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 922 毫秒
1.
Tk-subtilisin from the hyperthermophilic archaeon Thermococcus kodakarensis matures from Pro-Tk-subtilisin (Pro-TKS) upon autoprocessing and degradation of propeptide. Pro-TKS contains the insertion sequence (IS1) at the N-terminus of the mature domain as compared to bacterial pro-subtilisins. To analyze the role of IS1, the Pro-TKS derivative without IS1 (?IS1-Pro-TKS) and its active-site mutants (?IS1-Pro-S324A and ?IS1-Pro-S324C) were constructed and characterized. ?IS1-Pro-S324A and ?IS1-Pro-TKS represent an unautoprocessed and autoprocessed form of ?IS1-Pro-TKS, respectively. The CD and ANS fluorescence spectra of these proteins indicate that folding of ?IS1-Pro-TKS is not completed by binding of Ca2+ ions but is completed by the subsequent autoprocessing reaction. Thermal denaturation of these proteins analyzed by DSC and CD spectroscopy indicates that unautoprocessed ?IS1-Pro-TKS is less stable than autoprocessed ?IS1-Pro-TKS by 26.3?°C in T m. The stability of autoprocessed ?IS1-Pro-TKS is comparable to that of Pro-TKS, which is slightly lower than that of unautoprocessed Pro-TKS. These results suggest that ?IS1-Pro-TKS is fully folded and greatly stabilized by autoprocessing. ?IS1-Pro-TKS more slowly matured to ?IS1-Tk-subtilisin than Pro-TKS did, due to a decrease in the autoprocessing rate. We propose that IS1 is required not only for hyperstabilization of Pro-TKS but also for its rapid maturation.  相似文献   

2.
The crystal structure of an active site mutant of pro-Tk-subtilisin (pro-S324A) from the hyperthermophilic archaeon Thermococcus kodakaraensis was determined at 2.3 A resolution. The overall structure of this protein is similar to those of bacterial subtilisin-propeptide complexes, except that the peptide bond linking the propeptide and mature domain contacts with the active site, and the mature domain contains six Ca2+ binding sites. The Ca-1 site is conserved in bacterial subtilisins but is formed prior to autoprocessing, unlike the corresponding sites of bacterial subtilisins. All other Ca2+-binding sites are unique in the pro-S324A structure and are located at the surface loops. Four of them apparently contribute to the stability of the central alphabetaalpha substructure of the mature domain. The CD spectra, 1-anilino-8-naphthalenesulfonic acid fluorescence spectra, and sensitivities to chymotryptic digestion of this protein indicate that the conformation of pro-S324A is changed from an unstable molten globule-like structure to a stable native one upon Ca2+ binding. Another active site mutant, pro-S324C, was shown to be autoprocessed to form a propeptide-mature domain complex in the presence of Ca2+. The CD spectra of this protein indicate that the structure of pro-S324C is changed upon Ca2+ binding like pro-S324A but is not seriously changed upon subsequent autoprocessing. These results suggest that the maturation process of Tk-subtilisin is different from that of bacterial subtilisins in terms of the requirement of Ca2+ for folding of the mature domain and completion of the folding process prior to autoprocessing.  相似文献   

3.
Tk-subtilisin (a subtilisin homologue from Thermococcus kodakaraensis) is matured from Pro-Tk-subtilisin upon autoprocessing and degradation of Tk-propeptide. To analyze the folding mechanism of Tk-subtilisin, the crystal structure of the active site mutant of Tk-subtilisin (S324A-subtilisin), which was refolded in the presence of Ca2+ and absence of Tk-propeptide, was determined at 2.16 Å resolution. This structure is essentially the same as that of Tk-subtilisin matured from Pro-Tk-subtilisin. S324A-subtilisin was refolded with a rate constant of 0.17 and 1.8 min−1 at 30 °C in the absence and presence of Tk-propeptide, respectively, indicating that Tk-subtilisin does not require Tk-propeptide for folding but requires it for acceleration of folding.  相似文献   

