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 578 毫秒
1.
Melanoma cells express the chemokine receptor CXCR4 that confers high invasiveness upon binding to its ligand CXCL12. Melanoma cells at initial stages of the disease show reduction or loss of E-cadherin expression, but recovery of its expression is frequently found at advanced phases. We overexpressed E-cadherin in the highly invasive BRO lung metastatic cell melanoma cell line to investigate whether it could influence CXCL12-promoted cell invasion. Overexpression of E-cadherin led to defective invasion of melanoma cells across Matrigel and type I collagen in response to CXCL12. A decrease in individual cell migration directionality toward the chemokine and reduced adhesion accounted for the impaired invasion. A p190RhoGAP-dependent inhibition of RhoA activation was responsible for the impairment in chemokine-stimulated E-cadherin melanoma transfectant invasion. Furthermore, we show that p190RhoGAP and p120ctn associated predominantly on the plasma membrane of cells overexpressing E-cadherin, and that E-cadherin-bound p120ctn contributed to RhoA inactivation by favoring p190RhoGAP-RhoA association. These results suggest that melanoma cells at advanced stages of the disease could have reduced metastatic potency in response to chemotactic stimuli compared with cells lacking E-cadherin, and the results indicate that p190RhoGAP is a central molecule controlling melanoma cell invasion.Cadherins are a family of Ca2+-dependent adhesion molecules that mediate cell-cell contacts and are expressed in most solid tissues providing a tight control of morphogenesis (1, 2). Classical cadherins, such as epithelial (E) cadherin, are found in adherens junctions, forming core protein complexes with β-catenin, α-catenin, and p120 catenin (p120ctn). Both β-catenin and p120ctn directly interact with E-cadherin, whereas α-catenin associates with the complex through its binding to β-catenin, providing a link with the actin cytoskeleton (1, 2). E-cadherin is frequently lost or down-regulated in many human tumors, coincident with morphological epithelial to mesenchymal transition and acquisition of invasiveness (3-6).Although melanoma only accounts for 5% of skin cancers, when metastasis starts, it is responsible for 80% of deaths from skin cancers (7). Melanocytes express E-cadherin (8-10), but melanoma cells at early radial growth phase show a large reduction in the expression of this cadherin, and surprisingly, expression has been reported to be partially recovered by vertical growth phase and metastatic melanoma cells (9, 11, 12).Trafficking of cancer cells from primary tumor sites to intravasation into blood circulation and later to extravasation to colonize distant organs requires tightly regulated directional cues and cell migration and invasion that are mediated by chemokines, growth factors, and adhesion molecules (13). Solid tumor cells express chemokine receptors that provide guidance of these cells to organs where their chemokine ligands are expressed, constituting a homing model resembling the one used by immune cells to exert their immune surveillance functions (14). Most solid cancer cells express CXCR4, a receptor for the chemokine CXCL12 (also called SDF-1), which is expressed in lungs, bone marrow, and liver (15). Expression of CXCR4 in human melanoma has been detected in the vertical growth phase and on regional lymph nodes, which correlated with poor prognosis and increased mortality (16, 17). Previous in vivo experiments have provided evidence supporting a crucial role for CXCR4 in the metastasis of melanoma cells (18).Rho GTPases control the dynamics of the actin cytoskeleton during cell migration (19, 20). The activity of Rho GTPases is tightly regulated by guanine-nucleotide exchange factors (GEFs),4 which stimulate exchange of bound GDP by GTP, and inhibited by GTPase-activating proteins (GAPs), which promote GTP hydrolysis (21, 22), whereas guanine nucleotide dissociation inhibitors (GDIs) appear to mediate blocking of spontaneous activation (23). Therefore, cell migration is finely regulated by the balance between GEF, GAP, and GDI activities on Rho GTPases. Involvement of Rho GTPases in cancer is well documented (reviewed in Ref. 24), providing control of both cell migration and growth. RhoA and RhoC are highly expressed in colon, breast, and lung carcinoma (25, 26), whereas overexpression of RhoC in melanoma leads to enhancement of cell metastasis (27). CXCL12 activates both RhoA and Rac1 in melanoma cells, and both GTPases play key roles during invasion toward this chemokine (28, 29).Given the importance of the CXCL12-CXCR4 axis in melanoma cell invasion and metastasis, in this study we have addressed the question of whether changes in E-cadherin expression on melanoma cells might affect cell invasiveness. We show here that overexpression of E-cadherin leads to impaired melanoma cell invasion to CXCL12, and we provide mechanistic characterization accounting for the decrease in invasion.  相似文献   

2.
3.
4.
5.
6.
7.
The actin cytoskeleton has an important role in the organization and function of the immune synapse during antigen recognition. Dynamic rearrangement of the actin cytoskeleton in response to T cell receptor (TCR) triggering requires the coordinated activation of Rho family GTPases that cycle between active and inactive conformations. This is controlled by GTPase-activating proteins (GAP), which regulate inactivation of Rho GTPases, and guanine exchange factors, which mediate their activation. Whereas much attention has centered on guanine exchange factors for Rho GTPases in T cell activation, the identity and functional roles of the GAP in this process are largely unknown. We previously reported β2-chimaerin as a diacylglycerol-regulated Rac-GAP that is expressed in T cells. We now demonstrate Lck-dependent phosphorylation of β2-chimaerin in response to TCR triggering. We identify Tyr-153 as the Lck-dependent phosphorylation residue and show that its phosphorylation negatively regulates membrane stabilization of β2-chimaerin, decreasing its GAP activity to Rac. This study establishes the existence of TCR-dependent regulation of β2-chimaerin and identifies a novel mechanism for its inactivation.T cell activation requires presentation of an antigen by antigen-presenting cells (APC)2 to the T cell receptor (TCR); this event involves the reorganization of several scaffolds and signaling proteins, leading to formation of the immunological synapse (IS) (1). Correct protein redistribution during synapse formation is critical for an efficient T cell response, and it is largely regulated by actin polymerization at the T cell/APC contact site as a result of TCR-regulated Rac-dependent signals (2, 3). Like other Rho GTPases, Rac cycles between a GTP-bound active state and a GDP-bound inactive state. This continuous recycling is regulated by the concerted action of two proteins as follows: GEF, which activates Rac by mediating GDP/GTP exchange (4), and GAP, which induces Rac inactivation by accelerating intrinsic Rac GTPase activity, converting GTP to GDP (5).Vav-1 is the best studied GEF for Rac, and it has critical roles in T cell-dependent functions (6). In naive, unstimulated T cells, Vav-1 is in an inactive state through autoinhibition, as the GEF domain is blocked by the N-terminal region (7). This autoinhibition is relieved by TCR-mediated tyrosine phosphorylation (8, 9). Thymocytes from Vav-1-deficient mice have a developmental block, and their mature T cells show severe defects in IS formation, as well as reduced Ca2+ influx, IL-2 production, T cell proliferation, and cytotoxic activity (1013). Although several studies have shown a key role for Vav-1, the mechanisms that govern Rac inactivation downstream of the TCR remain elusive.The chimaerins are a family of Rho-GAP, with specific activity for Rac. In addition to their catalytic domain, they have an N-terminal SH2 domain and a C1 domain required for interaction with the lipid messenger diacylglycerol (DAG) and with phorbol esters (14). There are two mammalian chimaerin genes (CHN1 and CHN2), which encode the full-lengthα2-(ARHGAP2) and β2-chimaerins (ARHGAP3), and at least one splice variant each (α1 and β1) that lack the SH2 domain. The α-chimaerins are expressed abundantly in brain and are linked to neuritogenesis and axon guidance (1520). β2-Chimaerin expression is ubiquitous (21) and is involved in EGF-dependent Rac regulation (22, 23). Experiments in T cells showed that β2-chimaerin participates in chemokine-dependent regulation of T cell migration and adhesion (24). A very recent study implicates chimaerins in the modulation of Rac activity during T cell synapse formation, suggesting that this protein family contributes to DAG-mediated regulation of cytoskeletal remodeling during T cell activation (25).Determination of the β2-chimaerin crystal structure provided important clues regarding its mechanism of action. In the absence of stimulation, the protein is in an inactive state in which the N-terminal domain maintains a “closed” conformation, blocking Rac binding and concealing the C1 domain (26). These structural data were fully supported by experiments in live T lymphocytes showing that phorbol myristate acetate (PMA)-dependent translocation of β2-chimaerin was less effective than that of its isolated C1 domain (24). These data not only confirmed the lack of accessibility of the β2-chimaerin C1 domain but also suggested that there are negative regulatory mechanisms that promote β2-chimaerin release from the membrane.DAG-dependent signaling is critical for the modulation of T cell functions, by virtue of its ability to bind and regulate C1 domain-containing proteins such as protein kinase Cθ, protein kinase D, and RasGRP1 (27). An important issue is to determine how the different DAG-binding proteins discriminate between distinct DAG pools, and how DAG activates certain C1-containing proteins and not others. Some mechanisms that allow discrimination between DAG receptors include the distinct affinity of C1 domains for different DAG pools, association of C1 domain-containing proteins to specific scaffolds, and/or structural determinants in these proteins that limit C1 domain accessibility to membrane DAG (2830).To explore the events that contribute to the specific regulation of β2-chimaerin, we studied β2-chimaerin phosphorylation in the context of TCR stimulation. We show that β2-chimaerin is phosphorylated in tyrosine residues after TCR stimulation, and we identify Lck as the Tyr kinase responsible for this phosphorylation. Generation of point mutants identified Tyr-153, at the hinge of the SH2 and C1 domains, as the main tyrosine residue phosphorylated in response to TCR stimulation. Cells expressing a β2-chimaerin mutant defective for Tyr-153 phosphorylation show anomalies in TCR clustering, conjugate formation, NF-AT activation, and IL-2 production that correlate with elevated Rac-GAP activity in this mutant. Subcellular localization analysis of the β2-chimaerin mutants reveals that impairment of β2-chimaerin phosphorylation at Tyr-153 promotes C1-mediated β2-chimaerin stabilization at the plasma membrane, providing a mechanistic explanation for its higher Rac-GAP activity. In summary, our results demonstrate for the first time that tyrosine kinase-mediated negative regulation of β2-chimaerin is elicited by physiological stimulation in T lymphocytes, and suggest that TCR stimulation provides both positive and negative signals for β2-chimaerin activation.  相似文献   

8.
