首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
A central question in Wnt signaling is the regulation of β-catenin phosphorylation and degradation. Multiple kinases, including CKIα and GSK3, are involved in β-catenin phosphorylation. Protein phosphatases such as PP2A and PP1 have been implicated in the regulation of β-catenin. However, which phosphatase dephosphorylates β-catenin in vivo and how the specificity of β-catenin dephosphorylation is regulated are not clear. In this study, we show that PP2A regulates β-catenin phosphorylation and degradation in vivo. We demonstrate that PP2A is required for Wnt/β-catenin signaling in Drosophila. Moreover, we have identified PR55α as the regulatory subunit of PP2A that controls β-catenin phosphorylation and degradation. PR55α, but not the catalytic subunit, PP2Ac, directly interacts with β-catenin. RNA interference knockdown of PR55α elevates β-catenin phosphorylation and decreases Wnt signaling, whereas overexpressing PR55α enhances Wnt signaling. Taken together, our results suggest that PR55α specifically regulates PP2A-mediated β-catenin dephosphorylation and plays an essential role in Wnt signaling.Wnt/β-catenin signaling plays essential roles in development and tumorigenesis (13). Our previous work found that β-catenin is sequentially phosphorylated by CKIα4 and GSK3 (4), which creates a binding site for β-Trcp (5), leading to degradation via the ubiquitination/proteasome machinery (3). Mutations in β-catenin or APC genes that prevent β-catenin phosphorylation or ubiquitination/degradation lead ultimately to cancer (1, 2).In addition to the involvement of kinases, protein phosphatases, such as PP1, PP2A, and PP2C, are also implicated in Wnt/β-catenin regulation. PP2C and PP1 may regulate dephosphorylation of Axin and play positive roles in Wnt signaling (6, 7). PP2A is a multisubunit enzyme (810); it has been reported to play either positive or negative roles in Wnt signaling likely by targeting different components (1121). Toward the goal of understanding the mechanism of β-catenin phosphorylation, we carried out siRNA screening targeting several major phosphatases, in which we found that PP2A dephosphorylates β-catenin. This is consistent with a recent study where PP2A is shown to dephosphorylate β-catenin in a cell-free system (18).PP2A consists of a catalytic subunit (PP2Ac), a structure subunit (PR65/A), and variable regulatory B subunits (PR/B, PR/B′, PR/B″, or PR/B‴). The substrate specificity of PP2A is thought to be determined by its B subunit (9). By siRNA screening, we further identified that PR55α, a regulatory subunit of PP2A, specifically regulates β-catenin phosphorylation and degradation. Mechanistically, we found that PR55α directly interacts with β-catenin and regulates PP2A-mediated β-catenin dephosphorylation in Wnt signaling.  相似文献   

3.
4.
5.
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   

6.
7.
8.
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   

9.
The temperature-dependent rotation of F1-ATPase γ subunit was observed in Vmax conditions at low viscous drag using a 60-nm gold bead (Nakanishi-Matsui, M., Kashiwagi, S., Hosokawa, H., Cipriano, D. J., Dunn, S. D., Wada, Y., and Futai, M. (2006) J. Biol. Chem. 281, 4126–4131). The Arrhenius slopes of the speed of the individual 120° steps and reciprocal of the pause length between rotation steps were very similar, indicating a flat energy pathway followed by the rotationally coupled catalytic cycle. In contrast, the Arrhenius slope of the reciprocal pause length of the γM23K mutant F1 was significantly increased, whereas that of the rotation rate was similar to wild type. The effects of the rotor γM23K substitution and the counteracting effects of βE381D mutation in the interacting stator subunits demonstrate that the rotor-stator interactions play critical roles in the utilization of stored elastic energy. The γM23K enzyme must overcome an abrupt activation energy barrier, forcing it onto a less favored pathway that results in uncoupling catalysis from rotation.F-ATPase (FoF1), consisting of the catalytic sector F13β3γδϵ) and the transmembrane proton transport sector Fo (ab2c10), synthesizes or hydrolyzes ATP coupled with proton transport (for reviews, see Ref. 16). As Abrahams et al. (7) discovered in the first high resolution x-ray structure, a critical feature of the F1-ATPase is the inherent asymmetry of the three β subunits in different conformations, βTP, βDP, and βE, referring to the nucleotide bound in each catalytic site, ATP, ADP, or empty, respectively. A rotational mechanism has been firmly established mostly based on direct observation in single molecule experiments of the behavior of the rotor complex ϵγc10, relative to the stator complex α3β3δab2 (reviewed in Ref. 1). ATP hydrolysis-dependent rotation of the γ and ϵ subunits in purified bacterial F1 (8, 9), the ϵγc10 complex in detergent solubilized FoF1 (1013), and the ϵγc10 complexin FoF1 in lipid bilayers (14) were shown experimentally by single molecule observations using fluorescent actin filament as a probe. Relative rotation of the single copy Fo a subunit was also shown in F0F1, which was immobilized through the ring of ∼10 c subunits, suggesting that the rotor