首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Takagi J  Petre BM  Walz T  Springer TA 《Cell》2002,110(5):599-511
How ligand binding alters integrin conformation in outside-in signaling, and how inside-out signals alter integrin affinity for ligand, have been mysterious. We address this with electron microscopy, physicochemical measurements, mutational introduction of disulfides, and ligand binding to alphaVbeta3 and alphaIIbbeta3 integrins. We show that a highly bent integrin conformation is physiological and has low affinity for biological ligands. Addition of a high affinity ligand mimetic peptide or Mn(2+) results in a switchblade-like opening to an extended structure. An outward swing of the hybrid domain at its junction with the I-like domain shows conformational change within the headpiece that is linked to ligand binding. Breakage of a C-terminal clasp between the alpha and beta subunits enhances Mn(2+)-induced unbending and ligand binding.  相似文献   

2.
The glycosylphosphatidylinositol-linked urokinase-type plasminogen activator receptor (uPAR) interacts with the heterodimer cell adhesion molecules integrins to modulate cell adhesion and migration. Devoid of a cytoplasmic domain, uPAR triggers intracellular signaling via its associated molecules that contain cytoplasmic domains. Interestingly, uPAR changes the ectodomain conformation of one of its partner molecules, integrin alpha(5)beta(1), and elicits cytoplasmic signaling. The separation or reorientation of integrin transmembrane domains and cytoplasmic tails are required for integrin outside-in signaling. However, there is a lack of direct evidence showing these conformational changes of an integrin that interacts with uPAR. In this investigation we used reporter monoclonal antibodies and fluorescence resonance energy transfer analyses to show conformational changes in the alpha(M)beta(2) headpiece and reorientation of its transmembrane domains when alpha(M)beta(2) interacts with uPAR.  相似文献   

3.
Cysteine-rich repeats in the integrin beta subunit stalk region relay activation signals to the ligand-binding headpiece. The NMR solution structure and disulfide bond connectivity of Cys-rich module-3 of the integrin beta2 subunit reveal a nosecone-shaped variant of the EGF fold, termed an integrin-EGF (I-EGF) domain. Interdomain contacts between I-EGF domains 2 and 3 observed by NMR support a model in which the modules are related by an approximate two-fold screw axis in an extended arrangement. Our findings complement a 3.1 A crystal structure of the extracellular portion of integrin alphaVbeta3, which lacks an atomic model for I-EGF2 and a portion of I-EGF3. The disulfide connectivity of I-EGF3 chemically assigned here differs from the pairings suggested in the alphaVbeta3 structure. Epitopes that become exposed upon integrin activation and residues that restrain activation are defined in beta2 I-EGF domains 2 and 3. Superposition on the alphaVbeta3 structure reveals that they are buried. This observation suggests that the highly bent alphaVbeta3 structure represents the inactive conformation and that release of contacts with I-EGF modules 2 and 3 triggers a switchblade-like opening motion extending the integrin into its active conformation.  相似文献   

4.
Occupancy of integrin receptors induces conformational changes in the receptor, resulting in exposure of novel interactive sites termed ligand-induced binding sites (LIBS). We report here that Fab fragments of certain antibodies against LIBS on integrin alpha IIb beta 3 (platelet glycoprotein IIb-IIIa) block platelet aggregation. Thus, certain LIBS or the regions surrounding them may participate in events required for platelet aggregation. In addition, certain anti-alpha IIb beta 3 LIBS Fab fragments stimulated platelet aggregation. This was due to induction of fg binding to alpha IIb beta 3, apparently by shifting a conformational equilibrium between a "resting" and an "activated" state of alpha IIb beta 3. Some of the activating anti-LIBS Fab fragments also induced high affinity fibronectin binding to alpha IIb beta 3, whereas others did not. Thus, changes in the conformation of this integrin modulate both the specificity and affinity of ligand recognition.  相似文献   