It is known that platelet-activating factor (PAF) induces severe endothelial barrier leakiness, but the signaling mechanisms remain unclear. Here, using a wide range of biochemical and morphological approaches applied in both mouse models and cultured endothelial cells, we addressed the mechanisms of PAF-induced disruption of interendothelial junctions (IEJs) and of increased endothelial permeability. The formation of interendothelial gaps filled with filopodia and lamellipodia is the cellular event responsible for the disruption of endothelial barrier. We observed that PAF ligation of its receptor induced the activation of the Rho GTPase Rac1. Following PAF exposure, both Rac1 and its guanine nucleotide exchange factor Tiam1 were found associated with a membrane fraction from which they co-immunoprecipitated with PAF receptor. In the same time frame with Tiam1-Rac1 translocation, the junctional proteins ZO-1 and VE-cadherin were relocated from the IEJs, and formation of numerous interendothelial gaps was recorded. Notably, the response was independent of myosin light chain phosphorylation and thus distinct from other mediators, such as histamine and thrombin. The changes in actin status are driven by the PAF-induced localized actin polymerization as a consequence of Rac1 translocation and activation. Tiam1 was required for the activation of Rac1, actin polymerization, relocation of junctional associated proteins, and disruption of IEJs. Thus, PAF-induced IEJ disruption and increased endothelial permeability requires the activation of a Tiam1-Rac1 signaling module, suggesting a novel therapeutic target against increased vascular permeability associated with inflammatory diseases.The endothelial barrier is made up of endothelial cells (ECs)4 connected to each other by interendothelial junctions (IEJs) consisting of protein complexes organized as tight junctions (TJs) and adherens junctions (AJs). In addition, the focal adhesion complex located at the basal plasma membrane enables firm contact of ECs with the underlying basement membrane and also contributes to the barrier function (1-3). The glycocalyx, the endothelial monolayer, and the basement membrane all together constitute the vascular barrier.The structural integrity of the ECs along with their proper functionality are the two most important factors controlling the tightness of the endothelial barrier. Changes affecting these factors cause loss of barrier restrictiveness and leakiness. Therefore, defining and understanding the cellular and molecular mechanisms controlling these processes is of paramount importance. Increased width of IEJs in response to permeability-increasing mediators (4) regulates the magnitude of transendothelial exchange of fluid and solutes. Disruption of IEJs and the resultant barrier leakiness contribute to the genesis of diverse pathological conditions, such as inflammation (5), metastasis (6, 7), and uncontrolled angiogenesis (8, 9).Accumulated evidence demonstrated that IEJs changes are responsible for increased or decreased vascular permeability, and the generally accepted mechanism responsible for them was the myosin light chain (MLC)-mediated contraction of ECs (5, 10). However, published evidence showed that an increase in vascular permeability could be obtained without a direct involvement of any contractile mechanism (11-16).The main component of the vascular barrier, the ECs, has more than 10% of their total protein represented by actin (17), which under physiological salt concentrations subsists as monomers (G-actin) and assembled into filaments (F-actin). A large number of actin-interacting proteins may modulate the assembly, disassembly, and organization of G-actin and of actin filaments within a given cell type. Similar to the complexity of actin-interacting proteins found in other cell types, the ECs utilize their actin binding proteins to stabilize the endothelial monolayer in order to efficiently function as a selective barrier (11). In undisturbed ECs, the actin microfilaments are organized as different networks with distinctive functional and morphological characteristics: the peripheral filaments also known as peripheral dense band (PDB), the cytoplasmic fibers identified as stress fibers (SF), and the actin from the membrane cytoskeleton (18). The peripheral web, localized immediately under the membrane, is associated with (i) the luminal plasmalemma (on the apical side), (ii) the IEJ complexes on the lateral surfaces, and (iii) the focal adhesion complexes on the abluminal side (the basal part) of polarized ECs. The SF reside inside the endothelial cytoplasm and are believed to be directly connected with the plasmalemma proper on the luminal as well as on the abluminal side of the cell. As described, the endothelial actin cytoskeleton (specifically the SF) seems to be a stable structure helping the cells to remain flat under flow (19). It is also established that the actin fibers participate in correct localization of different junctional complexes while keeping them in place (20). However, it was suggested that the dynamic equilibrium between F- and G-actin might modulate the tightness of endothelial barrier in response to different challenges (13).Mediators effective at nanomolar concentrations or less that disrupt the endothelial barrier and increase vascular permeability include C2 toxin of Clostridium botulinum, vascular permeability factor, better known as vascular endothelial growth factor, and PAF (21). C2 toxin increases endothelial permeability by ribosylating monomeric G-actin at Arg-177 (22). This results in the impairment of actin polymerization (23), followed by rounding of ECs (16) and the disruption of junctional integrity. Vascular permeability factor was shown to open IEJs by redistribution of junctional proteins (24, 25) and by interfering with the equilibrium of actin pools (26). PAF (1-O-alkyl-2-acetyl-sn-glycero-3-phosphocoline), a naturally synthesized phospholipid is active at 10-10 m or less (27). PAF is synthesized by and acts on a variety of cell types, including platelets (28), neutrophils (29), monocytes (30), and ECs (31). PAF-mediated activation of ECs induced cell migration (32), angiogenesis (7), and vascular hyperpermeability (33) secondary to disassembly of IEJs (34). The effects of PAF on the endothelium are initiated through a G protein-coupled receptor (PAF-R) localized at the plasmalemma, in a large endosomal compartment inside the cell (34), and also in the nuclear membrane (35). In ECs, PAF-R was shown to signal through Gαq and downstream activation of phospholipase C isozymes (PLCβ3 and PLCγ1), and via cSrc (32, 36). Studies have shown that PAF challenge induced endothelial actin cytoskeletal rearrangement (37) and marked vascular leakiness (38); however, the signaling pathways have not been elucidated.Therefore, in the present study, we carried out a systematic analysis of PAF-induced morphological and biochemical changes of endothelial barrier in vivo and in cultured ECs. We found that the opening of endothelial barrier and the increased vascular leakiness induced by PAF are the result of a shift in actin pools without involvement of EC contraction, followed by a redistribution of tight junctional associated protein ZO-1 and adherens junctional protein VE-cadherin.  相似文献   

9.
