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.
Vasopressin controls renal water excretion largely through actions to regulate the water channel aquaporin-2 in collecting duct principal cells. Our knowledge of the mechanisms involved has increased markedly in recent years with the advent of methods for large-scale systems-level profiling such as protein mass spectrometry, yeast two-hybrid analysis, and oligonucleotide microarrays. Here we review this progress.Regulation of water excretion by the kidney is one of the most visible aspects of everyday physiology. An outdoor tennis game on a hot summer day can result in substantial water losses by sweating, and the kidneys respond by reducing water excretion. In contrast, excessive intake of water, a frequent occurrence in everyday life, results in excretion of copious amounts of clear urine. These responses serve to exact tight control on the tonicity of body fluids, maintaining serum osmolality in the range of 290–294 mosmol/kg of H2O through the regulated return of water from the pro-urine in the renal collecting ducts to the bloodstream.The importance of this process is highlighted when the regulation fails. For example, polyuria (rapid uncontrolled excretion of water) is a sometimes devastating consequence of lithium therapy for bipolar disorder. On the other side of the coin are water balance disorders that result from excessive renal water retention causing systemic hypo-osmolality or hyponatremia. Hyponatremia due to excessive water retention can be seen with severe congestive heart failure, hepatic cirrhosis, and the syndrome of inappropriate antidiuresis.The chief regulator of water excretion is the peptide hormone AVP,2 whereas the chief molecular target for regulation is the water channel AQP2. In this minireview, we describe new progress in the understanding of the molecular mechanisms involved in regulation of AQP2 by AVP in collecting duct cells, with emphasis on new information derived from “systems-level” approaches involving large-scale profiling and screening techniques such as oligonucleotide arrays, protein mass spectrometry, and yeast two-hybrid analysis. Most of the progress with these techniques is in the identification of individual molecules involved in AVP signaling and binding interactions with AQP2. Additional related issues are addressed in several recent reviews (14).  相似文献   

2.
3.
4.
5.
6.
7.
Although ligand-selective regulation of G protein-coupled receptor-mediated signaling and trafficking are well documented, little is known about whether ligand-selective effects occur on endogenous receptors or whether such effects modify the signaling response in physiologically relevant cells. Using a gene targeting approach, we generated a knock-in mouse line, in which N-terminal hemagglutinin epitope-tagged α2A-adrenergic receptor (AR) expression was driven by the endogenous mouse α2AAR gene locus. Exploiting this mouse line, we evaluated α2AAR trafficking and α2AAR-mediated inhibition of Ca2+ currents in native sympathetic neurons in response to clonidine and guanfacine, two drugs used for treatment of hypertension, attention deficit and hyperactivity disorder, and enhancement of analgesia through actions on the α2AAR subtype. We discovered a more rapid desensitization of Ca2+ current suppression by clonidine than guanfacine, which paralleled a more marked receptor phosphorylation and endocytosis of α2AAR evoked by clonidine than by guanfacine. Clonidine-induced α2AAR desensitization, but not receptor phosphorylation, was attenuated by blockade of endocytosis with concanavalin A, indicating a critical role for internalization of α2AAR in desensitization to this ligand. Our data on endogenous receptor-mediated signaling and trafficking in native cells reveal not only differential regulation of G protein-coupled receptor endocytosis by different ligands, but also a differential contribution of receptor endocytosis to signaling desensitization. Taken together, our data suggest that these HA-α2AAR knock-in mice will serve as an important model in developing ligands to favor endocytosis or nonendocytosis of receptors, depending on the target cell and pathophysiology being addressed.G protein-coupled receptors (GPCRs)4 represent the largest family of cell surface receptors mediating responses to hormones, cytokines, neurotransmitters, and therapeutic agents (1). In addition to regulating downstream signaling, ligand binding to a receptor can initiate phosphorylation of the active conformation of the receptor by G protein receptor kinases (GRKs) and subsequent binding of arrestins, thus restricting the magnitude and duration of the ligand-evoked signaling responses (2, 3). Binding of arrestins to GPCRs also leads to GPCR internalization (4, 5), a process that has been proposed as a means to desensitize receptor signaling at the cell surface, resensitize receptors, and/or initiate intracellular signaling (6, 7).Different ligands are able to induce distinct signaling and internalization profiles of the same receptor (8-14). However, the lack of available tools to study trafficking of endogenous GPCRs in native target cells has limited our understanding of ligand-selective endocytosis profiles and the relative contribution of receptor endocytosis to desensitization in native biological