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 109 毫秒
1.
A detailed structural analysis of interactions between denatured proteins and GroEL is essential for an understanding of its mechanism. Minichaperones constitute an excellent paradigm for obtaining high-resolution structural information about the binding site and conformation of substrates bound to GroEL, and are particularly suitable for NMR studies. Here, we used transferred nuclear Overhauser effects to study the interaction in solution between minichaperone GroEL(193-335) and a synthetic peptide (Rho), corresponding to the N-terminal alpha-helix (residues 11 to 23) of the mitochondrial rhodanese, a protein whose in vitro refolding is mediated by minichaperones. Using a 60 kDa maltose-binding protein (MBP)-GroEL(193-335) fusion protein to increase the sensitivity of the transferred NOEs, we observed characteristic sequential and mid-range transferred nuclear Overhauser effects. The peptide adopts an alpha-helical conformation upon binding to the minichaperone. Thus the binding site of GroEL is compatible with binding of alpha-helices as well as extended beta-strands. To locate the peptide-binding site on GroEL(193-335), we analysed changes in its chemical shifts on adding an excess of Rho peptide. All residues with significant chemical shift differences are localised in helices H8 and H9. Non-specific interactions were not observed. This indicates that the peptide Rho binds specifically to minichaperone GroEL(193-335). The binding region identified by NMR in solution agrees with crystallographic studies with small peptides and with fluorescence quenching studies with denatured proteins.  相似文献   

2.
Advances in understanding how GroEL binds to non-native proteins are reported. Conformational flexibility in the GroEL apical domain, which could account for the variety of substrates that GroEL binds, is illustrated by comparison of several independent crystallographic structures of apical domain constructs that show conformational plasticity in helices H and I. Additionally, ESI-MS indicates that apical domain constructs have co-populated conformations at neutral pH. To assess the ability of different apical domain conformers to bind co-chaperone and substrate, model peptides corresponding to the mobile loop of GroES and to helix D from rhodanese were studied. Analysis of apical domain-peptide complexes by ESI-MS indicates that only the folded or partially folded apical domain conformations form complexes that survive gas phase conditions. Fluorescence binding studies show that the apical domain can fully bind both peptides independently. No competition for binding was observed, suggesting the peptides have distinct apical domain-binding sites. Blocking the GroES-apical domain-binding site in GroEL rendered the chaperonin inactive in binding GroES and in assisting the folding of denatured rhodanese, but still capable of binding non-native proteins, supporting the conclusion that GroES and substrate proteins have, at least partially, distinct binding sites even in the intact GroEL tetradecamer.  相似文献   

3.
The chaperonin GroEL binds a variety of polypeptides that share no obvious sequence similarity. The precise structural, chemical and dynamic features that are recognised remain largely unknown. Structural models of the complex between GroEL and its co-chaperonin GroES, and of the isolated apical domain of GroEL (minichaperone; residues 191-376) with a 17 residue N-terminal tag show that a linear sequential sequence (extended beta-strand) can be bound. We have analysed characteristics of the motifs that bind to GroEL by using affinity panning of immobilised GroEL minichaperones for a library of bacteriophages that display the fungal cellulose-binding domain of the enzyme cellobiohydrolase I. This protein has seven non-sequential residues in its binding site that form a linear binding motif with similar dimensions and characteristics to the peptide tag that was bound to the minichaperone GroEL(191-376). The seven residues thus form a constrained scaffold. We find that GroEL does bind suitable mutants of these seven residues. The side-chains recognised do not have to be totally hydrophobic, but polar and positively charged chains can be accommodated. Further, the spatial distribution of the side-chains is also compatible with those in an alpha-helix. This implies that GroEL can bind a wide range of structures, from extended beta-strands and alpha-helices to folded states, with exposed side-chains. The binding site can accommodate substrates of approximately 18 residues when in a helical or seven when in an extended conformation. The data support two activities of GroEL: the ability to act as a temporary parking spot for sticky intermediates by binding many motifs; and an unfolding activity of GroEL by binding an extended sequential conformation of the substrate.  相似文献   

