamic radii obtained by NMR for A(beta)(1-9) and A(beta)(12-28), we found that the structure transition includes more than two states. The N-terminal hydrophilic A(beta)(1-9) populates PII-like conformations at 0 degrees C, then when the temperature increases, conformations with dihedral angles moving towards beta-strand at 20 degrees C, and approaches random coil at 60 degrees C. The residues in the central hydrophobic (18-28) segment show varying behaviour, but there is a significant contribution of beta-strand-like conformations at all temperatures below 20 degrees C. The C-terminal (29-40) segment was not studied by NMR, but from CD difference spectra we concluded that it is mainly in a random coil conformation at all studied temperatures. These results on structural preferences and transitions of the segments in the monomeric form of A(beta) may be related to the processes leading to the aggregation and formation of fibrils in the Alzheimer plaques.  相似文献   

5.
Smeller L  Rubens P  Heremans K 《Biochemistry》1999,38(12):3816-3820
This work demonstrates that pressure-induced partially unfolded states play a very important role in the aggregation of proteins. The high-pressure unfolding of horse heart metmyoglobin results in an intermediate form that shows a strong tendency to aggregate after pressure release. These aggregates are similar to those that are usually observed upon temperature denaturation. Infrared spectra in the amide I region indicate the formation of an intermolecular antiparallel beta-sheet stabilized by hydrogen bonding. The formation of the aggregates is temperature-dependent. Below 30 degrees C, no aggregation is taking place as seen from the infrared spectra. At 45 and 60 degrees C, two types of aggregates are formed: one that can be dissociated by moderate pressures and one that is pressure-insensitive. When precompressed at 5 degrees C, temperature-induced aggregation takes place at lower temperature (38 degrees C) than without pressure pretreatment (74 degrees C).  相似文献   

6.
The effect of temperature on the kinetic parameters of phosphoenolpyruvate carboxylase purified from Crassula argentea was such that both the Vmax and Km(MgPEP) values tended upward over the range from 11 to 35 degrees C. The increased rate at low temperatures due to the low Km is at least partially offset by the increased Vmax at higher temperatures, potentially leading to a broad plateau of enzyme activity and a relatively small effect of temperature on the enzyme. The cooperativity was negative at 11 degrees C, but above 15 degrees C it became positive. The presence of 5 mM glucose-6-phosphate has relatively little effect on Vmax but it clearly reduces Km and overcomes any effect of temperature on this parameter in the range studied. Positive cooperativity is observed only at temperatures above 25 degrees C. The size of the native enzyme, as determined by dynamic light scattering, was strongly toward the tetrameric form. At a temperature of 40 degrees C and above, a considerable oligomerization takes place. No loss of activity can be observed in this range of temperature. In the presence of either glucose-6-phosphate or magnesium phosphoenolpyruvate, at temperatures under 25 degrees C, the equilibrium is displaced toward higher levels of aggregation. Maximal accumulation of lead malate occurred at 10 to 12 degrees C in vivo with reduction to about 25% at 35 degrees C. Glucose-6-phosphate followed a similar curve in response to temperature, but the overall difference was about 50%. The sum of phosphoenolpyruvate plus pyruvate is level at night temperatures below 25 degrees C, doubling at 35 degrees C. Calculated concentrations of malate, glucose-6-phosphate, and phosphoenolpyruvate plus pyruvate indicate that the concentrations present are equal to or greater than Ki, Ka, and Km values for these metabolites, respectively.  相似文献   

7.
Thermal stability of Escherichia coli Fpg protein was studied using far-UV circular dichroism and intrinsic fluorescence. Experimental data indicate that Fpg irreversibly aggregates under heating above 35 degrees C. Heat aggregation is preceded by tertiary conformational changes of Fpg. However, the secondary structure of the fraction that does not aggregate remains unchanged up to approximately 60 degrees C. The kinetics of heat aggregation occurs with an activation enthalpy of approximately 21 kcal/mol. The fraction of monomers forming aggregates decreases with increasing urea concentration, with essentially no aggregation observed above approximately 3 M urea, suggesting that heat aggregation results from hydrophobic association of partially unfolded proteins. With increasing urea concentration, Fpg unfolds in a two-state reversible transition, with a stability of approximately 3.6 kcal/mol at 25 degrees C. An excellent correlation is observed between the unfolded fraction and loss of activity of Fpg. A simple kinetic scheme that describes both the rates and the extent of aggregation at each temperature is presented.  相似文献   

