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.
3.
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   

4.
Leptospira spp., the causative agents of leptospirosis, adhere to components of the extracellular matrix, a pivotal role for colonization of host tissues during infection. Previously, we and others have shown that Leptospira immunoglobulin-like proteins (Lig) of Leptospira spp. bind to fibronectin, laminin, collagen, and fibrinogen. In this study, we report that Leptospira can be immobilized by human tropoelastin (HTE) or elastin from different tissues, including lung, skin, and blood vessels, and that Lig proteins can bind to HTE or elastin. Moreover, both elastin and HTE bind to the same LigB immunoglobulin-like domains, including LigBCon4, LigBCen7′–8, LigBCen9, and LigBCen12 as demonstrated by enzyme-linked immunosorbent assay (ELISA) and competition ELISAs. The LigB immunoglobulin-like domain binds to the 17th to 27th exons of HTE (17–27HTE) as determined by ELISA (LigBCon4, KD = 0.50 μm; LigBCen7′–8, KD = 0.82 μm; LigBCen9, KD = 1.54 μm; and LigBCen12, KD = 0.73 μm). The interaction of LigBCon4 and 17–27HTE was further confirmed by steady state fluorescence spectroscopy (KD = 0.49 μm) and ITC (KD = 0.54 μm). Furthermore, the binding was enthalpy-driven and affected by environmental pH, indicating it is a charge-charge interaction. The binding affinity of LigBCon4D341N to 17–27HTE was 4.6-fold less than that of wild type LigBCon4. In summary, we show that Lig proteins of Leptospira spp. interact with elastin and HTE, and we conclude this interaction may contribute to Leptospira adhesion to host tissues during infection.Pathogenic Leptospira spp. are spirochetes that cause leptospirosis, a serious infectious disease of people and animals (1, 2). Weil syndrome, the severe form of leptospiral infection, leads to multiorgan damage, including liver failure (jaundice), renal failure (nephritis), pulmonary hemorrhage, meningitis, abortion, and uveitis (3, 4). Furthermore, this disease is not only prevalent in many developing countries, it is reemerging in the United States (3). Although leptospirosis is a serious worldwide zoonotic disease, the pathogenic mechanisms of Leptospira infection remain enigmatic. Recent breakthroughs in applying genetic tools to Leptospira may facilitate studies on the molecular pathogenesis of leptospirosis (58).The attachment of pathogenic Leptospira spp. to host tissues is critical in the early phase of Leptospira infection. Leptospira spp. adhere to host tissues to overcome mechanical defense systems at tissue surfaces and to initiate colonization of specific tissues, such as the lung, kidney, and liver. Leptospira invade hosts tissues through mucous membranes or injured epidermis, coming in contact with subepithelial tissues. Here, certain bacterial outer surface proteins serve as microbial surface components recognizing adhesive matrix molecules (MSCRAMMs)2 to mediate the binding of bacteria to different extracellular matrices (ECMs) of host cells (9). Several leptospiral MSCRAMMs have been identified (1018), and we speculate that more will be identified in the near future.Lig proteins are distributed on the outer surface of pathogenic Leptospira, and the expression of Lig protein is only found in low passage strains (14, 16, 17), probably induced by environmental cues such as osmotic or temperature changes (19). Lig proteins can bind to fibrinogen and a variety of ECMs, including fibronectin (Fn), laminin, and collagen, thereby mediating adhesion to host cells (2023). Lig proteins also constitute good vaccine candidates (2426).Elastin is a component of ECM critical to tissue elasticity and resilience and is abundant in skin, lung, blood vessels, placenta, uterus, and other tissues (2729). Tropoelastin is the soluble precursor of elastin (28). During the major phase of elastogenesis, multiple tropoelastin molecules associate through coacervation (3032). Because of the abundance of elastin or tropoelastin on the surface of host cells, several bacterial MSCRAMMs use elastin and/or tropoelastin to mediate adhesion during the infection process (3335).Because leptospiral infection is known to cause severe pulmonary hemorrhage (36, 37) and abortion (38), we hypothesize that some leptospiral MSCRAMMs may interact with elastin and/or tropoelastin in these elastin-rich tissues. This is the first report that Lig proteins of Leptospira interact with elastin and tropoelastin, and the interactions are mediated by several specific immunoglobulin-like domains of Lig proteins, including LigBCon4, LigBCen7′–8, LigBCen9, and LigBCen12, which bind to the 17th to 27th exons of human tropoelastin (HTE).  相似文献   

5.
According to the lipid raft theory, the plasma membrane contains small domains enriched in cholesterol and sphingolipid, which may serve as platforms to organize membrane proteins. Using methyl-β-cyclodextrin (MβCD) to deplete membrane cholesterol, many G protein-coupled receptors have been shown to depend on putative lipid rafts for proper signaling. Here we examine the hypothesis that treatment of HEK293 cells stably expressing FLAG-tagged μ-opioid receptors (HEK FLAG-μ) or δ-opioid receptors (HEK FLAG-δ) with MβCD will reduce opioid receptor signaling to adenylyl cyclase. The ability of the μ-opioid agonist [d-Ala2,N-Me-Phe4,Gly5-ol]enkephalin to acutely inhibit adenylyl cyclase or to cause sensitization of adenylyl cyclase following chronic treatment was attenuated with MβCD. These effects were due to removal of cholesterol, because replenishment of cholesterol restored [d-Ala2,N-Me-Phe4,Gly5-ol]enkephalin responses back to control values, and were confirmed in SH-SY5Y cells endogenously expressing μ-opioid receptors. The effects of MβCD may be due to uncoupling of the μ receptor from G proteins but were not because of decreases in receptor number and were not mimicked by cytoskeleton disruption. In contrast to the results in HEK FLAG-μ cells, MβCD treatment of HEK FLAG-δ cells had no effect on acute inhibition or sensitization of adenylyl cyclase by δ-opioid agonists. The differential responses of μ- and δ-opioid agonists to cholesterol depletion suggest that μ-opioid receptors are more dependent on cholesterol for efficient signaling than δ receptors and can be partly explained by localization of μ- but not δ-opioid receptors in cholesterol- and caveolin-enriched membrane domains.Membrane cholesterol can alter the function of integral proteins, such as G protein-coupled receptors, through cholesterol-protein interactions and by changes in membrane viscosity (1). In addition, cholesterol interacts with other lipids found in the bilayer, particularly sphingolipids (2), which allows for tight and organized packing that can precipitate the formation of specialized domains within the plasma membrane (3). These domains have become an area of intense research interest and have been termed lipid or membrane rafts (4). The study of membrane rafts in intact cells is controversial, due in part to the limitations of the current methods used to study rafts (5, 6). Regardless, the membrane environment formed in regions of high cholesterol and sphingolipids may be such that certain proteins have an affinity for these regions, especially proteins with a propensity to interact with cholesterol.Many G protein-coupled receptors and signaling proteins have been found to prefer cholesterol-enriched domains leading to the hypothesis that these domains can organize signaling molecules in the membrane to enhance or inhibit specific signaling events (7). This includes μ- (8, 9), δ- (10, 11), and κ-opioid receptors (12). In addition, Gαi (1217), Gαo (16), and adenylyl cyclase isoforms 3 (18), 5/6 (9, 18, 19), and 8 (20) have been found to associate with cholesterol and/or the cholesterol-binding protein caveolin. Activated opioid receptors couple to Gαi/o proteins and acutely inhibit the activity of adenylyl cyclase. Longer term exposure to opioid agonists causes sensitization of adenylyl cyclase and a rebound overshoot of cAMP production upon withdrawal of the agonist (21). Consequently, we sought to assess the role of cholesterol depletion on the ability of μ- and δ-opioid receptor agonists to inhibit and cause sensitization of adenylyl cyclase.There are conflicting data for the effect of changes in membrane cholesterol on opioid signaling. For example, an increase in plasma membrane microviscosity by addition of cholesteryl hemisuccinate to SH-SY5Y cell membranes increased μ-opioid receptor coupling to G proteins (22). Conversely, removal of membrane cholesterol from Chinese hamster ovary cells has been shown to either decrease (23) or increase (24) the coupling of μ-opioid receptors to G proteins, as measured by [35S]GTPγS3 binding stimulated by the μ-opioid agonist DAMGO. Furthermore, the effect of cholesterol removal on δ-opioid agonist-stimulated [35S]GTPγS binding varies by cell type (10, 25). In these previous studies, the variety of cell types utilized and the conflicting results make comparisons between opioid receptor types difficult. The objective of this study was to directly compare the role of membrane cholesterol in modulating acute and chronic μ- and δ-opioid signaling in the same cell systems using identical methods, including the following: 1) depletion of cholesterol by the cholesterol-sequestering agent methyl-β-cyclodextrin (MβCD); 2) separation of cholesterol-enriched membranes by sucrose gradient ultracentrifugation; and 3) clustering of lipid raft patches in whole cells with cholera toxin B subunit.In initial experiments using human embryonic kidney (HEK) cells heterologously expressing μ- or δ-opioid receptors, we found that δ-opioid receptors were located in caveolin-poor fractions following 1% Triton X-100 homogenization and sucrose gradient ultracentrifugation. This differs from studies using a detergent-free method to identify lipid raft fractions (10, 11). In contrast, we found that the μ-opioid receptor was found in both caveolin-poor and caveolin-rich fractions, in accordance with previous literature (8, 9). This differential localization of opioid receptors led us to test the hypothesis that, in contrast to the μ-opioid receptor, the δ-opioid receptor would not be dependent on cholesterol for signaling. The results show that μ- but not δ-opioid receptors have a dependence on cholesterol for signaling to adenylyl cyclase and that this effect is much more pronounced following chronic exposure to opioids.  相似文献   

6.