4.
Structural studies on minichaperones and GroEL imply a continuous ring of binding sites around the neck of GroEL. To investigate the importance of this ring, we constructed an artificial heptameric assembly of minichaperones to mimic their arrangement in GroEL. The heptameric Gp31 co-chaperonin from bacteriophage T4, an analogue of GroES, was used as a scaffold to display the GroEL minichaperones. A fusion protein, MC(7), was generated by replacing a part of the highly mobile loop of Gp31 (residues 23-44) with the sequence of the minichaperone (residues 191-376 of GroEL). The purified recombinant protein assembled into a heptameric ring composed of seven 30.6 kDa subunits. Although single minichaperones (residues 193-335 to 191-376 of GroEL) have certain chaperone activities in vitro and in vivo, they cannot refold heat and dithiothreitol-denatured mitochondrial malate dehydrogenase (mtMDH), a reaction that normally requires GroEL, its co-chaperonin GroES and ATP. But, MC(7) refolded MDH in vitro. The expression of MC(7) complements in vivo two temperature-sensitive Escherichia coli alleles, groEL44 and groEL673, at 43 degrees C. Although MC(7) could not compensate for the complete absence of GroEL in vivo, it enhanced the colony-forming ability of cells containing limiting amounts of wild-type GroEL at 37 degrees C. MC(7 )also reduces aggregate formation and cell death in mammalian cell models of Huntington's disease. The assembly of seven minichaperone subunits on a heptameric ring significantly improves their activity, demonstrating the importance of avidity in GroEL function.  相似文献   

5.
We are reconstructing the mechanism of action of GroEL by a reductionist approach of isolating its minimal fragment that has residual activity (the "minichaperone" core) and then identifying how additional elements of structure confer further activity and function. We report here the 2.0 A resolution crystal structure of the minichaperone GroEL(193-345). The structure provides further clues on the nature of GroEL-polypeptide substrate interactions, because two molecules in the asymmetric unit interact by the binding of one molecule in the active site of its partner, thus mimicking a chaperone-polypeptide substrate complex. The results may explain some experimental observations, including the preference of GroEL for net positive charges (mediated by Glu238 and Glu257) and the key role of Tyr203 in mediating polypeptide binding. The larger binding site identified by these studies forms a continuous surface near the opening of the central cavity of GroEL that can accommodate a wide range of non-native protein conformations that differ in size and in structural and chemical properties.  相似文献   

6.
The isolated apical domain of GroEL consisting of residues 191–345 (known as “minichaperone”) binds and assists the folding of a wide variety of client proteins without GroES and ATP, but the mechanism of its action is still unknown. In order to probe into the matter, we have examined minichaperone-mediated folding of a large aggregation prone protein Maltodextrin-glucosidase (MalZ). The key objective was to identify whether MalZ exists free in solution, or remains bound to, or cycling on and off the minichaperone during the refolding process. When GroES was introduced during refolding process, production of the native MalZ was inhibited. We also observed the same findings with a trap mutant of GroEL, which stably captures a predominantly non-native MalZ released from minichaperone during refolding process, but does not release it. Tryptophan and ANS fluorescence measurements indicated that refolded MalZ has the same structure as the native MalZ, but that its structure when bound to minichaperone is different. Surface plasmon resonance measurements provide an estimate for the equilibrium dissociation constant KD for the MalZ-minichaperone complex of 0.21 ± 0.04 μM, which are significantly higher than for most GroEL clients. This showed that minichaperone interacts loosely with MalZ to allow the protein to change its conformation and fold while bound during the refolding process. These observations suggest that the minichaperone works by carrying out repeated cycles of binding aggregation-prone protein MalZ in a relatively compact conformation and in a partially folded but active state, and releasing them to attempt to fold in solution.  相似文献   

