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 78 毫秒
1.
2.
3.
4.
5.
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   

6.
The acid-sensing ion channels (ASICs) open in response to extracellular acidic pH, and individual subunits display differential sensitivity to protons and calcium. ASIC1a acts as a high affinity proton sensor, whereas ASIC2a requires substantially greater proton concentrations to activate. Using chimeras composed of ASIC1a and ASIC2a, we determined that two regions of the extracellular domain (residues 87–197 and 323–431) specify the high affinity proton response of ASIC1a. These two regions appear to undergo intersubunit interactions within the multimeric channel to specify proton sensitivity. Single amino acid mutations revealed that amino acids around Asp357 play a prominent role in determining the pH dose response of ASIC1a. Within the same region, mutation F352L abolished PcTx1 modulation of ASIC1a. Surprisingly, we determined that another area of the extracellular domain was required for calcium-dependent regulation of ASIC1a activation, and this region functioned independently of high affinity proton sensing. These results indicate that specific regions play overlapping roles in pH-dependent gating and PcTx1-dependent modulation of ASIC1a activity, whereas a distinct region determines the calcium dependence of ASIC1a activation.The acid-sensing ion channels (ASICs)3 are proton-gated ion channels expressed in neurons throughout the central and peripheral nervous system (13). ASICs are activated by extracellular acidosis, and protons act as ligands triggering channel opening (4). Disruption of the accn2 gene (which encodes ASIC1) dramatically reduces proton-gated currents in central neurons and alters a variety of behaviors, including fear, learning, and memory (5, 6). ASIC1 also contributes to neuronal damage and death during the prolonged acidosis accompanying cerebral ischemia (7). Specifically, mice with disruptions in the accn2 gene display 60% smaller lesion size compared with normal mice in models of stroke (8). Application of PcTx1, a venom peptide that prevents ASIC1a activation, is similarly neuroprotective, even when administered hours after injury (8, 9). Thus, ASIC1a represents a novel pharmacological target for the prevention of neuronal death following stroke.Mammals have four genes encoding ASICs (accn1 to -4) that encode at least six different ASIC subunits (13, 10). Like all members of the DEG/ENaC family, individual ASIC subunits have two transmembrane regions separated by a large cysteine-rich extracellular region. Three ASIC subunits associate to form homomeric or heteromeric channels with distinct biophysical characteristics (1114). Specifically, ASIC1a homomeric channels activate at pH values much closer to neutral pH compared with ASIC2a homomeric channels. The high affinity proton sensitivity of ASIC1a plays a prominent role in acidosis-induced neuronal death, and modulators that alter the pH dose response of ASIC1a affect neuronal sensitivity to prolonged acidosis (8, 9, 15). For example, the neuroprotective venom peptide PcTx1 increases the proton sensitivity of the ASIC1a channel, allowing the channel to desensitize at neutral pH and become unresponsive to subsequent acidic shifts in pH (16, 17). The large extracellular region of ASIC1a is thought to be the site of proton/modulator interaction and governs the characteristics of channel gating (10, 11, 18). However, the exact molecular mechanisms defining ASIC1a activation and the protein domains that are responsible for the apparent proton sensitivity of ASIC1a remain unclear. Here, we used chimeras containing specific regions from both ASIC1a and ASIC2a to identify the specific protein regions that confer high affinity proton sensing, PcTx1 sensitivity, and calcium modulation to ASIC1a.  相似文献   

