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.
Post-translational modifications of ribosomal proteins are important for the accuracy of the decoding machinery. A recent in vivo study has shown that the rimO gene is involved in generation of the 3-methylthio derivative of residue Asp-89 in ribosomal protein S12 (Anton, B. P., Saleh, L., Benner, J. S., Raleigh, E. A., Kasif, S., and Roberts, R. J. (2008) Proc. Natl. Acad. Sci. U. S. A. 105, 1826–1831). This reaction is formally identical to that catalyzed by MiaB on the C2 of adenosine 37 near the anticodon of several tRNAs. We present spectroscopic evidence that Thermotoga maritima RimO, like MiaB, contains two [4Fe-4S] centers, one presumably bound to three invariant cysteines in the central radical S-adenosylmethionine (AdoMet) domain and the other to three invariant cysteines in the N-terminal UPF0004 domain. We demonstrate that holo-RimO can specifically methylthiolate the aspartate residue of a 20-mer peptide derived from S12, yielding a mixture of mono- and bismethylthio derivatives. Finally, we present the 2.0 Å crystal structure of the central radical AdoMet and the C-terminal TRAM (tRNA methyltransferase 2 and MiaB) domains in apo-RimO. Although the core of the open triose-phosphate isomerase (TIM) barrel of the radical AdoMet domain was conserved, RimO showed differences in domain organization compared with other radical AdoMet enzymes. The unusually acidic TRAM domain, likely to bind the basic S12 protein, is located at the distal edge of the radical AdoMet domain. The basic S12 protein substrate is likely to bind RimO through interactions with both the TRAM domain and the concave surface of the incomplete TIM barrel. These biophysical results provide a foundation for understanding the mechanism of methylthioation by radical AdoMet enzymes in the MiaB/RimO family.  相似文献   

2.
Nonribosomal peptide synthetases (NRPSs) are large, multidomain proteins that are involved in the biosynthesis of an array of secondary metabolites. We report the structure of the third adenylation domain from the siderophore-synthesizing NRPS, SidN, from the endophytic fungus Neotyphodium lolii. This is the first structure of a eukaryotic NRPS domain, and it reveals a large binding pocket required to accommodate the unusual amino acid substrate, Nδ-cis-anhydromevalonyl-Nδ-hydroxy-l-ornithine (cis-AMHO). The specific activation of cis-AMHO was confirmed biochemically, and an AMHO moiety was unambiguously identified as a component of the fungal siderophore using mass spectroscopy. The protein structure shows that the substrate binding pocket is defined by 17 amino acid residues, in contrast to both prokaryotic adenylation domains and to previous predictions based on modeling. Existing substrate prediction methods for NRPS adenylation domains fail for domains from eukaryotes due to the divergence of their signature sequences from those of prokaryotes. Thus, this new structure will provide a basis for improving prediction methods for eukaryotic NRPS enzymes that play important and diverse roles in the biology of fungi.  相似文献   

3.
A synthetic peptide (Nterm-E1p) is used to characterize the structure and function of the N-terminal region (amino acid residues 4-45) of the pyruvate dehydrogenase component (E1p) from the pyruvate dehydrogenase multienzyme complex (PDHC) from Azotobacter vinelandii. Activity and binding studies established that Nterm-E1p specifically competes with E1p for binding to the dihydrolipoyl transacetylase component (E2p) of PDHC. Moreover, the experiments show that the N-terminal region of E1p forms an independent folding domain that functions as a binding domain. CD measurements, two-dimensional (2D) (1)H NMR analysis, and secondary structure prediction all indicate that Nterm-E1p has a high alpha-helical content. Here a structural model of the N-terminal domain is proposed. The peptide is present in two conformations, the population of which depends on the sample conditions. The conformations are designated "unfolded" at pH > or =6 and "folded" at pH <5. The 2D (1)H TOCSY spectrum of a mixture of folded and unfolded Nterm-E1p shows exchange cross-peaks that "link" the folded and unfolded state of Nterm-E1p. The rate of exchange between the two species is in the range of 0.5-5 s(-1). Sharp resonances in the NMR spectra of wild-type E1p demonstrate that this 200 kDa enzyme contains highly flexible regions. The observed dynamic character of E1p and of Nterm-E1p is likely required for the binding of the E1p dimer to the two different binding sites on E2p. Moreover, the flexibility might be essential in sustaining the allosteric properties of the enzyme bound in the complex.  相似文献   