7.
GroEL is an essential Escherichia coli molecular chaperon that uses ATP to facilitate correct folding of a range of proteins in a cell. Central to the GroEL substrate diversity is how GroEL recognizes the substrates. The interaction between GroEL and substrate has been proposed to be largely hydrophobic because GroEL interacts with proteins in non-native conformations but not in native forms. Analysis of GroEL substrate proteins reveals that one of its main substrates are proteins with αβ folding domains, suggesting that GroEL may stabilize the collapsed αβ core by binding to hydrophobic surfaces that are usually buried between the α and β elements. In this study, we characterize the interaction between GroEL and a peptide derived from our previous selection via a phage display method. NMR studies map the peptide-binding site to the region containing Helices H and I, which is consistent with evidence that this region comprises the primary substrate-binding site. The peptide is largely unstructured in solution but adopts a helical conformation when bound to the GroEL apical domain with a moderate affinity (Kd = 17.1 ± 2.5 μm). The helical conformation aligns residues to form an amphipathic structure, and the hydrophobic side of this amphipathic helix interacts with GroEL as suggested by fluorescence quenching studies. Together with previous structural studies on the GroEL-peptide complexes, our work supports the notion that the amphipathic secondary elements in the substrate proteins may be the structural motif recognized by GroEL.The bacterial chaperonin GroEL and its co-chaperonin GroES are essential for cell viability by assisting folding of a wide range of proteins via an ATP-dependent mechanism (13). Structurally, fourteen 57-kDa GroEL subunits assemble into two back-to-back stacking heptameric rings, giving rise to two functionally independent central cavities (4). Each GroEL subunit folds into three distinctive domains: equatorial domain, intermediate domain, and apical domain. The equatorial domains contain the ATP-binding sites and provide most of the intra-ring interactions and all the inter-ring interactions. The apical domains form the rims of the central cavities and contain the binding sites for the substrate proteins and GroES. The intermediate domains link the apical domains and the equatorial domains. For the co-chaperonin GroES, seven GroES subunits, of 10 kDa each, assemble into a heptamer ring (5, 6). In forming the GroEL-GroES complex, GroES caps one end of GroEL, and large structural changes are observed in both GroEL and GroES (7). In GroEL, the apical domain is rotated 90° along its axis and 60° upwards, and the intermediate domain is closed down ∼25° to the equatorial domain. A loop in GroES (residues 17–33) that is unstructured in the isolated GroES adopts a β-turn structure and forms contact with the GroEL apical domain. Compared with the unliganded GroEL, the volume of the enclosed GroEL-GroES cavity is doubled, and the surface lining the wall of the GroEL cavity changes from hydrophobic to hydrophilic.A wealth of information derived from both intensive biochemical and structural characterizations has revealed a general role of GroEL-GroES in assisting protein folding (see reviews in Refs. 3, 8, and 9). Briefly, GroEL binds the substrate proteins in their aggregation-prone non-native states, preventing them from aggregating. Binding of ATP to the substrate occupied GroEL ring (cis-ring) presumably induces large conformational change in GroEL that promotes binding of GroES to the cis-ring. As a result of ATP and GroES binding, the substrate protein is displaced into the GroEL central cavity, initiating the folding process. Both hydrolysis of ATP in the cis-ring and binding of ATP to the substrate unoccupied ring (trans ring) weaken the GroES-GroEL interaction, and ATP binding to the trans ring results in the dissociation of GroES from GroEL, releasing substrate from the central cavity of GroEL. The released substrate may continue folding into the native state if in a folding competent state or may rebind to GroEL if it is still misfolded.One of the most intriguing aspects of the GroE-assisted folding is the substrate promiscuity. It has been shown that about 300 Escherichia coli proteins can interact with GroEL, and these proteins are diverse in terms of both structures and functions (10). A range of techniques have been applied to investigate this important yet complex aspect, and salient features regarding GroEL-substrate interactions have emerged. The apical domains, on the rim of the GroEL central cavity, contain the main substrate-binding site (1113). Structural flexibility, reflected by both high temperature factors of the apical domain in the crystal structure of tetradecameric GroEL (14) and conformational multiplicity around Helix H and I (15), is proposed to account for the diverse spectrum of GroEL substrates. Mutational studies on GroEL suggest that the GroEL-substrate interactions are largely hydrophobic (16). Structural study on GroEL-substrate interaction, however, is hindered mainly because of the multiple conformations of the bound substrate protein. Very recently, NMR techniques have been used to directly investigate the bound conformations of the substrate (17, 18); yet the nature of GroEL-substrate interaction is not revealed. Peptides may mimic segments of substrate proteins, and studies of GroEL-peptide interactions have uncovered detailed intermolecular interactions and provided insights into principles of substrate recognition by GroEL. The bound peptides may adopt α-helix (1923), β-hairpin (15), or extended conformations (24), and despite different conformations, they all appear to bind to Helix H and I of GroEL. Hydrophobic interaction dominates the interface between GroEL and peptides in either β-hairpin or extended structures and is proposed so between GroEL and α-helical peptides. These detailed structural characterizations on GroEL-peptide interactions have contributed to dissecting the complex nature of the substrate recognition by GroEL (25).We previously identified a high affinity peptide (strong binding peptide (SBP))2 for GroEL using a phage display method and found that SBP adopts a β-hairpin structure bound to GroEL (15, 26). To investigate the contribution of the β-turn in SBP to the GroEL-SBP interaction, we have created various SBP variants with the intension to disrupt the β-turn structure and have studied their binding to GroEL. One of the peptides (termed SBP-W2DP6V), however, adopts a helical conformation when bound to GroEL by NMR analysis. NMR results also map the peptide-binding site on GroEL to be a region formed by Helix H and I. The helical peptide has an amphipathic feature, and fluorescence studies provide direct evidence that the hydrophobic face is involved in the interaction with GroEL. Our structural analysis, combined with previous studies, suggests that GroEL recognizes the amphipathic property in the secondary structures of the substrate protein and binds preferably to the hydrophobic side of these structural elements to stabilize and preserve their structures.  相似文献   