7.
8.
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   

9.
10.
11.
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   

12.
13.
There are a large number of Rho guanine nucleotide exchange factors, most of which have no known functions. Here, we carried out a short hairpin RNA-based functional screen of Rho-GEFs for their roles in leukocyte chemotaxis and identified Arhgef5 as an important factor in chemotaxis of a macrophage phage-like RAW264.7 cell line. Arhgef5 can strongly activate RhoA and RhoB and weakly RhoC and RhoG, but not Rac1, RhoQ, RhoD, or RhoV, in transfected human embryonic kidney 293 cells. In addition, Gβγ interacts with Arhgef5 and can stimulate Arhgef5-mediated activation of RhoA in an in vitro assay. In vivo roles of Arhgef5 were investigated using an Arhgef-5-null mouse line. Arhgef5 deficiency did not affect chemotaxis of mouse macrophages, T and B lymphocytes, and bone marrow-derived mature dendritic cells (DC), but it abrogated MIP1α-induced chemotaxis of immature DCs and impaired migration of DCs from the skin to lymph node. In addition, Arhgef5 deficiency attenuated allergic airway inflammation. Therefore, this study provides new insights into signaling mechanisms for DC migration regulation.Leukocyte chemotaxis underlies leukocyte migration, infiltration, trafficking, and homing that are not only important for normal leukocyte functions, but also have a important role in inflammation-related diseases. Leukocyte chemotaxis is regulated by leukocyte chemoattractants that include bacterial by-products such as formylmethionylleucylphenylalanine, complement proteolytic fragments such as C5a, and the superfamily of chemotactic cytokines, chemokines. These chemoattractants bind to their specific cell G protein-coupled receptors and are primarily coupled to the Gi family of G proteins to regulate leukocyte chemotaxis. Previous studies have established that the Rho family of small GTPases regulates leukocyte migration (1, 2). Rac, Cdc42, and RhoA are the three best studied Rho small GTPases. In myeloid cells, Cdc42 regulates directionality by directing where F-actin and lamellipodia are formed, and Rac regulates F-actin formation in the lamellipodia, which provides a driving force for cell motility (36). On the other hand, RhoA regulates the formation and contractility of the actomyosin structure at the back that provides a pushing force (5, 7). Rho guanine nucleotide exchange factors (GEF)3 are key regulators for the activity of these small GTPases. GEFs activate small GTPases by promoting the loading of GTP to the small GTPases, a rate-limiting step in GTPase regulation (811). Previous biochemical and genetic studies have revealed how Cdc42 and Rac may be regulated by chemokine receptors in leukocytes. Chemokine receptors can regulate Cdc42 via a Rho-GEF PIXα, which is regulated by Gβγ from the Gi proteins via the interactions between Gβγ and Pak1 and between Pak1 and PIXα in myeloid cells 12. On the other hand, in neutrophils chemokine receptors regulate Rac2 via another Rho-GEF P-Rex1, which is directly regulated by Gβγ (1315). Two Rho-GEFs have been implicated in regulation of RhoA in neutrophils. GEF115 was found in the leading edges of polarized mouse neutrophils, whereas PDZ Rho-GEF was found in the uropods of differentiated HL-60 cells. Both Rho-GEFs were believed to mediate pertussis toxin-resistant activation of RhoA in these cells. However, a significant portion of RhoA activity in leukocytes are pertussis toxin-sensitive, which is presumably regulated by the α and/or βγ subunits from the Gi proteins. The signaling mechanism for this pertussis toxin-sensitive RhoA regulation by chemokine receptors remains largely elusive.Molecular cloning and genomic sequencing have identified more than 70 Rho-GEFs in mammals (1620). Many of these Rho-GEFs have been shown to activate RhoA in in vitro and overexpression assays (1620). However, it is not known if any of them regulate RhoA in vivo, we have found that PIXα is a specific GEF for Cdcd42 in neutrophils (12) despite its potent activity on Rac in in vitro and overexpression assays (21, 22). Therefore, we used a siRNA-based loss of function screen in an attempt to identify the GEFs that regulate myeloid cell migration and RhoA activity. One of the candidates, Arhgef5, was found to be directly activated by Gβγ to regulate RhoA and has an important role in immature DC migration. In addition, Arhgef5 deficiency attenuated allergic airway inflammation in a mouse model.  相似文献   

14.
15.
16.
Class I hydrophobins function in fungal growth and development by self-assembling at hydrophobic-hydrophilic interfaces into amyloid-like fibrils. SC3 of the mushroom-forming fungus Schizophyllum commune is the best studied class I hydrophobin. This protein spontaneously adopts the amyloid state at the water-air interface. In contrast, SC3 is arrested in an intermediate conformation at the interface between water and a hydrophobic solid such as polytetrafluoroethylene (PTFE; Teflon). This finding prompted us to study conditions that promote assembly of SC3 into amyloid fibrils. Here, we show that SC3 adopts the amyloid state at the water-PTFE interface at high concentration (300 μg ml−1) and prolonged incubation (16 h). Moreover, we show that amyloid formation at both the water-air and water-PTFE interfaces is promoted by the cell wall components schizophyllan (β(1–3),β(1–6)-glucan) and β(1–3)-glucan. Hydrophobin concentration and cell wall polysaccharides thus contribute to the role of SC3 in formation of aerial hyphae and in hyphal attachment.Hydrophobins are a class of surface active proteins that play diverse roles in fungal growth and development. For instance, they allow fungi to escape an aqueous environment, confer hydrophobicity to fungal surfaces in contact with air, and mediate attachment of fungi to hydrophobic surfaces (1, 2). They also play a role in the architecture of the cell wall (3).Hydrophobins share eight conserved cysteine residues, but otherwise their sequences are diverse (4). Class I and II hydrophobins are distinguished on the basis of differences in hydropathy patterns and biophysical properties (5). SC3 of Schizophyllum commune is the best characterized class I hydrophobin. It self-assembles at interfaces between water and air, water and oil, and water and hydrophobic solids (68). The four disulfide bridges of SC3 prevent spontaneous self-assembly in solution and thus account for the controlled assembly at hydrophobic-hydrophilic interfaces (9).The water-soluble form of SC3 is oligomeric (10) and rich in β-sheet (11). Upon assembly at the water-air interface, SC3 proceeds via an intermediate form that has increased α-helical structure (α-helical state) to a stable end form that has increased β-sheet structure (β-sheet state) (1113). SC3 in the β-sheet state initially has no clear ultrastructure (β-sheet I state) (12), but after prolonged incubation, the protein forms 10-nm wide amyloid-like fibrils (β-sheet II state) (1214) that are called rodlets (6, 15). Like other amyloid fibrils (16), rodlets of the hydrophobins SC3 of S. commune and EAS of Neurospora crassa increase fluorescence of thioflavin T and bind Congo red (14, 17, 18). Moreover, x-ray diffraction of rodlets of EAS showed reflections at 4.8 Å (distance between strands in a β-sheet) and 10–12 Å (spacing between β-sheets stacked perpendicular to the fibril long axis) (19), which are indicative for amyloid fibrils.Notably, SC3 does not spontaneously self-assemble into amyloid fibrils at an interface between water and a hydrophobic solid. Instead, SC3 is arrested in the intermediate α-helical state. Transition to the β-sheet state is observed only by heating the sample in the presence of detergent (11, 12). These observations prompted us to study conditions that promote assembly of SC3 into amyloid fibrils. Here, we show that amyloid formation of SC3 is promoted by increasing its concentration or by the presence of cell wall polysaccharides.  相似文献   