4.
Cytochrome c oxidase is a member of the heme-copper family of oxygen reductases in which electron transfer is linked to the pumping of protons across the membrane. Neither the redox center(s) associated with proton pumping nor the pumping mechanism presumably common to all heme-copper oxidases has been established. A possible conformational coupling between the catalytic center (Fea33+–CuB2+) and a protein site has been identified earlier from ligand binding studies, whereas a structural change initiated by azide binding to the protein has been proposed to facilitate the access of cyanide to the catalytic center of the oxidized bovine enzyme. Here we show that cytochrome oxidase pretreated with a low concentration of azide exhibits a significant increase in the apparent rate of cyanide binding relative to that of free enzyme. However, this increase in rate does not reflect a conformational change enhancing the rapid formation of a Fea33+–CN–CuB2+ complex. Instead the cyanide-induced transition of a preformed Fea33+–N3–CuB2+ to the ternary complex of Fea33+–N3 CuB2+–CN is the most likely reason for the observed acceleration. Significantly, the slow rate of azide release from the ternary complex indicates that cyanide ligated to CuB blocks a channel between the catalytic site and the solvent. The results suggest that there is a pathway that originates at CuB and that, during catalysis, ligands present at this copper center control access to the iron of heme a3 from the bulk medium.  相似文献   

5.
Previously we showed that the inactive form of p90 ribosomal S6 kinase 1 (RSK1) interacts with the regulatory subunit, PKARIα, of protein kinase A (PKA), whereas the active RSK1 interacts with the catalytic subunit (PKAc) of PKA. Herein, we demonstrate that the N-terminal kinase domain (NTK) of RSK1 is necessary for interactions with PKARIα. Substitution of the activation loop phosphorylation site (Ser-221) in the NTK with the negatively charged Asp residue abrogated the association between RSK1 and PKARIα. This explains the lack of an interaction between active RSK1 and PKARIα. Full-length RSK1 bound to PKARIα with an affinity of 0.8 nm. The NTK domain of RSK1 competed with PKAc for binding to the pseudosubstrate region (amino acids 93–99) of PKARIα. Overexpressed RSK1 dissociated PKAc from PKARIα, increasing PKAc activity, whereas silencing of RSK1 increased PKAc/PKARIα interactions and decreased PKAc activity. Unlike PKAc, which requires Arg-95 and -96 in the pseudosubstrate region of PKARIα for their interactions, RSK1/PKARIα association requires all four Arg residues (Arg-93–96) in the pseudosubstrate site of PKARIα. A peptide (Wt-PS) corresponding to residues 91–99 of PKARIα competed for binding of RSK1 with PKARIα both in vitro and in intact cells. Furthermore, peptide Wt-PS (but not control peptide Mut-PS), by dissociating RSK1 from PKARIα, activated RSK1 in the absence of any growth factors and protected cells from apoptosis. Thus, by competing for binding to the pseudosubstrate region of PKARIα, RSK1 regulates PKAc activity in a cAMP-independent manner, and PKARIα by associating with RSK1 regulates its activation and its biological functions.  相似文献   

