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 29 毫秒
1.
The nicotinic acetylcholine receptor (nAChR) subtype α6β2* (the asterisk denotes the possible presence of additional subunits) has been identified as an important molecular target for the pharmacotherapy of Parkinson disease and nicotine dependence. The α6 subunit is closely related to the α3 subunit, and this presents a problem in designing ligands that discriminate between α6β2* and α3β2* nAChRs. We used positional scanning mutagenesis of α-conotoxin PeIA, which targets both α6β2* and α3β2*, in combination with mutagenesis of the α6 and α3 subunits, to gain molecular insights into the interaction of PeIA with heterologously expressed α6/α3β2β3 and α3β2 receptors. Mutagenesis of PeIA revealed that Asn11 was located in an important position that interacts with the α6 and α3 subunits. Substitution of Asn11 with a positively charged amino acid essentially abolished the activity of PeIA for α3β2 but not for α6/α3β2β3 receptors. These results were used to synthesize a PeIA analog that was >15,000-fold more potent on α6/α3β2β3 than α3β2 receptors. Analogs with an N11R substitution were then used to show a critical interaction between the 11th position of PeIA and Glu152 of the α6 subunit and Lys152 of the α3 subunit. The results of these studies provide molecular insights into designing ligands that selectively target α6β2* nAChRs.  相似文献   

2.
α6β2 Nicotinic acetylcholine receptors (nAChRs) expressed by dopaminergic neurons in the CNS are potential therapeutic targets for the treatment of several neuropsychiatric diseases, including nicotine addiction and Parkinson disease. However, recent studies indicate that the α6 subunit can also associate with the β4 subunit to form α6β4 nAChRs that are difficult to pharmacologically distinguish from α6β2, α3β4, and α3β2 subtypes. The current study characterized a novel 16-amino acid α-conotoxin (α-CTx) TxIB from Conus textile whose sequence is GCCSDPPCRNKHPDLC-amide as deduced from gene cloning. The peptide and an analog with an additional C-terminal glycine were chemically synthesized and tested on rat nAChRs heterologously expressed in Xenopus laevis oocytes. α-CTx TxIB blocked α6/α3β2β3 nAChR with an IC50 of 28 nm. In contrast, the peptide showed little or no block of other tested subtypes at concentrations up to 10 μm. The three-dimensional solution structure of α-CTx TxIB was determined using NMR spectroscopy. α-CTx TxIB represents a uniquely selective ligand for probing the structure and function of α6β2 nAChRs.  相似文献   

3.
δ subunit-containing γ-aminobutyric acid, type A (GABAA)receptors are expressed extrasynaptically and mediate tonic inhibition. In cerebellar granule cells, they often form receptors together with α1 and/or α6 subunits. We were interested in determining the architecture of receptors containing both subunits. We predefined the subunit arrangement of several different GABAA receptor pentamers by concatenation. These receptors composed of α1, α6, β3, and δ subunits were expressed in Xenopus oocytes. Currents elicited in response to GABA were determined in the presence and absence of 3α,21-dihydroxy-5α-pregnan-20-one (THDOC) or ethanol, or currents were elicited by 4,5,6,7-tetrahydroisoxazolo[5,4-c]-pyridin-3-ol (THIP). Several subunit configurations formed active channels. We therefore conclude that δ can assume multiple positions in a receptor pentamer made up of α1, α6, β3, and δ subunits. The different receptors differ in their functional properties. Functional expression of one receptor type was only evident in the combined presence of the neurosteroid THDOC with the channel agonist GABA. Most, but not all, receptors active with GABA/THDOC responded to THIP. None of the receptors was modulated by ethanol concentrations up to 30 mm. Several observations point to a preferred position of δ subunits between two α subunits in α1α6β3δ receptors. This property is shared by α1β3δ and α6β3δ receptors, but there are differences in the additionally expressed isoforms.  相似文献   

