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 46 毫秒
1.
Activation of platelets by the serine protease thrombin is a critical event in haemostasis. This process involves the binding of thrombin to glycoprotein Ibα (GpIbα) and cleavage of protease-activated receptors (PARs). The N-terminal extracellular domain of GpIbα contains an acidic peptide stretch that has been identified as the main thrombin binding site, and both anion binding exosites of thrombin have been implicated in GpIbα binding, but it remains unclear how they are involved. This issue is of critical importance for the mechanism of platelet activation by thrombin. If both exosites bind to GpIbα, thrombin could potentially act as a platelet adhesion molecule or receptor dimerisation trigger. Alternatively, if only a single site is involved, GpIbα may serve as a cofactor for PAR-1 activation by thrombin. To determine the involvement of thrombin's two exosites in GpIbα binding, we employed the complementary methods of mutational analysis, binding studies, X-ray crystallography and NMR spectroscopy. Our results indicate that the peptide corresponding to the C-terminal portion of GpIbα and the entire extracellular domain bind exclusively to thrombin's exosite II. The interaction of thrombin with GpIbα thus serves to recruit thrombin activity to the platelet surface while leaving exosite I free for PAR-1 recognition.  相似文献   

2.
The substrate specificity of thrombin is regulated by binding of macromolecular substrates and effectors to exosites I and II. Exosites I and II have been reported to be extremely linked allosterically, such that binding of a ligand to one exosite results in near-total loss of affinity for ligands at the alternative exosite, whereas other studies support the independence of the interactions. An array of fluorescent thrombin derivatives and fluorescein-labeled hirudin(54-65) ([5F]Hir(54-65)(SO(3)(-))) were used as probes in quantitative equilibrium binding studies to resolve whether the affinities of the exosite I-specific ligands, Hir(54-65)(SO(3)(-)) and fibrinogen, and of the exosite II-specific ligands, prothrombin fragment 2 and a monoclonal antibody, were affected by alternate exosite occupation. Hir(54-65)(SO(3)(-)) and fibrinogen bound to exosite I with dissociation constants of 16-28 nm and 5-7 microm, respectively, which were changed < or =2-fold by fragment 2 binding. Native thrombin and four thrombin derivatives labeled with different probes bound fragment 2 and the antibody with dissociation constants of 3-12 microm and 1.8 nm, respectively, unaffected by Hir(54-65)(SO(3)(-)). The results support a ternary complex binding model in which exosites I and II can be occupied simultaneously. The thrombin catalytic site senses individual and simultaneous binding of exosite I and II ligands differently, resulting in unique active site environments for each thrombin complex. The results indicate significant, ligand-specific allosteric coupling between thrombin exosites I and II and catalytic site perturbations but insignificant inter-exosite thermodynamic linkage.  相似文献   

3.
Bothrojaracin, a 27-kDa C-type lectin from Bothrops jararaca venom, is a selective and potent thrombin inhibitor (K(d) = 0.6 nM) which interacts with the two thrombin anion-binding exosites (I and II) but not with its catalytic site. In the present study, we analyzed the allosteric effects produced in the catalytic site by bothrojaracin binding to thrombin exosites. Opposite effects were observed with alpha-thrombin, which possesses both exosites I and II, and with gamma-thrombin, which lacks exosite I. On the one hand, bothrojaracin altered both kinetic parameters K(m) and k(cat) of alpha-thrombin for small synthetic substrates, resulting in an increased efficiency of alpha-thrombin catalytic activity. This effect was similar to that produced by hirugen, a peptide based on the C-terminal hirudin sequence (residues 54-65) which interacts exclusively with exosite I. On the other hand, bothrojaracin decreased the amidolytic activity of gamma-thrombin toward chromogenic substrates, although this effect was observed with higher concentrations of bothrojaracin than those used with alpha-thrombin. In agreement with these observaions, bothrojaracin produced opposite effects on the fluorescence intensity of alpha- and gamma-thrombin derivatives labeled at the active site with fluorescein-Phe-Pro-Arg-chloromethylketone. These observations support the conclusion that bothrojaracin binding to thrombin produces two different structural changes in its active site, depending on whether it interacts exclusively with exosite II, as seen with gamma-thrombin, or with exosite I (or both I and II) as observed with alpha-thrombin. The ability of bothrojaracin to evoke distinct modifications in the thrombin catalytic site environment when interacting with exosites I and II make this molecule an interesting tool for the study of allosteric changes in the thrombin molecule.  相似文献   

