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 46 毫秒
1.
2.
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   

3.
4.
5.
6.
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   

7.
8.
Redox-active copper is implicated in the pathogenesis of Alzheimer disease (AD), β-amyloid peptide (Aβ) aggregation, and amyloid formation. Aβ·copper complexes have been identified in AD and catalytically oxidize cholesterol and lipid to generate H2O2 and lipid peroxides. The site and mechanism of this abnormality is not known. Growing evidence suggests that amyloidogenic processing of the β-amyloid precursor protein (APP) occurs in lipid rafts, membrane microdomains enriched in cholesterol. β- and γ-secretases, and Aβ have been identified in lipid rafts in cultured cells, human and rodent brains, but the role of copper in lipid raft amyloidogenic processing is presently unknown. In this study, we found that copper modulates flotillin-2 association with cholesterol-rich lipid raft domains, and consequently Aβ synthesis is attenuated via copper-mediated inhibition of APP endocytosis. We also found that total cellular copper is associated inversely with lipid raft copper levels, so that under intracellular copper deficiency conditions, Aβ·copper complexes are more likely to form. This explains the paradoxical hypermetallation of Aβ with copper under tissue copper deficiency conditions in AD.Imbalance of metal ions has been recognized as one of the key factors in the pathogenesis of Alzheimer disease (AD).2 Aberrant interactions between copper or zinc with the β-amyloid peptide (Aβ) released into the glutamatergic synaptic cleft vicinity could result in the formation of toxic Aβ oligomers and aggregation into plaques characteristic of AD brains (reviewed in Ref. 1). Copper, iron, and zinc are highly concentrated in extracellular plaques (2, 3), and yet brain tissues from AD (46) and human β-amyloid precursor protein (APP) transgenic mice (710) are paradoxically copper deficient compared with age-matched controls. Elevation of intracellular copper levels by genetic, dietary, and pharmacological manipulations in both AD transgenic animal and cell culture models is able to attenuate Aβ production (7, 9, 1115). However, the underlying mechanism is at present unclear.Abnormal cholesterol metabolism is also a contributing factor in the pathogenesis of AD. Hypercholesterolemia increases the risk of developing AD-like pathology in a transgenic mouse model (16). Epidemiological and animal model studies show that a hypercholesterolemic diet is associated with Aβ accumulation and accelerated cognitive decline, both of which are further aggravated by high dietary copper (17, 18). In contrast, biochemical depletion of cholesterol using statins, inhibitors of 3-hydroxy-3-methyglutaryl coenzyme A reductase, and methyl-β-cyclodextrin, a cholesterol sequestering agent, inhibit Aβ production in animal and cell culture models (1925).Cholesterol is enriched in lipid rafts, membrane microdomains implicated in Aβ generation from APP cleavage by β- and γ-secretases. Recruitment of BACE1 (β-secretase) into lipid rafts increases the production of sAPPβ and Aβ (23, 26). The β-secretase-cleaved APP C-terminal fragment (β-CTF), and γ-secretase, a multiprotein complex composed of presenilin (PS1 or PS2), nicastrin (Nct), PEN-2 and APH-1, colocalize to lipid rafts (27). The accumulation of Aβ in lipid rafts isolated from AD and APP transgenic mice brains (28) provided further evidence that cholesterol plays a role in APP processing and Aβ generation.Currently, copper and cholesterol have been reported to modulate APP processing independently. However, evidence indicates that, despite tissue copper deficiency, Aβ·Cu2+ complexes form in AD that catalytically oxidize cholesterol and lipid to generate H2O2 and lipid peroxides (e.g. hydroxynonenal and malondialdehyde), which contribute to oxidative damage observed in AD (2935). The underlying mechanism leading to the formation of pathological Aβ·Cu2+ complexes is unknown. In this study, we show that copper alters the structure of lipid rafts, and attenuates Aβ synthesis in lipid rafts by inhibition of APP endocytosis. We also identify a paradoxical inverse relationship between total cellular copper levels and copper distribution to lipid rafts, which appear to possess a privileged pool of copper where Aβ is more likely to interact with Cu2+ under copper-deficiency conditions to form Aβ·Cu2+ complexes. These data provide a novel mechanism by which cellular copper deficiency in AD could foster an environment for potentially adverse interactions between Aβ, copper, and cholesterol in lipid rafts.  相似文献   