8.
The chaperonin GroEL binds to a large number of polypeptides, prevents their self-association, and mediates appropriate folding in a GroES and adenosine triphosphate-dependent manner. But how the GroEL molecule actually recognizes the polypeptide and what are the exact GroEL recognition sites in the substrates are still poorly understood. We have examined more than 50 in vivo substrates as well as well-characterized in vitro substrates, for their binding characteristics with GroEL. While addressing the issue, we have been driven by the basic concept that GroES, being the cochaperonin of GroEL, is the best-suited substrate for GroEL, as well as by the fact that polypeptide substrate and GroES occupy the same binding sites on the GroEL apical domain. GroES interacts with GroEL through selective hydrophobic residues present on its mobile loop region, and we have considered the group of residues on the GroES mobile loop as the key element in choosing a substrate for GroEL. Considering the hydrophobic region on the GroES mobile loop as the standard, we have attempted to identify the homologous region on the peptide sequences in the proteins of our interest. Polypeptides have been judged as potential GroEL substrates on the basis of the presence of the GroES mobile loop-like hydrophobic segments in their amino acid sequences. We have observed 1 or more GroES mobile loop-like hydrophobic patches in the peptide sequence of some of the proteins of our interest, and the hydropathy index of most of these patches also seems to be approximately close to that of the standard. It has been proposed that the presence of hydrophobic patches having substantial degree of hydropathy index as compared with the standard segment is a necessary condition for a peptide sequence to be recognized by GroEL molecules. We also observed that the overall hydrophobicity is also close to 30% in these substrates, although this is not the sufficient criterion for a polypeptide to be assigned as a substrate for GroEL. We found that the binding of aconitase, alpha-lactalbumin, and murine dihydrofolate reductase to GroEL falls in line with our present model and have also predicted the exact regions of their binding to GroEL. On the basis of our GroEL substrate prediction, we have presented a model for the binding of apo form of some proteins to GroEL and the eventual formation of the holo form. Our observation also reveals that in most of the cases, the GroES mobile loop-like hydrophobic patch is present in the unstructured region of the protein molecule, specifically in the loop or beta-sheeted region. The outcome of our study would be an essential feature in identifying a potential substrate for GroEL on the basis of the presence of 1 or more GroES mobile loop-like hydrophobic segments in the amino acid sequence of those polypeptides and their location in three-dimensional space.  相似文献   