4.
Minimal photosynthetic catalytic F1() core complexes, containing equimolar ratios of the and subunits, were isolated from membrane-bound spinach chloroplast CF1 and Rhodospirillum rubrum chromatophore RrF1. A CF1-33 hexamer and RrF1-11 dimer, which were purified from the respective F1() complexes, exhibit lower rates and different properties from their parent F1-ATPases. Most interesting is their complete resistance to inhibition by the general F1 inhibitor azide and the specific CF1 inhibitor tentoxin. These inhibitors were earlier reported to inhibit multisite, but not unisite, catalysis in all sensitive F1-ATPases and were therefore suggested to block catalytic site cooperativity. The absence of this typical property of all F1-ATPases in the 11 dimer is consistant with the view that the dimer contains only a single catalytic site. The 33 hexamer contains however all F1 catalytic sites. Therefore the observation that CF1-33 can bind tentoxin and is stimulated by it suggests that the F1 subunit, which is required for obtaining inhibition by tentoxin as well as azide, plays an important role in the cooperative interactions between the F1-catalytic sites.Abbreviations CF0F1 chloroplast F0F1 - CF1 chloroplast F1 - CF1 chloroplast F1 subunit - CF1 chloroplast F1 subunit - CF1() a complex containing equal amounts of the CF1 and subunits - MF1 mitochondrial F1 - RrF0F1 Rhodospirillum rubrum F0F1 - RrF1 R. rubrum F1 - RrF1 R. rubrum F1 subunit - RrF1 R. rubrum F1 subunit - RrF1() a complex containing equal amounts of the RrF1 and subunits - Rubisco Ribulose-1,5-bisphosphate carboxylase - TF1 thermophilic bacterium PS3 F1  相似文献   

5.
Background information. Previous studies have reported that cross‐talk between integrins may be an important regulator of integrin—ligand binding and subsequent signalling events that control a variety of cell functions in many tissues. We previously demonstrated that αvβ5/β6 integrin represses α2β1‐dependent cell migration. The αv subunits undergo an endoproteolytic cleavage by protein convertases, whose role in tumoral invasion has remained controversial. Results. Inhibition of convertases by the convertase inhibitor α1‐PDX (α1‐antitrypsin Portland variant), leading to the cell‐surface expression of an uncleaved form of the αv integrin, stimulated cell migration toward type I collagen. Under convertase inhibition, α2β1 engagement led to enhanced phosphorylation of both FAK (focal adhesion kinase) and MAPK (mitogen‐activated protein kinase). This outside‐in signalling stimulation was associated with increased levels of activated β1 integrin located in larger than usual focal‐adhesion structures and a cell migration that was independent of the PI3K (phosphoinositide 3‐kinase)/Akt (also called protein kinase B) pathway. Conclusions. The increase in cell migration observed upon convertases inhibition appears to be due to the up‐regulation of β1 integrins and to their location in larger focal‐adhesion structures. The endoproteolytic cleavage of αv subunits is necessary for αvβ5/β6 integrin to control α2β1 function and could thus play an essential role in colon cancer cell migration.  相似文献   