8.
With decreasing temperature the reactivation yield of denatured D-glyceraldehyde 3-phosphate dehydrogenase (GAPDH) upon dilution increases but the reactivation rate decreases. Neither reactivation nor aggregation during refolding can be detected at 4 degrees C in 48 h, and at 3 degrees C even in 6 days. However, the reactivation takes place once the temperature is raised with little decrease of the yield after incubation for 6 days at 3 degrees C. A cold folding intermediate forms in a burst phase of refolding at 4 degrees C as shown by a fast change of the intrinsic fluorescence followed by further conformational adjustment to a stable state in about 1 h. The stable folding intermediate has been characterized to be a dimer of partially folded GAPDH subunit with secondary structure between that of the native and denatured enzymes, a hydrophobic cluster not found in either the native or the denatured state, and an active site similar to but different from that of the native state. Chaperonin 60 (GroEL) binds with all intermediates formed at 4 degrees C, but the intermediates formed at the early folding stage reactivate with higher yield than those formed after conformational adjustment when dissociated from GroEL in the presence of ATP and further folded and assembled into the native tetramer.  相似文献   

9.
tsJT60, a temperature-sensitive (ts) mutant cell line of Fischer rat, is viable at both permissive (34 degrees C) and non-permissive (39.5 degrees C) temperatures. The cells grow normally in exponential growth phase at both temperatures, but when stimulated with fetal bovine serum (FBS) from G0 phase they re-enter S phase at 34 degrees C but not at 39.5 degrees. When tsJT60 cells were transformed with adenovirus (Ad) 5 wild type, they grew well at both temperatures, expressed E1A and E1B genes, and formed colonies in soft agar. When tsJT60 cells were transformed with Ad5 dl313, that lacks E1B gene, the transformed cells grew well at 34 degrees C but failed to form colony in soft agar. They died very soon at 39.5 degrees C. 3Y1 cells (a parental line of tsJT60) transformed with dl313 grew well at both temperatures, although neither expressed E1B gene nor formed colonies in soft agar. The phenotype of being lethal at 39.5 degrees C of dl313-transformed tsJT60 cells was complemented by cell fusion with 3Y1BUr cells (5-BrdU-resistant 3Y1), but not with tsJT60TGr cells (6-thioguanine resistant tsJT60). These results indicate that the lethal phenotype is related to the ts mutation of tsJT60 cells and also to the deletion of E1B gene of Ad5.  相似文献   

10.
G Panick  R Malessa  R Winter 《Biochemistry》1999,38(20):6512-6519
We examined the temperature- and pressure-induced unfolding and aggregation of beta-lactoglobulin (beta-Lg) and its genetic variants A and B up to temperatures of 90 degrees C in the pressure range from 1 bar to 10 kbar. To achieve information simultaneously on the secondary, tertiary, and quaternary structures, we have applied Synchrotron small-angle X-ray diffraction and Fourier transform infrared spectroscopy. Upon heating a beta-Lg solution at pH 7.0, the radius of gyration Rg first decreases, indicating a partial dissociation of the dimer into the monomers, the secondary structures remaining essentially unchanged. Above 50 degrees C, the infrared spectroscopy data reveal a decrease in intramolecular beta-sheet and alpha-helical structures, whereas the contribution of disordered structures increases. Within the temperature range from 50 to 60 degrees C, the appearance of the pair distance distribution function is not altered significantly, whereas the amount of defined secondary structures declines approximately by 10%. Above 60 degrees C the aggregation process of 1% beta-Lg solutions is clearly detectable by the increase in Rg and intermolecular beta-sheet content. The irreversible aggregation is due to intermolecular S-H/S-S interchange reactions and hydrophobic interactions. Upon pressurization at room temperature, the equilibrium between monomers and dimers is also shifted and dissociation of dimers is induced. At pressures of approximately 1300 bar, the amount of beta-sheet and alpha-helical structures decreases and the content of disordered structures increases, indicating the beginning unfolding of the protein which enables aggregation. Contrary to the thermal denaturation process, intermolecular beta-sheet formation is of less importance in pressure-induced protein aggregation and gelation. The spatial extent of the resulting protein clusters is time- and concentration-dependent. The aggregation of a 1% (w/w) solution of A, B, and the mixture AB results in the formation of at least octameric units as can be deduced from the radius of gyration of about 36 A. No differences in the pressure stability of the different genetic variants of beta-Lg are detectable in our FT-IR and SAXS experiments. Even application of higher pressures (up to 10 kbar) does not result in complete unfolding of all beta-Lg variants.  相似文献   

