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 750 毫秒
1.
Rap1GAP expression is decreased in human tumors. The significance of its downregulation is unknown. We show that Rap1GAP expression is decreased in primary colorectal carcinomas. To elucidate the advantages conferred on tumor cells by loss of Rap1GAP, Rap1GAP expression was silenced in human colon carcinoma cells. Suppressing Rap1GAP induced profound alterations in cell adhesion. Rap1GAP-depleted cells exhibited defects in cell/cell adhesion that included an aberrant distribution of adherens junction proteins. Depletion of Rap1GAP enhanced adhesion and spreading on collagen. Silencing of Rap expression normalized spreading and restored E-cadherin, β-catenin, and p120-catenin to cell/cell contacts, indicating that unrestrained Rap activity underlies the alterations in cell adhesion. The defects in adherens junction protein distribution required integrin signaling as E-cadherin and p120-catenin were restored at cell/cell contacts when cells were plated on poly-l-lysine. Unexpectedly, Src activity was increased in Rap1GAP-depleted cells. Inhibition of Src impaired spreading and restored E-cadherin at cell/cell contacts. These findings provide the first evidence that Rap1GAP contributes to cell/cell adhesion and highlight a role for Rap1GAP in regulating cell/matrix and cell/cell adhesion. The frequent downregulation of Rap1GAP in epithelial tumors where alterations in cell/cell and cell/matrix adhesion are early steps in tumor dissemination supports a role for Rap1GAP depletion in tumor progression.Mammalian Rap proteins Rap1a/b and Rap2a/b/c are members of the Ras superfamily of small GTPases. Rap proteins are active when bound to GTP and inactive when bound to GDP. Cellular Rap activity is regulated by the concerted action of guanine nucleotide exchange factors that activate Rap (RapGEFs) and Rap-specific GTPase-activating proteins (RapGAPs) that inactivate Rap (reviewed in reference 10). The Rap1GAP family is composed of several members, including Rap1GAP, Rap1GAPII, Spa-1/SIPA1, and E6TP1/SIPA1L1. Several lines of evidence suggest that RapGAPs function as tumor and/or invasion suppressors. Downregulation of E6TP1 by human papillomavirus protein E6 contributes to cervical cancer (20, 21), and Spa-1 deficiency in mice induces a spectrum of myelodysplastic disorders similar to chronic myelogenous leukemia (26). The SPA1 gene was identified as a candidate for the metastasis efficiency modifier locus in mice (38). Although the relevance of this observation to humans is not yet clear, single-nucleotide polymorphisms in the SPA1 gene in human breast tumors have been associated with lymph node involvement and poor survival (15). Intriguingly, Spa-1 interacts with Brd4 (18) and Rrp-1b (13), the protein products of genes associated with patterns of extracellular matrix protein gene expression characteristic of metastatic tumors (14).The RAP1GAP gene maps to 1p35-36, a chromosomal region subject to copy number alterations in human tumors (36, 49). Rap1GAP protein levels are decreased in pancreatic adenocarcinomas (53), papillary thyroid carcinomas (37, 47, 57), and melanomas (56). Rap1GAP downregulation has been shown to arise as a consequence of proteasomal degradation (46), loss of heterozygosity (37, 53), and promoter methylation (56, 57). Mutations of unknown significance in RAP1GAP have been identified in breast cancer (42). Although downregulation of Rap1GAP is frequent in human tumors, the functional significance of decreased Rap1GAP expression is unknown. Up to now, studies assessing the role of Rap1GAP in tumor cells have relied exclusively on overexpression experiments. Overexpression of Rap1GAP in oropharyngeal squamous cell (54) and pancreatic (53) carcinoma lines impaired tumor formation in mouse xenograft models. In vitro, overexpression of Rap1GAP impaired tumor cell proliferation (34, 47, 53, 54, 56) and enhanced apoptosis (34, 53, 56). In some instances, overexpression of Rap1GAP inhibited tumor cell migration and invasion (3, 47, 53, 56), while in others, it enhanced invasion (34). While these studies provide insight into cellular processes that can be deregulated by overexpression, they do not assess the significance of depletion of endogenous Rap1GAP in human tumors.Colorectal cancer (CRC) is one of the leading causes of cancer deaths worldwide. The majority of CRC deaths arise as a consequence of distant metastases, most frequently to the liver. While the genetic basis of CRC is well understood (19, 48), less is known about the events that trigger the transition to metastatic disease. We report that Rap1GAP is highly expressed in normal colonic epithelium and that its expression is profoundly decreased in primary colorectal carcinomas. As one strategy to assess the significance of Rap1GAP depletion, the expression of Rap1GAP was silenced in human colon carcinoma cells. Silencing of Rap1GAP induced marked increases in Rap1 and Rap2 activity, the first evidence that Rap1GAP is an essential negative regulator of Rap GTPases in colon cancer. Rap1 regulates inside-out signaling through integrins (reviewed in references 8, 9, and 11) and is a target of outside-in signaling via cadherins (reviewed in reference 30). Downregulation of Rap1GAP induced profound alterations in cell/matrix and cell/cell adhesion. Suppressing Rap1GAP expression enhanced adhesion and spreading on collagen. Unexpectedly, based on the role of Rap1 in promoting cell/cell adhesion, silencing of Rap1GAP impaired cell/cell adhesion. These findings demonstrate a requirement for regulated Rap activity in the maintenance of epithelial cell structure and demonstrate a heretofore unappreciated role for Rap1GAP in the regulation of cell/cell adhesion. As the dissemination of tumor cells requires the weakening of cell/cell adhesion and an enhanced ability to adhere to collagen-rich interstitial matrices, our studies identify a potential mechanism through which loss of Rap1GAP contributes to tumor progression.  相似文献   

2.
Epac1 is a guanine nucleotide exchange factor (GEF) for the small G protein Rap and is directly activated by cyclic AMP (cAMP). Upon cAMP binding, Epac1 undergoes a conformational change that allows the interaction of its GEF domain with Rap, resulting in Rap activation and subsequent downstream effects, including integrin-mediated cell adhesion and cell-cell junction formation. Here, we report that cAMP also induces the translocation of Epac1 toward the plasma membrane. Combining high-resolution confocal fluorescence microscopy with total internal reflection fluorescence and fluorescent resonance energy transfer assays, we observed that Epac1 translocation is a rapid and reversible process. This dynamic redistribution of Epac1 requires both the cAMP-induced conformational change as well as the DEP domain. In line with its translocation, Epac1 activation induces Rap activation predominantly at the plasma membrane. We further show that the translocation of Epac1 enhances its ability to induce Rap-mediated cell adhesion. Thus, the regulation of Epac1-Rap signaling by cAMP includes both the release of Epac1 from autoinhibition and its recruitment to the plasma membrane.Cyclic AMP (cAMP) is an important second messenger that mediates many cellular hormone responses. It has become more and more appreciated that, along with the cAMP effector protein kinase A (PKA), Epac proteins also play pivotal roles in many cAMP-controlled processes, including insulin secretion (23, 39), cell adhesion (9, 17, 25, 49, 60), neurotransmitter release (22, 53, 63), heart function (13, 35, 54), and circadian rhythm (38). Epac1 and Epac2 are cAMP-dependent guanine nucleotide exchange factors (GEFs) for the small G proteins Rap1 and Rap2 (12, 24). They contain a regulatory region with one (Epac1) or two (Epac2) cAMP-binding domains, a Dishevelled, Egl-10, Pleckstrin (DEP) domain, and a catalytic region for GEF activity (11). The binding of cAMP is a prerequisite for catalytic activity in vitro and in vivo (11). Recently, the structures of both the inactive and active conformations of Epac2 were solved (51, 52). This revealed that in the inactive conformation, the regulatory region occludes the Rap binding site, which is relieved by a conformational change induced by cAMP binding.Like all G proteins of the Ras superfamily, Rap cycles between an inactive GDP-bound and active GTP-bound state in an equilibrium that is tightly regulated by specific GEFs and GTPase-activating proteins (GAPs). The GEF-induced dissociation of GDP results in the binding of the cellularly abundant GTP, whereas GAPs enhance the intrinsic GTPase activity of the G protein, thereby inducing the inactive GDP-bound state. Besides Epac, several other GEFs for Rap have been identified, including C3G, PDZ-GEF, and RasGRP, and these act downstream of different signaling pathways (7). Since Rap localizes to several membrane compartments, including the Golgi network, vesicular membranes, and the plasma membrane (PM) (2-4, 37, 42, 48), the spatial regulation of its activity is expected to be established by the differential distributions of its upstream GEFs, each activating distinct pools of Rap on specific intracellular locations.Similarly to Rap, Epac1 also is observed at many locations in the cell, including the cytosol, the nucleus, the nuclear envelope, endomembranes, and the PM (5, 11, 14, 21, 29, 47). These various locations may reflect the many different functions assigned to Epac1, such as the regulation of cell adhesion, cell junction formation, secretion, the regulation of DNA-dependent protein kinase by nuclear Epac1, and the regulation of the Na+/H+ exchanger NHE3 at the brush borders of kidney epithelium (19, 21, 26). Apparently, specific anchors are responsible for this spatial regulation of Epac1. Indeed, Epac1 was found to associate with phosphodiesterase 4 (PDE4) in a complex with mAKAP in cardiomyocytes (13), with MAP-LC bound to microtubules (62), and with Ezrin at the brush borders of polarized cells (M. Gloerich, J. Zhao, and J. L. Bos, unpublished data).In this study, we report the unexpected observation that, in addition to the temporal control of Epac1 activity, cAMP also induces the translocation of Epac1 toward the plasma membrane. Using confocal fluorescence microscopy, total internal reflection fluorescence (TIRF) microscopy, and fluorescence resonance energy transfer (FRET)-based assays for high spatial and temporal resolution, we observed that the translocation of Epac1 is immediate and that Epac1 approaches the PM to within ∼7 nm. In line with this, Epac1-induced Rap activation was registered predominantly on this compartment. Epac1 translocation results directly from the cAMP-induced conformational change and depends on the integrity of its DEP domain. We further show that Epac1 translocation is a prerequisite for cAMP-induced Rap activation at the PM and enhances Rap-mediated cell adhesion. Thus, cAMP exerts dual regulation on Epac1 for the activation of Rap, controlling both its GEF activity and targeting to the PM.  相似文献   

3.