6.
Under physiological conditions, circulating platelets are discoid in shape.1 On these platelets, the fibrinogen receptor (integrin αIIbβ3) is in a low-affinity state, unable to bind soluble fibrinogen (Fg). Activation by agonists such as ADP and thrombin leads to a change in the conformation of the integrin αIIbβ3 through a process known as inside-out signaling. This enables the integrin to bind soluble Fg, which initiates a cascade of events referred to as outside-in signaling.2 Outside-in signaling control processes, such as platelet spreading and clot retraction, by regulating small G-proteins such as RhoA, Rac and cdc42.Key words: platelets, integrin αIIbβ3, Galpha13, RhoA, clot retraction, thrombin, fibrinogenThe majority of the physiological platelet agonists (except collagen) induce inside-out signaling by binding to specific G-protein-coupled receptors (GPCRs). A G-protein plays a crucial role in translating the signal from GPCR to downstream effector molecules, ultimately leading to affinity modulation of integrin αIIbβ3. Platelets express nine Gα subunits; namely Gq, Gi1, Gi2, Gi3, Gz, G12, G13, Gs and G16. Previous studies have shown that a small G-protein, RhoA, is activated by the G12/13 family and plays a crucial role in calcium-independent platelet shape change.3 However, RhoA is also activated by αIIbβ3 and inhibits platelet spreading to trigger clot retraction.4 Recently, in a series of elegant experiments, Gong et al. have described the dynamic regulation of RhoA through a signaling crosstalk between Gα13 and αIIbβ3.5By generating mice in which the platelets were depleted of Gα13 using siRNA technology, Gong et al. investigated the role of Gα13-mediated signaling on platelet spreading on immobilized Fg.5 The confocal images very clearly showed that, in the absence of Gα13, platelets spread poorly on Fg, which was rescued by pretreatment with the Rho-kinase inhibitor Y27632, confirming previous findings that RhoA activated downstream of integrin αIIbβ3 inhibits platelet spreading. Interestingly, Gα13-depleted platelets failed to activate c-Src but accelerated RhoA activation. From these observations, the authors infer that Gα13 is important for integrin-mediated c-Src activation and RhoA inhibition, leading to increased cell spreading.5Since Gα13 regulates integrin-mediated cell spreading and c-Src activation, Gong et al. examined the interaction of Gα13 with αIIbβ3 using co-immunoprecipitation and GST pull-down assays.5 They found that the GTP-bound form of Gα13 shows enhanced interaction with the integrin β3 subunit. This interaction is required for the activation of c-Src and the inhibition of RhoA. However, they found that the inhibition of RhoA is transient. RhoA activation is suppressed for the first 15 min of platelet spreading, after which RhoA is activated. This initial suppression is rescued by blocking Gα13 and β3 cytoplasmic domain (β3-CD) interaction. Furthermore, they observed that RhoA activation parallels clot retraction.5 These findings indicate that Gα13 is a key regulator of platelet spreading and clot retraction phenomena.According to Gong et al., thrombin-induced inside-out signaling through GPCR leads to GTP loading of Gα13 (Fig. 1A). This GTP-bound Gα13 interacts with integrin β3-CD of ligand-bound integrin, thus facilitating c-Src activation, which leads to platelet spreading. Blockade of the interaction between Gα13 and β3-CD or cleavage of β3-CD by calpain results in clot retraction (Fig. 1B).Open in a separate windowFigure 1Schematic representation of the dynamic regulation of RhoA by Gα13 during platelet activation. (A) Activation of platelets by thrombin receptors coupled to Gα13 leads to the activation of RhoA, leading to platelet shape change. (B) The change in the conformation of integrin to a high-affinity form results in fibrinogen binding to αIIbβ3. Active Gα13 binds to the cytoplasmic domain of β3 leading to the activation of c-Src, resulting in platelet spreading. The rise in intracellular calcium activates calpain, which cleaves the β3 cytoplasmic domain, releasing c-Src, which, resulting in the activation of RhoA, leads to cell retraction. *Denotes GTP-bound active form of G-proteins.Perhaps the most significant and novel finding of the study is the identification of integrin αIIbβ3 as an effector of Gα13. The study also convincingly shows that Gα13 bound to integrin regulates RhoA via c-Src. Furthermore, achieving 80% knockdown of Gα13 in an in vivo setting using siRNA represents a technological advancement. Since Gα13 binds to integrin β3-CD in a 1:1 stoichiometry, it appears that only a small population of integrin is regulated by Gα13, as there are far less Gα13 molecules in a single platelet than the number of αIIbβ3 molecules. This will require further investigation. Gong et al. also finds that an appreciable amount of Gα13 is associated with β3 in resting platelets, which requires some explanation.5 It is also not clear if Gα13 remains bound to β3-CD or dissociates from the integrin during clot retraction.Overall, this is a paradigm-shifting study that establishes the importance of the dynamic regulation of RhoA by Gα13 in order to achieve efficient platelet spreading and clot retraction.  相似文献   

7.
Although type IV collagen is heavily glycosylated, the influence of this post-translational modification on integrin binding has not been investigated. In the present study, galactosylated and nongalactosylated triple-helical peptides have been constructed containing the α1(IV)382–393 and α1(IV)531–543 sequences, which are binding sites for the α2β1 and α3β1 integrins, respectively. All peptides had triple-helical stabilities of 37 °C or greater. The galactosylation of Hyl393 in α1(IV)382–393 and Hyl540 and Hyl543 in α1(IV)531–543 had a dose-dependent influence on melanoma cell adhesion that was much more pronounced in the case of α3β1 integrin binding. Molecular modeling indicated that galactosylation occurred on the periphery of α2β1 integrin interaction with α1(IV)382–393 but right in the middle of α3β1 integrin interaction with α1(IV)531–543. The possibility of extracellular deglycosylation of type IV collagen was investigated, but no β-galactosidase-like activity capable of collagen modification was found. Thus, glycosylation of collagen can modulate integrin binding, and levels of glycosylation could be altered by reduction in expression of glycosylation enzymes but most likely not by extracellular deglycosylation activity.  相似文献   

