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.
A conserved structural module following the KMSKS catalytic loop exhibits α-α-β-α topology in class Ia and Ib aminoacyl-tRNA synthetases. However, the function of this domain has received little attention. Here, we describe the effect this module has on the aminoacylation and editing capacities of leucyl-tRNA synthetases (LeuRSs) by characterizing the key residues from various species. Mutation of highly conserved basic residues on the third α-helix of this domain impairs the affinity of LeuRS for the anticodon stem of tRNALeu, which decreases both aminoacylation and editing activities. Two glycine residues on this α-helix contribute to flexibility, leucine activation, and editing of LeuRS from Escherichia coli (EcLeuRS). Acidic residues on the β-strand enhance the editing activity of EcLeuRS and sense the size of the tRNALeu D-loop. Incorporation of these residues stimulates the tRNA-dependent editing activity of the chimeric minimalist enzyme Mycoplasma mobile LeuRS fused to the connective polypeptide 1 editing domain and leucine-specific domain from EcLeuRS. Together, these results reveal the stem contact-fold to be a functional as well as a structural linker between the catalytic site and the tRNA binding domain. Sequence comparison of the EcLeuRS stem contact-fold domain with editing-deficient enzymes suggests that key residues of this module have evolved an adaptive strategy to follow the editing functions of LeuRS.  相似文献   

2.
Cyclodipeptide synthases (CDPSs) use two aminoacyl-tRNA substrates in a sequential ping-pong mechanism to form a cyclodipeptide. The crystal structures of three CDPSs have been determined and all show a Rossmann-fold domain similar to the catalytic domain of class-I aminoacyl-tRNA synthetases (aaRSs). Structural features and mutational analyses however suggest that CDPSs and aaRSs interact differently with their tRNA substrates. We used AlbC from Streptomyces noursei that mainly produces cyclo(l-Phe-l-Leu) to investigate the interaction of a CDPS with its substrates. We demonstrate that Phe-tRNAPhe is the first substrate accommodated by AlbC. Its binding to AlbC is dependent on basic residues located in the helix α4 that form a basic patch at the surface of the protein. AlbC does not use all of the Leu-tRNALeu isoacceptors as a second substrate. We show that the G1-C72 pair of the acceptor stem is essential for the recognition of the second substrate. Substitution of D163 located in the loop α6–α7 or D205 located in the loop β6–α8 affected Leu-tRNALeu isoacceptors specificity, suggesting the involvement of these residues in the binding of the second substrate. This is the first demonstration that the two substrates of CDPSs are accommodated in different binding sites.  相似文献   

3.
Proteins that participate in the import of cytosolic tRNAs into mitochondria have been identified in several eukaryotic species, but the details of their interactions with tRNA and other proteins are unknown. In the kinetoplastid protozoon Leishmania tropica, multiple proteins are organized into a functional import complex. RIC8A, a tRNA-binding subunit of this complex, has a C-terminal domain that functions as subunit 6b of ubiquinol cytochrome c reductase (complex III). We show that the N-terminal domain, unique to kinetoplastid protozoa, is structurally similar to the appended S15/NS1 RNA-binding domain of aminoacyl tRNA synthetases, with a helix–turn–helix motif. Structure-guided mutagenesis coupled with in vitro assays showed that helix α1 contacts tRNA whereas helix α2 targets the protein for assembly into the import complex. Inducible expression of a helix 1-deleted variant in L. tropica resulted in formation of an inactive import complex, while the helix 2-deleted variant was unable to assemble in vivo. Moreover, a protein-interaction assay showed that the C-terminal domain makes allosteric contacts with import receptor RIC1 complexed with tRNA. These results help explain the origin of the bifunctionality of RIC8A, and the allosteric changes accompanying docking and release of tRNA during import.  相似文献   