4.
Tk-subtilisin [the mature domain of Pro-Tk-subtilisin in active form (Gly70-Gly398)] from the hyperthermophilic archaeon Thermococcus kodakaraensis is matured from Pro-Tk-subtilisin [a subtilisin homologue from T. kodakaraensis in pro form (Gly1-Gly398)] upon autoprocessing and degradation of propeptide. Pro-Tk-subtilisin is characterized by extremely slow maturation at mild temperatures, but this maturation rate is greatly increased by a single Gly56 → Ser mutation in the propeptide region. To analyze the role of Gly56, which assumes a left-handed conformation, Pro-Tk-subtilisin variants with complete amino acid substitutions at Gly56 were constructed. A comparison of their halo-forming activities suggests that all variants, except for Pro-G56W [Pro-G56X, Pro-Tk-subtilisin with Gly56 → X mutation (X = any amino acid)], mature faster than WT. Pro-G56W and Pro-G56E with the lowest and highest maturation rates, respectively, among 19 variants, as well as WT and Pro-G56S, were overproduced, purified, and characterized. SDS-PAGE analyses and Tk-subtilisin activity assay indicated that their maturation rates increased in the order WT ≤ Pro-G56W < Pro-G56S < Pro-G56E. The propeptides of these variants were also overproduced, purified, and characterized. The stability and inhibitory potency of these propeptides decreased in the order Tk-propeptide [propeptide of Tk-subtilisin (Gly1-Leu69)] ≥ G56W-propeptide > G56S-propeptide > G56E-propeptide, indicating that they are inversely correlated with the maturation rates of Pro7-Tk-subtilisin and its derivatives. The crystal structures of these propeptides determined in complex with S324A-subtilisin indicate that the conformation of the propeptide is altered by the mutation, such that nonglycine residues at position 56 assume a right-handed conformation and hydrophobic interactions at the core region decrease. These results indicate that Gly56 is required in stabilizing the propeptide fold. Stabilization of this fold leads to strong binding of Tk-propeptide to Tk-subtilisin, high resistance of Tk-propeptide to proteolytic degradation, and slow maturation of Pro-Tk-subtilisin.  相似文献   

5.
Subtilisin is produced as a precursor that requires its N-terminal propeptide to chaperone the folding of its protease domain. Once folded, subtilisin adopts a remarkably stable conformation, which has been attributed to a high affinity Ca(2+) binding site. We investigated the role of the metal ligand in the maturation of pro-subtilisin, a process that involves folding, autoprocessing and partial degradation. Our results establish that although Ca(2+) ions can stabilize the protease domain, the folding and autoprocessing of pro-subtilisin take place independent of Ca(2+) ion. We demonstrate that the stabilizing effect of calcium is observed only after the completion of autoprocessing and that the metal ion appears to be responsible for shifting the folding equilibrium towards the native conformation in both mature subtilisin and the autoprocessed propeptide:subtilisin complex. Furthermore, the addition of active subtilisin to unautoprocessed pro-subtilisin in trans does not facilitate precursor maturation, but rather promotes rapid autodegradation. The primary cleavage site that initiates this autodegradation is at Gln19 in the N-terminus of mature subtilisin. This corresponds to the loop that links alpha-helix-2 and beta-strand-1 in mature subtilisin and has indirect effects on the formation of the Ca(2+) binding site. Our results show that the N-terminus of mature subtilisin undergoes rearrangement subsequent to propeptide autoprocessing. Since this structural change enhances the proteolytic stability of the precursor, our results suggest that the autoprocessing reaction must be completed before the release of active subtilisin in order to maximize folding efficiency.  相似文献   