Codon optimization was used to synthesize the blh gene from the uncultured marine bacterium 66A03 for expression in Escherichia coli. The expressed enzyme cleaved β-carotene at its central double bond (15,15′) to yield two molecules of all-trans-retinal. The molecular mass of the native purified enzyme was ∼64 kDa as a dimer of 32-kDa subunits. The Km, kcat, and kcat/Km values for β-carotene as substrate were 37 μm, 3.6 min−1, and 97 mm−1 min−1, respectively. The enzyme exhibited the highest activity for β-carotene, followed by β-cryptoxanthin, β-apo-4′-carotenal, α-carotene, and γ-carotene in decreasing order, but not for β-apo-8′-carotenal, β-apo-12′-carotenal, lutein, zeaxanthin, or lycopene, suggesting that the presence of one unsubstituted β-ionone ring in a substrate with a molecular weight greater than C35 seems to be essential for enzyme activity. The oxygen atom of retinal originated not from water but from molecular oxygen, suggesting that the enzyme was a β-carotene 15,15′-dioxygenase. Although the Blh protein and β-carotene 15,15′-monooxygenases catalyzed the same biochemical reaction, the Blh protein was unrelated to the mammalian β-carotene 15,15′-monooxygenases as assessed by their different properties, including DNA and amino acid sequences, molecular weight, form of association, reaction mechanism, kinetic properties, and substrate specificity. This is the first report of in vitro characterization of a bacterial β-carotene-cleaving enzyme.Vitamin A (retinol) is a fat-soluble vitamin and important for human health. In vivo, the cleavage of β-carotene to retinal is an important step of vitamin A synthesis. The cleavage can proceed via two different biochemical pathways (1, 2). The major pathway is a central cleavage catalyzed by mammalian β-carotene 15,15′-monooxygenases (EC 1.14.99.36). β-Carotene is cleaved by the enzyme symmetrically into two molecules of all-trans-retinal, and retinal is then converted to vitamin A in vivo (35). The second pathway is an eccentric cleavage that occurs at double bonds other than the central 15,15′-double bond of β-carotene to produce β-apo-carotenals with different chain lengths, which are catalyzed by carotenoid oxygenases from mammals, plants, and cyanobacteria (6). These β-apo-carotenals are degraded to one molecule of retinal, which is subsequently converted to vitamin A in vivo (2).β-Carotene 15,15′-monooxygenase was first isolated as a cytosolic enzyme by identifying the product of β-carotene cleavage as retinal (7). The characterization of the enzyme and the reaction pathway from β-carotene to retinal were also investigated (4, 8). The enzyme activity has been found in mammalian intestinal mucosa, jejunum enterocytes, liver, lung, kidney, and brain (5, 9, 10). Molecular cloning, expression, and characterization of β-carotene 15,15′-monooxygenase have been reported from various species, including chickens (11), fruit flies (12), humans (13), mice (14), and zebra fishes (15).Other proteins thought to convert β-carotene to retinal include bacterioopsin-related protein (Brp) and bacteriorhodopsin-related protein-like homolog protein (Blh) (16). Brp protein is expressed from the bop gene cluster, which encodes the structural protein bacterioopsin, consisting of at least three genes as follows: bop (bacterioopsin), brp (bacteriorhodopsin-related protein), and bat (bacterioopsin activator) (17). brp genes were reported in Haloarcula marismortui (18), Halobacterium sp. NRC-1 (19), Halobacterium halobium (17), Haloquadratum walsbyi, and Salinibacter ruber (20). Blh protein is expressed from the proteorhodopsin gene cluster, which contains proteorhodopsin, crtE (geranylgeranyl-diphosphate synthase), crtI (phytoene dehydrogenase), crtB (phytoene synthase), crtY (lycopene cyclase), idi (isopentenyl diphosphate isomerase), and blh gene (21). Sources of blh genes were previously reported in Halobacterium sp. NRC-1 (19), Haloarcula marismortui (18), Halobacterium salinarum (22), uncultured marine bacterium 66A03 (16), and uncultured marine bacterium HF10 49E08 (21). β-Carotene biosynthetic genes crtE, crtB, crtI, crtY, ispA, and idi encode the enzymes necessary for the synthesis of β-carotene from isopentenyl diphosphate, and the Idi, IspA, CrtE, CrtB, CrtI, and CrtY proteins have been characterized in vitro (2328). Blh protein has been proposed to catalyze or regulate the conversion of β-carotene to retinal (29, 30), but there is no direct proof of the enzymatic activity.In this study, we used codon optimization to synthesize the blh gene from the uncultured marine bacterium 66A03 for expression in Escherichia coli, and we performed a detailed biochemical and enzymological characterization of the expressed Blh protein. In addition, the properties of the enzyme were compared with those of mammalian β-carotene 15,15′-monooxygenases.  相似文献   

7.