and stator are interchangeable mechanical units (14). ATP synthesis by F-ATPase is believed to follow the reverse mechanism of ATP hydrolysis because mechanically induced rotation of the γ subunit in immobilized F1 in the presence of ADP and Pi results in net ATP synthesis (15, 16). There remain many questions about the mechanism of coupling between catalysis and transport via mechanical rotation. In particular, the mechanism of coupling H+ transport to rotation of the subunit c10 ring is still not well understood (4).In contrast, there is considerably more information on the mechanism of coupling catalysis to γ and ϵ subunit rotation. Observations of γ subunit rotation in the catalytic F1 sector are consistent with Boyer''s binding change model (17); thus coupling between the chemistry and rotation can be assessed by studies of the soluble F1, and these findings relate to the mechanism of the entire ATP synthase complex. The γ subunit rotates relative to the α3β3 hexamer in distinct 120° steps. A 120° rotation step consisting of pause and rotation substeps appears to correspond to the hydrolysis of one ATP, assuming that three ATP molecules are hydrolyzed per 360° revolution (18). Additional pauses observed at low ATP concentrations are attributed to the “ATP waiting” dwell (19). Yasuda et al. (19) and Shimabukuro et al. (20) further resolved that each 120° step occurred in two substeps: an 80° substep whose onset was dependent upon the Mg·ATP concentration, and a 40° substep, which was not affected by substrate concentration (19). The pause before the 80° substep, the ATP waiting dwell became shorter with increasing [Mg·ATP]. In contrast, the pause duration before the 40° rotation step was modulated by the slow hydrolysis rate of ATPγS2 or by the catalytic site mutant βE190D (in the Bacillus PS3 F1), which was found to significantly increase the length of the catalytic dwell (20). These data together indicate that the dwell before the 40° step is the “catalytic dwell” (20) and defines the order of the substeps during the 120° rotation step observed in high Mg·ATP concentrations (21).In this paper, we address the question of when the rate-limiting step of steady state catalysis occurs, with respect to the rotational behavior. Pre-steady state analysis of the burst kinetics of ATP hydrolysis at nearly Vmax conditions demonstrated that the rate-limiting transition state occurs after the reversible hydrolysis/synthesis step and before release of phosphate (Pi) (22, 23). The rate-limiting step is likely associated with a rotation step because a γ-β cross-linked enzyme is still able to undergo the initial ATP hydrolysis, but the rotation-impeded enzyme is unable to release Pi (23). Significantly, the kinetics of steady state hydrolysis can only be assessed when the Mg·ATP concentration is high enough to fill all three catalytic sites. The only model consistent with these data is one that involves all three catalytic sites. During each 120° catalytic cycle, one site binds ATP, a different site carries out reversible hydrolysis/synthesis, and the third site releases product Pi and ADP (22, 23).Steady state analyses, which take advantage of a particular γ subunit mutation γM23K (24), are consistent with this model. Replacement of the conserved γMet-23 with lysine causes an uncoupling between catalysis and γ subunit rotation caused by altered interactions between γ and β subunits (25). Importantly, Al-Shawi and Nakamoto (26) and Al-Shawi et al. (25, 27) found that the γM23K mutation strongly affected the rate-limiting transition state of steady state ATP hydrolysis and ATP synthesis. The slope of the Arrhenius plots and thus the energy of activation were significantly increased in the mutant enzyme. Several second site suppressor mutations, mostly in the γ subunit (28, 29) but also in the β subunits (30, 31), were genetically identified because they restored coupled ATP synthesis. Significantly, all were in the γ-β interface. Thermodynamic analyses found that the second site suppressors generally compensated for the primary γM23K mutations by reducing the increased activation energy (25, 27, 31). Although most of the second site mutations were found distant from the γM23K site, the x-ray crystal structures (7) suggested that γM23K may directly interact with conserved βGlu-381. As expected, replacement of βGlu-381 with aspartate also suppressed the uncoupling effects of γM23K (31).To identify the rate-limiting transition state step in the rotational behavior, we analyzed the temperature dependence of the γM23K mutant in Vmax conditions observed in single molecule experiments. Interestingly, direct observation of this mutant using the micron-length actin filaments did not detect differences in the rotation behavior at room temperature (9). In contrast, we find in the data presented here that there is dramatic effect of the mutation on the temperature dependence of the length of the catalytic dwell or pause between the 120° rotation steps. This is likely because of two factors: first, we used a bead small enough not to invoke a drag on the rotation (32), and second, the temperature dependence of the rate of the rotation steps is critical for the analyses of the mechanism.  相似文献   

10.
