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 31 毫秒
1.
A central question in Wnt signaling is the regulation of β-catenin phosphorylation and degradation. Multiple kinases, including CKIα and GSK3, are involved in β-catenin phosphorylation. Protein phosphatases such as PP2A and PP1 have been implicated in the regulation of β-catenin. However, which phosphatase dephosphorylates β-catenin in vivo and how the specificity of β-catenin dephosphorylation is regulated are not clear. In this study, we show that PP2A regulates β-catenin phosphorylation and degradation in vivo. We demonstrate that PP2A is required for Wnt/β-catenin signaling in Drosophila. Moreover, we have identified PR55α as the regulatory subunit of PP2A that controls β-catenin phosphorylation and degradation. PR55α, but not the catalytic subunit, PP2Ac, directly interacts with β-catenin. RNA interference knockdown of PR55α elevates β-catenin phosphorylation and decreases Wnt signaling, whereas overexpressing PR55α enhances Wnt signaling. Taken together, our results suggest that PR55α specifically regulates PP2A-mediated β-catenin dephosphorylation and plays an essential role in Wnt signaling.Wnt/β-catenin signaling plays essential roles in development and tumorigenesis (13). Our previous work found that β-catenin is sequentially phosphorylated by CKIα4 and GSK3 (4), which creates a binding site for β-Trcp (5), leading to degradation via the ubiquitination/proteasome machinery (3). Mutations in β-catenin or APC genes that prevent β-catenin phosphorylation or ubiquitination/degradation lead ultimately to cancer (1, 2).In addition to the involvement of kinases, protein phosphatases, such as PP1, PP2A, and PP2C, are also implicated in Wnt/β-catenin regulation. PP2C and PP1 may regulate dephosphorylation of Axin and play positive roles in Wnt signaling (6, 7). PP2A is a multisubunit enzyme (810); it has been reported to play either positive or negative roles in Wnt signaling likely by targeting different components (1121). Toward the goal of understanding the mechanism of β-catenin phosphorylation, we carried out siRNA screening targeting several major phosphatases, in which we found that PP2A dephosphorylates β-catenin. This is consistent with a recent study where PP2A is shown to dephosphorylate β-catenin in a cell-free system (18).PP2A consists of a catalytic subunit (PP2Ac), a structure subunit (PR65/A), and variable regulatory B subunits (PR/B, PR/B′, PR/B″, or PR/B‴). The substrate specificity of PP2A is thought to be determined by its B subunit (9). By siRNA screening, we further identified that PR55α, a regulatory subunit of PP2A, specifically regulates β-catenin phosphorylation and degradation. Mechanistically, we found that PR55α directly interacts with β-catenin and regulates PP2A-mediated β-catenin dephosphorylation in Wnt signaling.  相似文献   

2.
3.
4.
5.
6.
7.
8.
Collapsin response mediator protein 2 (CRMP2) is an intracellular protein that mediates signaling of Semaphorin3A (Sema3A), a repulsive axon guidance molecule. Fyn, a Src-type tyrosine kinase, is involved in the Sema3A signaling. However, the relationship between CRMP2 and Fyn in this signaling pathway is still unknown. In our research, we demonstrated that Fyn phosphorylated CRMP2 at Tyr32 residues in HEK293T cells. Immunohistochemical analysis using a phospho-specific antibody at Tyr32 of CRMP showed that Tyr32-phosphorylated CRMP was abundant in the nervous system, including dorsal root ganglion neurons, the molecular and Purkinje cell layer of adult cerebellum, and hippocampal fimbria. Overexpression of a nonphosphorylated mutant (Tyr32 to Phe32) of CRMP2 in dorsal root ganglion neurons interfered with Sema3A-induced growth cone collapse response. These results suggest that Fyn-dependent phosphorylation of CRMP2 at Tyr32 is involved in Sema3A signaling.Collapsin response mediator proteins (CRMPs)4 have been identified as intracellular proteins that mediate Semaphorin3A (Sema3A) signaling in the nervous system (1). CRMP2 is one of the five members of the CRMP family. CRMPs also mediate signal transduction of NT3, Ephrin, and Reelin (24). CRMPs interact with several intracellular molecules, including tubulin, Numb, kinesin1, and Sra1 (58). CRMPs are involved in axon guidance, axonal elongation, cell migration, synapse maturation, and the generation of neuronal polarity (1, 2, 4, 5).CRMP family proteins are known to be the major phosphoproteins in the developing brain (1, 9). CRMP2 is phosphorylated by several Ser/Thr kinases, such as Rho kinase, cyclin-dependent kinase 5 (Cdk5), and glycogen synthase kinase 3β (GSK3β) (2, 1013). The phosphorylation sites of CRMP2 by these kinases are clustered in the C terminus and have already been identified. Rho kinase phosphorylates CRMP2 at Thr555 (10). Cdk5 phosphorylates CRMP2 at Ser522, and this phosphorylation is essential for sequential phosphorylations by GSK3β at Ser518, Thr514, and Thr509 (2, 1113). These phosphorylations disrupt the interaction of CRMP2 with tubulin or Numb (2, 3, 13). The sequential phosphorylation of CRMP2 by Cdk5 and GSK3β is an essential step in Sema3A signaling (11, 13). Furthermore, the neurofibrillary tangles in the brains of people with Alzheimer disease contain hyperphosphorylated CRMP2 at Thr509, Ser518, and Ser522 (14, 15).CRMPs are also substrates of several tyrosine kinases. The phosphorylation of CRMP2 by Fes/Fps and Fer has been shown to be involved in Sema3A signaling (16, 17). Phosphorylation of CRMP2 at Tyr479 by a Src family tyrosine kinase Yes regulates CXCL12-induced T lymphocyte migration (18). We reported previously that Fyn is involved in Sema3A signaling (19). Fyn associates with PlexinA2, one of the components of the Sema3A receptor complex. Fyn also activates Cdk5 