9.
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   

10.
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   

11.
12.
13.
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   

14.
15.
16.
17.
18.
AMP-activated protein kinase (AMPK) is a heterotrimeric complex playing a crucial role in maintaining cellular energy homeostasis. Recently, homodimerization of mammalian AMPK and yeast ortholog SNF1 was shown by us and others. In SNF1, it involved specific hydrophobic residues in the kinase domain αG-helix. Mutation of the corresponding AMPK α-subunit residues (Val-219 and Phe-223) to glutamate reduced the tendency of the kinase to form higher order homo-oligomers, as was determined by the following three independent techniques in vitro: (i) small angle x-ray scattering, (ii) surface plasmon resonance spectroscopy, and (iii) two-dimensional blue native/SDS-PAGE. Recombinant protein as well as AMPK in cell lysates of primary cells revealed distinct complexes of various sizes. In particular, the assembly of very high molecular mass complexes was dependent on both the αG-helix-mediated hydrophobic interactions and kinase activation. In vitro and when overexpressed in double knock-out (α1−/−, α2−/−) mouse embryonic fibroblast cells, activation of mutant AMPK was impaired, indicating a critical role of the αG-helix residues for AMPK activation via its upstream kinases. Also inactivation by protein phosphatase 2Cα was affected in mutant AMPK. Importantly, activation of mutant AMPK by LKB1 was restored by exchanging the corresponding and conserved hydrophobic αG-helix residues of LKB1 (Ile-260 and Phe-264) to positively charged amino acids. These mutations functionally rescued LKB1-dependent activation of mutant AMPK in vitro and in cell culture. Our data suggest a physiological role for the hydrophobic αG-helix residues in homo-oligomerization of heterotrimers and cellular interactions, in particular with upstream kinases, indicating an additional level of AMPK regulation.The maintenance of energy homeostasis is a basic requirement of all living organisms. The AMP-activated protein kinase (AMPK)2 is crucially involved in this essential process by playing a central role in sensing and regulating energy metabolism on the cellular and whole body level (16). AMPK is also participating in several signaling pathways associated with cancer and metabolic diseases, like type 2 diabetes mellitus, obesity, and other metabolic disorders (79).Mammalian AMPK belongs to a highly conserved family of serine/threonine protein kinases with homologs found in all eukaryotic organisms examined (1, 3, 10). Its heterotrimeric structure includes a catalytic α-subunit and regulatory β- and γ-subunits. These subunits exist in different isoforms (α1, α2, β1, β2, γ1, γ2, and γ3) and splice variants (for γ2 and γ3) and can thus assemble to a broad variety of heterotrimeric isoform combinations. The α- and β-subunits possess multiple autophosphorylation sites, which have been implicated in regulation of subcellular localization and kinase activation (1115). The most critical step of AMPK activation, however, is phosphorylation of Thr-172 within the activation segment of the α-subunit kinase domain. At least two AMPK upstream kinases (AMPKKs) have been identified so far, namely the tumor suppressor kinase LKB1 in complex with MO25 and STRAD (16) and Ca2+/calmodulin-dependent protein kinase kinase-2 (CamKK2) (17). Furthermore, the transforming growth factor-β-activated kinase 1 was also shown to activate AMPK using a variety of in vitro approaches (18), but the physiological relevance of these findings remains unclear. Besides direct phosphorylation of Thr-172, AMPK activity is stimulated by the allosteric activator AMP, which can bind to two Bateman domains formed by two pairs of CBS domains within the γ-subunit (1922). Hereby bound AMP not only allosterically stimulates AMPK but also protects Thr-172 from dephosphorylation by protein phosphatase 2Cα (PP2Cα) and thus hinders inactivation of the kinase (19, 22, 23). Consequently, on the cellular level, AMPK is activated upon metabolic stress increasing the AMP/ATP ratio. Furthermore, AMPK activation can also be induced by several chemical compounds, like nucleoside 5-aminoimidazole-4-carboxamide-1-β-d-ribofuranoside (24) and the anti-diabetic drug Metformin (2528). In addition, the small molecule compound A-769662 was recently developed as a direct allosteric activator of AMPK (29, 30).Previous work in our groups proposed a model of AMPK regulation by AMP, which incorporates the major functional features and the latest structural information (31). The latter mainly included truncated core complexes of AMPK from different species (3235). Further valuable structural information is provided by the x-ray structures of the isolated catalytic domains, in particular of the human AMPK α2-subunit (Protein Data Bank code 2H6D) and its yeast ortholog SNF1 (36, 37). The kinase domain of SNF1 is capable of forming homodimers in the protein crystal, as well as in vitro in solution, in a unique way, which has not been observed previously in any other kinase (36). The dimer interface is predominantly formed by hydrophobic interactions of the loop-αG region, also known as subdomain X situated on the large kinase lobe (36, 38, 39), and it mainly involves Ile-257 and Phe-261. Because the T-loop activation segment was buried within the dimer interface, it was suggested that the dimeric state of the SNF1 catalytic domain represents the inactive form of the kinase. Intriguingly, it was shown in our groups by small angle x-ray scattering that AMPK self-organizes in a concentration-dependent manner to form homo-oligomers in solution (31). However, the interface responsible for oligomerization of the AMPK heterotrimer has remained elusive.Here we further investigate the distinct oligomeric states of the AMPK heterotrimer and suggest a possible regulatory function for this process. Most importantly, we provide conclusive evidence for participation of αG-helix residues in the recognition of AMPK by its upstream kinases LKB1 and CamKK2.  相似文献   

19.
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   

20.
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