9.
While support in protein folding by molecular chaperones is extremely efficient for endogenous polypeptides, it often fails for recombinant proteins in a bacterial host, thus constituting a major hurdle for protein research and biotechnology. To understand the reasons for this difference and to answer the question of whether it is feasible to design tailor-made chaperones, we investigated one of the most prominent bacterial chaperones, the GroEL/ES ring complex. On the basis of structural data, we designed and constructed a combinatorial GroEL library, where the substrate-binding site was randomized. Screening and selection experiments with this library demonstrated that substrate binding and release is supported by many variants, but the majority of the library members failed to assist in chaperonin-mediated protein folding under conditions where spontaneous folding is suppressed. These findings revealed a conflict between binding of substrate and binding of the co-chaperonin GroES. As a consequence, the window of mutational freedom in that region of GroEL is very small. In screening experiments, we could identify GroEL variants slightly improved for a given substrate, which were still promiscuous. As the substrate-binding site of the GroEL molecule overlaps strongly with the site of cofactor binding, the outcome of our experiments suggests that maintenance of cofactor binding affinity is more critical for chaperonin-mediated protein folding than energetically optimized substrate recognition.  相似文献   

10.
In the crystal structure of the native GroEL.GroES.substrate protein complex from Thermus thermophilus, one GroEL subunit makes contact with two GroES subunits. One contact is through the H-I helices, and the other is through a novel GXXLE region. The side chain of Leu, in the GXXLE region, forms a hydrophobic cluster with residues of the H helix (Shimamura, T., Koike-Takeshita, A., Yokoyama, K., Masui, R., Murai, N., Yoshida, M., Taguchi, H., and Iwata, S. (2004) Structure (Camb.) 12, 1471-1480). Here, we investigated the functional role of Leu in the GXXLE region, using Escherichia coli GroEL. The results are as follows: (i) cross-linking between introduced cysteines confirmed that the GXXLE region in the E. coli GroEL.GroES complex is also in contact with GroES; (ii) when Leu was replaced by Lys (GroEL(L309K)) or other charged residues, chaperone activity was largely lost; (iii) the GroEL(L309K).substrate complex failed to bind GroES to produce a stable GroEL(L309K).GroES.substrate complex, whereas free GroEL(L309K) bound GroES normally; (iv) the GroEL(L309K).GroES.substrate complex was stabilized with BeF(x), but the substrate protein in the complex was readily digested by protease, indicating that it was not properly encapsulated into the internal cavity of the complex. Thus, conformational communication between the two GroES contact sites, the H helix and the GXXLE region (through Leu(309)), appears to play a critical role in encapsulation of the substrate.  相似文献   