through the phosphorylation at Tyr15 of Cdk5 (19). In dorsal root ganglion (DRG) neurons from fyn-deficient mice, Sema3A-induced growth cone collapse response is attenuated compared with control mice (19). Furthermore, we recently found that Fyn phosphorylates CRMP1 and that this phosphorylation is involved in Reelin signaling (4). Although it has been shown that CRMP2 is involved in Sema3A signaling (1, 11, 13), the relationship between Fyn and CRMP2 in Sema3A signaling and the tyrosine phosphorylation site(s) of CRMPs remain unknown.Here, we show that Fyn phosphorylates CRMP2 at Tyr32. Using a phospho-specific antibody against Tyr32, we determined that the residue is phosphorylated in vivo. A nonphosphorylated mutant CRMP2Y32F inhibits Sema3A-induced growth cone collapse. These results indicate that tyrosine phosphorylation by Fyn at Tyr32 is involved in Sema3A signaling.  相似文献   

9.
10.
Complexes involved in the γ/ϵ-secretase-regulated intramembranous proteolysis of substrates such as the amyloid-β precursor protein are composed primarily of presenilin (PS1 or PS2), nicastrin, anterior pharynx defective-1 (APH1), and PEN2. The presenilin aspartyl residues form the catalytic site, and similar potentially functional polar transmembrane residues in APH1 have been identified. Substitution of charged (E84A, R87A) or polar (Q83A) residues in TM3 had no effect on complex assembly or activity. In contrast, changes to either of two highly conserved histidines (H171A, H197A) located in TM5 and TM6 negatively affected PS1 cleavage and altered binding to other secretase components, resulting in decreased amyloid generating activity. Charge replacement with His-to-Lys substitutions rescued nicastrin maturation and PS1 endoproteolysis leading to assembly of the formation of structurally normal but proteolytically inactive γ-secretase complexes. Substitution with a negatively charged side chain (His-to-Asp) or altering the structural location of the histidines also disrupted γ-secretase binding and abolished functionality of APH1. These results suggest that the conserved transmembrane histidine residues contribute to APH1 function and can affect presenilin catalytic activity.The anterior pharynx defective-1 (APH1)5 protein is an essential component of presenilin-dependent complexes required for the γ/ϵ-secretase activity (1). The multicomponent γ-secretase is responsible for the intramembrane proteolysis of a variety of substrates including the amyloid-β precursor protein (APP) and Notch receptor. Notch signaling is involved in a variety of important cell fate decisions during embryogenesis and adulthood (2). The γ/ϵ-secretase cleavage of APP protein is related to the pathogenesis of Alzheimer disease by releasing the 4-kDa amyloid β-peptide (Aβ) which accumulates as senile plaques in patients with Alzheimer disease (3, 4).The γ-complexes are composed of multispanning transmembrane proteins that include APH1 (5, 6), presenilin (PS1 or PS2) (710), PEN2 (5), and the type 1 transmembrane nicastrin (NCT) (11). All four components are essential for proteolytic activity, and loss of any single component destabilizes the complex, resulting in the loss of substrate cleavage. Conversely, co-expression of all four components increases γ-secretase activity (1214). During the maturation of the complexes, presenilins undergo an endoproteolytic cleavage to generate amino- and carboxyl-terminal fragments which remain associated as heterodimers in the active high molecular weight complexes (1518). Although the exact function of presenilins has been debated (19, 20), it has been proposed that the presenilins are aspartyl proteases with two transmembrane residues constituting the catalytic subunit (21). Analogous aspartyl catalytic dyads are found in the signal peptide peptidases (21, 22). Contributions from the other components are under investigation, and it has been shown, for example, that the large ectodomain of NCT plays a key role in substrate recognition (23, 24). It has also been shown that other proteins can regulate activity such as TMP21, a member of p24 cargo protein, which binds to the presenilin complexes and selectively modulates γ but not ϵ cleavage (25, 26).APH1 is a seven-transmembrane protein with a topology such that the amino terminus is oriented with the endoplasmic reticulum and the carboxyl terminus resides in the cytoplasm (6, 27). It is also expressed as different isoforms encoded by two genes in humans (APH1a on chromosome 1; APH1b on chromosome 15) or three genes in rodents (APH1a on chromosome 3; APH1b and APH1c on chromosome 9). APH1a has 55% sequence similarity with APH1b/APH1c, whereas APH1b and APH1c share 95% similarity. In addition to these different genes, APH1a is alternatively spliced to generate a short (APH1aS) and a long isoform (APH1aL). These two isoforms differ by the addition of 18 residues on the carboxyl-terminal part of APH1aL (28, 29). Deletion of APH1a in mice is embryonically lethal and is associated with developmental and patterning defects similar to those found in Notch, NCT, or PS1 null embryos (30, 31). In contrast to the essential nature of APH1a, the combined APH1b/c-deficient mice survive into adulthood (31). This suggests that APH1a is the major homologue involved in presenilin-dependent function during embryonic development. In addition, these different APH1 variants are constituents of distinct, proteolytically active presenilin-containing complexes and may, therefore, make unique contributions to γ-secretase activity (3032).Despite their importance to complex formation and function, the exact role of the APH1 isoforms in presenilin-dependent γ/ϵ-secretase activity remains under investigation. In the current study, several highly conserved polar and charged residues located within the transmembrane domains of APH1 were identified. Mutagenesis of two conserved histidine residues embedded in TM5 and TM6 (His-171 and His-197) lead to alterations in γ-secretase complex maturation and activity. The histidine residues contribute to APH1 function and are involved in stabilizing interactions with other γ-secretase components. These key histidines may also be physically localized near the presenilin active site and involved in the γ-secretase activity as shown by the decreased activity of γ-secretase complexes that are assembled with the His-mutants.  相似文献   

