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.
When chaperonins GroEL and GroES are incubated under functional conditions in the presence of ATP (5 mM) and K+ (150 mM), GroEL-GroES complexes appear in the incubation mixture, that are either asymmetric (1:1 GroEL:GroES oligomer ratio) or symmetric (1:2 GroEL:GroES oligomer ratio). The percentage of symmetric complexes present is directly related to the [ATP]/[ADP] ratio and to the K+ concentration. Kinetic analysis shows that there is a cycle of formation and disappearance of symmetric complexes. A correlation between the presence of symmetric complexes in the incubation mixture and its rhodanese folding activity suggests some active role of these complexes in the protein folding process. Accordingly, under functional conditions, symmetric complexes are found to contain denatured rhodanese. These data suggest that binding of substrate inside the GroEL cavity takes place before the symmetric complex is formed.  相似文献   

2.
The conserved residue Lys-34 in GroES was replaced by alanine and glutamic acid using site-directed mutagenesis. This residue is near the carboxy terminus of the mobile loop in GroES (residues 17-32) which becomes immobilized upon formation of the GroEL/GroES complex [Landry et al. (1993) Nature 364, 255-258]. Both charge neutralization (Lys-34-->Ala) and charge reversal (Lys-34-->Glu) at this position have little effect on the binding constant of GroES to GroEL, but they increase the enhancement by GroES of cooperativity in ATP hydrolysis by GroEL. This is reflected by a change in the Hill coefficient (at 10 mM K+) from 4.10 (+/- 0.22) in the presence of wild-type GroES to 5.17 (+/- 0.24) and 4.46 (+/- 0.14) in the presence of the GroES mutants Lys-34-->Ala and Lys-34-->Glu, respectively. The results are interpreted using the Monod-Wyman-Changeux (MWC) model for cooperativity [Monod et al. (1965) J. Mol. Biol. 12, 88-118]. They suggest that Lys-34 in GroES modulates the allosteric transition in GroEL by stabilizing a relaxed (R)-like state.  相似文献   

3.
Although the chaperonin GroEL/GroES complex binds and hydrolyzes ATP, its structure is unlike other known ATPases. In order to better characterize its nucleotide binding sites, we have photolabeled the complex with the affinity analog 2-azido-ATP. Three residues of GroEL, Pro137, Cys138 and Thr468, are labeled by the probe. The location of these residues in the GroEL crystal structure [Braig, K., Otwinowski, Z., Hedge, R., Boisvert, D., Joachimiak, A., Horwich, A. & Sigler, P. (1994) Nature 371, 578-586: Boisvert, D. C., Wang, J., Otwinowski, Z., Horwich, A. L. & Sigler, P. B. (1996) Nat. Struct. Biol. 3, 170-177] suggests that 2-azido-ATP binds to an alternative conformer of GroEL in the presence of GroES. The labeled site appears to be located at the GroEL/GroEL subunit interface since modification of Pro137 and Cys138 is most readily explained by attack of a probe molecule bound to the adjacent GroEL subunit. Labeling of the co-chaperonin, GroES, is clearly demonstrated on gels and the covalent tethering of nucleotide allows detection of a GroES dimer in the presence of SDS. However, no stable peptide derivative of GroES could be purified for sequencing. In contrast, the GroES homolog, yeast cpn10, does give a stable derivative. The modified amino acid is identified as the conserved Pro13, which corresponds to Pro5 in Escherichia coli GroES.  相似文献   