Epac1 is a guanine nucleotide exchange factor for the small G protein Rap and is involved in membrane-localized processes such as integrin-mediated cell adhesion and cell-cell junction formation. Cyclic AMP (cAMP) directly activates Epac1 by release of autoinhibition and in addition induces its translocation to the plasma membrane. Here, we show an additional mechanism of Epac1 recruitment, mediated by activated ezrin-radixin-moesin (ERM) proteins. Epac1 directly binds with its N-terminal 49 amino acids to ERM proteins in their open conformation. Receptor-induced activation of ERM proteins results in increased binding of Epac1 and consequently the clustered localization of Epac1 at the plasma membrane. Deletion of the N terminus of Epac1, as well as disruption of the Epac1-ERM interaction by an interfering radixin mutant or small interfering RNA (siRNA)-mediated depletion of the ERM proteins, impairs Epac1-mediated cell adhesion. We conclude that ERM proteins are involved in the spatial regulation of Epac1 and cooperate with cAMP- and Rap-mediated signaling to regulate adhesion to the extracellular matrix.Cyclic AMP (cAMP) is a second messenger that relays a wide range of hormone responses. The discovery of Epac as a direct effector of cAMP (15, 29) has triggered the elucidation of many cAMP-regulated processes that could not be explained by the previously known effectors protein kinase A (PKA) and cyclic nucleotide-regulated ion channels (21). Both Epac family members, Epac1 and Epac2, act as guanine nucleotide exchange factors (GEFs) for the small G proteins Rap1 and Rap2. Thereby, Epac functions in processes such as exocytosis (28, 48, 59), cell-cell junction formation (13, 20, 30, 53, 64), and cell-extracellular matrix (ECM) adhesion (55). Adhesion to the ECM induced by Epac1 and Rap is mediated by actin-linked integrin molecules and is implicated in diverse biological processes such as homing of endothelial progenitor cells to ischemic tissue (9), remodeling of the vasculature (10, 36), and transendothelial migration of leukocytes (37, 60).Epac1 and Epac2 are multidomain proteins containing a C-terminal catalytic region, which consists of a CDC25 homology domain responsible for GEF activity, a Ras exchange motif (REM), which stabilizes the CDC25 homology domain, and a Ras association (RA) domain. In the autoinhibited state, the catalytic site is sterically covered by the N-terminal regulatory region, which harbors a DEP (Dishevelled, Egl-10, and pleckstrin) domain and one or two cyclic nucleotide-binding domains in Epac1 and Epac2, respectively. As demonstrated by the crystal structures of both active and inactive Epac2, autoinhibition is released by a conformational change induced by the binding of cAMP (56, 57).After its production at the plasma membrane (PM) by adenylate cylases, cAMP becomes compartmentalized due to local degradation by spatially restricted phosphodiesterases (1). Further compartmentalization of cAMP signaling is established by the confined targeting of the cAMP effector proteins. Numerous adaptor proteins that target PKA to distinct subcellular locations and mediate the assembly of large signaling complexes have been identified (3). Similarly, cAMP-Epac signaling appears to be spatially regulated by diverse anchoring mechanisms, which may reflect the many different functions assigned to Epac. For instance, the DNA damage-responsive kinase DNA-PK is regulated by nuclear Epac1 (26), whereas membrane recruitment by activated Ras is essential for the role of Epac2 in neurite outgrowth (34, 35). Recently, we reported that Epac1 translocates to the PM upon the binding of cAMP and that this translocation contributes to Rap-mediated cell-ECM adhesion (51). Although the anchor at the PM remains elusive, it has become clear that the cAMP-dependent translocation of Epac1 involves its DEP domain (amino acids 50 to 148) and requires the cAMP-induced conformation.In this study, we reveal an additional targeting mechanism of Epac1 by showing that its N terminus interacts with members of the ezrin-radixin-moesin (ERM) family. ERM proteins show high sequence similarity and function as scaffolding proteins that link the actin cytoskeleton to the PM (18, 42, 47). Inactive ERM proteins reside in the cytoplasm in an autoinhibited state maintained by an intramolecular interaction between the N-terminal FERM (4.1 protein, ezrin, radixin, moesin) domain and the C-terminal actin binding domain (ABD). This autoinhibition is released by binding to phosphatidylinositol-4,5-bisphosphate (PIP2) and threonine phosphorylation of the ABD, which induce the open conformation of the protein (reviewed in reference 8). Several kinases have been implicated in phosphorylation of this threonine in the ABD, including protein kinase C α (PKC α), PKC θ, NIK, Mst4, and the Rho effector ROCK (2, 40, 46, 50, 61). Active ERM proteins directly link the actin cytoskeleton to the PM and allow the recruitment of multiple signaling proteins. In this manner, ERM proteins function in numerous processes, such as the formation of microvilli, adherens junction stabilization, and leukocyte polarization (12, 18, 42, 47). Here, we demonstrate that ERM proteins also function as PM anchors for Epac1. The underlying interaction is mediated by the N terminus (residues 1 to 49) of Epac1 and is independent of its conformational state. Instead, the interaction is regulated at the level of the ERM proteins, which bind Epac1 when they are in their active, open conformation. G protein-coupled receptor (GPCR)-mediated signaling that results in activation of ERM proteins increases binding of Epac1 and results in a clustered localization of Epac1 at the PM. Together with DEP domain-mediated PM translocation, ERM proteins control cell adhesion mediated by Epac1. In conclusion, our data show that ERM proteins mediate PM recruitment of Epac1 and couple Epac1 activity to integrin-mediated cell adhesion.  相似文献   

4.
GAP1IP4BP is a member of the GAP1 family of Ras GTPase-activating proteins (GAPs) that includes GAP1m, CAPRI, and RASAL. Composed of a central Ras GAP-related domain (RasGRD), surrounded by amino-terminal C2 domains and a carboxy-terminal PH/Btk domain, these proteins, with the notable exception of GAP1m, possess an unexpected arginine finger-dependent GAP activity on the Ras-related protein Rap1 (S. Kupzig, D. Deaconescu, D. Bouyoucef, S. A. Walker, Q. Liu, C. L. Polte, O. Daumke, T. Ishizaki, P. J. Lockyer, A. Wittinghofer, and P. J. Cullen, J. Biol. Chem. 281:9891-9900, 2006). Here, we have examined the mechanism through which GAP1IP4BP can function as a Rap1 GAP. We show that deletion of domains on either side of the RasGRD, while not affecting Ras GAP activity, do dramatically perturb Rap1 GAP activity. By utilizing GAP1IP4BP/GAP1m chimeras, we establish that although the C2 and PH/Btk domains are required to stabilize the RasGRD, it is this domain which contains the catalytic machinery required for Rap1 GAP activity. Finally, a key residue in Rap1-specific GAPs is a catalytic asparagine, the so-called asparagine thumb. By generating a molecular model describing the predicted Rap1-binding site in the RasGRD of GAP1IP4BP, we show that mutagenesis of individual asparagine or glutamine residues that lie in close proximity to the predicted binding site has no detectable effect on the in vivo Rap1 GAP activity of GAP1IP4BP. In contrast, we present evidence consistent with a model in which the RasGRD of GAP1IP4BP functions to stabilize the switch II region of Rap1, allowing stabilization of the transition state during GTP hydrolysis initiated by the arginine finger.The Ras-like family of small GTPases are ubiquitously expressed, evolutionarily conserved proteins that, by undergoing conformational changes in response to the alternate binding of GDP and GTP, function as binary switches (28, 31, 35). The GDP-bound “off” state and the GTP-bound “on” state recognize distinct effector proteins, thereby allowing the regulation of a variety of downstream signaling events (28, 31, 35). While Ras is the best-known and best-studied Ras-like GTPase, Rap1 has recently attracted considerable attention (reviewed in reference 20).Rap1 was originally identified through its ability, when overexpressed, to reverse the phenotype of K-Ras-transformed NIH 3T3 cells (19). As Ras and Rap1 have very similar effector regions, the ability of Rap1 to reverse the transformed phenotype appeared to arise through an ability to compete with K-Ras effectors. For example, Rap1 binds the Ras effector Raf1 but this does not lead to its activation (11). This is consistent with a simple model in which Rap1 functions as a Ras antagonist (6, 37). However, recent work has challenged this view. Increasing evidence points to Rap1 interacting with its own panel of effectors through which it controls cell-cell adhesion and cell-matrix interactions (reviewed in reference 20).Like that of other GTPases, the activation of Ras and Rap1 is regulated through guanine nucleotide exchange factors, which control activation by stimulating the exchange of GDP for GTP. Inactivation is driven by GTPase-activating proteins (GAPs). These enhance the intrinsic GTPase activity of Ras and Rap1, thereby leading to GTP hydrolysis. A wide variety of guanine nucleotide exchange factors and GAPs specific for these GTPases have been identified (14). Through the arrangement of different modular domains, these proteins are regulated following the activation of cell surface receptors. This occurs either through direct association with the activated receptor or indirectly through second messengers (4, 5, 14, 41).Mammalian proteins capable of functioning as Ras GAPs include NF1 (3, 27, 40), p120GAP (38), the semaphorin 4D receptor plexin-B1 (29), and members of the GAP1 (reviewed in reference 41) and SynGAP (DAB2IP, nGAP, and SynGAP) families (10, 18, 39). These function as Ras GAPs by supplying a catalytic arginine residue—the arginine finger—into the active site of Ras. This stabilizes the transition state of the GTPase reaction, increasing the reaction rate by more than 1,000-fold (1, 33, 34).Rap1 GAPs include Rap GAPs I and II, the SPA-1 family (SPA-1, SPAR, SPAL, and E6TP1), and tuberin (16, 17, 26, 32). Unlike Ras, Rap1 does not possess the catalytic glutamine residue that is critical for GTP hydrolysis in Ras. This fundamental difference means that the mechanisms by which Ras and Rap1 GAPs function are distinct. Rap1 GAPs do not employ a catalytic arginine residue (8, 9); instead, they provide a catalytic asparagine—the asparagine thumb—to stimulate GTP hydrolysis (15). Here the asparagine carboxamide side chain has a function similar to that of the glutamine residue in Ras, stabilizing the position of the nucleophilic water and γ-phosphate in the transition complex (15, 36).Given such distinct catalytic mechanisms, surprisingly, some Ras GAPs, while having no detectable sequence homology with any Rap1 GAPs, are capable of stimulating the GTPase activity of Rap1. The first protein found to display such dual activities was GAP1IP4BP (13) (also known as RASA3, GAPIII, and R-Ras GAP). This is a member of the GAP1 family, which also comprises GAP1m, CAPRI, and RASAL (2, 23-25). These proteins are characterized by a domain architecture comprising amino-terminal tandem C2 domains, a highly conserved central Ras GAP-related domain (RasGRD), and a carboxy-terminal pleckstrin homology (PH) domain that is associated with a Bruton''s tyrosine kinase (Btk) motif (41). Consistent with the presence of the RasGRD, all proteins display Ras GAP activity, although each is differentially regulated following receptor stimulation (41). With the notable exception of GAP1m, all GAP1 proteins also possess efficient Rap1 GAP activity (22). Such dual specificity is not restricted solely to GAP1 proteins. Recently, C2 domain-containing SynGAP—a neuronal Ras GAP—has also been shown to display Rap1 GAP activity (21), an activity that appears to require, alongside the RasGRD, the presence of a single C2 domain (30).Here we have examined the mechanism behind the dual Ras and Rap1 GAP activities of GAP1IP4BP. Through the generation of a series of GAP1IP4BP/GAP1m chimeras, we have established that while the C2 domains of GAP1IP4BP are required to stabilize the RasGRD, these domains do not supply catalytic residues required for Rap1 GAP activity. Rather, the Rap1 GAP catalytic machinery appears to reside solely within the RasGRD. By the site-directed mutagenesis of selected asparagine and glutamine residues within this domain—selected following the generation of a predicted molecular model of the GAP1IP4BP RasGRD-Ras(Rap1) complex—we establish that the ability of GAP1IP4BP to function as a Rap1 GAP does not occur via a mechanism that utilizes a classic asparagine thumb. Rather, we suggest that the GAP1IP4BP RasGRD functions to stabilize the switch II region of Rap1 in a manner that allows a catalytic arginine finger from GAP1IP4BP to drive the hydrolysis of GTP.  相似文献   

5.