ARAP1 is a phosphatidylinositol 3,4,5-trisphosphate (PtdIns(3,4,5)P3)-dependent Arf GTPase-activating protein (GAP) with five PH domains that regulates endocytic trafficking of the epidermal growth factor receptor (EGFR). Two tandem PH domains are immediately N-terminal of the Arf GAP domain, and one of these fits the consensus sequence for PtdIns(3,4,5)P3 binding. Here, we tested the hypothesis that PtdIns(3,4,5)P3-dependent recruitment mediated by the first PH domain of ARAP1 regulates the in vivo and in vitro function of ARAP1. We found that PH1 of ARAP1 specifically bound to PtdIns(3,4,5)P3, but with relatively low affinity (≈1.6 μm), and the PH domains did not mediate PtdIns(3,4,5)P3-dependent recruitment to membranes in cells. However, PtdIns(3,4,5)P3 binding to the PH domain stimulated GAP activity and was required for in vivo function of ARAP1 as a regulator of endocytic trafficking of the EGFR. Based on these results, we propose a variation on the model for the function of phosphoinositide-binding PH domains. In our model, ARAP1 is recruited to membranes independently of PtdIns(3,4,5)P3, the subsequent production of which triggers enzymatic activity.Pleckstrin homology (PH)2 domains are a common structural motif encoded by the human genome (1, 2). Approximately 10% of PH domains bind to phosphoinositides. These PH domains are thought to mediate phosphoinositide-dependent recruitment to membranes (13). Most PH domains likely have functions other than or in addition to phosphoinositide binding. For example, PH domains have been found to bind to protein and DNA (412). In addition, some PH domains have been found to be structurally and functionally integrated with adjacent domains (13, 14). A small fraction of PH domain-containing proteins (about 9% of the human proteins) have multiple PH domains arranged in tandem, which have been proposed to function as adaptors but have only been examined in one protein (15, 16). Arf GTPase-activating proteins (GAPs) of the ARAP family are phosphatidylinositol 3,4,5-trisphosphate (PtdIns(3,4,5)P3)-dependent Arf GAPs with tandem PH domains (17, 18). The function of specific PH domains in regulating Arf GAP activity and for biologic activity has not been described.Arf GAPs are proteins that induce the hydrolysis of GTP bound to Arfs (1923). The Arf proteins are members of the Ras superfamily of GTP-binding proteins (2427). The six Arf proteins in mammals (five in humans) are divided into three classes based on primary sequence: Arf1, -2, and -3 are class 1, Arf4 and -5 are class 2, and Arf6 is class 3 (23, 24, 2729). Class 1 and class 3 Arf proteins have been studied more extensively than class 2. They have been found to regulate membrane traffic and the actin cytoskeleton.The Arf GAPs are a family of proteins with diverse domain structures (20, 21, 23, 30). ARAPs, the most structurally complex of the Arf GAPs, contain, in addition to an Arf GAP domain, the sterile α motif (SAM), five PH, Rho GAP, and Ras association domains (17, 18, 31, 32). The first and second and the third and fourth PH domains are tandem (Fig. 1). The first and third PH domains of the ARAPs fit the consensus for PtdIns(3,4,5)P3 binding (3335). ARAPs have been found to affect actin and membrane traffic (21, 23). ARAP3 regulates growth factor-induced ruffling of porcine aortic endothelial cells (31, 36, 37). The function is dependent on the Arf GAP and Rho GAP domains. ARAP2 regulates focal adhesions, an actin cytoskeletal structure (17). ARAP2 function requires Arf GAP activity and a Rho GAP domain capable of binding RhoA·GTP. ARAP1 has been found to have a role in membrane traffic (18). The protein associates with pre-early endosomes involved in the attenuation of EGFR signals. The function of the tandem PH domains in the ARAPs has not been examined.Open in a separate windowFIGURE 1.ARAP1 binding to phospholipids. A, schematic of the recombinant proteins used in this study. Domain abbreviations: Ank, ankyrin repeat; PLCδ-PH, PH domain of phospholipase C δ; RA, Ras association motif; RhoGAP, Rho GTPase-activating domain. B, ARAP1 phosphoinositide binding specificity. 500 nm PH1-Ank recombinant protein was incubated with sucrose-loaded LUVs formed by extrusion through a 1-μm pore filter. LUVs contained PtdIns alone or PtdIns with 2.5 μm PtdIns(3,4,5)P3, 2.5 μm PtdIns(3)P, 2.5 μm PtdIns(4)P, 2.5 μm PtdIns(5)P, 2.5 μm PtdIns(3,4)P2, 2.5 μm PtdIns(3,5)P2, or 2.5 μm PtdIns(4,5)P2 with a total phosphoinositide concentration of 50 μm and a total phospholipid concentration of 500 μm. Vesicles were precipitated by ultracentrifugation, and associated proteins were separated by SDS-PAGE. The amount of precipitated protein was determined by densitometry of the Coomassie Blue-stained gels with standards on each gel. C, PtdIns(3,4,5)P3-dependent binding of ARAP1 to LUVs. 1 μm PH1-Ank or ArfGAP-Ank recombinant protein was incubated with 1 mm sucrose-loaded LUVs formed by extrusion through a 1-μm pore size filter containing varying concentration of PtdIns(3,4,5)P3. Precipitation of LUVs and analysis of associated proteins were performed as described in B. The average ± S.E. of three independent experiments is presented.Here we investigated the role of the first two PH domains of ARAP1 for catalysis and in vivo function. The first PH domain fits the consensus sequence for PtdIns(3,4,5)P3 binding (3335). The second does not fit a phosphoinositide binding consensus but is immediately N-terminal to the GAP domain. We have previously reported that the PH domain that occurs immediately N-terminal of the Arf GAP domain of ASAP1 is critical for the catalytic function of the protein (38, 39). We tested the hypothesis that the two PH domains of ARAP1 function independently; one recruits ARAP1 to PtdIns(3,4,5)P3-rich membranes, and the other functions with the catalytic domain. As predicted, PH1 interacted specifically with PtdIns(3,4,5)P3, and PH2 did not. However, both PH domains contributed to catalysis independently of recruitment to membranes. None of the PH domains in ARAP1 mediated PtdIns(3,4,5)P3-dependent targeting to plasma membranes (PM). PtdIns(3,4,5)P3 stimulated GAP activity, and the ability to bind PtdIns(3,4,5)P3 was required for ARAP1 to regulate membrane traffic. We propose that ARAP1 is recruited independently of PtdIns(3,4,5)P3 to the PM where PtdIns(3,4,5)P3 subsequently regulates its GAP activity to control endocytic events.  相似文献   

8.
9.
10.