17.
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   

18.
19.
Aldo-keto reductase family 1 member B10 (AKR1B10) is primarily expressed in the normal human colon and small intestine but overexpressed in liver and lung cancer. Our previous studies have shown that AKR1B10 mediates the ubiquitin-dependent degradation of acetyl-CoA carboxylase-α. In this study, we demonstrate that AKR1B10 is critical to cell survival. In human colon carcinoma cells (HCT-8) and lung carcinoma cells (NCI-H460), small-interfering RNA-induced AKR1B10 silencing resulted in caspase-3-mediated apoptosis. In these cells, the total and subspecies of cellular lipids, particularly of phospholipids, were decreased by more than 50%, concomitant with 2–3-fold increase in reactive oxygen species, mitochondrial cytochrome c efflux, and caspase-3 cleavage. AKR1B10 silencing also increased the levels of α,β-unsaturated carbonyls, leading to the 2–3-fold increase of cellular lipid peroxides. Supplementing the HCT-8 cells with palmitic acid (80 μm), the end product of fatty acid synthesis, partially rescued the apoptosis induced by AKR1B10 silencing, whereas exposing the HCT-8 cells to epalrestat, an AKR1B10 inhibitor, led to more than 2-fold elevation of the intracellular lipid peroxides, resulting in apoptosis. These data suggest that AKR1B10 affects cell survival through modulating lipid synthesis, mitochondrial function, and oxidative status, as well as carbonyl levels, being an important cell survival protein.Aldo-keto reductase family 1 member B10 (AKR1B10,2 also designated aldose reductase-like-1, ARL-1) is primarily expressed in the human colon, small intestine, and adrenal gland, with a low level in the liver (13). However, this protein is overexpressed in hepatocellular carcinoma, cervical cancer, lung squamous cell carcinoma, and lung adenocarcinoma in smokers, being a potential diagnostic and/or prognostic marker (1, 2, 46).The biological function of AKR1B10 in the intestine and adrenal gland, as well as its role in tumor development and progression, remains unclear. AKR1B10 is a monomeric enzyme that efficiently catalyzes the reduction to corresponding alcohols of a range of aromatic and aliphatic aldehydes and ketones, including highly electrophilic α,β-unsaturated carbonyls and antitumor drugs containing carbonyl groups, with NADPH as a co-enzyme (1, 712). The electrophilic carbonyls are constantly produced by lipid peroxidation, particularly in oxidative conditions, and are highly cytotoxic; through interaction with proteins, peptides, and DNA, the carbonyls cause protein dysfunction and DNA damage (breaks and mutations), resulting in mutagenesis, carcinogenesis, or apoptosis (10, 1319). AKR1B10 also shows strong enzymatic activity toward all-trans-retinal, 9-cis-retinal, and 13-cis-retinal, reducing them to the corresponding retinols, which may regulate the intracellular retinoic acid, a signaling molecule modulating cell proliferation and differentiation (6, 2023). In lung cancer, AKR1B10 expression is correlated with the patient smoking history and activates procarcinogens in cigarette smoke, such as polycyclic aromatic hydrocarbons, thus involved in lung tumorigenesis (2426).Recent studies have shown that in breast cancer cells, AKR1B10 associates with acetyl-CoA carboxylase-α (ACCA) and blocks its ubiquitination and proteasome degradation (27). ACCA is a rate-limiting enzyme of de novo synthesis of long chain fatty acids, catalyzing the ATP-dependent carboxylation of acetyl-CoA to form malonyl-CoA (28). Long chain fatty acids are the building blocks of biomembranes and the precursor of lipid second messengers, playing a critical role in cell growth and proliferation (29, 30). Therefore, ACCA activity is tightly regulated by both metabolite-mediated allosteric mechanisms and phosphorylation-dependent mechanisms; the latter are controlled by multiple hormones, such as insulin, glucagon, and growth factors (3133). ACCA activity is also regulated through physical protein-protein interaction. For instance, breast cancer 1 (BRCA1) protein associates with the ACCA and blocks its Ser79 residue from dephosphorylation (34, 35). The AKR1B10-mediated regulation on ACCA stability represents a novel regulatory mechanism, and this current study elucidated the biological significance of this regulation. The results show that AKR1B10 promotes cell survival via modulating lipid synthesis, mitochondrial function and oxidative stress, and carbonyl levels.  相似文献   

20.
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   

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