Integrin signaling plays critical roles in cell adhesion, spreading, and migration, and it is generally accepted that to regulate these integrin functions accurately, localized actin remodeling is required. However, the molecular mechanisms that control the targeting of actin regulation molecules to the proper sites are unknown. We previously demonstrated that integrin-mediated cell spreading and migration on fibronectin are dependent on the localized activation of phospholipase D (PLD). However, the mechanism underlying PLD activation by integrin is largely unknown. Here we demonstrate that protein kinase Cδ (PKCδ) is required for integrin-mediated PLD signaling. After integrin stimulation, PKCδ is activated and translocated to the edges of lamellipodia, where it colocalizes with PLD2. The abrogation of PKCδ activity inhibited integrin-induced PLD activation and cell spreading. Finally, we show that Thr566 of PLD2 is directly phosphorylated by PKCδ and that PLD2 mutation in this region prevents PLD2 activation, PLD2 translocation to the edge of lamellipodia, Rac translocation, and cell spreading after integrin activation. Together, these results suggest that PKCδ is a primary regulator of integrin-mediated PLD activation via the direct phosphorylation of PLD, which is essential for directing integrin-induced cell spreading.Integrin-mediated cell adhesion, spreading, and migration, which are essential for cellular differentiation, proliferation, survival, chemotaxis, and wound healing, require cell polarization with an environmental stimulus (32). To regulate these integrin-mediated functions accurately, coordinated and spatial control of localized cytoskeletal rearrangement is required. The key downstream signaling molecules of integrin-mediated actin cytoskeletal rearrangements include small GTPases of the Rho family, such as Rho, Cdc42, and Rac (57, 58). Recently it was suggested that integrin indirectly regulates the recruitment of small G proteins and their localized activation at a specific plasma membrane region called the cholesterol-enriched membrane microdomain. Furthermore, the membrane targeting of these molecules appears to be required for the activation of downstream effectors that induce actin reorganization (8, 9, 48). However, in the absence of integrin signaling, despite the GTP loading status, activated Rac and Cdc42 remain in the cytosol and cannot activate downstream effectors, such as p21-activated kinase (PAK) (8). This regulation of the localization of small GTPases to a specific site is supported by the observation that Rac1 is localized and activated at the leading edges of migrating cells, while Cdc42 is also activated in cellular protrusions and in the peripheral region (33, 51). The differentially localized activation of small GTPases results in coordinated spatially confined signaling leading to cytoskeletal rearrangement, which is critical for the regulation of integrin-mediated cell spreading and directional cell migration.The hydrolysis of phosphatidylcholine by phospholipase D1 (PLD1) and PLD2 generates the messenger lipid phosphatidic acid (PA) in response to a variety of signals, which include hormones, neurotransmitters, and growth factors (17). It has been shown that PA affects actin cytoskeletal rearrangement and hence lamellipodium extension and integrin-mediated cell spreading as well as migration. PLD activity has been found in detergent-insoluble membrane fractions in which a wide variety of cytoskeletal proteins, such as F-actin, α-actinin, vinculin, paxillin, and talin, were enriched (34). Furthermore, the stimulation of PLD with physiologic and pharmacologic agonists results in its association with actin filaments (34). In addition, actin polymerization and stress fiber formation are tightly coupled to the activation of PLD (14). The formation of lamellipodium structures and membrane ruffles is blocked by PLD inhibition (53, 60), and PLD activity is critical for epithelial cell, leukocyte, and neutrophil adhesion and migration (41, 43, 52). Furthermore, we have previously shown that the activity of PLD is upregulated, and that the activated PLD is translocated to lamellipodia, after integrin activation (3). The PLD product PA acts as a lipid anchor for the membrane translocation of Rac, and this PA-mediated localized activation of Rac is critical for integrin-mediated cell spreading and migration through Rac downstream signaling activation and actin cytoskeleton rearrangement (3). However, the mechanisms that regulate the activation and localization of PLD, which induce the localized downstream activation of integrin signaling, have not been elucidated.Members of the protein kinase C (PKC) family of serine-threonine kinases are known to play important roles in the transduction of signals from the activation of integrin to cell adhesion and spreading, as well as in cell migration via actin reorganization (11, 25, 61, 66). Several studies have shown that the activities of several PKC isozymes are modulated and are crucially required for integrin-mediated cell spreading and migration. The PKCα, -δ, and -ɛ isotypes were activated and then translocated from the cytosol to the membrane after integrin activation, and inhibition of these PKC isozymes prevented cell spreading (10, 47, 66). In addition, the activation of PKCα, -δ, and -ɛ rescued the spreading of α5 integrin-deficient cells on fibronectin (10), and PKCβΙ mediated platelet cell spreading and migration on fibrinogen (2). It has also been demonstrated that PKCθ activity is involved in endothelial cell migration (65). These results suggest that the kinase activities of diverse members of PKC are involved in the integrin-mediated signaling pathway leading to the actin cytoskeletal rearrangement required for cell spreading and migration. Several PKC substrates are known to influence the actin cytoskeleton directly (42). However, the natures of the isoform-specific functions of PKC members and of their specific downstream effectors for actin cytoskeletal rearrangement induction by integrin signaling remain to be elucidated.In this study, we found that PKCδ is an upstream modulator of localized PLD activation in the integrin signaling pathway. We demonstrate for the first time that PKCδ activity (not PKCα or PKCɛ activity) is critical for integrin-mediated PLD activation, and we found that PLD2 is phosphorylated at Thr566 by PKCδ in the integrin signaling pathway. Furthermore, we show that this phosphorylation is critical for integrin-mediated targeting of PLD to membrane ruffles, Rac translocation to the membrane, and lamellipodium formation during cell spreading. These findings strongly suggest a bridge between PKCδ and the signaling of actin cytoskeletal rearrangement by the integrin signaling pathway via PLD activation, and they provide a novel molecular mechanism for localized PLD activation via PKCδ phosphorylation, which is critical for the actin cytoskeletal rearrangements required for integrin-mediated cell spreading.  相似文献   

6.