4.
We have previously shown that the C-terminal sequence of GroES is required for oligomerization [Seale and Horowitz (1995), J. Biol. Chem. 270, 30268-30270]. In this report, we have generated a C-terminal deletion mutant of GroES with a significantly destabilized oligomer and have investigated its function in the chaperonin-assisted protein folding cycle. Removal of the two C-terminal residues of GroES results in a cochaperonin [GroESD(96-97)] that is monomeric at concentrations where GroES function is assessed. Using equilibrium ultracentrifugation, we measured the dissociation constant for the oligomer-monomer equilibrium to be 7.3 x 10(-34)M6. The GroESD(96-97) is fully active as a cochaperonin. This mutant is able to inhibit the ATPase activity of GroEL to levels comparable to wild-type GroES. It is also able to assist the refolding of urea-denatured rhodanese by GroEL. While GroESD(96-97) can function at levels comparable to wild-type GroES, higher concentrations of mutant are required to produce the same effect. These results support the idea that the performed GroES heptamer is not required for function, but they suggest that the oligomeric cochaperonin is most efficient.  相似文献   

5.
The chaperonins GroEL and GroES of Escherichia coli facilitate protein folding in an adenosine triphosphate (ATP)-dependent reaction cycle. The kinetic parameters for the formation and dissociation of GroEL-GroES complexes were analyzed by surface plasmon resonance. Association of GroES and subsequent ATP hydrolysis in the interacting GroEL toroid resulted in the formation of a stable GroEL:ADP:GroES complex. The complex dissociated as a result of ATP hydrolysis in the opposite GroEL toroid, without formation of a symmetrical GroEL:(GroES)2 intermediate. Dissociation was accelerated by the addition of unfolded polypeptide. Thus, the functional chaperonin unit is an asymmetrical GroEL:GroES complex, and substrate protein plays an active role in modulating the chaperonin reaction cycle.  相似文献   

6.
Mitochondrial malate dehydrogenase (mMDH) folds more rapidly in the presence of GroEL, GroES and ATP than it does unassisted. The increase in folding rate as a function of the concentration of GroEL-ES reaches a maximum at a stoichiometry which is approximately equimolar (mMDH subunits:GroEL oligomer) and with an apparent dissociation constant K' for the GroE acceptor state of at least 1 x 10(-8) M. However, even at chaperonin concentrations which are 4000 x K', i.e. at negligible concentrations of free mMDH, the observed folding rate of the substrate remains at its optimum, showing not only that folding occurs in the chaperonin-mMDH complex but also that this rate is uninhibited by any interactions with sites on GroEL. Despite the ability of mMDH to fold on the chaperonin, trapping experiments show that its dwell time on the complex is only 20 seconds. This correlates with both the rate of ATP turnover and the dwell time of GroES on the complex and is only approximately 5% of the time taken for the substrate to commit to the folded state. The results imply that ATP drives the chaperonin complex through a cycle of three functional states: (1) an acceptor complex in which the unfolded substrate is bound tightly; (2) an encapsulation state in which it is sequestered but direct protein-protein contact is lost so that folding can proceed unhindered; and (3) an ejector state which forces dissociation of the substrate whether folded or not.  相似文献   

7.
We have analyzed the effects of different components of the GroE chaperonin system on protein folding by using a nonpermissive substrate (i.e., one that has very low spontaneous refolding yield) for which rate data can be acquired. In the absence of GroES and nucleotides, the rate of GroEL-mediated refolding of heat- and DTT-denatured mitochondrial malate dehydrogenase was extremely low, but some three times higher than the spontaneous rate. This GroEL-mediated rate was increased 17-fold by saturating concentrations of ATP, 11-fold by ADP and GroES, and 465-fold by ATP and GroES. Optimal refolding activity was observed when the dissociation of GroES from the chaperonin complex was dramatically reduced. Although GroEL minichaperones were able to bind denatured mitochondrial malate dehydrogenase, they were ineffective in enhancing the refolding rate. The spectrum of mechanisms for GroE-mediated protein folding depends on the nature of the substrate. The minimal mechanism for permissive substrates (i.e., having significant yields of spontaneous refolding), requires only binding to the apical domain of GroEL. Slow folding rates of nonpermissive substrates are limited by the transitions between high- and low-affinity states of GroEL alone. The optimal mechanism, which requires holoGroEL, physiological amounts of GroES, and ATP hydrolysis, is necessary for the chaperonin-mediated folding of nonpermissive substrates at physiologically relevant rates under conditions in which retention of bound GroES prevents the premature release of aggregation-prone folding intermediates from the chaperonin complex. The different mechanisms are described in terms of the structural features of mini- and holo-chaperones.  相似文献   

