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.
Syndecans function as co-receptors for integrins on different matrixes. Recently, syndecan-1 has been shown to be important for α2β1 integrin-mediated adhesion to collagen in tumor cells by regulating cell adhesion and migration on two-dimensional collagen. However, the function of syndecans in supporting α2β1 integrin interactions with three-dimensional (3D) collagen is less well studied. Using loss-of-function and overexpression experiments we show that in 3D collagen syndecan-4 supports α2β1-mediated collagen matrix contraction. Cell invasion through type I collagen containing 3D extracellular matrix (ECM) is driven by α2β1 integrin and membrane type-1 matrix metalloproteinase (MT1-MMP). Here we show that mutational activation of K-ras correlates with increased expression of α2β1 integrin, MT1-MMP, syndecan-1, and syndecan-4. While K-ras-induced α2β1 integrin and MT1-MMP are positive regulators of invasion, silencing and overexpression of syndecans demonstrate that these proteins inhibit cell invasion into collagen. Taken together, these data demonstrate the existence of a complex interplay between integrin α2β1, MT1-MMP, and syndecans in the invasion of K-ras mutant cells in 3D collagen that may represent a mechanism by which tumor cells become more invasive and metastatic.  相似文献   

2.
One of the hallmarks of cells undergoing mitotic division is their rounded morphology and reduced adhesion to the substratum. We have studied and compared the attachment of interphase and mitotic cells to substrata coated with fibronectin and vitronectin. We have found that adhesion of mitotic cells, as compared to interphase cells, is significantly reduced to fibronectin, but is higher to vitronectin. These results correlate well with the expression of α5β1 and αVβ3 integrins, the respective receptors for fibronectin and vitronectin, on the cell surface. Mitotic cells show higher levels of αVβ3 and very low levels of α5β1 proteins on the cell surface as compared to interphase cells. This difference in the levels of these integrins also reflects in the total amounts of fibronectin and vitronectin present on the cell surface of these cells. We have further shown, by flow cytometry, that binding of vitronectin, or the synthetic peptide-GRGDSP-, causes an increase in the intracellular levels of Ca2− in mitotic cells, but no change is seen in the interphase cells. Binding of fibronectin to either of these cells fails to elicit any response. One interesting feature of our results is that the levels of total, i.e., cytoplasmic plus membrane bound, α5β1 and αVβ3 integrins of mitotic and interphase cells remain the same, thus implying an alteration in the distribution of integrin chains between the plasma membrane and the cytoplasm during the conversion of interphase cells into the mitotic phase. © 1996 Wiley-Liss, Inc.  相似文献   

3.
Extracellular divalent cations are important regulators of integrin ligand binding activity. In this study we evaluated how divalent cations affect the organization of integrins into focal adhesion sites. Integrins αvβ3 and αvβ5 were compared because they share a high degree of structural homology and because both integrins mediate cell adhesion to vitronectin. On MG-63 osteosarcoma cells, we found that both the extent and pattern of integrin organization was regulated by the type of extracellular divalent ion. Integrin αvβ3 organized in focal contacts when Mn2+ or Mg2+ was present, but not in Ca2+. In contrast, αvβ5 organized in focal contacts only when Ca2+ or Mg2+ was present. Integrin αvβ5 clustered in a centrally located punctate field on the ventral surface of the cell in the presence of Mn2+. These observations reveal a previously unappreciated role for divalent ions in regulating the organization of integrins into focal adhesion sites. © 1996 Wiley-Liss, Inc.  相似文献   

