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 579 毫秒
1.
The acid-sensing ion channel 1a (ASIC1a) is widely expressed in central and peripheral neurons where it generates transient cation currents when extracellular pH falls. ASIC1a confers pH-dependent modulation on postsynaptic dendritic spines and has critical effects in neurological diseases associated with a reduced pH. However, knowledge of the proteins that interact with ASIC1a and influence its function is limited. Here, we show that α-actinin, which links membrane proteins to the actin cytoskeleton, associates with ASIC1a in brain and in cultured cells. The interaction depended on an α-actinin-binding site in the ASIC1a C terminus that was specific for ASIC1a versus other ASICs and for α-actinin-1 and -4. Co-expressing α-actinin-4 altered ASIC1a current density, pH sensitivity, desensitization rate, and recovery from desensitization. Moreover, reducing α-actinin expression altered acid-activated currents in hippocampal neurons. These findings suggest that α-actinins may link ASIC1a to a macromolecular complex in the postsynaptic membrane where it regulates ASIC1a activity.Acid-sensing ion channels (ASICs)2 are H+-gated members of the DEG/ENaC family (13). Members of this family contain cytosolic N and C termini, two transmembrane domains, and a large cysteine-rich extracellular domain. ASIC subunits combine as homo- or heterotrimers to form cation channels that are widely expressed in the central and peripheral nervous systems (14). In mammals, four genes encode ASICs, and two subunits, ASIC1 and ASIC2, have two splice forms, a and b. Central nervous system neurons express ASIC1a, ASIC2a, and ASIC2b (57). Homomeric ASIC1a channels are activated when extracellular pH drops below 7.2, and half-maximal activation occurs at pH 6.5–6.8 (810). These channels desensitize in the continued presence of a low extracellular pH, and they can conduct Ca2+ (9, 1113). ASIC1a is required for acid-evoked currents in central nervous system neurons; disrupting the gene encoding ASIC1a eliminates H+-gated currents unless extracellular pH is reduced below pH 5.0 (5, 7).Previous studies found ASIC1a enriched in synaptosomal membrane fractions and present in dendritic spines, the site of excitatory synapses (5, 14, 15). Consistent with this localization, ASIC1a null mice manifested deficits in hippocampal long term potentiation, learning, and memory, which suggested that ASIC1a is required for normal synaptic plasticity (5, 16). ASICs might be activated during neurotransmission when synaptic vesicles empty their acidic contents into the synaptic cleft or when neuronal activity lowers extracellular pH (1719). Ion channels, including those at the synapse often interact with multiple proteins in a macromolecular complex that incorporates regulators of their function (20, 21). For ASIC1a, only a few interacting proteins have been identified. Earlier work indicated that ASIC1a interacts with another postsynaptic scaffolding protein, PICK1 (15, 22, 23). ASIC1a also has been reported to interact with annexin II light chain p11 through its cytosolic N terminus to increase cell surface expression (24) and with Ca2+/calmodulin-dependent protein kinase II to phosphorylate the channel (25). However, whether ASIC1a interacts with additional proteins and with the cytoskeleton remain unknown. Moreover, it is not known whether such interactions alter ASIC1a function.In analyzing the ASIC1a amino acid sequence, we identified cytosolic residues that might bind α-actinins. α-Actinins cluster membrane proteins and signaling molecules into macromolecular complexes and link membrane proteins to the actincytoskeleton (for review, Ref. 26). Four genes encode α-actinin-1, -2, -3, and -4 isoforms. α-Actinins contain an N-terminal head domain that binds F-actin, a C-terminal region containing two EF-hand motifs, and a central rod domain containing four spectrin-like motifs (2628). The C-terminal portion of the rod segment appears to be crucial for binding to membrane proteins. The α-actinins assemble into antiparallel homodimers through interactions in their rod domain. α-Actinins-1, -2, and -4 are enriched in dendritic spines, concentrating at the postsynaptic membrane (2935). In the postsynaptic membrane of excitatory synapses, α-actinin connects the NMDA receptor to the actin cytoskeleton, and this interaction is key for Ca2+-dependent inhibition of NMDA receptors (3638). α-Actinins can also regulate the membrane trafficking and function of several cation channels, including L-type Ca2+ channels, K+ channels, and TRP channels (3941).To better understand the function of ASIC1a channels in macromolecular complexes, we asked if ASIC1a associates with α-actinins. We were interested in the α-actinins because they and ASIC1a, both, are present in dendritic spines, ASIC1a contains a potential α-actinin binding sequence, and the related epithelial Na+ channel (ENaC) interacts with the cytoskeleton (42, 43). Therefore, we hypothesized that α-actinin interacts structurally and functionally with ASIC1a.  相似文献   

2.
3.
Protein kinase D (PKD) is a serine/threonine protein kinase rapidly activated by G protein-coupled receptor (GPCR) agonists via a protein kinase C (PKC)-dependent pathway. Recently, PKD has been implicated in the regulation of long term cellular activities, but little is known about the mechanism(s) of sustained PKD activation. Here, we show that cell treatment with the preferential PKC inhibitors GF 109203X or Gö 6983 blocked rapid (1–5-min) PKD activation induced by bombesin stimulation, but this inhibition was greatly diminished at later times of bombesin stimulation (e.g. 45 min). These results imply that GPCR-induced PKD activation is mediated by early PKC-dependent and late PKC-independent mechanisms. Western blot analysis with site-specific antibodies that detect the phosphorylated state of the activation loop residues Ser744 and Ser748 revealed striking PKC-independent phosphorylation of Ser748 as well as Ser744 phosphorylation that remained predominantly but not completely PKC-dependent at later times of bombesin or vasopressin stimulation (20–90 min). To determine the mechanisms involved, we examined activation loop phosphorylation in a set of PKD mutants, including kinase-deficient, constitutively activated, and PKD forms in which the activation loop residues were substituted for alanine. Our results show that PKC-dependent phosphorylation of the activation loop Ser744 and Ser748 is the primary mechanism involved in early phase PKD activation, whereas PKD autophosphorylation on Ser748 is a major mechanism contributing to the late phase of PKD activation occurring in cells stimulated by GPCR agonists. The present studies identify a novel mechanism induced by GPCR activation that leads to late, PKC-independent PKD activation.A rapid increase in the synthesis of lipid-derived second messengers with subsequent activation of protein phosphorylation cascades has emerged as a fundamental signal transduction mechanism triggered by multiple extracellular stimuli, including hormones, neurotransmitters, chemokines, and growth factors (1). Many of these agonists bind to G protein-coupled receptors (GPCRs),4 activate heterotrimeric G proteins and stimulate isoforms of the phospholipase C family, including β, γ, δ, and ε (reviewed in Refs. 1 and 2). Activated phospholipase Cs catalyze the hydrolysis of phosphatidylinositol 4,5-bisphosphate to produce the second messengers inositol 1,4,5-trisphosphate and diacylglycerol (DAG). Inositol 1,4,5-trisphosphate mobilizes Ca2+ from intracellular stores (3, 4) whereas DAG directly activates the classic (α, β, and γ) and novel (δ, ε, η, and θ) isoforms of PKC (57). Although it is increasingly recognized that each PKC isozyme has specific functions in vivo (58), the mechanisms by which PKC-mediated signals are propagated to critical downstream targets remain incompletely defined.PKD, also known initially as PKCμ (9, 10), and two recently identified serine protein kinases termed PKD2 (11) and PKCν/PKD3 (12, 13), which are similar in overall structure and primary amino acid sequence to PKD (14), constitute a new protein kinase family within the Ca2+/calmodulin-dependent protein kinase group (15) and separate from the previously identified PKCs (14). Salient features of PKD structure include an N-terminal regulatory region containing a tandem repeat of cysteine-rich zinc finger-like motifs (termed the cysteine-rich domain) that confers high affinity binding to phorbol esters and DAG (9, 16, 17), followed by a pleckstrin homology (PH) domain that negatively regulates catalytic activity (18, 19). The C-terminal region of the PKDs contains its catalytic domain, which is distantly related to Ca2+-regulated kinases.In unstimulated cells, PKD is in a state of low kinase catalytic activity maintained by the N-terminal domain, which represses the catalytic activity of the enzyme by autoinhibition. Consistent with this model, deletions or single amino acid substitutions in the PH domain result in constitutive kinase activity (1820). Physiological activation of PKD within cells occurs via a phosphorylation-dependent mechanism first identified in our laboratory (21). In response to cellular stimuli, PKD is converted from a low activity form into a persistently active form that is retained during isolation from cells, as shown by in vitro kinase assays performed in the absence of lipid co-activators (21, 22). PKD activation has been demonstrated in response to engagement of specific GPCRs either by regulatory peptides (2330) or lysophosphatidic acid (27, 31, 32); signaling through Gq, G12, Gi, and Rho (27, 3134); activation of receptor tyrosine kinases, such as the platelet-derived growth factor receptor (23, 35, 36); cross-linking of B-cell receptor and T-cell receptor in B and T lymphocytes, respectively (3740); and oxidative stress (4144).Throughout these studies, multiple lines of evidence indicated that PKC activity is necessary for rapid PKD activation within intact cells. For example, rapid PKD activation was selectively and potently blocked by cell treatment with preferential PKC inhibitors (e.g. GF 109203X or Gö 6983) that do not directly inhibit PKD catalytic activity (21, 22), implying that PKD activation in intact cells is mediated, directly or indirectly, through PKCs. In line with this conclusion, cotransfection of PKD with active mutant forms of “novel” PKCs (PKCs δ, ε, η, and θ) resulted in robust PKD activation in the absence of cell stimulation (21, 4446). Many reports demonstrated the operation of a rapid PKC/PKD signaling cascade in response to multiple GPCR agonists in a broad range of cell types, including normal and cancer cells (reviewed in Ref. 14). Our previous studies identified Ser744 and Ser748 in the PKD activation loop (also referred as the activation segment or T-loop) as phosphorylation sites critical for PKC-mediated PKD activation (reviewed in Ref. 14). Collectively, these findings demonstrated the existence of rapidly activated PKC-PKD protein kinase cascade(s) and raised the possibility that some PKC-dependent biological responses involve PKD acting as a downstream effector.PKD has been reported recently to mediate several important cellular activities and processes, including signal transduction (30, 4749), chromatin modification (50), Golgi organization and function (51, 52), c-Jun function (47, 53, 54), NFκB-mediated gene expression (43, 55, 56), and cell survival, migration, and differentiation and DNA synthesis and proliferation (reviewed in Ref. 14). Thus, mounting evidence indicates that PKD has a remarkable diversity of both its signal generation and distribution and its potential for complex regulatory interactions with multiple downstream pathways, leading to multiple responses, including long term cellular events. Despite increasing recognition of its importance, very little is known about the mechanism(s) of sustained PKD activation as opposed to the well documented rapid, PKC-dependent PKD activation.The results presented here demonstrate that prolonged GPCR-induced PKD activation is mediated by sequential PKC-dependent and PKC-independent phases of regulation. We report here, for the first time, that PKD autophosphorylation on Ser748 is a major mechanism contributing to the late phase of PKD activation occurring in cells stimulated by GPCR agonists. The present studies expand previous models of PKD regulation by identifying a novel mechanism induced by GPCR activation that leads to late, PKC-independent PKD activation.  相似文献   

4.
5.
We have investigated the possible biochemical basis for enhancements in NO production in endothelial cells that have been correlated with agonist- or shear stress-evoked phosphorylation at Ser-1179. We have found that a phosphomimetic substitution at Ser-1179 doubles maximal synthase activity, partially disinhibits cytochrome c reductase activity, and lowers the EC50(Ca2+) values for calmodulin binding and enzyme activation from the control values of 182 ± 2 and 422 ± 22 nm to 116 ± 2 and 300 ± 10 nm. These are similar to the effects of a phosphomimetic substitution at Ser-617 (Tran, Q. K., Leonard, J., Black, D. J., and Persechini, A. (2008) Biochemistry 47, 7557–7566). Although combining substitutions at Ser-617 and Ser-1179 has no additional effect on maximal synthase activity, cooperativity between the two substitutions completely disinhibits reductase activity and further reduces the EC50(Ca2+) values for calmodulin binding and enzyme activation to 77 ± 2 and 130 ± 5 nm. We have confirmed that specific Akt-catalyzed phosphorylation of Ser-617 and Ser-1179 and phosphomimetic substitutions at these positions have similar functional effects. Changes in the biochemical properties of eNOS produced by combined phosphorylation at Ser-617 and Ser-1179 are predicted to substantially increase synthase activity in cells at a typical basal free Ca2+ concentration of 50–100 nm.The nitric-oxide synthases catalyze formation of NO and l-citrulline from l-arginine and O2, with NADPH as the electron donor (1). The role of NO generated by endothelial nitricoxide synthase (eNOS)2 in the regulation of smooth muscle tone is well established and was the first of several physiological roles for this small molecule that have so far been identified (2). The nitric-oxide synthases are homodimers of 130–160-kDa subunits. Each subunit contains a reductase and oxygenase domain (1). A significant difference between the reductase domains in eNOS and nNOS and the homologous P450 reductases is the presence of inserts in these synthase isoforms that appear to maintain them in their inactive states (3, 4). A calmodulin (CaM)-binding domain is located in the linker that connects the reductase and oxygenase domains, and the endothelial and neuronal synthases both require Ca2+ and exogenous CaM for activity (5, 6). When CaM is bound, it somehow counteracts the effects of the autoinhibitory insert(s) in the reductase. The high resolution structure for the complex between (Ca2+)4-CaM and the isolated CaM-binding domain from eNOS indicates that the C-ter and N-ter lobes of CaM, which each contain a pair of Ca2+-binding sites, enfold the domain, as has been observed in several other such CaM-peptide complexes (7). Consistent with this structure, investigations of CaM-dependent activation of the neuronal synthase suggest that both CaM lobes must participate (8, 9).Bovine eNOS can be phosphorylated in endothelial cells at Ser-116, Thr-497, Ser-617, Ser-635, and Ser-1179 (1012). There are equivalent phosphorylation sites in the human enzyme (1012). Phosphorylation of the bovine enzyme at Thr-497, which is located in the CaM-binding domain, blocks CaM binding and enzyme activation (7, 11, 13, 14). Ser-116 can be basally phosphorylated in cells (10, 11, 13, 15), and dephosphorylation of this site has been correlated with increased NO production (13, 15). However, it has also been reported that a phosphomimetic substitution at this position has no effect on enzyme activity measured in vitro (13). Ser-1179 is phosphorylated in response to a variety of stimuli, and this has been reliably correlated with enhanced NO production in cells (10, 11). Indeed, NO production is elevated in transgenic endothelium expressing an eNOS mutant containing an S1179D substitution, but not in tissue expressing an S1179A mutant (16). Shear stress or insulin treatment is correlated with Akt-catalyzed phosphorylation of Ser-1179 in endothelial cells, and this is correlated with increased NO production in the absence of extracellular Ca2+ (1719). Akt-catalyzed phosphorylation or an S1179D substitution has also been correlated with increased synthase activity in cell extracts at low intracellular free [Ca2+] (17). Increased NO production has also been observed in cells expressing an eNOS mutant containing an S617D substitution, and physiological stimuli such as shear-stress, bradykinin, VEGF, and ATP appear to stimulate Akt-catalyzed phosphorylation of Ser-617 and Ser-1179 (12, 13, 20). Although S617D eNOS has been reported to have the same maximum activity in vitro as the wild type enzyme (20), in our hands an S617D substitution increases the maximal CaM-dependent synthase activity of purified mutant enzyme ∼2-fold, partially disinhibits reductase activity, and reduces the EC50(Ca2+) values for CaM binding and enzyme activation (21).In this report, we describe the effects of a phosphomimetic Asp substitution at Ser-1179 in eNOS on the Ca2+ dependence of CaM binding and CaM-dependent activation of reductase and synthase activities. We also describe the effects on these properties of combining this substitution with one at Ser-617. Finally, we demonstrate that Akt-catalyzed phosphorylation and Asp substitutions at Ser-617 and Ser-1179 have similar functional effects. Our results suggest that phosphorylation of eNOS at Ser-617 and Ser-1179 can substantially increase synthase activity in cells at a typical basal free Ca2+ concentration of 50–100 nm, while single phosphorylations at these sites produce smaller activity increases, and can do so only at higher free Ca2+ concentrations.  相似文献   

6.