Endostatin is an endogenous inhibitor of angiogenesis. Although several endothelial cell surface molecules have been reported to interact with endostatin, its molecular mechanism of action is not fully elucidated. We used surface plasmon resonance assays to characterize interactions between endostatin, integrins, and heparin/heparan sulfate. α5β1 and αvβ3 integrins form stable complexes with immobilized endostatin (KD = ∼1.8 × 10−8 m, two-state model). Two arginine residues (Arg27 and Arg139) are crucial for the binding of endostatin to integrins and to heparin/heparan sulfate, suggesting that endostatin would not bind simultaneously to integrins and to heparan sulfate. Experimental data and molecular modeling support endostatin binding to the headpiece of the αvβ3 integrin at the interface between the β-propeller domain of the αv subunit and the βA domain of the β3 subunit. In addition, we report that α5β1 and αvβ3 integrins bind to heparin/heparan sulfate. The ectodomain of the α5β1 integrin binds to haparin with high affinity (KD = 15.5 nm). The direct binding between integrins and heparin/heparan sulfate might explain why both heparan sulfate and α5β1 integrin are required for the localization of endostatin in endothelial cell lipid rafts.Endostatin is an endogenous inhibitor of angiogenesis that inhibits proliferation and migration of endothelial cells (13). This C-fragment of collagen XVIII has also been shown to inhibit 65 different tumor types and appears to down-regulate pathological angiogenesis without side effects (2). Endostatin regulates angiogenesis by complex mechanisms. It modulates embryonic vascular development by enhancing proliferation, migration, and apoptosis (4). It also has a biphasic effect on the inhibition of endothelial cell migration in vitro, and endostatin therapy reveals a U-shaped curve for antitumor activity (5, 6). Short term exposure of endothelial cells to endostatin may be proangiogenic, unlike long term exposure, which is anti-angiogenic (7). The effect of endostatin depends on its concentration and on the type of endothelial cells (8). It exerts the opposite effects on human umbilical vein endothelial cells and on endothelial cells derived from differentiated embryonic stem cells. Furthermore, two different mechanisms (heparin-dependent and heparin-independent) may exist for the anti-proliferative activity of endostatin depending on the growth factor used to induce cell proliferation (fibroblast growth factor 2 or vascular endothelial growth factor). Its anti-proliferative effect on endothelial cells stimulated by fibroblast growth factor 2 is mediated by the binding of endostatin to heparan sulfate (9), whereas endostatin inhibits vascular endothelial growth factor-induced angiogenesis independently of its ability to bind heparin and heparan sulfate (9, 10). The broad range of molecular targets of endostatin suggests that multiple signaling systems are involved in mediating its anti-angiogenic action (11), and although several endothelial cell surface molecules have been reported to interact with endostatin, its molecular mechanisms of action are not as fully elucidated as they are for other endogenous angiogenesis inhibitors (11).Endostatin binds with relatively low affinity to several membrane proteins including α5β1 and αvβ3 integrins (12), heparan sulfate proteoglycans (glypican-1 and -4) (13), and KDR/Flk1/vascular endothelial growth factor receptor 2 (14), but no high affinity receptor(s) has been identified so far. The identification of molecular interactions established by endostatin at the cell surface is a first step toward the understanding of the mechanisms by which endostatin regulates angiogenesis. We have previously characterized the binding of endostatin to heparan sulfate chains (9). In the present study we have focused on characterizing the interactions between endostatin, α5β1, αvβ3, and αvβ5 integrins and heparan sulfate. Although interactions between several integrins and endostatin have been studied previously in solid phase assays (12) and in cell models (12, 15, 16), no molecular data are available on the binding site of endostatin to the integrins. We found that two arginine residues of endostatin (Arg27 and Arg139) participate in binding to integrins and to heparan sulfate, suggesting that endostatin is not able to bind simultaneously to these molecules displayed at the cell surface. Furthermore, we have demonstrated that α5β1, αvβ3, and αvβ5 integrins bind to heparan sulfate. This may explain why both heparan sulfate and α5β1 integrins are required for the localization of endostatin in lipid rafts, in support of the model proposed by Wickström et al. (15).  相似文献   

11.
12.
Lysophosphatidic acid (LPA), a bioactive phospholipid, induces a wide range of cellular effects, including gene expression, cytoskeletal rearrangement, and cell survival. We have previously shown that LPA stimulates secretion of pro- and anti-inflammatory cytokines in bronchial epithelial cells. This study provides evidence that LPA enhances pulmonary epithelial barrier integrity through protein kinase C (PKC) δ- and ζ-mediated E-cadherin accumulation at cell-cell junctions. Treatment of human bronchial epithelial cells (HBEpCs) with LPA increased transepithelial electrical resistance (TER) by ∼2.0-fold and enhanced accumulation of E-cadherin to the cell-cell junctions through Gαi-coupled LPA receptors. Knockdown of E-cadherin with E-cadherin small interfering RNA or pretreatment with EGTA (0.1 mm) prior to LPA (1 μm) treatment attenuated LPA-induced increases in TER in HBEpCs. Furthermore, LPA induced tyrosine phosphorylation of focal adhesion kinase (FAK) and overexpression of the FAK inhibitor, and FAK-related non-kinase-attenuated LPA induced increases in TER and E-cadherin accumulation at cell-cell junctions. Overexpression of dominant negative protein kinase δ and ζ attenuated LPA-induced phosphorylation of FAK, accumulation of E-cadherin at cell-cell junctions, and an increase in TER. Additionally, lipopolysaccharide decreased TER and induced E-cadherin relocalization from cell-cell junctions to cytoplasm in a dose-dependent fashion, which was restored by LPA post-treatment in HBEpCs. Intratracheal post-treatment with LPA (5 μm) reduced LPS-induced neutrophil influx, protein leak, and E-cadherin shedding in bronchoalveolar lavage fluids in a murine model of acute lung injury. These data suggest a protective role of LPA in airway inflammation and remodeling.The airway epithelium is the site of first contact for inhaled environmental stimuli, functions as a physical barrier to environmental insult, and is an essential part of innate immunity. Epithelial barrier disruption is caused by inhaled allergens, dust, and irritants, resulting in inflammation, bronchoconstriction, and edema as seen in asthma and other respiratory diseases (14). Furthermore, increased epithelial permeability also results in para-cellular leakage of large proteins, such as albumin, immunoglobulin G, and polymeric immunoglobulin A, into the airway lumen (5, 6). The epithelial cell-cell junctional complex is composed of tight junctions, adherens junctions, and desmosomes. These adherens junctions play a pivotal role in regulating the activity of the entire junctional complex because the formation of adherens junctions subsequently leads to the formation of other cell-cell junctions (79). The major adhesion molecules in the adherens junctions are the cadherins. E-cadherin is a member of the cadherin family that mediates calcium-dependent cell-cell adhesion. The N-terminal ectodomain of E-cadherin contains homophilic interaction specificity, and the cytoplasmic domain binds to catenins, which interact with actin (1013). Plasma membrane localization of E-cadherin is critical for the maintenance of epithelial cell-cell junctions and airway epithelium integrity (7, 10, 14). A decrease of adhesive properties of E-cadherin is related to the loss of differentiation and the subsequent acquisition of a higher motility and invasiveness of epithelial cells (10, 14, 15). Dislocation or shedding of E-cadherin in the airway epithelium induces epithelial shedding and increases airway permeability in lung airway diseases (10, 14, 16). In an ovalbumin-challenged guinea pig model of asthma, it has been demonstrated that E-cadherin is dislocated from the lateral margins of epithelial cells (10). Histamine increases airway para-cellular permeability and results in an increased susceptibility of airway epithelial cells to adenovirus infection by interrupting E-cadherin adhesion (14). Serine phosphorylation of E-cadherin by casein kinase II, GSK-3β, and PKD1/PKC2 μ enhanced E-cadherin-mediated cell-cell adhesion in NIH3T3 fibroblasts and LNCaP prostate cancer cells (11, 17). However, the regulation and mechanism by which E-cadherin is localized within the pulmonary epithelium is not fully known, particularly during airway remodeling.LPA, a naturally occurring bioactive lipid, is present in body fluids, such as plasma, saliva, follicular fluid, malignant effusions, and bronchoalveolar lavage (BAL) fluids (1820). Six distinct high affinity cell-surface LPA receptors, LPA-R1–6, have been cloned and described in mammals (2126). Extracellular activities of LPA include cell proliferation, motility, and cell survival (2730). LPA exhibits a wide range of effects on differing cell types, including pulmonary epithelial, smooth muscle, fibroblasts, and T cells (3135). LPA augments migration and cytokine synthesis in lymphocytes and induces chemotaxis of Jurkat T cells through Matrigel membranes (34). LPA induces airway smooth muscle cell contractility, proliferation, and airway repair and remodeling (35, 36). LPA also potently stimulates IL-8 (31, 3739), IL-13 receptor α2 (IL-13Rα2) (40), and COX-2 gene expression and prostaglandin E2 release (41) in HBEpCs. Prostaglandin E2 and IL-13Rα2 have anti-inflammatory properties in pulmonary inflammation (42, 43). These results suggest that LPA may play a protective role in lung disease by stimulating an innate immune response while simultaneously attenuating the adaptive immune response. Furthermore, intravenous injection with LPA attenuated bacterial endotoxin-induced plasma tumor necrosis factor-α production and myeloperoxidase activity in the lungs of mice (44), suggesting an anti-inflammatory role for LPA in a murine model of sepsis.We have reported that LPA induces E-cadherin/c-Met accumulation in cell-cell contacts and increases TER in HBEpCs (45). Here, for the first time, we report that LPA-induced increases in TER are dependent on PKCδ, PKCζ, and FAK-mediated E-cadherin accumulation at cell-cell junctions. Furthermore, we demonstrate that post-treatment of LPA rescues LPS-induced airway epithelial disruption in vitro and reduces E-cadherin shedding in a murine model of ALI. This study identifies the molecular mechanisms linking the LPA and LPA receptors to maintaining normal pulmonary epithelium barrier function, which is critical in developing novel therapies directed at ameliorating pulmonary diseases.  相似文献   

13.
