首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
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.
Carbon dioxide is fundamental to the physiology of all organisms. There is considerable interest in the precise molecular mechanisms that organisms use to directly sense CO2. Here we demonstrate that a mammalian recombinant G-protein-activated adenylyl cyclase and the related Rv1625c adenylyl cyclase of Mycobacterium tuberculosis are specifically stimulated by CO2. Stimulation occurred at physiological concentrations of CO2 through increased kcat. CO2 increased the affinity of enzyme for metal co-factor, but contact with metal was not necessary as CO2 interacted directly with apoenzyme. CO2 stimulated the activity of both G-protein-regulated adenylyl cyclases and Rv1625c in vivo. Activation of G-protein regulated adenylyl cyclases by CO2 gave a corresponding increase in cAMP-response element-binding protein (CREB) phosphorylation. Comparison of the responses of the G-protein regulated adenylyl cyclases and the molecularly, and biochemically distinct mammalian soluble adenylyl cyclase revealed that whereas G-protein-regulated enzymes are responsive to CO2, the soluble adenylyl cyclase is responsive to both CO2 and bicarbonate ion. We have, thus, identified a signaling enzyme by which eukaryotes can directly detect and respond to fluctuating CO2.Inorganic carbon (Ci)3 is central to prokaryotic and eukaryotic physiology. The predominant biologically active forms of Ci are CO2 and and their relative contributions to the total Ci pool are pH-dependent. Biological roles for CO2 and include photosynthetic carbon fixation (1), pH homeostasis (2), carbon metabolism (3), activation of virulence in pathogenic organisms (4), sperm maturation (5), and as an alarmone in Drosophila (6, 7).Given its importance in biology, the identification of CO2 responsive signaling pathways is key to understanding how organisms cope with fluctuating CO2. Two seven transmembrane receptors, Gr21a and Gr63a, have been shown to confer CO2 responsiveness in Drosophila neurons (6, 7). Guanylyl cyclase D expressing olfactory neurons also mediate sensitivity to CO2 in mice (8). A role for cGMP-activated channels in CO2 sensing has been observed in CO2 avoidance behavior in Caenorhabditis (9, 10). Despite these impressive advances, no eukaryotic signaling enzymes unequivocally demonstrated to respond to CO2 have been identified.The mammalian soluble adenylyl cyclase (sAC) synthesizes the second messenger 3′,5′-cAMP and is directly stimulated by (1113). Stimulation of sAC by has an unequivocal role in sperm maturation (5, 1416). sAC is a member of the Class III family of adenylyl cyclases (ACs), a family that also includes the G-protein-regulated ACs and many examples from prokaryotic genomes (17, 18). The Class III ACs can be divided into four subclasses (a–d) based upon polymorphisms within the active site (19). sAC is a member of Class IIIb, a subclass characterized partly by replacement of a substrate binding Asp with Thr. The Class IIIa ACs include the mammalian G-protein-stimulated ACs and numerous prokaryotic examples. These have been previously assumed to be non-responsive to Ci (12).All prokaryotic Class IIIb ACs examined to date respond to Ci including enzymes from organisms as diverse as Anabaena PCC 7120, Mycobacterium tuberculosis, Stigmatella aurantiaca, and Chloroflexus aurantiacus (20, 21). Two Class IIIb ACs, Slr1991 of Synechocystis PCC 6803 and CyaB1 of Anabaena PCC 7120, have been proven to respond to CO2 and not , giving rise to the idea of AC as a true gas-sensing molecule (22, 23). The finding that Class IIIb ACs respond to CO2 and not necessitates an examination of the assumption that G-protein-regulated ACs and related prokaryotic enzymes do not respond to Ci.Here we demonstrate, contrary to previous work, that a recombinant G-protein-regulated AC and the Class IIIa Rv1625c AC of M. tuberculosis H37Rv show a pH-dependent response to Ci due to specific stimulation by CO2 at physiologically relevant concentrations. CO2 interacted directly with the apoprotein and modulated the activity of both the prokaryotic enzyme and G-protein-regulated AC in vivo. Finally, we contrasted the responses of sAC- and G-protein-regulated ACs to different species of Ci and propose that the mammalian cAMP signaling pathway is able to discriminate between CO2 and in vivo.  相似文献   

3.