6.
Uehara R  Takeuchi Y  Tanaka S  Takano K  Koga Y  Kanaya S 《Biochemistry》2012,51(26):5369-5378
Tk-subtilisin, a hyperthermostable subtilisin-like serine protease from Thermococcus kodakarensis, matures from the inactive precursor, Pro-Tk-subtilisin (Pro-TKS), upon autoprocessing and degradation of the propeptide (Tkpro). It contains seven Ca(2+) ions. Four of them (Ca2-Ca5) are responsible for folding of Tk-subtilisin. In this study, to clarify the role of the other three Ca(2+) ions (Ca1, Ca6, and Ca7), we constructed Pro-TKS derivatives lacking the Ca1 ion (Pro-TKS/ΔCa1), Ca6 ion (Pro-TKS/ΔCa6), and Ca7 ion (Pro-TKS/ΔCa7), and their active site mutants (Pro-S324A/ΔCa1, Pro-S324A/ΔCa6, and Pro-S324A/ΔCa7, respectively). Pro-TKS/ΔCa6 and Pro-TKS/ΔCa7 fully matured into their active forms upon incubation at 80 °C for 30 min as did Pro-TKS. The mature enzymes were as active as Tk-subtilisin at 80 °C, indicating that the Ca6 and Ca7 ions are not important for activity. In contrast, Pro-TKS/ΔCa1 matured poorly at 80 °C because of the instability of its mature domain. The enzymatic activity of Tk-subtilisin/ΔCa1 was determined to be 50% of that of Tk-subtilisin using the refolded protein. This result suggests that the Ca1 ion is required for the maximal activity of Tk-subtilisin. The refolding rates of all Pro-S324A derivatives were comparable to that of Pro-S324A (active site mutant of Pro-TKS), indicating that these Ca(2+) ions are not needed for folding of Tk-subtilisin. The stabilities of Pro-S324A/ΔCa1 and Pro-S324A/ΔCa6 were decreased by 26.6 and 11.7 °C, respectively, in T(m) compared to that of Pro-S324A. The half-lives of Tk-subtilisin/ΔCa6 and Tk-subtilisin/ΔCa7 at 95 °C were 8- and 4-fold lower than that of Tk-subtilisin, respectively. These results suggest that the Ca1, Ca6, and Ca7 ions, especially the Ca1 ion, contribute to the hyperthermostabilization of Tk-subtilisin.  相似文献   

7.
Severe acute respiratory syndrome (SARS) is an emerging infectious disease caused by a novel human coronavirus. Viral maturation requires a main protease (3CL(pro)) to cleave the virus-encoded polyproteins. We report here that the 3CL(pro) containing additional N- and/or C-terminal segments of the polyprotein sequences undergoes autoprocessing and yields the mature protease in vitro. The dimeric three-dimensional structure of the C145A mutant protease shows that the active site of one protomer binds with the C-terminal six amino acids of the protomer from another asymmetric unit, mimicking the product-bound form and suggesting a possible mechanism for maturation. The P1 pocket of the active site binds the Gln side chain specifically, and the P2 and P4 sites are clustered together to accommodate large hydrophobic side chains. The tagged C145A mutant protein served as a substrate for the wild-type protease, and the N terminus was first digested (55-fold faster) at the Gln(-1)-Ser1 site followed by the C-terminal cleavage at the Gln306-Gly307 site. Analytical ultracentrifuge of the quaternary structures of the tagged and mature proteases reveals the remarkably tighter dimer formation for the mature enzyme (K(d) = 0.35 nm) than for the mutant (C145A) containing 10 extra N-terminal (K(d) = 17.2 nM) or C-terminal amino acids (K(d) = 5.6 nM). The data indicate that immature 3CL(pro) can form dimer enabling it to undergo autoprocessing to yield the mature enzyme, which further serves as a seed for facilitated maturation. Taken together, this study provides insights into the maturation process of the SARS 3CL(pro) from the polyprotein and design of new structure-based inhibitors.  相似文献   

8.
GABARAP recognizes and binds the gamma2 subunit of the GABA(A) receptor, interacts with microtubules and the N-ethyl maleimide sensitive factor, and is proposed to function in GABA(A) receptor trafficking and postsynaptic localization. We have determined the crystal structure of human GABARAP at 1.6 A resolution. The structure comprises an N-terminal helical subdomain and a ubiquitin-like C-terminal domain. Structure-based mutational analysis demonstrates that the N-terminal subdomain is responsible for tubulin binding while the C-terminal domain contains the binding site for the GABA(A). A second GABARAP crystal form was determined at 1.9 A resolution and documents that GABARAP can self-associate in a head-to-tail manner. The structural details of this oligomerization reveal how GABARAP can both promote tubulin polymerization and facilitate GABA(A) receptor clustering.  相似文献   