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   

11.
12.
The scorpion α-toxin Lqh2 (from Leiurus quinquestriatus hebraeus) is active at various mammalian voltage-gated sodium channels (Navs) and is inactive at insect Navs. To resolve the molecular basis of this preference we used the following strategy: 1) Lqh2 was expressed in recombinant form and key residues important for activity at the rat brain channel rNav1.2a were identified by mutagenesis. These residues form a bipartite functional surface made of a conserved “core domain” (residues of the loops connecting the secondary structure elements of the molecule core), and a variable “NC domain” (five-residue turn and the C-tail) as was reported for other scorpion α-toxins. 2) The functional role of the two domains was validated by their stepwise construction on the similar scaffold of the anti-insect toxin LqhαIT. Analysis of the activity of the intermediate constructs highlighted the critical role of Phe15 of the core domain in toxin potency at rNav1.2a, and has suggested that the shape of the NC-domain is important for toxin efficacy. 3) Based on these findings and by comparison with other scorpion α-toxins we were able to eliminate the activity of Lqh2 at rNav1.4 (skeletal muscle), hNav1.5 (cardiac), and rNav1.6 channels, with no hindrance of its activity at Nav1.1–1.3. These results suggest that by employing a similar approach the design of further target-selective sodium channel modifiers is imminent.The pivotal role of voltage-gated sodium channels (Navs)4 in excitability mark them as major targets for a large variety of toxins that bind at distinct receptor sites and modify their gating (1). These channels are large membrane proteins made of a pore-forming α-subunit of ∼260 kDa and auxiliary β-subunits of ∼30 kDa. The α-subunit is composed of four homologous domains (D1–D4), each consisting of six α-helical transmembrane segments (S1–S6) connected by intracellular and extracellular loops. A key feature in Navs function is their ability to rapidly activate and inactivate, leading to transient increase in Na+ conductance through the cell membrane. This mechanism is attributed to the ability of the positively charged S4 voltage sensors to move across the membrane in response to changes in membrane potential (1, 2).In mammals, at least nine genes encode a variety of Nav subtypes (1, 3), whose expression varies greatly in different tissues (Nav1.1–1.3 mainly in the central nervous system; Nav1.6 in both central and peripheral neurons; Nav1.7 in the peripheral nervous system; Nav1.8 and Nav1.9 in sensory neurons; Nav1.4 and Nav1.5 in skeletal and cardiac muscles, respectively). Nav subtypes are distributed heterogeneously in the human brain and their expression is regulated under developmental and pathological conditions (1, 35). In addition, many disorders in humans result from abnormal function due to mutations in various Nav genes (68). Thus, ligands that show specificity for Nav subtypes may be used for their identification at various tissues and as leads for design of specific drugs. This requires that the bioactive surfaces of these ligands be resolved along with molecular details that determine their specificity.Among the wide range of Nav modifiers, those derived from scorpion venoms play an important role in studying channel activation (β-toxins) and inactivation (α-toxins) (911). The channel site of interaction with scorpion α-toxins, named neurotoxin receptor site-3 (12), is shared also by structurally unrelated toxins from sea anemone and spider venoms (13, 14), which raises questions as to its architecture and boundaries. Based on the findings that site-3 toxins eliminate a gating charge component associated with the movement of D4/S4 (15, 16), and that this segment plays a critical role in coupling channel inactivation to activation (17), scorpion α-toxins were postulated to inhibit channel inactivation by hindering the outward movement of this segment during depolarization (9).Scorpion α-toxins constitute a class of structurally and functionally related 61–67-residue long polypeptides reticulated by four conserved disulfide bridges. Despite a common βαββ core (10, 18, 19) these toxins are highly diverse in sequence and preference for insect and mammalian