4.
Cyclic strain has been shown to modulate endothelial cell (EC) morphology, proliferation, and function. We have recently reported that the focal adhesion proteins focal adhesion kinase (pp125FAK) and paxillin, are tyrosine phosphorylated in EC exposed to strain and these events regulate the morphological change and migration induced by cyclic strain. Integrins are also localized on focal adhesion sites and have been reported to induce tyrosine phosphorylation of pp125FAK under a variety of stimuli. To study the involvement of different integrins in signaling induced by cyclic strain, we first observed the redistribution of α and β integrins in EC subjected to 4 h cyclic strain. Human umbilical vein endothelial cells (HUVEC) seeded on either fibronectin or collagen surfaces were subjected to 10% average strain at a frequency 60 cycles/min. Confocal microscopy revealed that β1 integrin reorganized in a linear pattern parallel with the long axis of the elongated cells creating a fusion of focal adhesion plaques in EC plated on either fibronectin (a ligand for α5β1) or collagen (a ligand for α2β1) coated plates after 4 h exposure to cyclic strain. β3 integrin, which is a vitronectin receptor, did not redistribute in EC exposed to cyclic strain. Cyclic strain also led to a reorganization of α5 and α2 integrins in a linear pattern in HUVEC seeded on fibronectin or collagen, respectively. The expression of integrins α5, α2, and β1 did not change even after 24 h exposure to strain when assessed by immunoprecipitation of these integrins. Cyclic strain-induced tyrosine phosphorylation of pp125FAK occurred concomitant with the reorganization of β1 integrin. We concluded that α5β1 and α2β1 integrins play an important role in transducing mechanical stimuli into intracellular signals. J. Cell. Biochem. 64:505–513. © 1997 Wiley-Liss, Inc.  相似文献   

5.
Heparin/heparan sulfate interact with growth factors, chemokines, extracellular proteins, and receptors. Integrins are αβ heterodimers that serve as receptors for extracellular proteins, regulate cell behavior, and participate in extracellular matrix assembly. Heparin binds to RGD‐dependent integrins (αIIbβ3, α5β1, αvβ3, and αvβ5) and to RGD‐independent integrins (α4β1, αXβ2, and αMβ2), but their binding sites have not been located on integrins. We report the mapping of heparin binding sites on the ectodomain of αvβ3 integrin by molecular modeling. The surface of the ectodomain was scanned with small rigid probes mimicking the sulfated domains of heparan sulfate. Docking results were clustered into binding spots. The best results were selected for further docking simulations with heparin hexasaccharide. Six potential binding spots containing lysine and/or arginine residues were identified on the ectodomain of αvβ3 integrin. Heparin would mostly bind to the top of the genu domain, the Calf‐I domain of the α subunit, and the top of the β subunit of RGD‐dependent integrins. Three spots were close enough from each other on the integrin surface to form an extended binding site that could interact with heparin/heparan sulfate chains. Because heparin does not bind to the same integrin site as protein ligands, no steric hindrance prevents the formation of ternary complexes comprising the integrin, its protein ligand, and heparin/heparan sulfate. The basic amino acid residues predicted to interact with heparin are conserved in the sequences of RGD‐dependent but not of RGD‐independent integrins suggesting that heparin/heparan sulfate could bind to different sites on these two integrin subfamilies. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

6.
Cellular receptors for collagens belong to the family of β(1) integrins. In the epidermis, integrin α(2)β(1) is the only collagen-binding integrin present. Its expression is restricted to basal keratinocytes with uniform distribution on the cell surface of those cells. Although α(2)β(1) receptors localized at the basal surface interact with basement membrane proteins collagen IV and laminin 111 and 332, no interaction partners have been reported for these integrin molecules at the lateral and apical membranes of basal keratinocytes. Solid phase binding and surface plasmon resonance spectroscopy demonstrate that collagen XXIII, a member of the transmembrane collagens, directly interacts with integrin α(2)β(1) in an ion- and conformation-dependent manner. The two proteins co-localize on the surface of basal keratinocytes. Furthermore, collagen XXIII is sufficient to induce adhesion and spreading of keratinocytes, a process that is significantly reduced in the absence of functional integrin α(2)β(1).  相似文献   

7.
8.
There is a critical need for compounds that target cell surface integrin receptors for applications in cancer therapy and diagnosis. We used directed evolution to engineer the Ecballium elaterium trypsin inhibitor (EETI‐II), a knottin peptide from the squash family of protease inhibitors, as a new class of integrin‐binding agents. We generated yeast‐displayed libraries of EETI‐II by substituting its 6‐amino acid trypsin binding loop with 11‐amino acid loops containing the Arg‐Gly‐Asp integrin binding motif and randomized flanking residues. These libraries were screened in a high‐throughput manner by fluorescence‐activated cell sorting to identify mutants that bound to αvβ3 integrin. Select peptides were synthesized and were shown to compete for natural ligand binding to integrin receptors expressed on the surface of U87MG glioblastoma cells with half‐maximal inhibitory concentration values of 10–30 nM. Receptor specificity assays demonstrated that engineered knottin peptides bind to both αvβ3 and αvβ5 integrins with high affinity. Interestingly, we also discovered a peptide that binds with high affinity to αvβ3, αvβ5, and α5β1 integrins. This finding has important clinical implications because all three of these receptors can be coexpressed on tumors. In addition, we showed that engineered knottin peptides inhibit tumor cell adhesion to the extracellular matrix protein vitronectin, and in some cases fibronectin, depending on their integrin binding specificity. Collectively, these data validate EETI‐II as a scaffold for protein engineering, and highlight the development of unique integrin‐binding peptides with potential for translational applications in cancer. Proteins 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

