首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Chaperonin GroEL from Escherichia coli consists of two heptameric rings stacked back-to-back to form a cagelike structure. It assists in the folding of substrate proteins in concert with the co-chaperonin GroES by incorporating them into its large cavity. The mechanism underlying the incorporation of substrate proteins currently remains unclear. The flexible C-terminal residues of GroEL, which are invisible in the x-ray crystal structure, have recently been suggested to play a key role in the efficient encapsulation of substrates. These C-terminal regions have also been suggested to separate the double rings of GroEL at the bottom of the cavity. To elucidate the role of the C-terminal regions of GroEL on the efficient encapsulation of substrate proteins, we herein investigated the effects of C-terminal truncation on GroE-mediated folding using the green fluorescent protein (GFP) as a substrate. We demonstrated that the yield of in-cage folding mediated by a single ring GroEL (SR1) was markedly decreased by truncation, whereas that mediated by a double ring football-shaped complex was not affected. These results suggest that the C-terminal region of GroEL functions as a barrier between rings, preventing the leakage of GFP through the bottom space of the cage. We also found that once GFP folded into its native conformation within the cavity of SR1 it never escaped even in the absence of the C-terminal tails. This suggests that GFP molecules escaped through the pore only when they adopted a denatured conformation. Therefore, the folding and escape of GFP from C-terminally truncated SR1·GroES appeared to be competing with each other.  相似文献   

2.
To elucidate the exact role of the C-terminal region of GroEL in its functional cycle, the C-terminal 20-amino acid truncated mutant of GroEL was constructed. The steady-state ATPase rate and duration of GroES binding showed that the functional cycle of the truncated GroEL is extended by approximately 2 s in comparison with that of the wild type, without interfering with the basic functions of GroEL. We have proposed a model for the functional cycle of GroEL, which consists of two rate-limiting steps of approximately 3- and approximately 5-s duration (Ueno, T., Taguchi, H., Tadakuma, H., Yoshida, M., and Funatsu, T. (2004) Mol. Cell 14, 423-434 g). According to the model, detailed kinetic studies were performed. We found that a 20-residue truncation of the C terminus extends the time until inorganic phosphate is generated and the time for arresting protein folding in the central cavity, i.e. the lifetime of the first rate-limiting step in the functional cycle, to an approximately 5-s duration. These results suggest that the integrity of the C-terminal region facilitates the transition from the first to the second rate-limiting state.  相似文献   

3.
The alpha-amylase from Bacillus sp. strain TS-23 is a secreted starch hydrolase with a domain organization similar to that of other microbial alpha-amylases and an additional functionally unknown domain (amino acids 517-613) in the C-terminal region. By sequence comparison, we found that this latter domain contained a sequence motif typical for raw-starch binding. To investigate the functional role of the C-terminal region of the alpha-amylase of Bacillus sp. strain TS-23, four His(6)-tagged mutants with extensive deletions in this region were constructed and expressed in Escherichia coli. SDS-PAGE and activity staining analyses showed that the N- and C-terminally truncated alpha-amylases had molecular masses of approximately 65, 58, 54, and 49 kDa. Progressive loss of raw-starch-binding activity occurred upon removal of C-terminal amino acid residues, indicating the requirement for the entire region in formation of a functional starch-binding domain. Up to 98 amino acids from the C-terminal end of the alpha-amylase could be deleted without significant effect on the raw-starch hydrolytic activity or thermal stability. Furthermore, the active mutants hydrolyzed raw corn starch to produce maltopentaose as the main product, suggesting that the raw-starch hydrolytic activity of the Bacillus sp. strain TS-23 alpha-amylase is functional and independent from the starch-binding domain.  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.
The mechanism of GroEL (chaperonin)-mediated protein folding is only partially understood. We have analysed structural and functional properties of the interaction between GroEL and the co-chaperonin GroES. The stoichiometry of the GroEL 14mer and the GroES 7mer in the functional holo-chaperonin is 1:1. GroES protects half of the GroEL subunits from proteolytic truncation of the approximately 50 C-terminal residues. Removal of this region results in an inhibition of the GroEL ATPase, mimicking the effect of GroES on full-length GroEL. Image analysis of electron micrographs revealed that GroES binding triggers conspicuous conformational changes both in the GroES adjacent end and at the opposite end of the GroEL cylinder. This apparently prohibits the association of a second GroES oligomer. Addition of denatured polypeptide leads to the appearance of irregularly shaped, stain-excluding masses within the GroEL double-ring, which are larger with bound alcohol oxidase (75 kDa) than with rhodanese (35 kDa). We conclude that the functional complex of GroEL and GroES is characterized by asymmetrical binding of GroES to one end of the GroEL cylinder and suggest that binding of the substrate protein occurs within the central cavity of GroEL.  相似文献   