9.
Substrate selectivity and proteolytic activity for the E. coli ATP-dependent protease, ClpAP, is modulated by an adaptor protein, ClpS. ClpS binds to ClpA, the regulatory component of the ClpAP complex. We report the crystal structure of ClpS in complex with the isolated N-terminal domain of ClpA in two different crystal forms at 2.3- and 3.3-A resolution. The ClpS structure forms an alpha/beta-sandwich and is topologically analogous to the C-terminal domain of the ribosomal protein L7/L12. ClpS contacts two surfaces on the N-terminal domain in both crystal forms; the more extensive interface was shown to be favored in solution by protease protection experiments. The N-terminal 20 residues of ClpS are not visible in the crystal structures; the removal of the first 17 residues produces ClpSDeltaN, which binds to the ClpA N-domain but no longer inhibits ClpA activity. A zinc binding site involving two His and one Glu residue was identified crystallographically in the N-terminal domain of ClpA. In a model of ClpS bound to hexameric ClpA, ClpS is oriented with its N terminus directed toward the distal surface of ClpA, suggesting that the N-terminal region of ClpS may affect productive substrate interactions at the apical surface or substrate entry into the ClpA translocation channel.  相似文献   

10.
The crystal structure of the P1/Mahoney poliovirus empty capsid has been determined at 2.9 A resolution. The empty capsids differ from mature virions in that they lack the viral RNA and have yet to undergo a stabilizing maturation cleavage of VP0 to yield the mature capsid proteins VP4 and VP2. The outer surface and the bulk of the protein shell are very similar to those of the mature virion. The major differences between the 2 structures are focused in a network formed by the N-terminal extensions of the capsid proteins on the inner surface of the shell. In the empty capsids, the entire N-terminal extension of VP1, as well as portions corresponding to VP4 and the N-terminal extension of VP2, are disordered, and many stabilizing interactions that are present in the mature virion are missing. In the empty capsid, the VP0 scissile bond is located some 20 A away from the positions in the mature virion of the termini generated by VP0 cleavage. The scissile bond is located on the rim of a trefoil-shaped depression in the inner surface of the shell that is highly reminiscent of an RNA binding site in bean pod mottle virus. The structure suggests plausible (and ultimately testable) models for the initiation of encapsidation, for the RNA-dependent autocatalytic cleavage of VP0, and for the role of the cleavage in establishing the ordered N-terminal network and in generating stable virions.  相似文献   

11.
Glycine N-methyltransferase (S-adenosyl-l-methionine: glycine methyltransferase, EC 2.1.1.20; GNMT) catalyzes the AdoMet-dependent methylation of glycine to form sarcosine (N-methylglycine). Unlike most methyltransferases, GNMT is a tetrameric protein showing a positive cooperativity in AdoMet binding and weak inhibition by S-adenosylhomocysteine (AdoHcy). The first crystal structure of GNMT complexed with AdoMet showed a unique "closed" molecular basket structure, in which the N-terminal section penetrates and corks the entrance of the adjacent subunit. Thus, the apparent entrance or exit of the active site is not recognizable in the subunit structure, suggesting that the enzyme must possess a second, enzymatically active, "open" structural conformation. A new crystalline form of the R175K enzyme has been grown in the presence of an excess of AdoHcy, and its crystal structure has been determined at 3.0 A resolution. In this structure, the N-terminal domain (40 amino acid residues) of each subunit has moved out of the active site of the adjacent subunit, and the entrances of the active sites are now opened widely. An AdoHcy molecule has entered the site occupied in the "closed" structure by Glu15 and Gly16 of the N-terminal domain of the adjacent subunit. An AdoHcy binds to the consensus AdoMet binding site observed in the other methyltransferase. This AdoHcy binding site supports the glycine binding site (Arg175) deduced from a chemical modification study and site-directed mutagenesis (R175K). The crystal structures of WT and R175K enzymes were also determined at 2.5 A resolution. These enzyme structures have a closed molecular basket structure and are isomorphous to the previously determined AdoMet-GNMT structure. By comparing the open structure to the closed structure, mechanisms for auto-inhibition and for the forced release of the product AdoHcy have been revealed in the GNMT structure. The N-terminal section of the adjacent subunit occupies the AdoMet binding site and thus inhibits the methyltransfer reaction, whereas the same N-terminal section forces the departure of the potentially potent inhibitor AdoHcy from the active site and thus facilitates the methyltransfer reaction. Consequently GNMT is less active at a low level of AdoMet concentration, and is only weakly inhibited by AdoHcy. These properties of GNMT are particularly suited for regulation of the cellular AdoMet/AdoHcy ratio.  相似文献   