14.
The C-terminal domain of subunit ε of the bacterial FoF1 ATP synthase is reported to be an intrinsic inhibitor of ATP synthesis/hydrolysis activity in vitro, preventing wasteful hydrolysis of ATP under low-energy conditions. Mutants defective in this regulatory domain exhibited no significant difference in growth rate, molar growth yield, membrane potential, or intracellular ATP concentration under a wide range of growth conditions and stressors compared to wild-type cells, suggesting this inhibitory domain is dispensable for growth and survival of Escherichia coli.FoF1 ATP synthases are ubiquitous enzymes that synthesize ATP using a transmembrane electrochemical potential of protons or proton motive force (PMF) generated by the respiratory chain across the cytoplasmic membrane of bacteria, the thylakoid membrane of chloroplasts, or the mitochondrial inner membrane (4, 5, 37). The enzyme consists of two parts: membrane-embedded Fo subcomplex (a complex of subunits a, b, and c in bacteria) and hydrophilic F1 subcomplex (composed of subunits α, β, γ, δ, and ε). The enzyme is also known as a molecular motor, which is composed of the stator subcomplex (α, β, δ, a, and b) and the rotor subcomplex (γ, ε, and c), and its rotation is coupled to ATP synthesis and proton flow across the membrane (20, 31, 52). The reaction of the enzyme is reversible; ATP is hydrolyzed into ADP and inorganic phosphate, the rotor subcomplex rotates in reverse, and protons are extruded to the periplasmic side, resulting in the generation of PMF. Although some bacteria utilize the reverse reaction under particular conditions, the primary function of FoF1 ATP synthase is generation of ATP from the PMF. Therefore, the direction of the activity of FoF1 ATP synthase is regulated to avoid wasteful ATP hydrolysis.Subunit ε in bacterial FoF1 has been known to be an intrinsic inhibitor of F1 and FoF1 complex (18, 21, 23) and is proposed to have a regulatory function (10, 11, 42). Although the inhibitory effects of subunit ε vary among species, in general, ε inhibits ATP hydrolysis activity while repressing ATP synthesis activity to a lesser degree (14, 27). This regulatory function of the ε subunit is mediated almost exclusively by the C-terminal region of ε, which is comprised of two antiparallel α-helices (18, 49, 50). Biochemical and crystallographic studies have revealed that the C-terminal helices can adopt two different conformations (34, 46, 47, 48). In the retracted conformation, the α-helices form a hairpin-like structure and sit on the N-terminal β-sandwich domain of the ε subunit. When the ε subunit exhibits an inhibitory effect, it adopts a more extended conformation in which the C-terminal α-helices extend along the γ subunit, which composes the central stalk. It has also been shown that basic, positively charged residues on the second α-helix of the ε subunit interact with negatively charged residues in the DELSEED segment of subunit β to exert the inhibitory effect (12).Escherichia coli mutants deleted in the entire ε subunit exhibit a reduced growth rate and growth yield, and this effect is proposed to be a result of a deficiency in assembly of the Fo and F1 complexes (21). The N-terminal β-sandwich domain of the ε subunit is responsible for the assembly of Fo and F1 and is therefore important for efficient coupling between proton translocation through Fo and ATP synthesis/hydrolysis in F1 (15, 39). Deletion of the ε subunit leads to dissociation of the FoF1 complex and wasteful ATP hydrolysis by free (cytoplasmic) F1 and dissipation of PMF through free Fo (21, 22, 51).While the importance of the entire ε subunit in the whole-cell physiology of E. coli is fairly well established, the role of the regulatory C-terminal region of ε has received little attention and warrants investigation to determine if the regulatory functions (e.g., inhibition of ATP hydrolysis) observed in vitro are manifested in the physiology of E. coli under various growth conditions. To address this question, we constructed isogenic E. coli mutants that were deleted in the C-terminal region of ε subunit (εDC) and used these strains to compare physiological properties of wild-type versus εDC cells under a wide range of environmental conditions and stressors.  相似文献   

15.
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   

16.
Regulator of G-protein signaling (RGS) proteins are a family of molecules that control the duration of G protein signaling. A variety of RGS proteins have been reported to modulate opioid receptor signaling. Here we show that RGS4 is abundantly expressed in human neuroblastoma SH-SY5Y cells that endogenously express μ- and δ-opioid receptors and test the hypothesis that the activity of opioids in these cells is modulated by RGS4. Endogenous RGS4 protein was reduced by ∼90% in SH-SY5Y cells stably expressing short hairpin RNA specifically targeted to RGS4. In these cells, the potency and maximal effect of δ-opioid receptor agonist (SNC80)-mediated inhibition of forskolin-stimulated cAMP accumulation was increased compared with control cells. This effect was reversed by transient transfection of a stable RGS4 mutant (HA-RGS4C2S). Furthermore, MAPK activation by SNC80 was increased in cells with knockdown of RGS4. In contrast, there was no change in the μ-opioid (morphine) response at adenylyl cyclase or MAPK. FLAG-tagged opioid receptors and HA-RGS4C2S were transiently expressed in HEK293T cells, and co-immunoprecipitation experiments showed that the δ-opioid receptor but not the μ-opioid receptor could be precipitated together with the stable RGS4. Using chimeras of the δ- and μ-opioid receptors, the C-tail and third intracellular domain of the δ-opioid receptor were suggested to be the sites of interaction with RGS4. The findings demonstrate a role for endogenous RGS4 protein in modulating δ-opioid receptor signaling in SH-SY5Y cells and provide evidence for a receptor-specific effect of RGS4.μ- and δ-opioid receptors are members of the G protein-coupled receptor family and interact with Gαi/o proteins (1, 2). This results in signaling to a variety of downstream effectors, including adenylyl cyclase and the mitogen-activated protein kinase (MAPK)2 cascade. Signaling of opioid receptors is regulated negatively by regulator of G protein signaling (RGS) proteins (3, 4). These are a family of molecules containing a “RGS consensus” domain that bind to Gα subunits and act as GTPase-accelerating proteins to increase the rate of GTP hydrolysis. This results in a decrease in the lifetime of the active Gα-GTP and free Gβγ subunits and limits signaling to downstream effectors (58). The mechanisms by which RGS proteins selectively modulate G protein-mediated receptor signal transduction pathways, especially opioid receptor signaling, are beginning to unfold (912). The foundation for the function and selectivity of RGS proteins in regulating opioid signaling lies in their ability to interact with opioid receptors and their cognate G proteins. In general, the selectivity or the preference of an RGS protein for a particular receptor is determined by a variety of factors, including tissue-specific expression and precise interaction with the intracellular domains of receptor proteins, G protein subunits, and effectors as well as other pathway-specific components (13).The effects of RGS proteins on opioid receptor signaling have been examined in several systems. The findings are not always consistent, probably due to the different methodologies used. It has been shown that members of the RZ, R4, and R7 subfamilies (7) of RGS proteins play crucial roles not only in terminating acute opioid agonist action but also in opioid receptor desensitization, internalization, recycling, and degradation (3, 14), thereby affecting opioid tolerance and dependence (1518). Much work has been performed with RGS4, because it is a smaller RGS protein with a structure consisting of the RGS consensus (box) sequence and a small N terminus (19, 20). It also has a wide distribution in the brain, especially in brain regions important for opioid actions, including the striatum, locus coeruleus, dorsal horn of the spinal