settings.To specifically test hypotheses regarding ligand-selective effects on GPCR internalization, and functional consequences of this trafficking on signaling, we utilized a homologous recombination gene targeting strategy to introduce a hemagglutinin (HA) epitope-tagged wild type α2A-adrenergic receptor (AR) into the mouse ADRA2A gene locus (“knock-in”). The α2AAR is a prototypical GPCR that couples to the Gi/o subfamily of G proteins (15). Studies on genetically engineered mice made null or mutant for the α2AAR have revealed that this subtype mediates the therapeutic effects of α2-adrenergic agents on blood pressure, pain perception, volatile anesthetic sparing, analgesia, and working memory enhancement (16-18). Two classic α2-ligands, clonidine and guanfacine, have been widely used to treat hypertension (19), attention deficit and hyperactivity disorder (20), and to elicit analgesia (19, 21) mediated via the α2AAR. Clinically guanfacine has a much longer duration of action than clonidine (22-24); this longer duration of action cannot be accounted for by the pharmacokinetic profile of these agents in human beings, as both drugs have a half-life of 12-14 h (25, 26). Because ligand-induced desensitization and trafficking of GPCRs have been implicated as critical mechanisms for modulating response duration in vivo (3), one hypothesis underlying the difference in duration between clonidine and guanfacine is that clonidine provokes accelerated desensitization of the α2AAR via one or several mechanisms, whereas guanfacine does not. Signaling desensitization in response to these two agonists has not been compared under the same experimental settings. To specifically test this hypothesis, we have exploited our HA-α2AAR knock-in mice so that we could examine these properties of guanfacine and clonidine in native target cells.We compared internalization of the α2AAR and inhibition of Ca2+ currents induced by clonidine and guanfacine in primary superior cervical ganglia (SCG) neurons, where the α2AAR is the major adrenergic receptor subtype controlling norepinephrine release and sympathetic tone (17, 27). Our data revealed a differential regulation of α2AAR trafficking and signaling duration by clonidine versus guanfacine, i.e. clonidine induced a more dramatic desensitization of the α2AAR than guanfacine, and this desensitization was largely because of α2AAR internalization. These studies reveal the powerful tool that the HA-α2AAR knock-in mice provide for identifying endocytosis-dependent and -independent physiological phenomena for this receptor subtype as a first step in defining novel loci for therapeutic intervention in the α2AAR signaling/trafficking cascade.  相似文献   

8.
Recently we reported that N-glycans on the β-propeller domain of the integrin α5 subunit (S-3,4,5) are essential for α5β1 heterodimerization, expression, and cell adhesion. Herein to further investigate which N-glycosylation site is the most important for the biological function and regulation, we characterized the S-3,4,5 mutants in detail. We found that site-4 is a key site that can be specifically modified by N-acetylglucosaminyltransferase III (GnT-III). The introduction of bisecting GlcNAc into the S-3,4,5 mutant catalyzed by GnT-III decreased cell adhesion and migration on fibronectin, whereas overexpression of N-acetylglucosaminyltransferase V (GnT-V) promoted cell migration. The phenomenon is similar to previous observations that the functions of the wild-type α5 subunit were positively and negatively regulated by GnT-V and GnT-III, respectively, suggesting that the α5 subunit could be duplicated by the S-3,4,5 mutant. Interestingly GnT-III specifically modified the S-4,5 mutant but not the S-3,5 mutant. This result was confirmed by erythroagglutinating phytohemagglutinin lectin blot analysis. The reduction in cell adhesion was consistently observed in the S-4,5 mutant but not in the S-3,5 mutant cells. Furthermore mutation of site-4 alone resulted in a substantial decrease in erythroagglutinating phytohemagglutinin lectin staining and suppression of cell spread induced by GnT-III compared with that of either the site-3 single mutant or wild-type α5. These results, taken together, strongly suggest that N-glycosylation of site-4 on the α5 subunit is the most important site for its biological functions. To our knowledge, this is the first demonstration that site-specific modification of N-glycans by a glycosyltransferase results in functional regulation.Glycosylation is a crucial post-translational modification of most secreted and cell surface proteins (1). Glycosylation is involved in a variety of physiological and pathological events, including cell growth, migration, differentiation, and tumor invasion. It is well known that glycans play important roles in cell-cell communication, intracellular signal transduction, protein folding, and stability (2, 3).Integrins comprise a family of receptors that are important for cell adhesion. The major function of integrins is to connect cells to the extracellular matrix, activate intracellular signaling pathways, and regulate cytoskeletal formation (4). Integrin α5β1 is well known as a fibronectin (FN)3 receptor. The interaction between integrin α5 and FN is essential for cell migration, cell survival, and development (58). In addition, integrins