Recent studies have revealed that in G protein-coupled receptor signalings switching between G protein- and β-arrestin (βArr)-dependent pathways occurs. In the case of opioid receptors, the signal is switched from the initial inhibition of adenylyl cyclase (AC) to an increase in AC activity (AC activation) during prolonged agonist treatment. The mechanism of such AC activation has been suggested to involve the switching of G proteins activated by the receptor, phosphorylation of signaling molecules, or receptor-dependent recruitment of cellular proteins. Using protein kinase inhibitors, dominant negative mutant studies and mouse embryonic fibroblast cells isolated from Src kinase knock-out mice, we demonstrated that μ-opioid receptor (OPRM1)-mediated AC activation requires direct association and activation of Src kinase by lipid raft-located OPRM1. Such Src activation was independent of βArr as indicated by the ability of OPRM1 to activate Src and AC after prolonged agonist treatment in mouse embryonic fibroblast cells lacking both βArr-1 and -2. Instead the switching of OPRM1 signals was dependent on the heterotrimeric G protein, specifically Gi2 α-subunit. Among the Src kinase substrates, OPRM1 was phosphorylated at Tyr336 within NPXXY motif by Src during AC activation. Mutation of this Tyr residue, together with mutation of Tyr166 within the DRY motif to Phe, resulted in the complete blunting of AC activation. Thus, the recruitment and activation of Src kinase by OPRM1 during chronic agonist treatment, which eventually results in the receptor tyrosine phosphorylation, is the key for switching the opioid receptor signals from its initial AC inhibition to subsequent AC activation.Classical G protein-coupled receptor (GPCR)2 signaling involves the activation of specific heterotrimeric G proteins and the subsequent dissociation of α- and βγ-subunits. These G protein subunits serve as the activators and/or inhibitors of several effector systems, including adenylyl cyclases, phospholipases, and ion channels (1). However, recent studies have shown that GPCR signaling deviates from such a classical linear model. For example, in kidney and colonic epithelial cells, protease-activated receptor 1 can transduce its signals through either Gαi/o or Gαq subunits via inhibition of small GTPase RhoA or activation of RhoD. Thus, RhoA and RhoD act as molecular switches between the negative and positive signaling activity of protease-activated receptor 1 (2). Another example is the ability of β2-adrenergic receptor to switch from Gs-dependent pathways to non-classical signaling pathways by coupling to pertussis toxin-sensitive Gi proteins in a cAMP-dependent protein kinase/protein kinase C phosphorylation-dependent manner. In this case, the phosphorylation-induced switch in G protein coupling provides the receptor access to alternative signaling pathways. For β2-adrenergic receptors, this leads to a Gi-dependent activation of MAP kinase (3, 4). Furthermore the involvement of protein scaffolds, such as β-arrestins in the MAP kinase cascade, could also alter the GPCR signaling (58). Hence the formation of “signaling units” or “receptosomes” would influence the GPCR signaling process and destination.For opioid receptors, which are members of the rhodopsin GPCR subfamily receptors, signal switching is also observed. Normally opioid receptors inhibit AC activity, activate the MAP kinases and Kir3 K+ channels, inhibit the voltage-dependent Ca2+ channels, and regulate other effectors such as phospholipase C (9). However, during prolonged agonist treatment, not only is there a blunting of these cellular responses but also a compensatory increase in intracellular cAMP level, which is particularly significant upon the removal of the agonist or the addition of an antagonist such as naloxone (1012). This compensatory adenylyl cyclase activation phenomenon has been postulated to be responsible for the development of drug tolerance and