The mammalian target of rapamycin (mTOR) is a conserved Ser/Thr kinase that forms two functionally distinct complexes important for nutrient and growth factor signaling. While mTOR complex 1 (mTORC1) regulates mRNA translation and ribosome biogenesis, mTORC2 plays an important role in the phosphorylation and subsequent activation of Akt. Interestingly, mTORC1 negatively regulates Akt activation, but whether mTORC1 signaling directly targets mTORC2 remains unknown. Here we show that growth factors promote the phosphorylation of Rictor (rapamycin-insensitive companion of mTOR), an essential subunit of mTORC2. We found that Rictor phosphorylation requires mTORC1 activity and, more specifically, the p70 ribosomal S6 kinase 1 (S6K1). We identified several phosphorylation sites in Rictor and found that Thr1135 is directly phosphorylated by S6K1 in vitro and in vivo, in a rapamycin-sensitive manner. Phosphorylation of Rictor on Thr1135 did not affect mTORC2 assembly, kinase activity, or cellular localization. However, cells expressing a Rictor T1135A mutant were found to have increased mTORC2-dependent phosphorylation of Akt. In addition, phosphorylation of the Akt substrates FoxO1/3a and glycogen synthase kinase 3α/β (GSK3α/β) was found to be increased in these cells, indicating that S6K1-mediated phosphorylation of Rictor inhibits mTORC2 and Akt signaling. Together, our results uncover a new regulatory link between the two mTOR complexes, whereby Rictor integrates mTORC1-dependent signaling.The mammalian target of rapamycin (mTOR) is an evolutionarily conserved phosphatidylinositol 3-kinase (PI3K)-related Ser/Thr kinase that integrates signals from nutrients, energy sufficiency, and growth factors to regulate cell growth as well as organ and body size in a variety of organisms (reviewed in references 4, 38, 49, and 77). mTOR was discovered as the molecular target of rapamycin, an antifungal agent used clinically as an immunosuppressant and more recently as an anticancer drug (5, 20). Recent evidence indicates that deregulation of the mTOR pathway occurs in a majority of human cancers (12, 18, 25, 46), suggesting that rapamycin analogs may be potent antineoplastic therapeutic agents.mTOR forms two distinct multiprotein complexes, the rapamycin-sensitive and -insensitive mTOR complexes 1 and 2 (mTORC1 and mTORC2), respectively (6, 47). In cells, rapamycin interacts with FKBP12 and targets the FKBP12-rapamycin binding (FRB) domain of mTORC1, thereby inhibiting some of its function (13, 40, 66). mTORC1 is comprised of the mTOR catalytic subunit and four associated proteins, Raptor (regulatory associated protein of mTOR), mLST8 (mammalian lethal with sec13 protein 8), PRAS40 (proline-rich Akt substrate of 40 kDa), and Deptor (28, 43, 44, 47, 59, 73, 74). The small GTPase Rheb (Ras homolog enriched in brain) is a key upstream activator of mTORC1 that is negatively regulated by the tuberous sclerosis complex 1 (TSC1)/TSC2 GTPase-activating protein complex (reviewed in reference 35). mTORC1 is activated by PI3K and Ras signaling through direct phosphorylation and inactivation of TSC2 by Akt, extracellular signal-regulated kinase (ERK), and p90 ribosomal protein S6 kinase (RSK) (11, 37, 48, 53, 63). mTORC1 activity is also regulated at the level of Raptor. Whereas low cellular energy levels negatively regulate mTORC1 activity through phosphorylation of Raptor by AMP-activated protein kinase (AMPK) (27), growth signaling pathways activating the Ras/ERK pathway positively regulate mTORC1 activity through direct phosphorylation of Raptor by RSK (10). More recent evidence has also shown that mTOR itself positively regulates mTORC1 activity by directly phosphorylating Raptor at proline-directed sites (20a, 75). Countertransport of amino acids (55) and amino acid signaling through the Rag GTPases were also shown to regulate mTORC1 activity (45, 65). When activated, mTORC1 phosphorylates two main regulators of mRNA translation and ribosome biogenesis, the AGC (protein kinase A, G, and C) family kinase p70 ribosomal S6 kinase 1 (S6K1) and eukaryotic initiation factor 4E-binding protein 1 (4E-BP1), and thus stimulates protein synthesis and cellular growth (50, 60).The second mTOR complex, mTORC2, is comprised of mTOR, Rictor (rapamycin-insensitive companion of mTOR), mSin1 (mammalian stress-activated mitogen-activated protein kinase-interacting protein 1), mLST8, PRR5 (proline-rich region 5), and Deptor (21, 39, 58, 59, 66, 76, 79). Rapamycin does not directly target and inhibit mTORC2, but long-term treatment with this drug was shown to correlate with mTORC2 disassembly and cytoplasmic accumulation of Rictor (21, 39, 62, 79). Whereas mTORC1 regulates hydrophobic motif phosphorylation of S6K1, mTORC2 has been shown to phosphorylate other members of the AGC family of kinases. Biochemical and genetic evidence has demonstrated that mTORC2 phosphorylates Akt at Ser473 (26, 39, 68, 70), thereby contributing to growth factor-mediated Akt activation (6, 7, 52). Deletion or knockdown of the mTORC2 components mTOR, Rictor, mSin1, and mLST8 has a dramatic effect on mTORC2 assembly and Akt phosphorylation at Ser473 (26, 39, 79). mTORC2 was also shown to regulate protein kinase Cα (PKCα) (26, 66) and, more recently, serum- and glucocorticoid-induced protein kinase 1 (SGK1) (4, 22). Recent evidence implicates mTORC2 in the regulation of Akt and PKCα phosphorylation at their turn motifs (19, 36), but whether mTOR directly phosphorylates these sites remains a subject of debate (4).Activation of mTORC1 has been shown to negatively regulate Akt phosphorylation in response to insulin or insulin-like growth factor 1 (IGF1) (reviewed in references 30 and 51). This negative regulation is particularly evident in cell culture models with defects in the TSC1/TSC2 complex, where mTORC1 and S6K1 are constitutively activated. Phosphorylation of insulin receptor substrate-1 (IRS-1) by mTORC1 (72) and its downstream target S6K1 has been shown to decrease its stability and lead to an inability of insulin or IGF1 to activate PI3K and Akt (29, 69). Although the mechanism is unknown, platelet-derived growth factor receptor β (PDGF-Rβ) has been found to be downregulated in TSC1- and TSC2-deficient murine embryonic fibroblasts (MEFs), contributing to a reduction of PI3K signaling (80). Interestingly, inhibition of Akt phosphorylation by mTORC1 has also been observed in the presence of growth factors other than IGF-1, insulin, or PDGF, suggesting that there are other mechanisms by which mTORC1 activation restricts Akt activity in cells (reviewed in references 6 and 31). Recent evidence demonstrates that rapamycin treatment causes a significant increase in Rictor electrophoretic mobility (2, 62), suggesting that phosphorylation of the mTORC2 subunit Rictor may be regulated by mTORC1 or downstream protein kinases.Herein, we demonstrate that Rictor is phosphorylated by S6K1 in response to mTORC1 activation. We demonstrate that Thr1135 is directly phosphorylated by S6K1 and found that a Rictor mutant lacking this phosphorylation site increases Akt phosphorylation induced by growth factor stimulation. Cells expressing the Rictor T1135A mutant were found to have increased Akt signaling to its substrates compared to Rictor wild-type- and T1135D mutant-expressing cells. Together, our results suggest that Rictor integrates mTORC1 signaling via its phosphorylation by S6K1, resulting in the inhibition of mTORC2 and Akt signaling.  相似文献   

7.
We report the molecular mechanisms that underlie chemotaxis of macrophages and cell migration of fibroblasts, cells that are essential during the body''s innate immune response and during wound repair, respectively. Silencing of phospholipase D1 (PLD1) and PLD2 reduced cell migration (both chemokinesis and chemotaxis) by ∼60% and >80%, respectively; this migration was restored by cell transfection with PLD2 constructs refractory to small interfering RNA (siRNA). Cells overexpressing active phospholipase D1 (PLD1) but, mostly, active PLD2 exhibited cell migration capabilities that were elevated over those elicited by chemoattractants alone. The mechanism for this enhancement is complex. It involves two pathways: one that is dependent on the activity of the lipase (and signals through its product, phosphatidic acid [PA]) and another that involves protein-protein interactions. The first is evidenced by partial abrogation of chemotaxis with lipase activity-defective constructs (PLD2-K758R) and by n-butanol treatment of cells. The second is evidenced by PLD association with the growth factor receptor-bound protein 2 (Grb2) through residue Y169, located within a Src homology 2 (SH2) consensus site. The association Grb2-PLD2 could be visualized by fluorescence microscopy in RAW/LR5 macrophages concentrated in actin-rich membrane ruffles, making possible that Grb2 serves as a docking or intermediary protein. The Grb2/PLD2-mediated chemotaxis process also depends on Grb2''s ability to recognize other motility proteins, like the Wiskott-Aldrich syndrome protein (WASP). Cell transfection with WASP, PLD2, and Grb2 constructs yields the highest levels of cell migration response, particularly in a macrophage cell line (RAW/LR5) and only modestly in the fibroblast cell line COS-7. Further, RAW/LR5 macrophages utilize for cell migration an additional pathway that involves S6 kinase (S6K) through PLD2-Y296, known to be phosphorylated by epidermal growth factor receptor (EGFR) kinase. Thus, both fibroblasts and macrophages use activity-dependent and activity-independent signaling mechanisms. However, highly mobile cells like macrophages use all signaling machinery available to them to accomplish their required function in rapid immune response, which sets them apart from fibroblasts, cells normally nonmobile that are only briefly involved in wound healing.Normalcy of migration is found in the ability of leukocytes to move toward foreign invaders of the body in phagocytic and immunogenic responses. Fibroblasts and endothelial cells migrate to aid wound repair and deposit collagen around the wounded area (32, 41). Migration is also important in embryogenesis and angiogenesis (37). Cellular migration may be involved not only in the normal physiological state of some cells but also in the pathological state, as in metastasis, whereby tumor cells migrate from the initial tumor site into the circulatory system and establish new colonies (32). Before migration can occur, cells must orient themselves to establish a distinct cell front and rear. The orientation of the cell depends on the inflammation site and location of the chemical stimuli or chemoattractants being released (32, 35). A localized enrichment of filamentous actin (F-actin) under the cell membrane occurs shortly after cell stimulation and is responsible for the formation of the lamellipodia and other morphological changes, such as membrane protrusions or ruffles that occur right before chemotaxis (28).Monocyte chemoattractant protein 1 (MCP-1), macrophage inflammatory protein 1α (MIP-1α), and colony-stimulating factor (CSF-1, also known as macrophage colony-stimulating factor [M-CSF]) are chemoattractants found to increase cell motility of monocytes and macrophages (52). MIP-1α acts through G-protein-coupled cell surface receptors that are highly abundant on those cells (42) and activate the phosphatidylinositol 3-kinase (PI3K) pathway and phospholipase C (PLC) (49). CSF-1 plays essential roles in the ability of monocytes to survive, proliferate, differentiate, and mature (49) and is also a monocyte/macrophage chemoattractant. CSF-1 acts by binding to cell surface receptors (CSF-1R) encoded by the c-fms proto-oncogene (48). Upon the binding of M-CSF, the CSF-1R dimerizes, and the tyrosine kinase domain is activated, resulting in transphosphorylation of the receptor. Grb2 and PI3K bind to two of the four phosphotyrosyl residues created, and the signal is transmitted to the cell interior (8).Epidermal growth factor (EGF) triggers proliferation and differentiation in fibroblasts (6, 37). EGF also serves other biological functions, such as cell rounding, ruffling, actin cytoskeletal reorganization, filopodium extension, and cell motility (37). Tyrosines that are autophosphorylated at the EGF receptor (EGFR) C terminus upon ligand binding enable binding to Src homology 2 (SH2) and phosphotyrosine binding (PTB) domains (37). Receptor kinase activity, along with at least one of the C-terminal tyrosine autophosphorylation sites, is required for cell movement (9). In addition, PLCγ and protein kinase C (PKC) have been linked to EGF and its ability to enhance cell motility (9, 10).As for the downstream signaling mechanism of these cell surface receptors, the recognized key players are the small GTPases, Rho, Rac, and cdc42. They are activated during actin cytoskeleton rearrangement and in cellular migration (7). New evidence has implicated other signaling proteins. Phospholipase D (PLD) has been found to play a role in leukocyte chemotaxis and adhesion (35). PLD is also involved in the regulation of essential cellular functions, largely due to the production of second messengers such as phosphatidic acid (PA) and ultimately diacylglycerol (DAG) (2, 11, 21, 22, 25, 38, 47, 53). Once produced, PA is involved in many cellular functions, including cytoskeletal rearrangement, phagocytosis, vesicle trafficking, exocytosis, and neuronal and cardiac stimulation (1, 11, 21, 30, 38). PA mediates chemotaxis, as increasing concentrations of PA enhanced the rate of cell migration of phagocytes (35). In the murine lymphoma cell line EL4, Knoepp et al. found that activated PLD2 promotes phosphorylation of FAK and Akt, leading to cell-substrate adhesion (7, 29). However, while inactivated PLD2 inhibits adhesion, migration, proliferation, and tumor invasion, it does not alter the basal level of FAK and Akt phosphorylation (29). Although PA does play a role in cell migration, the specific mechanisms involved are not completely understood, and more precise structure-function studies are needed.We have reported earlier an association of PLD2 and Grb2 important for DNA synthesis/cell proliferation at the level of Y179 (13) and Y511 (23). This study uncovers that migrating cells also use PLD2 and Grb2 but through a different mechanism, involving residue Y169 and the presence of the Wiskott-Aldrich syndrome protein (WASP). We also present new evidence implicating PLD2-Y296, known to be phosphorylated by EGFR kinase (24), in a pathway utilizing S6 kinase (S6K). This additional pathway serves to boost the ability of RAW/LR5 cells to migrate more readily than the other type of cell utilized in this study, COS-7 fibroblasts.  相似文献   

8.