6.
The human pyruvate dehydrogenase complex (PDC) is a 9.5-megadalton catalytic machine that employs three catalytic components, i.e. pyruvate dehydrogenase (E1p), dihydrolipoyl transacetylase (E2p), and dihydrolipoamide dehydrogenase (E3), to carry out the oxidative decarboxylation of pyruvate. The human PDC is organized around a 60-meric dodecahedral core comprising the C-terminal domains of E2p and a noncatalytic component, E3-binding protein (E3BP), which specifically tethers E3 dimers to the PDC. A central issue concerning the PDC structure is the subunit stoichiometry of the E2p/E3BP core; recent studies have suggested that the core is composed of 48 copies of E2p and 12 copies of E3BP. Here, using an in vitro reconstituted PDC, we provide densitometry, isothermal titration calorimetry, and analytical ultracentrifugation evidence that there are 40 copies of E2p and 20 copies of E3BP in the E2p/E3BP core. Reconstitution with saturating concentrations of E1p and E3 demonstrated 40 copies of E1p heterotetramers and 20 copies of E3 dimers associated with the E2p/E3BP core. To corroborate the 40/20 model of this core, the stoichiometries of E3 and E1p binding to their respective binding domains were reexamined. In these binding studies, the stoichiometries were found to be 1:1, supporting the 40/20 model of the core. The overall maximal stoichiometry of this in vitro assembled PDC for E2p:E3BP:E1p:E3 is 40:20:40:20. These findings contrast a previous report that implicated that two E3-binding domains of E3BP bind simultaneously to a single E3 dimer (Smolle, M., Prior, A. E., Brown, A. E., Cooper, A., Byron, O., and Lindsay, J. G. (2006) J. Biol. Chem. 281, 19772–19780).The human pyruvate dehydrogenase complex (PDC)3 resides in mitochondria and catalyzes the oxidative decarboxylation of pyruvate to yield acetyl-CoA and reducing equivalents (NADH), serving as a link between glycolysis and the Krebs cycle (13). The PDC is a large (∼9.5 MDa) catalytic machine comprising multiple protein components. The three catalytic components are pyruvate dehydrogenase (E1p), dihydrolipoyl transacetylase (E2p), and dihydrolipoamide dehydrogenase (E3), with E3 being a common component between different α-keto acid dehydrogenase complexes. The two regulatory enzymes in the PDC are the isoforms of pyruvate dehydrogenase kinase and pyruvate dehydrogenase phosphatase.The PDC is organized around a structural core, which includes the C-terminal domains of E2p and a noncatalytic component that specifically binds E3, i.e. the E3-binding protein (E3BP). To this E2p/E3BP core, multiple copies of the other PDC components are tethered through noncovalent interactions. Each E2p subunit contains two consecutive N-terminal lipoic acid-bearing domains (LBDs), termed L1 and L2, followed by the E1p-binding domain (E1pBD) and the C-terminal inner-core/catalytic domain, with these independent domains connected by unstructured linkers. Similarly, each E3BP subunit consists of a single N-terminal LBD (referred to as L3), the E3-binding domain (E3BD), and the noncatalytic inner core domain. Together, the inner core domains of E2p and E3BP assemble to form the dodecahedral 60-meric E2p/E3BP core. The role of the E1pBD and E3BD domains is to tether E1p and E3, respectively, to the periphery of the E2p/E3BP core. It is presumed that the LBDs (L1, L2, and L3) shuttle between the active sites of the three catalytic components of the PDC during the oxidative decarboxylation cycle (4). The eukaryotic PDC is unique among α-keto acid dehydrogenase complexes in its requirement for E3BP; prokaryotic PDCs employ the single subunit-binding domain to secure either E1p or E3 to the complex (5).Using a “divide-and-conquer” approach, a wealth of structural information on the PDC has been accumulated recently. High-resolution crystal structures are available for the human E1p (68) and E3 components (9). A model for the human E2p has been constructed based on an 8.8-Å electron density map available from cryo-electron microscopy (10). Additionally, solution and crystal structures of the L1 and L2 domains of E2p have been determined (1113), and the high-resolution crystal structures of the E3BD (14, 15), pyruvate dehydrogenase kinase isoforms 1–4 (12, 1618), and pyruvate dehydrogenase phosphatase isoform 1 (19) are known. Therefore, atomic models are available for almost all components and domains of the mammalian PDC.With the successes of the above structural approach, attention has turned to the overall structure of the PDC. There are two outstanding questions as follows. What are the subunit and overall catalytic component stoichiometries? What are the positions and orientations of the components in this large catalytic machine? Yu et al. (10) recently determined the cryo-EM structure of a PDC core comprising only human E2p subunits. Like yeast E2p, human E2p adopts a dodecahedral structure composed of 60 E2p proteins; each face of the dodecahedron has a large gap. Although this structure is highly informative, the composition of this core deviates substantially from that of the native PDC, because no E3BP subunits are present in the core structure. Based on the similar structure of the dodecahedral yeast PDC, a hypothesis was formed that, in human PDC, 12 copies of E3BP bind in the 12 gaps, which is termed the “60/12” model (20). Biophysical studies on complexes of E2p and E3BP later negated the 60/12 model; Hiromasa et al. (21) therefore posited an alternative, the “48/12” model, in which the dodecahedral core includes 48 E2p subunits and 12 E3BP proteins. A further source of conjecture is how many E1p and E3 components bind to the periphery of the PDC. If one binding domain binds to one peripheral catalytic component, a maximally occupied 60/12 PDC would harbor 60 E1p heterotetramers and 12 E3 dimers (or 48 E1ps and 12 E3s in the 48/12 model). The notion of such 1:1 binding is supported by the preponderance of available biophysical evidence. Specifically, two crystal structures, site-directed mutagenesis, and calorimetric measurements describe a 1:1 interaction between E3BD and E3 (14, 15). Also, although no structures are available for the human E1p-E1pBD complex, a crystal structure of the homologs of these proteins from Bacillus stearothermophilus also demonstrates a 1:1 interaction between the E1pBD of E2p and the E1p heterotetramer (22). In addition, ITC experiments performed on the bacterial E1p and the cognate subunit-binding domain indicate a 1:1 association (23). At variance with the above observations, a different subunit stoichiometry has been proposed by Smolle et al. (24, 25). Their evidence suggests that two binding domains bind for every peripheral component; such an arrangement potentially yields a PDC with half as many peripheral components bound.This study was undertaken to ascertain the subunit and component stoichiometries of the human PDC, particularly with regard to interactions between the E3BD and the E3 dimer. We show that quantification of bands on an SDS-polyacrylamide gel of a PDC reconstituted at saturating E1p and E3 concentrations supports neither the 60/12 nor the 48/12 model. Instead, a “40/20” model is proposed, and subsequent ITC and analytical ultracentrifugation (AUC) data corroborate this new model. In addition, results from electrophoretic mobility shift assays, ITC, and AUC presented here uniformly show a 1:1 interaction between E3BD and the E3 dimer as well as between E1pBD and the E1p heterotetramer. The implications of this 1:1 binding stoichiometry for the macromolecular assembly of the PDC are discussed.  相似文献   