cord, and cerebral cortex (21). In vitro RGS4 has been shown to reverse δ-opioid receptor agonist-induced inhibition of cAMP synthesis in membranes prepared from NG108-15 cells (6). Overexpression of RGS4 in HEK293 cells also attenuated morphine-, [d-Ala2,N-Me-Phe4,Gly-ol5]enkephalin (DAMGO)-, and [d-Pen2,d-Pen5]enkephalin (DPDPE)-induced inhibition of adenylyl cyclase (22, 23). Co-expression of RGS4 with GIRK1/GIRK2 channels in Xenopus oocytes reduced the basal K+ current and accelerated the deactivation of GIRK channels activated by κ-opioid receptor agonist U69593 (24). Although these previous studies have provided evidence that RGS4 can negatively regulate opioid receptor signaling, they do not confirm a functional role for endogenous RGS4 in endogenous, nontransfected systems.Human neuroblastoma SH-SY5Y cells endogenously express μ- and δ-opioid receptors and a variety of Gαi/o proteins (2527). Here we show that RGS4 is abundantly found at both the mRNA and protein levels in these cells. Consequently, we used SH-SY5Y cells to examine the hypothesis that RGS4 negatively modulates opioid receptor signaling under physiological conditions. The endogenously expressed RGS4 level in SH-SY5Y cells was reduced using lentiviral delivery of short hairpin RNA (shRNA) targeting the RGS4 gene. This resulted in changes in δ- but not μ-opioid receptor-mediated signaling to adenylyl cyclase and the MAPK pathway. These findings argue for a selective interaction of RGS4 with the δ-opioid receptor. To test this, we expressed FLAG-tagged μ- and δ-opioid receptors together with a construct for a stable, proteosome-resistant RGS4 protein in HEK293T cells. Co-immunoprecipitation indicated that the δ-opioid but not the μ-opioid receptor was closely associated with RGS4, providing further evidence for a selective interaction between RGS4 and δ-opioid receptor signaling.  相似文献   

17.
Streptokinase (SK) conformationally activates the central zymogen of the fibrinolytic system, plasminogen (Pg). The SK·Pg* catalytic complex binds Pg as a specific substrate and cleaves it into plasmin (Pm), which binds SK to form the SK·Pm complex that propagates Pm generation. Catalytic complex formation is dependent on lysine-binding site (LBS) interactions between a Pg/Pm kringle and the SK COOH-terminal Lys414. Pg substrate recognition is also LBS-dependent, but the kringle and SK structural element(s) responsible have not been identified. SK mutants lacking Lys414 with Ala substitutions of charged residues in the SK β-domain 250-loop were evaluated in kinetic studies that resolved conformational and proteolytic Pg activation. Activation of [Lys]Pg and mini-Pg (containing only kringle 5 of Pg) by SK with Ala substitutions of Arg253, Lys256, and Lys257 showed decreases in the bimolecular rate constant for Pm generation, with nearly total inhibition for the SK Lys256/Lys257 double mutant. Binding of bovine Pg (BPg) to the SK·Pm complex containing fluorescently labeled Pm demonstrated LBS-dependent assembly of a SK·labeled Pm·BPg ternary complex, whereas BPg did not bind to the complex containing the SK Lys256/Lys257 mutant. BPg was activated by SK·Pm with a Km indistinguishable from the KD for BPg binding to form the ternary complex, whereas the SK Lys256/Lys257 mutant did not support BPg activation. We conclude that SK residues Arg253, Lys256, and Lys257 mediate Pg substrate recognition through kringle 5 of the [Lys]Pg and mini-Pg substrates. A molecular model of the SK·kringle 5 complex identifies the putative interactions involved in LBS-dependent Pg substrate recognition.Streptokinase (SK)6 activates the human fibrinolytic system by activating plasminogen (Pg) through a unique mechanism that is responsible for the use of SK as a thrombolytic drug and its role as a key pathogenicity factor in Group A streptococcal infection (1, 2). The crystal structure of SK bound to the catalytic domain of plasmin (μPm) shows that SK consists of three β-grasp, tightly folded domains, α, β, and γ, linked by flexible segments (3). In solution, SK is highly flexible and behaves hydrodynamically like three beads on a string (4). When bound to μPm, SK assumes a highly ordered structure resembling a three-sided crater surrounding the catalytic site that provides an exosite(s) for binding the catalytic domain of Pg as a substrate (3, 5). In the first step of the SK-mediated Pg activation pathway, SK binds the catalytic domain of the Pg zymogen in a rapid equilibrium process and inserts its NH2-terminal Ile1 residue into the NH2-terminal binding cleft of Pg, activating the catalytic site nonproteolytically (610). Although structural proof is lacking, SK Ile1 presumably forms a critical salt bridge with Asp740(194) (plasminogen numbering; chymotrypsinogen numbering is in parentheses) that initiates conformational activation of the substrate binding site and oxyanion hole required for proteolytic activity (6, 810). The activated SK·Pg* complex binds a second molecule of Pg as a specific substrate and cleaves it at Arg561(15)-Val562(16) to form the fibrin-degrading proteinase, plasmin (Pm) (1014). Proteolytic generation of Pm is propagated by formation of a high affinity SK·Pm complex that converts the remaining free Pg into Pm (5, 11).[Glu]Pg, the full-length form of Pg circulating in blood, consists of an NH2-terminal PAN (Pg/Apple/Nematode (15, 16)) module, followed by five kringle domains (K1–K5), and the trypsin-like serine proteinase catalytic domain (17). Formation of the SK·Pg* and SK·Pm catalytic complexes and Pg substrate binding are inhibited by the lysine analog, 6-aminohexanoic acid (6-AHA), which binds to lysine-binding sites (LBS) located primarily in kringles K1, K4, and K5 of Pg and Pm (10, 11, 1823). Cleavage of the Lys77-Lys78 peptide bond in [Glu]Pg by Pm releases the PAN module and generates the truncated form, [Lys]Pg. Formation of [Lys]Pg is accompanied by a conformational change of [Glu]Pg from a compact, closed α-conformation to a partially extended β-conformation with expression of higher affinity LBS for 6-AHA (24, 25). The fourth kringle module mediates a second conformational change, from the β-conformation to the extended γ-conformation (25).Binding of SK to [Glu]Pg is independent of LBS, with a dissociation constant of 100–150 nm, whereas formation of SK·[Lys]Pg is LBS-dependent with a 13–20-fold higher affinity that is reduced to that of [Glu]Pg by saturating concentrations of 6-AHA (10, 21). Activation of the catalytic domain in [Lys]Pm increases affinity for SK about 830-fold, which is reduced 11–20-fold by 6-AHA (5, 21). Interaction of the COOH-terminal Lys414 residue of SK with a Pg/Pm kringle domain is responsible for the LBS-dependent enhancement of the affinity of SK·[Lys]Pg* and SK·Pm catalytic complex formation (22). Recent rapid reaction kinetic studies of the SK·Pm binding pathway demonstrated that interaction of Lys414 with a Pm kringle enhances formation of an initial rapid equilibrium SK·Pm encounter complex, succeeded by two sequential, tightening conformational changes, to achieve an overall dissociation constant of ∼12 pm (26). The Pg/Pm kringle domain responsible for the enhancement of SK·Pg* and SK·Pm complex formation is not known. Productive interaction of Pg as a substrate of the SK·Pg*/Pm complexes is also greatly inhibited by saturating 6-AHA (11). Kinetic and equilibrium binding studies of SK-mediated Pm formation resolved the conformational activation process from the coupled proteolytic generation of Pm (10, 11). The kinetic approach demonstrated that Lys414 deletion reduced the affinity of formation of the SK·Pg* catalytic complex specifically, whereas the subsequent LBS-dependent proteolytic formation of Pm was unaffected, indicating that Pg substrate recognition is mediated by a structurally distinct region of SK and an unknown kringle (22).Previous structure-function studies have yielded diverse interpretations and conclusions regarding the structural basis of LBS-dependent Pg substrate recognition (23, 2734). Each of the three domains of SK has been implicated in this regard (29, 30, 35, 36), and binding of two Pg molecules to the residue 1–59 sequence