12.
The precursor mRNA retention and splicing (RES) complex mediates nuclear retention and enhances splicing of precursor mRNAs. The RES complex from yeast comprises three proteins, Snu17p, Bud13p and Pml1p. Snu17p acts as a central platform that concomitantly binds the Bud13p and Pml1p subunits via short peptide epitopes. As a step to decipher the molecular architecture of the RES complex, we have determined crystal structures of full-length Pml1p and N-terminally truncated Pml1p. The first 50 residues of full-length Pml1p, encompassing the Snu17p-binding region, are disordered, showing that Pml1p binds to Snu17p via an intrinsically unstructured region. The remainder of Pml1p folds as a forkhead-associated (FHA) domain, which is expanded by a number of noncanonical elements compared with known FHA domains from other proteins. An atypical N-terminal appendix runs across one β-sheet and thereby stabilizes the domain as shown by deletion experiments. FHA domains are thought to constitute phosphopeptide-binding elements. Consistently, a sulfate ion was found at the putative phosphopeptide-binding loops of full-length Pml1p. The N-terminally truncated version of the protein lacked a similar phosphopeptide mimic but retained an almost identical structure. A long loop neighboring the putative phosphopeptide-binding site was disordered in both structures. Comparison with other FHA domain proteins suggests that this loop adopts a defined conformation upon ligand binding and thereby confers ligand specificity. Our results show that in the RES complex, an FHA domain of Pml1p is flexibly tethered via an unstructured N-terminal region to Snu17p.  相似文献   

13.
PA protease (pro-aminopeptidase processing protease) is an extracellular zinc metalloprotease produced by the Gram-negative bacterium Aeromonas caviae T-64. The 590-amino-acid precursor of PA protease is composed of a putative 19-amino-acid signal sequence, a 165-amino-acid N-terminal propeptide, a 33 kDa mature protease domain and an 11 kDa C-terminal propeptide. The proform of PA protease, which was produced as inclusion bodies in Escherichia coli, was subjected to in vitro refolding. It was revealed that the processing of the proform involved a stepwise autoprocessing mechanism. Firstly, the N-terminal propeptide was autocatalytically removed on completion of refolding and secondly, the C-terminal propeptide was autoprocessed after the degradation of the N-terminal propeptide. Both the N- and C-terminal propeptides existed as intact peptides after their successive removal, and they were subsequently degraded gradually. The degradation of the N-terminal propeptide appears to be the rate-limiting step in the maturation of the proform of PA protease.  相似文献   

14.
15.
Aldehyde dehydrogenases catalyze the oxidation of aldehyde substrates to the corresponding carboxylic acids. Lactaldehyde dehydrogenase from Escherichia coli (aldA gene product, P25553) is an NAD(+)-dependent enzyme implicated in the metabolism of l-fucose and l-rhamnose. During the heterologous expression and purification of taxadiene synthase from the Pacific yew, lactaldehyde dehydrogenase from E. coli was identified as a minor (相似文献   

16.
Two crystal structures of yeast translation elongation factor 2 (eEF2) were determined: the apo form at 2.9 A resolution and eEF2 in the presence of the translocation inhibitor sordarin at 2.1 A resolution. The overall conformation of apo eEF2 is similar to that of its prokaryotic homolog elongation factor G (EF-G) in complex with GDP. Upon sordarin binding, the three tRNA-mimicking C-terminal domains undergo substantial conformational changes, while the three N-terminal domains containing the nucleotide-binding site form an almost rigid unit. The conformation of eEF2 in complex with sordarin is entirely different from known conformations observed in crystal structures of EF-G or from cryo-EM studies of EF-G-70S complexes. The domain rearrangements induced by sordarin binding and the highly ordered drug-binding site observed in the eEF2-sordarin structure provide a high-resolution structural basis for the mechanism of sordarin inhibition. The two structures also emphasize the dynamic nature of the ribosomal translocase.  相似文献   