6.
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   

7.
A weak reversibly acting neurotoxin, fasciatoxin, was found in the venom of Bungarus fasciatus. The sequencing was completed by manual and automated Edman analyses of the reduced and carboxymethylated protein and of the peptides obtained from enzyme digestions. It is composed of 63 amino acid residues with four disulphide bonds and a unique sequence at the C-terminal end. According to the criteria set by Ryden, Gabel & Eaker [(1973) Int. J. Pept. Protein Res. 5, 261-273], fasciatoxin lacks all of the five functionally invariant residues of neurotoxins. The hydropathy index indicates that fasciatoxin is devoid of a strong hydrophilicity domain for binding to the receptor site. Structural comparison with some typical neurotoxins also reveals the uniqueness of fasciatoxin in that the extent of similarity is only about 30%.  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.
Luteoviruses avoid degradation in the hemolymph of their aphid vector by interacting with a GroEL homolog from the aphid's primary endosymbiotic bacterium (Buchnera sp.). Mutational analysis of GroEL from the primary endosymbiont of Myzus persicae (MpB GroEL) revealed that the amino acids mediating binding of Potato leafroll virus (PLRV; Luteoviridae) are located within residues 9 to 19 and 427 to 457 of the N-terminal and C-terminal regions, respectively, of the discontinuous equatorial domain. Virus overlay assays with a series of overlapping synthetic decameric peptides and their derivatives demonstrated that R13, K15, L17, and R18 of the N-terminal region and R441 and R445 of the C-terminal region of the equatorial domain of GroEL are critical for PLRV binding. Replacement of R441 and R445 by alanine in full-length MpB GroEL and in MpB GroEL deletion mutants reduced but did not abolish PLRV binding. Alanine substitution of either R13 or K15 eliminated the PLRV-binding capacity of the other and those of L17 and R18. In the predicted tertiary structure of GroEL, the determinants mediating virus binding are juxtaposed in the equatorial plain.  相似文献   

10.
p23 is a co-chaperone of the heat shock protein Hsp90. p23 binds to Hsp90 in its ATP-bound state and, on its own, interacts specifically with non-native proteins. In our attempt to correlate these functions to specific regions of p23 we have identified an unstructured region in p23 that maps to the C-terminal part of the protein sequence. This unstructured region is dispensible for interaction of p23 with Hsp90, since truncated p23 can still form complexes with Hsp90. In contrast, however, truncation of the C-terminal 30 amino acid residues of p23 affects the ability of p23 to bind non-native proteins and to prevent their non-specific aggregation. The isolated C-terminal region itself is not able to act as a chaperone nor is it possible to complement truncated p23 by addition of this peptide. These results imply that the binding site for Hsp90 is contained in the folded domain of p23 and that for efficient interaction of p23 with non-native proteins both the folded domain and the C-terminal unstructured region are required.  相似文献   