4.
The integrin receptor αMβ2 mediates phagocytosis of complement-opsonized objects, adhesion to the extracellular matrix, and transendothelial migration of leukocytes. However, the mechanistic aspects of αMβ2 signaling upon ligand binding are unclear. Here, we present the first atomic structure of the human αMβ2 headpiece fragment in complex with the nanobody (Nb) hCD11bNb1 at a resolution of 3.2 Å. We show that the receptor headpiece adopts the closed conformation expected to exhibit low ligand affinity. The crystal structure indicates that in the R77H αM variant, associated with systemic lupus erythematosus, the modified allosteric relationship between ligand binding and integrin outside–inside signaling is due to subtle conformational effects transmitted over a distance of 40 Å. Furthermore, we found the Nb binds to the αI domain of the αM subunit in an Mg2+-independent manner with low nanomolar affinity. Biochemical and biophysical experiments with purified proteins demonstrated that the Nb acts as a competitive inhibitor through steric hindrance exerted on the thioester domain of complement component iC3b attempting to bind the αM subunit. Surprisingly, we show that the Nb stimulates the interaction of cell-bound αMβ2 with iC3b, suggesting that it may represent a novel high-affinity proteinaceous αMβ2-specific agonist. Taken together, our data suggest that the iC3b–αMβ2 complex may be more dynamic than predicted from the crystal structure of the core complex. We propose a model based on the conformational spectrum of the receptor to reconcile these observations regarding the functional consequences of hCD11bNb1 binding to αMβ2.  相似文献   

5.
Calcium-dependent lectin I from Pseudomonas aeruginosa (PA-IL) binds specifically to oligosaccharides presenting an α-galactose residue at their nonreducing end, such as the disaccharides αGal1–2βGalOMe, αGal1–3βGalOMe, and αGal1–4βGalOMe. This provides a unique model for studying the effect of the glycosidic linkage of the ligands on structure and thermodynamics of the complexes by means of experimental and theoretical tools. The structural features of PA-IL in complex with the three disaccharides were established by docking and molecular dynamics simulations and compared with those observed in available crystal structures, including PA-IL·αGal1–2βGalOMe complex, which was solved at 2.4 Å resolution and reported herein. The role of a structural bridge water molecule in the binding site of PA-IL was also elucidated through molecular dynamics simulations and free energy calculations. This water molecule establishes three very stable hydrogen bonds with O6 of nonreducing galactose, oxygen from Pro-51 main chain, and nitrogen from Gln-53 main chain of the lectin binding site. Binding free energies for PA-IL in complex with the three disaccharides were investigated, and the results were compared with the experimental data determined by titration microcalorimetry. When the bridge water molecule was included in the free energy calculations, the simulations predicted the correct binding affinity trends with the 1–2-linked disaccharide presenting three times stronger affinity ligand than the other two. These results highlight the role of the water molecule in the binding site of PA-IL and indicate that it should be taken into account when designing glycoderivatives active against P. aeruginosa adhesion.  相似文献   