11.
Production of laccase activity by nine strains of Chalara paradoxa isolated from olive mill wastewater disposal ponds were studied. Enzyme extracts obtained from cultured broths by adsorption on hydroxyapatite showed a single band of laccase activity on ABTS after polyacrylamide gel electrophoresis (PAGE). They showed small mobility differences, with molecular masses of 67 to 68 kDa. Enzymes from the different strains oxidized a variety of phenolic and non-phenolic substances, and they could be divided into two groups according to their relative activities on substrates. All laccases showed a dual pH dependence of activity, with a maximum in the range of pH 3.0 to 4.5 for ABTS, o-dianisidine and 2,6-dimethoxyphenol, and pH 6.0 (Group 1) or pH 6.5 (Group 2) for syringaldazine and other substrates. Optimal temperatures were in the range of 10 to 28 degrees C for two strains (maximum at 10 degrees C) and 10 to 37 degrees C for the rest. The different enzymes were partially inactivated by heating at 60 degrees C and totally inactivated at 70 degrees C. Laccases were stable in a pH range of 3.0 to 9.0 (except for strain 36A, which was partially inactivated at pH 3.0), but became inactivated at pH 2.0. Altogether these data suggest that Ch. paradoxa strains produce different laccase isoenzymes.  相似文献   

12.
Heat stress inhibits photosynthesis by reducing the activation of Rubisco by Rubisco activase. To determine if loss of activase function is caused by protein denaturation, the thermal stability of activase was examined in vitro and in vivo and compared with the stabilities of two other soluble chloroplast proteins. Isolated activase exhibited a temperature optimum for ATP hydrolysis of 44 degrees C compared with > or =60 degrees C for carboxylation by Rubisco. Light scattering showed that unfolding/aggregation occurred at 45 degrees C and 37 degrees C for activase in the presence and absence of ATPgammaS, respectively, and at 65 degrees C for Rubisco. Addition of chemically denatured rhodanese to heat-treated activase trapped partially folded activase in an insoluble complex at treatment temperatures that were similar to those that caused increased light scattering and loss of activity. To examine thermal stability in vivo, heat-treated tobacco (Nicotiana rustica cv Pulmila) protoplasts and chloroplasts were lysed with detergent in the presence of rhodanese and the amount of target protein that aggregated was determined by immunoblotting. The results of these experiments showed that thermal denaturation of activase in vivo occurred at temperatures similar to those that denatured isolated activase and far below those required to denature Rubisco or phosphoribulokinase. Edman degradation analysis of aggregated proteins from tobacco and pea (Pisum sativum cv "Little Marvel") chloroplasts showed that activase was the major protein that denatured in response to heat stress. Thus, loss of activase activity during heat stress is caused by an exceptional sensitivity of the protein to thermal denaturation and is responsible, in part, for deactivation of Rubisco.  相似文献   

13.
Subcellular locations and chaperone functions of Hsp60 and Hsp70 with flagellin were investigated in Borrelia burgdorferi. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis and Western blot (immunoblot) analysis of fractionated cells showed Hsp60 to be present in the soluble fractions and the Triton X-100 detergent-soluble membrane fraction at growth temperatures ranging from 20 to 37 degrees C. The relative amount of Hsp60 associated with the membrane increased with growth temperature. Hsp70 was found in soluble fractions at growth temperatures between 28 and 37 degrees C, but at 20 degrees C it was also present in the Triton X-100-insoluble membrane fraction. Immunoelectron microscopy revealed that the majority of Hsp60 was localized in the cytoplasm but a detectable fraction (approximately 30%) was associated with the cell envelope. The chaperone functions of Hsp60 and Hsp70 were analyzed by immunoprecipitation of [35S]methionine-labeled cell lysates under nondenaturing conditions in the presence or absence of ATP. Hsp70 was found to bind flagellin at all temperatures tested between 33 and 41 degrees C. This association could be decreased with ATP when cells had been incubated at 41 degrees C during radioactive labeling but not at lower temperatures. Both flagellin and Hsp70 were found to associate with Hsp60, forming a complex of the three proteins. Hsp70 association with this complex could be decreased with ATP, but flagellin binding to Hsp60 was ATP independent at all temperatures studied. Both Hsp70 and flagellin were inaccessible to monoclonal antibodies against them when bound to Hsp60. These studies suggest that in B. burgdorferi, a major function of Hsp60 and Hsp70 is in the molecular processing of flagellin.  相似文献   