11.
We have previously determined the crystal structure of a non-structural 5B (NS5B) protein, an RNA-dependent RNA polymerase (RdRp) of hepatitis C virus (HCV). NS5B protein with the hydrophobic C-terminal 21 amino acid residues truncated, designated NS5B(570), shows a typical nucleotide polymerase structure resembling a right-hand shape. In the crystal structure, a C-terminal region between Leu545 and His562 occupies a putative RNA-binding cleft of this polymerase and seems to inhibit the polymerase activity. Varieties of recombinant NS5B proteins (NS5B(552), NS5B(544), NS5B(536) or NS5B(531), with C-terminal 39, 47, 55 or 60 amino acid residues truncated, respectively) were systematically constructed to elucidate effects of the region on the polymerase activity. NS5B(544), NS5B(536) and NS5B(531) showed markedly higher RdRp activities compared to the activities of NS5B(570) or NS5B(552). Furthermore, when the hydrophobic amino acid residues Leu547, Trp550 and Phe551 (LWF) in NS5B(570) and NS5B(552) were changed to alanine, their activities were higher than that of the original NS5B(570). The crystal structures of the various recombinant NS5B proteins were also determined. Structural comparison of the NS5B proteins indicates that the activation was caused by elimination of a unique hydrophobic interaction between the three C-terminal residues and a shallowly concave pocket consisting of thumb and palm domains.  相似文献   

12.
The carboxyl-terminal region of theras oncogene-encoded p21 protein is critical to the protein's function, since membrane binding through the C-terminus is necessary for its cellular activity. X-ray crystal structures for truncated p21 proteins are available, but none of these include the C-terminal region of the protein (from residues 172–189). Using conformational energy analysis, we determined the preferred three-dimensional structures for this C-terminal octadecapeptide of the H-ras oncogene p21 protein and generated these structures onto the crystal structure of the remainder of the protein. The results indicate that, like other membrane-associated proteins, the membrane-binding C-terminus of p21 assumes a helical hairpin conformation. In several low-energy orientations, the C-terminal structure is in close proximity to other critical locales of p21. These include the central transforming region (around Gln 61) and the amino terminal transforming region (around Gly 12), indicating that extracellular signals can be transduced through the C-terminal helical hairpin to the effector regions of the protein. This finding is consistent with the results of recent genetic experiments.  相似文献   

13.
Site-directed mutagenesis of smooth muscle myosin light chain kinase was applied to define its autoinhibitory domain. Mutants were all initiated at Leu-447 but contained varying lengths of C-terminal sequence. Those containing the complete C-terminal sequence to Glu-972 possessed kinase activities that were calmodulin-dependent. Removal of the putative inhibitory domain by truncation to Thr-778 resulted in generation of a constitutively active (calmodulin-independent) species. Thus, the inhibitory domain lies to the C-terminal side of Thr-778. Truncation to Lys-793 and to Trp-800 also resulted in constitutively active mutants, although the specific activity of the latter was less than the other mutants. None of the truncated mutants bound calmodulin. For each mutant, the Km values with respect to ATP and to the 20,000-dalton light chain were similar to values obtained with the native enzyme. The presence of the inhibitory domain was detected by activation of kinase activity following limited proteolysis with trypsin. Using this procedure, it was determined that the inhibitory domain was manifest only in the mutant truncated to Trp-800 and was absent from that ending at Lys-793. These results indicate that a critical region of the inhibitory domain is contained within the sequence Tyr-794 to Trp-800. This region overlaps with the calmodulin-binding site for five residues. Our assignment of the inhibitory sequence is consistent with autoinhibition via a pseudosubstrate domain.  相似文献   