7.
The apolipoprotein E family contains three major isoforms (ApoE4, E3, and E2) that are directly involved with lipoprotein metabolism and cholesterol transport. ApoE3 and apoE4 differ in only a single amino acid with an arginine in apoE4 changed to a cysteine at position 112 in apoE3. Yet only apoE4 is recognized as a risk factor for Alzheimer''s disease. Here we used 19F NMR to examine structural differences between apoE4 and apoE3 and the effect of the C-terminal domain on the N-terminal domain. After incorporation of 5-19F-tryptophan the 1D 19F NMR spectra were compared for the N-terminal domain and for the full length proteins. The NMR spectra of the N-terminal region (residues 1–191) are reasonably well resolved while those of the full length wild-type proteins are broad and ill-defined suggesting considerable conformational heterogeneity. At least four of the seven tryptophan residues in the wild type protein appear to be solvent exposed. NMR spectra of the wild-type proteins were compared to apoE containing four mutations in the C-terminal region that gives rise to a monomeric form either of apoE3 under native conditions (Zhang et al., Biochemistry 2007; 46: 10722–10732) or apoE4 in the presence of 1 M urea. For either wild-type or mutant proteins the differences in tryptophan resonances in the N-terminal region of the protein suggest structural differences between apoE3 and apoE4. We conclude that these differences occur both as a consequence of the Arg158Cys mutation and as a consequence of the interaction with the C-terminal domain.  相似文献   

8.
A key structural element in the ion translocating F-, A-, and V-ATPases is the peripheral stalk, an assembly of two polypeptides that provides a structural link between the ATPase and ion channel domains. Previously, we have characterized the peripheral stalk forming subunits E and H of the A-ATPase from Thermoplasma acidophilum and demonstrated that the two polypeptides interact to form a stable heterodimer with 1:1 stoichiometry (Kish-Trier, E., Briere, L. K., Dunn, S. D., and Wilkens, S. (2008) J. Mol. Biol. 375, 673–685). To define the domain architecture of the A-ATPase peripheral stalk, we have now generated truncated versions of the E and H subunits and analyzed their ability to bind each other. The data show that the N termini of the subunits form an α-helical coiled-coil, ∼80 residues in length, whereas the C-terminal residues interact to form a globular domain containingα- and β-structure. We find that the isolated C-terminal domain of the E subunit exists as a dimer in solution, consistent with a recent crystal structure of the related Pyrococcus horikoshii A-ATPase E subunit (Lokanath, N. K., Matsuura, Y., Kuroishi, C., Takahashi, N., and Kunishima, N. (2007) J. Mol. Biol. 366, 933–944). However, upon the addition of a peptide comprising the C-terminal 21 residues of the H subunit (or full-length H subunit), dimeric E subunit C-terminal domain dissociates to form a 1:1 heterodimer. NMR spectroscopy was used to show that H subunit C-terminal peptide binds to E subunit C-terminal domain via the terminal α-helices, with little involvement of the β-sheet region. Based on these data, we propose a structural model of the A-ATPase peripheral stalk.The archaeal ATP synthase (A1A0-ATPase),2 along with the related F1F0- and V1V0-ATPases (proton pumping vacuolar ATPases), is a rotary molecular motor (14). The rotary ATPases are bilobular in overall architecture, with one lobe comprising the water-soluble A1, F1, or V1 and the other comprising the membrane-bound A0, F0, or V0 domain, respectively. The subunit composition of the A-ATPase is A3B3DE2FH2 for the A1 and CIKx for the A0. In the A1 domain, the three A and B subunits come together in an alternating fashion to form a hexamer with a hydrophobic inner cavity into which part of the D subunit is inserted. Subunits D and F comprise the central stalk connection to A0, whereas two heterodimeric EH complexes are thought to form the peripheral stalk attachment to A0 seen in electron microscopy reconstructions (5, 6). In the A0 domain (subunits CIKx), the K subunits (proteolipids) form a ring that is linked to the central stalk by the C subunit, whereas the cytoplasmic N-terminal domain of the I subunit probably mediates the binding of the EH peripheral stalks to A0, as suggested for the bacterial A/V-type enzyme (7). Although closer in structure to the proton-pumping V-ATPase, the A-ATPase functions in vivo as an ATP synthase, coupling ion motive force to ATP synthesis, most likely via a similar rotary mechanism as demonstrated for the bacterial A/V- and the vacuolar type enzymes (8, 9). During catalysis, substrate binding occurs sequentially on the three catalytic sites, which are formed predominantly by the A subunits. This is accompanied by conformation changes in the A3B3 hexamer that are linked to the rotation of the embedded D subunit together with the rotor subunits F, C, and the proteolipid ring. Each copy of K contains a lipid-exposed carboxyl residue (Asp or Glu), which is transiently interfaced with the membrane-bound domain of I during rotation, thereby catalyzing ion translocation. The EH peripheral stalks function to stabilize the A3B3 hexamer against the torque generated during rotation of the central stalk. Much work has been accomplished to elucidate the architectural features of the rotational and catalytic domains, especially in the related F- and V-type enzymes. However, the peripheral stalk complexes in the A- and V-type enzymes remain an area open to question. Although the stoichiometry of the peripheral stalks in the A/V-type and the vacuolar type ATPases have recently been resolved to two and three, respectively (6, 10), the overall structure of the peripheral stalk, including the nature of attachment to the A3B3 hexamer and I subunit (called subunit a in the F- and V-ATPase), is not well understood. Some structural information exists in the form of the A-ATPase E subunit C-terminal domain (11), although isolation from its binding partner H may have influenced its conformation.Previously, our lab has characterized the Thermoplasma acidophilum A-ATPase E and H subunits individually and in complex (12). We found that despite their tendency to oligomerize when isolated separately, upon mixing, E and H form a tight heterodimer that was monodisperse and elongated in solution, which is consistent with its role as the peripheral stalk element in the A-ATPase. Here, we have expanded our study of the A-ATPase EH complex through the production of various N- and C-terminal truncation mutants of both binding partners. The data show that the EH complex is comprised of two distinct domains, one that contains both N termini interacting via a coiled-coil and a second that contains both C termini folded in a globular structure containing mixed secondary structure. Consistent with recent crystallographic data for the related A-ATPase from Pyrococcus horikoshii (11), we found that the isolated C-terminal domain of the E subunit exists as a stable homodimer in solution. However, the addition of subunit H or a peptide consisting of the 21 C-terminal residues of the subunit to the dimeric C-terminal domain of subunit E resulted in dissociation of the homodimer with concomitant formation of a 1:1 heterodimer containing the C termini of both polypeptides. This study delineates and characterizes the two domains of the EH complex and will aid in the further exploration of the nature of peripheral stalk attachment and function in the intact A1A0-ATPase.  相似文献   