14.
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   

15.
The range of water temperatures in which Perkinsus species can survive and proliferate remains ill-defined, particularly at lower temperatures. The in vitro viability, metabolic activity, and proliferation of 3 isolates each of P. marinus and P. olseni trophozoites at 28 degrees C, and at 15 and 4 degrees C, after transfer from 28 degrees C, were compared. Both species showed declines in metabolic activity and proliferation from 28 degrees C to 15 degrees C. At 4 degrees C, both species had viability after 30 days incubation time (P. marinus 49%, P. olseni 58%), but limited metabolic activity and no proliferation. Perkinsus marinus viability was further compared when transferred directly from 28 degrees C, 18 degrees C and progressively from 18 degrees C (0.5 degrees C/day) to 2, 4 and 6 degrees C and maintained for up to 4 months. Viability was highest under progressive transfer (77% and 54% after 30 and 60 days exposure to test temperatures). The decrease in P. marinus viability at the lower temperatures in vitro only partially explains decreasing parasite infection intensities in eastern oysters in the colder months of the year. Moreover, the significant decrease in parasite infection intensities in late winter and early spring, as temperatures increase, is likely due to an active process of elimination by oyster host defences.  相似文献   

16.
The short-term effects of temperature on methanogenesis from acetate or CO(2) in a thermophilic (58 degrees C) anaerobic digestor were studied by incubating digestor sludge at different temperatures with C-labeled methane precursors (CH(3)COO or CO(2)). During a period when Methanosarcina sp. was numerous in the sludge, methanogenesis from acetate was optimal at 55 to 60 degrees C and was completely inhibited at 65 degrees C. A Methanosarcina culture isolated from the digestor grew optimally on acetate at 55 to 58 degrees C and did not grow or produce methane at 65 degrees C. An accidental shift of digestor temperature from 58 to 64 degrees C during this period caused a sharp decrease in gas production and a large increase in acetate concentration within 24 h, indicating that the aceticlastic methanogens in the digestor were the population most susceptible to this temperature increase. During a later period when Methanothrix sp. was numerous in the digestor, methanogenesis from CH(3)COO was optimal at 65 degrees C and completely inhibited at 75 degrees C. A partially purified Methanothrix enrichment culture derived from the digestor had a maximum growth temperature near 70 degrees C. Methanogenesis from CO(2) in the sludge was optimal at 65 degrees C and still proceeded at 75 degrees C. A CO(2)-reducing Methanobacterium sp. isolated from the digestor was capable of methanogenesis at 75 degrees C. During the period when Methanothix sp. was apparently dominant, sludge incubated for 24 h at 65 degrees C produced more methane than sludge incubated at 60 degrees C, and no acetate accumulated at 65 degrees C. Methanogenesis was severely inhibited in sludge incubated at 70 degrees C, but since neither acetate nor H(2) accumulated, production of these methanogenic substrates by fermentative bacteria was probably the most temperature-sensitive process. Thus, there was a correlation between digestor performance at different temperatures and responses to temperature by cultures of methanogens believed to play important roles in the digestor.  相似文献   

17.
alpha-Crystallin, a major lens protein of approximately 800 kDa with subunits of approximately 20 kDa has previously been shown to act as a chaperone protecting other proteins from stress-induced aggregation. Here it is demonstrated that alpha-crystallin can bind to partially denatured enzymes at 42-43 degrees C and prevent their irreversible aggregation, but cannot prevent loss of enzyme activity. However, the alpha-crystallin-bound enzymes regain activity on interaction with other chaperones. The data indicate that the re-activated enzymes are no longer associated with the alpha-crystallin, and ATP is required for re-activation. When inactive luciferase bound to alpha-crystallin was treated with reticulocyte lysate, a rich source of chaperones, up to 60% of the original luciferase activity could be recovered. Somewhat less re-activation was observed when the alpha-crystallin-bound enzyme was treated with heat-shock protein (HSP)70, HSP40, HSP60 and an ATP-generating system. Similar results were also obtained with citrate synthase. The overall results suggest that alpha-crystallin acts to stabilize denaturing proteins so that they can later be re-activated by other chaperones.  相似文献   