of the α-domain has been reported (36). In particular, segments 16–36, 41–48, 48–59, and 88–97 of the SK α-domain have been concluded to play a role in Pg substrate recognition (32, 33, 37, 38). For several SK mutants, a complex mixture of functional effects on their binding to [Glu]Pg and its conformational and proteolytic activation has been reported (28, 31, 33). Some of these effects may result from the inherent flexibility of SK when bound to Pg or Pm (39), and others may be due to the use of kinetic approaches that do not clearly discriminate between conformational and proteolytic activation.Some observations implicate a protruding hairpin loop called the 250-loop (residues Ala251–Ile264) in the SK β-domain in Pg substrate recognition (27, 28, 31, 34). This loop is disordered in the structure of the SK·μPm complex but is ordered in the structure of the isolated β-domain (3, 40). Deletion of the 250-loop, Ala substitution of Lys256 and Lys257 at the apex of the loop, and substitution of multiple residues near and within the loop resulted in disparate effects on Km and kcat for [Glu]Pg activation (27, 28, 31). The conclusions of these studies were that Lys256 and Lys257 are involved in SK binding and conformational activation of [Glu]Pg in addition to proteolytic processing of Pg as a substrate. Some of these studies are problematic because the natural NH2-terminal Ile1 residue necessary for conformational activation is preceded either by an additional methionine (27, 31) or maltose-binding protein (28) in the recombinant SK species used.Because of the diverse conclusions regarding the functional properties of the 250-loop mutations and the possibility of other potential Pg substrate binding sites, the present studies were undertaken to resolve the function of residues in the 250-loop in LBS-dependent Pg substrate recognition by the SK·Pg* complex. The kringle domain of Pg involved in Pg substrate recognition has not been clearly identified but has been suggested to be K5 (27) on the basis that the isolated β-domain bound Pg (30) and K5 (29) in an LBS-dependent manner. Given the general specificity of Pg kringles for COOH-terminal Lys residues and zwitterionic ligands, such as 6-AHA, and the internal sequence of the 250-loop, it appeared possible that a pseudolysine motif on SK was involved. In the binding of a 30-residue peptide from plasminogen binding Group A streptococcal M-like protein (PAM), VEK-30, to K2 of Pg, Castellino and co-workers (41, 42) showed by crystallography and mutagenesis that residues with cationic (Arg and His) and anionic side chains (Glu) arranged spatially on a helix constituted a pseudolysine structure similar to 6-AHA that binds specifically to the LBS of K2. Additional evidence for pseudolysine structures in Pg binding comes from studies of α-enolase from Streptococcus pneumoniae, which has a 9-residue internal binding site for Pg containing essential basic (two Lys residues) and acidic (Asp and Glu residues) located on a surface loop (43, 44).To determine whether a similar SK structure is involved in [Lys]Pg substrate recognition, anionic and cationic residues in the 250-loop were substituted with Ala and characterized in kinetic studies using methods that resolve conformational and proteolytic activation. Studies with [Lys]Pg and mini-Pg, which contains only K5 and the catalytic domain, showed that Arg253, Lys256, and Lys257 facilitate LBS-dependent substrate recognition through interactions with K5. The absence of evidence for a pseudolysine structure in the 250-loop is compatible with the established atypical specificity of K5 for cationic ligands, such as benzamidine, Nα-acetyl-Lys-methyl ester, 6-aminohexane, and 5-aminopentane, in addition to zwitterionic ligands (19, 4547). The studies resolve for the first time the structural features of SK that mediate the LBS-dependent interactions that enhance affinity of SK·Pg* and SK·Pm catalytic complex formation and those that facilitate binding of Pg as a substrate of these complexes.  相似文献   

18.
Transient exposure of β-cells to oxidative stress interrupts the transduction of signals normally coupling glucose metabolism to insulin secretion. We investigated putative persistence of effects induced by one transient oxidative stress (200 μm H2O2, 10 min) on insulin secreting cells following recovery periods of days and weeks. Three days after oxidative stress INS-1E cells and rat islets exhibited persistent dysfunction. In particular, the secretory response to 15 mm glucose was reduced by 40% in INS-1E cells stressed 3 days before compared with naïve cells. Compared with non-stressed INS-1E cells, we observed reduced oxygen consumption (−43%) and impaired glucose-induced ATP generation (−46%). These parameters correlated with increased mitochondrial reactive oxygen species formation (+60%) accompanied with down-regulation of subunits of the respiratory chain and decreased expression of genes responsible for mitochondrial biogenesis (TFAM, −24%; PGC-1α, −67%). Three weeks after single oxidative stress, both mitochondrial respiration and secretory responses were recovered. Moreover, such recovered INS-1E cells exhibited partial resistance to a second transient oxidative stress and up-regulation of UCP2 (+78%) compared with naïve cells. In conclusion, one acute oxidative stress induces β-cell dysfunction lasting over days, explained by persistent damages in mitochondrial components.Pancreatic β-cells are poised to sense blood glucose to regulate insulin exocytosis and thereby glucose homeostasis. The conversion from metabolic signals to secretory responses is mediated through mitochondrial metabolism (1). Failure of the insulin secreting β-cells, a common characteristic of both type 1 and type 2 diabetes, derives from various origins, among them mitochondrial impairment secondary to oxidative stress is a proposed mechanism (2).Oxidative stress is characterized by a persistent imbalance between excessive production of reactive oxygen species (ROS)3 and limited antioxidant defenses. Examples of ROS include superoxide (O2−̇), hydroxyl radical (OH), and hydrogen peroxide (H2O2). Superoxide can be converted to less reactive H2O2 by superoxide dismutase (SOD) and then to oxygen and water by catalase (CAT), glutathione peroxidase (GPx), and peroxiredoxin, which constitute antioxidant defenses. Increased oxidative stress and free radical damages have been proposed to participate in the diabetic state (3). In type 1 diabetes, ROS are implicated in β-cell dysfunction caused by autoimmune reactions and inflammatory cytokines (4). In the context of type 2 diabetes, excessive ROS could promote deficient insulin synthesis (5, 6) and apoptotic pathways in β-cells (5, 7). Of note, ROS fluctuations may also contribute to physiological control of cell functions (8), including the control of insulin secretion (9). It should also be stressed that metabolism of physiological nutrient increases ROS without causing deleterious effects on cell function. However, uncontrolled increase of oxidants, or reduction of their detoxification, leads to free radical-mediated chain reactions ultimately triggering pathogenic events. Pancreatic β-cells are relatively weak in expressing free radical-quenching enzymes SOD, CAT, and GPx (10, 11), rendering those cells particularly susceptible to oxidative attacks (12). Mitochondria are not only the main source of cellular oxidants, they are also the primary target of ROS (13, 14).Mitochondria are essential for pancreatic β-cell function, and damages to these organelles are well known to markedly alter glucose-stimulated insulin secretion (15). The mitochondrial genome constitutes one of the targets, encoding for 13 polypeptides essential for the integrity of electron transport chain (16). Damages to mitochondrial DNA (mtDNA) induce mutations that in turn may favor ROS generation, although the contribution of mtDNA mutations to ROS generation remains unclear. We previously reported that patient-derived mitochondrial A3243G mutation, causing mitochondrial inherited diabetes, is responsible for defective mitochondrial metabolism associated with elevated ROS levels and