Navs. Indeed, the α-toxin class is divided to pharmacological groups according to their toxicity in insects and mice brain and ability to compete on binding at insect and mammalian Navs (10) (supplemental Fig. S1): (i) classical anti-mammalian toxins, such as Aah2 (from Androctonus australis hector) and Lqh2 (from Leiurus quinquestriatus hebraeus), which bind with high affinity to Navs at rat brain synaptosomes and are practically non-toxic to insects; (ii) α-toxins, such as LqhαIT, which strongly affect insect Navs and are weak in mammalian brain; and (iii) α-like toxins, such as Lqh3 and BmKM1 (from Buthus martensii Karsch), which are active in both mammalian brain and insects.Efforts to identify α-toxin residues involved in the interaction with the Nav receptor site-3 revealed a generally common bioactive surface divided to two topologically distinct domains: a conserved “core domain” formed by residues of the loops connecting the secondary structure elements of the molecule core, and a variable “NC domain” formed by the five-residue turn (residues 8–12) and the C-tail (2023). These analyses raised the hypothesis that a protruding conformation of the NC domain correlates with high activity at insect Navs, whereas a flat conformation of this domain appears in α-toxins active at the brain channel rNav1.2a (21). The correlation of this structural difference with toxin preference for Nav subtypes was corroborated by constructing the bioactive surface of LqhαIT on the scaffold of the anti-mammalian α-toxin Aah2 ending up with a chimera (Aah2LqhαIT(face)) active on insects, whose NC domain is in the protruding conformation (21). Despite this result, the molecular requirements that enable high affinity binding of classical α-toxins to mammalian Navs have not been clarified, and only initial data about the channel region that constitutes receptor site-3 is available (Refs. 2426; also see Ref. 10 for review).Lqh2 is a 64-residue long toxin from L. quinquestriatus hebraeus (Israeli yellow scorpion) (27) that is almost identical in sequence (96% identity) to the most active anti-mammalian toxin, Aah2, whose structure and action are documented (18, 28, 29). By functional expression and mutagenesis we uncovered residues on the Lqh2 exterior that are putatively involved in bioactivity. By construction of these residues on the scaffold of the anti-insect toxin LqhαIT we confirmed their bioactive role and differentiated those that determine toxin potency from those contributing to toxin efficacy. Comparison to other α-toxins was then instrumental for the design of an Lqh2 mutant that exhibits high specificity for the neuronal channels hNav1.1, rNav1.2a, and rNav1.3.  相似文献   

13.
14.
Low glutathione levels are associated with crystallin oxidation in age-related nuclear cataract. To understand the role of cysteine residue oxidation, we used the novel approach of comparing human cataracts with glutathione-depleted LEGSKO mouse lenses for intra- versus intermolecular disulfide crosslinks using 2D-PAGE and proteomics, and then systematically identified in vivo and in vitro all disulfide forming sites using ICAT labeling method coupled with proteomics. Crystallins rich in intramolecular disulfides were abundant at young age in human and WT mouse lens but shifted to multimeric intermolecular disulfides at older age. The shift was ∼4x accelerated in LEGSKO lens. Most cysteine disulfides in β-crystallins (except βA4 in human) were highly conserved in mouse and human and could be generated by oxidation with H2O2, whereas γ-crystallin oxidation selectively affected γC23/42/79/80/154, γD42/33, and γS83/115/130 in human cataracts, and γB79/80/110, γD19/109, γF19/79, γE19, γS83/130, and γN26/128 in mouse. Analysis based on available crystal structure suggests that conformational changes are needed to expose Cys42, Cys79/80, Cys154 in γC; Cys42, Cys33 in γD, and Cys83, Cys115, and Cys130 in γS. In conclusion, the β-crystallin disulfidome is highly conserved in age-related nuclear cataract and LEGSKO mouse, and reproducible by in vitro oxidation, whereas some of the disulfide formation sites in γ-crystallins necessitate prior conformational changes. Overall, the LEGSKO mouse model is closely reminiscent of age-related nuclear cataract.Aging lens crystallins