A key set of reactions for the initiation of new DNA strands during herpes simplex virus-1 replication consists of the primase-catalyzed synthesis of short RNA primers followed by polymerase-catalyzed DNA synthesis (i.e. primase-coupled polymerase activity). Herpes primase (UL5-UL52-UL8) synthesizes products from 2 to ∼13 nucleotides long. However, the herpes polymerase (UL30 or UL30-UL42) only elongates those at least 8 nucleotides long. Surprisingly, coupled activity was remarkably inefficient, even considering only those primers at least 8 nucleotides long, and herpes polymerase typically elongated <2% of the primase-synthesized primers. Of those primers elongated, only 4–26% of the primers were passed directly from the primase to the polymerase (UL30-UL42) without dissociating into solution. Comparing RNA primer-templates and DNA primer-templates of identical sequence showed that herpes polymerase greatly preferred to elongate the DNA primer by 650–26,000-fold, thus accounting for the extremely low efficiency with which herpes polymerase elongated primase-synthesized primers. Curiously, one of the DNA polymerases of the host cell, polymerase α (p70-p180 or p49-p58-p70-p180 complex), extended herpes primase-synthesized RNA primers much more efficiently than the viral polymerase, raising the possibility that the viral polymerase may not be the only one involved in herpes DNA replication.Herpes simplex virus 1 (HSV-1)2 encodes seven proteins essential for replicating its double-stranded DNA genome; five of these encode the heterotrimeric helicase-primase (UL5-UL52-UL8 gene products) and the heterodimeric polymerase (UL30-UL42 gene products) (1, 2). The helicase-primase unwinds the DNA at the replication fork and generates single-stranded DNA for both leading and lagging strand synthesis. Primase synthesizes short RNA primers on the lagging strand that the polymerase presumably elongates using dNTPs (i.e. primase-coupled polymerase activity). These two protein complexes are thought to replicate the viral genome on both the leading and lagging strands (1, 2).Previous studies have focused on the helicase-primase and polymerase separately. The helicase-primase contains three subunits, UL5, UL52, and UL8 in a 1:1:1 ratio (35). The UL5 subunit has helicase-like motifs and the UL52 subunit has primase-like motifs, yet the minimal active complex that demonstrates either helicase or primase activities contains both UL5 and UL52 (6, 7). Although the UL8 subunit has no known catalytic activity, several functions have been proposed, including enhancing helicase and primase activities, enhancing primer synthesis on ICP8 (the HSV-1 single-stranded binding protein)-coated DNA strands, and facilitating formation of the replisome (812). Although primase will synthesize short (23 nucleotides long) primers on a variety of template sequences, synthesis of longer primers up to 13 nucleotides long requires the template sequence, 3′-deoxyguanidine-pyrimidine-pyrimidine-5′ (13). Primase initiates synthesis at the first pyrimidine via the polymerization of two purine NTPs (13). Even after initiation at this sequence, however, the vast majority of products are only 2–3 nucleotides long (13, 14).The herpes polymerase consists of the UL30 subunit, which has polymerase and 3′ → 5′ exonuclease activities (1, 2), and the UL42 subunit, which serves as a processivity factor (1517). Unlike most processivity factors that encircle the DNA, the UL42 protein binds double-stranded DNA and thus directly tethers the polymerase to the DNA (18). Using pre-existing DNA primer-templates as the substrate, the heterodimeric polymerase (UL30-UL42) incorporates dNTPs at a rate of 150 s–1, a rate much faster than primer synthesis (for primers >7 nucleotides long, 0.0002–0.01 s–1) (19, 20).We examined primase-coupled polymerase activity by the herpes primase and polymerase complexes. Although herpes primase synthesizes RNA primers 2–13 nucleotides long, the polymerase only effectively elongates those at least 8 nucleotides long. Surprisingly, the polymerase elongated only a small fraction of the primase-synthesized primers (<1–2%), likely because of the polymerase elongating RNA primer-templates much less efficiently than DNA primer-templates. In contrast, human DNA polymerase α (pol α) elongated the herpes primase-synthesized primers very efficiently. The biological significance of these data is discussed.  相似文献   

7.
The RAD51 protein is a central player in homologous recombinational repair. The RAD51B protein is one of five RAD51 paralogs that function in the homologous recombinational repair pathway in higher eukaryotes. In the present study, we found that the human EVL (Ena/Vasp-like) protein, which is suggested to be involved in actin-remodeling processes, unexpectedly binds to the RAD51 and RAD51B proteins and stimulates the RAD51-mediated homologous pairing and strand exchange. The EVL knockdown cells impaired RAD51 assembly onto damaged DNA after ionizing radiation or mitomycin C treatment. The EVL protein alone promotes single-stranded DNA annealing, and the recombination activities of the EVL protein are further enhanced by the RAD51B protein. The expression of the EVL protein is not ubiquitous, but it is significantly expressed in breast cancer-derived MCF7 cells. These results suggest that the EVL protein is a novel recombination factor that may be required for repairing specific DNA lesions, and that may cause tumor malignancy by its inappropriate expression.Chromosomal DNA double strand breaks (DSBs)2 are potential inducers of chromosomal aberrations and tumorigenesis, and they are accurately repaired by the homologous recombinational repair (HRR) pathway, without base substitutions, deletions, and insertions (13). In the HRR pathway (4, 5), single-stranded DNA (ssDNA) tails are produced at the DSB sites. The RAD51 protein, a eukaryotic homologue of the bacterial RecA protein, binds to the ssDNA tail and forms a helical nucleoprotein filament. The RAD51-ssDNA filament then binds to the intact double-stranded DNA (dsDNA) to form a three-component complex, containing ssDNA, dsDNA, and the RAD51 protein. In this three-component complex, the RAD51 protein promotes recombination reactions, such as homologous pairing and strand exchange (69).The RAD51 protein requires auxiliary proteins to promote the homologous pairing and strand exchange reactions efficiently in cells (1012). In humans, the RAD52, RAD54, and RAD54B proteins directly interact with the RAD51 protein (1317) and stimulate the RAD51-mediated homologous pairing and/or strand exchange reactions in vitro (1821). The human RAD51AP1 protein, which directly binds to the RAD51 protein (22), was also found to stimulate RAD51-mediated homologous pairing in vitro (23, 24). The BRCA2 protein contains ssDNA-binding, dsDNA-binding, and RAD51-binding motifs (2533), and the Ustilago maydis BRCA2 ortholog, Brh2, reportedly stimulated RAD51-mediated strand exchange (34, 35). Most of these RAD51-interacting factors are known to be required for efficient RAD51 assembly onto DSB sites in cells treated with ionizing radiation (1012).The RAD51B (RAD51L1, Rec2) protein is a member of the RAD51 paralogs, which share about 20–30% amino acid sequence similarity with the RAD51 protein (3638). RAD51B-deficient cells are hypersensitive to DSB-inducing agents, such as cisplatin, mitomycin C (MMC), and γ-rays, indicating that the RAD51B protein is involved in the HRR pathway (3944). Genetic experiments revealed that RAD51B-deficient cells exhibited impaired RAD51 assembly onto DSB sites (39, 44), suggesting that the RAD51B protein functions in the early stage of the HRR pathway. Biochemical experiments also suggested that the RAD51B protein participates in the early to late stages of the HRR pathway (4547).In the present study, we found that the human EVL (Ena/Vasp-like) protein binds to the RAD51 and RAD51B proteins in a HeLa cell extract. The EVL protein is known to be involved in cytoplasmic actin remodeling (48) and is also overexpressed in breast cancer (49). Like the RAD51B knockdown cells, the EVL knockdown cells partially impaired RAD51 foci formation after DSB induction, suggesting that the EVL protein enhances RAD51 assembly onto DSB sites. The purified EVL protein preferentially bound to ssDNA and stimulated RAD51-mediated homologous pairing and strand exchange. The EVL protein also promoted the annealing of complementary strands. These recombination reactions that were stimulated or promoted by the EVL protein were further enhanced by the RAD51B protein. These results strongly suggested that the EVL protein is a novel factor that activates RAD51-mediated recombination reactions, probably with the RAD51B protein. We anticipate that, in addition to its involvement in cytoplasmic actin dynamics, the EVL protein may be required in homologous recombination for repairing specific DNA lesions, and it may cause tumor malignancy by inappropriate recombination enhanced by EVL overexpression in certain types of tumor cells.  相似文献   

8.
9.