8.
Electron microscopy of the tetradecameric double-ring complex of GroEL reveals a four-layered structure, indicating that the 58 kDa subunits are composed of two major morphological domains. We have used immuno-electron microscopy to assign these domains to the corresponding segments of the GroEL sequence. Upon chemical modification of GroEL with N-ethylmaleimide, protease treatment in the presence of ATP or ADP generates GroEL fragments of 15 kDa (N15; residues 1-141) and 40 kDa (C40; residues 153-531). As visualized by scanning transmission electron microscopy, affinity-purified antibodies directed against C40 recognize the outer layers, whereas antibodies against N15 interact with the equatorial portions of the GroEL double-ring. Thus, the two major domains of the subunits in the chaperonin complex are arranged in the order C40-N15:N15-C40. The single-ring chaperonin co-factor GroES interacts with the C40 domain while the ATP-binding site of GroEL is probably close to the junction between N15 and C40.  相似文献   

9.
Chaperonins GroEL and GroES form two types of hetero-oligomers in vitro that can mediate the folding of proteins. Chemical cross-linking and electron microscopy showed that in the presence of adenosine triphosphate (ATP), two GroES7 rings can successively bind a single GroEL14 core oligomer. The symmetric GroEL14(GroES7)2 chaperonin, whose central cavity appears obstructed by two GroES7 rings, can nonetheless stably bind and assist the ATP-dependent refolding of RuBisCO enzyme. Thus, unfolded proteins first bind and possibly fold on the external envelope of the chaperonin hetero-oligomer.  相似文献   

10.
Fragments encompassing the apical domain of GroEL, called minichaperones, facilitate the refolding of several proteins in vitro without requiring GroES, ATP, or the cage-like structure of multimeric GroEL. We have identified the smallest minichaperone that is active in vitro in chaperoning the refolding of rhodanese and cyclophilin A: GroEL(193-335). This finding raises the question of whether the minichaperones are active under more stringent conditions in vivo. The smallest minichaperones complement two temperature-sensitive Escherichia coli groEL alleles, EL44 and EL673, at 43 degreesC. Although they cannot replace GroEL in cells in which the chromosomal groEL gene has been deleted by P1 transduction, GroEL(193-335) enhances the colony-forming ability of such cells when limiting amounts of GroEL are expressed from a tightly regulated plasmid. Surprisingly, we found that overexpression of GroEL prevents plaque formation by bacteriophage lambda and inhibits replication of the lambda origin-dependent plasmid, Lorist6. The minichaperones also inhibit Lorist6 replication, but less markedly. The complex quaternary structure of GroEL, its central cavity, and the structural allosteric changes that take place on the binding of nucleotides and GroES are not essential for all of its functions in vivo.  相似文献   

11.
In this work we show that the GroEL (Hsp60 equivalent) chaperone protein can protected purified Escherichia coli RNA polymerase (RNAP) holoenzyme from heat inactivation better than the DnaK (Hsp70 equivalent) chaperone can. In this protection reaction, the GroES protein is not essential, but its presence reduces the amount of GroEL required. GroEL and GroES can also reactivate heat-inactivated RNAP in the presence of ATP. The mutant GroEL673 protein, with or without GroES, is incapable of reactivating heat-inactivated RNAP. GroEL673 can only protect RNAP, and this protecting ability is not stimulated by GroES. The mechanism by which the DnaJ and GrpE heat shock proteins contribute to DnaK's ability to reactivate heat-inactivated RNAP GroEL673 has also been investigated. We found that the DnaJ protein substantially reduces the levels of DnaK protein needed in this reactivation assay. However, the observed lag in reactivation is diminished only in the additional presence of the GrpE protein. Hence, DnaJ and GrpE are involved in both steps of this reactivation reaction (recognition of substrate and release of chaperone from the substrate-chaperone complex) while, in the case of the GroEL-dependent reaction, GroES is involved only during the release of chaperone from the substrate-chaperone complex.  相似文献   