dependence (13). The observed change from receptor-mediated AC inhibition to receptor-mediated AC activation reflects possible receptor signal switching. Although the exact mechanism for such signal changes has yet to be elucidated, activation of specific protein kinases and subsequent phosphorylation of AC isoforms (14, 15) and other signaling molecules (16) have been suggested to be the key for observed AC activation. Among all the protein kinases studied, involvement of protein kinase C, MAP kinase, and Raf-1 has been implicated in the activation of AC (1719). Alternative mechanisms, such as agonist-induced receptor internalization and the increase in the constitutive activities of the receptor, also have been suggested to play a role in increased AC activity after prolonged opioid agonist treatment (20). Earlier studies also implicated the switching of the opioid receptor from Gi/Go to Gs coupling during chronic agonist treatment (21). Regardless of the mechanism, the exact molecular events that lead to the switching of opioid receptor from an inhibitory response to a stimulatory response remain elusive.Src kinases, which are members of the nonreceptor tyrosine kinase family, have been implicated in GPCR function because several Src family members such as cSrc, Fyn, and Yes have been reported to be activated by several GPCRs, including β2- (22) and β3 (23)-adrenergic, M2- (24) and M3 (25)-muscarinic, and bradykinin receptors (26). The GPCRs that are capable of activating Src predominantly couple to Gi/o family G proteins (27). Src kinases appear to associate with, and be activated by, GPCRs themselves either through direct interaction with intracellular receptor domains or by binding to GPCR-associated proteins, such as G protein subunits or β-arrestins (27). Src kinase has been reported to be activated by κ- (28) and δ (29)-opioid receptors and regulate the c-Jun kinase and MAP kinase activities. Src kinase within the nucleus accumbens has been implicated in the rewarding effect and hyperlocomotion induced by morphine in mice (30). However, it is not clear whether the Src kinase is activated and involved in the signal transduction in AC activation after chronic opioid agonist administration.Previously we reported that the lipid raft location of the receptor and the Gαi2 proteins are two prerequisites for the observed increase in AC activity during prolonged agonist treatment (31, 32). Because various protein kinases including Src kinases and G proteins have been shown to be enriched in lipid rafts (33), the roles of these cellular proteins in the eventual switching of opioid receptor signals from inhibition to stimulation of AC activity were examined in the current studies. We were able to demonstrate that the association with and subsequent activation of Src kinase by the μ-opioid receptor (OPRM1), which leads to eventual tyrosine phosphorylation of OPRM1, are the cellular events required for the switching of opioid receptor signaling upon chronic agonist treatment.  相似文献   

4.
5.
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   

6.
7.
8.
Perioperative factors including hypoxia, hypocapnia, and certain anesthetics have been suggested to contribute to Alzheimer disease (AD) neuropathogenesis. Desflurane is one of the most commonly used inhalation anesthetics. However, the effects of desflurane on AD neuropathogenesis have not been previously determined. Here, we set out to assess the effects of desflurane and hypoxia on caspase activation, amyloid precursor protein (APP) processing, and amyloid β-protein (Aβ) generation in H4 human neuroglioma cells (H4 naïve cells) as well as those overexpressing APP (H4-APP cells). Neither 12% desflurane nor hypoxia (18% O2) alone affected caspase-3 activation, APP processing, and Aβ generation. However, treatment with a combination of 12% desflurane and hypoxia (18% O2) (desflurane/hypoxia) for 6 h induced caspase-3 activation, altered APP processing, and