The Ca2+ release-activated Ca2+ channel is a principal regulator of intracellular Ca2+ rise, which conducts various biological functions, including immune responses. This channel, involved in store-operated Ca2+ influx, is believed to be composed of at least two major components. Orai1 has a putative channel pore and locates in the plasma membrane, and STIM1 is a sensor for luminal Ca2+ store depletion in the endoplasmic reticulum membrane. Here we have purified the FLAG-fused Orai1 protein, determined its tetrameric stoichiometry, and reconstructed its three-dimensional structure at 21-Å resolution from 3681 automatically selected particle images, taken with an electron microscope. This first structural depiction of a member of the Orai family shows an elongated teardrop-shape 150Å in height and 95Å in width. Antibody decoration and volume estimation from the amino acid sequence indicate that the widest transmembrane domain is located between the round extracellular domain and the tapered cytoplasmic domain. The cytoplasmic length of 100Å is sufficient for direct association with STIM1. Orifices close to the extracellular and intracellular membrane surfaces of Orai1 seem to connect outside the molecule to large internal cavities.Ca2+ is an intracellular second messenger that plays important roles in various physiological functions such as immune response, muscle contraction, neurotransmitter release, and cell proliferation. Intracellular Ca2+ is mainly stored in the endoplasmic reticulum (ER).2 This ER system is distributed through the cytoplasm from around the nucleus to the cell periphery close to the plasma membrane. In non-excitable cells, the ER releases Ca2+ through the inositol 1,4,5-trisphosphate (IP3) receptor channel in response to various signals, and the Ca2+ store is depleted. Depletion of Ca2+ then induces Ca2+ influx from outside the cell to help in refilling the Ca2+ stores and to continue Ca2+ rise for several minutes in the cytoplasm (1, 2). This Ca2+ influx was first proposed by Putney (3) and was named store-operated Ca2+ influx. In the immune system, store-operated Ca2+ influx is mainly mediated by the Ca2+ release-activated Ca2+ (CRAC) current, which is a highly Ca2+-selective inwardly rectified current with low conductance (4, 5). Pathologically, the loss of CRAC current in T cells causes severe combined immunodeficiency (6) where many Ca2+ signal-dependent gene expressions, including cytokines, are interrupted (7). Therefore, CRAC current is necessary for T cell functions.Recently, Orai1 (also called CRACM1) and STIM1 have been physiologically characterized as essential components of the CRAC channel (812). They are separately located in the plasma membrane and in the ER membrane; co-expression of these proteins presents heterologous CRAC-like currents in various types of cells (10, 1315). Both of them are shown to be expressed ubiquitously in various tissues (1618). STIM1 senses Ca2+ depletion in the ER through its EF hand motif (19) and transmits a signal to Orai1 in the plasma membrane. Although Orai1 is proposed as a regulatory component for some transient receptor potential canonical channels (20, 21), it is believed from the mutation analyses to be the pore-forming subunit of the CRAC channel (8, 2224). In the steady state, both Orai1 and STIM1 molecules are dispersed in each membrane. When store depletion occurs, STIM1 proteins gather into clusters to form puncta in the ER membrane near the plasma membrane (11, 19). These clusters then trigger the clustering of Orai1 in the plasma membrane sites opposite the puncta (25, 26), and CRAC channels are activated (27).Orai1 has two homologous genes, Orai2 and Orai3 (8). They form the Orai family and have in common the four transmembrane (TM) segments with relatively large N and C termini. These termini are demonstrated to be in the cytoplasm, because both N- and C-terminally introduced tags are immunologically detected only in the membrane-permeabilized cells (8, 9). The subunit stoichiometry of Orai1 is as yet controversial: it is believed to be an oligomer, presumably a dimer or tetramer even in the steady state (16, 2830).Despite the accumulation of biochemical and electrophysiological data, structural information about Orai1 is limited due to difficulties in purification and crystallization. In this study, we have purified Orai1 in its tetrameric form and have reconstructed the three-dimensional structure from negatively stained electron microscopic (EM) images.  相似文献   

10.
11.
12.
Recently, this laboratory identified a proton-coupled folate transporter (PCFT), with optimal activity at low pH. PCFT is critical to intestinal folate absorption and transport into the central nervous system because there are loss-of-function mutations in this gene in the autosomal recessive disorder, hereditary folate malabsorption. The current study addresses the role PCFT might play in another transport pathway, folate receptor (FR)-mediated endocytosis. FRα cDNA was transfected into novel PCFT+ and PCFT HeLa sublines. FRα was shown to bind and trap folates in vesicles but with minimal export into the cytosol in PCFT cells. Cotransfection of FRα and PCFT resulted in enhanced folate transport into cytosol as compared with transfection of FRα alone. Probenecid did not inhibit folate binding to FR, but inhibited PCFT-mediated transport at endosomal pH, and blocked FRα-mediated transport into the cytosol. FRα and PCFT co-localized to the endosomal compartment. These observations (i) indicate that PCFT plays a role in FRα-mediated endocytosis by serving as a route of export of folates from acidified endosomes and (ii) provide a functional role for PCFT in tissues in which it is expressed, such as the choroid plexus, where the extracellular milieu is at neutral pH.Loss of function mutations of the proton-coupled folate transporter (PCFT),2 which functions optimally at low pH, are the molecular basis for the autosomal recessive disorder, hereditary folate malabsorption (HFM) (14). Infants present with this disorder several months after birth with marked folate deficiency anemia, hypogammaglobulinemia with immune deficiency and infections, neurological deficits, and often seizures (5). PCFT is highly expressed at the apical brush-border membrane of the duodenum and proximal jejunum (69) where the pH at the microclimate of the surface of this epithelium is low (pH 5.8–6.0), and folates are absorbed (1, 7, 10, 11). Hence, the failure to absorb folates in the absence of this transporter in HFM is expected. However, PCFT expression, and its associated folate transport activity at low pH, is observed in many tissues where the transport interface is presumed to be at pH 7.4 (12). Of particular interest is the mechanism by which PCFT mediates transport of folates into the central nervous system (CNS) where this transporter is expressed in brain and choroid plexus (1, 7, 13). Transport into the CNS is impaired in patients with HFM who have very low cerebrospinal fluid (CSF) folate levels and marked reversal of the blood:CSF folate gradient which is normally 2–3:1 (5).Folates are also transported into cells by a receptor-mediated process. Folate receptor-α (FRα) is anchored to cell membranes via a glycosylphosphatidylinositol domain. Uptake begins with folate binding to receptor at the cell surface followed by invagination of the membrane and the formation of endosomes that traffic along microtubules to a perinuclear compartment before returning to the plasma membrane (1416). During transit in the cytoplasm, endosomes acidify to a pH of ∼6.0–6.5 (17), folate is released from the receptor and exported from the intact endosome into the cytoplasm. This putative exporter was shown to require a trans-endosomal pH gradient (1820).The current report addresses the hypothesis that PCFT is an endosomal folate exporter and thereby plays a role in FRα-mediated endocytosis (1, 2, 21, 22), that the ubiquitous expression of PCFT in mammalian tissues may be related to this function, and that loss of this function may be a basis for the low CSF folate levels in HFM. The experimental approach utilized a series of HeLa sublines, developed in this laboratory, in which constitutive expression of FRα is negligible. HeLa R5 cells lack reduced folate carrier (RFC) function due to a genomic deletion of this gene (23). A derivative of R5 cells, HeLa R1-11 cells lack, in addition, PCFT expression, while an R1-11 revertant re-expresses PCFT (24). The impact of PCFT on FRα-mediated endocytosis, achieved by transfection of the receptor into these cell lines, was assessed under conditions in which there was negligible PCFT-mediated transport directly across the plasma membrane into cells.  相似文献   

13.