11.
The present study tests the hypothesis that the structure of extracellular domain Loop 2 can markedly affect ethanol sensitivity in glycine receptors (GlyRs) and γ-aminobutyric acid type A receptors (GABAARs). To test this, we mutated Loop 2 in the α1 subunit of GlyRs and in the γ subunit of α1β2γ2GABAARs and measured the sensitivity of wild type and mutant receptors expressed in Xenopus oocytes to agonist, ethanol, and other agents using two-electrode voltage clamp. Replacing Loop 2 of α1GlyR subunits with Loop 2 from the δGABAAR (δL2), but not the γGABAAR subunit, reduced ethanol threshold and increased the degree of ethanol potentiation without altering general receptor function. Similarly, replacing Loop 2 of the γ subunit of GABAARs with δL2 shifted the ethanol threshold from 50 mm in WT to 1 mm in the GABAA γ-δL2 mutant. These findings indicate that the structure of Loop 2 can profoundly affect ethanol sensitivity in GlyRs and GABAARs. The δL2 mutations did not affect GlyR or GABAAR sensitivity, respectively, to Zn2+ or diazepam, which suggests that these δL2-induced changes in ethanol sensitivity do not extend to all allosteric modulators and may be specific for ethanol or ethanol-like agents. To explore molecular mechanisms underlying these results, we threaded the WT and δL2 GlyR sequences onto the x-ray structure of the bacterial Gloeobacter violaceus pentameric ligand-gated ion channel homologue (GLIC). In addition to being the first GlyR model threaded on GLIC, the juxtaposition of the two structures led to a possible mechanistic explanation for the effects of ethanol on GlyR-based on changes in Loop 2 structure.Alcohol abuse and dependence are significant problems in our society, with ∼14 million people in the United States being affected (1, 2). Alcohol causes over 100,000 deaths in the United States, and alcohol-related issues are estimated to cost nearly 200 billion dollars annually (2). To address this, considerable attention has focused on the development of medications to prevent and treat alcohol-related problems (35). The development of such medications would be aided by a clear understanding of the molecular structures on which ethanol acts and how these structures influence receptor sensitivity to ethanol.Ligand-gated ion channels (LGICs)2 have received substantial attention as putative sites of ethanol action that cause its behavioral effects (612). Research in this area has focused on investigating the effects of ethanol on two large superfamilies of LGICs: 1) the Cys-loop superfamily of LGICs (13, 14), whose members include nicotinic acetylcholine, 5-hydroxytryptamine3, γ-aminobutyric acid type A (GABAA), γ-aminobutyric acid type C, and glycine receptors (GlyRs) (10, 11, 1520) and 2) the glutamate superfamily, including N-methyl d-aspartate, α-amino-3-hydroxyisoxazolepropionic acid, and kainate receptors (21, 22). Recent studies have also begun investigating ethanol action in the ATP-gated P2X superfamily of LGICs (2325).A series of studies that employed chimeric and mutagenic strategies combined with sulfhydryl-specific labeling identified key regions within Cys-loop receptors that appear to be initial targets for ethanol action that also can determine the sensitivity of the receptors to ethanol (712, 18, 19, 2630). This work provides several lines of evidence that position 267 and possibly other sites in the transmembrane (TM) domain of GlyRs and homologous sites in GABAARs are targets for ethanol action and that mutations at these sites can influence ethanol sensitivity (8, 9, 26, 31).Growing evidence from GlyRs indicates that ethanol also acts on the extracellular domain. The initial findings came from studies demonstrating that α1GlyRs are more sensitive to ethanol than are α2GlyRs despite the high (∼78%) sequence homology between α1GlyRs and α2GlyRs (32). Further work found that an alanine to serine exchange at position 52 (A52S) in Loop 2 can eliminate the difference in ethanol sensitivity between α1GlyRs and α2GlyRs (18, 20, 33). These studies also demonstrated that mutations at position 52 in α1GlyRS and the homologous position 59 in α2GlyRs controlled the sensitivity of these receptors to a novel mechanistic ethanol antagonist (20). Collectively, these studies suggest that there are multiple sites of ethanol action in α1GlyRs, with one site located in the TM domain (e.g. position 267) and another in the extracellular domain (e.g. position 52).Subsequent studies revealed that the polarity of the residue at position 52 plays a key role in determining the sensitivity of GlyRs to ethanol (20). The findings with polarity in the extracellular domain contrast with the findings at position 267 in the TM domain, where molecular volume, but not polarity, significantly affected ethanol sensitivity (9). Taken together, these findings indicate that the physical-chemical parameters of residues at positions in the extracellular and TM domains that modulate ethanol effects and/or initiate ethanol action in GlyRs are not uniform. Thus, knowledge regarding the physical-chemical properties that control agonist and ethanol sensitivity is key for understanding the relationship