9.
The collagen-binding integrins α1β1 and α2β1 have profoundly different functions, yet they are often co-expressed in epithelial cells. When both integrins are expressed in the same cell, it has been suggested that α1β1 negatively regulates integrin α2β1-dependent functions. In this study we utilized murine ureteric bud (UB) epithelial cells, which express no functionally detectable levels of endogenous integrins α1β1 and α2β1, to determine the mechanism whereby this regulation occurs. We demonstrate that UB cells expressing integrin α2β1, but not α1β1 adhere, migrate and proliferate on collagen I as well as form cellular cords in 3D collagen I gels. Substitution of the transmembrane domain of the integrin α2 subunit with that of α1 results in decreased cell adhesion, migration and cord formation. In contrast, substitution of the integrin α2 cytoplasmic tail with that of α1, decreases cell migration and cord formation, but increases proliferation. When integrin α1 and α2 subunits are co-expressed in UB cells, the α1 subunit negatively regulates integrin α2β1-dependent cord formation, adhesion and migration and this inhibition requires expression of both α1 and α2 tails. Thus, we provide evidence that the transmembrane and cytoplasmic domains of the α2 integrin subunit, as well as the α1 integrin subunit, regulate integrin α2β1 cell function.  相似文献   

10.
A severe re‐emergingzoonosis, leptospirosis, is caused by pathogenic spirochetes of the genus Leptospira. Several studies have identified leptospiral surface proteins with the ability to bind ECM and plasma components, which could mediate adhesion and invasion through the hosts. It has been shown that Mce of pathogenic Leptospira spp. is an RGD (Arg‐Gly‐Asp)‐motif‐dependent virulence factor, responsible for infection of cells and animals. In the present article, we decided to further study the repertoire of the Mce activities in leptospiral biological properties. We report that the recombinant Mce is a broad‐spectrum ECM‐binding protein, capable of interacting with laminin, cellular and plasma fibronectin and collagen IV. Dose­–r­esponse interaction was observed for all the components, fulfilling ligand­–receptor requirements. Mce is a PLG binding protein capable to recruit this component from NHS, generating PLA in the presence of PLG activator. Binding of Mce was also observed with the leukocyte cell receptors αLβ2 [(CD11a/CD18)‐LFA‐1] and αMβ2 [(CD11b/CD18)‐Mac‐1], suggesting the involvement of this protein in the host immune response. Indeed, virulent Leptospira L1‐130 was capable of binding both integrins, whereas culture‐attenuated M‐20 strain only bind to αMβ2 [(CD11b/CD18)‐Mac‐1]. To the best of our knowledge, this is the first work to describe that Mce surface protein could mediate the attachment of Leptospira interrogans to human cell receptors αLβ2(CD11a/CD18) and αMβ2(CD11b/CD18).  相似文献   

11.
Integrins are well characterized cell surface receptors for extracellular matrix proteins. Mapping integrin-binding sites within the fibrillar collagens identified GFOGER as a high affinity site recognized by α2β1, but with lower affinity for α1β1. Here, to identify specific ligands for α1β1, we examined binding of the recombinant human α1 I domain, the rat pheochromocytoma cell line (PC12), and the rat glioma Rugli cell line to our collagen Toolkit II and III peptides using solid-phase and real-time label-free adhesion assays. We observed Mg(2+)-dependent binding of the α1 I domain to the peptides in the following rank order: III-7 (GLOGEN), II-28 (GFOGER), II-7 and II-8 (GLOGER), II-18 (GAOGER), III-4 (GROGER). PC12 cells showed a similar profile. Using antibody blockade, we confirmed that binding of PC12 cells to peptide III-7 was mediated by integrin α1β1. We also identified a new α1β1-binding activity within peptide II-27. The sequence GVOGEA bound weakly to PC12 cells and strongly to activated Rugli cells or to an activated α1 I domain, but not to the α2 I domain or to C2C12 cells expressing α2β1 or α11β1. Thus, GVOGEA is specific for α1β1. Although recognized by both α2β1 and α11β1, GLOGEN is a better ligand for α1β1 compared with GFOGER. Finally, using biosensor assays, we show that although GLOGEN is able to compete for the α1 I domain from collagen IV (IC(50) ~3 μm), GFOGER is much less potent (IC(50) ~90 μm), as shown previously. These data confirm the selectivity of GFOGER for α2β1 and establish GLOGEN as a high affinity site for α1β1.  相似文献   