5.
Expression in OV10 cells of either wild-type CD47 or its extracellular IgV domain linked to a glycosylphosphatidylinositol anchor-(IgV-GPI) enhanced ligand-induced alpha(v)beta(3) activation as detected by the binding of LIBS1 and LIBS6 mAbs. The amplitude of LIBS binding was greater with both CD47 and IgV-GPI expression, indicating an increase in the population of "activable" integrin molecules. Expression of either CD47 species also increased alpha(v)beta(3)-mediated adhesion to vitronectin, and to surfaces coated with the anti-beta(3) antibody AP3, because of enhanced clustering of alpha(v)beta(3) as confirmed by chemical cross-linking. Cholesterol depletion with methyl-beta-cyclodextrin did not prevent the increase in anti-LIBS binding, but reduced cell adhesion to vitronectin and AP3. However, cells expressing CD47 were partially insulated against this disruption, and IgV-GPI was even more effective. Both CD47 and IgV-GPI were found in cholesterol-rich rafts prepared in the absence of detergent, but only CD47 could recruit alpha(v)beta(3) and its associated signaling molecules to these domains. Thus CD47-alpha(v)beta(3) complexes in cholesterol-rich raft domains appear to engage in G(i)-dependent signaling whereas CD47-alpha(v)beta(3) interactions that lead to integrin clustering are also detergent resistant, but are insensitive to cholesterol depletion and do not require the transmembrane region of CD47.  相似文献   

6.
Integrins are adhesion molecules that convey signals both to and from the cytoplasm across the plasma membrane. In resting cells, integrins in a low affinity state can be activated by 'inside-out signaling', in which signals affecting integrin heterodimer cytoplasmic domains cause a conformational change in the integrin ligand-binding headpiece connected to the membrane by two long, approximately 16 nm stalks. Here we demonstrate a mechanism for conveying a conformational change over the long distance from the plasma membrane to the headpiece. We prepared soluble, alpha5beta1 integrin heterodimer extracellular fragments in which interactions between alpha- and beta-subunit cytoplasmic domains were replaced with an artificial clasp. Release of this C-terminal clasp by specific protease cleavage resulted in an approximately 14 nm separation of the stalks coupled to increased binding to fibronectin. This activation did not require any associated molecules or clustering and was observed with physiological concentrations of divalent cations. These findings suggest that the overall mechanism for integrin inside-out activation involves the spatial separation of the cytoplasmic and/or transmembrane domains.  相似文献   

7.
Conformational change in the integrin extracellular domain is required for high affinity ligand binding and is also involved in post-ligand binding cellular signaling. Although there is evidence to the contrary, electron microscopic studies showing that ligand binding triggers alpha- and beta-subunit dissociation in the integrin headpiece have gained popularity and support the hypothesis that head separation activates integrins. To test directly the head separation hypothesis, we enforced head association by introducing disulfide bonds across the interface between the alpha-subunit beta-propeller domain and the beta-subunit I-like domain. Basal and activation-dependent ligand binding by alpha(IIb)beta(3) and alpha(V)beta(3) was unaffected. The covalent linkage prevented dissociation of alpha(IIb)beta(3) into its subunits on EDTA-treated cells. Whereas EDTA dissociated wild type alpha(IIb)beta(3) on the cell surface, a ligand-mimetic Arg-Gly-Asp peptide did not, as judged by binding of complex-specific antibodies. Finally, a high affinity ligand-mimetic compound stabilized noncovalent association between alpha(IIb) and beta(3) headpiece fragments in the presence of SDS, indicating that ligand binding actually stabilized subunit association at the head, as opposed to the suggested subunit separation. The mechanisms of conformational regulation of integrin function should therefore be considered in the context of the associated alphabeta headpiece.  相似文献   

8.
The membrane-distal headpiece of integrins has evolved to specifically bind large extracellular protein ligands, but the molecular architecture of the resulting complexes has not been determined. We used molecular electron microscopy to determine the three-dimensional structure of the ligand-binding headpiece of integrin alpha5beta1 complexed with fragments of its physiological ligand fibronectin. The density map for the unliganded alpha5beta1 headpiece shows a 'closed' conformation similar to that seen in the alphaVbeta3 crystal structure. By contrast, binding to fibronectin induces an 'open' conformation with a dramatic, approximately 80 degrees change in the angle of the hybrid domain of the beta subunit relative to its I-like domain. The fibronectin fragment binds to the interface between the beta-propeller and I-like domains in the integrin headpiece through the RGD-containing module 10, but direct contact of the synergy-region-containing module 9 to integrin is not evident. This finding is corroborated by kinetic analysis of real-time binding data, which shows that the synergy site greatly enhances k(on) but has little effect on the stability or k(off) of the complex.  相似文献   