9.
Small guanosine triphosphatases (GTPases) become activated when GDP is replaced by GTP at the highly conserved nucleotide binding site. This process is intrinsically very slow in most GTPases but is significantly accelerated by guanine nucleotide exchange factors (GEFs). Nucleotide exchange in small GTPases has been widely studied using spectroscopy with fluorescently tagged nucleotides. However, this method suffers from effects of the bulky fluorescent moiety covalently attached to the nucleotide. Here, we have used a newly developed real-time NMR-based assay to monitor small GTPase RhoA nucleotide exchange by probing the RhoA conformation. We compared RhoA nucleotide exchange from GDP to GTP and GTP analogues in the absence and presence of the catalytic DH-PH domain of PDZ-RhoGEF (DH-PHPRG). Using the non-hydrolyzable analogue guanosine-5′-O-(3-thiotriphosphate), which we found to be a reliable mimic of GTP, we obtained an intrinsic nucleotide exchange rate of 5.5 × 10−4 min−1. This reaction is markedly accelerated to 1179 × 10−4 min−1 in the presence of DH-PHPRG at a ratio of 1:8,000 relative to RhoA. Mutagenesis studies confirmed the importance of Arg-868 near a conserved region (CR3) of the Dbl homology (DH) domain and revealed that Glu-741 in CR1 is critical for full activity of DH-PHPRG, together suggesting that the catalytic mechanism of PDZ-RhoGEF is similar to Tiam1. Mutation of the single RhoA (E97A) residue that contacts the pleckstrin homology (PH) domain rendered the mutant 10-fold less sensitive to the activity of DH-PHPRG. Interestingly, this mutation does not affect RhoA activation by leukemia-associated RhoGEF (LARG), indicating that the PH domains of these two homologous GEFs may play different roles.  相似文献   

10.
The role of the distal histidine in regulating ligand binding to adult human hemoglobin (HbA) was re-examined systematically by preparing His(E7) to Gly, Ala, Leu, Gln, Phe, and Trp mutants of both Hb subunits. Rate constants for O2, CO, and NO binding were measured using rapid mixing and laser photolysis experiments designed to minimize autoxidation of the unstable apolar E7 mutants. Replacing His(E7) with Gly, Ala, Leu, or Phe causes 20–500-fold increases in the rates of O2 dissociation from either Hb subunit, demonstrating unambiguously that the native His(E7) imidazole side chain forms a strong hydrogen bond with bound O2 in both the α and β chains (ΔGHis(E7)H-bond ≈ −8 kJ/mol). As the size of the E7 amino acid is increased from Gly to Phe, decreases in kO2′, kNO′, and calculated bimolecular rates of CO entry (kentry′) are observed. Replacing His(E7) with Trp causes further decreases in kO2′, kNO′, and kentry′ to 1–2 μm−1 s−1 in β subunits, whereas ligand rebinding to αTrp(E7) subunits after photolysis is markedly biphasic, with fast kO2′, kCO′, and kNO′ values ≈150 μm−1 s−1 and slow rate constants ≈0.1 to 1 μm−1 s−1. Rapid bimolecular rebinding to an open α subunit conformation occurs immediately after photolysis of the αTrp(E7) mutant at high ligand concentrations. However, at equilibrium the closed αTrp(E7) side chain inhibits the rate of ligand binding >200-fold. These data suggest strongly that the E7 side chain functions as a gate for ligand entry in both HbA subunits.  相似文献   