Oxygen and glucose deprivation (OGD) induces delayed cell death in hippocampal CA1 neurons via Ca2+/Zn2+-permeable, GluR2-lacking AMPA receptors (AMPARs). Following OGD, synaptic AMPAR currents in hippocampal neurons show marked inward rectification and increased sensitivity to channel blockers selective for GluR2-lacking AMPARs. This occurs via two mechanisms: a delayed down-regulation of GluR2 mRNA expression and a rapid internalization of GluR2-containing AMPARs during the OGD insult, which are replaced by GluR2-lacking receptors. The mechanisms that underlie this rapid change in subunit composition are unknown. Here, we demonstrate that this trafficking event shares features in common with events that mediate long term depression and long term potentiation and is initiated by the activation of N-methyl-d-aspartic acid receptors. Using biochemical and electrophysiological approaches, we show that peptides that interfere with PICK1 PDZ domain interactions block the OGD-induced switch in subunit composition, implicating PICK1 in restricting GluR2 from synapses during OGD. Furthermore, we show that GluR2-lacking AMPARs that arise at synapses during OGD as a result of PICK1 PDZ interactions are involved in OGD-induced delayed cell death. This work demonstrates that PICK1 plays a crucial role in the response to OGD that results in altered synaptic transmission and neuronal death and has implications for our understanding of the molecular mechanisms that underlie cell death during stroke.Oxygen and glucose deprivation (OGD)3 associated with transient global ischemia induces delayed cell death, particularly in hippocampal CA1 pyramidal cells (13), a phenomenon that involves Ca2+/Zn2+-permeable, GluR2-lacking AMPARs (4). AMPARs are heteromeric complexes of subunits GluR1–4 (5), and most AMPARs in the hippocampus contain GluR2, which renders them calcium-impermeable and results in a marked inward rectification in their current-voltage relationship (68). Ischemia induces a delayed down-regulation of GluR2 mRNA and protein expression (4, 911), resulting in enhanced AMPAR-mediated Ca2+ and Zn2+ influx into CA1 neurons (10, 12). In these neurons, AMPAR-mediated postsynaptic currents (EPSCs) show marked inward rectification 1–2 days following ischemia and increased sensitivity to 1-naphthyl acetyl spermine (NASPM), a channel blocker selective for GluR2-lacking AMPARs (1316). Blockade of these channels at 9–40 h following ischemia is neuroprotective, indicating a crucial role for Ca2+-permeable AMPARs in ischemic cell death (16).In addition to delayed changes in AMPAR subunit composition as a result of altered mRNA expression, it was recently reported that Ca2+-permable, GluR2-lacking AMPARs are targeted to synaptic sites via membrane trafficking at much earlier times during OGD (17). This subunit rearrangement involves endocytosis of AMPARs containing GluR2 complexed with GluR1/3, followed by exocytosis of GluR2-lacking receptors containing GluR1/3 (17). However, the molecular mechanisms behind this trafficking event are unknown, and furthermore, it is not known whether these trafficking-mediated changes in AMPAR subunit composition contribute to delayed cell death.AMPAR trafficking is a well studied phenomenon because of its crucial involvement in long term depression (LTD) and long term potentiation (LTP), activity-dependent forms of synaptic plasticity thought to underlie learning and memory. AMPAR endocytosis, exocytosis, and more recently subunit-switching events (brought about by trafficking that involves endo/exocytosis) are central to the necessary changes in synaptic receptor complement (7, 1820). It is possible that similar mechanisms regulate AMPAR trafficking during OGD.PICK1 is a PDZ and BAR (Bin-amphiphysin-Rus) domain-containing protein that binds, via the PDZ domain, to a number of membrane proteins including AMPAR subunits GluR2/3. This interaction is required for AMPAR internalization from the synaptic plasma membrane in response to Ca2+ influx via NMDAR activation in hippocampal neurons (2123). This process is the major mechanism that underlies the reduction in synaptic strength in LTD. Furthermore, PICK1-mediated trafficking has recently emerged as a mechanism that regulates the GluR2 content of synaptic receptors, which in turn determines their Ca2+ permeability (7, 20). This is likely to be of profound importance in both plasticity and pathological mechanisms. Importantly, PICK1 overexpression has been shown to induce a shift in synaptic AMPAR subunit composition in hippocampal CA1 neurons, resulting in inwardly rectifying AMPAR EPSCs via reduced surface GluR2 and no change in GluR1 (24). This suggests that PICK1 may mediate the rapid switch in subunit composition occurring during OGD (17). Here, we demonstrate that the OGD-induced switch in AMPAR subunit composition is dependent on PICK1 PDZ interactions, and importantly, that this early trafficking event that occurs during OGD contributes to the signaling that results in delayed neuronal death.  相似文献   

14.
In Alzheimer disease (AD), frontotemporal dementia and parkinsonism linked to chromosome 17 (FTDP-17) and other tauopathies, tau accumulates and forms paired helical filaments (PHFs) in the brain. Tau isolated from PHFs is phosphorylated at a number of sites, migrates as ∼60-, 64-, and 68-kDa bands on SDS-gel, and does not promote microtubule assembly. Upon dephosphorylation, the PHF-tau migrates as ∼50–60-kDa bands on SDS-gels in a manner similar to tau that is isolated from normal brain and promotes microtubule assembly. The site(s) that inhibits microtubule assembly-promoting activity when phosphorylated in the diseased brain is not known. In this study, when tau was phosphorylated by Cdk5 in vitro, its mobility shifted from ∼60-kDa bands to ∼64- and 68-kDa bands in a time-dependent manner. This mobility shift correlated with phosphorylation at Ser202, and Ser202 phosphorylation inhibited tau microtubule-assembly promoting activity. When several tau point mutants were analyzed, G272V, P301L, V337M, and R406W mutations associated with FTDP-17, but not nonspecific mutations S214A and S262A, promoted Ser202 phosphorylation and mobility shift to a ∼68-kDa band. Furthermore, Ser202 phosphorylation inhibited the microtubule assembly-promoting activity of FTDP-17 mutants more than of WT. Our data indicate that FTDP-17 missense mutations, by promoting phosphorylation at Ser202, inhibit the microtubule assembly-promoting activity of tau in vitro, suggesting that Ser202 phosphorylation plays a major role in the development of NFT pathology in AD and related tauopathies.Neurofibrillary tangles (NFTs)4 and senile plaques are the two characteristic neuropathological lesions found in the brains of patients suffering from Alzheimer disease (AD). The major fibrous component of NFTs are paired helical filaments (PHFs) (for reviews see Refs. 13). Initially, PHFs were found to be composed of a protein component referred to as “A68” (4). Biochemical analysis reveled that A68 is identical to the microtubule-associated protein, tau (4, 5). Some characteristic features of tau isolated from PHFs (PHF-tau) are that it is abnormally hyperphosphorylated (phosphorylated on more sites than the normal brain tau), does not bind to microtubules, and does not promote microtubule assembly in vitro. Upon dephosphorylation, PHF-tau regains its ability to bind to and promote microtubule assembly (6, 7). Tau hyperphosphorylation is suggested to cause microtubule instability and PHF formation, leading to NFT pathology in the brain (13).PHF-tau is phosphorylated on at least 21 proline-directed and non-proline-directed sites (8, 9). The individual contribution of these sites in converting tau to PHFs is not entirely clear. However, some sites are only partially phosphorylated in PHFs (8), whereas phosphorylation on specific sites inhibits the microtubule assembly-promoting activity of tau (6, 10). These observations suggest that phosphorylation on a few sites may be responsible and sufficient for causing tau dysfunction in AD.Tau purified from the human brain migrates as ∼50–60-kDa bands on SDS-gel due to the presence of six isoforms that are phosphorylated to different extents (2). PHF-tau isolated from AD brain, on the other hand, displays ∼60-, 64-, and 68 kDa-bands on an SDS-gel (4, 5, 11). Studies have shown that ∼64- and 68-kDa tau bands (the authors have described the ∼68-kDa tau band as an ∼69-kDa band in these studies) are present only in brain areas affected by NFT degeneration (12, 13). Their amount is correlated with the NFT densities at the affected brain regions. Moreover, the increase in the amount of ∼64- and 68-kDa band tau in the brain correlated with a decline in the intellectual status of the patient. The ∼64- and 68-kDa tau bands were suggested to be the pathological marker of AD (12, 13). Biochemical analyses determined that ∼64- and 68-kDa bands are hyperphosphorylated tau, which upon dephosphorylation, migrated as normal tau on SDS-gel (4, 5, 11). Tau sites involved in the tau mobility shift to ∼64- and 68-kDa bands were suggested to have a role in AD pathology (12, 13). It is not known whether phosphorylation at all 21 PHF-sites is required for the tau mobility shift in AD. However, in vitro the tau mobility shift on SDS-gel is sensitive to phosphorylation only on some sites (6, 14). It is therefore possible that in the AD brain, phosphorylation on some sites also causes a tau mobility shift. Identification of such sites will significantly enhance our knowledge of how NFT pathology develops in the brain.PHFs are also the major component of NFTs found in the brains of patients suffering from a group of neurodegenerative disorders collectively called tauopathies (2, 11). These disorders include frontotemporal dementia and Parkinsonism linked to chromosome 17 (FTDP-17), corticobasal degeneration, progressive supranuclear palsy, and Pick disease. Each PHF-tau isolated from autopsied brains of patients suffering from various tauopathies is hyperphosphorylated, displays ∼60-, 64-, and 68-kDa bands on SDS-gel, and is incapable of binding to microtubules. Upon dephosphorylation, the above referenced PHF-tau migrates as a normal tau on SDS-gel, binds to microtubules, and promotes microtubule assembly (2, 11). These observations suggest that the mechanisms of NFT pathology in various tauopathies may be similar and the phosphorylation-dependent mobility shift of tau on SDS-gel may be an indicator of the disease. The tau gene is mutated in familial FTDP-17, and these mutations accelerate NFT pathology in the brain (1518). Understanding how FTDP-17 mutations promote tau phosphorylation can provide a better understanding of how NFT pathology develops in AD and various tauopathies. However, when expressed in CHO cells, G272V, R406W, V337M, and P301L tau mutations reduce tau phosphorylation (19, 20). In COS cells, although G272V, P301L, and V337M mutations do not show any significant affect, the R406W mutation caused a reduction in tau phosphorylation (21, 22). When expressed in SH-SY5Y cells subsequently differentiated into neurons, the R406W, P301L, and V337M mutations reduce tau phosphorylation (23). In contrast, in hippocampal neurons, R406W increases tau phosphorylation (24). When phosphorylated by recombinant GSK3β in vitro, the P301L and V337M mutations do not have any effect, and the R406W mutation inhibits phosphorylation (25). However, when incubated with rat brain extract, all of the G272V, P301L, V337M, and R406W mutations stimulate tau phosphorylation (26). The mechanism by which FTDP-17 mutations promote tau phosphorylation leading to development of NFT pathology has remained unclear.Cyclin-dependent protein kinase 5 (Cdk5) is one of the major kinases that phosphorylates tau in the brain (27, 28). In this study, to determine how FTDP-17 missense mutations affect tau phosphorylation, we phosphorylated four FTDP-17 tau mutants (G272V, P301L, V337M, and R406W) by Cdk5. We have found that phosphorylation of tau by Cdk5 causes a tau mobility shift to ∼64- and 68 kDa-bands. Although the mobility shift to a ∼64-kDa band is achieved by phosphorylation at Ser396/404 or Ser202, the mobility shift to a 68-kDa band occurs only in response to phosphorylation at Ser202. We show that in vitro, FTDP-17 missense mutations, by promoting phosphorylation at Ser202, enhance the mobility shift to ∼64- and 68-kDa bands and inhibit the microtubule assembly-promoting activity of tau. Our data suggest that Ser202 phosphorylation is the major event leading to NFT pathology in AD and related tauopathies.  相似文献   

15.