9.
Integrins are postulated to undergo structural rearrangement from a low affinity bent conformer to a high affinity extended conformer upon activation. However, some reports have shown that a bent conformer is capable of binding a ligand, whereas another report has shown that integrin extension does not absolutely lead to activation. To clarify whether integrin affinity is indeed regulated by the so-called switchblade-like movement, we have engineered a series of mutant αIIbβ3 integrins that are constrained specifically in either a bent or an extended conformation. These mutant αIIbβ3 integrins were expressed in mammalian cells, and fibrinogen binding to these cells was examined. The bent integrins were created through the introduction of artificial disulfide bridges in the β-head/β-tail interface. Cells expressing bent integrins all failed to bind fibrinogen unless pretreated with DTT to disrupt the disulfide bridges. The extended integrins were created by introducing N-glycosylation sites in amino acid residues located close to the α-genu, where the integrin legs fold backward. Among these mutants, activation was maximized in one integrin with an N-glycosylation site located behind the α-genu. This extension-induced activation was completely blocked when the swing-out of the hybrid domain was prevented. These results suggest that the bent and extended conformers represent low affinity and high affinity conformers, respectively, and that extension-induced activation depends on the swing-out of the hybrid domain. Taken together, these results are consistent with the current hypothesis that integrin affinity is regulated by the switchblade-like movement of the integrin legs.  相似文献   

10.
Although integrin alpha subunit I domains exist in multiple conformations, it is controversial whether integrin beta subunit I-like domains undergo structurally analogous movements of the alpha7-helix that are linked to affinity for ligand. Disulfide bonds were introduced into the beta(3) integrin I-like domain to lock its beta6-alpha7 loop and alpha7-helix in two distinct conformations. Soluble ligand binding, ligand mimetic mAb binding and cell adhesion studies showed that disulfide-bonded receptor alpha(IIb)beta(3)(T329C/A347C) was locked in a low affinity state, and dithiothreitol treatment restored the capability of being activated to high affinity binding; by contrast, disulfide-bonded alpha(IIb)beta(3)(V332C/M335C) was locked in a high affinity state. The results suggest that activation of the beta subunit I-like domain is analogous to that of the alpha subunit I domain, i.e. that axial movement in the C-terminal direction of the alpha7-helix is linked to rearrangement of the I-like domain metal ion-dependent adhesion site into a high affinity conformation.  相似文献   

11.
We recently reported a computational method (CHAMP) for designing sequence-specific peptides that bind to the membrane-embedded portions of transmembrane proteins. We successfully applied this method to design membrane-spanning peptides targeting the transmembrane domains of the alpha IIb subunit of integrin alpha IIbbeta 3. Previously, we demonstrated that these CHAMP peptides bind specifically with reasonable affinity to isolated transmembrane helices of the targeted transmembrane region. These peptides also induced integrin alpha IIbbeta 3 activation due to disruption of the helix-helix interactions between the transmembrane domains of the alpha IIb and beta 3 subunits. In this paper, we show the direct interaction of the designed anti-alpha IIb CHAMP peptide with isolated full-length integrin alpha IIbbeta 3 in detergent micelles. Further, the behavior of the designed peptides in phospholipid bilayers is essentially identical to their behavior in detergent micelles. In particular, the peptides assume a membrane-spanning alpha-helical conformation that does not disrupt bilayer integrity. The activity and selectivity of the CHAMP peptides were further explored in platelets, comfirming that anti-alpha IIb activates wild-type alpha IIbbeta 3 in whole cells as a result of its disruption of the protein-protein interactions between the alpha and beta subunits in the transmembrane regions. These results demonstrate that CHAMP is a successful chemical biology approach that can provide specific tools for probing the transmembrane domains of proteins.  相似文献   