The mammalian target of rapamycin (mTOR) complex 1 (mTORC1) functions as a rapamycin-sensitive environmental sensor that promotes cellular biosynthetic processes in response to growth factors and nutrients. While diverse physiological stimuli modulate mTORC1 signaling, the direct biochemical mechanisms underlying mTORC1 regulation remain poorly defined. Indeed, while three mTOR phosphorylation sites have been reported, a functional role for site-specific mTOR phosphorylation has not been demonstrated. Here we identify a new site of mTOR phosphorylation (S1261) by tandem mass spectrometry and demonstrate that insulin-phosphatidylinositol 3-kinase signaling promotes mTOR S1261 phosphorylation in both mTORC1 and mTORC2. Here we focus on mTORC1 and show that TSC/Rheb signaling promotes mTOR S1261 phosphorylation in an amino acid-dependent, rapamycin-insensitive, and autophosphorylation-independent manner. Our data reveal a functional role for mTOR S1261 phosphorylation in mTORC1 action, as S1261 phosphorylation promotes mTORC1-mediated substrate phosphorylation (e.g., p70 ribosomal protein S6 kinase 1 [S6K1] and eukaryotic initiation factor 4E binding protein 1) and cell growth to increased cell size. Moreover, Rheb-driven mTOR S2481 autophosphorylation and S6K1 phosphorylation require S1261 phosphorylation. These data provide the first evidence that site-specific mTOR phosphorylation regulates mTORC1 function and suggest a model whereby insulin-stimulated mTOR S1261 phosphorylation promotes mTORC1 autokinase activity, substrate phosphorylation, and cell growth.The mammalian target of rapamycin (mTOR), an evolutionarily conserved serine/threonine protein kinase, senses and integrates signals from diverse environmental cues (14, 31, 50, 74). mTOR associates with different partner proteins to form functionally distinct signaling complexes (4). The immunosuppressive drug rapamycin acutely inhibits signaling by mTOR complex 1 (mTORC1) (22), which contains mTOR, mLST8/GβL, raptor, and PRAS40 (24, 33, 34, 54, 67). Rapamycin fails to acutely inhibit signaling by mTORC2, which contains mTOR, mLST8/GβL, rictor, mSin1, and PRR5/Protor (18, 32, 47, 55, 73, 76). mTORC1 promotes various biosynthetic processes, including protein synthesis, cell growth (an increase in cell mass and size), and cell proliferation (an increase in cell number) (14, 40, 74). During growth factor (e.g., insulin) and nutrient (e.g., amino acids and glucose) sufficiency, mTORC1 phosphorylates the translational regulators p70 ribosomal protein S6 kinase 1 (S6K1) and eukaryotic initiation factor 4E binding protein 1 (4EBP1) to coordinately upregulate protein biosynthesis (40). Both S6K1 and 4EBP1 contain a TOR signaling motif, which mediates their interaction with raptor and thus facilitates their recruitment to the mTOR kinase (10, 44, 57, 58). In addition to regulating protein synthesis, mTORC1-mediated phosphorylation of S6K1 and 4EBP also promotes cell growth and cell cycle progression (15, 16). While more recently identified and thus less well characterized than mTORC1, mTORC2 mediates the phosphorylation of AGC kinase family members (e.g., Akt [also known as protein kinase B, PKB], PKCα, and SGK1) on their hydrophobic motifs and modulates the organization of the actin cytoskeleton (20, 26, 32, 55, 56).The insulin pathway represents the best-characterized activator of mTORC1 signaling to date, and thus many signaling intermediates that link insulin receptor activation to mTORC1 have been identified (12, 31). Complementary work using Drosophila melanogaster genetics and mammalian cell culture identified TSC1 (hamartin) and TSC2 (tuberin) as upstream negative regulators of mTORC1 (27). Inactivation of either the TSC1 or TSC2 genes, whose protein products heterodimerize to form a tumor suppressor complex, causes the development of benign tumors in diverse organs in both humans and rodents, a disease known as tuberous sclerosis complex (TSC) (36). TSC2 contains a GTPase-activating protein domain that acts on Rheb, a Ras-like GTP binding protein that activates mTORC1 (27). Thus, in TSC-deficient cells, constitutive Rheb-GTP leads to chronically high mTORC1 signaling. While the mechanism by which Rheb-GTP activates mTORC1 remains incompletely understood, Rheb coimmunoprecipitates with mTOR and directly activates mTORC1 kinase activity in vivo and in vitro when GTP bound (2, 38, 54). Rheb has been reported to augment the activity of PLD1, an enzyme that catalyzes the production of the lipid second messenger phosphatidic acid, which contributes to the mitogenic activation of mTORC1 signaling (13, 62). Additionally, Rheb-GTP was reported to induce the dissociation of the endogenous mTOR inhibitor FKBP38 (3), although aspects of this model have been questioned (72). Insulin/phosphatidylinositol 3-kinase (PI3K) signaling reduces the inhibitory effect of TSC on mTORC1 via Akt-mediated phosphorylation of TSC2 (29, 42, 64). Additionally, Ras-regulated signaling via mitogen-activated protein kinase (MAPK) and RSK also inhibits TSC via PI3K/Akt-independent phosphorylation of TSC2 (39, 51, 63). In contrast, glucose deprivation enhances TSC''s inhibitory effect on mTORC1 signaling via AMP-activated protein kinase (AMPK)-mediated phosphorylation of TSC2 (on different sites) (30). Thus, TSC functions as a central nexus of diverse physiological signals to fine-tune mTORC1 signaling depending on environmental conditions (27). While the mechanism by which amino acids promote mTORC1 signaling has remained elusive, compelling new data reveal that the Rag GTPases link amino acid sensing to mTORC1 activation (35, 52, 53). During amino acid sufficiency, GTP-bound Rag heterodimers bind raptor and recruit mTORC1 to an endomembrane compartment that contains the mTORC1 activator Rheb; thus, amino acid sufficiency may function to prime mTORC1 for subsequent growth factor-mediated activation via a dynamic subcellular redistribution mechanism (52).Despite the well-characterized regulation of mTORC1 signaling by growth factors (e.g., insulin), nutrients (e.g., amino acids and glucose), and cellular stress (e.g., hypoxia) and the identification of numerous signaling mediators of these pathways, the direct molecular mechanisms by which cellular signals modulate mTORC1 action remain obscure (31). While three phosphorylation sites (P-sites) on mTOR have been reported to date (T2446, S2448, and S2481), no function has yet been ascribed to any site (7, 43, 49, 59). Here we identify S1261 as a novel mTOR phosphorylation site in vivo in cultured mammalian cells and provide the first evidence that site-specific mTOR phosphorylation regulates mTORC1 function. We show that insulin signals via the PI3K/TSC/Rheb pathway in an amino acid-dependent and rapamycin-insensitive manner to promote mTOR S1261 phosphorylation, which regulates mTORC1 autokinase activity, biochemical signaling to downstream substrates, and cell growth to increased cell size, a major cellular function of mTORC1. Elucidation of the molecular mechanisms underlying mTORC1 regulation will enable us to better understand how mTORC1 senses environmental stimuli to control cellular physiology. As aberrantly upregulated mTORC1 signaling likely contributes to cancer, insulin-resistant diabetes, and cardiovascular diseases, understanding mTORC1 regulation may aid in the development of novel therapeutics for these prevalent human diseases (11, 21, 28).  相似文献   

9.