accumulate post-synthetic modifications that can be broadly classified into three categories, namely (1) protein backbone changes, such as racemization and truncation (13), (2) conversion of one amino acid into another, such as deamidation of asparagine into aspartate or deguanidination of arginine into ornithine, deamination of lysine into allysine and 2-aminoadipic acid (46), and (3) amino acid residue damage from reactive carbonyls and reactive oxygen species (7,8). Carbonyl damage results from the Maillard Reaction by glucose, methylglyoxal, or oxidation products of ascorbate, tryptophan or lipids which form adducts and crosslinks with nucleophilic group of lysine, arginine and cysteine. Examples include carboxymethyl-lysine, pentosidine, methylglyoxal hydroimidazolones, HNE-cysteine adducts and kynurenine (7,912). Oxidative damage results from reactive oxygen species that directly damage amino acid residues, e.g. oxidizing tryptophan into N-formyl kynurenine and kynurenine, methionine into its sulfoxide, and cysteine into cysteine disulfides or cysteic acid (1315).Because of their relevance to age-related cataract, the impact of each of these modifications on crystallin structure and stability is the subject of intense investigation. Importantly, Benedek proposed that high molecular weight (HMW)1 crystallin aggregates the size of 50 million daltons are needed in order for lens opacification to be visible(16,17). Crystallin aggregation conceivably occurs by one of several mechanisms that include conformational changes as a consequence amino acid mutations (18) or physical-chemical protein modifications. Of the latter, one mechanism that is dominant in several types of cataract involves oxidation of cysteines into protein disulfides (18) and formation of HMW aggregates that scatter light (19).In order to mimic the oxidative process and formation of protein disulfides linked to low concentrations of glutathione (GSH) in the nucleus of the human lens, we recently created the LEGSKO mouse in which lenticular GSH was lowered by knocking out the γ-glutamyl cysteine ligase subunit Gclc (20). These mice develop full-blown nuclear cataract by about 9 months and represent an important model for the development of drugs that might block or reverse the oxidation of crystallin sulfhydryls and presumably protein aggregation. However, this assumption in part depends on whether the sites of disulfide bond formation are similar in mouse and human age-related cataract. To test this hypothesis we performed the first comparative analysis of the cataract prone LEGSKO mouse and human aging and cataractous lens crystallin disulfidome, and compared the results with the disulfidome from mouse lens homogenate oxidized in vitro with H2O2 as a model of crystallin aggregation and opacification.  相似文献   

15.
Understanding the structural and assembly dynamics of the amyloid β-protein (Aβ) has direct relevance to the development of therapeutic agents for Alzheimer disease. To elucidate these dynamics, we combined scanning amino acid substitution with a method for quantitative determination of the Aβ oligomer frequency distribution, photo-induced cross-linking of unmodified proteins (PICUP), to perform “scanning PICUP.” Tyr, a reactive group in PICUP, was substituted at position 1, 10, 20, 30, or 40 (for Aβ40) or 42 (for Aβ42). The effects of these substitutions were probed using circular dichroism spectroscopy, thioflavin T binding, electron microscopy, PICUP, and mass spectrometry. All peptides displayed a random coil → α/β → β transition, but substitution-dependent alterations in assembly kinetics and conformer complexity were observed. Tyr1-substituted homologues of Aβ40 and Aβ42 assembled the slowest and yielded unusual patterns of oligomer bands in gel electrophoresis experiments, suggesting oligomer compaction had occurred. Consistent with this suggestion was the observation of relatively narrow [Tyr1]Aβ40 fibrils. Substitution of Aβ40 at the C terminus decreased the population conformational complexity and substantially extended the highest order of oligomers observed. This latter effect was observed in both Aβ40 and Aβ42 as the Tyr substitution position number increased. The ability of a single substitution (Tyr1) to alter Aβ assembly kinetics and the oligomer