Mitochondrial dysregulation is strongly implicated in Parkinson disease. Mutations in PTEN-induced kinase 1 (PINK1) are associated with familial parkinsonism and neuropsychiatric disorders. Although overexpressed PINK1 is neuroprotective, less is known about neuronal responses to loss of PINK1 function. We found that stable knockdown of PINK1 induced mitochondrial fragmentation and autophagy in SH-SY5Y cells, which was reversed by the reintroduction of an RNA interference (RNAi)-resistant plasmid for PINK1. Moreover, stable or transient overexpression of wild-type PINK1 increased mitochondrial interconnectivity and suppressed toxin-induced autophagy/mitophagy. Mitochondrial oxidant production played an essential role in triggering mitochondrial fragmentation and autophagy in PINK1 shRNA lines. Autophagy/mitophagy served a protective role in limiting cell death, and overexpressing Parkin further enhanced this protective mitophagic response. The dominant negative Drp1 mutant inhibited both fission and mitophagy in PINK1-deficient cells. Interestingly, RNAi knockdown of autophagy proteins Atg7 and LC3/Atg8 also decreased mitochondrial fragmentation without affecting oxidative stress, suggesting active involvement of autophagy in morphologic remodeling of mitochondria for clearance. To summarize, loss of PINK1 function elicits oxidative stress and mitochondrial turnover coordinated by the autophagic and fission/fusion machineries. Furthermore, PINK1 and Parkin may cooperate through different mechanisms to maintain mitochondrial homeostasis.Parkinson disease is an age-related neurodegenerative disease that affects ∼1% of the population worldwide. The causes of sporadic cases are unknown, although mitochondrial or oxidative toxins such as 1-methyl-4-phenylpyridinium, 6-hydroxydopamine (6-OHDA),3 and rotenone reproduce features of the disease in animal and cell culture models (1). Abnormalities in mitochondrial respiration and increased oxidative stress are observed in cells and tissues from parkinsonian patients (2, 3), which also exhibit increased mitochondrial autophagy (4). Furthermore, mutations in parkinsonian genes affect oxidative stress response pathways and mitochondrial homeostasis (5). Thus, disruption of mitochondrial homeostasis represents a major factor implicated in the pathogenesis of sporadic and inherited parkinsonian disorders (PD).The PARK6 locus involved in autosomal recessive and early-onset PD encodes for PTEN-induced kinase 1 (PINK1) (6, 7). PINK1 is a cytosolic and mitochondrially localized 581-amino acid serine/threonine kinase that possesses an N-terminal mitochondrial targeting sequence (6, 8). The primary sequence also includes a putative transmembrane domain important for orientation of the PINK1 domain (8), a conserved kinase domain homologous to calcium calmodulin kinases, and a C-terminal domain that regulates autophosphorylation activity (9, 10). Overexpression of wild-type PINK1, but not its PD-associated mutants, protects against several toxic insults in neuronal cells (6, 11, 12). Mitochondrial targeting is necessary for some (13) but not all of the neuroprotective effects of PINK1 (14), implicating involvement of cytoplasmic targets that modulate mitochondrial pathobiology (8). PINK1 catalytic activity is necessary for its neuroprotective role, because a kinase-deficient K219M substitution in the ATP binding pocket of PINK1 abrogates its ability to protect neurons (14). Although PINK1 mutations do not seem to impair mitochondrial targeting, PD-associated mutations differentially destabilize the protein, resulting in loss of neuroprotective activities (13, 15).Recent studies indicate that PINK1 and Parkin interact genetically (3, 16-18) to prevent oxidative stress (19, 20) and regulate mitochondrial morphology (21). Primary cells derived from PINK1 mutant patients exhibit mitochondrial fragmentation with disorganized cristae, recapitulated by RNA interference studies in HeLa cells (3).Mitochondria are degraded by macroautophagy, a process involving sequestration of cytoplasmic cargo into membranous autophagic vacuoles (AVs) for delivery to lysosomes (22, 23). Interestingly, mitochondrial fission accompanies autophagic neurodegeneration elicited by the PD neurotoxin 6-OHDA (24, 25). Moreover, mitochondrial fragmentation and increased autophagy are observed in neurodegenerative diseases including Alzheimer and Parkinson diseases (4, 26-28). Although inclusion of mitochondria in autophagosomes was once believed to be a random process, as observed during starvation, studies involving hypoxia, mitochondrial damage, apoptotic stimuli, or limiting amounts of aerobic substrates in facultative anaerobes support the concept of selective mitochondrial autophagy (mitophagy) (29, 30). In particular, mitochondrially localized kinases may play an important role in models involving oxidative mitochondrial injury (25, 31, 32).Autophagy is involved in the clearance of protein aggregates (33-35) and normal regulation of axonal-synaptic morphology (36). Chronic disruption of lysosomal function results in accumulation of subtly impaired mitochondria with decreased calcium buffering capacity (37), implicating an important role for autophagy in mitochondrial homeostasis (37, 38). Recently, Parkin, which complements the effects of PINK1 deficiency on mitochondrial morphology (3), was found to promote autophagy of depolarized mitochondria (39). Conversely, Beclin 1-independent autophagy/mitophagy contributes to cell death elicited by the PD toxins 1-methyl-4-phenylpyridinium and 6-OHDA (25, 28, 31, 32), causing neurite retraction in cells expressing a PD-linked mutation in leucine-rich repeat kinase 2 (40). Whereas properly regulated autophagy plays a homeostatic and neuroprotective role, excessive or incomplete autophagy creates a condition of “autophagic stress” that can contribute to neurodegeneration (28).As mitochondrial fragmentation (3) and increased mitochondrial autophagy (4) have been described in human cells or tissues of PD patients, we investigated whether or not the engineered loss of PINK1 function could recapitulate these observations in human neuronal cells (SH-SY5Y). Stable knockdown of endogenous PINK1 gave rise to mitochondrial fragmentation and increased autophagy and mitophagy, whereas stable or transient overexpression of PINK1 had the opposite effect. Autophagy/mitophagy was dependent upon increased mitochondrial oxidant production and activation of fission. The data indicate that PINK1 is important for the maintenance of mitochondrial networks, suggesting that coordinated regulation of mitochondrial dynamics and autophagy limits cell death associated with loss of PINK1 function.  相似文献   

10.
These studies explore the connections between simvastatin, Rac1, and AMP-activated protein kinase (AMPK) pathways in cultured vascular endothelial cells and in arterial preparations isolated from statin-treated mice. In addition to their prominent effects on lipoprotein metabolism, statins can regulate the small GTPase Rac1, and may also affect the phosphorylation of the ubiquitous AMPK. We explored pathways of statin-modulated Rac1 and AMPK activation both in arterial preparations from statin-treated mice as well as in cultured endothelial cells. We treated adult mice with simvastatin daily for 2 weeks and then harvested and analyzed arterial preparations. Simvastatin treatment of mice led to a significant increase in AMPK and LKB1 phosphorylation and to a decrease in protein kinase A activity relative to control animals, associated with a marked increase in Rac1 activation. Exposure of bovine aortic endothelial cells to simvastatin for 24 h strikingly increased GTP-bound Rac1 and led to increased phosphorylation of AMPK as well as the AMPK kinase LKB1. These responses to simvastatin were blocked by mevalonate or geranylgeranyl pyrophosphate but not by farnesyl pyrophosphate. Small interfering RNA (siRNA)-mediated knockdown of AMPK abrogated simvastatin-induced Rac1 activation and LKB1 phosphorylation. Importantly, siRNA-mediated knockdown of the key AMPK kinase, calcium/calmodulin-dependent protein kinase kinase β, completely blocked simvastatin-induced endothelial cell migration and also abrogated statin-promoted phosphorylation of AMPK and LKB1, as did pharmacological inhibition with the specific calcium/calmodulin-dependent protein kinase β inhibitor STO-609. Moreover, siRNA-mediated knockdown of Rac1 completely blocked simvastatin-induced LKB1 phosphorylation, but without affecting simvastatin-induced AMPK phosphorylation. These findings establish a key role for simvastatin in activation of a novel Rac1-dependent signaling pathway in the vascular wall.HMG-CoA2 reductase inhibitors, commonly known as statins, are widely prescribed for the prevention and treatment of hypercholesterolemia and cardiovascular diseases (1, 2). The salutary clinical effects of these drugs derive in part from their effects on the levels of serum lipoproteins, yet other statin responses appear to be mediated by alterations in vascular function involving the endothelial isoform of nitric-oxide synthase (3) and related signaling pathways. Inhibition of HMG-CoA reductase suppresses the cellular levels of its enzymatic product mevalonate, thereby attenuating formation both of cholesterol as well as the synthesis of distinct isoprenoid compounds such as farnesyl pyrophosphate (Fpp) and geranylgeranyl pyrophosphate (GGpp). Many key signaling proteins are covalently modified by these isoprenoids, which are the products of a metabolic pathway that diverges from the pathway that leads to cholesterol synthesis downstream of HMG-CoA reductase. These isoprenoid compounds can provide lipophilic anchors that facilitate membrane targeting and modulate protein-protein interactions of many key signaling proteins. One such iso-prenylated signaling protein is the GTP-binding cytoskeletonassociated protein Rac1, a member of the Rho GTPase small G protein family that undergoes geranylgeranylation at its C terminus. Statins also affect post-translational modification of another small GTPase, RhoA, that, like Rac1, is a geranylgeranylated protein that is an important determinant of vascular signaling (48). Rac1 has particularly important roles in vascular endothelial cells, where this cytoskeleton regulatory protein modulates activity of the endothelial isoform of nitric-oxide synthase (eNOS), a key determinant of vascular homeostasis (9). Rac1 activation in endothelial cells is influenced by the AMP-activated protein kinase (AMPK) (6), which itself is phosphorylated by the protein kinase LKB1 and by the calcium-calmodulin-dependent protein kinase β (CaMKKβ) (see review (10)). In recent years, numerous reports have described effects of statins on variety of these signaling proteins in different experimental systems (1114).Statins have been shown to promote the phosphorylation of AMPK (13), a heterotrimeric enzyme involved in the modulation of cellular energy pathways that has also been implicated in eNOS regulation (3, 1517). AMPK was originally discovered and characterized as a cellular “energy sensor” that can be activated by increases in the intracellular AMP:ATP ratio (18). However, in recent years, it has become clear that AMPK is also regulated through AMP-independent pathways involving enzyme phosphorylation on threonine 172 of the enzyme''s α subunit, leading to marked enzyme activation (19). Protein kinases that phosphorylate AMPK include the tumor suppressor LKB1 and the calcium/calmodulin-dependent kinase CaMKKβ. LKB1 itself is a phosphoprotein. The pathways that regulate LKB1 are incompletely understood, and a variety of upstream protein kinases have been implicated in LKB1 regulation (see review (20)). CaMKKβ is principally regulated by calcium binding, but this kinase may also be phosphorylated by the cAMP-dependent protein kinase PKA (21, 22). Another substrate for PKA in vascular cells is the actin-binding phosphoprotein VASP (23, 24); the phosphorylation state of VASP at its PKA site can serve as a surrogate marker for the activity of cAMP-dependent signaling pathways in the vascular wall (25). CaMKKβ has been shown to be involved in AMPK regulation in endothelial cells in response to receptor tyrosine kinase activation and via G protein-coupled receptor pathways (6). Activated AMPK directly phosphorylates eNOS, and this kinase thereby appears be an important determinant of NO-dependent signaling in endothelial cells. However, much remains to be learned about the molecular mechanisms whereby statins enhance AMPK activation.In cultured cells, statins have been shown to inhibit the geranylgeranylation of Rac1, associated with an increase in Rac1 GTP binding and activation (26). The activation of Rac1 is a key step in eNOS activation: siRNA-mediated Rac1 “knockdown” in endothelial cells markedly suppresses receptor signaling to eNOS (5, 7). siRNA-mediated AMPK knockdown suppresses Rac1 activation, again leading to the attenuation of receptor-dependent activation of eNOS (6). The relationships among these various statin-modulated signaling pathways are incompletely characterized. The present studies identify CaMKKβ and LKB1 as critical determinants of simvastatin-dependent activation of AMPK- and Rac1-modulated signaling and reveal that Rac1 in turn regulates LKB1 phosphorylation.  相似文献   

11.