between the structure and the actions of ethanol in LGICs (19, 31, 3440).GlyRs and GABAARs, which differ significantly in their sensitivities to ethanol, offer a potential method for identifying the structures that control ethanol sensitivity. For example, α1GlyRs do not reliably respond to ethanol concentrations less than 10 mm (32, 33, 41). Similarly, γ subunit-containing GABAARs (e.g. α1β2γ2), the most predominantly expressed GABAARs in the central nervous system, are insensitive to ethanol concentrations less than 50 mm (42, 43). In contrast, δ subunit-containing GABAARs (e.g. α4β3δ) have been shown to be sensitive to ethanol concentrations as low as 1–3 mm (4451). Sequence alignment of α1GlyR, γGABAAR, and δGABAAR revealed differences between the Loop 2 regions of these receptor subunits. Since prior studies found that mutations of Loop 2 residues can affect ethanol sensitivity (19, 20, 39), the non-conserved residues in Loop 2 of GlyR and GABAAR subunits could provide the physical-chemical and structural bases underlying the differences in ethanol sensitivity between these receptors.The present study tested the hypothesis that the structure of Loop 2 can markedly affect the ethanol sensitivity of GlyRs and GABAARs. To accomplish this, we performed multiple mutations that replaced the Loop 2 region of the α1 subunit in α1GlyRs and the Loop 2 region of the γ subunit of α1β2γ2 GABAARs with corresponding non-conserved residues from the δ subunit of GABAAR and tested the sensitivity of these receptors to ethanol. As predicted, replacing Loop 2 of WT α1GlyRs with the homologous residues from the δGABAAR subunit (δL2), but not the γGABAAR subunit (γL2), markedly increased the sensitivity of the receptor to ethanol. Similarly, replacing the non-conserved residues of the γ subunit of α1β2γ2 GABAARs with δL2 also markedly increased ethanol sensitivity of GABAARs. These findings support the hypothesis and suggest that Loop 2 may play a role in controlling ethanol sensitivity across the Cys-loop superfamily of receptors. The findings also provide the basis for suggesting structure-function relationships in a new molecular model of the GlyR based on the bacterial Gloeobacter violaceus pentameric LGIC homologue (GLIC).  相似文献   

12.
13.
Laminins are large heterotrimeric glycoproteins with many essential functions in basement membrane assembly and function. Cell adhesion to laminins is mediated by a tandem of five laminin G-like (LG) domains at the C terminus of the α chain. Integrin binding requires an intact LG1-3 region, as well as contributions from the coiled coil formed by the α, β, and γ chains. We have determined the crystal structure at 2.8-Å resolution of the LG1-3 region of the laminin α2 chain (α2LG1-3). The three LG domains adopt typical β-sandwich folds, with canonical calcium binding sites in LG1 and LG2. LG2 and LG3 interact through a substantial interface, but LG1 is completely dissociated from the LG2-3 pair. We suggest that the missing γ chain tail may be required to stabilize the interaction between LG1 and LG2-3 in the biologically active conformation. A global analysis of N-linked glycosylation sites shows that the β-sandwich faces of LG1 are free of carbohydrate modifications in all five laminin α chains, suggesting that these surfaces may harbor the integrin binding site. The α2LG1-3 structure provides the first atomic view of the integrin binding region of laminins.The laminins constitute a major class of cell-adhesive glycoproteins that are intimately involved in basement membrane assembly and function. Their essential roles in embryo development and tissue function have been demonstrated by numerous genetic studies and the analysis of severe human diseases resulting from mutations in laminin genes (14). All laminins are heterotrimers composed of three different gene products, termed α, β, and γ chains. At present, 16 mouse and human laminins are known, assembled from five α, three β, and three γ chains. The different laminins have characteristic expression patterns and functions in the embryo and adult animal (1). Laminins are cross-shaped molecules: the three short arms are composed of one chain each, while the long arm is a coiled coil of all three chains, terminating in a tandem of five laminin G-like (LG)2 domains, LG1-5, contributed by the α chain (2). Basement membrane assembly requires polymerization via the short arms and cell attachment via the LG1-5 region (5, 6).Cell adhesion to laminins is mediated by multiple receptors: integrins bind to the LG1-3 region, whereas α-dystroglycan, heparan sulfate proteoglycans, and sulfated glycolipids bind predominantly to sites in the LG4-5 pair (7). Integrins are heterodimers with a large extracellular domain consisting of one α and one β chain, which both span the cell membrane and engage in transmembrane signaling (8). Of the 24 mouse and human integrins, the major laminin binding integrins are α3β1, α6β1, α7β1, and α6β4, which have distinct affinities for the different laminin isoforms (9). Although some studies have reported integrin binding or integrin-mediated cell adhesion to isolated LG domains or tandems (1012), there is strong evidence to suggest that the coiled coil region and an intact γ chain tail are required for full integrin binding to the laminin LG1-3 region (1318). Compared with integrin binding to collagen and fibronectin, which is understood in atomic detail (19, 20), the laminin-integrin interaction remains poorly characterized in structural terms. We previously determined crystal structures of the LG4-5 region of the laminin α1 and α2 chains and defined their receptor binding sites (2123). Here, we report the crystal structure of the remainder of the laminin α2 receptor binding region, LG1-3.  相似文献   

14.