11.
2,4-Diacetylphloroglucinol hydrolase PhlG from Pseudomonas fluorescens catalyzes hydrolytic carbon-carbon (C–C) bond cleavage of the antibiotic 2,4-diacetylphloroglucinol to form monoacetylphloroglucinol, a rare class of reactions in chemistry and biochemistry. To investigate the catalytic mechanism of this enzyme, we determined the three-dimensional structure of PhlG at 2.0 Å resolution using x-ray crystallography and MAD methods. The overall structure includes a small N-terminal domain mainly involved in dimerization and a C-terminal domain of Bet v1-like fold, which distinguishes PhlG from the classical α/β-fold hydrolases. A dumbbell-shaped substrate access tunnel was identified to connect a narrow interior amphiphilic pocket to the exterior solvent. The tunnel is likely to undergo a significant conformational change upon substrate binding to the active site. Structural analysis coupled with computational docking studies, site-directed mutagenesis, and enzyme activity analysis revealed that cleavage of the 2,4-diacetylphloroglucinol C–C bond proceeds via nucleophilic attack by a water molecule, which is coordinated by a zinc ion. In addition, residues Tyr121, Tyr229, and Asn132, which are predicted to be hydrogen-bonded to the hydroxyl groups and unhydrolyzed acetyl group, can finely tune and position the bound substrate in a reactive orientation. Taken together, these results revealed the active sites and zinc-dependent hydrolytic mechanism of PhlG and explained its substrate specificity as well.  相似文献   

12.
Protein-protein interactions represent an important post-translational mechanism for endothelial nitric-oxide synthase (eNOS) regulation. We have previously reported that β-actin is associated with eNOS oxygenase domain and that association of eNOS with β-actin increases eNOS activity and nitric oxide (NO) production. In the present study, we found that β-actin-induced increase in NO production was accompanied by decrease in superoxide formation. A synthetic actin-binding sequence (ABS) peptide 326 with amino acid sequence corresponding to residues 326–333 of human eNOS, one of the putative ABSs, specifically bound to β-actin and prevented eNOS association with β-actin in vitro. Peptide 326 also prevented β-actin-induced decrease in superoxide formation and increase in NO and l-citrulline production. A modified peptide 326 replacing hydrophobic amino acids leucine and tryptophan with neutral alanine was unable to interfere with eNOS-β-actin binding and to prevent β-actin-induced changes in NO and superoxide formation. Site-directed mutagenesis of the actin-binding domain of eNOS replacing leucine and tryptophan with alanine yielded an eNOS mutant that exhibited reduced eNOS-β-actin association, decreased NO production, and increased superoxide formation in COS-7 cells. Disruption of eNOS-β-actin interaction in endothelial cells using ABS peptide 326 resulted in decreased NO production, increased superoxide formation, and decreased endothelial monolayer wound repair, which was prevented by PEG-SOD and NO donor NOC-18. Taken together, this novel finding indicates that β-actin binding to eNOS through residues 326–333 in the eNOS protein results in shifting the enzymatic activity from superoxide formation toward NO production. Modulation of NO and superoxide formation from eNOS by β-actin plays an important role in endothelial function.  相似文献   

13.
The β-hydroxyacid dehydrogenases form a large family of ubiquitous enzymes that catalyze oxidation of various β-hydroxy acid substrates to corresponding semialdehydes. Several known enzymes include β-hydroxyisobutyrate dehydrogenase, 6-phosphogluconate dehydrogenase, 2-(hydroxymethyl)glutarate dehydrogenase, and phenylserine dehydrogenase, but the vast majority of β-hydroxyacid dehydrogenases remain uncharacterized. Here, we demonstrate that the predicted β-hydroxyisobutyrate dehydrogenase PA0743 from Pseudomonas aeruginosa catalyzes an NAD+-dependent oxidation of l-serine and methyl-l-serine but exhibits low activity against β-hydroxyisobutyrate. Two crystal structures of PA0743 were solved at 2.2–2.3-Å resolution and revealed an N-terminal Rossmann fold domain connected by a long α-helix to the C-terminal all-α domain. The PA0743 apostructure showed the presence of additional density modeled as HEPES bound in the interdomain cleft close to the predicted catalytic Lys-171, revealing the molecular details of the PA0743 substrate-binding site. The structure of the PA0743-NAD+ complex demonstrated that the opposite side of the enzyme active site accommodates the cofactor, which is also bound near Lys-171. Site-directed mutagenesis of PA0743 emphasized the critical role of four amino acid residues in catalysis including the primary catalytic residue Lys-171. Our results provide further insight into the molecular mechanisms of substrate selectivity and activity of β-hydroxyacid dehydrogenases.  相似文献   