Mutations in SHP-2 phosphatase (PTPN11) that cause hyperactivation of its catalytic activity have been identified in Noonan syndrome and various childhood leukemias. Recent studies suggest that the gain-of-function (GOF) mutations of SHP-2 play a causal role in the pathogenesis of these diseases. However, the molecular mechanisms by which GOF mutations of SHP-2 induce these phenotypes are not fully understood. Here, we show that GOF mutations in SHP-2, such as E76K and D61G, drastically increase spreading and migration of various cell types, including hematopoietic cells, endothelial cells, and fibroblasts. More importantly, in vivo angiogenesis in SHP-2 D61G knock-in mice is also enhanced. Mechanistic studies suggest that the increased cell migration is attributed to the enhanced β1 integrin outside-in signaling. In response to β1 integrin cross-linking or fibronectin stimulation, activation of ERK and Akt kinases is greatly increased by SHP-2 GOF mutations. Also, integrin-induced activation of RhoA and Rac1 GTPases is elevated. Interestingly, mutant cells with the SHP-2 GOF mutation (D61G) are more sensitive than wild-type cells to the suppression of cell motility by inhibition of these pathways. Collectively, these studies reaffirm the positive role of SHP-2 phosphatase in cell motility and suggest a new mechanism by which SHP-2 GOF mutations contribute to diseases.SHP-2, a multifunctional SH2 domain-containing protein-tyrosine phosphatase implicated in diverse cell signaling processes (13), plays a critical role in cellular function. Homozygous deletion of Exon 2 (4) or Exon 3 (5) of the SHP-2 gene (PTPN11) in mice leads to early embryonic lethality prior to and at midgestation, respectively. SHP-2 null mutant mice die much earlier, at peri-implantation (4). Exon 3 deletion mutation of SHP-2 blocks hematopoietic potential of embryonic stem cells both in vitro and in vivo (68), whereas SHP-2 null mutation causes inner cell mass death and diminished trophoblast stem cell survival (4). Recent studies on SHP-2 conditional knock-out or tissue-specific knock-out mice have further revealed an array of important functions of this phosphatase in various physiological processes (912). The phenotypes demonstrated by loss of SHP-2 function are apparently attributed to the role of SHP-2 in the cell signaling pathways induced by growth factors/cytokines. SHP-2 generally promotes signal transmission in growth factor/cytokine signaling in both catalytic-dependent and -independent fashion (13). The positive role of SHP-2 in the intracellular signaling processes, in particular, the ERK3 and PI3K/Akt kinase pathways, has been well established, although the underlying mechanism remains elusive, in particular, the signaling function of the catalytic activity of SHP-2 in these pathways is poorly understood.In addition to the role of SHP-2 in cell proliferation and differentiation, the phenotypes induced by loss of SHP-2 function may be associated with its role in cell migration. Indeed, dominant negative SHP-2 disrupts Xenopus gastrulation, causing tail truncations (13, 14). Targeted Exon 3 deletion mutation in SHP-2 results in decreased cell spreading, migration (15, 16), and impaired limb development in the chimeric mice (7). The role of SHP-2 in cell adhesion and migration has also been demonstrated by catalytically inactive mutant SHP-2-overexpressing cells (1720). The molecular mechanisms by which SHP-2 regulates these cellular processes, however, have not been well defined. For example, the role of SHP-2 in the activation of the Rho family small GTPases that is critical for cell motility is still controversial. Both positive (19, 21, 22) and negative roles (18, 23) for SHP-2 in this context have been reported. Part of the reason for this discrepancy might be due to the difference in the cell models used. Catalytically inactive mutant SHP-2 was often used to determine the role of SHP-2 in cell signaling. In the catalytically inactive mutant SHP-2-overexpressing cells, the catalytic activity of endogenous SHP-2 is inhibited. However, as SHP-2 also functions independent of its catalytic activity, overexpression of catalytically deficient SHP-2 may also increase its scaffolding function, generating complex effects.The critical role of SHP-2 in cellular function is further underscored by the identification of SHP-2 mutations in human diseases. Genetic lesions in PTPN11 that cause hyperactivation of SHP-2 catalytic activity have been identified in the developmental disorder Noonan syndrome (24) and various childhood leukemias, including juvenile myelomonocytic leukemia (JMML), B cell acute lymphoblastic leukemia, and acute myeloid leukemia (25, 26). In addition, activating mutations in SHP-2 have been identified in sporadic solid tumors (27). The SHP-2 mutations appear to play a causal role in the development of these diseases as SHP-2 mutations and other JMML-associated Ras or Neurofibromatosis 1 mutations are mutually exclusive in the patients (2427). Moreover, single SHP-2 gain-of-function (GOF) mutations are sufficient to induce Noonan syndrome, cytokine hypersensitivity in hematopoietic progenitor cells, and JMML-like myeloproliferative disease in mice (2832). Gain-of-function cell models derived from the newly available SHP-2 GOF mutation (D61G) knock-in mice (28) now provide us with a good opportunity to clarify the role of SHP-2 in cell motility. Unlike the dominant negative approach in which overexpression of mutant forms of SHP-2 generates complex effects, the SHP-2 D61G knock-in model eliminates this possibility as the mutant SHP-2 is expressed at the physiological level (28). Additionally, defining signaling functions of GOF mutant SHP-2 in cell movement can also help elucidate the molecular mechanisms by which SHP-2 mutations contribute to the relevant diseases.  相似文献   

12.
As obligate intracellular parasites, viruses exploit diverse cellular signaling machineries, including the mitogen-activated protein-kinase pathway, during their infections. We have demonstrated previously that the open reading frame 45 (ORF45) of Kaposi sarcoma-associated herpesvirus interacts with p90 ribosomal S6 kinases (RSKs) and strongly stimulates their kinase activities (Kuang, E., Tang, Q., Maul, G. G., and Zhu, F. (2008) J. Virol. 82 ,1838 -1850). Here, we define the mechanism by which ORF45 activates RSKs. We demonstrated that binding of ORF45 to RSK increases the association of extracellular signal-regulated kinase (ERK) with RSK, such that ORF45, RSK, and ERK formed high molecular mass protein complexes. We further demonstrated that the complexes shielded active pERK and pRSK from dephosphorylation. As a result, the complex-associated RSK and ERK were activated and sustained at high levels. Finally, we provide evidence that this mechanism contributes to the sustained activation of ERK and RSK in Kaposi sarcoma-associated herpesvirus lytic replication.The extracellular signal-regulated kinase (ERK)2 mitogen-activated protein kinase (MAPK) signaling pathway has been implicated in diverse cellular physiological processes including proliferation, survival, growth, differentiation, and motility (1-4) and is also exploited by a variety of viruses such as Kaposi sarcoma-associated herpesvirus (KSHV), human cytomegalovirus, human immunodeficiency virus, respiratory syncytial virus, hepatitis B virus, coxsackie, vaccinia, coronavirus, and influenza virus (5-17). The MAPK kinases relay the extracellular signaling through sequential phosphorylation to an array of cytoplasmic and nuclear substrates to elicit specific responses (1, 2, 18). Phosphorylation of MAPK is reversible. The kinetics of deactivation or duration of signaling dictates diverse biological outcomes (19, 20). For example, sustained but not transient activation of ERK signaling induces the differentiation of PC12 cells into sympathetic-like neurons and transformation of NIH3T3 cells (20-22). During viral infection, a unique biphasic ERK activation has been observed for some viruses (an early transient activation triggered by viral binding or entry and a late sustained activation correlated with viral gene expression), but the responsible viral factors and underlying mechanism for the sustained ERK activation remain largely unknown (5, 8, 13, 23).The p90 ribosomal S6 kinases (RSKs) are a family of serine/threonine kinases that lie at the terminus of the ERK pathway (1, 24-26). In mammals, four isoforms are known, RSK1 to RSK4. Each one has two catalytically functional kinase domains, the N-terminal kinase domain (NTKD) and C-terminal kinase domain (CTKD) as well as a linker region between the two. The NTKD is responsible for phosphorylation of exogenous substrates, and the CTKD and linker region regulate RSK activation (1, 24, 25). In quiescent cells ERK binds to the docking site in the C terminus of RSK (27-29). Upon mitogen stimulation, ERK is activated by its upstream MAPK/ERK kinase (MEK). The active ERK phosphorylates Thr-359/Ser-363 of RSK in the linker region (amino acid numbers refer to human RSK1) and Thr-573 in the CTKD activation loop. The activated CTKD then phosphorylates Ser-380 in the linker region, creating a docking site for 3-phosphoinositide-dependent protein kinase-1. The 3-phosphoinositide-dependent protein kinase-1 phosphorylates Ser-221 of RSK in the activation loop and activates the NTKD. The activated NTKD autophosphorylates the serine residue near the ERK docking site, causing a transient dissociation of active ERK from RSK (25, 26, 28). The stimulation of quiescent cells by a mitogen such as epidermal growth factor or a phorbol ester such as 12-O-tetradecanoylphorbol-13-acetate (TPA) usually results in a transient RSK activation that lasts less than 30 min. RSKs have been implicated in regulating cell survival, growth, and proliferation. Mutation or aberrant expression of RSK has been implicated in several human diseases including Coffin-Lowry syndrome and prostate and breast cancers (1, 24, 25, 30-32).KSHV is a human DNA tumor virus etiologically linked to Kaposi sarcoma, primary effusion lymphoma, and a subset of multicentric Castleman disease (33, 34). Infection and reactivation of KSHV activate multiple MAPK pathways (6, 12, 35). Noticeably, the ERK/RSK activation is sustained late during KSHV primary infection and reactivation from latency (5, 6, 12, 23), but the mechanism of the sustained ERK/RSK activation is unclear. Recently, we demonstrated that ORF45, an immediate early and also virion tegument protein of KSHV, interacts with RSK1 and RSK2 and strongly stimulates their kinase activities (23). We also demonstrated that the activation of RSK plays an essential role in KSHV lytic replication (23). In the present study we determined the mechanism of ORF45-induced sustained ERK/RSK activation. We found that ORF45 increases the association of RSK with ERK and protects them from dephosphorylation, causing sustained activation of both ERK and RSK.  相似文献   

13.
In archaea and eukarya, box C/D ribonucleoprotein (RNP) complexes are responsible for 2′-O-methylation of tRNAs and rRNAs. The archaeal box C/D small RNP complex requires a small RNA component (sRNA) possessing Watson-Crick complementarity to the target RNA along with three proteins: L7Ae, Nop5p, and fibrillarin. Transfer of a methyl group from S-adenosylmethionine to the target RNA is performed by fibrillarin, which by itself has no affinity for the sRNA-target duplex. Instead, it is targeted to the site of methylation through association with Nop5p, which in turn binds to the L7Ae-sRNA complex. To understand how Nop5p serves as a bridge between the targeting and catalytic functions of the box C/D small RNP complex, we have employed alanine scanning to evaluate the interaction between the Pyrococcus horikoshii Nop5p domain and an L7Ae box C/D RNA complex. From these data, we were able to construct an isolated RNA-binding domain (Nop-RBD) that folds correctly as demonstrated by x-ray crystallography and binds to the L7Ae box C/D RNA complex with near wild type affinity. These data demonstrate that the Nop-RBD is an autonomously folding and functional module important for protein assembly in a number of complexes centered on the L7Ae-kinkturn RNP.Many biological RNAs require extensive modification to attain full functionality in the cell (1). Currently there are over 100 known RNA modification types ranging from small functional group substitutions to the addition of large multi-cyclic ring structures (2). Transfer RNA, one of many functional RNAs targeted for modification (3-6), possesses the greatest modification type diversity, many of which are important for proper biological function (7). Ribosomal RNA, on the other hand, contains predominantly two types of modified nucleotides: pseudouridine and 2′-O-methylribose (8). The crystal structures of the ribosome suggest that these modifications are important for proper folding (9, 10) and structural stabilization (11) in vivo as evidenced by their strong tendency to localize to regions associated with function (8, 12, 13). These roles have been verified biochemically in a number of cases (14), whereas newly emerging functional modifications are continually being investigated.Box C/D ribonucleoprotein (RNP)3 complexes serve as RNA-guided site-specific 2′-O-methyltransferases in both archaea and eukaryotes (15, 16) where they are referred to as small RNP complexes and small nucleolar RNPs, respectively. Target RNA pairs with the sRNA guide sequence and is methylated at the 2′-hydroxyl group of the nucleotide five bases upstream of either the D or D′ box motif of the sRNA (Fig. 1, star) (17, 18). In archaea, the internal C′ and D′ motifs generally conform to a box C/D consensus sequence (19), and each sRNA contains two guide regions ∼12 nucleotides in length (20). The bipartite architecture of the RNP potentially enables the complex to methylate two distinct RNA targets (21) and has been shown to be essential for site-specific methylation (22).Open in a separate windowFIGURE 1.Organization of the archaeal box C/D complex. The protein components of this RNP are L7Ae, Nop5p, and fibrillarin, which together bind a box C/D sRNA. The regions of the Box C/D sRNA corresponding to the conserved C, D, C′, and D′ boxes are labeled. The target RNA binds the sRNA through Watson-Crick pairing and is methylated by fibrillarin at the fifth nucleotide from the D/D′ boxes (star).In addition to the sRNA, the archaeal box C/D complex requires three proteins for activity (23): the ribosomal protein L7Ae (24, 25), fibrillarin, and the Nop56/Nop58 homolog Nop5p (Fig. 1). L7Ae binds to both box C/D and the C′/D′ motifs (26), which respectively comprise kink-turn (27) or k-loop structures (28), to initiate the assembly of the RNP (29, 30). Fibrillarin performs the methyl group transfer from the cofactor S-adenosylmethionine to the target RNA (31-33). For this to occur, the active site of fibrillarin must be positioned precisely over the specific 2′-hydroxyl group to be methylated. Although fibrillarin methylates this functional group in the context of a Watson-Crick base-paired helix (guide/target), it has little to no binding affinity for double-stranded RNA or for the L7Ae-sRNA complex (22, 26, 33, 34). Nop5p serves as an intermediary protein bringing fibrillarin to the complex through its association with both the L7Ae-sRNA complex and fibrillarin (22). Along with its role as an intermediary between fibrillarin and the L7Ae-sRNA complex, Nop5p possesses other functions not yet fully understood. For example, Nop5p self-dimerizes through a coiled-coil domain (35) that in most archaea and eukaryotic homologs includes a small insertion sequence of unknown function (36, 37). However, dimerization and fibrillarin binding have been shown to be mutually exclusive in Methanocaldococcus jannaschii Nop5p, potentially because of the presence of this insertion sequence (36). Thus, whether Nop5p is a monomer or a dimer in the active RNP is still under debate.In this study, we focus our attention on the Nop5p protein to investigate its interaction with a L7Ae box C/D RNA complex because both the fibrillarin-Nop5p and the L7Ae box C/D RNA interfaces are known from crystal structures (29, 35, 38). Individual residues on the surface of a monomeric form of Nop5p (referred to as mNop5p) (22) were mutated to alanine, and the effect on binding affinity for a L7Ae box C/D motif RNA complex was assessed through the use of electrophoretic mobility shift assays. These data reveal that residues important for binding cluster within the highly conserved NOP domain (39, 40). To demonstrate that this domain is solely responsible for the affinity of Nop5p for the preassembled L7Ae box C/D RNA complex, we expressed and purified it in isolation from the full Nop5p protein. The isolated Nop-RBD domain binds to the L7Ae box C/D RNA complex with nearly wild type affinity, demonstrating that the Nop-RBD is truly an autonomously folding and functional module. Comparison of our data with the crystal structure of the homologous spliceosomal hPrp31-15.5K protein-U4 snRNA complex (41) suggests the adoption of a similar mode of binding, further supporting a crucial role for the NOP domain in RNP complex assembly.  相似文献   

14.