increased Aβ generation in H4-APP cells. Desflurane/hypoxia also increased levels of β-site APP-cleaving enzyme in H4-APP cells. In addition, desflurane/hypoxia-induced Aβ generation could be reduced by the broad caspase inhibitor benzyloxycarbonyl-VAD. Finally, the Aβ aggregation inhibitor clioquinol and γ-secretase inhibitor L-685,458 attenuated caspase-3 activation induced by desflurane/hypoxia. In summary, desflurane can induce Aβ production and caspase activation, but only in the presence of hypoxia. Pending in vivo confirmation, these data may have profound implications for anesthesia care in elderly patients, and especially those with AD.An estimated 200 million patients worldwide undergo surgery each year. Several reports have suggested that anesthesia and surgery may facilitate development of Alzheimer disease (AD)4 (13). A recent study also reported that patients having coronary artery bypass graft surgery under general anesthesia are at increased risk for AD as compared with those having percutaneous transluminal coronary angioplasty under local anesthesia (4).Genetic evidence, confirmed by neuropathological and biochemical findings, indicates that excessive production and/or accumulation of amyloid β-protein (Aβ) play a fundamental role in the pathology of AD (reviewed in Refs. 5 and 6). Aβ is produced via serial proteolysis of amyloid precursor protein (APP) by aspartyl protease β-site APP-cleaving enzyme (BACE), or β-secretase, andγ-secretase. BACE cleaves APP to generate a 99-residue membrane-associated C terminus fragment (APP-C99). APP-C99 is further cleaved by γ-secretase to release 4-kDa Aβ and β-amyloid precursor protein intracellular domain (79). Presenilin and γ-secretase co-fractionate as a detergent-sensitive, high molecular weight complex (10) that includes at least three other proteins, nicastrin/APH-2, APH-1, and PEN-2, all of which are necessary and sufficient for γ-secretase activity (1113). Increasing evidence indicates that apoptosis is associated with a variety of neurodegenerative disorders, including AD (Refs. 1417; reviewed in Ref. 18). Aβ has been shown to cause caspase activation and apoptosis, which can in turn potentiate Aβ generation (16, 1928). Finally, fibrillar aggregates of Aβ and oligomeric species of Aβ are more neurotoxic (2937).Perioperative factors, including hypoxia (3842), hypocapnia (43), and anesthetics (4447), have been reported to potentially contribute to AD neuropathogenesis. These perioperative factors may also cause post-operative cognitive dysfunction, a dementia associated with surgery and anesthesia, by triggering AD neuropathogenesis.Isoflurane, sevoflurane, and desflurane are the most commonly used inhalation anesthetics. It has been reported that isoflurane enhances the oligomerization and cytotoxicity of Aβ (44) and induces apoptosis (4851). Our recent studies have shown that a clinically relevant concentration of isoflurane can lead to caspase-3 activation, decrease cell viability, alter APP processing, and increase Aβ generation in human H4 neuroglioma cells overexpressing human APP (4547). Loop et al. (49) reported that isoflurane and sevoflurane, but not desflurane, can induce caspase activation and apoptosis in human T lymphocytes. However, effects of desflurane and desflurane plus other perioperative risk factors, e.g. hypoxia, on APP processing and Aβ generation have not been assessed.In the present study, we set out to determine effects of desflurane, hypoxia, and the combination of the two (desflurane/hypoxia) on caspase-3 activation, APP processing, and Aβ generation in H4 human neuroglioma cells (H4 naïve cells) and H4 naïve cells stably transfected to express full-length (FL) APP (H4-APP cells). We also investigated whether the caspase inhibitor, Z-VAD, the γ-secretase inhibitor L-685,458, and the Aβ aggregation inhibitor clioquinol could attenuate desflurane/hypoxia-induced caspase-3 activation and Aβ generation.  相似文献   

9.