Accumulation of amyloid β (Aβ) oligomers in the brain is toxic to synapses and may play an important role in memory loss in Alzheimer disease. However, how these toxins are built up in the brain is not understood. In this study we investigate whether impairments of insulin and insulin-like growth factor-1 (IGF-1) receptors play a role in aggregation of Aβ. Using primary neuronal culture and immortal cell line models, we show that expression of normal insulin or IGF-1 receptors confers cells with abilities to reduce exogenously applied Aβ oligomers (also known as ADDLs) to monomers. In contrast, transfection of malfunctioning human insulin receptor mutants, identified originally from patient with insulin resistance syndrome, or inhibition of insulin and IGF-1 receptors via pharmacological reagents increases ADDL levels by exacerbating their aggregation. In healthy cells, activation of insulin and IGF-1 receptor reduces the extracellular ADDLs applied to cells via seemingly the insulin-degrading enzyme activity. Although insulin triggers ADDL internalization, IGF-1 appears to keep ADDLs on the cell surface. Nevertheless, both insulin and IGF-1 reduce ADDL binding, protect synapses from ADDL synaptotoxic effects, and prevent the ADDL-induced surface insulin receptor loss. Our results suggest that dysfunctions of brain insulin and IGF-1 receptors contribute to Aβ aggregation and subsequent synaptic loss.Abnormal protein misfolding and aggregation are common features in neurodegenerative diseases such as Alzheimer (AD),2 Parkinson, Huntington, and prion diseases (13). In the AD brain, intracellular accumulation of hyperphosphorylated Tau aggregates and extracellular amyloid deposits comprise the two major pathological hallmarks of the disease (1, 4). Aβ aggregation has been shown to initiate from Aβ1–42, a peptide normally cleaved from the amyloid precursor protein (APP) via activities of α- and γ-secretases (5, 6). A large body of evidence in the past decade has indicated that accumulated soluble oligomers of Aβ1–42, likely the earliest or intermediate forms of Aβ deposition, are potently toxic to neurons. The toxic effects of Aβ oligomers include synaptic structural deterioration (7, 8) and functional deficits such as inhibition of synaptic transmission (9) and synaptic plasticity (1013), as well as memory loss (11, 14, 15). Accumulation of high levels of these oligomers may also trigger inflammatory processes and oxidative stress in the brain probably due to activation of astrocytes and microglia (16, 17). Thus, to understand how a physiologically produced peptide becomes a misfolded toxin has been one of the key issues in uncovering the molecular pathogenesis of the disease.Aβ accumulation and aggregation could derive from overproduction or impaired clearance. Mutations of APP or presenilins 1 and 2, for example, are shown to cause overproduction of Aβ1–42 and amyloid deposits in the brain of early onset AD (18, 19). Because early onset AD accounts for less than 5% of entire AD population, APP and presenilin mutations cannot represent a universal mechanism for accumulation/aggregation of Aβ in the majority of AD cases. With respect to clearance, Aβ is normally removed by both global and local mechanisms, with the former requiring vascular transport across the blood-brain barrier (20, 21) and the latter via local enzymatic digestions by several metalloproteases, including neprilysin, insulin-degrading enzyme (IDE), and endothelin converting enzymes 1 and 2 (2224).The fact that insulin is a common substrate for most of the identified Aβ-degrading enzymes has drawn attention of investigators to roles of insulin signaling in Aβ clearance. Increases in insulin levels frequently seen in insulin resistance may compete for these enzymes and thus contribute to Aβ accumulation. Indeed, insulin signaling has been shown to regulate expression of metalloproteases such as IDE (25, 26), and influence aspects of Aβ metabolism and catabolism (27). In the endothelium of the brain-blood barrier and glial cells, insulin signaling is reported to regulate protein-protein interactions in an uptake cascade involving low density lipoprotein receptor-related protein and its ligands ApoE and α2-macroglobulin, a system known to bind and clear Aβ via endocytosis and/or vascular transport (28, 29). Similarly, circulating IGF-1 has been reported to play a role in Aβ clearance probably via facilitating brain-blood barrier transportation (30, 31).In the brain, insulin signaling plays a role in learning and memory (3234), potentially linking insulin resistance to AD dementia. Recently we and others have shown that Aβ oligomers interact with neuronal insulin receptors to cause impairments of the receptor expression and function (3537). These impairments mimic the Aβ oligomer-induced synaptic long term potentiation inhibition and can be overcome by insulin treatment (35, 38). Consistently, impairments of both IR and IGF-1R have been reported in the AD brain (3941).Based on these results, we ask whether impairment of insulin and IGF-1 signaling contribute to Aβ oligomer build-up in brain cells. To address this question, we set out to test roles of IR and IGF-1R in cellular clearance and transport of Aβ oligomers (ADDLs) applied to primary neuronal cultures and cell lines overexpressing IR and IGF-1R. Our results show that insulin and IGF-1 receptors function to reduce Aβ oligomers to monomers, and prevent Aβ oligomer-induced synaptic toxicity both at the level of synapse composition and structure. By contrast, receptor impairments resulting from “kinase-dead” insulin receptor mutations, a tyrosine kinase inhibitor of the insulin and IGF-1 receptor, or an inhibitory IGF-1 receptor antibody increase ADDL aggregation in the extracellular medium. Our results provide cellular evidence linking insulin and IGF-1 signaling to amyloidogenesis.  相似文献   

15.
Protein kinase A-anchoring proteins (AKAPs) play important roles in the compartmentation of cAMP signaling, anchoring protein kinase A (PKA) to specific cellular organelles and serving as scaffolds that assemble localized signaling cascades. Although AKAPs have been recently shown to bind adenylyl cyclase (AC), the functional significance of this association has not been studied. In cardiac myocytes, the muscle protein kinase A-anchoring protein β (mAKAPβ) coordinates cAMP-dependent, calcium, and MAP kinase pathways and is important for cellular hypertrophy. We now show that mAKAPβ selectively binds type 5 AC in the heart and that mAKAPβ-associated AC activity is absent in AC5 knock-out hearts. Consistent with its known inhibition by PKA phosphorylation, AC5 is inhibited by association with mAKAPβ-PKA complexes. AC5 binds to a unique N-terminal site on mAKAP-(245–340), and expression of this peptide disrupts endogenous mAKAPβ-AC association. Accordingly, disruption of mAKAPβ-AC5 complexes in neonatal cardiac myocytes results in increased cAMP and hypertrophy in the absence of agonist stimulation. Taken together, these results show that the association of AC5 with the mAKAPβ complex is required for the regulation of cAMP second messenger controlling cardiac myocyte hypertrophy.The formation of multimolecular protein complexes contributes to the specificity of intracellular signaling pathways, including those regulating cardiac myocyte hypertrophy. The cAMP-dependent protein kinase (PKA)3 is targeted to specific intracellular domains by protein kinase A-anchoring proteins (AKAPs) that often serve as scaffolding proteins for diverse signaling enzymes (1). In the heart, global disruption of PKA anchoring affects cardiac contractility, while the inhibited expression of individual AKAPs such as mAKAPβ or AKAP-Lbc attenuates adrenergic-induced hypertrophy of cultured neonatal myocytes (24). We have recently shown that specific AKAPs, namely AKAP79 and Yotiao, bind adenylyl cyclases (AC) (5, 6). However, the functional significance of AC-AKAP complexes has not been demonstrated.mAKAPβ, expressed in striated myocytes, is one of two known splice variants encoded by the single mAKAP (AKAP6) gene (7). We previously published that mAKAPβ is primarily localized to the outer membrane of the nuclear envelope via direct binding to nesprin-1α (4, 8). In cardiac myocytes, mAKAPβ serves as the scaffold for a multimolecular signaling complex that in addition to PKA includes the ryanodine receptor (RyR2), the protein phosphatases PP2A and calcineurin, phosphodiesterase 4D3 (PDE4D3), exchange protein activated by cAMP (Epac1), ERK5, and MEK5 mitogen-activated protein kinases, molecules implicated in the regulation of cardiac hypertrophy (4, 713). mAKAPβ complexes facilitate cross-talk between MAP kinase, calcium, and cAMP signaling pathways, permitting feedback inhibition of cAMP levels and the dynamic regulation of PKA and ERK5 activity (4, 913). Accordingly, mAKAPβ RNAi attenuates adrenergic and cytokine-induced hypertrophy of cultured rat neonatal ventricular myocytes (4, 11).Because mAKAPβ forms a complex with two cAMP effectors and a metabolizing enzyme for cAMP, we considered whether AC might also be an integral part of the mAKAPβ complex. We now demonstrate that type 5 adenylyl cyclase (AC5) binds directly a unique N-terminal site on mAKAPβ and is the predominant AC isoform associated with mAKAPβ in the heart. We show that AC5 bound to mAKAPβ is inhibited by PKA-dependent negative feedback. Importantly, inhibition of endogenous mAKAPβ-AC5 binding revealed the functional importance of these complexes for the regulation of cAMP-dependent myocyte hypertrophy.  相似文献   

16.