frequency distribution suggests that the N terminus is not a benign peptide segment, but rather that Aβ conformational dynamics and assembly are affected significantly by the competition between the N and C termini to form a stable complex with the central hydrophobic cluster.Alzheimer disease (AD)4 is the most common cause of late-life dementia (1) and is estimated to afflict more than 27 million people worldwide (2). An important etiologic hypothesis is that amyloid β-protein (Aβ) oligomers are the proximate neurotoxins in AD. Substantial in vivo and in vitro evidence supports this hypothesis (312). Neurotoxicity studies have shown that Aβ assemblies are potent neurotoxins (5, 1320), and the toxicity of some oligomers can be greater than that of the corresponding fibrils (21). Soluble Aβ oligomers inhibit hippocampal long term potentiation (4, 5, 13, 15, 17, 18, 22) and disrupt cognitive function (23). Compounds that bind and disrupt the formation of oligomers have been shown to block the neurotoxicity of Aβ (24, 25). Importantly, recent studies in higher vertebrates (dogs) have shown that substantial reduction in amyloid deposits in the absence of decreases in oligomer concentration has little effect on recovery of neurological function (26).Recent studies of Aβ oligomers have sought to correlate oligomer size and biological activity. Oligomers in the supernatants of fibril preparations centrifuged at 100,000 × g caused sustained calcium influx in rat hippocampal neurons, leading to calpain activation and dynamin 1 degradation (27). Aβ-derived diffusible ligand-like Aβ42 oligomers induced inflammatory responses in cultured rat astrocytes (28). A 90-kDa Aβ42 oligomer (29) has been shown to activate ERK1/2 in rat hippocampal slices (30) and bind avidly to human cortical neurons (31), in both cases causing apoptotic cell death. A comparison of the time dependence of the toxic effects of the 90-kDa assembly with that of Aβ-derived diffusible ligands revealed a 5-fold difference, Aβ-derived diffusible ligands requiring more time for equivalent effects (31). A 56-kDa oligomer, “Aβ*56,” was reported to cause memory impairment in middle-aged transgenic mice expressing human amyloid precursor protein (32). A nonamer also had adverse effects. Impaired long term potentiation in rat brain slices has been attributed to Aβ trimers identified in media from cultured cells expressing human amyloid precursor protein (33). Dimers and trimers from this medium also have been found to cause progressive loss of synapses in organotypic rat hippocampal slices (10). In mice deficient in neprilysin, an enzyme that has been shown to degrade Aβ in vivo (34), impairment in neuronal plasticity and cognitive function correlated with significant increases in Aβ dimer levels and synapse-associated Aβ oligomers (35).The potent pathologic effects of Aβ oligomers provide a compelling reason for elucidating the mechanism(s) of their formation. This has been a difficult task because of the metastability and polydispersity of Aβ assemblies (36). To obviate these problems, we introduced the use of the method of photo-induced cross-linking of unmodified proteins (PICUP) to rapidly (<1 s) and covalently stabilize oligomer mixtures (for reviews see Refs. 37, 38). Oligomers thus stabilized no longer exist in equilibrium with monomers or each other, allowing determination of oligomer frequency distributions by simple techniques such as SDS-PAGE (37). Recently, to obtain population-average information on contributions to fibril formation of amino acid residues at specific sites in Aβ, we employed a scanning intrinsic fluorescence approach (39). Tyr was used because it is a relatively small fluorophore, exists natively in Aβ, and possesses the side chain most reactive in the PICUP chemistry (40). Using this approach, we found that the central hydrophobic cluster region (Leu17–Ala21) was particularly important in controlling fibril formation of Aβ40, whereas the C terminus was the predominant structural element controlling Aβ42 assembly (39). Here we present results of studies in which key strategic features of the two methods have been combined to enable execution of “scanning PICUP” and the consequent revelation of site-specific effects on Aβ oligomerization.  相似文献   

16.