18.
Age-related cataract is a result of crystallins, the predominant lens proteins, forming light-scattering aggregates. In the low protein turnover environment of the eye lens, the crystallins are susceptible to modifications that can reduce stability, increasing the probability of unfolding and aggregation events occurring. It is hypothesized that the alpha-crystallin molecular chaperone system recognizes and binds these proteins before they can form the light-scattering centres that result in cataract, thus maintaining the long-term transparency of the lens. In the present study, we investigated the unfolding and aggregation of (wild-type) human and calf betaB2-crystallins and the formation of a complex between alpha-crystallin and betaB2-crystallins under destabilizing conditions. Human and calf betaB2-crystallin unfold through a structurally similar pathway, but the increased stability of the C-terminal domain of human betaB2-crystallin relative to calf betaB2-crystallin results in the increased population of a partially folded intermediate during unfolding. This intermediate is aggregation-prone and prevents constructive refolding of human betaB2-crystallin, while calf betaB2-crystallin can refold with high efficiency. alpha-Crystallin can effectively chaperone both human and calf betaB2-crystallins from thermal aggregation, although chaperone-bound betaB2-crystallins are unable to refold once returned to native conditions. Ordered secondary structure is seen to increase in alpha-crystallin with elevated temperatures up to 60 degrees C; structure is rapidly lost at temperatures of 70 degrees C and above. Our experimental results combined with previously reported observations of alpha-crystallin quaternary structure have led us to propose a structural model of how activated alpha-crystallin chaperones unfolded betaB2-crystallin.  相似文献   

19.
In vitro refolding of the monomeric mitochondrial enzyme, rhodanese (thiosulfate sulfurtransferase; EC 2.8.1.1) is facilitated by molecular chaperonins. The four components: two proteins from Escherichia coli, chaperonin 60 (groEL) and chaperonin 10 (groES), MgATP, and K+, are necessary for the in vitro folding of rhodanese. These were previously shown to be necessary for the in vitro folding of ribulose-1,5-bisphosphate carboxylase at temperatures in excess of 25 degrees C (Viitanen, P. V., Lubben, T. H., Reed, J., Goloubinoff, P., O'Keefe, D. P., and Lorimer, G. H. (1990) Biochemistry 29, 5665-5671). The labile folding intermediate, rhodanese-I, which rapidly aggregates at 37 degrees C in the absence of the chaperonins, can be stabilized by forming a binary complex with chaperonin 60. The discharge of the binary chaperonin 60-rhodanese-I complex, results in the formation of active rhodanese, and requires the presence of chaperonin 10. Optimal refolding is associated with a K(+)-dependent hydrolysis of ATP. At lower protein concentrations and 25 degrees C, where aggregation is reduced, a fraction of the rhodanese refolds to an active form in the absence of the chaperonins. This spontaneous refolding can be arrested by chaperonin 60. There is some refolding (approximately equal to 20%) when ATP is replaced by nonhydrolyzable analogs, but there is no refolding in the presence of ADP or AMP. ATP analogs may interfere with the interaction of rhodanese-I with the chaperonins. Nondenaturing detergents facilitate rhodanese refolding by interacting with exposed hydrophobic surfaces of folding intermediates and thereby prevent aggregation (Tandon, S., and Horowitz, P. (1986) J. Biol. Chem. 261, 15615-15618). The chaperonin proteins appear to play a similar role in as much as they can replace the detergents. Consistent with this view, chaperonin 60, but not chaperonin 10, binds 2-3 molecules of the hydrophobic fluorescent reporter, 1,1'-bi(4-anilino)naphthalene-S,5'-disulfonic acid, indicating the presence of hydrophobic surfaces on chaperonin 60. The number of bound probe molecules is reduced to 1-2 molecules when chaperonin 10 and MgATP are added. The results support a model in which chaperonins facilitate folding, at least in part, by interacting with partly folded intermediates, thus preventing the interactions of hydrophobic surfaces that lead to aggregation.  相似文献   

20.
The aggregation state of the gonadotropin receptor has been examined by coupling fluorescence energy transfer donor and acceptor fluorophores to hCG and LH. Energy transfer is observed at low (4 degrees C) but not at high (37 degrees C) temperature. Energy transfer could also be detected with receptor solubilized in the presence of hormone at the lower temperature only. Solubilization of receptor in the absence of hormone and subsequent addition of hormone conjugates revealed no energy transfer. These results are consistent with stabilization of receptor complexes at low temperatures, but presumptive hormone induced receptor dissociation under physiological conditions.  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