4.
F1-ATPase is the catalytic complex of rotary nanomotor ATP synthases. Bacterial ATP synthases can be autoinhibited by the C-terminal domain of subunit ϵ, which partially inserts into the enzyme''s central rotor cavity to block functional subunit rotation. Using a kinetic, optical assay of F1·ϵ binding and dissociation, we show that formation of the extended, inhibitory conformation of ϵ (ϵX) initiates after ATP hydrolysis at the catalytic dwell step. Prehydrolysis conditions prevent formation of the ϵX state, and post-hydrolysis conditions stabilize it. We also show that ϵ inhibition and ADP inhibition are distinct, competing processes that can follow the catalytic dwell. We show that the N-terminal domain of ϵ is responsible for initial binding to F1 and provides most of the binding energy. Without the C-terminal domain, partial inhibition by the ϵ N-terminal domain is due to enhanced ADP inhibition. The rapid effects of catalytic site ligands on conformational changes of F1-bound ϵ suggest dynamic conformational and rotational mobility in F1 that is paused near the catalytic dwell position.  相似文献   

5.
DNA replication of phage-plasmid P4 in its host Escherichia coli depends on its replication protein α. In the plasmid state, P4 copy number is controlled by the regulator protein Cnr (copy number regulation). Mutations in α (αcr) that prevent regulation by Cnr cause P4 over-replication and cell death. Using the two-hybrid system in Saccharomyces cerevisiae and a system based on λ immunity in E.coli for in vivo detection of protein–protein interactions, we found that: (i) α protein interacts with Cnr, whereas αcr proteins do not; (ii) both α–α and αcr–αcr interactions occur and the interaction domain is located within the C-terminal of α; (iii) Cnr–Cnr interaction also occurs. Using an in vivo competition assay, we found that Cnr interferes with both α–α and αcr–αcr dimerization. Our data suggest that Cnr and α interact in at least two ways, which may have different functional roles in P4 replication control.  相似文献   

6.
In sequenced genomes, genes belonging to the cluster of orthologous group COG1041 are exclusively, and almost ubiquitously, found in Eukaryota and Archaea but never in Bacteria. The corresponding gene products exhibit a characteristic Rossmann fold, S-adenosylmethionine-dependent methyltransferase domain in the C terminus and a predicted RNA-binding THUMP (thiouridine synthases, RNA methyltransferases, and pseudouridine synthases) domain in the N terminus. Recombinant PAB1283 protein from the archaeon Pyrococcus abyssi GE5, a member of COG1041, was purified and shown to behave as a monomeric 39-kDa entity. This protein (EC 2.1.1.32), now renamed (Pab)Trm-G10, which is extremely thermostable, forms a 1:1 complex with tRNA and catalyzes the adenosylmethionine-dependent methylation of the exocyclic amino group (N(2)) of guanosine located at position 10. Depending on the experimental conditions used, as well as the tRNA substrate tested, the enzymatic reaction leads to the formation of either N(2)-monomethyl (m(2)G) or N(2)-dimethylguanosine (m(2)(2)G). Interestingly, (Pab)Trm-G10 exhibits different domain organization and different catalytic site architecture from another, earlier characterized, tRNA-dimethyltransferase from Pyrococcus furiosus ((Pfu)Trm-G26, also known as (Pfu)Trm1, a member of COG1867) that catalyzes an identical two-step dimethylation of guanosine but at position 26 in tRNAs and is also conserved among all sequenced Eukaryota and Archaea. The co-occurrence of these two guanosine dimethyltransferases in both Archaea and Eukaryota but not in Bacteria is a hallmark of distinct tRNAs maturation strategies between these domains of life.  相似文献   

7.
In archaeal rRNAs, the isomerization of uridine into pseudouridine (Ψ) is achieved by the H/ACA sRNPs and the minimal set of proteins required for RNA:Ψ-synthase activity is the aCBF5–aNOP10 protein pair. The crystal structure of the aCBF5–aNOP10 heterodimer from Pyrococcus abyssi was solved at 2.1 Å resolution. In this structure, protein aNOP10 has an extended shape, with a zinc-binding motif at the N-terminus and an α-helix at the C-terminus. Both motifs contact the aCBF5 catalytic domain. Although less efficiently as does the full-length aNOP10, the aNOP10 C-terminal domain binds aCBF5 and stimulates the RNA-guided activity. We show that the C-terminal domain of aCBF5 (the PUA domain), which is wrapped by an N-terminal extension of aCBF5, plays a crucial role for aCBF5 binding to the guide sRNA. Addition of this domain in trans partially complement particles assembled with an aCBF5ΔPUA truncated protein. In the crystal structure, the aCBF5–aNOP10 complex forms two kinds of heterotetramers with parallel and perpendicular orientations of the aNOP10 terminal α-helices, respectively. By gel filtration assay, we showed that aNOP10 can dimerize in solution. As both residues Y41 and L48 were needed for dimerization, the dimerization likely takes place by interaction of parallel α-helices.  相似文献   