Clinically, amniotic membrane (AM) suppresses inflammation, scarring, and angiogenesis. AM contains abundant hyaluronan (HA) but its function in exerting these therapeutic actions remains unclear. Herein, AM was extracted sequentially with buffers A, B, and C, or separately by phosphate-buffered saline (PBS) alone. Agarose gel electrophoresis showed that high molecular weight (HMW) HA (an average of ∼3000 kDa) was predominantly extracted in isotonic Extract A (70.1 ± 6.0%) and PBS (37.7 ± 3.2%). Western blot analysis of these extracts with hyaluronidase digestion or NaOH treatment revealed that HMW HA was covalently linked with the heavy chains (HCs) of inter-α-inhibitor (IαI) via a NaOH-sensitive bond, likely transferred by the tumor necrosis factor-α stimulated gene-6 protein (TSG-6). This HC·HA complex (nHC·HA) could be purified from Extract PBS by two rounds of CsCl/guanidine HCl ultracentrifugation as well as in vitro reconstituted (rcHC·HA) by mixing HMW HA, serum IαI, and recombinant TSG-6. Consistent with previous reports, Extract PBS suppressed transforming growth factor-β1 promoter activation in corneal fibroblasts and induced mac ro phage apo pto sis. However, these effects were abolished by hyaluronidase digestion or heat treatment. More importantly, the effects were retained in the nHC·HA or rcHC·HA. These data collectively suggest that the HC·HA complex is the active component in AM responsible in part for clinically observed anti-inflammatory and anti-scarring actions.Hyaluronan (HA)4 is widely distributed in extracellular matrices, tissues, body fluids, and even in intracellular compartments (reviewed in Refs. 1 and 2). The molecular weight of HA ranges from 200 to 10,000 kDa depending on the source (3), but can also exist as smaller fragments and oligosaccharides under certain physiological or pathological conditions (1). Investigations over the last 15 years have suggested that low Mr HA can induce the gene expression of proinflammatory mediators and proangiogenesis, whereas high molecular weight (HMW) HA inhibits these processes (47).Several proteins have been shown to bind to HA (8) such as aggrecan (9), cartilage link protein (10), versican (11), CD44 (12, 13), inter-α-inhibitor (IαI) (14, 15), and tumor necrosis factor-α stimulated gene-6 protein (TSG-6) (16, 17). IαI consists of two heavy chains (HCs) (HC1 and HC2), both of which are linked through ester bonds to a chondroitin sulfate chain that is attached to the light chain, i.e. bikunin. Among all HA-binding proteins, only the HCs of IαI have been clearly demonstrated to be covalently coupled to HA (14, 18). However, TSG-6 has also been reported to form stable, possibly covalent, complexes with HA, either alone (19, 20) or when associated with HC (21).The formation of covalent bonds between HCs and HA is mediated by TSG-6 (2224) where its expression is often induced by inflammatory mediators such as tumor necrosis factor-α and interleukin-1 (25, 26). TSG-6 is also expressed in inflammatory-like processes, such as ovulation (21, 27, 28) and cervical ripening (29). TSG-6 interacts with both HA (17) and IαI (21, 24, 3033), and is essential for covalently transferring HCs on to HA (2224). The TSG-6-mediated formation of the HC·HA complex has been demonstrated to play a crucial role in female fertility in mice. The HC·HA complex is an integral part of an expanded extracellular “cumulus” matrix around the oocyte, which plays a critical role in successful ovulation and fertilization in vivo (22, 34). HC·HA complexes have also been found at sites of inflammation (3538) where its pro- or anti-inflammatory role remain arguable (39, 40).Immunostaining reveals abundant HA in the avascular stromal matrix of the AM (41, 42).5 In ophthalmology, cryopreserved AM has been widely used as a surgical graft for ocular surface reconstruction and exerts clinically observable actions to promote epithelial wound healing and to suppress inflammation, scarring, and angiogenesis (for reviews see Refs. 4345). However, it is not clear whether HA in AM forms HC·HA complex, and if so whether such an HC·HA complex exerts any of the above therapeutic actions. To address these questions, we extracted AM with buffers of increasing salt concentration. Because HMW HA was found to form the HC·HA complex and was mainly extractable by isotonic solutions, we further purified it from the isotonic AM extract and reconstituted it in vitro from three defined components, i.e. HMW HA, serum IαI, and recombinant TSG-6. Our results showed that the HC·HA complex is an active component in AM responsible for the suppression of TGF-β1 promoter activity, linkable to the scarring process noted before by AM (4648) and by the AM soluble extract (49), as well as for the promotion of macrophage death, linkable to the inflammatory process noted by AM (50) and the AM soluble extract (51).  相似文献   

17.
18.
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   

19.
20.