17.
Bacterial tyrosyl-tRNA synthetases (TyrRS) possess a flexibly linked C-terminal domain of approximately 80 residues, which has hitherto been disordered in crystal structures of the enzyme. We have determined the structure of Thermus thermophilus TyrRS at 2.0 A resolution in a crystal form in which the C-terminal domain is ordered, and confirm that the fold is similar to part of the C-terminal domain of ribosomal protein S4. We have also determined the structure at 2.9 A resolution of the complex of T.thermophilus TyrRS with cognate tRNA(tyr)(G Psi A). In this structure, the C-terminal domain binds between the characteristic long variable arm of the tRNA and the anti-codon stem, thus recognizing the unique shape of the tRNA. The anticodon bases have a novel conformation with A-36 stacked on G-34, and both G-34 and Psi-35 are base-specifically recognized. The tRNA binds across the two subunits of the dimeric enzyme and, remarkably, the mode of recognition of the class I TyrRS for its cognate tRNA resembles that of a class II synthetase in being from the major groove side of the acceptor stem.  相似文献   

18.
Gupta GD  Makde RD  Rao BJ  Kumar V 《The FEBS journal》2008,275(16):4235-4249
Translin protein is highly conserved in eukaryotes. Human translin binds both ssDNA and RNA. Its nucleic acid binding site results from a combination of basic regions in the octameric structure. We report here the first biochemical characterization of wild-type Drosophila melanogaster (drosophila) translin and a chimeric translin, and present 3.5 A resolution crystal structures of drosophila P168S mutant translin from two crystal forms. The wild-type drosophila translin most likely exists as an octamer/decamer, and binds to the ssDNA Bcl-CL1 sequence. In contrast, ssDNA binding-incompetent drosophila P168S mutant translin exists as a tetramer. The structures of the mutant translin are identical in both crystal forms, and their C-terminal residues are disordered. The chimeric protein, possessing two nucleic acid binding motifs of drosophila translin, the C-terminal residues of human translin, and serine at position 168, attains the octameric state and binds to ssDNA. The present studies suggest that the oligomeric status of translin critically influences the DNA binding properties of translin proteins.  相似文献   

19.
Protein glutaminase, which converts a protein glutamine residue to a glutamate residue, is expected to be useful as a new food-processing enzyme. The crystal structures of the mature and pro forms of the enzyme were refined at 1.15 and 1.73 ? resolution, respectively. The overall structure of the mature enzyme has a weak homology to the core domain of human transglutaminase-2. The catalytic triad (Cys-His-Asp) common to transglutaminases and cysteine proteases is located in the bottom of the active site pocket. The structure of the recombinant pro form shows that a short loop between S2 and S3 in the proregion covers and interacts with the active site of the mature region, mimicking the protein substrate of the enzyme. Ala-47 is located just above the pocket of the active site. Two mutant structures (A47Q-1 and A47Q-2) refined at 1.5 ? resolution were found to correspond to the enzyme-substrate complex and an S-acyl intermediate. Based on these structures, the catalytic mechanism of protein glutaminase is proposed.  相似文献   

20.
Replication protein A (RPA) is the ubiquitous, eukaryotic single-stranded DNA (ssDNA) binding protein and is essential for DNA replication, recombination, and repair. Here, crystal structures of the soluble RPA heterodimer, composed of the RPA14 and RPA32 subunits, have been determined for the full-length protein in multiple crystal forms. In all crystals, the electron density for the N-terminal (residues 1-42) and C-terminal (residues 175-270) regions of RPA32 is weak and of poor quality indicating that these regions are disordered and/or assume multiple positions in the crystals. Hence, the RPA32 N terminus, that is hyperphosphorylated in a cell-cycle-dependent manner and in response to DNA damaging agents, appears to be inherently disordered in the unphosphorylated state. The C-terminal, winged helix-loop-helix, protein-protein interaction domain adopts several conformations perhaps to facilitate its interaction with various proteins. Although the ordered regions of RPA14/32 resemble the previously solved protease-resistant core crystal structure, the quaternary structures between the heterodimers are quite different. Thus, the four-helix bundle quaternary assembly noted in the original core structure is unlikely to be related to the quaternary structure of the intact heterotrimer. An organic ligand binding site between subunits RPA14 and RPA32 was identified to bind dioxane. Comparison of the ssDNA binding surfaces of RPA70 with RPA14/32 showed that the lower affinity of RPA14/32 can be attributed to a shallower binding crevice with reduced positive electrostatic charge.  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