8.
The ATPase activity of the F1-ATPase from the thermophilic bacterium PS3 is stimulated at concentrations of rhodamine 6G up to about 10 µM where 70% stimulation is observed at 36°C. Half maximal stimulation is observed at about 3 µM dye. At rhodamine 6G concentrations greater than 10 µM, ATPase activity declines with 50% inhibition observed at about 75 µM dye. The ATPase activities of the 33 and 33 complexes assembled from isolated subunits of TF1 expressed inE. coli deleted of theunc operon respond to increasing concentrations of rhodamine 6G nearly identically to the response of TF1. In contrast, the ATPase activities of the 33 and 33 complexes are only inhibited by rhodamine 6G with 50% inhibition observed, respectively, at 35 and 75 µM dye at 36°C. The ATPase activity of TF1 is stimulated up to 4-fold by the neutral detergent, LDAO. In the presence of stimulating concentrations of LDAO, the ATPase activity of TF1 is no longer stimulated by rhodamine 6G, but rather, it is inhibited with 50% inhibition observed at about 30 µM dye at 30°C. One interpretation of these results is that binding of rhodamine 6G to a high-affinity site on TF1 stimulates ATPase activity and unmasks a low-affinity, inhibitory site for the dye which is also exposed by LDAO.  相似文献   

9.
The stereoselective syntheses of 2-cyclopropyl- and (2S)-2-hydroxymethyl-(3R,4S)-4-hydroxy-β3-homoproline are described. The reported amino acids were constructed through 1,3-dipolar cycloaddition of strained alkylidenecyclopropanes with enantiopure pyrroline N-oxides derived from malic acid followed by thermal rearrangement of the adducts in the presence of trifluoroacetic acid. The two-step sequence afforded the homoprolines suitably protected to be directly used as building blocks in peptidomimetic synthesis as proved by the synthesis of the two model mixed α/β/α tripeptides Phe-β3-HPro-Val.  相似文献   

10.
Human (α6β2)(α4β2)β3 nicotinic acetylcholine receptors (AChRs) are essential for addiction to nicotine and a target for drug development for smoking cessation. Expressing this complex AChR is difficult, but has been achieved using subunit concatamers. In order to determine what limits expression of α6* AChRs and to efficiently express α6* AChRs using free subunits, we investigated expression of the simpler (α6β2)2β3 AChR. The concatameric form of this AChR assembles well, but is transported to the cell surface inefficiently. Various chimeras of α6 with the closely related α3 subunit increased expression efficiency with free subunits and produced pharmacologically equivalent functional AChRs. A chimera in which the large cytoplasmic domain of α6 was replaced with that of α3 increased assembly with β2 subunits and transport of AChRs to the oocyte surface. Another chimera replacing the unique methionine 211 of α6 with leucine found at this position in transmembrane domain 1 of α3 and other α subunits increased assembly of mature subunits containing β3 subunits within oocytes. Combining both α3 sequences in an α6 chimera increased expression of functional (α6β2)2β3 AChRs to 12-fold more than with concatamers. This is pragmatically useful, and provides insights on features of α6 subunit structure that limit its expression in transfected cells.  相似文献   

11.
Acetylcholine-based neurotransmission is regulated by cationic, ligand-gated ion channels called nicotinic acetylcholine receptors (nAChRs). These receptors have been linked to numerous neurological diseases and disorders such as Alzheimer's disease, Parkinson's disease, and nicotine addiction. Recently, a class of compounds has been discovered that antagonize nAChR function in an allosteric fashion. Models of human α4β2 and α3β4 nicotinic acetylcholine receptor (nAChR) extracellular domains have been developed to computationally explore the binding of these compounds, including the dynamics and free energy changes associated with ligand binding. Through a blind docking study to multiple receptor conformations, the models were used to determine a putative binding mode for the negative allosteric modulators. This mode, in close proximity to the agonist binding site, is presented in addition to a hypothetical mode of antagonism that involves obstruction of C loop closure. Molecular dynamics simulations and MM-PBSA free energy of binding calculations were used as computational validation of the predicted binding mode, while functional assays on wild-type and mutated receptors provided experimental support. Based on the proposed binding mode, two residues on the β2 subunit were independently mutated to the corresponding residues found on the β4 subunit. The T58K mutation resulted in an eight-fold decrease in the potency of KAB-18, a compound that exhibits preferential antagonism for human α4β2 over α3β4 nAChRs, while the F118L mutation resulted in a loss of inhibitory activity for KAB-18 at concentrations up to 100 μM. These results demonstrate the selectivity of KAB-18 for human α4β2 nAChRs and validate the methods used for identifying the nAChR modulator binding site. Exploitation of this site may lead to the development of more potent and subtype-selective nAChR antagonists which may be used in the treatment of a number of neurological diseases and disorders.  相似文献   