16.
Glycodelin is a human glycoprotein with four reported glycoforms, namely glycodelin-A (GdA), glycodelin-F (GdF), glycodelin-C (GdC), and glycodelin-S (GdS). These glycoforms have the same protein core and appear to differ in their N-glycosylation. The glycosylation of GdA is completely different from that of GdS. GdA inhibits proliferation and induces cell death of T cells. However, the glycosylation and immunomodulating activities of GdF and GdC are not known. This study aimed to use ultra-high sensitivity mass spectrometry to compare the glycomes of GdA, GdC, and GdF and to study the relationship between the immunological activity and glycosylation pattern among glycodelin glycoforms. Using MALDI-TOF strategies, the glycoforms were shown to contain an enormous diversity of bi-, tri-, and tetra-antennary complex-type glycans carrying Galβ1–4GlcNAc (lacNAc) and/or GalNAcβ1–4GlcNAc (lacdiNAc) antennae backbones with varying levels of fucose and sialic acid substitution. Interestingly, they all carried a family of Sda (NeuAcα2–3(GalNAcβ1–4)Gal)-containing glycans, which were not identified in the earlier study because of less sensitive methodologies used. Among the three glycodelins, GdA is the most heavily sialylated. Virtually all the sialic acid on GdC is located on the Sda antennae. With the exception of the Sda epitope, the GdC N-glycome appears to be the asialylated counterpart of the GdA/GdF glycomes. Sialidase activity, which may be responsible for transforming GdA/GdF to GdC, was detected in cumulus cells. Both GdA and GdF inhibited the proliferation, induced cell death, and suppressed interleukin-2 secretion of Jurkat cells and peripheral blood mononuclear cells. In contrast, no immunosuppressive effect was observed for GdS and GdC.Glycodelin is a member of the lipocalin family. It consists of 180 amino acid residues (1) with two sites of N-linked glycosylation. There are four reported glycodelin isoforms, namely glycodelin-A (amniotic fluid isoform, GdA),4 glycodelin-F (follicular fluid, GdF), glycodelin-C (cumulus matrix, GdC) and glycodelin-S (seminal plasma, GdS) (25). Among the four glycodelin isoforms, only the N-glycan structures of GdA and GdS have been previously determined. This was achieved using fast atom bombardment mass spectrometry (6, 7). The glycan structures of GdA and GdS are completely different. In GdA, the Asn-28 site carries high mannose, hybrid, and complex-type structures, whereas the second Asn-63 site is exclusively occupied by complex-type glycans (6). The major non-reducing epitopes characterized in the complex-type glycans are Galβ1–4GlcNAc (lacNAc), GalNAcβ1–4GlcNAc (lacdiNAc), NeuAcα2–6Galβ1–4GlcNAc (sialylated lacNAc), NeuAcα2–6GalNAcβ1–4GlcNAc (sialylated lacdiNAc), Galβ1–4(Fucα1–3)GlcNAc (Lewis-x), and GalNAcβ1–4(Fucα1–3)GlcNAc (lacdiNAc analog of the blood group substance Lewis-x) (6). Many of these oligosaccharides are rare in other human glycoproteins. GdS glycans are unusually fucose-rich, and the major complex type glycan structures are bi-antennary glycans with Lewis-x and Lewis-y antennae. Glycosylation of GdS is highly site-specific. Asn-28 contains only high mannose structures, whereas Asn-63 contains only complex type glycans. More than 80% of the complex glycans have 3–5 fucose residues/glycan, and none of the glycans is sialylated, which is unusual for a secreted human glycoprotein (7). The glycan structures of GdF and GdC are not known, although they differ in lectin-binding properties and isoelectric point from the other two glycodelin isoforms (5).Glycans are involved in various intracellular, intercellular, and cell-matrix recognition events (8, 9). Glycosylation determines the biological activities of the glycodelin isoforms (2, 10). For example, both GdA and GdF inhibit the spermatozoa-zona pellucida binding (11) via fucosyltransferase-5 (12), but only the latter inhibits progesterone-induced acrosome reaction, thus preventing a premature acrosome reaction of the spermatozoa. There is evidence that cumulus cells can convert exogenous GdA and -F to GdC, the physicochemical properties of which suggest that it is differently glycosylated compared with GdA/F (5). Moreover, GdC stimulated spermatozoa-zona pellucida binding in a dose-dependent manner, and it effectively displaced sperm-bound GdA and -F (4, 5). GdS suppresses capacitation probably via its inhibitory activity on cholesterol efflux from spermatozoa (13).Except for the effects on fertilization, GdA is involved in fetomaternal defense. This glycodelin isoform suppresses proliferation and induces apoptosis of T cells (2) and inhibits natural killer cell (14) and B-cell (15) activities. Glycosylation is involved in the binding of GdA to receptors on T cells (16). The sialic acid of GdA contributes to the apoptotic activity in T cells (17, 18) and binding to CD45, a potential GdA receptor (16). The importance of glycosylation in glycodelin is further shown by the absence of immunosuppressive activities in GdS with different glycosylation (18). The immunomodulating activities of GdF and GdC are unknown.Our previous work showed that glycans are indispensable for the different glycodelins to exhibit their binding activities and biological effects (13, 19, 20). The present study aims to identify the effect of all four glycodelin isoforms on lymphocyte viability, cell death, and interleukin-2 (IL-2) secretion and to correlate these bioactivities with their glycosylation patterns determined by mass spectrometry.  相似文献   

17.