6.
Although exosites 1 and 2 regulate thrombin activity by binding substrates and cofactors and by allosterically modulating the active site, it is unclear whether there is direct allosteric linkage between the two exosites. To begin to address this, we first titrated a thrombin variant fluorescently labeled at exosite 1 with exosite 2 ligands, HD22 (a DNA aptamer), γ′-peptide (an analog of the COOH terminus of the γ′-chain of fibrinogen) or heparin. Concentration-dependent and saturable changes in fluorescence were elicited, supporting inter-exosite linkage. To explore the functional consequences of this phenomenon, we evaluated the capacity of exosite 2 ligands to inhibit thrombin binding to γAA-fibrin, an interaction mediated solely by exosite 1. When γAA-fibrinogen was clotted with thrombin in the presence of HD22, γ′-peptide, or prothrombin fragment 2 there was a dose-dependent and saturable decrease in thrombin binding to the resultant fibrin clots. Furthermore, HD22 reduced the affinity of thrombin for γAA-fibrin 6-fold and accelerated the dissociation of thrombin from preformed γAA-fibrin clots. Similar responses were obtained when surface plasmon resonance was used to monitor the interaction of thrombin with γAA-fibrinogen or fibrin. There is bidirectional communication between the exosites, because exosite 1 ligands, HD1 (a DNA aptamer) or hirudin-(54–65) (an analog of the COOH terminus of hirudin), inhibited the exosite 2-mediated interaction of thrombin with immobilized γ′-peptide. These findings provide evidence for long range allosteric linkage between exosites 1 and 2 on thrombin, revealing further complexity to the mechanisms of thrombin regulation.As the central effector of hemostasis, thrombin is engaged in procoagulant, anticoagulant, and fibrinolytic processes. These seemingly contrasting roles are regulated, at least in part, by thrombin''s interactions with other factors in the blood and vasculature. The binding of ligands to thrombin is promoted by exosites 1 and 2, which are positively charged domains that flank the active site. These exosites facilitate the binding of substrates or cofactors and align them for optimal interaction with the active site (1).Exosite 1 is predominantly used to gain access to the active site by substrates such as fibrinogen (2), factors V (3) and VIII (4), and the protease-activated receptors (PARs)2 on platelets (5). Effectors that modulate thrombin activity, including thrombomodulin (6), hirudin (7), and heparin cofactor II (8), also utilize exosite 1. Thrombomodulin alters the specificity of thrombin by hindering access of other substrates to exosite 1 (9) and by providing new binding sites for protein C and thrombin-activable fibrinolysis inhibitor, thereby promoting anticoagulant and antifibrinolytic pathways, respectively (10, 11). Fewer processes are mediated by exosite 2, which serves largely as a tether that anchors thrombin for participation in other reactions. Thus, heparin binds exosite 2 (12) and catalyzes thrombin inhibition by antithrombin and heparin cofactor II (13, 14). Exosite 2 also is used by glycoprotein 1bα on platelets to localize thrombin for activation of PARs (1517).Although the prevailing role of the exosites is to bring substrates and cofactors into proximity with thrombin, there is evidence that the exosites also serve as allosteric regulators of thrombin activity. Crystallographic studies reveal that, when peptides derived from PAR1 or PAR3 are bound to exosite 1 on thrombin, an obstructing surface loop moves out of the active site pocket, thereby providing access to substrates (18). The binding of a thrombomodulin fragment to exosite 1 was shown to alter the environment of an active site fluorescent probe (19), which accelerates the rate of protein C and thrombin-activable fibrinolysis inhibitor activation in an allosteric fashion. In contrast, exosite 1-binding peptides from heparin cofactor II or fibrinogen decrease the rate of protein C activation (20). Additionally, the binding of ligands to exosite 1 alters the rates of chromogenic substrate hydrolysis (21, 22). Allosteric effects are not limited to exosite 1, because prothrombin fragment 2 (F2), a cleavage product of prothrombin, binds exosite 2 and decreases the rate at which thrombin converts fibrinogen to fibrin (23, 24) and is inhibited by antithrombin (25, 26). In support of the concept that these alterations are allosteric in origin, fluorescent probes bound to the active site of thrombin undergo a change in fluorescence intensity when exosite 2 is occupied (24, 27).Although there is good evidence for allosteric regulation of the active site by the exosites, it remains unclear whether there is direct allosteric connection between the exosites. Reciprocal effects between exosites 1 and 2 have been observed by some investigators (2830), but not by others (25, 31). The aim of the current study was to use different techniques and additional ligands to resolve this controversy. First, we examined the effect of exosite 2-directed ligands on the fluorescence intensity of a thrombin variant that was labeled in exosite 1. Next, we examined the effect of these ligands on thrombin binding to fibrin. To exploit the observation that thrombin binds γAA-fibrinogen exclusively via exosite 1 (2, 32), leaving exosite 2 accessible, this subpopulation was isolated (32). We used intact fibrin clots and surface plasmon resonance (SPR) to examine the influence of exosite 2-directed ligands on thrombin binding to γAA-fibrin. In addition, diffusion studies were performed to examine the effect of exosite-directed ligands on the rate of thrombin dissociation from preformed fibrin clots. Finally, we explored whether exosite 1-directed ligands modulate the binding of thrombin to an exosite 2-directed ligand.  相似文献   