11.
The next step in our reductional analysis of GroEL was to study the activity of an isolated single seven-membered ring of the 14-mer. A known single-ring mutant, GroEL(SR1), contains four point mutations that prevent the formation of double-rings. That heptameric complex is functionally inactive because it is unable to release GroES. We found that the mutation E191G, which is responsible for the temperature sensitive (ts) Escherichia coli allele groEL44 and is located in the hinge region between the intermediate and apical domains of GroEL, appears to function by weakening the binding of GroES, without destabilizing the overall structure of GroEL44 mutant. We introduced, therefore, the mutation E191G into GroEL(SR1) in order to generate a single-ring mutant that may have weaker binding of GroES and hence be active. The new single-ring mutant, GroEL(SR44), was indeed effective in refolding both heat and dithiothreitol-denatured mitochondrial malate dehydrogenase with great efficiency. Further, unlike all smaller constructs of GroEL, the expression of GroEL(SR44) in E. coli that contained no endogenous GroEL restored biological viability, but not as efficiently as does wild-type GroEL. We envisage the notional evolution of the structure and properties of GroEL. The minichaperone core acts as a primitive chaperone by providing a binding surface for denatured states that prevents their self-aggregation. The assembly of seven minichaperones into a ring then enhances substrate binding by introducing avidity. The acquisition of binding sites for ATP then allows the modulation of substrate binding by introducing the allosteric mechanism that causes cycling between strong and weak binding sites. This is accompanied by the acquisition by the heptamer of the binding of GroES, which functions as a lid to the central cavity and competes for peptide binding sites. Finally, dimerization of the heptamer enhances its biological activity.  相似文献   

12.
The GroES binding site at the apical domain of GroEL, mostly consisting of hydrophobic residues, overlaps largely with the substrate polypeptide binding site. Essential contribution of hydrophobic interaction to the binding of both GroES and polypeptide was exemplified by the mutant GroEL(L237Q) which lost the ability to bind either of them. The binding site, however, contains three hydrophilic residues, E238, T261, and N265. For GroES binding, N265 is essential since GroEL(N265A) is unable to bind GroES. E238 contributes to rapid GroES binding to GroEL because GroEL(E238A) is extremely sluggish in GroES binding. Polypeptide binding was not impaired by any mutations of E238A, T261A, and N265A. Rather, these mutants, especially GroEL(N265A), showed stronger polypeptide binding affinity than wild-type GroEL. Thus, these hydrophilic residues have a dual role; they help GroES binding on one hand but attenuate polypeptide binding on the other hand.  相似文献   

13.
The extent of hydrophobic exposure upon bis-ANS binding to the functional apical domain fragment of GroEL, or minichaperone (residues 191-345), was investigated and compared with that of the GroEL tetradecamer. Although a total of seven molecules of bis-ANS bind cooperatively to this minichaperone, most of the hydrophobic sites were induced following initial binding of one to two molecules of probe. From the equilibrium and kinetics studies at low bis-ANS concentrations, it is evident that the native apical domain is converted to an intermediate conformation with increased hydrophobic surfaces. This intermediate binds additional bis-ANS molecules. Tyrosine fluorescence detected denaturation demonstrated that bis-ANS can destabilize the apical domain. The results from (i) bis-ANS titrations, (ii) urea denaturation studies in the presence and absence of bis-ANS, and (iii) intrinsic tyrosine fluorescence studies of the apical domain are consistent with a model in which bis-ANS binds tightly to the intermediate state, relatively weakly to the native state, and little to the denatured state. The results suggest that the conformational changes seen in apical domain fragments are not seen in the intact GroEL oligomer due to restrictions imposed by connections of the apical domain to the intermediate domain and suppression of movement due to quaternary structure.  相似文献   

14.
A double-heptamer ring chaperonin GroEL binds denatured substrate protein, ATP, and GroES to the same heptamer ring and encapsulates substrate into the central cavity underneath GroES where productive folding occurs. GroES is a disk-shaped heptamer, and each subunit has a GroEL-binding loop. The residues of the GroEL subunit responsible for GroES binding largely overlap those involved in substrate binding, and the mechanism by which GroES can replace the substrate when GroES binds to GroEL/substrate complex remains to be clarified. To address this question, we generated single polypeptide GroES by fusing seven subunits with various combinations of active and GroEL binding-defective subunits. Functional tests of the fused GroES variants indicated that four active GroES subunits were required for efficient formation of the stable GroEL/GroES complex and five subunits were required for the productive GroEL/substrate/GroES complex. An increase in the number of defective GroES subunits resulted in a slowing of encapsulation and folding. These results indicate the presence of an intermediate GroEL/substrate/GroES complex in which the substrate and GroES bind to GroEL by sharing seven common binding sites.  相似文献   