Rapid protein kinase D (PKD) activation and phosphorylation via protein kinase C (PKC) have been extensively documented in many cell types cells stimulated by multiple stimuli. In contrast, little is known about the role and mechanism(s) of a recently identified sustained phase of PKD activation in response to G protein-coupled receptor agonists. To elucidate the role of biphasic PKD activation, we used Swiss 3T3 cells because PKD expression in these cells potently enhanced duration of ERK activation and DNA synthesis in response to Gq-coupled receptor agonists. Cell treatment with the preferential PKC inhibitors GF109203X or Gö6983 profoundly inhibited PKD activation induced by bombesin stimulation for <15 min but did not prevent PKD catalytic activation induced by bombesin stimulation for longer times (>60 min). The existence of sequential PKC-dependent and PKC-independent PKD activation was demonstrated in 3T3 cells stimulated with various concentrations of bombesin (0.3–10 nm) or with vasopressin, a different Gq-coupled receptor agonist. To gain insight into the mechanisms involved, we determined the phosphorylation state of the activation loop residues Ser744 and Ser748. Transphosphorylation targeted Ser744, whereas autophosphorylation was the predominant mechanism for Ser748 in cells stimulated with Gq-coupled receptor agonists. We next determined which phase of PKD activation is responsible for promoting enhanced ERK activation and DNA synthesis in response to Gq-coupled receptor agonists. We show, for the first time, that the PKC-independent phase of PKD activation mediates prolonged ERK signaling and progression to DNA synthesis in response to bombesin or vasopressin through a pathway that requires epidermal growth factor receptor-tyrosine kinase activity. Thus, our results identify a novel mechanism of Gq-coupled receptor-induced mitogenesis mediated by sustained PKD activation through a PKC-independent pathway.The understanding of the mechanisms that control cell proliferation requires the identification of the molecular pathways that govern the transition of quiescent cells into the S phase of the cell cycle. In this context the activation and phosphorylation of protein kinase D (PKD),4 the founding member of a new protein kinase family within the Ca2+/calmodulin-dependent protein kinase (CAMK) group and separate from the previously identified PKCs (for review, see Ref. 1), are attracting intense attention. In unstimulated cells, PKD is in a state of low catalytic (kinase) activity maintained by autoinhibition mediated by the N-terminal domain, a region containing a repeat of cysteinerich zinc finger-like motifs and a pleckstrin homology (PH) domain (14). Physiological activation of PKD within cells occurs via a phosphorylation-dependent mechanism first identified in our laboratory (57). In response to cellular stimuli (1), including phorbol esters, growth factors (e.g. PDGF), and G protein-coupled receptor (GPCR) agonists (6, 816) that signal through Gq, G12, Gi, and Rho (11, 1519), PKD is converted into a form with high catalytic activity, as shown by in vitro kinase assays performed in the absence of lipid co-activators (5, 20).During these studies multiple lines of evidence indicated that PKC activity is necessary for rapid PKD activation within intact cells. For example, rapid PKD activation was selectively and potently blocked by cell treatment with preferential PKC inhibitors (e.g. GF109203X or Gö6983) that do not directly inhibit PKD catalytic activity (5, 20), implying that PKD activation in intact cells is mediated directly or indirectly through PKCs. Many reports demonstrated the operation of a rapid PKC/PKD signaling cascade induced by multiple GPCR agonists and other receptor ligands in a range of cell types (for review, see Ref. 1). Our previous studies identified Ser744 and Ser748 in the PKD activation loop (also referred as activation segment or T-loop) as phosphorylation sites critical for PKC-mediated PKD activation (1, 4, 7, 17, 21). Collectively, these findings demonstrated the existence of a rapidly activated PKC-PKD protein kinase cascade(s). In a recent study we found that the rapid PKC-dependent PKD activation was followed by a late, PKC-independent phase of catalytic activation and phosphorylation induced by stimulation of the bombesin Gq-coupled receptor ectopically expressed in COS-7 cells (22). This study raised the possibility that PKD mediates rapid biological responses downstream of PKCs, whereas, in striking contrast, PKD could mediate long term responses through PKC-independent pathways. Despite its potential importance for defining the role of PKC and PKD in signal transduction, this hypothesis has not been tested in any cell type.Accumulating evidence demonstrates that PKD plays an important role in several cellular processes and activities, including signal transduction (14, 2325), chromatin organization (26), Golgi function (27, 28), gene expression (2931), immune regulation (26), and cell survival, adhesion, motility, differentiation, DNA synthesis, and proliferation (for review, see Ref. 1). In Swiss 3T3 fibroblasts, a cell line used extensively as a model system to elucidate mechanisms of mitogenic signaling (3234), PKD expression potently enhances ERK activation, DNA synthesis, and cell proliferation induced by Gq-coupled receptor agonists (8, 14). Here, we used this model system to elucidate the role and mechanism(s) of biphasic PKD activation. First, we show that the Gq-coupled receptor agonists bombesin and vasopressin, in contrast to phorbol esters, specifically induce PKD activation through early PKC-dependent and late PKC-independent mechanisms in Swiss 3T3 cells. Subsequently, we demonstrate for the first time that the PKC-independent phase of PKD activation is responsible for promoting ERK signaling and progression to DNA synthesis through an epidermal growth factor receptor (EGFR)-dependent pathway. Thus, our results identify a novel mechanism of Gq-coupled receptor-induced mitogenesis mediated by sustained PKD activation through a PKC-independent pathway.  相似文献   

18.
19.
Thimet oligopeptidase (EC 3.4.24.15; EP24.15) is an intracellular enzyme that has been proposed to metabolize peptides within cells, thereby affecting antigen presentation and G protein-coupled receptor signal transduction. However, only a small number of intracellular substrates of EP24.15 have been reported previously. Here we have identified over 100 peptides in human embryonic kidney 293 (HEK293) cells that are derived from intracellular proteins; many but not all of these peptides are substrates or products of EP24.15. First, cellular peptides were extracted from HEK293 cells and incubated in vitro with purified EP24.15. Then the peptides were labeled with isotopic tags and analyzed by mass spectrometry to obtain quantitative data on the extent of cleavage. A related series of experiments tested the effect of overexpression of EP24.15 on the cellular levels of peptides in HEK293 cells. Finally, synthetic peptides that corresponded to 10 of the cellular peptides were incubated with purified EP24.15 in vitro, and the cleavage was monitored by high pressure liquid chromatography and mass spectrometry. Many of the EP24.15 substrates identified by these approaches are 9–11 amino acids in length, supporting the proposal that EP24.15 can function in the degradation of peptides that could be used for antigen presentation. However, EP24.15 also converts some peptides into products that are 8–10 amino acids, thus contributing to the formation of peptides for antigen presentation. In addition, the intracellular peptides described here are potential candidates to regulate protein interactions within cells.Intracellular protein turnover is a crucial step for cell functioning, and if this process is impaired, the elevated levels of aged proteins usually lead to the formation of intracellular insoluble aggregates that can cause severe pathologies (1). In mammalian cells, most proteins destined for degradation are initially tagged with a polyubiquitin chain in an energy-dependent process and then digested to small peptides by the 26 S proteasome, a large proteolytic complex involved in the regulation of cell division, gene expression, and other key processes (2, 3). In eukaryotes, 30–90% of newly synthesized proteins may be degraded by proteasomes within minutes of synthesis (3, 4). In addition to proteasomes, other extralysosomal proteolytic systems have been reported (5, 6). The proteasome cleaves proteins into peptides that are typically 2–20 amino acids in length (7). In most cases, these peptides are thought to be rapidly hydrolyzed into amino acids by aminopeptidases (810). However, some intracellular peptides escape complete degradation and are imported into the endoplasmic reticulum where they associate with major histocompatibility complex class I (MHC-I)3 molecules and traffic to the cell surface for presentation to the immune system (1012). Additionally, based on the fact that free peptides added to the intracellular milieu can regulate cellular functions mediated by protein interactions such as gene regulation, metabolism, cell signaling, and protein targeting (13, 14), intracellular peptides generated by proteasomes that escape degradation have been suggested to play a role in regulating protein interactions (15). Indeed, oligopeptides isolated from rat brain tissue using the catalytically inactive EP24.15 (EC 3.4.24.15) were introduced into Chinese hamster ovarian-S and HEK293 cells and were found capable of altering G protein-coupled receptor signal transduction (16). Moreover, EP24.15 overexpression itself changed both angiotensin II and isoproterenol signal transduction, suggesting a physiological function for its intracellular substrates/products (16).EP24.15 is a zinc-dependent peptidase of the metallopeptidase M3 family that contains the HEXXH motif (17). This enzyme was first described as a neuropeptide-degrading enzyme present in the soluble fraction of brain homogenates (18). Whereas EP24.15 can be secreted (19, 20), its predominant location in the cytosol and nucleus suggests that the primary function of this enzyme is not the extracellular degradation of neuropeptides and hormones (21, 22). EP24.15 was shown in vivo to participate in antigen presentation through MHC-I (2325) and in vitro to bind (26) or degrade (27) some MHC-I associated peptides. EP24.15 has also been shown in vitro to degrade peptides containing 5–17 amino acids produced after proteasome digestion of β-casein (28). EP24.15 shows substrate size restriction to peptides containing from 5 to 17 amino acids because of its catalytic center that is located in a deep channel (29). Despite the size restriction, EP24.15 has a broad substrate specificity (30), probably because a significant portion of the enzyme-binding site is lined with potentially flexible loops that allow reorganization of the active site following substrate binding (29). Recently, it has also been suggested that certain substrates may be cleaved by an open form of EP24.15 (31). This characteristic is supported by the ability of EP24.15 to accommodate different amino acid residues at subsites S4 to S3′, which even includes the uncommon post-proline cleavage (30). Such biochemical and structural features make EP24.15 a versatile enzyme to degrade structurally unrelated oligopeptides.Previously, brain peptides that bound to catalytically inactive EP24.15 were isolated and identified using mass spectrometry (22). The majority of peptides captured by the inactive enzyme were intracellular protein fragments that efficiently interacted with EP24.15; the smallest peptide isolated in these assays contained 5 and the largest 17 amino acids (15, 16, 22, 32), which is within the size range previously reported for natural and synthetic substrates of EP24.15 (18, 30, 33, 34). Interestingly, the peptides released by the proteasome are in the same size range of EP24.15 competitive inhibitors/substrates (7, 35, 36). Taken altogether, these data suggest that in the intracellular environment EP24.15 could further cleave proteasome-generated peptides unrelated to MHC-I antigen presentation (15).Although the mutated inactive enzyme “capture” assay was successful in identifying several cellular protein fragments that were substrates for EP24.15, it also found some interacting peptides that were not substrates. In this study, we used several approaches to directly screen for cellular peptides that were cleaved by EP24.15. The first approach involved the extraction of cellular peptides from the HEK293 cell line, incubation in vitro with purified EP24.15, labeling with isotopic tags, and analysis by mass spectrometry to obtain quantitative data on the extent of cleavage. The second approach examined the effect of EP24.15 overexpression on the cellular levels of peptides in the HEK293 cell line. The third set of experiments tested synthetic peptides with purified EP24.15 in vitro, and examined cleavage by high pressure liquid chromatography and mass spectrometry. Collectively, these studies have identified a large number of intracellular peptides, including those that likely represent the endogenous substrates and products of EP24.15, and this original information contributes to a better understanding of the function of this enzyme in vivo.  相似文献   

20.