12.
Bovine lactadherin binds to the alpha(v)beta(3) and alpha(v)beta(5) integrins in an RGD-dependent manner and also to anionic phospholipids. During the affinity purification of lactadherin binding receptors, a 35-kDa protein persistently coeluted with the alpha(v)beta(5) integrin receptor. Subsequently, peptide mapping, amino acid sequencing, and mass spectrometry analysis identified this protein as bovine annexin-V. Annexin-V accompanied the integrin receptor eluted with either RGD peptide or with EDTA suggesting that annexin-V bound specifically to the alpha(v)beta(5) integrin. To further investigate this putative interaction of annexin-V with the alpha(v)beta(5) integrin receptor, human annexin-V and intracellular domains of the human alpha(v)beta(5) integrin subunits were used in ligand blotting assays. Radiolabeled annexin-V showed weak binding to the intracellular part of beta(5) integrin subunit. However, by adding the aminophospholipid, phosphatidyl serine, the interaction with the beta(5) cytoplasmic peptide was enhanced many fold. Furthermore, the interaction was shown to be independent of phosphorylation, as annexin-V bound to unphosphorylated beta(5) peptide at a similar level to the phosphorylated peptide. Since binding of annexin-V to the alpha(v) integrin subunit tail was not detected, annexin-V was shown to associate specifically with the beta(5) cytoplasmic tail. Together these findings suggest a novel link between annexins and the integrin receptor family.  相似文献   

13.
The leukocyte-restricted integrin alpha(L)beta(2) is required in immune processes such as leukocyte adhesion, migration, and immune synapse formation. Activation of alpha(L)beta(2) by conformational changes promotes alpha(L)beta(2) binding to its ligands, ICAMs. It was reported that different affinity states of alpha(L)beta(2) are required for binding ICAM-1 and ICAM-3. Recently, the bent, extended with a closed headpiece, and extended with open headpiece conformations of alpha(L)beta(2), was reported. To address the overall conformational requirements of alpha(L)beta(2) that allow selective binding of these ICAMs, we examined the adhesion properties of these alpha(L)beta(2) conformers. alpha(L)beta(2) with different conformations were generated by mutations, and verified by using a panel of reporter mAbs that detect alpha(L)beta(2) extension, hybrid domain movement, or I-like domain activation. We report a marked difference between extended alpha(L)beta(2) with closed and open headpieces in their adhesive properties to ICAM-1 and ICAM-3. Our data show that the extension of alpha(L)beta(2) alone is sufficient to mediate ICAM-1 adhesion. By contrast, an extended alpha(L)beta(2) with an open headpiece is required for ICAM-3 adhesion.  相似文献   

14.
Integrin cytoplasmic domains mediate inside-out signal transduction   总被引:35,自引:10,他引:25       下载免费PDF全文
《The Journal of cell biology》1994,124(6):1047-1059
We analyzed the binding of fibronectin to integrin alpha 5 beta 1 in various cells; in some cells fibronectin bound with low affinity (e.g., K562 cells) whereas in others (e.g., CHO), it bound with high affinity (Kd approximately 100 nM) in an energy-dependent manner. We constructed chimeras of the extracellular and transmembrane domains of alpha IIb beta 3 joined to the cytoplasmic domains of alpha 5 beta 1. The affinity state of these chimeras was assessed by binding of fibrinogen or the monoclonal antibody, PAC1. The cytoplasmic domains of alpha 5 beta 1 conferred an energy-dependent high affinity state on alpha IIb beta 3 in CHO but not K562 cells. Three additional alpha cytoplasmic domains (alpha 2, alpha 6A, alpha 6B) conferred PAC1 binding in CHO cells, while three others (alpha M, alpha L, alpha v) did not. In the high affinity alpha chimeras, cotransfection with a truncated (beta 3 delta 724) or mutated (beta 3(S752-->P)) beta 3 subunit abolished high affinity binding. Thus, both cytoplasmic domains are required for energy-dependent, cell type-specific affinity modulation. In addition, mutations that disrupted a highly conserved alpha subunit GFFKR motif, resulted in high affinity binding of ligands to alpha IIb beta 3. In contrast to the chimeras, the high affinity state of these mutants was independent of cellular metabolism, cell type, and the bulk of the beta subunit cytoplasmic domain. Thus, integrin cytoplasmic domains mediate inside-out signaling. Furthermore, the highly conserved GFFKR motif of the alpha subunit cytoplasmic domain maintains the default low affinity state.  相似文献   