12.
13.
Condensation of benzyl 2-acetamido-4,6-O-benzylidene-2-deoxy-α-d-galactopyranoside with 2,3,4-tri-O-acetyl-α-d-fucopyranosyl bromide in 1:1 nitromethane-benzene, in the presence of powdered mercuric cyanide, afforded benzyl 2-acetamido-4,6-O-benzylidene-2-deoxy-3-O-(2,3,4-tri-O-acetyl-β-d-fucopyranosyl)-α-d-galactopyranoside (3). Cleavage of the benzylidene group of 3 with hot, 60% aqueous acetic acid afforded diol 4, which, on deacetylation, furnished the disaccharide 5. Condensation of diol 4 with 2-methyl-(3,4,6-tri-O-acetyl-1,2-di-deoxy-α-d-glucopyrano)-[2,1-d]-2-oxazoline in 1,2-dichloroethane afforded the trisaccharide derivative (7). Deacetylation of 7 with Amberlyst A-26 (OH?) anion-exchange resin in methanol gave the title trisaccharide (8). The structures of 5 and 8 were confirmed by 13C-n.m.r. spectroscopy.  相似文献   

14.
The basic structures of the catalytic portion (F1, 33) of ATP synthase are the 33 hexamer (oligomer with cooperativity) and 11 heterodimer (protomer). These were reconstituted from the and subunits of thermophilic F1 (TF1), and the 33 hexamer was crystallized. On electrophoresis, both the dimer and hexamer showed bands with ATPase activity. Using the dimer and hexamer, we studied the nucleotide-dependent rapid molecular dynamics. The formation of the hexamer required neither nucleotide nor Mg. The hexamer was dissociated into the dimer in the presence of MgADP, while the dimer was associated into the hexamer in the presence of MgATP. The hexamer, like mitochondrial F1 and TF1, showed two kinds of ATPase activity: one was cooperative and was inhibited by only one BzADP per hexamer, and the other was inhibited by three BzADP per hexamer.  相似文献   