12.
Using stopped-flow fluorescence techniques, we have examined both the refolding and unfolding reactions of four structurally homologous dihydrofolate reductases (murine DHFR, wild-type E. coli DHFR, and two E. coli DHFR mutants) in the presence and absence of the molecular chaperonin GroEL. We show that GroEL binds the unfolded conformation of each DHFR with second order rate constants greater than 3 x 10(7) M(-1)s(-1) at 22 degrees C. Once bound to GroEL, the proteins refold with rate constants similar to those for folding in the absence of GroEL. The overall rate of formation of native enzyme is decreased by the stability of the complex between GroEL and the last folding intermediate. For wild-type E. coli DHFR, complex formation is transient while for the others, a stable complex is formed. The stable complexes are the same regardless of whether they are formed from the unfolded or folded DHFR. When complex formation is initiated from the native conformation, GroEL binds to a pre-existing non-native conformation, presumably a late folding intermediate, rather than to the native state, thus shifting the conformational equilibrium toward the non-native species by mass action. The model presented here for the interaction of these four proteins with GroEL quantitatively describes the difference between the formation of a transient complex and a stable complex as defined by the rate constants for release and rebinding to GroEL relative to the rate constant for the last folding step. Due to this kinetic partitioning, three different mechanisms can be proposed for the formation of stable complexes between GroEL and either murine DHFR or the two E. coli DHFR mutants. These data show that productive folding of GroEL-bound proteins can occur in the absence of nucleotides or the co-chaperonin GroES and suggest that transient complex formation may be the functional role of GroEL under normal conditions.  相似文献   

13.
Two models are being considered for the mechanism of chaperonin-assisted protein folding in E. coli: (i) GroEL/GroES act primarily by enclosing substrate polypeptide in a folding cage in which aggregation is prevented during folding. (ii) GroEL mediates the repetitive unfolding of misfolded polypeptides, returning them onto a productive folding track. Both models are not mutually exclusive, but studies with the polypeptide-binding domain of GroEL have suggested that unfolding is the primary mechanism, enclosure being unnecessary. Here we investigate the capacity of the isolated apical polypeptide-binding domain to functionally replace the complete GroEL/GroES system. We show that the apical domain binds aggregation-sensitive polypeptides but cannot significantly assist their refolding in vitro and fails to replace the groEL gene or to complement defects of groEL mutants in vivo. A single-ring version of GroEL cannot substitute for GroEL. These results strongly support the view that sequestration of aggregation-prone intermediates in a folding cage is an important element of the chaperonin mechanism.  相似文献   