15.
Neff S  Mason PW  Baxt B 《Journal of virology》2000,74(16):7298-7306
We have previously reported that Foot-and-mouth disease virus (FMDV), which is virulent for cattle and swine, can utilize the integrin alpha(v)beta(3) as a receptor on cultured cells. Since those studies were performed with the human integrin, we have molecularly cloned the bovine homolog of the integrin alpha(v)beta(3) and have compared the two receptors for utilization by FMDV. Both the alpha(v) and beta(3) subunits of the bovine integrin have high degrees of amino acid sequence similarity to their corresponding human subunits in the ectodomains (96%) and essentially identical transmembrane and cytoplasmic domains. Within the putative ligand-binding domains, the bovine and human alpha(v) subunits have a 98.8% amino acid sequence similarity while there is only a 93% similarity between the beta(3) subunits of these two species. COS cell cultures, which are not susceptible to FMDV infection, become susceptible if cotransfected with alpha(v) and beta(3) subunit cDNAs from a bovine or human source. Cultures cotransfected with the bovine alpha(v)beta(3) subunit cDNAs and infected with FMDV synthesize greater amounts of viral proteins than do infected cultures cotransfected with the human integrin subunits. Cells cotransfected with a bovine alpha(v) subunit and a human beta(3) subunit synthesize viral proteins at levels equivalent to those in cells expressing both human subunits. However, cells cotransfected with the human alpha(v) and the bovine beta(3) subunits synthesize amounts of viral proteins equivalent to those in cells expressing both bovine subunits, indicating that the bovine beta(3) subunit is responsible for the increased effectiveness of this receptor. By engineering chimeric bovine-human beta(3) subunits, we have shown that this increase in receptor efficiency is due to sequences encoding the C-terminal one-third of the subunit ectodomain, which contains a highly structured cysteine-rich repeat region. We postulate that amino acid sequence differences within this region may be responsible for structural differences between the human and bovine beta(3) subunit, leading to more efficient utilization of the bovine receptor by this bovine pathogen.  相似文献   

16.
We show that the three conformational states of integrin α5β1 have discrete free energies and define activation by measuring intrinsic affinities for ligand of each state and the equilibria linking them. The 5,000‐fold higher affinity of the extended‐open state than the bent‐closed and extended‐closed states demonstrates profound regulation of affinity. Free energy requirements for activation are defined with protein fragments and intact α5β1. On the surface of K562 cells, α5β1 is 99.8% bent‐closed. Stabilization of the bent conformation by integrin transmembrane and cytoplasmic domains must be overcome by cellular energy input to stabilize extension. Following extension, headpiece opening is energetically favored. N‐glycans and leg domains in each subunit that connect the ligand‐binding head to the membrane repel or crowd one another and regulate conformational equilibria in favor of headpiece opening. The results suggest new principles for regulating signaling in the large class of receptors built from extracellular domains in tandem with single‐span transmembrane domains.  相似文献   

17.
L-plastin (LPL) is a leukocyte actin binding protein previously implicated in the activation of the integrin alpha(M)beta(2) on polymorphonuclear neutrophils. To determine the role for LPL in integrin activation, K562 cell adhesion to vitronectin via alpha(v)beta(3), a well-studied model for activable integrins, was examined. Cell permeant versions of peptides based on the N-terminal sequence of LPL and the LPL headpiece domain both activated alpha(v)beta(3)-mediated adhesion. In contrast to adhesion induced by treatment with phorbol 12-myristate 13-acetate (PMA), LPL peptide-activated adhesion was independent of integrin beta(3) cytoplasmic domain tyrosines and was not inhibited by cytochalasin D. Also in contrast to PMA, LPL peptides synergized with RGD ligand or Mn(2+) for generation of a conformational change in alpha(v)beta(3) associated with the high affinity state of the integrin, as determined by binding of a ligand-induced binding site antibody. Although LPL and ligand showed synergy for ligand-induced binding site expression when actin depolymerization was inhibited by jasplakinolide, LPL peptide-induced adhesion was inhibited. Thus, both actin depolymerization and ligand-induced integrin conformational change are required for LPL peptide-induced adhesion. We hypothesize that the critical steps of increased integrin diffusion and affinity enhancement may be linked via modulation of the function of the actin binding protein L-plastin.  相似文献   