are N-glycan carrier proteins. For example, α5β1 integrin contains 14 and 12 putative N-glycosylation sites on the α5 and β1 subunits, respectively. Several studies suggest that N-glycosylation is essential for functional integrin α5β1. When human fibroblasts were cultured in the presence of 1-deoxymannojirimycin, which prevents N-linked oligosaccharide processing, immature α5β1 integrin appeared on the cell surface, and FN-dependent adhesion was greatly reduced (9). Treatment of purified integrin α5β1 with N-glycosidase F, which cleaves between the innermost N-acetylglucosamine (GlcNAc) and asparagine N-glycan residues of N-linked glycoproteins, prevented the inherent association between subunits and blocked α5β1 binding to FN (10).A growing body of evidence indicates that the presence of the appropriate oligosaccharide can modulate integrin activation. N-Acetylglucosaminyltransferase III (GnT-III) catalyzes the addition of GlcNAc to mannose that is β1,4-linked to an underlying N-acetylglucosamine, producing what is known as a “bisecting” GlcNAc linkage as shown in Fig. 1B. GnT-III is generally regarded as a key glycosyltransferase in N-glycan biosynthetic pathways and contributes to inhibition of metastasis. The introduction of a bisecting GlcNAc catalyzed by GnT-III suppresses additional processing and elongation of N-glycans. These reactions, which are catalyzed in vitro by other glycosyltransferases, such as N-acetylglucosaminyltransferase V (GnT-V), which catalyzes the formation of β1,6 GlcNAc branching structures (Fig. 1B) and plays important roles in tumor metastasis, do not proceed because the enzymes cannot utilize the bisected N-glycans as a substrate. Introduction of the bisecting GlcNAc to integrin α5 by overexpression of GnT-III resulted in decreased in ligand binding and down-regulation of cell adhesion and migration (1113). Contrary to the functions of GnT-III, overexpression of GnT-V promoted integrin α5β1-mediated cell migration on FN (14). These observations clearly demonstrate that the alteration of N-glycan structure affected the biological functions of integrin α5β1. Similarly characterization of the carbohydrate moieties in integrin α3β1 from non-metastatic and metastatic human melanoma cell lines showed that expression of β1,6 GlcNAc branched structures was higher in metastatic cells compared with non-metastatic cells, confirming the notion that the β1,6 GlcNAc branched structure confers invasive and metastatic properties to cancer cells. In fact, Partridge et al. (15) reported that GnT-V-modified N-glycans containing poly-N-acetyllactosamine, the preferred ligand for galectin-3, on surface receptors oppose their constitutive endocytosis, promoting intracellular signaling and consequently cell migration and tumor metastasis.Open in a separate windowFIGURE 1.Potential N-glycosylation sites on the α5 subunit and its modification by GnT-III and GnT-V. A, schematic diagram of potential N-glycosylation sites on the α5 subunit. Putative N-glycosylation sites are indicated by triangles, and point mutations are indicated by crosses (N84Q, N182Q, N297Q, N307Q, N316Q, N524Q, N530Q, N593Q, N609Q, N675Q, N712Q, N724Q, N773Q, and N868Q). B, illustration of the reaction catalyzed by GnT-III and GnT-V. Square, GlcNAc; circle, mannose. TM, transmembrane domain.In addition, sialylation on the non-reducing terminus of N-glycans of α5β1 integrin plays an important role in cell adhesion. Colon adenocarcinomas express elevated levels of α2,6 sialylation and increased activity of ST6GalI sialyltransferase. Elevated ST6GalI positively correlated with metastasis and poor survival. Therefore, ST6GalI-mediated hypersialylation likely plays a role in colorectal tumor invasion (16, 17). In fact, oncogenic ras up-regulated ST6GalI and, in turn, increased sialylation of β1 integrin adhesion receptors in colon epithelial cells (18). However, this is not always the case. The expression of hyposialylated integrin α5β1 was induced by phorbol esterstimulated differentiation in myeloid cells in which the expression of the ST6GalI was down-regulated by the treatment, increasing FN binding (19). A similar phenomenon was also observed in hematopoietic or other epithelial cells. In these cells, the increased sialylation of the β1 integrin subunit was correlated with reduced adhesiveness and metastatic potential (2022). In contrast, the enzymatic removal of α2,8-linked oligosialic acids from the α5 integrin subunit inhibited cell adhesion to FN (23). Collectively these findings suggest that the interaction of integrin α5β1 with FN is dependent on its N-glycosylation and the processing status of N-glycans.Because integrin α5β1 contains multipotential N-glycosylation sites, it is important to determine the sites that are crucial for its biological function and regulation. Recently we found that N-glycans on the β-propeller domain (sites 3, 4, and 5) of the integrin α5 subunit are essential for α5β1 heterodimerization, cell surface expression, and biological function (24). In this study, to further investigate the underlying molecular mechanism of GnT-III-regulated biological functions, we characterized the N-glycans on the α5 subunit in detail using genetic and biochemical approaches and found that site-4 is a key site that can be specifically modified by GnT-III.  相似文献   

9.
10.