14.
The aim of this study was to compare the effects of increased concentrations of MgADP, inorganic phosphate (Pi) and H+ ([MgADP], [Pi] and [H+], respectively) on the rate of relaxation in two different muscle types: skinned muscle fibres from the frog Rana temporaria and myofibrillar bundles from the giant Pacific acorn barnacle Balanus nubilus. Relaxation transients are produced by the photolysis of diazo-2 and are well fitted with a double exponential curve, giving two rate constants: k1 [5.6+/-0.1 s-1 for barnacle, n=30; 26.3+/-0.7 s-1 for frog, n=14 (mean+/-SEM)] and k2 [0.6+/-0.1 s-1 in barnacle, n=30; 10.4+/-1.0 s-1 in frog, n=14 (mean+/-SEM)], at 10 degrees C. Decreasing the pH by 0.5 pH units did not significantly affect k1 for barnacle relaxation [5.6+/-0.1 s-1 (mean+/-SEM), n=15] compared to the decrease in k1 of 40% seen in frog. Use of the Ca2+-sensitive fluorescent label acrylodan on barnacle wild-type troponin C demonstrated that decreasing the pH from 7.0 to 6.6 only alters the pCa50 value by 0.23 in the cuvette, while stopped-flow experiments with acrylodan revealed no significant change in koff from the labelled protein [322+/-32 s-1 at pH 7.0 and 381+/-24 s-1 (mean+/-SEM) at pH 6.6]. Increasing [MgADP] by 20 microM (50 microM added ADP) from control values of 50 microM in frog decreased k1 to 12.3+/-0.4 s-1 (mean+/-SEM, n=8), and at 400 microM MgADP, k1=9.6+/-0.1 s-1 (mean+/-SEM, n=12). In barnacle, 500 microM MgADP had a much smaller effect on k1 (4.0+/-0. 9 s-1, mean+/-SEM, n=8). Increasing the free [Pi] from the contaminant level of 0.36 mM to 1.9 mM slowed k1 by approximately 15% in barnacle [4.8+/-0.8 s-1, mean+/-SEM, n=7], compared to a approximately 30% reduction seen in frog. We conclude that the differences between barnacle and frog seen here are most probably due to different isomers of the contractile proteins, and that events underlying the crossbridge cycle are the same or similar. We interpret our results according to a model of crossbridge transitions during relaxation.  相似文献   

15.
We propose a mechanism for the role of the bacterial chaperonin GroEL in folding proteins. The principal assumptions of the mechanism are (i) that many unfolded proteins bind to GroEL because GroEL preferentially binds small unstructured regions of the substrate protein, (ii) that substrate protein within the cavity of GroEL folds by the same kinetic mechanism and rate processes as in bulk solution, (iii) that stable or transient complexes with GroEL during the folding process are defined by a kinetic partitioning between formation and dissociation of the complex and the rate of folding and unfolding of the protein, and (iv) that dissociation from the complex in early stages of folding may lead to aggregation but dissociation at a late stage leads to correct folding. The experimental conditions for refolding may play a role in defining the function of GroEL in the folding pathway. We propose that the role of GroES and MgATP, either binding or hydrolysis, is to regulate the association and dissociation processes rather than affecting the rate of folding.  相似文献   

16.
As a basic principle, assisted protein folding by GroEL has been proposed to involve the disruption of misfolded protein structures through ATP hydrolysis and interaction with the cofactor GroES. Here, we describe chaperonin subreactions that prompt a re-examination of this view. We find that GroEL-bound substrate polypeptide can induce GroES cycling on and off GroEL in the presence of ADP. This mechanism promotes efficient folding of the model protein rhodanese, although at a slower rate than in the presence of ATP. Folding occurs when GroES displaces the bound protein into the sequestered volume of the GroEL cavity. Resulting native protein leaves GroEL upon GroES release. A single-ring variant of GroEL is also fully functional in supporting this reaction cycle. We conclude that neither the energy of ATP hydrolysis nor the allosteric coupling of the two GroEL rings is directly required for GroEL/GroES-mediated protein folding. The minimal mechanism of the reaction is the binding and release of GroES to a polypeptide-containing ring of GroEL, thereby closing and opening the GroEL folding cage. The role of ATP hydrolysis is mainly to induce conformational changes in GroEL that result in GroES cycling at a physiologically relevant rate.  相似文献   