17.
18.
19.
Paneth cells are a secretory epithelial lineage that release dense core granules rich in host defense peptides and proteins from the base of small intestinal crypts. Enteric α-defensins, termed cryptdins (Crps) in mice, are highly abundant in Paneth cell secretions and inherently resistant to proteolysis. Accordingly, we tested the hypothesis that enteric α-defensins of Paneth cell origin persist in a functional state in the mouse large bowel lumen. To test this idea, putative Crps purified from mouse distal colonic lumen were characterized biochemically and assayed in vitro for bactericidal peptide activities. The peptides comigrated with cryptdin control peptides in acid-urea-PAGE and SDS-PAGE, providing identification as putative Crps. Matrix-assisted laser desorption ionization time-of-flight mass spectrometry experiments showed that the molecular masses of the putative α-defensins matched those of the six most abundant known Crps, as well as N-terminally truncated forms of each, and that the peptides contain six Cys residues, consistent with identities as α-defensins. N-terminal sequencing definitively revealed peptides with N termini corresponding to full-length, (des-Leu)-truncated, and (des-Leu-Arg)-truncated N termini of Crps 1–4 and 6. Crps from mouse large bowel lumen were bactericidal in the low micromolar range. Thus, Paneth cell α-defensins secreted into the small intestinal lumen persist as intact and functional forms throughout the intestinal tract, suggesting that the peptides may mediate enteric innate immunity in the colonic lumen, far from their upstream point of secretion in small intestinal crypts.Antimicrobial peptides (AMPs)2 are released by epithelial cells onto mucosal surfaces as effectors of innate immunity (15). In mammals, most AMPs derive from two major families, the cathelicidins and defensins (6). The defensins comprise the α-, β-, and θ-defensin subfamilies, which are defined by the presence of six cysteine residues paired in characteristic tridisulfide arrays (7). α-Defensins are highly abundant in two primary cell lineages: phagocytic leukocytes, primarily neutrophils, of myeloid origin and Paneth cells, which are secretory epithelial cells located at the base of the crypts of Lieberkühn in the small intestine (810). Neutrophil α-defensins are stored in azurophilic granules and contribute to non-oxidative microbial cell killing in phagolysosomes (11, 12), except in mice whose neutrophils lack defensins (13). In the small bowel, α-defensins and other host defense proteins (1418) are released apically as components of Paneth cell secretory granules in response to cholinergic stimulation and after exposure to bacterial antigens (19). Therefore, the release of Paneth cell products into the crypt lumen is inferred to protect mitotically active crypt cells from colonization by potential pathogens and confer protection against enteric infection (7, 20, 21).Under normal, homeostatic conditions, Paneth cells are not found outside the small bowel, although they may appear ectopically in response to local inflammation throughout the gastrointestinal tract (22, 23). Paneth cell numbers increase progressively throughout the small intestine, occurring at highest numbers in the distal ileum (24). Mouse Paneth cells express numerous α-defensin isoforms, termed cryptdins (Crps) (25), that have broad spectrum antimicrobial activities (6, 26). Collectively, α-defensins constitute approximately seventy percent of the bactericidal peptide activity in mouse Paneth cell secretions (19), selectively killing bacteria by membrane-disruptive mechanisms (2730). The role of Paneth cell α-defensins in gastrointestinal mucosal immunity is evident from studies of mice transgenic for human enteric α-defensin-5, HD-5, which are immune to infection by orally administered Salmonella enterica sv. typhimurium (S. typhimurium) (31).The biosynthesis of mature, bactericidal α-defensins from their inactive precursors requires activation by lineage-specific proteolytic convertases. In mouse Paneth cells, inactive ∼8.4-kDa Crp precursors are processed intracellularly into microbicidal ∼4-kDa Crps by specific cleavage events mediated by matrix metalloproteinase-7 (MMP-7) (32, 33). MMP-7 null mice exhibit increased susceptibility to systemic S. typhimurium infection and decreased clearance of orally administered non-invasive Escherichia coli (19, 32). Although the α-defensin proregions are sensitive to proteolysis, the mature, disulfide-stabilized peptides resist digestion by their converting enzymes in vitro, whether the convertase is MMP-7 (32), trypsin (34), or neutrophil serine proteinases (35). Because α-defensins resist proteolysis in vitro, we hypothesized that Paneth cell α-defensins resist degradation and remain in a functional state in the large bowel, a complex, hostile environment containing varied proteases of both host and microbial origin.Here, we report on the isolation and characterization of a population of enteric α-defensins from the mouse colonic lumen. Full-length and N-terminally truncated Paneth cell α-defensins were identified and are abundant in the distal large bowel lumen.  相似文献   

20.