15.
Human parechovirus (HPEV) infections are very common in early childhood and can be severe in neonates. It has been shown that integrins are important for cellular infectivity of HPEV1 through experiments using peptide blocking assays and function-blocking antibodies to αV integrins. The interaction of HPEV1 with αV integrins is presumably mediated by a C-terminal RGD motif in the capsid protein VP1. We characterized the binding of integrins αVβ3 and αVβ6 to HPEV1 by biochemical and structural studies. We showed that although HPEV1 bound efficiently to immobilized integrins, αVβ6 bound more efficiently than αVβ3 to immobilized HPEV1. Moreover, soluble αVβ6, but not αVβ3, blocked HPEV1 cellular infectivity, indicating that it is a high-affinity receptor for HPEV1. We also showed that HPEV1 binding to integrins in vitro could be partially blocked by RGD peptides. Using electron cryo-microscopy and image reconstruction, we showed that HPEV1 has the typical T=1 (pseudo T=3) organization of a picornavirus. Complexes of HPEV1 and integrins indicated that both integrin footprints reside between the 5-fold and 3-fold symmetry axes. This result does not match the RGD position predicted from the coxsackievirus A9 X-ray structure but is consistent with the predicted location of this motif in the shorter C terminus found in HPEV1. This first structural characterization of a parechovirus indicates that the differences in receptor binding are due to the amino acid differences in the integrins rather than to significantly different viral footprints.Picornaviruses consist of a positive-sense, single-stranded infectious RNA genome of approximately 7.3 kb enclosed in a capsid composed of 60 copies of each of the three or four capsid proteins (VP1 to VP4). Human parechovirus 1 (HPEV1) is a member of the Parechovirus genus of the Picornaviridae family (38, 70). There are currently eight completely sequenced human parechovirus types and 14 described types (4, 19, 24, 30, 38, 39, 51, 58, 78). In addition, the Parechovirus genus currently has four Ljungan virus members that infect rodents. HPEV1 exhibits several distinct molecular characteristics compared to other picornaviruses (38, 71). These include the lack of the maturation cleavage of the capsid proteins VP0 to VP4 (N-terminal) and VP2 (C-terminal), existence of an approximately 30-amino-acid-long extension to the N terminus of VP3, a unique nonstructural protein 2A, and a 5′ untranslated region that is more closely related to picornaviruses infecting animals than those infecting humans.HPEV infections are common during the first years of life and are often mild or asymptomatic (20, 28, 42, 73, 80). Recently, a number of new types have been identified, and their prevalence in stool samples, for example, highlights their clinical importance. Normally, they cause gastroenteritis and respiratory infections, but severe illnesses, such as infections of the central nervous system, generalized infections of neonates, and myocarditis, have also been associated with HPEV infections (1, 8, 10, 28, 80). Currently, the role of the unique molecular, structural, and antigenic characteristics of HPEVs in the pathogenesis of infection is unknown.HPEV types 1, 2, 4, 5, and 6 are known to possess an RGD motif near the C terminus of VP1 that is known to facilitate binding of cellular ligands (e.g., fibronectin) to αv integrins. The motif is in an analogous position to motifs in coxsackievirus A9 (CAV9) and echovirus 9 (EV9; Barty strain) (Fig. (Fig.1).1). The role of the RGD sequence in cellular entry and subsequent replication of HPEV1 has been shown through blocking assays with RGD-containing peptides, mutation of the sequence, and function-blocking antibodies to αv integrins (11, 43, 62, 71). These results strongly suggested that αv integrins play a central role in the initiation of HPEV1 infection. Direct involvement of αv integrins in the infectious entry of HPEV1 was further confirmed by overexpression of human αvβ1 and αvβ3 integrins in Chinese hamster ovary (CHO) cells, allowing successful virus infection (74). There are no reports yet on the identification of receptors for the HPEV types lacking the RGD motif (HPEV3, HPEV7, and HPEV8) (19, 39, 51).Open in a separate windowFIG. 1.Sequence alignments. Amino acid sequence alignment of the viral coat protein VP1 from different picornaviruses with the CAV9 secondary structure derived from the atomic model displayed above the alignment (34). The columns boxed in blue with red letters signify similarity, and the red column signifies identity. There is limited similarity between HPEV and other picornaviruses. C-terminal RGD motifs are boxed in red.Although the crystal structures of several picornaviruses have been determined (3, 26, 34, 35, 44, 57, 59, 65, 68, 72) and the receptor interactions have been studied in detail by X-ray crystallography, electron cryo-microscopy (cryo-EM), and three-dimensional (3D) image reconstruction (6, 9, 23, 31, 32, 47, 83), there is no structural information available for the parechoviruses or parechovirus-receptor complexes. Here, we compare the binding of αVβ3 and αVβ6 to HPEV1 in vitro by biochemical assays and determine the structures of HPEV1 and the corresponding HPEV1-integrin complexes.  相似文献   

16.
The neutral urinary excretion products of 17β-hydroxy-2α,3α-cyclopropano-5α-androstane from the rabbit, dosed orally, were investigated. Column chromatography yielded five crystalline metabolites which were identified by GLC and spectroscopic measurements. Three of these substances were hydroxylated in the 4α-position and one in the 6a-position with the cyclopropane ring intact. The fifth substance, 17β-hydroxy-3β-methyl-5α-androstan-2-one, can be derived from initial hydroxylation of the cyclopropane ring at C-2 followed by ring opening. The dosed substance and triol material was shown to be present by GLC and m.s. measurements. GLC determinations show that hydroxylation has occurred at C-4?C-6>C-2.  相似文献   