Rap1 small GTPases interact with Rap1-GTP-interacting adaptor molecule (RIAM), a member of the MRL (Mig-10/RIAM/Lamellipodin) protein family, to promote talin-dependent integrin activation. Here, we show that MRL proteins function as scaffolds that connect the membrane targeting sequences in Ras GTPases to talin, thereby recruiting talin to the plasma membrane and activating integrins. The MRL proteins bound directly to talin via short, N-terminal sequences predicted to form amphipathic helices. RIAM-induced integrin activation required both its capacity to bind to Rap1 and to talin. Moreover, we constructed a minimized 50-residue Rap-RIAM module containing the talin binding site of RIAM joined to the membrane-targeting sequence of Rap1A. This minimized Rap-RIAM module was sufficient to target talin to the plasma membrane and to mediate integrin activation, even in the absence of Rap1 activity. We identified a short talin binding sequence in Lamellipodin (Lpd), another MRL protein; talin binding Lpd sequence joined to a Rap1 membrane-targeting sequence is sufficient to recruit talin and activate integrins. These data establish the mechanism whereby MRL proteins interact with both talin and Ras GTPases to activate integrins.Increased affinity (“activation”) of cellular integrins is central to physiological events such as cell migration, assembly of the extracellular matrix, the immune response, and hemostasis (1). Each integrin comprises a type I transmembrane α and β subunit, each of which has a large extracellular domain, a single transmembrane domain, and a cytoplasmic domain (tail). Talin binds to most integrin β cytoplasmic domains and the binding of talin to the integrin β tail initiates integrin activation (24). A small, PTB-like domain of talin mediates activation via a two-site interaction with integrin β tails (5), and this PTB domain is functionally masked in the intact talin molecule (6). A central question in integrin biology is how the talin-integrin interaction is regulated to control integrin activation; recent work has implicated Ras GTPases as critical signaling modules in this process (7).Ras proteins are small monomeric GTPases that cycle between the GTP-bound active form and the GDP-bound inactive form. Guanine nucleotide exchange factors (GEFs) promote Ras activity by exchanging bound GDP for GTP, whereas GTPase-activating proteins (GAPs)3 enhance the hydrolysis of Ras-bound GTP to GDP (for review, see Ref. 8). The Ras subfamily members Rap1A and Rap1B stimulate integrin activation (9, 10). For example, expression of constitutively active Rap1 activates integrin αMβ2 in macrophage, and inhibition of Rap1 abrogated integrin activation induced by inflammatory agonists (1113). Murine T-cells expressing constitutively active Rap1 manifest enhanced integrin dependent cell adhesion (14). In platelets, Rap1 is rapidly activated by platelet agonists (15, 16). A knock-out of Rap1B (17) or of the Rap1GEF, RasGRP2 (18), resulted in impairment of αIIbβ3-dependent platelet aggregation, highlighting the importance of Rap1 in platelet aggregation in vivo. Thus, Rap1 GTPases play important roles in the activation of several integrins in multiple biological contexts.Several Rap1 effectors have been implicated in integrin activation (1921). Rap1-GTP-interacting adaptor molecule (RIAM) is a Rap1 effector that is a member of the MRL (Mig-10/RIAM/Lamellipodin) family of adaptor proteins (20). RIAM contains Ras association (RA) and pleckstrin homology (PH) domains and proline-rich regions, which are defining features of the MRL protein family. In Jurkat cells, RIAM overexpression induces β1 and β2 integrin-mediated cell adhesion, and RIAM knockdown abolishes Rap1-dependent cell adhesion (20), indicating RIAM is a downstream regulator of Rap1-dependent signaling. RIAM regulates actin dynamics as RIAM expression induces cell spreading; conversely, its depletion reduces cellular F-actin content (20). Whereas RIAM is greatly enriched in hematopoietic cells, Lamellipodin (Lpd) is a paralogue present in fibroblasts and other somatic cells (22).Recently we used forward, reverse, and synthetic genetics to engineer and order an integrin activation pathway in Chinese hamster ovary cells expressing a prototype activable integrin, platelet αIIbβ3. We found that Rap1 induced formation of an “integrin activation complex” containing RIAM and talin (23). Here, we have established the mechanism whereby Ras GTPases cooperate with MRL family proteins, RIAM and Lpd, to regulate integrin activation. We find that MRL proteins function as scaffolds that connect the membrane targeting sequences in Ras GTPases to talin, thereby recruiting talin to integrins at the plasma membrane.  相似文献   

15.
Bacillus cereus and other Gram-positive bacteria elaborate pili via a sortase D-catalyzed transpeptidation mechanism from major and minor pilin precursor substrates. After cleavage of the LPXTG sorting signal of the major pilin, BcpA, sortase D forms an amide bond between the C-terminal threonine and the amino group of lysine within the YPKN motif of another BcpA subunit. Pilus assembly terminates upon sortase A cleavage of the BcpA sorting signal, resulting in a covalent bond between BcpA and the cell wall cross-bridge. Here, we show that the IPNTG sorting signal of BcpB, the minor pilin, is cleaved by sortase D but not by sortase A. The C-terminal threonine of BcpB is amide-linked to the YPKN motif of BcpA, thereby positioning BcpB at the tip of pili. Thus, unique attributes of the sorting signals of minor pilins provide Gram-positive bacteria with a universal mechanism ordering assembly of pili.Sortases catalyze transpeptidation reactions to assemble proteins in the envelope of Gram-positive bacteria (1). Secreted proteins require a C-terminal sorting signal for sortase recognition such that sortase cleaves the substrate at a short peptide motif and forms a thioester-linked intermediate to its active site cysteine (24). Nucleophilic attack by an amino group within the bacterial envelope resolves the thioester intermediate, generating an amide bond tethering surface proteins at their C terminus onto Gram-positive bacteria (5). Four classes of sortases can be distinguished on the basis of sequence homology and substrate recognition (6, 7). Sortase A cleaves secreted protein at LPXTG sorting signals and recognizes the amino group of lipid II peptidoglycan precursors as a nucleophile (8, 9). Sortase B cleaves protein substrates at NPQTN sorting signals (10). This enzyme immobilizes proteins within fully assembled cell walls, utilizing the cell wall cross-bridge as a nucleophile (11). Sortase C cuts LPNTA sorting signals and anchors proteins to the peptidoglycan cross-bridges in sporulating bacteria (12, 13). Finally, sortase D catalyzes transpeptidation reactions in the assembly of pili (14, 15). Sortase D recognizes the amino group of lysine residues within the YPKN motif of pilin subunits as nucleophiles (16). The resultant sortase D-catalyzed amide bond links adjacent pilin subunits to grow the pilus fiber (16, 17).Pili of Gram-positive bacteria comprised either two or three different pilin subunits synthesized as cytoplasmic precursors with N-terminal signal peptides and C-terminal sorting signals (P1 precursors) (14, 18). After translocation across the plasma membrane, P2 precursor species arise from removal of the signal peptide from P1 precursors by a signal peptidase (16). Bacillus cereus pili are composed of two subunits; that is, the major pilin, BcpA, and the minor pilin, BcpB (15). In contrast to BcpA, which is deposited throughout the pilus, BcpB is found at fiber tip (15). Sortase D cleaves the BcpA LPXTG motif sorting signal between the threonine and glycine residues to form an amide bond to the ε-amino group of the lysine within the YPKN motif of adjacent BcpA subunits (16). However, sortase A also cleaves BcpA precursors, which are subsequently linked to the side chain amino group of meso-diaminopimelic acid within lipid II (19). The latter reaction serves to terminate fiber elongation, immobilizing BcpA pili in the cell wall envelope (19).The conservation of sortase D, the YPKN motif, and C-terminal sorting signal in major pilin subunits suggest a universal pilus assembly mechanism among Gram-positive bacteria (14, 20). However, the molecular mechanism whereby bacilli deposit BcpB, the minor pilin, at the tip of BcpA pili is not known. Although the BcpB precursor harbors an N-terminal signal peptide and a C-terminal IPNTG sorting signal, it lacks the YPKN pilin motif of the major subunit (15). Furthermore, the substrate properties of the BcpB IPNTG sorting signal for the four classes of sortases expressed by bacilli has yet to be established.  相似文献   

16.