7.
Endothelial dysfunction is a hallmark of inflammation and is mediated by inflammatory factors that signal through G protein–coupled receptors including protease-activated receptor-1 (PAR1). PAR1, a receptor for thrombin, signals via the small GTPase RhoA and myosin light chain intermediates to facilitate endothelial barrier permeability. PAR1 also induces endothelial barrier disruption through a p38 mitogen-activated protein kinase–dependent pathway, which does not integrate into the RhoA/MLC pathway; however, the PAR1-p38 signaling pathways that promote endothelial dysfunction remain poorly defined. To identify effectors of this pathway, we performed a global phosphoproteome analysis of thrombin signaling regulated by p38 in human cultured endothelial cells using multiplexed quantitative mass spectrometry. We identified 5491 unique phosphopeptides and 2317 phosphoproteins, four distinct dynamic phosphoproteome profiles of thrombin-p38 signaling, and an enrichment of biological functions associated with endothelial dysfunction, including modulators of endothelial barrier disruption and a subset of kinases predicted to regulate p38-dependent thrombin signaling. Using available antibodies to detect identified phosphosites of key p38-regulated proteins, we discovered that inhibition of p38 activity and siRNA-targeted depletion of the p38α isoform increased basal phosphorylation of extracellular signal–regulated protein kinase 1/2, resulting in amplified thrombin-stimulated extracellular signal–regulated protein kinase 1/2 phosphorylation that was dependent on PAR1. We also discovered a role for p38 in the phosphorylation of α-catenin, a component of adherens junctions, suggesting that this phosphorylation may function as an important regulatory process. Taken together, these studies define a rich array of thrombin- and p38-regulated candidate proteins that may serve important roles in endothelial dysfunction.  相似文献   

8.
Batroxobin is a thrombin-like serine protease from the venom of Bothrops atrox moojeni that clots fibrinogen. In contrast to thrombin, which releases fibrinopeptide A and B from the NH2-terminal domains of the Aα- and Bβ-chains of fibrinogen, respectively, batroxobin only releases fibrinopeptide A. Because the mechanism responsible for these differences is unknown, we compared the interactions of batroxobin and thrombin with the predominant γAA isoform of fibrin(ogen) and the γA/γ′ variant with an extended γ-chain. Thrombin binds to the γ′-chain and forms a higher affinity interaction with γA/γ′-fibrin(ogen) than γAA-fibrin(ogen). In contrast, batroxobin binds both fibrin(ogen) isoforms with similar high affinity (Kd values of about 0.5 μm) even though it does not interact with the γ′-chain. The batroxobin-binding sites on fibrin(ogen) only partially overlap with those of thrombin because thrombin attenuates, but does not abrogate, the interaction of γAA-fibrinogen with batroxobin. Furthermore, although both thrombin and batroxobin bind to the central E-region of fibrinogen with a Kd value of 2–5 μm, the α(17–51) and Bβ(1–42) regions bind thrombin but not batroxobin. Once bound to fibrin, the capacity of batroxobin to promote fibrin accretion is 18-fold greater than that of thrombin, a finding that may explain the microvascular thrombosis that complicates envenomation by B. atrox moojeni. Therefore, batroxobin binds fibrin(ogen) in a manner distinct from thrombin, which may contribute to its higher affinity interaction, selective fibrinopeptide A release, and prothrombotic properties.  相似文献   

9.
α1-Protease inhibitor Portland (α1PDX) is an engineered serpin family inhibitor of the proprotein convertase (PC), furin, that exhibits high specificity but limited selectivity for inhibiting furin over other PC family proteases. Here, we characterize serpin B8, a natural inhibitor of furin, together with α1PDX-serpin B8 and furin-PC chimeras to identify determinants of serpin specificity and selectivity for furin inhibition. Replacing reactive center loop (RCL) sequences of α1PDX with those of serpin B8 demonstrated that both the P4–P1 RXXR recognition sequence as well as the P1′–P5′ sequence are critical determinants of serpin specificity for furin. Alignments of PC catalytic domains revealed four variable active-site loops whose role in furin reactivity with serpin B8 was tested by engineering furin-PC loop chimeras. The furin(298–300) loop but not the other loops differentially affected furin reactivity with serpin B8 and α1PDX in a manner that depended on the serpin RCL-primed sequence. Modeling of the serpin B8-furin Michaelis complex identified serpin exosites in strand 3C close to the 298–300 loop whose substitution in α1PDX differentially affected furin reactivity depending on the furin loop and serpin RCL-primed sequences. These studies demonstrate that RCL-primed residues, strand 3C exosites, and the furin(298–300) loop are critical determinants of serpin reactivity with furin, which may be exploited in the design of specific and selective α1PDX inhibitors of PCs.  相似文献   