14.
The bacterial phosphoenolpyruvate (PEP) sugar phosphotransferase system mediates sugar uptake and controls the carbon metabolism in response to carbohydrate availability. Enzyme I (EI), the first component of the phosphotransferase system, consists of an N-terminal protein binding domain (EIN) and a C-terminal PEP binding domain (EIC). EI transfers phosphate from PEP by double displacement via a histidine residue on EIN to the general phosphoryl carrier protein HPr. Here we report the 2.4 Å crystal structure of the homodimeric EI from Staphylococcus aureus. EIN consists of the helical hairpin HPr binding subdomain and the phosphorylatable βα phospho-histidine (P-His) domain. EIC folds into an (βα)8 barrel. The dimer interface of EIC buries 1833 Å2 of accessible surface per monomer and contains two Ca2+ binding sites per dimer. The structures of the S. aureus and Escherichia coli EI domains (Teplyakov, A., Lim, K., Zhu, P. P., Kapadia, G., Chen, C. C., Schwartz, J., Howard, A., Reddy, P. T., Peterkofsky, A., and Herzberg, O. (2006) Proc. Natl. Acad. Sci. U.S.A. 103, 16218–16223) are very similar. The orientation of the domains relative to each other, however, is different. In the present structure the P-His domain is docked to the HPr binding domain in an orientation appropriate for in-line transfer of the phosphate to the active site histidine of the acceptor HPr. In the E. coli structure the phospho-His of the P-His domain projects into the PEP binding site of EIC. In the S. aureus structure the crystallographic temperature factors are lower for the HPr binding domain in contact with the P-His domain and higher for EIC. In the E. coli structure it is the reverse.  相似文献   

15.
Pyranose 2-oxidase (P2O) catalyzes the oxidation by O2 of d-glucose and several aldopyranoses to yield the 2-ketoaldoses and H2O2. Based on crystal structures, in one rotamer conformation, the threonine hydroxyl of Thr169 forms H-bonds to the flavin-N5/O4 locus, whereas, in a different rotamer, it may interact with either sugar or other parts of the P2O·sugar complex. Transient kinetics of wild-type (WT) and Thr169 → S/N/G/A replacement variants show that d-Glc binds to T169S, T169N, and WT with the same Kd (45–47 mm), and the hydride transfer rate constants (kred) are similar (15.3–9.7 s−1 at 4 °C). kred of T169G with d-glucose (0.7 s−1, 4 °C) is significantly less than that of WT but not as severely affected as in T169A (kred of 0.03 s−1 at 25 °C). Transient kinetics of WT and mutants using d-galactose show that P2O binds d-galactose with a one-step binding process, different from binding of d-glucose. In T169S, T169N, and T169G, the overall turnover with d-Gal is faster than that of WT due to an increase of kred. In the crystal structure of T169S, Ser169 Oγ assumes a position identical to that of Oγ1 in Thr169; in T169G, solvent molecules may be able to rescue H-bonding. Our data suggest that a competent reductive half-reaction requires a side chain at position 169 that is able to form an H-bond within the ES complex. During the oxidative half-reaction, all mutants failed to stabilize a C4a-hydroperoxyflavin intermediate, thus suggesting that the precise position and geometry of the Thr169 side chain are required for intermediate stabilization.  相似文献   

16.
Bacillus anthracis is a Gram-positive spore-forming bacterium that causes anthrax. With the increased threat of anthrax in biowarfare, there is an urgent need to characterize new antimicrobial targets from B. anthracis. One such target is dihydrodipicolinate synthase (DHDPS), which catalyzes the committed step in the pathway yielding meso-diaminopimelate and lysine. In this study, we employed CD spectroscopy to demonstrate that the thermostability of DHDPS from B. anthracis (Ba-DHDPS) is significantly enhanced in the presence of the substrate, pyruvate. Analytical ultracentrifugation studies show that the tetramer-dimer dissociation constant of the enzyme is 3-fold tighter in the presence of pyruvate compared with the apo form. To examine the significance of this substrate-mediated stabilization phenomenon, a dimeric mutant of Ba-DHDPS (L170E/G191E) was generated and shown to have markedly reduced activity compared with the wild-type tetramer. This demonstrates that the substrate, pyruvate, stabilizes the active form of the enzyme. We next determined the high resolution (2.15 Å) crystal structure of Ba-DHDPS in complex with pyruvate (3HIJ) and compared this to the apo structure (1XL9). Structural analyses show that there is a significant (91 Å2) increase in buried surface area at the tetramerization interface of the pyruvate-bound structure. This study describes a new mechanism for stabilization of the active oligomeric form of an antibiotic target from B. anthracis and reveals an “Achilles heel” that can be exploited in structure-based drug design.  相似文献   