10.
Cell migration is critical for normal development and for pathological processes including cancer cell metastasis. Dynamic remodeling of focal adhesions and the actin cytoskeleton are crucial determinants of cell motility. The Rho family and the mitogen-activated protein kinase (MAPK) module consisting of MEK-extracellular signal-regulated kinase (ERK) are important regulators of these processes, but mechanisms for the integration of these signals during spreading and motility are incompletely understood. Here we show that ERK activity is required for fibronectin-stimulated Rho-GTP loading, Rho-kinase function, and the maturation of focal adhesions in spreading cells. We identify p190A RhoGAP as a major target for ERK signaling in adhesion assembly and identify roles for ERK phosphorylation of the C terminus in p190A localization and activity. These observations reveal a novel role for ERK signaling in adhesion assembly in addition to its established role in adhesion disassembly.Cell migration is a highly coordinated process essential for physiological and pathological processes (69). Signaling through Rho family GTPases (e.g., Rac, Cdc42, and Rho) is crucial for cell migration. Activated Rac and Cdc42 are involved in the production of a dominant lamellipodium and filopodia, respectively, whereas Rho-stimulated contractile forces are required for tail retraction and to maintain adhesion to the matrix (57, 58, 68). Rac- and Cdc42-dependent membrane protrusions are driven by the actin cytoskeleton and the formation of peripheral focal complexes; Rho activation stabilizes protrusions by stimulating the formation of mature focal adhesions and stress fibers. Active Rho influences cytoskeletal dynamics through effectors including the Rho kinases (ROCKs) (2, 3).Rho activity is stimulated by GEFs that promote GTP binding and attenuated by GTPase-activating proteins (GAPs) that enhance Rho''s intrinsic GTPase activity. However, due to the large number of RhoGEFs and RhoGAPs expressed in mammalian cells, the molecular mechanisms responsible for regulation of Rho activity in time and space are incompletely understood. p190A RhoGAP (hereafter p190A) is implicated in adhesion and migration signaling. p190A contains an N-terminal GTPase domain, a large middle domain juxtaposed to the C-terminal GAP domain, and a short C-terminal tail (74). The C-terminal tail of ∼50 amino acids is divergent between p190A and the closely related family member p190B (14) and thus may specify the unique functional roles for p190A and p190B revealed in gene knockout studies (10, 11, 41, 77, 78). p190A activity is dynamically regulated in response to external cues during cell adhesion and migration (5, 6, 59). Arthur et al. (5) reported that p190A activity is required for the transient decrease in RhoGTP levels seen in fibroblasts adhering to fibronectin. p190A activity is positively regulated by tyrosine phosphorylation (4, 5, 8, 17, 31, 39, 40, 42): phosphorylation at Y1105 promotes its association with p120RasGAP and subsequent recruitment to membranes or cytoskeleton (8, 17, 27, 31, 71, 75, 84). However, Y1105 phosphorylation is alone insufficient to activate p190A GAP activity (39). While the functions of p190A can be irreversibly terminated by ubiquitinylation in a cell-cycle-dependent manner (80), less is known about reversible mechanisms that negatively regulate p190A GAP activity during adhesion and motility.The integration of Rho family GTPase and extracellular signal-regulated kinase (ERK) signaling is important for cell motility (48, 50, 63, 76, 79). Several studies have demonstrated a requirement for ERK signaling in the disassembly of focal adhesions in migrating cells, in part through the activation of calpain proteases (36, 37) that can downregulate focal adhesion kinase (FAK) signaling (15), locally suppress Rho activity (52), and sever cytoskeletal linkers to focal adhesions (7, 33). Inhibition of ERK signaling increases focal adhesion size and retards disassembly of focal adhesions in adherent cells (57, 64, 85, 86). It is also recognized that ERK modulates Rho-dependent cellular processes, including membrane protrusion and migration (18, 25, 64, 86). Interestingly, ERK activated in response to acute fibronectin stimulation localizes not only to mature focal adhesions, but also to peripheral focal complexes (32, 76). Since these complexes can either mature or be turned over (12), ERK may play a distinct role in focal adhesion assembly. ERK is proposed to promote focal adhesion formation by activating myosin light chain kinase (MLCK) (21, 32, 50).Here we find that ERK activity is required for Rho activation and focal adhesion formation during adhesion to fibronectin and that p190A is an essential target of ERK signaling in this context. Inspection of the p190A C terminus reveals a number of consensus ERK sites and indeed p190A is phosphorylated by recombinant ERK only on its C terminus in vitro, and on the same C-terminal peptide in vivo. Mutation of the C-terminal ERK phosphorylation sites to alanine increases the biochemical and biological activity of p190A. Finally, inhibition of MEK or mutation of the C-terminal phosphorylation sites enhances retention of p190A in peripheral membranes during spreading on fibronectin. Our data support the conclusion that ERK phosphorylation inhibits p190A allowing increases in RhoGTP and cytoskeletal changes necessary for focal adhesion formation.  相似文献   

11.
12.
13.
Endothelial cell (EC) migration, cell-cell adhesion, and the formation of branching point structures are considered hallmarks of angiogenesis; however, the underlying mechanisms of these processes are not well understood. Lipid phosphate phosphatase 3 (LPP3) is a recently described p120-catenin-associated integrin ligand localized in adherens junctions (AJs) of ECs. Here, we tested the hypothesis that LPP3 stimulates β-catenin/lymphoid enhancer binding factor 1 (β-catenin/LEF-1) to induce EC migration and formation of branching point structures. In subconfluent ECs, LPP3 induced expression of fibronectin via β-catenin/LEF-1 signaling in a phosphatase and tensin homologue (PTEN)-dependent manner. In confluent ECs, depletion of p120-catenin restored LPP3-mediated β-catenin/LEF-1 signaling. Depletion of LPP3 resulted in destabilization of β-catenin, which in turn reduced fibronectin synthesis and deposition, which resulted in inhibition of EC migration. Accordingly, reexpression of β-catenin but not p120-catenin in LPP3-depleted ECs restored de novo synthesis of fibronectin, which mediated EC migration and formation of branching point structures. In confluent ECs, however, a fraction of p120-catenin associated and colocalized with LPP3 at the plasma membrane, via the C-terminal cytoplasmic domain, thereby limiting the ability of LPP3 to stimulate β-catenin/LEF-1 signaling. Thus, our study identified a key role for LPP3 in orchestrating PTEN-mediated β-catenin/LEF-1 signaling in EC migration, cell-cell adhesion, and formation of branching point structures.Angiogenesis, the formation of new blood vessels, involves several well-coordinated cellular processes, including endothelial cell (EC) migration, synthesis and deposition of extracellular matrix proteins, such as fibronectin, cell-cell adhesion, and formation of branching point structures (1-3, 19, 33); however, less is known about the underlying mechanisms of these processes (6, 8, 12, 14, 16, 17). For example, adherens junctions (AJs), which mediate cell-cell adhesion between ECs, may be involved in limiting the extent of cell migration (2, 14, 38, 40). VE-cadherin, a protein found in AJs, is a single-pass transmembrane polypeptide responsible for calcium-dependent homophilic interactions through its extracellular domains (2, 38, 40). The VE-cadherin cytoplasmic domain interacts with the Armadillo domain-containing proteins, β-catenin, γ-catenin (plakoglobin), and p120-catenin (p120ctn) (2, 15, 38, 40, 43). Genetic and biochemical evidence documents a crucial role of β-catenin in regulating cell adhesion as well as proliferation secondary to the central position of β-catenin in the Wnt signaling pathway (13, 16, 25, 31, 44). In addition, the juxtamembrane protein p120ctn regulates AJ stability via binding to VE-cadherin (2, 7, 9, 15, 21, 28, 32, 43). The absence of regulation or inappropriate regulation of β-catenin and VE-cadherin functions is linked to cardiovascular disease and tumor progression (2, 6).We previously identified lipid phosphate phosphatase 3 (LPP3), also known as phosphatidic acid phosphatase 2b (PAP2b), in a functional assay of angiogenesis (18, 19, 41, 42). LPP3 not only exhibits lipid phosphatase activity but also functions as a cell-associated integrin ligand (18, 19, 35, 41, 42). The known LPPs (LPP1, LPP2, and LPP3) (20-23) are six transmembrane domain-containing plasma membrane-bound enzymes that dephosphorylate sphingosine-1-phosphate (S1P) and its structural homologues, and thus, these phosphatases generate lipid mediators (4, 5, 23, 35, 39). All LPPs, which contain a single N-glycosylation site and a putative lipid phosphatase motif, are situated such that their N and C termini are within the cell (4, 5, 22, 23, 35, 39). Only the LPP3 isoform contains an Arg-Gly-Asp (RGD) sequence in the second extracellular loop, and this RGD sequence enables LPP3 to bind integrins (18, 19, 22). Transfection experiments with green fluorescent protein (GFP)-tagged LPP1 and LPP3 showed that LPP1 is apically sorted, whereas LPP3 colocalized with E-cadherin at cell-cell contact sites with other Madin-Darby canine kidney (MDCK) cells (22). Mutagenesis and domain swapping experiments established that LPP1 contains an apical targeting signal sequence (FDKTRL) in its N-terminal segment. In contrast, LPP3 contains a dityrosine (109Y/110Y) basolateral sorting motif (22). Interestingly, conventional deletion of Lpp3 is embryonic lethal, since the Lpp3 gene plays a critical role in extraembryonic vasculogenesis independent of its lipid phosphatase activity (11). In addition, an LPP3-neutralizing antibody was shown to prevent cell-cell interactions (19, 42) and angiogenesis (42). Here, we addressed the hypothesis that LPP3 plays a key role in EC migration, cell-cell adhesion, and formation of branching point structures by stimulating β-catenin/lymphoid enhancer binding factor 1 (β-catenin/LEF-1) signaling.  相似文献   

14.
Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding to CD4 and a chemokine receptor, most commonly CCR5. CXCR4 is a frequent alternative coreceptor (CoR) in subtype B and D HIV-1 infection, but the importance of many other alternative CoRs remains elusive. We have analyzed HIV-1 envelope (Env) proteins from 66 individuals infected with the major subtypes of HIV-1 to determine if virus entry into highly permissive NP-2 cell lines expressing most known alternative CoRs differed by HIV-1 subtype. We also performed linear regression analysis to determine if virus entry via the major CoR CCR5 correlated with use of any alternative CoR and if this correlation differed by subtype. Virus pseudotyped with subtype B Env showed robust entry via CCR3 that was highly correlated with CCR5 entry efficiency. By contrast, viruses pseudotyped with subtype A and C Env proteins were able to use the recently described alternative CoR FPRL1 more efficiently than CCR3, and use of FPRL1 was correlated with CCR5 entry. Subtype D Env was unable to use either CCR3 or FPRL1 efficiently, a unique pattern of alternative CoR use. These results suggest that each subtype of circulating HIV-1 may be subject to somewhat different selective pressures for Env-mediated entry into target cells and suggest that CCR3 may be used as a surrogate CoR by subtype B while FPRL1 may be used as a surrogate CoR by subtypes A and C. These data may provide insight into development of resistance to CCR5-targeted entry inhibitors and alternative entry pathways for each HIV-1 subtype.Human immunodeficiency virus type 1 (HIV-1) infects target cells by binding first to CD4 and then to a coreceptor (CoR), of which C-C chemokine receptor 5 (CCR5) is the most common (6, 53). CXCR4 is an additional CoR for up to 50% of subtype B and D HIV-1 isolates at very late stages of disease (4, 7, 28, 35). Many other seven-membrane-spanning G-protein-coupled receptors (GPCRs) have been identified as alternative CoRs when expressed on various target cell lines in vitro, including CCR1 (76, 79), CCR2b (24), CCR3 (3, 5, 17, 32, 60), CCR8 (18, 34, 38), GPR1 (27, 65), GPR15/BOB (22), CXCR5 (39), CXCR6/Bonzo/STRL33/TYMSTR (9, 22, 25, 45, 46), APJ (26), CMKLR1/ChemR23 (49, 62), FPLR1 (67, 68), RDC1 (66), and D6 (55). HIV-2 and simian immunodeficiency virus SIVmac isolates more frequently show expanded use of these alternative CoRs than HIV-1 isolates (12, 30, 51, 74), and evidence that alternative CoRs other than CXCR4 mediate infection of primary target cells by HIV-1 isolates is sparse (18, 30, 53, 81). Genetic deficiency in CCR5 expression is highly protective against HIV-1 transmission (21, 36), establishing CCR5 as the primary CoR. The importance of alternative CoRs other than CXCR4 has remained elusive despite many studies (1, 30, 70, 81). Expansion of CoR use from CCR5 to include CXCR4 is frequently associated with the ability to use additional alternative CoRs for viral entry (8, 16, 20, 63, 79) in most but not all studies (29, 33, 40, 77, 78). This finding suggests that the sequence changes in HIV-1 env required for use of CXCR4 as an additional or alternative CoR (14, 15, 31, 37, 41, 57) are likely to increase the potential to use other alternative CoRs.We have used the highly permissive NP-2/CD4 human glioma cell line developed by Soda et al. (69) to classify virus entry via the alternative CoRs CCR1, CCR3, CCR8, GPR1, CXCR6, APJ, CMKLR1/ChemR23, FPRL1, and CXCR4. Full-length molecular clones of 66 env genes from most prevalent HIV-1 subtypes were used to generate infectious virus pseudotypes expressing a luciferase reporter construct (19, 57). Two types of analysis were performed: the level of virus entry mediated by each alternative CoR and linear regression of entry mediated by CCR5 versus all other alternative CoRs. We thus were able to identify patterns of alternative CoR use that were subtype specific and to determine if use of any alternative CoR was correlated or independent of CCR5-mediated entry. The results obtained have implications for the evolution of env function, and the analyses revealed important differences between subtype B Env function and all other HIV-1 subtypes.  相似文献   

15.
mTOR, the mammalian target of rapamycin, is a critical node for control of cell growth and survival and has widely been implicated in cancer survival signals. mTOR exists in two complexes: mTORC1 and mTORC2. Phospholipase D (PLD) and its metabolite phosphatidic acid (PA) have been implicated in the regulation of mTOR; however, their role has been controversial. We report here that suppression of PLD prevents phosphorylation of the mTORC1 substrate S6 kinase (S6K) at Thr389 and the mTORC2 substrate Akt at Ser473. Suppression of PLD also blocked insulin-stimulated Akt phosphorylation at Ser473 and the mTORC2-dependent phosphorylation of PRAS40. Importantly, PA was required for the association of mTOR with Raptor to form mTORC1 and that of mTOR with Rictor to form mTORC2. The effect of PA was competitive with rapamycin—with much higher concentrations of rapamycin needed to compete with the PA-mTORC2 interaction than with PA-mTORC1. Suppressing PA production substantially increased the sensitivity of mTORC2 to rapamycin. Data provided here demonstrate a PA requirement for the stabilization of both mTORC1 and mTORC2 complexes and reveal a mechanism for the inhibitory effect of rapamycin on mTOR. This study also suggests that by suppressing PLD activity, mTORC2 could be targeted therapeutically with rapamycin.It has become apparent during the past decade that a critical aspect of tumor progression is the suppression of default apoptotic programs that constitute what is likely the most important protection against cancer. Cellular signals that suppress apoptosis have come to be known as “survival signals.” A common node for survival signals is mTOR, the mammalian target of rapamycin (5, 13, 14, 25). mTOR exists in two distinct complexes, mTORC1 and mTORC2 (21), that differ in their subunit composition and sensitivity to rapamycin. mTORC1 consists of a complex that includes mTOR and a protein known as Raptor (regulatory associated protein of mTOR), whereas mTORC2 consists of a complex that includes mTOR and a protein known as Rictor (rapamycin-insensitive companion of mTOR) (13, 14). There are also mTORC2 complexes that can be distinguished by association with different isoforms of mSin1 (9). While much is known about the regulation of mTORC1 (21), very little is known about the regulation of mTORC2.mTORC1 is highly sensitive to rapamycin, whereas mTORC2 is relatively insensitive to rapamycin (21). However, it was recently reported that long-term exposure to rapamycin prevented the formation of mTORC2 complexes and blocked the phosphorylation of the mTORC2 substrate Akt at Ser473 (24, 38). Rapamycin, in association with FK506 binding protein 12 (FKBP12), has been reported to interact with mTOR in a manner that is competitive with phosphatidic acid (PA), the metabolic product of phospholipase D (PLD) (2, 4). PLD, like mTOR, has been implicated in survival signals in several human cancer cell lines (1, 10, 11, 27, 32, 39). Since rapamycin-FKBP12 competes with PA for binding to mTOR, the sensitivity of mTORC2 complex formation to rapamycin suggests that PA facilitates the assembly of mTORC2—and ultimately the activation of mTORC2. We report here that the assembly of both mTORC1 and mTORC2 complexes is dependent upon PLD and its metabolite PA. The study also provides mechanistic insight into how rapamycin impacts on mTOR-mediated signals and how PLD regulates mTOR by facilitating the formation of mTOR complexes.  相似文献   

16.
17.