PHLPP2 (PH domain leucine-rich repeat protein phosphatase 2) terminates Akt and protein kinase C (PKC) activity by specifically dephosphorylating these kinases at a key regulatory site, the hydrophobic motif (Ser-473 in Akt1). Here we identify a polymorphism that results in an amino acid change from a Leu to Ser at codon 1016 in the phosphatase domain of PHLPP2, which reduces phosphatase activity toward Akt both in vitro and in cells, in turn resulting in reduced apoptosis. Depletion of endogenous PHLPP2 variants in breast cancer cells revealed the Ser-1016 variant is less functional toward both Akt and PKC. In pair-matched high grade breast cancer samples we observed retention of only the Ser allele from heterozygous patients (identical results were observed in a pair-matched normal and tumor cell line). Thus, we have identified a functional polymorphism that impairs the activity of PHLPP2 and correlates with elevated Akt phosphorylation and increased PKC levels.Breast cancer is diagnosed in ∼180,000 women and is the cause of 40,000 deaths each year in the U.S.2 A prevalent underlying mechanism driving tumorigenesis is aberrant signal transduction pathways that result in constitutive activation of cell growth, proliferation, and survival pathways (2). A well characterized signal transduction pathway in breast cancer that promotes cellular survival, growth, and proliferation is the phosphatidylinositol 3-kinase/Akt pathway (3). This pathway is activated by a number of mechanisms, including gene amplification or gain of function mutations in upstream receptor protein-tyrosine kinases (4, 5), constitutive activation of hormone receptors (6), activating mutations in phosphatidylinositol 3-kinase and Akt (7, 8), and loss of function mutations in the regulatory phosphatase PTEN3 (phosphatase and tensin homolog on chromosome ten) (9). Thus, Akt is a major regulator of breast tumorigenesis.There are three isoforms of Akt present in humans. All three isoforms contain activating phosphorylation sites in the activation loop (Thr-308 in Akt1) and in the C-terminal hydrophobic motif (Ser-473 in Akt1) (10). Upon growth factor receptor stimulation, phosphatidylinositol 3-kinase becomes activated and phosphorylates the D3 position of, typically, phosphatidylinositol (4, 5) bisphosphate to generate phosphatidylinositol (3,4,5)-trisphosphate (11). This 3′-phosphorylated lipid recruits Akt to the plasma membrane by binding to its PH domain, resulting in conformational changes that allow access to the activation loop phosphorylation site (11). Constitutively bound phosphatidylinositol-dependent kinase-1 then phosphorylates Akt at Thr-308, accompanied by phosphorylation at Ser-473 resulting in a catalytically active kinase (12). Phosphorylation of Ser-473 depends on the mTORC2 complex (13-16). Signaling through this pathway is terminated by removal of the lipid second messenger phosphatidylinositol (3,4,5)-trisphosphate catalyzed by the phosphatase PTEN and by direct dephosphorylation of Akt by the recently-identified PHLPP family of phosphatases and protein phosphatase 2A-type phosphatases (17-20).The PHLPP family of phosphatases comprise three variants, the alternatively spliced PHLPP1α and PHLPP1β, and PHLPP2 (21). PHLPP1 and PHLPP2 specifically dephosphorylate the hydrophobic motif of specific Akt isozymes, thus decreasing Akt activity and promoting apoptosis (18, 19). PHLPP2 binds and dephosphorylates Akt1 and Akt3, whereas PHLPP1 binds and dephosphorylates Akt2 and Akt3 (18, 22). Their role in inactivating Akt suggests that both PHLPP1 and PHLPP2 could be potential tumor suppressors. Consistent with such a role, these phosphatases also dephosphorylate the hydrophobic motif of PKC, resulting in degradation of PKC. For this kinase, phosphorylation stabilizes the enzyme, so that the effect of depletion of the PHLPP phosphatases is to increase PKC protein levels (23). PKC is a well characterized oncogene, and loss of function of the PHLPP phosphatases could increase PKC protein levels and promote tumorigenesis (24). Providing further rationale that PHLPP2 could be a potential tumor suppressor, the phosphatase is located on chromosome 16q22.3, a region that encounters frequent loss of heterozygosity (LOH) in many primary and malignant breast tumors (25).Here we identify a non-synonymous polymorphism that results in an amino acid change from a Leu to a Ser at codon 1016 in the PP2C phosphatase domain of PHLPP2. Overexpression studies reveal the Ser-1016 variant has impaired phosphatase activity and is less effective at inducing apoptosis than the Leu-1016 variant. When comparing a pair-matched normal and breast cancer cell line or pair-matched normal and high grade tumor patient samples that are heterozygous, we observe preferential loss of the Leu allele in the tumor tissue or breast cancer cell line. This observation provides evidence that PHLPP2 could be one of the elusive tumor suppressor genes on chromosome 16q, and for heterozygous patients, loss of the more catalytically active Leu-1016 may promote breast tumorigenesis.  相似文献   

17.
The causative agent of Legionnaires disease, Legionella pneumophila, forms a replicative vacuole in phagocytes by means of the intracellular multiplication/defective organelle trafficking (Icm/Dot) type IV secretion system and translocated effector proteins, some of which subvert host GTP and phosphoinositide (PI) metabolism. The Icm/Dot substrate SidC anchors to the membrane of Legionella-containing vacuoles (LCVs) by specifically binding to phosphatidylinositol 4-phosphate (PtdIns(4)P). Using a nonbiased screen for novel L. pneumophila PI-binding proteins, we identified the Rab1 guanine nucleotide exchange factor (GEF) SidM/DrrA as the predominant PtdIns(4)P-binding protein. Purified SidM specifically and directly bound to PtdIns(4)P, whereas the SidM-interacting Icm/Dot substrate LidA preferentially bound PtdIns(3)P but also PtdIns(4)P, and the L. pneumophila Arf1 GEF RalF did not bind to any PIs. The PtdIns(4)P-binding domain of SidM was mapped to the 12-kDa C-terminal sequence, termed “P4M” (PtdIns4P binding of SidM/DrrA). The isolated P4M domain is largely helical and displayed higher PtdIns(4)P binding activity in the context of the α-helical, monomeric full-length protein. SidM constructs containing P4M were translocated by Icm/Dot-proficient L. pneumophila and localized to the LCV membrane, indicating that SidM anchors to PtdIns(4)P on LCVs via its P4M domain. An L. pneumophila ΔsidM mutant strain displayed significantly higher amounts of SidC on LCVs, suggesting that SidM and SidC compete for limiting amounts of PtdIns(4)P on the vacuole. Finally, RNA interference revealed that PtdIns(4)P on LCVs is specifically formed by host PtdIns 4-kinase IIIβ. Thus, L. pneumophila exploits PtdIns(4)P produced by PtdIns 4-kinase IIIβ to anchor the effectors SidC and SidM to LCVs.The Gram-negative pathogen Legionella pneumophila is the causative agent of Legionnaires disease, but it evolved as a parasite of various species of environmental predatory protozoa, including the social amoeba Dictyostelium discoideum (1, 2). The human disease is linked to the inhalation of contaminated aerosols, followed by replication in alveolar macrophages. To accommodate the transfer between host cells, L. pneumophila alternates between replicative and transmissive phases, the regulation of which includes an apparent quorum-sensing system (35).In macrophages and amoebae, L. pneumophila forms a replicative compartment, the Legionella-containing vacuole (LCV).3 LCVs avoid fusion with lysosomes (6), intercept vesicular traffic at endoplasmic reticulum (ER) exit sites (7), and fuse with the ER (810). The uptake of L. pneumophila and formation of LCVs in macrophages and amoebae depends on the Icm/Dot type IV secretion system (T4SS) (1114). Although more than 100 Icm/Dot substrates (“effector” proteins) have been identified to date, only few are functionally characterized, including effectors that interfere with host cell signal transduction, vesicle trafficking, or apoptotic pathways (1518).Two Icm/Dot-translocated substrates, SidM/DrrA (19, 20) and RalF (21), have been characterized as guanine nucleotide exchange factors (GEFs) for the Rho subfamily of small GTPases. These bacterial GEFs are recruited to and activate their targets on LCVs. Small GTPases of the Rho subfamily are involved in many eukaryotic signal transduction pathways and in actin cytoskeleton regulation (22). Inactive Rho GTPases bind GDP and a guanine nucleotide dissociation inhibitor (GDI). The GTPases are activated by removal of the GDI and the exchange of GDP with GTP by GEFs, which promotes the interaction with downstream effector proteins, such as protein or lipid kinases and various adaptor proteins. The cycle is closed by hydrolysis of the bound GTP, which is mediated by GTPase-activating proteins.SidM is a GEF for Rab1, which is essential for ER to Golgi vesicle transport, and additionally, SidM acts as a GDI displacement factor (GDF) to activate Rab1 (23, 24). The function of SidM is assisted by the Icm/Dot substrate LidA, which also localizes to LCVs. LidA preferentially binds to activated Rab1, thus supporting the recruitment of early secretory vesicles by SidM (19, 20, 23, 25, 26). Another Icm/Dot substrate, LepB (27), contributes to Rab1-mediated membrane cycling by inactivating Rab1 through its GTPase-activating protein function, thus acting as an antagonist of SidM (24).The Icm/Dot substrate RalF recruits and activates the small GTPase ADP-ribosylation factor 1 (Arf1), which is involved in retrograde vesicle transport from Golgi to ER (21). Dominant negative Arf1 (7, 28) or knockdown of Arf1 by RNA interference (29) impairs the formation of LCVs, as well as the recruitment of the Icm/Dot substrate SidC to the LCV (30).SidC and its paralogue SdcA localize to the LCV membrane (31), where the proteins specifically bind to the host cell lipid phosphatidylinositol 4-phosphate (PtdIns(4)P) (32, 33). Phosphoinositides (PIs) regulate eukaryotic receptor-mediated signal transduction, actin remodeling, and membrane dynamics (34, 35). PtdIns(4)P is present on the cytoplasmic membrane, but localizes preferentially to the trans-Golgi network (TGN), where this PI is produced by an Arf-dependent recruitment of PtdIns(4)P kinase IIIβ (PI4K IIIβ) (36) to promote trafficking along the secretory pathway. Recently, PtdIns(4)P was found to also mediate the export of early secretory vesicles from ER exit sites (37). At present, the L. pneumophila effector proteins that mediate exploitation of host PI signaling remain ill defined.In a nonbiased screen for L. pneumophila PI-binding proteins using different PIs coupled to agarose beads, we identified SidM as a major PtdIns(4)P-binding effector. We mapped its PtdIns(4)P binding activity to a novel P4M domain within a 12-kDa C-terminal sequence. SidM constructs, including the P4M domain, were found to be translocated and bind the LCV membrane, where the levels of PtdIns(4)P are controlled by PI4K IIIβ.  相似文献   

18.
19.