Several neurological diseases, including Parkinson disease and dementia with Lewy bodies, are characterized by the accumulation of α-synuclein phosphorylated at Ser-129 (p-Ser-129). The kinase or kinases responsible for this phosphorylation have been the subject of intense investigation. Here we submit evidence that polo-like kinase 2 (PLK2, also known as serum-inducible kinase or SNK) is a principle contributor to α-synuclein phosphorylation at Ser-129 in neurons. PLK2 directly phosphorylates α-synuclein at Ser-129 in an in vitro biochemical assay. Inhibitors of PLK kinases inhibited α-synuclein phosphorylation both in primary cortical cell cultures and in mouse brain in vivo. Finally, specific knockdown of PLK2 expression by transduction with short hairpin RNA constructs or by knock-out of the plk2 gene reduced p-Ser-129 levels. These results indicate that PLK2 plays a critical role in α-synuclein phosphorylation in central nervous system.The importance of α-synuclein to the pathogenesis of Parkinson disease (PD)4 and the related disorder, dementia with Lewy bodies (DLB), is suggested by its association with Lewy bodies and Lewy neurites, the inclusions that characterize these diseases (13), and demonstrated by the existence of mutations that cause syndromes mimicking sporadic PD and DLB (46). Furthermore, three separate mutations cause early onset forms of PD and DLB. It is particularly telling that duplications or triplications of the gene (79), which increase levels of α-synuclein with no alteration in sequence, also cause PD or DLB.α-Synuclein has been reported to be phosphorylated on serine residues, at Ser-87 and Ser-129 (10), although to date only the Ser-129 phosphorylation has been identified in the central nervous system (11, 12). Phosphorylation at tyrosine residues has been observed by some investigators (13, 14) but not by others (1012). Phosphorylation at Ser-129 (p-Ser-129) is of particular interest because the majority of synuclein in Lewy bodies contains this modification (15). In addition, p-Ser-129 was found to be the most extensive and consistent modification in a survey of synuclein in Lewy bodies (11). Results have been mixed from studies investigating the function of phosphorylation using S129A and S129D mutations to respectively block and mimic the modification. Although the phosphorylation mimic was associated with pathology in studies in Drosophila (16) and in transgenic mouse models (17, 18), studies using adeno-associated virus vectors to overexpress α-synuclein in rat substantia nigra found an exacerbation of pathology with the S129A mutation, whereas the S129D mutation was benign, if not protective (19). Interpretation of these studies is complicated by a recent study showing that the S129D and S129A mutations themselves have effects on the aggregation properties of α-synuclein independent of their effects on phosphorylation, with the S129A mutation stimulating fibril formation (20). Clearly, determination of the role of p-Ser-129 phosphorylation would be helped by identification of the responsible kinase. In addition, identification will provide a pathologically relevant way to increase phosphorylation in a cell or animal model.Several kinases have been proposed to phosphorylate α-synuclein, including casein kinases 1 and 2 (10, 12, 21) and members of the G-protein-coupled receptor kinase family (22). In this report, we offer evidence that a member of the polo-like kinase (PLK) family, PLK2 (or serum-inducible kinase, SNK), functions as an α-synuclein kinase. The ability of PLK2 to directly phosphorylate α-synuclein at Ser-129 is established by overexpression in cell culture and by in vitro reaction with the purified kinase. We show that PLK2 phosphorylates α-synuclein in cells, including primary neuronal cultures, using a series of kinase inhibitors as well as inhibition of expression with RNA interference. In addition, inhibitor and knock-out studies in mouse brain support a role for PLK2 as an α-synuclein kinase in vivo.  相似文献   

11.
12.
13.
14.