17.
Lee C  Lee SH  Kim DH  Han KH 《BMB reports》2012,45(5):275-280
Nicotinic acetylcholine receptors (nAChRs) are a diverse family of homo- or heteropentameric ligand-gated ion channels. Understanding the physiological role of each nAChR subtype and the key residues responsible for normal and pathological states is important. α-Conotoxin neuropeptides are highly selective probes capable of discriminating different subtypes of nAChRs. In this study, we performed homology modeling to generate the neuronal α3, β2 and β4 subunits using the x-ray structure of the α1 subunit as a template. The structures of the extracellular domains containing ligand binding sites in the α3β2 and α3β4 nAChR subtypes were constructed using MD simulations and ligand docking processes in their free and ligand-bound states using α-conotoxin GIC, which exhibited the highest α3β2 vs. α3β4 discrimination ratio. The results provide a reasonable structural basis for such a discriminatory ability, supporting the idea that the present strategy can be used for future investigations on nAChR-ligand complexes.  相似文献   

18.
Mouse B lymphocytes express several nicotinic acetylcholine receptor (nAChR) subtypes, their exact functions being not clearly understood. Here we show that α7 nAChR was present in about 60%, while α4β2 and α9(α10) nAChRs in about 10% and 20% of mouse spleen B lymphocytes, respectively; Balb/c and C57Bl/6 mice possessed different relative amounts of these nAChR subtypes. α4β2 and α7, but not α9(α10) nAChRs, were up-regulated upon B lymphocyte activation in vitro. Flow cytometry and sandwich ELISA studies demonstrated that α7 and α9(α10) nAChRs are coupled to CD40, whereas α4β2 nAChR is coupled to IgM. B lymphocytes of both α7(-/-) and β2(-/-) mice responded to anti-CD40 stronger than those of the wild-type mice, whereas the cells of β2(-/-) mice responded to anti-IgM worse than those of the wild-type or α7(-/-) mice. Inhibition of α7 and α9(α10) nAChRs with methyllicaconitine resulted in considerable augmentation of CD40-mediated B lymphocyte proliferation in cells of all genotypes; stimulation of α4β2 nAChRs with epibatidine increased the IgM-mediated proliferation of the wild-type and α7(-/-), but not β2(-/-) cells. Inhibition of α9(α10) nAChRs with α-conotoxin PeAI exerted weak stimulating effect on CD40-mediated proliferation. This nAChR subtype was up-regulated in α7(-/-) B-cells. α7 nAChRs were found recruited to immune synapses between human T and B lymphocytes, both of which produced acetylcholine. It is concluded that α7 nAChR fulfills inhibitory CD40-related mitogenic function, α4β2 nAChR produces a stimulatory IgM-related effect, while α9α10 nAChR is a "reserve" receptor, which partly compensates the absence of α7 nAChR in α7(-/-) cells. Acetylcholine is an additional mediator to modulate activation of interacting T and B lymphocytes.  相似文献   

19.
Datura meteloides; plants were fed with tiglic acid-[-14C] via the roots and after 2 days the percentage incorporation into the alkaloids 3α-tigloyloxytropane, 3α,6β-ditigloyloxytropane, meteloidine and 3α,6β-ditigloyloxytropan-7β-ol were 15·2, 1·82, 2·2 and 1·8 respectively. 3α,6β-Ditigloyloxytropane was partially hydrolysed to 6β-hydroxy-3α-tigloyloxytropane which contained 58·1% of the radioactivity of the original base, whereas 3α,6β-ditigloyloxytropan-7β-ol gave meteloidine containing only 9·2% of the original activity. The results suggest that the di- and tri-hydroxytropanes may be formed by different routes.  相似文献   

20.
姜泓  白雪帆 《病毒学报》2007,23(6):494-497
整合素αvβ3是一类表达于细胞表面的跨膜糖蛋白粘附分子,由α和β两种Ⅰ型膜蛋白亚单位以非共价键形式连接形成异源二聚体分子。整合素αvβ3可表达于多种细胞,细胞外信号通过不同分子可与其发生相互作用,经整合素αvβ3将细胞外信号传递至细胞内,引起钙离子、Pyk2和磷脂酰肌醇-3(PI-3)激酶等细胞内信号发生变化。整合素αvβ3在血管生成、胚胎发育、肿瘤转移、免疫应答等多种生理和病理过和中发挥着重要的作用。近几年,整合素αvβ3与病毒感染的相关研究进展迅速,本文就整合素αvβ3与病毒感染作一综述。1整合素αvβ3概述1·1整合素αv…  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