15.
The double ring-shaped chaperonin GroEL binds a wide range of non-native polypeptides within its central cavity and, together with its cofactor GroES, assists their folding in an ATP-dependent manner. The conformational cycle of GroEL/ES has been studied extensively but little is known about how the environment in the central cavity affects substrate conformation. Here, we use the von Hippel-Lindau tumor suppressor protein VHL as a model substrate for studying the action of the GroEL/ES system on a bound polypeptide. Fluorescent labeling of pairs of sites on VHL for fluorescence (Förster) resonant energy transfer (FRET) allows VHL to be used to explore how GroEL binding and GroEL/ES/nucleotide binding affect the substrate conformation. On average, upon binding to GroEL, all pairs of labeling sites experience compaction relative to the unfolded protein while single-molecule FRET distributions show significant heterogeneity. Upon addition of GroES and ATP to close the GroEL cavity, on average further FRET increases occur between the two hydrophobic regions of VHL, accompanied by FRET decreases between the N- and C-termini. This suggests that ATP- and GroES-induced confinement within the GroEL cavity remodels bound polypeptides by causing expansion (or racking) of some regions and compaction of others, most notably, the hydrophobic core. However, single-molecule observations of the specific FRET changes for individual proteins at the moment of ATP/GroES addition reveal that a large fraction of the population shows the opposite behavior; that is, FRET decreases between the hydrophobic regions and FRET increases for the N- and C-termini. Our time-resolved single-molecule analysis reveals the underlying heterogeneity of the action of GroES/EL on a bound polypeptide substrate, which might arise from the random nature of the specific binding to the various identical subunits of GroEL, and might help explain why multiple rounds of binding and hydrolysis are required for some chaperonin substrates.  相似文献   

16.
GroEL C138W is a mutant form of Escherichia coli GroEL, which forms an arrested ternary complex composed of GroEL, the co-chaperonin GroES and the refolding protein molecule rhodanese at 25 degrees C. This state of arrest could be reversed with a simple increase in temperature. In this study, we found that GroEL C138W formed both stable trans- and cis-ternary complexes with a number of refolding proteins in addition to bovine rhodanese. These complexes could be reactivated by a temperature shift to obtain active refolded protein. The simultaneous binding of GroES and substrate to the cis ring suggested that an efficient transfer of substrate protein into the GroEL central cavity was assured by the binding of GroES prior to complete substrate release from the apical domain. Stopped-flow fluorescence spectroscopy of the mutant chaperonin revealed a temperature-dependent conformational change in GroEL C138W that acts as a trigger for complete protein release. The behavior of GroEL C138W was reflected closely in its in vivo characteristics, demonstrating the importance of this conformational change to the overall activity of GroEL.  相似文献   

17.
GroEL encapsulates nonnative substrate proteins in a central cavity capped by GroES, providing a safe folding cage. Conventional models assume that a single timer lasting approximately 8 s governs the ATP hydrolysis-driven GroEL chaperonin cycle. We examine single molecule imaging of GFP folding within the cavity, binding release dynamics of GroEL-GroES, ensemble measurements of GroEL/substrate FRET, and the initial kinetics of GroEL ATPase activity. We conclude that the cycle consists of two successive timers of approximately 3 s and approximately 5 s duration. During the first timer, GroEL is bound to ATP, substrate protein, and GroES. When the first timer ends, the substrate protein is released into the central cavity and folding begins. ATP hydrolysis and phosphate release immediately follow this transition. ADP, GroES, and substrate depart GroEL after the second timer is complete. This mechanism explains how GroES binding to a GroEL-substrate complex encapsulates the substrate rather than allowing it to escape into solution.  相似文献   