10.
Leucine-rich repeat G-protein-coupled receptors (LGRs) are a unique class of G-protein-coupled receptors characterized by a large extracellular domain to recognize ligands and regulate many important developmental processes. Among the three groups of LGRs, group B members (LGR4–6) recognize R-spondin family proteins (Rspo1–4) to stimulate Wnt signaling. In this study, we successfully utilized the “hybrid leucine-rich repeat technique,” which fused LGR4 with the hagfish VLR protein, to obtain two recombinant human LGR4 proteins, LGR415 and LGR49. We determined the crystal structures of ligand-free LGR415 and the LGR49-Rspo1 complex. LGR4 exhibits a twisted horseshoe-like structure. Rspo1 adopts a flat and β-fold architecture and is bound in the concave surface of LGR4 in the complex through electrostatic and hydrophobic interactions. All the Rspo1-binding residues are conserved in LGR4–6, suggesting that LGR4–6 bind R-spondins through an identical surface. Structural analysis of our LGR4-Rspo1 complex with the previously determined LGR4 and LGR5 structures revealed that the concave surface of LGR4 is the sole binding site for R-spondins, suggesting a one-site binding model of LGR4–6 in ligand recognition. The molecular mechanism of LGR4–6 is distinct from the two-step mechanism of group A receptors LGR1–3 and the multiple-interface binding model of group C receptors LGR7–8, suggesting LGRs utilize the divergent mechanisms for ligand recognition. Our structures, together with previous reports, provide a comprehensive understanding of the ligand recognition by LGRs.  相似文献   

11.
Carefully soaking crystals with Arg-Gly-Asp (RGD) peptides, we captured eight distinct RGD-bound conformations of the αIIbβ3 integrin headpiece. Starting from the closed βI domain conformation, we saw six intermediate βI conformations and finally the fully open βI with the hybrid domain swung out in the crystal lattice. The β1-α1 backbone that hydrogen bonds to the Asp side chain of RGD was the first element to move followed by adjacent to metal ion-dependent adhesion site Ca2+, α1 helix, α1’ helix, β6-α7 loop, α7 helix, and hybrid domain. We define in atomic detail how conformational change was transmitted over long distances in integrins, 40 Å from the ligand binding site to the opposite end of the βI domain and 80 Å to the far end of the hybrid domain. During these movements, RGD slid in its binding groove toward αIIb, and its Arg side chain became ordered. RGD concentration requirements in soaking suggested a >200-fold higher affinity after opening. The thermodynamic cycle shows how higher affinity pays the energetic cost of opening.  相似文献   

12.
Protease-activated receptor 1 (PAR1) is a G protein–coupled receptor (GPCR) for thrombin and promotes inflammatory responses through multiple pathways including p38 mitogen-activated protein kinase signaling. The mechanisms that govern PAR1-induced p38 activation remain unclear. Here, we define an atypical ubiquitin-dependent pathway for p38 activation used by PAR1 that regulates endothelial barrier permeability. Activated PAR1 K63-linked ubiquitination is mediated by the NEDD4-2 E3 ubiquitin ligase and initiated recruitment of transforming growth factor-β–activated protein kinase-1 binding protein-2 (TAB2). The ubiquitin-binding domain of TAB2 was essential for recruitment to PAR1-containing endosomes. TAB2 associated with TAB1, which induced p38 activation independent of MKK3 and MKK6. The P2Y1 purinergic GPCR also stimulated p38 activation via NEDD4-2–mediated ubiquitination and TAB1–TAB2. TAB1–TAB2-dependent p38 activation was critical for PAR1-promoted endothelial barrier permeability in vitro, and p38 signaling was required for PAR1-induced vascular leakage in vivo. These studies define an atypical ubiquitin-mediated signaling pathway used by a subset of GPCRs that regulates endosomal p38 signaling and endothelial barrier disruption.  相似文献   

13.
Two aptamers that bind separately with exosite I or exosite II of thrombin were studied for better understanding of the binding effect of aptamers on thrombin. CD and intrinsic fluorescence spectra indicated that after binding with aptamers the secondary structure of thrombin seemed unchanged, but the whole conformation of thrombin changed. The binding of aptamers on thrombin also made the catalytic activity of thrombin toward the chromogenic substrate (β-Ala-Gly-Arg-p-nitroanilide diacetate) increased. The present study indicated that the allostery of the two exosites seemed to be independent.  相似文献   