The novel α1D L-type Ca2+ channel is expressed in supraventricular tissue and has been implicated in the pacemaker activity of the heart and in atrial fibrillation. We recently demonstrated that PKA activation led to increased α1D Ca2+ channel activity in tsA201 cells by phosphorylation of the channel protein. Here we sought to identify the phosphorylated PKA consensus sites on the α1 subunit of the α1D Ca2+ channel by generating GST fusion proteins of the intracellular loops, N terminus, proximal and distal C termini of the α1 subunit of α1D Ca2+ channel. An in vitro PKA kinase assay was performed for the GST fusion proteins, and their phosphorylation was assessed by Western blotting using either anti-PKA substrate or anti-phosphoserine antibodies. Western blotting showed that the N terminus and C terminus were phosphorylated. Serines 1743 and 1816, two PKA consensus sites, were phosphorylated by PKA and identified by mass spectrometry. Site directed mutagenesis and patch clamp studies revealed that serines 1743 and 1816 were major functional PKA consensus sites. Altogether, biochemical and functional data revealed that serines 1743 and 1816 are major functional PKA consensus sites on the α1 subunit of α1D Ca2+ channel. These novel findings provide new insights into the autonomic regulation of the α1D Ca2+ channel in the heart.L-type Ca2+ channels are essential for the generation of normal cardiac rhythm, for induction of rhythm propagation through the atrioventricular node and for the contraction of the atrial and ventricular muscles (15). L-type Ca2+ channel is a multisubunit complex including α1, β and α2/δ subunits (57). The α1 subunit contains the voltage sensor, the selectivity filter, the ion conduction pore, and the binding sites for all known Ca2+ channel blockers (69). While α1C Ca2+ channel is expressed in the atria and ventricles of the heart (1013), expression of α1D Ca2+ channel is restricted to the sinoatrial (SA)2 and atrioventricular (AV) nodes, as well as in the atria, but not in the adult ventricles (2, 3, 10).Only recently it has been realized that α1D along with α1C Ca2+ channels contribute to L-type Ca2+ current (ICa-L) and they both play important but unique roles in the physiology/pathophysiology of the heart (69). Compared with α1C, α1D L-type Ca2+ channel activates at a more negative voltage range and shows slower current inactivation during depolarization (14, 15). These properties may allow α1D Ca2+ channel to play critical roles in SA and AV nodes function. Indeed, α1D Ca2+ channel knock-out mice exhibit significant SA dysfunction and various degrees of AV block (12, 1619).The modulation of α1C Ca2+ channel by cAMP-dependent PKA phosphorylation has been extensively studied, and the C terminus of α1 was identified as the site of the modulation (2022). Our group was the first to report that 8-bromo-cAMP (8-Br-cAMP), a membrane-permeable cAMP analog, increased α1D Ca2+ channel activity using patch clamp studies (2). However, very little is known about potential PKA phosphorylation consensus motifs on the α1D Ca2+ channel. We therefore hypothesized that the C terminus of the α1 subunit of the α1D Ca2+ channel mediates its modulation by cAMP-dependent PKA pathway.  相似文献   

10.
Many G protein-coupled receptors (GPCRs) recycle after agonist-induced endocytosis by a sequence-dependent mechanism, which is distinct from default membrane flow and remains poorly understood. Efficient recycling of the β2-adrenergic receptor (β2AR) requires a C-terminal PDZ (PSD-95/Discs Large/ZO-1) protein-binding determinant (PDZbd), an intact actin cytoskeleton, and is regulated by the endosomal protein Hrs (hepatocyte growth factor-regulated substrate). The PDZbd is thought to link receptors to actin through a series of protein interaction modules present in NHERF/EBP50 (Na+/H+ exchanger 3 regulatory factor/ezrin-binding phosphoprotein of 50 kDa) family and ERM (ezrin/radixin/moesin) family proteins. It is not known, however, if such actin connectivity is sufficient to recapitulate the natural features of sequence-dependent recycling. We addressed this question using a receptor fusion approach based on the sufficiency of the PDZbd to promote recycling when fused to a distinct GPCR, the δ-opioid receptor, which normally recycles inefficiently in HEK293 cells. Modular domains mediating actin connectivity promoted receptor recycling with similarly high efficiency as the PDZbd itself, and recycling promoted by all of the domains was actin-dependent. Regulation of receptor recycling by Hrs, however, was conferred only by the PDZbd and not by downstream interaction modules. These results suggest that actin connectivity is sufficient to mimic the core recycling activity of a GPCR-linked PDZbd but not its cellular regulation.G protein-coupled receptors (GPCRs)2 comprise the largest family of transmembrane signaling receptors expressed in animals and transduce a wide variety of physiological and pharmacological information. While these receptors share a common 7-transmembrane-spanning topology, structural differences