12.
Twelve of sixteen different cell types including fibroblasts and tumor cells were able to attach and spread on substrates of pepsin-solubilized or intact collagen VI, and on its triple helical domain. Attachment and spreading were independent of soluble mediator proteins (fibronectin, laminin) and collagen VI was distinct from collagens I, IV and V in the cells with which it interacted. Many of the same cells bound and spread on substrates prepared from unfolded α2(VI) and α3(VI) chains but not on the α1(VI) chain. The interactions with the chains were inhibited by low concentrations (10–100 μM) of synthetic RGDS and RGDT but not RGES peptides while the binding of cells to pepsin-solubilized collagen VI was more than 20-fold less sensitive to these peptides. The data incidate that cells have the ability to bind to collagen VI in a specific manner suggesting a similar function for collagen VI in situ.  相似文献   

13.
The maturation of connective tissue involves the organization of collagen fibres by resident fibroblasts. Fibroblast attachment to collagen has been demonstrated to involve cell surface receptors, integrins of the β1 family. Integrins are associated with cytoplasmic actin of microfilaments either directly or through focal adhesions. The major actin isoform of fibroblast microfilaments is β actin and to a lesser extent α smooth muscle (α SM) actin. Cultured human dermal fibroblasts derived from adult dermis, newborn foreskin or keloid scar were grown on either uncoated or collagen-coated surfaces. The expression and synthesis of both α2β1 integrin and α SM actin were followed by immunohistology and immunoprecipitation. Fibroblasts on uncoated surfaces expressed little α2β1 integrin on their surface, while 20 per cent of them demonstrated α SM actin within microfilaments. Fibroblasts grown on a collagen-coated surface minimally expressed α SM actin in microfilament structures and a majority of the cells were positive for α2β1 integrin on their membranes. Using [35S]-methionine incorporation and immunoprecipitation, it was shown that fibroblasts grown in uncoated dishes synthesized more α SM actin than fibroblasts grown on collagen-coated dishes. In contrast, fibroblasts grown on collagen coated dishes synthesized more α2β1 integrin compared to the same cells grown on uncoated dishes. Fibroblasts maintained on a type I collagen upregulate the expression and synthesis of α2β1 integrin, and downregulate the expression and synthesis of α SM actin. © 1998 John Wiley & Sons, Ltd.  相似文献   

14.
Malignant cell transformation is generally accompanied by changes in their interactions with environing matrix proteins in a way to facilitate their migration and generate invasion. Our results show the binding of rat colon adenocarcinoma PROb cells to fibronectin strongly reduced when compared to normal rat intestine epithelial cells. This decrease was not due to the level of α(s)β1 integrins expressed at the surface of the cell line. However, β1- and α(s)-associated subunits appeared to be structurally altered as shown by immunoprecipitation followed by electrophoresis. Pulse chase experiments using 35S methionine evidenced differences in the biosynthesis of β1- and α (s) associated integrins: normal epithelial IEC18 cells required 16 h for maximal biosynthesis of the completely mature β1 subunit, while PROb cells did it within 4-6 h. Studies using endoglycosidases O, H, D, and N glycanase confirmed that the molecular weight alterations were due to abnormal glycosylation and suggested that α(s)β1 integrins of PROb cells could bear both mature complex and immature high mannose types while IEC18 cells borne only mature complex type oligosaccharidic chains. Treatment of both cell types with castanospermine, an inhibitor of N-glycosylation, reduced the differences observed in their adhesion to the fibronectin without significantly affecting β1 receptors expression at the cell surface. These results strongly suggest a role of the glycosylation of β1 receptors in the adhesion of rat colon adenocarcinoma PROb cells to fibronectin substrata. © 1996 Wiley-Liss, Inc.  相似文献   