18.
Small-angle neutron scattering and contrast variation were used to study the solution structure of GroEL and GroEL/GroES chaperonins complexed with a nonnative variant of the polypeptide substrate, subtilisin (PJ9). The subtilisin was 86% deuterated (dPJ9) so that it contrasted sufficiently with the chaperonin, allowing the contrast variation technique to be used to separate the scattering from the two components bound in the complex. Both the native double-ring GroEL and a single-ring mutant were used with dPJ9 bound in a 1:1 stoichiometry per GroEL toroid. This allowed both the position and the shape of dPJ9 in the GroEL/dPJ9 complexes to be determined. A single-ring GroEL/GroES variant complexed with one dPJ9 molecule was used to study the structural changes of dPJ9 in GroEL/GroES/dPJ9 complexes formed with ADP and with ATP. It was found that both the shape and the position of the bound dPJ9 in the GroEL/GroES/dPJ9 complex with ADP were the same as those in the GroEL/dPJ9 complex. However, dPJ9 assumed a more symmetric shape when bound in the GroEL/GroES/dPJ9 complex with ATP. This important observation reflects the relative ability of ATP to promote refolding of protein substrates relative to that of ADP.  相似文献   

19.
The GroEL–GroES is an essential molecular chaperon system that assists protein folding in cell. Binding of various substrate proteins to GroEL is one of the key aspects in GroEL‐assisted protein folding. Small peptides may mimic segments of the substrate proteins in contact with GroEL and allow detailed structural analysis of the interactions. A model peptide SBP has been shown to bind to a region in GroEL that is important for binding of substrate proteins. Here, we investigated whether the observed GroEL–SBP interaction represented those of GroEL–substrate proteins, and whether SBP was able to mimic various aspects of substrate proteins in GroE‐assisted protein folding cycle. We found that SBP competed with substrate proteins, including α‐lactalbumin, rhodanese, and malate dehydrogenase, in binding to GroEL. SBP stimulated GroEL ATP hydrolysis rate in a manner similar to that of α‐lactalbumin. SBP did not prevent GroES from binding to GroEL, and GroES association reduced the ATPase rates of GroEL/SBP and GroEL/α‐lactalbumin to a comparable extent. Binding of both SBP and α‐lactalbumin to apo GroEL was dominated by hydrophobic interaction. Interestingly, association of α‐lactalbumin to GroEL/GroES was thermodynamically distinct from that to GroEL with reduced affinity and decreased contribution from hydrophobic interaction. However, SBP did not display such differential binding behaviors to apo GroEL and GroEL/GroES, likely due to the lack of a contiguous polypeptide chain that links all of the bound peptide fragments. Nevertheless, studies using peptides provide valuable information on the nature of GroEL–substrate protein interaction, which is central to understand the mechanism of GroEL‐assisted protein folding. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
GroEL assists protein folding by preventing the interaction of partially folded molecules with other non-native proteins. It binds them, sequesters them, and then releases them so that they can fold in an ATP-driven cycle. Previous studies have also shown that protein substrates, GroES, and oligopeptides bind to partially overlapped sites on the apical domain surfaces of GroEL. In this study, we have determined the crystal structure at 3.0A resolution of a symmetric (GroEL-peptide)(14) complex. The binding of each of these small 12 amino acid residue peptides to GroEL involves interactions between three adjacent apical domains of GroEL. Each peptide interacts primarily with a single GroEL subunit. Residues R231 and R268 from adjacent subunits isolate each substrate-binding pocket, and prevent bound substrates from sliding into adjacent binding pockets. As a consequence of peptide binding, domains rotate and inter-domain interactions are greatly enhanced. The direction of rotation of the apical domain of each GroEL subunit is opposite to that of its intermediate domain. Viewed from outside, the apical domains rotate clockwise within one GroEL ring, while the ATP-induced apical domain rotation is counter-clockwise.  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