8.
The pathogenesis of Alzheimer disease appears to be strongly linked to the aggregation of amyloid-β (Aβ) peptide and, especially, formation of soluble Aβ1–42 oligomers. It was recently demonstrated that the cellular prion protein, PrPC, binds with high affinity to these oligomers, acting as a putative receptor that mediates at least some of their neurotoxic effects. Here we show that the soluble (i.e. glycophosphatidylinositol anchor-free) prion protein and its N-terminal fragment have a strong effect on the aggregation pathway of Aβ1–42, inhibiting its assembly into amyloid fibrils. Furthermore, the prion protein prevents formation of spherical oligomers that normally occur during Aβ fibrillogenesis, acting as a potent inhibitor of Aβ1–42 toxicity as assessed in experiments with neuronal cell culture. These findings may provide a molecular level foundation to explain the reported protective action of the physiologically released N-terminal N1 fragment of PrPC against Aβ neurotoxicity. They also suggest a novel approach to pharmacological intervention in Alzheimer disease.  相似文献   

9.
α-1,4-Glucan lyase (EC 4.2.2.13) from the red seaweed Gracilariopsis lemaneiformis cleaves α-1,4-glucosidic linkages in glycogen, starch, and malto-oligosaccharides, yielding the keto-monosaccharide 1,5-anhydro-d-fructose. The enzyme belongs to glycoside hydrolase family 31 (GH31) but degrades starch via an elimination reaction instead of hydrolysis. The crystal structure shows that the enzyme, like GH31 hydrolases, contains a (β/α)8-barrel catalytic domain with B and B′ subdomains, an N-terminal domain N, and the C-terminal domains C and D. The N-terminal domain N of the lyase was found to bind a trisaccharide. Complexes of the enzyme with acarbose and 1-dexoynojirimycin and two different covalent glycosyl-enzyme intermediates obtained with fluorinated sugar analogues show that, like GH31 hydrolases, the aspartic acid residues Asp553 and Asp665 are the catalytic nucleophile and acid, respectively. However, as a unique feature, the catalytic nucleophile is in a position to act also as a base that abstracts a proton from the C2 carbon atom of the covalently bound subsite −1 glucosyl residue, thus explaining the unique lyase activity of the enzyme. One Glu to Val mutation in the active site of the homologous α-glucosidase from Sulfolobus solfataricus resulted in a shift from hydrolytic to lyase activity, demonstrating that a subtle amino acid difference can promote lyase activity in a GH31 hydrolase.  相似文献   

10.
During blood coagulation, the protease factor XIa (fXIa) activates factor IX (fIX). We describe a new mechanism for this process. FIX is cleaved initially after Arg145 to form fIXα, and then after Arg180 to form the protease fIXaβ. FIXα is released from fXIa, and must rebind for cleavage after Arg180 to occur. Catalytic efficiency of cleavage after Arg180 is 7-fold greater than for cleavage after Arg145, limiting fIXα accumulation. FXIa contains four apple domains (A1–A4) and a catalytic domain. Exosite(s) on fXIa are required for fIX binding, however, there is lack of consensus on their location(s), with sites on the A2, A3, and catalytic domains described. Replacing the A3 domain with the prekallikrein A3 domain increases Km for fIX cleavage after Arg145 and Arg180 25- and ≥90-fold, respectively, and markedly decreases kcat for cleavage after Arg180. Similar results were obtained with the isolated fXIa catalytic domain, or fXIa in the absence of Ca2+. Forms of fXIa lacking the A3 domain exhibit 15-fold lower catalytic efficiency for cleavage after Arg180 than for cleavage after Arg145, resulting in fIXα accumulation. Replacing the A2 domain does not affect fIX activation. The results demonstrate that fXIa activates fIX by an exosite- and Ca2+-mediated release-rebind mechanism in which efficiency of the second cleavage is enhanced by conformational changes resulting from the first cleavage. Initial binding of fIX and fIXα requires an exosite on the fXIa A3 domain, but not the A2 or catalytic domain.  相似文献   