15.
Collagen is a potent adhesive substrate for cells, an event essentially mediated by the integrins alpha 1 beta 1 and alpha 2 beta 1. Collagen fibrils also bind to the integrin alpha 2 beta 1 and the platelet receptor glycoprotein VI to activate and aggregate platelets. The distinct triple helical recognition motifs for these receptors, GXOGER and (GPO)n, respectively, all contain hydroxyproline. Using unhydroxylated collagen I produced in transgenic plants, we investigated the role of hydroxyproline in the receptor-binding properties of collagen. We show that alpha 2 beta 1 but not alpha 1 beta 1 mediates cell adhesion to unhydroxylated collagen. Soluble recombinant alpha 1 beta 1 binding to unhydroxylated collagen is considerably reduced compared with bovine collagens, but binding can be restored by prolyl hydroxylation of recombinant collagen. We also show that platelets use alpha 2 beta 1 to adhere to the unhydroxylated recombinant molecules, but the adhesion is weaker than on fully hydroxylated collagen, and the unhydroxylated collagen fibrils fail to aggregate platelets. Prolyl hydroxylation is thus required for binding of collagen to platelet glycoprotein VI and to cells by alpha 1 beta 1. These observations give new insights into the molecular basis of collagen-receptor interactions and offer new selective applications for the recombinant unhydroxylated collagen I.  相似文献   

16.
Dermal fibroblasts are essential for the repair of cutaneous wounds. Fibroblasts presumably use cell surface receptors of the integrin family during migration into a wound from the adjacent uninjured tissue and for the subsequent matrix repairs. We have investigated the possible roles of platelet-derived growth factor and inflammatory cytokines in the regulation of integrin expression on wound fibroblasts using a porcine cutaneous wound model and cultured human cells. Tissue specimens collected from 4-day pig wounds were stained with antibodies specific for the α1 and α5 integrin subunits. Staining for α1 was markedly decreased on fibroblasts adjacent to the wound and in the granulation tissue, while staining for α5 was clearly enhanced in both locations. Normal adult human dermal fibroblasts in culture express the integrins α1β1, a collagen receptor, and α5β1, a fibronectin receptor. Quantitative flow cytometry was used to measure cell surface integrin expression after treatment with platelet-derived growth factor (PDGF)-AA, PDGF-AB, or PDGF-BB. Each isoform of PDGF produced a significant decrease in the level of α1 present on the cell surface and an increase in the level of α5. Furthermore, PDGF-BB produced a corresponding decrease in α1 mRNA and an increase in α5 mRNA. In contrast, treatment with three inflammatory cytokines, IL-1β, TNF-α, and IFN-γ, produced clear increases in the levels of α1 and α5 present on the cell surface. Our observations suggest that the differential effects of PDGF and inflammatory cytokines may be part of the mechanism regulating the expression of α1 and α5 integrins by dermal fibroblasts during wound repair. © 1996 Wiley-Liss, Inc.  相似文献   

17.
Rotaviruses utilize integrins during virus-cell interactions that lead to infection. Cell binding and infection by simian rotavirus SA11 were inhibited by antibodies (Abs) to the inserted (I) domain of the alpha2 integrin subunit. To determine directly which integrins or other proteins bind rotaviruses, cell surface proteins precipitated by rotaviruses were compared with those precipitated by anti-alpha2beta1 Abs. Two proteins precipitated by SA11 and rhesus rotavirus RRV from MA104 and Caco-2 cells migrated indistinguishably from alpha2beta1 integrin, and SA11 precipitated beta1 from alpha2beta1-transfected CHO cells. These viruses specifically precipitated two MA104 cell proteins only, but an additional 160- to 165-kDa protein was precipitated by SA11 from Caco-2 cells. The role of the alpha2 I domain in rotavirus binding, infection, and growth was examined using CHO cell lines expressing wild-type or mutated human alpha2 or alpha2beta1. Infectious SA11 and RRV, but not human rotavirus Wa, specifically bound CHO cell-expressed human alpha2beta1 and, to a lesser extent, human alpha2 combined with hamster beta1. Binding was inhibited by anti-alpha2 I domain monoclonal Abs (MAbs), but not by non-I domain MAbs to alpha2, and required the presence of the alpha2 I domain. Amino acid residues 151, 221, and 254 in the metal ion-dependent adhesion site of the alpha2 I domain that are necessary for type I collagen binding to alpha2beta1 were not essential for rotavirus binding. Rotavirus-alpha2beta1 binding led to increased virus infection and RRV growth. SA11 and RRV require the alpha2 I domain for binding to alpha2beta1, and their binding to this integrin is distinguishable from that of collagen.  相似文献   