Autophagy is a degradative process that recycles long-lived and faulty cellular components. It is linked to many diseases and is required for normal development. ULK1, a mammalian serine/threonine protein kinase, plays a key role in the initial stages of autophagy, though the exact molecular mechanism is unknown. Here we report identification of a novel protein complex containing ULK1 and two additional protein factors, FIP200 and ATG13, all of which are essential for starvation-induced autophagy. Both FIP200 and ATG13 are critical for correct localization of ULK1 to the pre-autophagosome and stability of ULK1 protein. Additionally, we demonstrate by using both cellular experiments and a de novo in vitro reconstituted reaction that FIP200 and ATG13 can enhance ULK1 kinase activity individually but both are required for maximal stimulation. Further, we show that ATG13 and ULK1 are phosphorylated by the mTOR pathway in a nutrient starvation-regulated manner, indicating that the ULK1·ATG13·FIP200 complex acts as a node for integrating incoming autophagy signals into autophagosome biogenesis.Macroautophagy (herein referred to as autophagy) is a catabolic process whereby long-lived proteins and damaged organelles are shuttled to lysosomes for degradation. This process is conserved in all eukaryotes. Under normal growth conditions a housekeeping level of autophagy exists. Under stress, such as nutrient starvation, autophagy is strongly induced resulting in the engulfment of cytosolic components and organelles in specialized double-membrane structures termed autophagosomes. Following fusion of the outer autophagosomal membrane with lysosomes, the inner membrane and its cytoplasmic cargo are degraded and recycled (13). Recent work has implicated autophagy in many disease pathologies, including cancer, neurodegeneration, as well as in eliminating intracellular pathogens (48).The morphology of autophagy was first described in mammalian cells over 50 years ago (9). However, it is only recently through yeast genetic screens, that multiple autophagy-related (ATG) genes have been identified (1012). The yeast ATG proteins have been classified into four major groups: the Atg1 protein kinase complex, the Vps34 phosphatidylinositol 3-phosphate kinase complex, the Atg8/Atg12 conjugation systems, and the Atg9 recycling complex (13). Even though many ATG genes are now known, most of which have functional homologs in mammalian cells (14, 15), the molecular mechanism by which they sense the initial triggers and subsequently dictate autophagy-specific intracellular membrane events is far from understood.In yeast, one of the earliest autophagy-specific events is believed to involve the Atg1 protein kinase complex. Atg1 is a serine/threonine protein kinase and a key autophagy-regulator (16). Atg1 is complexed to at least two other proteins during autophagy, Atg13 and Atg17, both of which are required for normal Atg1 function and autophagosome generation (1719). Classical signaling pathways such as the cAMP-dependent kinase (PKA) pathway or the Tor kinase pathway appear to converge upon this complex, placing Atg1 at an early stage during autophagosome biogenesis (2022). Atg1 phosphorylation by PKA blocks its association with the forming autophagosome (21), while the Tor pathway hyperphosphorylates Atg13 causing a reduced affinity of Atg13 for Atg1, resulting in repression of autophagy (17, 19). In contrast, nutrient starvation or inhibition of Tor leads to dephosphorylation of Atg13 thus increased Atg1 complex formation and kinase activity, resulting in stimulation of autophagy (19). Surprisingly, the physiological substrates of Atg1 kinase have not been identified; thus how Atg1 transduces upstream autophagic signaling is undefined. Recently, mammalian homologs of Atg1 have been identified as ULK1 and ULK2 (Unc-51-like kinase)2 (2325). ULK1 and ULK2 are ubiquitously expressed and localize to the isolation membrane, or forming autophagosome, upon nutrient starvation (25); RNAi-mediated depletion of ULK1 in HEK293 cells compromises autophagy (23, 24). The exact role of ULK1 versus ULK2 in autophagy is unclear, and it is possible some redundancy exists between the two isoforms (26).Given the conservation of autophagy from yeast to man, it is interesting to note that no mammalian counterpart to yeast Atg13 or Atg17 had been identified until very recently. The protein FIP200 (focal adhesion kinase family-interacting protein of 200 kDa) was identified as an autophagy-essential binding partner of both ULK1 and ULK2 (25), and it has been speculated that FIP200 might be the equivalent of yeast Atg17, despite low sequence similarity (25, 27).In this study, we delve deeper into the molecular regulation of ULK1 to gain a better insight into how mammalian signaling pathways affect autophagy initiation. We describe here the identification of a triple complex consisting of ULK1, FIP200, and the mammalian equivalent of Atg13. This complex is required not only for localization of ULK1 to the isolation membrane but also for maximal kinase activity. In addition, both ATG13 and ULK1 are kinase substrates in the mTOR pathway and thus might function to sense nutrient starvation. Therefore, this study defines the role of mammalian ULK1-ATG13-FIP200 complex in mediating the initial autophagic triggers and to transduce the signal to the core autophagic machinery.  相似文献   

15.