Differential inhibitors of Gβγ-effector regions are required to dissect the biological contribution of specific Gβγ-initiated signaling pathways. Here, we characterize PhLP-M1-G149, a Gβγ-interacting construct derived from phosducin-like protein 1 (PhLP) as a differential inhibitor of Gβγ, which, in endothelial cells, prevented sphingosine 1-phosphate-induced phosphorylation of AKT, glycogen synthase kinase 3β, cell migration, and tubulogenesis, while having no effect on ERK phosphorylation or hepatocyte growth factor-dependent responses. This construct attenuated the recruitment of phosphoinositide 3-kinase γ (PI3Kγ) to the plasma membrane and the signaling to AKT in response to Gβγ overexpression. In coimmunoprecipitation experiments, PhLP-M1-G149 interfered with the interaction between PI3Kγ and Gβγ. Other PhLP-derived constructs interacted with Gβγ but were not effective inhibitors of Gβγ signaling to AKT or ERK. Our results indicate that PhLP-M1-G149 is a suitable tool to differentially modulate the Gβγ-initiated pathway linking this heterodimer to AKT, endothelial cell migration, and in vitro angiogenesis. It can be also useful to further characterize the molecular determinants of the Gβγ-PI3Kγ interaction.Heterotrimeric G protein signaling depends on the actions of GTP-loaded Gα and free Gβγ, the two functional components of the heterotrimer, leading to the generation of second messengers and cell specific functional events (1, 2). Differential inhibitors of Gβγ are required to dissect the biological impact of different Gβγ-dependent effectors. Gβγ actions can be blocked by competition with peptides derived from its effectors. For example, the effect of Gβγ on adenylyl cyclase II, G protein-activated inward rectifier K+ channel, G protein-coupled receptor kinase 2, and phospholipase Cβ3, is attenuated by a peptide from adenylyl cyclase II (3). In addition, RACK1 (receptor for activated C kinase 1) selectively inhibits the effect the chemokine receptor CXCR2 on the activation of phospholipase Cβ2 and adenylyl cyclase II in HEK293 cells, without affecting other functions of Gβγ (4). Recently, Smrcka and colleagues characterized the effect of small molecule inhibitors of Gβγ, suggesting their potential application in therapeutic strategies targeting particular Gβγ-dependent pathways (5). Emerging possibilities to target this heterodimer in pathological situations such as inflammation and angiogenesis are based on the role of Gβγ in cell survival and chemotaxis. To the best of our knowledge, no molecular tool is yet available to differentially inhibit Gβγ signaling to AKT.3Gβγ is a key transducer of sphingosine 1-phosphate (S1P)-elicited angiogenic signals promoting endothelial cell migration, proliferation, and survival (612). Multiple Gβγ-dependent effectors are potentially involved in the molecular events required for endothelial cell migration. These include lipid kinases such as PI3Kγ and PI3Kβ (13), and a novel family of Rac guanine nucleotide exchange factors, represented by P-REX1, which is activated by Gβγ and phosphatidylinositol 3,4,5-trisphosphate (1416). Gβγ signaling is frequently attributed to pertussis toxin-sensitive Gi coupled receptors, and it has been consistently revealed by the antagonistic effect of the carboxyl-terminal region of G protein-coupled receptor kinase 2, which sequesters Gβγ thereby inhibiting all its intracellular actions (17). In addition, mutational analysis of Gβ revealed that different residues, all of them mapping to the interface of contact between Gβγ and Gα, are important for the activation of distinct Gβγ effector molecules (18).Phosducin was originally identified as a phosphoprotein restricted to the retina and pineal gland forming a complex with Gβγ (19, 20). It was considered a protein kinase A-sensitive regulator of G protein-mediated signaling (21, 22). Further studies identified a family of phosducin-like proteins (PhLPs) (23, 24). Phosducin and Gα share affinity for the same region of Gβγ, as revealed by the structural analysis of Gβγ in complex with Gα or phosducin and by in vitro binding experiments (25). This area of interaction includes some of the residues considered necessary for the activation of Gβγ-dependent effectors (18, 26). It was initially postulated that phosducin and related proteins, by interfering with the availability of free Gβγ, exert an inhibitory role on Gβγ signaling. However, recent genetic evidence raised an apparently conflicting situation; the knockout of PhLP in fungi resulted in a phenotype equivalent to the absence of Gβγ, contrary to its expected role as an inhibitor (27). Novel experimental evidence indicated that PhLP has a positive effect on Gβγ signaling due to its participation in the assembly of the heterodimer, helping to stabilize free Gβ subunits leaving the ribosome after synthesis (2831).Despite the positive role of full-length PhLP in the assembly of Gβγ heterodimers, it is still possible that different fragments of this protein, which could retain their interaction with distinct regions of Gβγ, might function as inhibitors of Gβγ signaling. Accordingly, we characterized here the effect of different PhLP-derived constructs on the signaling pathways elicited by S1P or HGF in endothelial cells. In addition, we explored the mechanism by which PhLP-M1-G149 interferes with Gβγ preventing the activation of AKT.  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