17.
BRCA1 C-terminal domain (BRCT)-containing proteins are found widely throughout the animal and bacteria kingdoms where they are exclusively involved in cell cycle regulation and DNA metabolism. Whereas most BRCT domains are involved in protein-protein interactions, a small subset has bona fide DNA binding activity. Here, we present the solution structure of the BRCT region of the large subunit of replication factor C bound to DNA and a model of the structure-specific complex with 5′-phosphorylated double-stranded DNA. The replication factor C BRCT domain possesses a large basic patch on one face, which includes residues that are structurally conserved and ligate the phosphate in phosphopeptide binding BRCT domains. An extra α-helix at the N terminus, which is required for DNA binding, inserts into the major groove and makes extensive contacts to the DNA backbone. The model of the protein-DNA complex suggests 5′-phosphate recognition by the BRCT domains of bacterial NAD+-dependent ligases and a nonclamp loading role for the replication factor C complex in DNA transactions.  相似文献   

18.
The Escherichia coli pyruvate dehydrogenase multienzyme complex contains multiple copies of three enzymatic components, E1p, E2p, and E3, that sequentially carry out distinct steps in the overall reaction converting pyruvate to acetyl-CoA. Efficient functioning requires the enzymatic components to assemble into a large complex, the integrity of which is maintained by tethering of the displaced, peripheral E1p and E3 components to the E2p core through non-covalent binding. We here report the crystal structure of a subcomplex between E1p and an E2p didomain containing a hybrid lipoyl domain along with the peripheral subunit-binding domain responsible for tethering to the core. In the structure, a region at the N terminus of each subunit in the E1p homodimer previously unseen due to crystallographic disorder was observed, revealing a new folding motif involved in E1p-E2p didomain interactions, and an additional, unexpected, flexibility was discovered in the E1p-E2p didomain subcomplex, both of which probably have consequences in the overall multienzyme complex assembly. This represents the first structure of an E1p-E2p didomain subcomplex involving a homodimeric E1p, and the results may be applicable to a large range of complexes with homodimeric E1 components. Results of HD exchange mass spectrometric experiments using the intact, wild type 3-lipoyl E2p and E1p are consistent with the crystallographic data obtained from the E1p-E2p didomain subcomplex as well as with other biochemical and NMR data reported from our groups, confirming that our findings are applicable to the entire E1p-E2p assembly.  相似文献   

19.
Platelet attachment to von Willebrand factor (vWF) requires the interaction between the platelet GP1bα and exposed vWF-A1 domains. Structural insights into the mechanism of the A1-GP1bα interaction have been limited to an N-terminally truncated A1 domain that lacks residues Q1238E1260 that make up the linker between the D3 and A1 domains of vWF. We have demonstrated that removal of these residues destabilizes quaternary interactions in the A1A2A3 tridomain and contributes to platelet activation under high shear (Auton et al., J Biol Chem 2012;287:14579–14585). In this study, we demonstrate that removal of these residues from the single A1 domain enhances platelet pause times on immobilized A1 under rheological shear. A rigorous comparison between the truncated A1-1261 and full length A1-1238 domains demonstrates a kinetic stabilization of the A1 domain induced by these N-terminal residues as evident in the enthalpy of the unfolding transition. This stabilization occurs through site and sequence-specific binding of the N-terminal peptide to A1. Binding of free N-terminal peptide to A1-1261 has an affinity and this binding although free to dissociate is sufficient to suppress the platelet pause times to levels comparable to A1-1238 under shear stress. Our results support a dual-structure/function role for this linker region involving a conformational equilibria that maintains quaternary A domain associations in the inactive state of vWF at low shear and an intra-A1-domain conformation that regulates the strength of platelet GP1bα-vWF A1 domain associations in the active state of vWF at high shear.  相似文献   

20.
Type IV secretion systems (T4SS) are used by Gram-negative bacteria to translocate protein and DNA substrates across the cell envelope and into target cells. Translocation across the outer membrane is achieved via a ringed tetradecameric outer membrane complex made up of a small VirB7 lipoprotein (normally 30 to 45 residues in the mature form) and the C-terminal domains of the VirB9 and VirB10 subunits. Several species from the genera of Xanthomonas phytopathogens possess an uncharacterized type IV secretion system with some distinguishing features, one of which is an unusually large VirB7 subunit (118 residues in the mature form). Here, we report the NMR and 1.0 Å X-ray structures of the VirB7 subunit from Xanthomonas citri subsp. citri (VirB7XAC2622) and its interaction with VirB9. NMR solution studies show that residues 27–41 of the disordered flexible N-terminal region of VirB7XAC2622 interact specifically with the VirB9 C-terminal domain, resulting in a significant reduction in the conformational freedom of both regions. VirB7XAC2622 has a unique C-terminal domain whose topology is strikingly similar to that of N0 domains found in proteins from different systems involved in transport across the bacterial outer membrane. We show that VirB7XAC2622 oligomerizes through interactions involving conserved residues in the N0 domain and residues 42–49 within the flexible N-terminal region and that these homotropic interactions can persist in the presence of heterotropic interactions with VirB9. Finally, we propose that VirB7XAC2622 oligomerization is compatible with the core complex structure in a manner such that the N0 domains form an extra layer on the perimeter of the tetradecameric ring.  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