Proinflammatory cytokines induce nitric oxide-dependent DNA damage and ultimately β-cell death. Not only does nitric oxide cause β-cell damage, it also activates a functional repair process. In this study, the mechanisms activated by nitric oxide that facilitate the repair of damaged β-cell DNA are examined. JNK plays a central regulatory role because inhibition of this kinase attenuates the repair of nitric oxide-induced DNA damage. p53 is a logical target of JNK-dependent DNA repair; however, nitric oxide does not stimulate p53 activation or accumulation in β-cells. Further, knockdown of basal p53 levels does not affect DNA repair. In contrast, expression of growth arrest and DNA damage (GADD) 45α, a DNA repair gene that can be regulated by p53-dependent and p53-independent pathways, is stimulated by nitric oxide in a JNK-dependent manner, and knockdown of GADD45α expression attenuates the repair of nitric oxide-induced β-cell DNA damage. These findings show that β-cells have the ability to repair nitric oxide-damaged DNA and that JNK and GADD45α mediate the p53-independent repair of this DNA damage.Insulin-dependent diabetes mellitus is an autoimmune disease characterized by the selective destruction of insulin-secreting pancreatic β-cells found in the islets of Langerhans (1). Cytokines, released from invading leukocytes during insulitis, are believed to participate in the initial destruction of β-cells, precipitating the autoimmune response (2, 3). Treatment of rat islets with the macrophage-derived cytokine interleukin-1 (IL-1)2 results in the inhibition of glucose-stimulated insulin secretion and oxidative metabolism and in the induction of DNA damage that ultimately results in β-cell death (46). Nitric oxide, produced in micromolar levels following enhanced expression of the inducible nitric-oxide synthase in β-cells, mediates the damaging actions of cytokines on β-cell function (79). Nitric oxide inhibits insulin secretion by attenuating the oxidation of glucose to CO2, reducing cellular levels of ATP and, thereby, attenuating ATP-inhibited K+ channel activity (10, 11). The net effect is the inhibition of β-cell depolarization, calcium entry, and calcium-dependent exocytosis. In addition to the inhibition of β-cell function, nitric oxide induces DNA damage in β-cells (4, 12, 13). Nitric oxide or the oxidation products N2O3 and ONOO induce DNA damage through direct strand breaks and base modification (1416) and by inhibition of DNA repair enzymes, thereby enhancing the damaging actions of nitric oxide (17, 18).Recent studies have shown that β-cells maintain a limited ability to recover from cytokine-mediated damage (19, 20). The addition of a nitric-oxide synthase inhibitor to islets treated for 24 h with cytokine and continued culture with the nitric-oxide synthase inhibitor and cytokine results in a time-dependent restoration of insulin secretion, mitochondrial aconitase activity, and the repair of nitric oxide-damaged DNA (20, 21). Nitric oxide plays a dual role in modifying β-cell responses to cytokines. Nitric oxide induces β-cell damage and also activates a JNK-dependent recovery response that requires new gene expression (22). The ability of β-cells to recover from cytokine-mediated damage is temporally limited because cytokine-induced β-cell damage becomes irreversible following a 36-h incubation, and islets at this point are committed to degeneration (19).The purpose of this study was to determine the mechanisms by which β-cells repair nitric oxide-damaged DNA. Previous reports have shown that DNA damage induced by oxidizing agents, such as nitric oxide, is repaired through the base excision repair pathway (23), but how this pathway is activated in response to nitric oxide is unknown. Similar to the recovery of metabolic function, we now show that the activation of JNK by nitric oxide is required for repair of cytokine-induced DNA damage in β-cells. p53 is a logical candidate to mediate this repair because it plays a central role in DNA repair, is a target of JNK, and is activated by nitric oxide (2427). However, we show that cytokines do not stimulate p53 phosphorylation, and nitric oxide fails to stimulate p53 accumulation and phosphorylation. Growth arrest and DNA damage (GADD) 45α is a DNA damage-inducible gene that can be regulated by both p53-dependent and p53-independent mechanisms (2831). In contrast to p53, we show that cytokines stimulate GADD45α expression in a nitric oxide- and JNK-dependent manner and that siRNA-mediated knockdown of GADD45α results in an attenuation in the repair of nitric oxide-mediated DNA damage. These findings support a role for JNK in the regulation of GADD45α-dependent and p53-independent repair of nitric oxide-damaged β-cell DNA.  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