14.
Chuang HH  Lin HY  Lin FP 《The FEBS journal》2008,275(9):2240-2254
The functional and structural significance of the C-terminal region of Bacillus licheniformis chitinase was explored using C-terminal truncation mutagenesis. Comparative studies between full-length and truncated mutant molecules included initial rate kinetics, fluorescence and CD spectrometric properties, substrate binding and hydrolysis abilities, thermostability, and thermodenaturation kinetics. Kinetic analyses revealed that the overall catalytic efficiency, k(cat)/K(m), was slightly increased for the truncated enzymes toward the soluble 4-methylumbelliferyl-N-N'-diacetyl chitobiose or 4-methylumbelliferyl-N-N'-N'-triacetyl chitotriose or insoluble alpha-chitin substrate. By contrast, changes to substrate affinity, K(m), and turnover rate, k(cat), varied considerably for both types of chitin substrates between the full-length and truncated enzymes. Both truncated enzymes exhibited significantly higher thermostabilities than the full-length enzyme. The truncated mutants retained similar substrate-binding specificities and abilities against the insoluble substrate but only had approximately 75% of the hydrolyzing efficiency of the full-length chitinase molecule. Fluorescence spectroscopy indicated that both C-terminal deletion mutants retained an active folding conformation similar to the full-length enzyme. However, a CD melting unfolding study was able to distinguish between the full-length and truncated mutant molecules by the two phases of apparent transition temperatures in the mutants. These results indicate that up to 145 amino acid residues, including the putative C-terminal chitin-binding region and the fibronectin (III) motif of B. licheniformis chitinase, could be removed without causing a seriously aberrant change in structure and a dramatic decrease in insoluble chitin hydrolysis. The results of the present study provide evidence demonstrating that the binding and hydrolyzing of insoluble chitin substrate for B. licheniformis chitinase was not dependent solely on the putative C-terminal chitin-binding region and the fibronectin (III) motif.  相似文献   

15.
Recombinant subunit IV mutants which identify the regions essential for restoration of bc(1) activity to the three-subunit core complex of Rhodobacter sphaeroides were generated and characterized. Four C-terminal truncated mutants: IV(1-109), IV(1-85), IV(1-76), and IV(1-40) had 100, 0, 0, and 0% of reconstitutive activity of the wild-type IV, indicating that residues 86-109 are essential. IV(1-109) is associated with the core complex in the same manner as the wild-type IV while mutants IV(1-85), IV(1-76), and IV(1-40) do not associate with the core complex, indicating that subunit IV requires its transmembrane helix region (residues 86-109) for assembly into the bc(1) complex. Since GST-IV(86-109) fusion protein has little reconstitutive activity, some region(s) in residues 1-85 are required for bc(1) activity restoration after subunit IV is incorporated into the complex through the transmembrane helix, presumably by interaction with cytochrome b in the core complex. The interacting regions are identified as residues 41-53 and 77-85, since mutants IV(21-109), IV(41-109), IV(54-109), and IV(77-109) had 95, 98, 53, and 53% of the reconstitutive activity of the wild-type IV. These two interacting regions are on the cytoplasmic side of the chromatophore membrane and closed to the DE loop and helix G of cytochrome b, respectively.  相似文献   

16.
Oryzacystatin, a cysteine proteinase inhibitor occurring in rice seeds, contains a particular glycine residue (Gly5) near the NH2-terminal position, and the sequence Gln53-Val54-Val55-Ala56-Gly57 in a central part of the molecule. Both are conserved among most members of the cystatin superfamily. We have found from Escherichia coli expression studies that the NH2-terminal 21 residues of oryzacystatin are not essential for its papain-inhibitory activity, and that the conserved pentapeptide region may be indispensable [Abe, K., Emori, Y., Kondo, H., Arai, S., & Suzuki, K. (1988) J. Biol. Chem. 263, 7655-7659]. Here we present more detailed data based on quantitative analyses of the inhibitory activities of NH2- and COOH-terminally truncated oryzacystatin and site-directed mutants at the Gln-Val-Val-Ala-Gly region. The data indicate the following results. (1) The truncated mutants lacking the NH2-terminal 21 residues or the COOH-terminal 11 residues exhibit potent papain-inhibitory activity equivalent to the activity of wild oryzacystatin. (2) However, neither the mutant lacking the NH2-terminal 38 residues nor that lacking the COOH-terminal 35 residues is completely able to inhibit papain. (3) Site-directed mutants at the Gln residue of the Gln-Val-Val-Ala-Gly region have drastically reduced papain-inhibitory activities: the Gln----Pro mutant is completely inactive and the Gln----Leu mutant has an approximately 150 times higher Ki value than wild-type oryzacystatin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
Many essential cellular proteins fold only with the assistance of chaperonin machines like the GroEL-GroES system of Escherichia coli. However, the mechanistic details of assisted protein folding by GroEL-GroES remain the subject of ongoing debate. We previously demonstrated that GroEL-GroES enhances the productive folding of a kinetically trapped substrate protein through unfolding, where both binding energy and the energy of ATP hydrolysis are used to disrupt the inhibitory misfolded states. Here, we show that the intrinsically disordered yet highly conserved C-terminal sequence of the GroEL subunits directly contributes to substrate protein unfolding. Interactions between the C terminus and the non-native substrate protein alter the binding position of the substrate protein on the GroEL apical surface. The C-terminal tails also impact the conformational state of the substrate protein during capture and encapsulation on the GroEL ring. Importantly, removal of the C termini results in slower overall folding, reducing the fraction of the substrate protein that commits quickly to a productive folding pathway and slowing several kinetically distinct folding transitions that occur inside the GroEL-GroES cavity. The conserved C-terminal tails of GroEL are thus important for protein folding from the beginning to the end of the chaperonin reaction cycle.  相似文献   