The actin cytoskeleton has an important role in the organization and function of the immune synapse during antigen recognition. Dynamic rearrangement of the actin cytoskeleton in response to T cell receptor (TCR) triggering requires the coordinated activation of Rho family GTPases that cycle between active and inactive conformations. This is controlled by GTPase-activating proteins (GAP), which regulate inactivation of Rho GTPases, and guanine exchange factors, which mediate their activation. Whereas much attention has centered on guanine exchange factors for Rho GTPases in T cell activation, the identity and functional roles of the GAP in this process are largely unknown. We previously reported β2-chimaerin as a diacylglycerol-regulated Rac-GAP that is expressed in T cells. We now demonstrate Lck-dependent phosphorylation of β2-chimaerin in response to TCR triggering. We identify Tyr-153 as the Lck-dependent phosphorylation residue and show that its phosphorylation negatively regulates membrane stabilization of β2-chimaerin, decreasing its GAP activity to Rac. This study establishes the existence of TCR-dependent regulation of β2-chimaerin and identifies a novel mechanism for its inactivation.T cell activation requires presentation of an antigen by antigen-presenting cells (APC)2 to the T cell receptor (TCR); this event involves the reorganization of several scaffolds and signaling proteins, leading to formation of the immunological synapse (IS) (1). Correct protein redistribution during synapse formation is critical for an efficient T cell response, and it is largely regulated by actin polymerization at the T cell/APC contact site as a result of TCR-regulated Rac-dependent signals (2, 3). Like other Rho GTPases, Rac cycles between a GTP-bound active state and a GDP-bound inactive state. This continuous recycling is regulated by the concerted action of two proteins as follows: GEF, which activates Rac by mediating GDP/GTP exchange (4), and GAP, which induces Rac inactivation by accelerating intrinsic Rac GTPase activity, converting GTP to GDP (5).Vav-1 is the best studied GEF for Rac, and it has critical roles in T cell-dependent functions (6). In naive, unstimulated T cells, Vav-1 is in an inactive state through autoinhibition, as the GEF domain is blocked by the N-terminal region (7). This autoinhibition is relieved by TCR-mediated tyrosine phosphorylation (8, 9). Thymocytes from Vav-1-deficient mice have a developmental block, and their mature T cells show severe defects in IS formation, as well as reduced Ca2+ influx, IL-2 production, T cell proliferation, and cytotoxic activity (1013). Although several studies have shown a key role for Vav-1, the mechanisms that govern Rac inactivation downstream of the TCR remain elusive.The chimaerins are a family of Rho-GAP, with specific activity for Rac. In addition to their catalytic domain, they have an N-terminal SH2 domain and a C1 domain required for interaction with the lipid messenger diacylglycerol (DAG) and with phorbol esters (14). There are two mammalian chimaerin genes (CHN1 and CHN2), which encode the full-lengthα2-(ARHGAP2) and β2-chimaerins (ARHGAP3), and at least one splice variant each (α1 and β1) that lack the SH2 domain. The α-chimaerins are expressed abundantly in brain and are linked to neuritogenesis and axon guidance (1520). β2-Chimaerin expression is ubiquitous (21) and is involved in EGF-dependent Rac regulation (22, 23). Experiments in T cells showed that β2-chimaerin participates in chemokine-dependent regulation of T cell migration and adhesion (24). A very recent study implicates chimaerins in the modulation of Rac activity during T cell synapse formation, suggesting that this protein family contributes to DAG-mediated regulation of cytoskeletal remodeling during T cell activation (25).Determination of the β2-chimaerin crystal structure provided important clues regarding its mechanism of action. In the absence of stimulation, the protein is in an inactive state in which the N-terminal domain maintains a “closed” conformation, blocking Rac binding and concealing the C1 domain (26). These structural data were fully supported by experiments in live T lymphocytes showing that phorbol myristate acetate (PMA)-dependent translocation of β2-chimaerin was less effective than that of its isolated C1 domain (24). These data not only confirmed the lack of accessibility of the β2-chimaerin C1 domain but also suggested that there are negative regulatory mechanisms that promote β2-chimaerin release from the membrane.DAG-dependent signaling is critical for the modulation of T cell functions, by virtue of its ability to bind and regulate C1 domain-containing proteins such as protein kinase Cθ, protein kinase D, and RasGRP1 (27). An important issue is to determine how the different DAG-binding proteins discriminate between distinct DAG pools, and how DAG activates certain C1-containing proteins and not others. Some mechanisms that allow discrimination between DAG receptors include the distinct affinity of C1 domains for different DAG pools, association of C1 domain-containing proteins to specific scaffolds, and/or structural determinants in these proteins that limit C1 domain accessibility to membrane DAG (2830).To explore the events that contribute to the specific regulation of β2-chimaerin, we studied β2-chimaerin phosphorylation in the context of TCR stimulation. We show that β2-chimaerin is phosphorylated in tyrosine residues after TCR stimulation, and we identify Lck as the Tyr kinase responsible for this phosphorylation. Generation of point mutants identified Tyr-153, at the hinge of the SH2 and C1 domains, as the main tyrosine residue phosphorylated in response to TCR stimulation. Cells expressing a β2-chimaerin mutant defective for Tyr-153 phosphorylation show anomalies in TCR clustering, conjugate formation, NF-AT activation, and IL-2 production that correlate with elevated Rac-GAP activity in this mutant. Subcellular localization analysis of the β2-chimaerin mutants reveals that impairment of β2-chimaerin phosphorylation at Tyr-153 promotes C1-mediated β2-chimaerin stabilization at the plasma membrane, providing a mechanistic explanation for its higher Rac-GAP activity. In summary, our results demonstrate for the first time that tyrosine kinase-mediated negative regulation of β2-chimaerin is elicited by physiological stimulation in T lymphocytes, and suggest that TCR stimulation provides both positive and negative signals for β2-chimaerin activation.  相似文献   

16.
17.
18.