11.
Exo-1,5-α-l-arabinofuranosidases belonging to glycoside hydrolase family 43 have strict substrate specificity. These enzymes hydrolyze only the α-1,5-linkages of linear arabinan and arabino-oligosaccharides in an exo-acting manner. The enzyme from Streptomyces avermitilis contains a core catalytic domain belonging to glycoside hydrolase family 43 and a C-terminal arabinan binding module belonging to carbohydrate binding module family 42. We determined the crystal structure of intact exo-1,5-α-l-arabinofuranosidase. The catalytic module is composed of a 5-bladed β-propeller topologically identical to the other family 43 enzymes. The arabinan binding module had three similar subdomains assembled against one another around a pseudo-3-fold axis, forming a β-trefoil-fold. A sugar complex structure with α-1,5-l-arabinofuranotriose revealed three subsites in the catalytic domain, and a sugar complex structure with α-l-arabinofuranosyl azide revealed three arabinose-binding sites in the carbohydrate binding module. A mutagenesis study revealed that substrate specificity was regulated by residues Asn-159, Tyr-192, and Leu-289 located at the aglycon side of the substrate-binding pocket. The exo-acting manner of the enzyme was attributed to the strict pocket structure of subsite −1, formed by the flexible loop region Tyr-281–Arg-294 and the side chain of Tyr-40, which occupied the positions corresponding to the catalytic glycon cleft of GH43 endo-acting enzymes.  相似文献   

12.
In this work we are proposing Homology modeled structures of Mycobacterium leprae 18kDa heat shock protein and its mutant. The more closely related structure of the small heat shock protein (sHSP) belonging to the eukaryotic species from wheat sHSP16.9 and 16.3kDa ACR1 protein from Mycobacterium tuberculosis were used as template structures. Each model contains an N-terminal domain, alpha-crystalline domain and a C-terminal tail. The models showed that a single point mutation from serine to proline at 52nd position causes structural changes. The structural changes are observed in N-terminal region and alpha-crystalline domains. Serine in 52nd position is observed in β4 strand and Proline in 52nd position is observed in loop. The number of residues contributing α helix at N-terminal region varies in both models. In 18S more number of residues is present in α helix when compared to 18P. The loop regions between β3 and β4 strands of both models vary in number of residues present in it. Number of residues contributing β4 strand in both models vary. β6 strand is absent in both models. Major functional peptide region of alpha crystalline domains of both models varies. These differences observed in secondary structures support their distinct functional roles. It also emphasizes that a point mutation can cause structural variation.  相似文献   