18.
The chaperonin GroEL contains two seven-subunit rings, and allosteric signals between them are required to complete the GroEL reaction cycle. For this reason SR1, a mutant of GroEL that forms only single rings, cannot function as a chaperone. Mutations in SR1 that restore chaperone function weaken its interaction with the cochaperonin GroES. We predicted that GroES mutants with reduced affinity for GroEL would also restore function to SR1. To test this, we mutated residues in GroES in and near its contact site with GroEL. Nearly half of the mutants showed partial function with SR1. Two mutants were confirmed to have reduced affinity for GroEL. Intriguingly, some GroES mutants were able to function with active single ring mutants of GroEL.  相似文献   

19.
Topologies of a substrate protein bound to the chaperonin GroEL   总被引:3,自引:0,他引:3  
The chaperonin GroEL assists polypeptide folding through sequential steps of binding nonnative protein in the central cavity of an open ring, via hydrophobic surfaces of its apical domains, followed by encapsulation in a hydrophilic cavity. To examine the binding state, we have classified a large data set of GroEL binary complexes with nonnative malate dehydrogenase (MDH), imaged by cryo-electron microscopy, to sort them into homogeneous subsets. The resulting electron density maps show MDH associated in several characteristic binding topologies either deep inside the cavity or at its inlet, contacting three to four consecutive GroEL apical domains. Consistent with visualization of bound polypeptide distributed over many parts of the central cavity, disulfide crosslinking could be carried out between a cysteine in a bound substrate protein and cysteines substituted anywhere inside GroEL. Finally, substrate binding induced adjustments in GroEL itself, observed mainly as clustering together of apical domains around sites of substrate binding.  相似文献   

20.
Vasodilator-stimulated phosphoprotein (VASP) is a member of the Ena/VASP family of proteins that are implicated in regulation of the actin cytoskeleton. All family members share a tripartite structural organization, comprising an N-terminal Ena/VASP homology (EVH) 1 domain, a more divergent proline-rich central part, and a common C-terminal EVH2 region of about 160-190 amino acids. Using chemical cross-linking, sucrose gradient sedimentation, and gel filtration analyses of different truncated VASP constructs, we demonstrate that the VASP EVH2 region is both necessary and sufficient for tetramerization. Moreover, co-sedimentation and fluorescent phalloidin staining showed that the EVH2 region binds and bundles F-actin in vitro and localizes to stress fibers in transfected cells. Analysis of the functional contribution of highly conserved blocks within this region indicated that residues 259-276 of human VASP are essential for the interaction with F-actin, whereas residues 343-380 are required for tetramerization, probably via coiled-coil formation. Interactions with F-actin are enhanced by VASP tetramerization. The results demonstrate that the C-terminal EVH2 segment is not only conserved in sequence but also forms a distinct functional entity. The data suggest that the EVH2 segment represents a novel oligomerization and F-actin binding domain.  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