γ-Secretase is known to play a pivotal role in the pathogenesis of Alzheimer disease through production of amyloidogenic Aβ42 peptides. Early onset familial Alzheimer disease mutations in presenilin (PS), the catalytic core of γ-secretase, invariably increase the Aβ42:Aβ40 ratio. However, the mechanism by which these mutations affect γ-secretase complex formation and cleavage specificity is poorly understood. We show that our in vitro assay system recapitulates the effect of PS1 mutations on the Aβ42:Aβ40 ratio observed in cell and animal models. We have developed a series of small molecule affinity probes that allow us to characterize active γ-secretase complexes. Furthermore we reveal that the equilibrium of PS1- and PS2-containing active complexes is dynamic and altered by overexpression of Pen2 or PS1 mutants and that formation of PS2 complexes is positively correlated with increased Aβ42:Aβ40 ratios. These data suggest that perturbations to γ-secretase complex equilibrium can have a profound effect on enzyme activity and that increased PS2 complexes along with mutated PS1 complexes contribute to an increased Aβ42:Aβ40 ratio.β-Amyloid (Aβ)5 peptides are believed to play a causative role in Alzheimer disease (AD). Aβ peptides are generated from the processing of the amyloid precursor protein (APP) by two proteases, β-secretase and γ-secretase. Although γ-secretase generates heterogenous Aβ peptides ranging from 37 to 46 amino acids in length, significant work has focused mainly on the Aβ40 and Aβ42 peptides that are the major constituents of amyloid plaques. γ-Secretase is a multisubunit membrane aspartyl protease comprised of at least four known subunits: presenilin (PS), nicastrin (Nct), anterior pharynx-defective (Aph), and presenilin enhancer 2 (Pen2). Presenilin is thought to contain the catalytic core of the complex (14), whereas Aph and Nct play critical roles in the assembly, trafficking, and stability of γ-secretase as well as substrate recognition (5, 6). Lastly Pen2 facilitates the endoproteolysis of PS into its N-terminal (NTF) and C-terminal (CTF) fragments thereby yielding a catalytically competent enzyme (5, 710). All four proteins (PS, Nct, Aph1, and Pen2) are obligatory for γ-secretase activity in cell and animal models (11, 12). There are two homologs of PS, PS1 and PS2, and three isoforms of Aph1, Aph1aS, Aph1aL, and Aph1b. At least six active γ-secretase complexes have been reported (two presenilins × three Aph1s) (13, 14). The sum of apparent molecular masses of the four proteins (PS1-NTF/CTF ≈ 53 kDa, Nct ≈ 120 kDa, Aph1 ≈ 30 kDa, and Pen2 ≈ 10kDa) is ∼200 kDa. However, active γ-secretase complexes of varying sizes, ranging from 250 to 2000 kDa, have been reported (1519). Recently a study suggested that the γ-secretase complex contains only one of each subunit (20). Collectively these studies suggest that a four-protein complex around 200–250 kDa may be the minimal functional γ-secretase unit with additional cofactors and/or varying stoichiometry of subunits existing in the high molecular weight γ-secretase complexes. CD147 and TMP21 have been found to be associated with the γ-secretase complex (21, 22); however, their role in the regulation of γ-secretase has been controversial (23, 24).Mutations of PS1 or PS2 are associated with familial early onset AD (FAD), although it is debatable whether these familial PS mutations act as “gain or loss of function” alterations in regard to γ-secretase activity (2527). Regardless the overall outcome of these mutations is an increased ratio of Aβ42:Aβ40. Clearly these mutations differentially affect γ-secretase activity for the production of Aβ40 and Aβ42. Despite intensive studies of Aβ peptides and γ-secretase, the molecular mechanism controlling the specificity of γ-secretase activity for Aβ40 and Aβ42 production has not been resolved. It has been found that PS1 mutations affect the formation of γ-secretase complexes (28). However, the precise mechanism by which individual subunits alter the dynamics of γ-secretase complex formation and activity is largely unresolved. A better mechanistic understanding of γ-secretase activity associated with FAD mutations has been hindered by the lack of suitable assays and probes that are necessary to recapitulate the effect of these mutations seen in cell models and to characterize the active γ-secretase complex.In our present studies, we have determined the overall effect of Pen2 and PS1 expression on the dynamics of PS1- and PS2-containing complexes and their association with γ-secretase activity. Using newly developed biotinylated small molecular probes and activity assays, we revealed that expression of Pen2 or PS1 FAD mutants markedly shifts the equilibrium of PS1-containing active complexes to that of PS2-containing complexes and results in an overall increase in the Aβ42:Aβ40 ratio in both stable cell lines and animal models. Our studies indicate that perturbations to the equilibrium of active γ-secretase complexes by an individual subunit can greatly affect the activity of the enzyme. Moreover they serve as further evidence that there are multiple and distinct γ-secretase complexes that can exist within the same cells and that their equilibrium is dynamic. Additionally the affinity probes developed here will facilitate further study of the expression and composition of endogenous active γ-secretase from a variety of model systems.  相似文献   

19.
20.