between individual GPCR family members confer diverse functional and regulatory properties (1-4). A fundamental mechanism of GPCR regulation involves agonist-induced endocytosis of receptors via clathrin-coated pits (4). Regulated endocytosis can have multiple functional consequences, which are determined in part by the specificity with which internalized receptors traffic via divergent downstream membrane pathways (5-7).Trafficking of internalized GPCRs to lysosomes, a major pathway traversed by the δ-opioid receptor (δOR), contributes to proteolytic down-regulation of receptor number and produces a prolonged attenuation of subsequent cellular responsiveness to agonist (8, 9). Trafficking of internalized GPCRs via a rapid recycling pathway, a major route traversed by the β2-adrenergic receptor (β2AR), restores the complement of functional receptors present on the cell surface and promotes rapid recovery of cellular signaling responsiveness (6, 10, 11). When co-expressed in the same cells, the δOR and β2AR are efficiently sorted between these divergent downstream membrane pathways, highlighting the occurrence of specific molecular sorting of GPCRs after endocytosis (12).Recycling of various integral membrane proteins can occur by default, essentially by bulk membrane flow in the absence of lysosomal sorting determinants (13). There is increasing evidence that various GPCRs, such as the β2AR, require distinct cytoplasmic determinants to recycle efficiently (14). In addition to requiring a cytoplasmic sorting determinant, sequence-dependent recycling of the β2AR differs from default recycling in its dependence on an intact actin cytoskeleton and its regulation by the conserved endosomal sorting protein Hrs (hepatocyte growth factor receptor substrate) (11, 14). Compared with the present knowledge regarding protein complexes that mediate sorting of GPCRs to lysosomes (15, 16), however, relatively little is known about the biochemical basis of sequence-directed recycling or its regulation.The β2AR-derived recycling sequence conforms to a canonical PDZ (PSD-95/Discs Large/ZO-1) protein-binding determinant (henceforth called PDZbd), and PDZ-mediated protein association(s) with this sequence appear to be primarily responsible for its endocytic sorting activity (17-20). Fusion of this sequence to the cytoplasmic tail of the δOR effectively re-routes endocytic trafficking of engineered receptors from lysosomal to recycling pathways, establishing the sufficiency of the PDZbd to function as a transplantable sorting determinant (18). The β2AR-derived PDZbd binds with relatively high specificity to the NHERF/EBP50 family of PDZ proteins (21, 22). A well-established biochemical function of NHERF/EBP50 family proteins is to associate integral membrane proteins with actin-associated cytoskeletal elements. This is achieved through a series of protein-interaction modules linking NHERF/EBP50 family proteins to ERM (ezrin-radixin-moesin) family proteins and, in turn, to actin filaments (23-26). Such indirect actin connectivity is known to mediate other effects on plasma membrane organization and function (23), however, and NHERF/EBP50 family proteins can bind to additional proteins potentially important for endocytic trafficking of receptors (23, 25). Thus it remains unclear if actin connectivity is itself sufficient to promote sequence-directed recycling of GPCRs and, if so, if such connectivity recapitulates the normal cellular regulation of sequence-dependent recycling. In the present study, we took advantage of the modular nature of protein connectivity proposed to mediate β2AR recycling (24, 26), and extended the opioid receptor fusion strategy used successfully for identifying diverse recycling sequences in GPCRs (27-29), to address these fundamental questions.Here we show that the recycling activity of the β2AR-derived PDZbd can be effectively bypassed by linking receptors to ERM family proteins in the absence of the PDZbd itself. Further, we establish that the protein connectivity network can be further simplified by fusing receptors to an interaction module that binds directly to actin filaments. We found that bypassing the PDZ-mediated interaction using either domain is sufficient to mimic the ability of the PDZbd to promote efficient, actin-dependent recycling of receptors. Hrs-dependent regulation, however, which is characteristic of sequence-dependent recycling of wild-type receptors, was recapitulated only by the fused PDZbd and not by the proposed downstream interaction modules. These results support a relatively simple architecture of protein connectivity that is sufficient to mimic the core recycling activity of the β2AR-derived PDZbd, but not its characteristic cellular regulation. Given that an increasing number of GPCRs have been shown to bind PDZ proteins that typically link directly or indirectly to cytoskeletal elements (17, 27, 30-32), the present results also suggest that actin connectivity may represent a common biochemical principle underlying sequence-dependent recycling of various GPCRs.  相似文献   

11.