LFA-1 regulates T cell activation and signal transduction through the immunological synapse. T cell receptor (TCR) stimulation rapidly activates LFA-1, which provides unique LFA-1-dependent signals to promote T cell activation. However, the detailed molecular pathways that regulate these processes and the precise mechanism by which LFA-1 contributes to TCR activation remain unclear. We found LFA-1 directly participates in Erk1/2 signaling upon TCR stimulation in CD8+ T cells. The presence of LFA-1, not ligand binding, is required for the TCR-mediated Erk1/2 signal pathway. LFA-1-deficient T cells have defects in sustained Erk1/2 signaling and TCR/CD3 clustering, which subsequently prevents MTOC reorientation, cell cycle progression, and mitosis. LFA-1 regulates the TCR-mediated Erk1/2 signal pathway in the context of immunological synapse for recruitment and amplification of the Erk1/2 signal. In addition, LFA-1 ligation with ICAM-1 generates an additional Erk1/2 signal, which synergizes with the existing TCR-mediated Erk1/2 signal to enhance T cell activation. Thus, LFA-1 contributes to CD8+ T cell activation through two distinct signal pathways. We demonstrated that the function of LFA-1 is to enhance TCR signaling through the immunological synapse and deliver distinct signals in CD8+ T cell activation.Leukocyte function-associated antigen-1 (LFA-1)2 plays an important role in regulating leukocyte adhesion and T cell activation (1, 2). LFA-1 consists of the αL (CD11a) and β2 (CD18) subunits. The ligands for LFA-1 include intercellular adhesion molecular-1 (ICAM-1), ICAM-2, and ICAM-3 (3). LFA-1 participates in the formation of the immunological synapse, which regulates T cell activation synergistically with TCR engagement. The immunological synapse is a specialized structure that forms between the T cell and the APC or target cell (1, 2, 4). The function of the immunological synapse is to facilitate T cell activation and signal transduction. Mice deficient in LFA-1 (CD11a KO) have defects in leukocyte adhesion, lymphocyte proliferation, and tumor rejection (57).Upon TCR stimulation, the nascent immunological synapse is initiated with surface receptor clustering and cytoskeleton rearrangement, then followed by mature synapse formation after prolonged stimulation (8, 9). In the mature immunological synapse, LFA-1 forms a ring-like pattern at the peripheral supramolecular activation cluster (pSMAC), which surrounds the central supramolecular activation cluster (cSMAC) containing TCR/CD3/lipid rafts (10, 11). The structure of the mature synapse is stable for hours and thought to be important for sustained TCR signaling (1214). LFA-1 functions via pSMAC to stabilize the cSMAC and is associated with the induction of T cell proliferation, cytokine production, and lytic granule migration toward cSMAC (1, 15). Although LFA-1-containing pSMAC is self-evident in lipid bilayer systems and cell lines, whether it is required for T cell activation under physiological conditions remains controversial (15).TCR stimulation rapidly induces the functional activity of LFA-1, which then provides unique LFA-1-dependent signals to promote T cell activation (16). The process can be divided into two steps. First, the intracellular signaling from TCR regulating LFA-1 activation is known as “inside-out” signaling; second, activated LFA-1, as a signaling receptor, can feedback to transduce the intracellular signal, the “outside-in” signaling (1, 17). It is widely accepted that TCR stimulation activates LFA-1 through affinity and/or avidity regulation, as supported by increased adhesion to ICAM-1 and pSMAC formation (16, 17). The “inside-out” signal process has been investigated extensively (1821). The TCR proximal signal molecules, Lck, ZAP-70, and PI3K, are known to be important for TCR signaling to LFA-1 activation (2226). The molecular mechanisms of LFA-1 “outside-in” signaling have been explored only recently. Perez et al. (27) have demonstrated that LFA-1 and ICAM-1 ligation activates the downstream Erk1/2 MAPK signaling pathway upon TCR stimulation, which ultimately leads to the qualitative modulation of CD4+ T cell activation through distinct LFA-1-dependent signals. Another recent study provided compelling evidence that LFA-1 reshapes the Ras MAPK pathway downstream of TCR (28). However, the detailed molecular pathways that regulate these processes are poorly defined. Especially, the evidence in support of a distinctive role for LFA-1 in the T cell signaling pathway has lagged behind; whether the function of LFA-1 is to enhance TCR signaling through the immunological synapse and/or deliver distinct signal in T cell activation and whether LFA-1 is indispensable for or merely assists the existing TCR signal pathway. Furthermore, whether and how TCR proximal signal molecules regulate LFA-1 function remains unknown. Further studies are required to understand the LFA-1 and TCR signaling network.In this study, we found that LFA-1 directly participates in CD8+ T cell activation. Upon TCR stimulation, LFA-1 regulates both TCR-mediated and LFA-1-mediated Erk1/2 signal pathways. First, the presence of LFA-1, not ligand binding, is required for the sustained Erk1/2 signaling and TCR/CD3 clustering on the surface of CD8+ T cells, subsequently leading to MTOC reorientation, cell cycle progression, and mitosis. Second, LFA-1 ligation with ICAM-1 enhances Erk1/2 signaling, which promotes T cell activation with increased IL-2 production and cell proliferation. This LFA-1-mediated Erk1/2 signal pathway integrates with the existing TCR-mediated Erk1/2 signal pathway to enhance T cell activation.  相似文献   

18.
Signal transduction is regulated by the lateral segregation of proteins into nanodomains on the plasma membrane. However, the molecular mechanisms that regulate the lateral segregation of cell surface receptors, such as receptor tyrosine kinases, upon ligand binding are unresolved. Here we used high-resolution spatial mapping to investigate the plasma membrane nanoscale organization of the epidermal growth factor (EGF) receptor (EGFR). Our data demonstrate that in serum-starved cells, the EGFR exists in preformed, cholesterol-dependent, actin-independent nanoclusters. Following stimulation with EGF, the number and size of EGFR nanoclusters increase in a time-dependent manner. Our data show that the formation of EGFR nanoclusters requires receptor tyrosine kinase activity. Critically, we show for the first time that production of phosphatidic acid by phospholipase D2 (PLD2) is essential for ligand-induced EGFR nanocluster formation. In accordance with its crucial role in regulating EGFR nanocluster formation, we demonstrate that modulating PLD2 activity tunes the degree of EGFR nanocluster formation and mitogen-activated protein kinase signal output. Together, these data show that EGFR activation drives the formation of signaling domains by regulating the production of critical second-messenger lipids and modifying the local membrane lipid environment.The epidermal growth factor (EGF) receptor (EGFR) is a single transmembrane domain protein that possesses intrinsic tyrosine kinase (TK) activity. Ligand binding to the extracellular domain induces conformational changes that promote activation of the intracellular TK domain. The kinase domain then autophosphorylates a number of tyrosine residues in the C-terminal region of the protein, creating docking sites for adapter and effector proteins. Thus, the active form of the EGFR could reasonably be expected to be a dimer. However, recent studies using single-molecule imaging, image correlation spectroscopy (ICS), fluorescence correlation spectroscopy (FCS), and immunoelectron microscopy (immuno-EM) show that the EGFR is, in fact, nonrandomly organized into oligomers on the plasma membrane (6, 7, 16, 34, 44). ICS measurements estimate that, in the absence of ligand, there are, on average, 2.2 EGFRs per cluster, which increases to 3.7 receptors per cluster upon stimulation (7). Single-molecule tracking experiments also suggest that unliganded EGFRs continually fluctuate between monomers and dimers that are primed for activation (5). Furthermore, the organization of the EGFR is dynamic and clustering of the EGFR increases over time after EGF stimulation (7, 16). However, neither the precise role of EGFR oligomerization in signal transduction nor the mechanisms driving oligomer formation have been resolved.The organization of the EGFR into oligomers is dependent upon cellular cholesterol. Saffarian et al., using FCS, estimated that 70% of EGFRs exist as monomers, 20% as dimers, and 10% as oligomers (34). However, depletion of cholesterol decreases the percentage of monomeric receptors and increases the proportion of oligomeric receptors. Cholesterol depletion and actin depolymerization also alter the diffusion coefficient of the EGFR and the confinement area size (22). The finding that EGFR membrane organization is dependent upon cholesterol is of particular interest because a number of studies have demonstrated that EGFR activity is negatively regulated by cholesterol (4, 23, 28, 32).Phospholipase D2 (PLD2) hydrolyzes phosphatidylcholine (PC) to produce choline and phosphatidic acid (PA). PLD2 is localized to the plasma membrane (10), associates with the EGFR (39), and is rapidly activated upon EGF stimulation, leading to increased production of PA (15, 38, 39). A number of lines of evidence suggest that PA is an important mediator of EGFR action. First, exogenous PA is mitogenic when incubated with cells (17, 19, 42, 45). Second, direct interaction with membrane PA regulates the activity of a number of components downstream of the EGFR, including Sos (47) and Raf (12, 13, 30, 31).In the current study, we used high-resolution spatial analysis techniques to investigate EGFR plasma membrane organization. Using these approaches, we identified PA as the key molecular component responsible for driving EGFR nanocluster formation in response to EGF binding and demonstrated that the level of PLD2 activity regulates the duration of mitogen-activated protein kinase (MAPK) signal output.  相似文献   

19.
20.
KSR1 is a mitogen-activated protein (MAP) kinase scaffold that enhances the activation of the MAP kinase extracellular signal-regulated kinase (ERK). The function of KSR1 in NK cell function is not known. Here we show that KSR1 is required for efficient NK-mediated cytolysis and polarization of cytolytic granules. Single-cell analysis showed that ERK is activated in an all-or-none fashion in both wild-type and KSR1-deficient cells. In the absence of KSR1, however, the efficiency of ERK activation is attenuated. Imaging studies showed that KSR1 is recruited to the immunological synapse during T-cell activation and that membrane recruitment of KSR1 is required for recruitment of active ERK to the synapse.Kinase suppressor of Ras was originally identified in Drosophila melanogaster (53) and Caenorhabditis elegans (19, 32, 52) as a positive regulator of the extracellular signal-regulated kinase (ERK) mitogen-activated protein (MAP) kinase signaling pathway. It is thought to function as a MAP kinase scaffold because it can bind to Raf, MEK, and ERK (18, 19, 27, 28, 44, 59). While the exact function of KSR is unknown, preassembling the three components of the ERK MAP kinase cascade could function to enhance the efficiency of ERK activation, potentially regulate the subcellular location of ERK activation, and promote access to specific subcellular substrates (16, 45, 46).While only one isoform of KSR is expressed in Drosophila (53), two KSR isoforms have been identified in C. elegans (19, 32, 52) and most higher organisms. They are referred to as KSR1 and KSR2 (32, 43). While KSR1 mRNA and protein are detectable in a wide variety of cells and tissues, including brain, thymus, and muscle (10, 11, 29), little is known about the expression pattern of KSR2.We previously reported the phenotype of KSR1-deficient mice (30). These mice are born at Mendelian ratios and develop without any obvious defects. Using gel filtration, we showed that KSR1 promotes the formation of large signaling complexes containing KSR1, Raf, MEK, and ERK (30). Using both primary T cells stimulated with antibodies to the T-cell receptor as well as fibroblasts stimulated with growth factors, we showed that KSR1-deficient cells exhibit an attenuation of ERK activation with defects in cell proliferation.Here we explored the role of KSR1 in NK cell-mediated cytolysis. The killing of a target cell by a cytolytic T cell or NK cell is a complicated process that involves cell polarization with microtubule-dependent movement of cytolytic granules to an area that is proximal to the contact surface or immunological synapse (7, 33, 34, 48-50, 54). A variety of different signaling molecules are also involved, including calcium (23), phosphatidylinositol-3,4,5-triphosphate (13, 17), and activation of the ERK MAP kinase (6, 42, 56). Recently, the recruitment of activated ERK to the immunological synapse (IS) has been shown to be a feature of successful killing of a target by cytotoxic T lymphocytes (58).How active ERK is recruited to the synapse is not known. Since KSR1 is known to be recruited to the plasma membrane by Ras activation (24), and since the immunological synapse is one of the major sites of Ras activation (26, 41), it seemed plausible to test the hypothesis that KSR1 recruitment to the plasma membrane functions to recruit ERK to the immunological synapse and facilitate its activation. We found that KSR1 was recruited to the immunological synapse and that KSR1 appeared to be required for the localization of active ERK at the contact site. As KSR1-deficient cells exhibit a defect in killing, this suggests that KSR1 recruitment to the synapse may be important in the cytolytic killing of target cells.  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