17.
The ATP-sensitive potassium (KATP) channels in pancreatic beta cells are critical in the regulation of glucose-induced insulin secretion. Although electrophysiological studies provide clues to the complex control of KATP channels by ATP, MgADP, and pharmacological agents, the molecular mechanism of KATP-channel regulation remains unclear. The KATP channel is a heterooligomeric complex of SUR1 subunits of the ATP-binding-cassette superfamily with two nucleotide-binding folds (NBF1 and NBF2) and the pore-forming Kir6.2 subunits. Here, we report that MgATP and MgADP, but not the Mg salt of gamma-thio-ATP, stabilize the binding of prebound 8-azido-[alpha-32P]ATP to SUR1. Mutation in the Walker A and B motifs of NBF2 of SUR1 abolished this stabilizing effect of MgADP. These results suggest that SUR1 binds 8-azido-ATP strongly at NBF1 and that MgADP, either by direct binding to NBF2 or by hydrolysis of bound MgATP at NBF2, stabilizes prebound 8-azido-ATP binding at NBF1. The sulfonylurea glibenclamide caused release of prebound 8-azido-[alpha-32P]ATP from SUR1 in the presence of MgADP or MgATP in a concentration-dependent manner. This direct biochemical evidence of cooperative interaction in nucleotide binding of the two NBFs of SUR1 suggests that glibenclamide both blocks this cooperative binding of ATP and MgADP and, in cooperation with the MgADP bound at NBF2, causes ATP to be released from NBF1.  相似文献   

18.
Modification of the Escherichia coli chaperonin GroEL with N-ethylmaleimide at residue Cys138 affects the structural and functional integrity of the complex. Nucleotide affinity and ATPase activity of the modified chaperonin are increased, whereas cooperativity of ATP hydrolysis and affinity for GroES are reduced. As a consequence, release and folding of substrate proteins are strongly impaired and uncoupled from ATP hydrolysis in a temperature-dependent manner. Folding of dihydrofolate reductase at 25 degrees C becomes dependent on GroES, whereas folding of typically GroES-dependent proteins is blocked completely. At 37 degrees C, GroES binding is restored to normal levels, and the modified GroEL regains its chaperone activity to some extent. These results assign a central role to the intermediate GroEL domain for transmitting conformational changes between apical and central domains, and for coupling ATP hydrolysis to productive protein release.  相似文献   

19.
The Gp31 protein from bacteriophage T4 functionally substitutes for the bacterial co-chaperonin GroES in assisted protein folding reactions both in vitro and in vivo. But Gp31 is required for the folding and/or assembly of the T4 major capsid protein Gp23, and this requirement cannot be satisfied by GroES. The 2.3 A crystal structure of Gp31 shows that its tertiary and quaternary structures are similar to those of GroES despite the existence of only 14% sequence identity between the two proteins. However, Gp31 shows a series of structural adaptations which will increase the size and the hydrophilicity of the "Anfinsen cage," the enclosed cavity within the GroEL/GroES complex that is the location of the chaperonin-assisted protein folding reaction.  相似文献   

20.
The shape and subunit arrangement of the Escherichia coli F1 ATPase (ECF1 ATPase) was investigated by synchrotron radiation x-ray solution scattering. The radius of gyration and the maximum dimension of the enzyme complex are 4.61 +/- 0.03 nm and 15.5 +/- 0.05 nm, respectively. The shape of the complex was determined ab initio from the scattering data at a resolution of 3 nm, which allowed unequivocal identification of the volume occupied by the alpha3beta3 subassembly and further positioning of the atomic models of the smaller subunits. The delta subunit was positioned near the bottom of the alpha3beta3 hexamer in a location consistent with a beta-delta disulfide formation in the mutant ECF1 ATPase, betaY331W:betaY381C:epsilonS108C, when MgADP is bound to the enzyme. The position and orientation of the epsilon subunit were found by interactively fitting the solution scattering data to maintain connection of the two-helix hairpin with the alpha3beta3 complex and binding of the beta-sandwich domain to the gamma subunit. Nucleotide-dependent changes of the delta subunit were investigated by stopped-flow fluorescence technique at 12 degrees C using N-[4-[7-(dimethylamino)-4-methyl]coumarin-3-yl]maleimide (CM) as a label. Fluorescence quenching monitored after addition of MgATP was rapid [k = 6.6 s-1] and then remained constant. Binding of MgADP and the noncleavable nucleotide analog AMP . PNP caused an initial fluorescent quenching followed by a slower decay back to the original level. This suggests that the delta subunit undergoes conformational changes and/or rearrangements in the ECF1 ATPase during ATP hydrolysis.  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