18.
Previously identified high affinity integrin-binding motifs in collagens, GFOGER and GLOGER, are not present in type III collagen. Here, we first characterized the binding of recombinant I domains from integrins alpha(1) and alpha(2) (alpha(1)I and alpha(2)I) to fibrillar collagen types I-III and showed that each I domain bound to the three types of collagens with similar affinities. Using rotary shadowing followed by electron microscopy, we identified a high affinity binding region in human type III collagen recognized by alpha(1)I and alpha(2)I. Examination of the region revealed the presence of two sequences that contain the critical GER motif, GROGER and GAOGER. Collagen-like peptides containing these two motifs were synthesized, and their triple helical nature was confirmed by circular dichroism spectroscopy. Experiments show that the GROGER-containing peptide was able to bind both alpha(1)I and alpha(2)I with high affinity and effectively inhibit the binding of alpha(1)I and alpha(2)I to type III and I collagens, whereas the GAOGER-containing peptide was considerably less effective. Furthermore, the GROGER-containing peptide supported adhesion of human lung fibroblast cells when coated on a culture dish. Thus, we have identified a novel high affinity binding sequence for the collagen-binding integrin I domains.  相似文献   

19.
Collagen receptor integrins alpha 1 beta 1 and alpha 2 beta 1 can selectively recognize different collagen subtypes. Here we show that their alpha I domains can discriminate between laminin isoforms as well: alpha 1I and alpha 2I recognized laminin-111, -211 and -511, whereas their binding to laminin-411 was negligible. Residue Arg-218 in alpha1 was found to be instrumental in high-avidity binding. The gain-of-function mutation E318W makes the alpha 2I domain to adopt the "open" high-affinity conformation, while the wild-type alpha 2I domain favors the "closed" low-affinity conformation. The E318W mutation markedly increased alpha 2I domain binding to the laminins (-111, -211 and -511), leading us to propose that the activation state of the alpha 2 beta 1 integrin defines its role as a laminin receptor. However, neither wild-type nor alpha 2IE318W domain could bind to laminin-411. alpha 2IE318W also bound tighter to all collagens than alpha 2I wild-type, but it showed reduced ability to discriminate between collagens I, IV and IX. The corresponding mutation, E317A, in the alpha 1I domain transformed the domain into a high-avidity binder of collagens I and IV. Thus, our results indicate that conformational activation of integrin alpha 1I and alpha 2I domains leads to high-avidity binding to otherwise disfavored collagen subtypes.  相似文献   

20.
Osteopontin (OPN) is highly expressed by macrophages and plays a key role in the pathology of several chronic inflammatory diseases including atherosclerosis and the foreign body reaction. However, the molecular mechanism behind OPN regulation of macrophage functions is not well understood. OPN is a secreted molecule and interacts with several integrins via two domains: the RGD sequence binding to αv‐containing integrins, and the SLAYGLR sequence binding to α4β1, α4β7, and α9β1 integrins. Here we determined the role of OPN in macrophage survival, chemotaxis, and activation state. For survival studies, OPN treated‐bone marrow derived macrophages (BMDMs) were challenged with growth factor withdrawal and neutralizing integrin antibodies. We found that survival in BMDMs is mediated primarily through the α4 integrin. In chemotaxis studies, we observed that migration to OPN was blocked by neutralizing α4 and α9 integrin antibodies. Further, OPN did not affect macrophage activation as measured by IL‐12 production. Finally, the relative contributions of the RGD and the SLAYGLR functional domains of OPN to leukocyte recruitment were evaluated in an in vivo model. We generated chimeric mice expressing mutated forms of OPN in myeloid‐derived leukocytes, and found that the SLAYGLR functional domain of OPN, but not the RGD, mediates macrophage accumulation in response to thioglycollate‐elicited peritonitis. Collectively, these data indicate that α4 and α9 integrins interacting with OPN via the SLAYGLR domain play a key role in macrophage biology by regulating migration, survival, and accumulation. J. Cell. Biochem. 114: 1194–1202, 2013. © 2012 Wiley Periodicals, Inc.  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