14.
Streptococcus pneumoniae is an important human pathogen that causes a range of disease states. Sialidases are important bacterial virulence factors. There are three pneumococcal sialidases: NanA, NanB, and NanC. NanC is an unusual sialidase in that its primary reaction product is 2-deoxy-2,3-didehydro-N-acetylneuraminic acid (Neu5Ac2en, also known as DANA), a nonspecific hydrolytic sialidase inhibitor. The production of Neu5Ac2en from α2–3-linked sialosides by the catalytic domain is confirmed within a crystal structure. A covalent complex with 3-fluoro-β-N-acetylneuraminic acid is also presented, suggesting a common mechanism with other sialidases up to the final step of product formation. A conformation change in an active site hydrophobic loop on ligand binding constricts the entrance to the active site. In addition, the distance between the catalytic acid/base (Asp-315) and the ligand anomeric carbon is unusually short. These features facilitate a novel sialidase reaction in which the final step of product formation is direct abstraction of the C3 proton by the active site aspartic acid, forming Neu5Ac2en. NanC also possesses a carbohydrate-binding module, which is shown to bind α2–3- and α2–6-linked sialosides, as well as N-acetylneuraminic acid, which is captured in the crystal structure following hydration of Neu5Ac2en by NanC. Overall, the pneumococcal sialidases show remarkable mechanistic diversity while maintaining a common structural scaffold.  相似文献   

15.
Spectrin dimer-tetramer interconversion is a critical contributor to red cell membrane stability, but some properties of spectrin tetramer formation cannot be studied effectively using monomeric recombinant domains. To address these limitations, a fused αβ mini-spectrin was produced that forms wild-type divalent tetramer complexes. Using this mini-spectrin, a medium-resolution structure of a seven-repeat bivalent tetramer was produced using homology modeling coupled with chemical cross-linking. Inter- and intramolecular cross-links provided critical distance constraints for evaluating and optimizing the best conformational model and appropriate docking interfaces. The two strands twist around each other to form a super-coiled, rope-like structure with the AB helix face of one strand associating with the opposing AC helix face. Interestingly, two tetramer site hereditary anemia mutations that exhibit wild-type binding in univalent head-to-head assays are located in the interstrand region. This suggests that perturbations of the interstrand region can destabilize spectrin tetramers and the membrane skeleton. The α subunit N-terminal cross-links to multiple sites on both strands, demonstrating that this non-homologous tail remains flexible and forms heterogeneous structures in the tetramer complex. Although no cross-links were observed involving the β subunit non-homologous C-terminal tail, several cross-links were observed only when this domain was present, suggesting it induces subtle conformational changes to the tetramer site region. This medium-resolution model provides a basis for further studies of the bivalent spectrin tetramer site, including analysis of functional consequences of interstrand interactions and mutations located at substantial molecular distances from the tetramer site.  相似文献   

16.
The periplasmic binding protein (PBP) FepB plays a key role in transporting the catecholate siderophore ferric enterobactin from the outer to the inner membrane in Gram-negative bacteria. The solution structures of the 34-kDa apo- and holo-FepB from Escherichia coli, solved by NMR, represent the first solution structures determined for the type III class of PBPs. Unlike type I and II PBPs, which undergo large “Venus flytrap” conformational changes upon ligand binding, both forms of FepB maintain similar overall folds; however, binding of the ligand is accompanied by significant loop movements. Reverse methyl cross-saturation experiments corroborated chemical shift perturbation results and uniquely defined the binding pocket for gallium enterobactin (GaEnt). NMR relaxation experiments indicated that a flexible loop (residues 225–250) adopted a more rigid and extended conformation upon ligand binding, which positioned residues for optimal interactions with the ligand and the cytoplasmic membrane ABC transporter (FepCD), respectively. In conclusion, this work highlights the pivotal role that structural dynamics plays in ligand binding and transporter interactions in type III PBPs.  相似文献   