18.
The integrin alpha(v)beta(3) is the major receptor mediating the attachment of osteoclasts to the extracellular matrix in bone and plays a critical role in bone resorption and bone remodeling. Most of the ligands interacting with the alpha(v)beta(3) receptor contain an Arg-Gly-Asp (RGD) motif. Recently, we have identified two small RGD peptides, containing a benzophenone moiety at either the carboxyl or amino terminus, that photo-cross-linked within the beta(3)[99-118] [Bitan, G., et al. (1999) Biochemistry 38, 3414-3420] or the beta(3)[167-171] [Bitan, G., et al. (2000) Biochemistry 39, 11014-11023] sequence, respectively, of the alpha(v)beta(3) receptor in a selective fashion. Here, we report the synthesis of a photoreactive analogue of echistatin (a 49-amino acid peptide), a potent RGD-containing antagonist of the alpha(v)beta(3) receptor both in vitro and in vivo. This bioactive analogue is substituted at position 45 with a p-benzoyl moiety (pBz(2)), located within the flexible C-terminal domain and removed 20 amino acid residues from the R(24)GD(26) triad. This C-terminal domain was reported to contribute to receptor binding affinity by acting as an auxiliary binding site. The radiolabeled (125)I-[Arg(35),Lys(45)(N(epsilon)-pBz(2))]-echistatin photo-cross-links effectively to a site within the beta(3)[209-220] sequence. Residues in this domain have been reported to be part of the metal ion-dependent adhesion site (MIDAS). Receptor fragments overlapping this domain were reported to bind to fibrinogen and block fibrinogen binding to alpha(IIb)beta(3), the platelet integrin receptor. Taken together, position 45 in echistatin, located within an auxiliary binding site in echistatin, cross-links to a site distinct from the two previously reported sites, beta(3)[99-118] and beta(3)[167-171], which cross-link to photophores flanking the RGD triad. These cross-linking data support the hypothesis that the ligand-bound conformation of the integrin beta(3) subunit differs from the known conformation of I domains.  相似文献   

19.
Conformational communication across the plasma membrane between the extracellular and intracellular domains of integrins is beginning to be defined by structural work on both domains. However, the role of the α and β subunit transmembrane domains and the nature of signal transmission through these domains have been elusive. Disulfide bond scanning of the exofacial portions of the integrin αIIβ and β3 transmembrane domains reveals a specific heterodimerization interface in the resting receptor. This interface is lost rather than rearranged upon activation of the receptor by cytoplasmic mutations of the α subunit that mimic physiologic inside-out activation, demonstrating a link between activation of the extracellular domain and lateral separation of transmembrane helices. Introduction of disulfide bridges to prevent or reverse separation abolishes the activating effect of cytoplasmic mutations, confirming transmembrane domain separation but not hinging or piston-like motions as the mechanism of transmembrane signaling by integrins.  相似文献   

20.
Integrin inside-out signaling and the immunological synapse   总被引:1,自引:0,他引:1  
Integrins dynamically equilibrate between three conformational states on cell surfaces. A bent conformation has a closed headpiece. Two extended conformations contain either a closed or an open headpiece. Headpiece opening involves hybrid domain swing-out and a 70 ? separation at the integrin knees, which is conveyed by allostery from the hybrid-proximal end of the βI domain to a 3 ? rearrangement of the ligand-binding site at the opposite end of the βI domain. Both bent-closed and extended-closed integrins have low affinity, whereas extended-open integrin affinity is 10(3) to 10(4) higher. Integrin-mediated adhesion requires the extended-open conformation, which in physiological contexts is stabilized by post-ligand binding events. Integrins thus discriminate between substrate-bound and soluble ligands. Analysis of LFA-1-ICAM-1 interactions in the immunological synapse suggests that bond lifetimes are on the order of seconds, which is consistent with high affinity interactions subjected to cytoskeletal forces that increase the dissociation rate. LFA-1 βI domain antagonists abrogate function in the immunological synapse, further supporting a critical role for high affinity LFA-1.  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