13.
Exopeptidases, including dipeptidyl- and tripeptidylpeptidase, are crucial for the growth of Porphyromonas gingivalis, a periodontopathic asaccharolytic bacterium that incorporates amino acids mainly as di- and tripeptides. In this study, we identified a novel exopeptidase, designated acylpeptidyl oligopeptidase (AOP), composed of 759 amino acid residues with active Ser615 and encoded by PGN_1349 in P. gingivalis ATCC 33277. AOP is currently listed as an unassigned S9 family peptidase or prolyl oligopeptidase. Recombinant AOP did not hydrolyze a Pro-Xaa bond. In addition, although sequence similarities to human and archaea-type acylaminoacyl peptidase sequences were observed, its enzymatic properties were apparently distinct from those, because AOP scarcely released an N-acyl-amino acid as compared with di- and tripeptides, especially with N-terminal modification. The kcat/Km value against benzyloxycarbonyl-Val-Lys-Met-4-methycoumaryl-7-amide, the most potent substrate, was 123.3 ± 17.3 μm−1 s−1, optimal pH was 7–8.5, and the activity was decreased with increased NaCl concentrations. AOP existed predominantly in the periplasmic fraction as a monomer, whereas equilibrium between monomers and oligomers was observed with a recombinant molecule, suggesting a tendency of oligomerization mediated by the N-terminal region (Met16–Glu101). Three-dimensional modeling revealed the three domain structures (residues Met16–Ala126, which has no similar homologue with known structure; residues Leu127–Met495 (β-propeller domain); and residues Ala496–Phe736 (α/β-hydrolase domain)) and further indicated the hydrophobic S1 site of AOP in accord with its hydrophobic P1 preference. AOP orthologues are widely distributed in bacteria, archaea, and eukaryotes, suggesting its importance for processing of nutritional and/or bioactive oligopeptides.  相似文献   

14.
XRCC1 (X-ray cross-complementing group 1) is a DNA repair protein that forms complexes with DNA polymerase β (β-Pol), DNA ligase III and poly-ADP-ribose polymerase in the repair of DNA single strand breaks. The domains in XRCC1 have been determined, and characterization of the domain–domain interaction in the XRCC1-β-Pol complex has provided information on the specificity and mechanism of binding. The domain structure of XRCC1, determined using limited proteolysis, was found to include an N-terminal domain (NTD), a central BRCT-I (breast cancer susceptibility protein-1) domain and a C-terminal BRCT-II domain. The BRCT-Ilinker–BRCT-II C-terminal fragment and the linker–BRCT-II C-terminal fragment were relatively stable to proteolysis suggestive of a non-random conformation of the linker. A predicted inner domain was found not to be stable to proteolysis. Using cross-linking experiments, XRCC1 was found to bind intact β-Pol and the β-Pol 31 kDa domain. The XRCC1-NTD1–183 (residues 1183) was found to bind β-Pol, the β-Pol 31 kDa domain and the β-Pol C-terminal palm-thumb (residues 140–335), and the interaction was further localized to XRCC1-NTD1–157 (residues 1–157). The XRCC1-NTD1–183-β-Pol 31 kDa domain complex was stable at high salt (1 M NaCl) indicative of a hydrophobic contribution. Using a yeast two-hybrid screen, polypeptides expressed from two XRCC1 constructs, which included residues 36–355 and residues 1–159, were found to interact with β-Pol, the β-Pol 31 kDa domain, and the β-Pol C-terminal thumb-only domain polypeptides expressed from the respective β-Pol constructs. Neither the XRCC1-NTD1–159, nor the XRCC136–355 polypeptide was found to interact with a β-Pol thumbless polypeptide. A third XRCC1 polypeptide (residues 75–212) showed no interaction with β-Pol. In quantitative gel filtration and analytical ultracentrifugation experiments, the XRCC1-NTD1–183 was found to bind β-Pol and its 31 kDa domain in a 1:1 complex with high affinity (Kd of 0.4–2.4 µM). The combined results indicate a thumb-domain specific 1:1 interaction between the XRCC1-NTD1–159 and β-Pol that is of an affinity comparable to other binding interactions involving β-Pol.  相似文献   