Hepatocellular carcinoma (HCC) is one of the most common and aggressive human malignancies. Recombinant tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) is a promising anti-tumor agent. However, many HCC cells show resistance to TRAIL-induced apoptosis. In this study, we showed that bortezomib, a proteasome inhibitor, overcame TRAIL resistance in HCC cells, including Huh-7, Hep3B, and Sk-Hep1. The combination of bortezomib and TRAIL restored the sensitivity of HCC cells to TRAIL-induced apoptosis. Comparing the molecular change in HCC cells treated with these agents, we found that down-regulation of phospho-Akt (P-Akt) played a key role in mediating TRAIL sensitization of bortezomib. The first evidence was that bortezomib down-regulated P-Akt in a dose- and time-dependent manner in TRAIL-treated HCC cells. Second, LY294002, a PI3K inhibitor, also sensitized resistant HCC cells to TRAIL-induced apoptosis. Third, knocking down Akt1 by small interference RNA also enhanced TRAIL-induced apoptosis in Huh-7 cells. Finally, ectopic expression of mutant Akt (constitutive active) in HCC cells abolished TRAIL sensitization effect of bortezomib. Moreover, okadaic acid, a protein phosphatase 2A (PP2A) inhibitor, reversed down-regulation of P-Akt in bortezomib-treated cells, and PP2A knockdown by small interference RNA also reduced apoptosis induced by the combination of TRAIL and bortezomib, indicating that PP2A may be important in mediating the effect of bortezomib on TRAIL sensitization. Together, bortezomib overcame TRAIL resistance at clinically achievable concentrations in hepatocellular carcinoma cells, and this effect is mediated at least partly via inhibition of the PI3K/Akt pathway.Hepatocellular carcinoma (HCC)2 is currently the fifth most common solid tumor worldwide and the fourth leading cause of cancer-related death. To date, surgery is still the only curative treatment but is only feasible in a small portion of patients (1). Drug treatment is the major therapy for patients with advanced stage disease. Unfortunately, the response rate to traditional chemotherapy for HCC patients is unsatisfactory (1). Novel pharmacological therapy is urgently needed for patients with advanced HCC. In this regard, the approval of sorafenib might open a new era of molecularly targeted therapy in the treatment of HCC patients.Tumor necrosis factor (TNF)-related apoptosis-inducing ligand (TRAIL), a type II transmembrane protein and a member of the TNF family, is a promising anti-tumor agent under clinical investigation (2). TRAIL functions by engaging its receptors expressed on the surface of target cells. Five receptors specific for TRAIL have been identified, including DR4/TRAIL-R1, DR5/TRAIL-R2, DcR1, DcR2, and osteoprotegerin. Among TRAIL receptors, only DR4 and DR5 contain an effective death domain that is essential to formation of death-inducing signaling complex (DISC), a critical step for TRAIL-induced apoptosis. Notably, the trimerization of the death domains recruits an adaptor molecule, Fas-associated protein with death domain (FADD), which subsequently recruits and activates caspase-8. In type I cells, activation of caspase-8 is sufficient to activate caspase-3 to induce apoptosis; however, in another type of cells (type II), the intrinsic mitochondrial pathway is essential for apoptosis characterized by cleavage of Bid and release of cytochrome c from mitochondria, which subsequently activates caspase-9 and caspase-3 (3).Although TRAIL induces apoptosis in malignant cells but sparing normal cells, some tumor cells are resistant to TRAIL-induced apoptosis. Mechanisms responsible for the resistance include receptors and intracellular resistance. Although the cell surface expression of DR4 or DR5 is absolutely required for TRAIL-induced apoptosis, tumor cells expressing these death receptors are not always sensitive to TRAIL due to intracellular mechanisms. For example, the cellular FLICE-inhibitory protein (c-FLIP), a homologue to caspase-8 but without protease activity, has been linked to TRAIL resistance in several studies (4, 5). In addition, inactivation of Bax, a proapoptotic Bcl-2 family protein, resulted in resistance to TRAIL in MMR-deficient tumors (6, 7), and reintroduction of Bax into Bax-deficient cells restored TRAIL sensitivity (8), indicating that the Bcl-2 family plays a critical role in intracellular mechanisms for resistance of TRAIL.Bortezomib, a proteasome inhibitor approved clinically for multiple myeloma and mantle cell lymphoma, has been investigated intensively for many types of cancer (9). Accumulating studies indicate that the combination of bortezomib and TRAIL overcomes the resistance to TRAIL in various types of cancer, including acute myeloid leukemia (4), lymphoma (1013), prostate (1417), colon (15, 18, 19), bladder (14, 16), renal cell carcinoma (20), thyroid (21), ovary (22), non-small cell lung (23, 24), sarcoma (25), and HCC (26, 27). Molecular targets responsible for the sensitizing effect of bortezomib on TRAIL-induced cell death include DR4 (14, 27), DR5 (14, 20, 2223, 28), c-FLIP (4, 11, 2123, 29), NF-κB (12, 24, 30), p21 (16, 21, 25), and p27 (25). In addition, Bcl-2 family also plays a role in the combinational effect of bortezomib and TRAIL, including Bcl-2 (10, 21), Bax (13, 22), Bak (27), Bcl-xL (21), Bik (18), and Bim (15).Recently, we have reported that Akt signaling is a major molecular determinant in bortezomib-induced apoptosis in HCC cells (31). In this study, we demonstrated that bortezomib overcame TRAIL resistance in HCC cells through inhibition of the PI3K/Akt pathway.  相似文献   

20.
The acid-sensing ion channel 1a (ASIC1a) is widely expressed in central and peripheral neurons where it generates transient cation currents when extracellular pH falls. ASIC1a confers pH-dependent modulation on postsynaptic dendritic spines and has critical effects in neurological diseases associated with a reduced pH. However, knowledge of the proteins that interact with ASIC1a and influence its function is limited. Here, we show that α-actinin, which links membrane proteins to the actin cytoskeleton, associates with ASIC1a in brain and in cultured cells. The interaction depended on an α-actinin-binding site in the ASIC1a C terminus that was specific for ASIC1a versus other ASICs and for α-actinin-1 and -4. Co-expressing α-actinin-4 altered ASIC1a current density, pH sensitivity, desensitization rate, and recovery from desensitization. Moreover, reducing α-actinin expression altered acid-activated currents in hippocampal neurons. These findings suggest that α-actinins may link ASIC1a to a macromolecular complex in the postsynaptic membrane where it regulates ASIC1a activity.Acid-sensing ion channels (ASICs)2 are H+-gated members of the DEG/ENaC family (13). Members of this family contain cytosolic N and C termini, two transmembrane domains, and a large cysteine-rich extracellular domain. ASIC subunits combine as homo- or heterotrimers to form cation channels that are widely expressed in the central and peripheral nervous systems (14). In mammals, four genes encode ASICs, and two subunits, ASIC1 and ASIC2, have two splice forms, a and b. Central nervous system neurons express ASIC1a, ASIC2a, and ASIC2b (57). Homomeric ASIC1a channels are activated when extracellular pH drops below 7.2, and half-maximal activation occurs at pH 6.5–6.8 (810). These channels desensitize in the continued presence of a low extracellular pH, and they can conduct Ca2+ (9, 1113). ASIC1a is required for acid-evoked currents in central nervous system neurons; disrupting the gene encoding ASIC1a eliminates H+-gated currents unless extracellular pH is reduced below pH 5.0 (5, 7).Previous studies found ASIC1a enriched in synaptosomal membrane fractions and present in dendritic spines, the site of excitatory synapses (5, 14, 15). Consistent with this localization, ASIC1a null mice manifested deficits in hippocampal long term potentiation, learning, and memory, which suggested that ASIC1a is required for normal synaptic plasticity (5, 16). ASICs might be activated during neurotransmission when synaptic vesicles empty their acidic contents into the synaptic cleft or when neuronal activity lowers extracellular pH (1719). Ion channels, including those at the synapse often interact with multiple proteins in a macromolecular complex that incorporates regulators of their function (20, 21). For ASIC1a, only a few interacting proteins have been identified. Earlier work indicated that ASIC1a interacts with another postsynaptic scaffolding protein, PICK1 (15, 22, 23). ASIC1a also has been reported to interact with annexin II light chain p11 through its cytosolic N terminus to increase cell surface expression (24) and with Ca2+/calmodulin-dependent protein kinase II to phosphorylate the channel (25). However, whether ASIC1a interacts with additional proteins and with the cytoskeleton remain unknown. Moreover, it is not known whether such interactions alter ASIC1a function.In analyzing the ASIC1a amino acid sequence, we identified cytosolic residues that might bind α-actinins. α-Actinins cluster membrane proteins and signaling molecules into macromolecular complexes and link membrane proteins to the actincytoskeleton (for review, Ref. 26). Four genes encode α-actinin-1, -2, -3, and -4 isoforms. α-Actinins contain an N-terminal head domain that binds F-actin, a C-terminal region containing two EF-hand motifs, and a central rod domain containing four spectrin-like motifs (2628). The C-terminal portion of the rod segment appears to be crucial for binding to membrane proteins. The α-actinins assemble into antiparallel homodimers through interactions in their rod domain. α-Actinins-1, -2, and -4 are enriched in dendritic spines, concentrating at the postsynaptic membrane (2935). In the postsynaptic membrane of excitatory synapses, α-actinin connects the NMDA receptor to the actin cytoskeleton, and this interaction is key for Ca2+-dependent inhibition of NMDA receptors (3638). α-Actinins can also regulate the membrane trafficking and function of several cation channels, including L-type Ca2+ channels, K+ channels, and TRP channels (3941).To better understand the function of ASIC1a channels in macromolecular complexes, we asked if ASIC1a associates with α-actinins. We were interested in the α-actinins because they and ASIC1a, both, are present in dendritic spines, ASIC1a contains a potential α-actinin binding sequence, and the related epithelial Na+ channel (ENaC) interacts with the cytoskeleton (42, 43). Therefore, we hypothesized that α-actinin interacts structurally and functionally with ASIC1a.  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