reduced antioxidant enzyme expression (17). On the other hand, mtDNA mutator mice exhibit accelerated aging without changes in superoxide levels in embryonic fibroblasts (18), showing that ROS generation can be dissociated from mtDNA mutations.In humans, mitochondrial defects typically appear with aging (19), accompanied by sustained ROS generation and progressive oxidant-induced damages (20). In support of this “mitochondrial theory of aging” (21), accumulating evidence shows that in older individuals mitochondria are altered, both morphologically and functionally (22). These age-related mitochondrial changes are foreseen to play a role in the late onset diabetes. In a rat model of intrauterine growth retardation, a vicious cycle between accumulation of mtDNA mutations and elevation of ROS production has been associated to β-cell abnormalities and the onset of type 2 diabetes in adulthood (23). Similarly, mitochondrion-derived ROS impair β-cell function in the Zucker diabetic fatty rat (24). Altogether, these observations point to ROS action as a triggering event inducing mitochondrial dysfunction and ultimately resulting in the loss of the secretory response in β-cells (14).In vitro, oxidative stress applied to β-cells rapidly interrupts the transduction of signals normally coupling glucose metabolism to insulin secretion (12, 25). Specifically, we reported that INS-1E β-cells and rat islets subjected to a 10-min H2O2 exposure exhibit impaired secretory response associated with mitochondrial dysfunction appearing already during the first minutes of oxidative stress (12). In the context of the mitochondrial theory of aging (21, 26), it is important to know whether transient exposure to H2O2 could possibly induce persistent modifications of mitochondrial function. Cells surviving an oxidative stress might carry defects leading to progressive loss of β-cell function. In the present study, we asked the simple but unanswered question if a short transient oxidative stress could induce durable alterations of the mitochondria and thereby chronically impair β-cell function. INS-1E β-cells and rat islets were transiently exposed to H2O2 for 10 min and analyzed after days and weeks of standard tissue culture.  相似文献   

19.
The SLC25 carrier family mediates solute transport across the inner mitochondrial membrane, a process that is still poorly characterized regarding both the mechanisms and proteins implicated. This study investigated mitochondrial glutamate carrier GC1 in insulin-secreting β-cells. GC1 was cloned from insulin-secreting cells, and sequence analysis revealed hydropathy profile of a six-transmembrane protein, characteristic of mitochondrial solute carriers. GC1 was found to be expressed at the mRNA and protein levels in INS-1E β-cells and pancreatic rat islets. Immunohistochemistry showed that GC1 was present in mitochondria, and ultrastructural analysis by electron microscopy revealed inner mitochondrial membrane localization of the transporter. Silencing of GC1 in INS-1E β-cells, mediated by adenoviral delivery of short hairpin RNA, reduced mitochondrial glutamate transport by 48% (p < 0.001). Insulin secretion at basal 2.5 mm glucose and stimulated either by intermediate 7.5 mm glucose or non-nutrient 30 mm KCl was not modified by GC1 silencing. Conversely, insulin secretion stimulated with optimal 15 mm glucose was reduced by 23% (p < 0.005) in GC1 knocked down cells compared with controls. Adjunct of cell-permeant glutamate (5 mm dimethyl glutamate) fully restored the secretory response at 15 mm glucose (p < 0.005). Kinetics of insulin secretion were investigated in perifused isolated rat islets. GC1 silencing in islets inhibited the secretory response induced by 16.7 mm glucose, both during first (−25%, p < 0.05) and second (−33%, p < 0.05) phases. This study demonstrates that insulin-secreting cells depend on GC1 for maximal glucose response, thereby assigning a physiological function to this newly identified mitochondrial glutamate carrier.Functions of mitochondria require regulated flux of molecules across the two membranes surrounding the matrix. Mitochondrial solute carriers (SLC25) are a large family of nuclearly encoded membrane-embedded proteins that promote solute transport across the inner mitochondrial membrane (14). The human genome contains 48 members of the SLC25 family, among them about 30 have been identified and characterized biochemically (1, 58). In particular, very little is known on solute carrier proteins transporting metabolites, such as glutamate. The two isoforms of the glutamate carrier GC1 and GC2 (encoded by SLC25A22 and SLC25A18, respectively) catalyze the transport of glutamate across the inner mitochondrial membrane, either by proton co-transport or in exchange for hydroxyl ions. To date, one human pathology has been associated with GC1, exhibiting a correlation between GC1 mutation and neonatal myoclonic epilepsy (9). Of interest, the high Km isoform GC1 was shown to be expressed in different tissues, especially in the brain, liver, and pancreas (10). Despite the importance of these studies, we still lack subcellular localization and demonstration of the physiological function of glutamate carriers. The elevated expression levels in the pancreas triggered our interest, given that the glutamate pathway has been highlighted over the last years in the endocrine pancreas in general and the β-cell in particular (11). Still, the putative mechanisms responsible for mitochondrial glutamate transport have not yet been characterized in specialized tissues such as insulin-secreting cells. Only two carriers involved in mitochondrial shuttles have been shown to play an important role in the control of insulin secretion, i.e. the aspartate/glutamate carrier (AGC1 or Aralar1) (12) and the citrate/isocitrate carrier (13).It is well founded that mitochondrial metabolism is crucial in pancreatic β-cells by generating signals involved in metabolism-secretion coupling (14). Upon glucose stimulation, generation of ATP through mitochondrial activation leads to the closure of ATP-sensitive K+ channels and depolarization of the plasma membrane (15). This, in turn, induces the opening of voltage-dependent calcium channels resulting in elevation of cytosolic Ca2+ (16). Ca2+ is necessary but not sufficient for the full development of the insulin secretory response (17). Other messengers have been proposed to contribute to stimulation of insulin exocytosis, such as protein kinases A and C, long chain acyl-CoAs, nucleotides, and glutamate (18). The involvement of the latter amino acid was deduced from experiments performed under conditions of intracellular [Ca2+] clamped at permissive concentrations, during which intracellular provision of glutamate directly stimulated insulin exocytosis (1921). Based on these results, it was proposed that glutamate could act downstream of mitochondrial function, participating in the coupling of glucose metabolism to insulin secretion (21). The importance of the glutamate pathway for β-cell function is illustrated in transgenic mice (named βGlud1/) with conditional β-cell-specific deletion of the mitochondrial enzyme glutamate dehydrogenase, resulting in about 40% reduction of glucose-stimulated insulin secretion (22). The exact role of glutamate in β-cell function is still debated as the glutamate pathway might raise insulin release by participating in the amplifying pathway (1921) and/or by relaying signals of protein abundance to mitochondria (2325). In both models, glutamate should be transported in and out of the mitochondria by some putative mitochondrial carrier that remains to be identified in β-cells. Overall, better characterization of mitochondrial glutamate handling will contribute to our comprehension of mechanisms implicated in the control of insulin secretion.In this study, we identified glutamate carrier GC1 as being expressed in the inner mitochondrial membrane of insulinoma INS-1E cells as well as in primary rat islets. Adenovirus-mediated knockdown of GC1 by shRNA2 demonstrated physiological functionality of GC1 in insulin secretion.  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