15.
The crystal structure of the full-length rat peroxisomal multifunctional enzyme, type 1 (rpMFE1), has been determined at 2.8 Å resolution. This enzyme has three catalytic activities and two active sites. The N-terminal part has the crotonase fold, which builds the active site for the Δ32-enoyl-CoA isomerase and the Δ2-enoyl-CoA hydratase-1 catalytic activities, and the C-terminal part has the (3S)-hydroxyacyl-CoA dehydrogenase fold and makes the (3S)-hydroxyacyl-CoA dehydrogenase active site. rpMFE1 is a multidomain protein having five domains (A–E). The crystal structure of full-length rpMFE1 shows a flexible arrangement of the A-domain with respect to the B–E-domains. Because of a hinge region near the end of the A-domain, two different positions of the A-domain were observed for the two protein molecules (A and B) of the asymmetric unit. In the most closed conformation, the mode of binding of CoA is stabilized by domains A and B (helix-10), as seen in other crotonase fold members. Domain B, although functionally belonging to the N-terminal part, is found tightly associated with the C-terminal part, i.e. fixed to the E-domain. The two active sites of rpMFE1 are ∼40 Å apart, separated by a tunnel, characterized by an excess of positively charged side chains. Comparison of the structures of rpMFE1 with the monofunctional crotonase and (3S)-hydroxyacyl-CoA dehydrogenase superfamily enzymes, as well as with the bacterial α2β2-fatty acid oxidation multienzyme complex, reveals that this tunnel could be important for substrate channeling, as observed earlier on the basis of the kinetics of rpMFE1 purified from rat liver.  相似文献   

16.
Bacillus circulans T-3040 cycloisomaltooligosaccharide glucanotransferase belongs to the glycoside hydrolase family 66 and catalyzes an intramolecular transglucosylation reaction that produces cycloisomaltooligosaccharides from dextran. The crystal structure of the core fragment from Ser-39 to Met-738 of B. circulans T-3040 cycloisomaltooligosaccharide glucanotransferase, devoid of its N-terminal signal peptide and C-terminal nonconserved regions, was determined. The structural model contained one catalytic (β/α)8-barrel domain and three β-domains. Domain N with an immunoglobulin-like β-sandwich fold was attached to the N terminus; domain C with a Greek key β-sandwich fold was located at the C terminus, and a carbohydrate-binding module family 35 (CBM35) β-jellyroll domain B was inserted between the 7th β-strand and the 7th α-helix of the catalytic domain A. The structures of the inactive catalytic nucleophile mutant enzyme complexed with isomaltohexaose, isomaltoheptaose, isomaltooctaose, and cycloisomaltooctaose revealed that the ligands bound in the catalytic cleft and the sugar-binding site of CBM35. Of these, isomaltooctaose bound in the catalytic site extended to the second sugar-binding site of CBM35, which acted as subsite −8, representing the enzyme·substrate complex when the enzyme produces cycloisomaltooctaose. The isomaltoheptaose and cycloisomaltooctaose bound in the catalytic cleft with a circular structure around Met-310, representing the enzyme·product complex. These structures collectively indicated that CBM35 functions in determining the size of the product, causing the predominant production of cycloisomaltooctaose by the enzyme. The canonical sugar-binding site of CBM35 bound the mid-part of isomaltooligosaccharides, indicating that the original function involved substrate binding required for efficient catalysis.  相似文献   

17.
18.
19.
20.

Background

The serine protease autotransporter EspP is a proposed virulence factor of Shiga toxin-producing Escherichia coli (STEC). We recently distinguished four EspP subtypes (EspPα, EspPβ, EspPγ, and EspPδ), which display large differences in transport and proteolytic activities and differ widely concerning their distribution within the STEC population. The mechanisms underlying these functional variations in EspP subtypes are, however, unknown.

Methodology/Principal Findings

The structural basis of proteolytic and autotransport activity was investigated using transposon-based linker scanning mutagenesis, site-directed mutagenesis and structure-function analysis derived from homology modelling of the EspP passenger domain. Transposon mutagenesis of the passenger domain inactivated autotransport when pentapeptide linker insertions occurred in regions essential for overall correct folding or in a loop protruding from the β-helical core. Loss of proteolytic function was limited to mutations in Domain 1 in the N-terminal third of the EspP passenger. Site-directed mutagenesis demonstrated that His127, Asp156 and Ser263 in Domain 1 form the catalytic triad of EspP.

Conclusions/Significance

Our data indicate that in EspP i) the correct formation of the tertiary structure of the passenger domain is essential for efficient autotransport, and ii) an elastase-like serine protease domain in the N-terminal Domain 1 is responsible for the proteolytic phenotype. Lack of stabilizing interactions of Domain 1 with the core structure of the passenger domain ablates proteolytic activity in subtypes EspPβ and EspPδ.  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