Mammalian defensins are cationic antimicrobial peptides that play a central role in host innate immunity and as regulators of acquired immunity. In animals, three structural defensin subfamilies, designated as α, β, and θ, have been characterized, each possessing a distinctive tridisulfide motif. Mature α- and β-defensins are produced by simple proteolytic processing of their prepropeptide precursors. In contrast, the macrocyclic θ-defensins are formed by the head-to-tail splicing of nonapeptides excised from a pair of prepropeptide precursors. Thus, elucidation of the θ-defensin biosynthetic pathway provides an opportunity to identify novel factors involved in this unique process. We incorporated the θ-defensin precursor, proRTD1a, into a bait construct for a yeast two-hybrid screen that identified rhesus macaque stromal cell-derived factor 2-like protein 1 (SDF2L1), as an interactor. SDF2L1 is a component of the endoplasmic reticulum (ER) chaperone complex, which we found to also interact with α- and β-defensins. However, analysis of the SDF2L1 domain requirements for binding of representative α-, β-, and θ-defensins revealed that α- and β-defensins bind SDF2L1 similarly, but differently from the interactions that mediate binding of SDF2L1 to pro-θ-defensins. Thus, SDF2L1 is a factor involved in processing and/or sorting of all three defensin subfamilies.Mammalian defensins are tridisulfide-containing antimicrobial peptides that contribute to innate immunity in all species studied to date. Defensins are comprised of three structural subfamilies: the α-, β-, and θ-defensins (1). α- and β-Defensins are peptides of about 29–45-amino acid residues with similar three-dimensional structures. Despite their similar tertiary conformations, the disulfide motifs of α- and β-defensins differ. Expression of human α-defensins is tissue-specific. Four myeloid α-defensins (HNP1–4) are expressed predominantly by neutrophils and monocytes wherein they are packaged in granules, while two enteric α-defensins (HD-5 and HD-6) are expressed at high levels in Paneth cells of the small intestine. Myeloid α-defensins constitute about 5% of the protein mass of human neutrophils. HNPs are discharged into the phagosome during phagocytic ingestion of microbial particles. HD-5 and HD-6 are produced and stored as propeptides in Paneth cell granules and are processed extracellularly by intestinal trypsin (2). β-Defensins are produced primarily by various epithelia (e.g. skin, urogenital tract, airway) and are secreted by the producing cells in their mature forms. In contrast to pro-α-defensins, which contain a conserved prosegment of ∼40 amino acids, the prosegments in β-defensins vary in length and sequence. θ-Defensins are found only in Old World monkeys and orangutans and are the only known circular peptides in animals. These 18-residue macrocyclic peptides are formed by ligation of two nonamer sequences excised from two precursor polypeptides, which are truncated versions of ancestral α-defensins. Like myeloid α-defensins, θ-defensins are stored primarily in neutrophil and monocyte granules (3).Numerous laboratories have demonstrated that the antimicrobial properties of defensins derive from their ability to bind and disrupt target cell membranes (4), and studies have shown defensins to be active against Gram-positive and -negative bacteria (5), viruses (69), fungi (10, 11), and parasites such as Giardia lamblia (12). Defensins also play a regulatory role in acquired immunity as they are known to chemoattract T lymphocytes, monocytes, and immature dendritic cells (13, 14), act as adjuvants, stimulate B cell responses, and up-regulate proliferation and cytokine production by spleen cells and T helper cells (15, 16).Defensins are produced as pre-propeptides and undergo post-translational processing to form the mature peptides. While much has been learned about regulation of defensin expression, little is known about the factors involved in their biosynthesis. Valore and Ganz (17) investigated the processing of defensins in cultured cells and demonstrated that maturation of HNPs occurs through two proteolytic steps that lead to formation of mature α-defensins, but the proteases involved have yet to be identified. Moreover, there are virtually no published data regarding endoplasmic reticulum (ER)2 factors that are responsible for the folding, processing, and sorting steps necessary for defensin maturation and secretion or trafficking to the proper subcellular compartment. It is likely that several chaperones, proteases, and protein-disulfide isomerase (PDI) family proteins are involved. Consistent with this possibility, Gruber et al. (18) recently demonstrated the role of a PDI in biosynthesis of cyclotides, small ∼30-residue macrocyclic peptides produced by plants.The primary structures of α- and θ-defensin precursors are closely related. We therefore undertook studies to identify proteins that interact with representative propeptides of each defensin subfamily with the goal of determining common and unique processes that regulate biosynthesis of α- and θ-defensins. We used two-hybrid analysis to first identify interactors of the θ-defensin precursor, proRTD1a. As described, we identified SDF2L1, a component of the ER-chaperone complex as an interactor, and showed that it also specifically interacts with α- and β-defensins. This suggests that SDF2L1 is involved in the maturation/trafficking of defensins at a step common to all three subfamilies of mammalian defensins.  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