17.
Human microsomal cytochrome P450 (CYP) 2E1 is widely known for its ability to oxidize >70 different, mostly compact, low molecular weight drugs and other xenobiotic compounds. In addition CYP2E1 oxidizes much larger C9–C20 fatty acids that can serve as endogenous signaling molecules. Previously structures of CYP2E1 with small molecules revealed a small, compact CYP2E1 active site, which would be insufficient to accommodate medium and long chain fatty acids without conformational changes in the protein. In the current work we have determined how CYP2E1 can accommodate a series of fatty acid analogs by cocrystallizing CYP2E1 with ω-imidazolyl-octanoic fatty acid, ω-imidazolyl-decanoic fatty acid, and ω-imidazolyl-dodecanoic fatty acid. In each structure direct coordination of the imidazole nitrogen to the heme iron mimics the position required for native fatty acid substrates to yield the ω-1 hydroxylated metabolites that predominate experimentally. In each case rotation of a single Phe298 side chain merges the active site with an adjacent void, significantly altering the active site size and topology to accommodate fatty acids. The binding of these fatty acid ligands is directly opposite the channel to the protein surface and the binding observed for fatty acids in the bacterial cytochrome P450 BM3 (CYP102A1) from Bacillus megaterium. Instead of the BM3-like binding mode in the CYP2E1 channel, these structures reveal interactions between the fatty acid carboxylates and several residues in the F, G, and B′ helices at successive distances from the active site.  相似文献   

18.
Thrombin interaction with platelet membrane glycoprotein Ib alpha   总被引:1,自引:0,他引:1  
The interaction of thrombin with platelet glycoprotein Ibalpha (GPIb alpha) is required for optimal platelet activation. The crystal structures of platelet GPIb alpha bound to thrombin reported by Dumas et al. and Celikel et al. both reveal the simultaneous interaction of GPIb alpha with thrombin exosites I and II but differ markedly regarding how the two proteins interact. The possible consequences on thrombus formation of thrombin interacting with GPIb alpha are discussed in light of these new data.  相似文献   

19.
Thrombin exists in two allosteric forms, slow (S) and fast (F), that recognize natural substrates and inhibitors with significantly different affinities. Because under physiologic conditions the two forms are almost equally populated, investigation of thrombin function must address the contribution from the S and F forms and the molecular origin of their differential recognition of ligands. Using a panel of 79 Ala mutants, we have mapped for the first time the epitopes of thrombin recognizing a macromolecular ligand, hirudin, in the S and F forms. Hirudin binding is a relevant model for the interaction of thrombin with fibrinogen and PAR1 and is likewise influenced by the allosteric S-->F transition. The epitopes are nearly identical and encompass two hot spots, one in exosite I and the other in the Na+ site at the opposite end of the protein. The higher affinity of the F form is due to the preferential interaction of hirudin with Lys-36, Leu-65, Thr-74, and Arg-75 in exosite I; Gly-193 in the oxyanion hole; and Asp-221 and Asp-222 in the Na+ site. Remarkably, no correlation is found between the energetic and structural involvements of thrombin residues in hirudin recognition, which invites caution in the analysis of protein-protein interactions in general.  相似文献   

20.
Remodeling of collagen by matrix metalloproteinases (MMPs) is crucial to tissue homeostasis and repair. MMP-13 is a collagenase with a substrate preference for collagen II over collagens I and III. It recognizes a specific, well-known site in the tropocollagen molecule where its binding locally perturbs the triple helix, allowing the catalytic domain of the active enzyme to cleave the collagen α chains sequentially, at Gly775–Leu776 in collagen II. However, the specific residues upon which collagen recognition depends within and surrounding this locus have not been systematically mapped. Using our triple-helical peptide Collagen Toolkit libraries in solid-phase binding assays, we found that MMP-13 shows little affinity for Collagen Toolkit III, but binds selectively to two triple-helical peptides of Toolkit II. We have identified the residues required for the adhesion of both proMMP-13 and MMP-13 to one of these, Toolkit peptide II-44, which contains the canonical collagenase cleavage site. MMP-13 was unable to bind to a linear peptide of the same sequence as II-44. We also discovered a second binding site near the N terminus of collagen II (starting at helix residue 127) in Toolkit peptide II-8. The pattern of binding of the free hemopexin domain of MMP-13 was similar to that of the full-length enzyme, but the free catalytic subunit bound none of our peptides. The susceptibility of Toolkit peptides to proteolysis in solution was independent of the very specific recognition of immobilized peptides by MMP-13; the enzyme proved able to cleave a range of dissolved collagen peptides.  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