We report here a biochemical and structural characterization of domain 2 of the nonstructural 5A protein (NS5A) from the JFH1 Hepatitis C virus strain and its interactions with cyclophilins A and B (CypA and CypB). Gel filtration chromatography, circular dichroism spectroscopy, and finally NMR spectroscopy all indicate the natively unfolded nature of this NS5A-D2 domain. Because mutations in this domain have been linked to cyclosporin A resistance, we used NMR spectroscopy to investigate potential interactions between NS5A-D2 and cellular CypA and CypB. We observed a direct molecular interaction between NS5A-D2 and both cyclophilins. The interaction surface on the cyclophilins corresponds to their active site, whereas on NS5A-D2, it proved to be distributed over the many proline residues of the domain. NMR heteronuclear exchange spectroscopy yielded direct evidence that many proline residues in NS5A-D2 form a valid substrate for the enzymatic peptidyl-prolyl cis/trans isomerase (PPIase) activity of CypA and CypB.Hepatitis C virus (HCV)4 is a small, positive strand, RNA-enveloped virus belonging to the Flaviviridae family and the genus Hepacivirus. With 120–180 million chronically infected individuals worldwide, hepatitis C virus infection represents a major cause of chronic hepatitis, liver cirrhosis, and hepatocellular carcinoma (1). The HCV viral genome (∼9.6 kb) codes for a unique polyprotein of ∼3000 amino acids (recently reviewed in Refs. 24). Following processing via viral and cellular proteases, this polyprotein gives rise to at least 10 viral proteins, divided into structural (core, E1, and E2 envelope glycoproteins) and nonstructural proteins (p7, NS2, NS3, NS4A, NS4B, NS5A, NS5B). Nonstructural proteins are involved in polyprotein processing and viral replication. The set composed of NS3, NS4A, NS4B, NS5A, and NS5B constitutes the minimal protein component required for viral replication (5).Cyclophilins are cellular proteins that have been identified first as CsA-binding proteins (6). As FK506-binding proteins (FKBP) and parvulins, cyclophilins are peptidyl-prolyl cis/trans isomerases (PPIase) that catalyze the cis/trans isomerization of the peptide linkage preceding a proline (6, 7). Several subtypes of cyclophilins are present in mammalian cells (8). They share a high sequence homology and a well conserved three-dimensional structure but display significant differences in their primary cellular localization and in abundance (9). CypA, the most abundant of the cyclophilins, is primarily cytoplasmic, whereas CypB is directed to the endoplasmic reticulum lumen or the secretory pathway. CypD, on the other hand, is the mitochondrial cyclophilin. Cyclophilins are involved in numerous physiological processes such as protein folding, immune response, and apoptosis and also in the replication cycle of viruses including vaccinia virus, vesicular stomatitis virus, severe acute respiratory syndrome (SARS)-coronavirus, and human immunodeficiency virus (HIV) (for review see Ref. 10). For HIV, CypA has been shown to interact with the capsid domain of the HIV Gag precursor polyprotein (11). CypA thereby competes with capsid domain/TRIM5 interaction, resulting in a loss of the antiviral protective effect of the cellular restriction factor TRIM5α (12, 13). Moreover, it has been shown that CypA catalyzes the cis/trans isomerization of Gly221-Pro222 in the capsid domain and that it has functional consequences for HIV replication efficiency (1416). For HCV, Watashi et al. (17) have described a molecular and functional interaction between NS5B, the viral RNA-dependent RNA polymerase (RdRp), and cyclophilin B (CypB). CypB may be a key regulator in HCV replication by modulating the affinity of NS5B for RNA. This regulation is abolished in the presence of cyclosporin A (CsA), an inhibitor of cyclophilins (6). These results provided for the first time a molecular mechanism for the early-on observed anti-HCV activity of CsA (1820). Although this initial report suggests that only CypB would be involved in the HCV replication process (17), a growing number of studies have recently pointed out a role for other cyclophilins (2125).In vitro selection of CsA-resistant HCV mutants indicated the importance of two HCV nonstructural proteins, NS5B and NS5A (26), with a preponderant effect for mutations in the C-terminal half of NS5A. NS5A is a large phosphoprotein (49 kDa), indispensable for HCV replication and particle assembly (2729), but for which the exact function(s) in the HCV replication cycle remain to be elucidated. This nonstructural protein is anchored to the cytoplasmic leaflet of the endoplasmic reticulum membrane via an N-terminal amphipathic α-helix (residues 1–27) (30, 31). Its cytoplasmic sequence can be divided into three domains: D1 (residues 27–213), D2 (residues 250–342), and D3 (residues 356–447), all connected by low complexity sequences (32). D1, a zinc-binding domain, adopts a dimeric claw-shaped structure, which is proposed to interact with RNA (33, 34). NS5A-D2 is essential for HCV replication, whereas NS5A-D3 is a key determinant for virus infectious particle assembly (27, 35). NS5A-D2 and -D3, for which sequence conservation among HCV genotypes is significantly lower than for D1, have been proposed to be natively unfolded domains (28, 32). Molecular and structural characterization of NS5A-D2 from HCV genotype 1a has confirmed the disordered nature of this domain (36, 37).As it is still not clear which cyclophilins are cofactors for HCV replication, and as mutations in HCV NS5A protein have been associated with CsA resistance, we decided to examine the interaction between both CypA and CypB and domain 2 of the HCV NS5A protein. We first characterized, at the molecular level, NS5A-D2 from the HCV JFH1 infectious strain (genotype 2a) and showed by NMR spectroscopy that this natively unfolded domain indeed interacts with both cyclophilin A and cyclophilin B. Our NMR chemical shift mapping experiments indicated that the interaction occurs at the level of the cyclophilin active site, whereas it lacks a precise localization on NS5A-D2. A peptide derived from the only well conserved amino acid motif in NS5A-D2 did interact with cyclophilin A but only with a 10-fold lower affinity than the full domain. We concluded from this that the many proline residues form multiple anchoring points, especially when they adopt the cis conformation. NMR exchange spectroscopy further demonstrated that NS5A-D2 is a substrate for the PPIase activities of both CypA and CypB. Both the NS5A/cyclophilin interaction and the PPIase activity of the cyclophilins on NS5A-D2 were abolished by CsA, underscoring the specificity of the interaction.  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