12.
Excessive accumulation of β-amyloid peptides in the brain is a major cause for the pathogenesis of Alzheimer disease. β-Amyloid is derived from β-amyloid precursor protein (APP) through sequential cleavages by β- and γ-secretases, whose enzymatic activities are tightly controlled by subcellular localization. Delineation of how intracellular trafficking of these secretases and APP is regulated is important for understanding Alzheimer disease pathogenesis. Although APP trafficking is regulated by multiple factors including presenilin 1 (PS1), a major component of the γ-secretase complex, and phospholipase D1 (PLD1), a phospholipid-modifying enzyme, regulation of intracellular trafficking of PS1/γ-secretase and β-secretase is less clear. Here we demonstrate that APP can reciprocally regulate PS1 trafficking; APP deficiency results in faster transport of PS1 from the trans-Golgi network to the cell surface and increased steady state levels of PS1 at the cell surface, which can be reversed by restoring APP levels. Restoration of APP in APP-deficient cells also reduces steady state levels of other γ-secretase components (nicastrin, APH-1, and PEN-2) and the cleavage of Notch by PS1/γ-secretase that is more highly correlated with cell surface levels of PS1 than with APP overexpression levels, supporting the notion that Notch is mainly cleaved at the cell surface. In contrast, intracellular trafficking of β-secretase (BACE1) is not regulated by APP. Moreover, we find that PLD1 also regulates PS1 trafficking and that PLD1 overexpression promotes cell surface accumulation of PS1 in an APP-independent manner. Our results clearly elucidate a physiological function of APP in regulating protein trafficking and suggest that intracellular trafficking of PS1/γ-secretase is regulated by multiple factors, including APP and PLD1.An important pathological hallmark of Alzheimer disease (AD)4 is the formation of senile plaques in the brains of patients. The major components of those plaques are β-amyloid peptides (Aβ), whose accumulation triggers a cascade of neurodegenerative steps ending in formation of senile plaques and intraneuronal fibrillary tangles with subsequent neuronal loss in susceptible brain regions (1, 2). Aβ is proteolytically derived from the β-amyloid precursor protein (APP) through sequential cleavages by β-secretase (BACE1), a novel membrane-bound aspartyl protease (3, 4), and by γ-secretase, a high molecular weight complex consisting of at least four components: presenilin (PS), nicastrin (NCT), anterior pharynx-defective-1 (APH-1), and presenilin enhancer-2 (PEN-2) (5, 6). APP is a type I transmembrane protein belonging to a protein family that includes APP-like protein 1 (APLP1) and 2 (APLP2) in mammals (7, 8). Full-length APP is synthesized in the endoplasmic reticulum (ER) and transported through the Golgi apparatus. Most secreted Aβ peptides are generated within the trans-Golgi network (TGN), also the major site of steady state APP in neurons (911). APP can be transported to the cell surface in TGN-derived secretory vesicles if not proteolyzed to Aβ or an intermediate metabolite. At the cell surface APP is either cleaved by α-secretase to produce soluble sAPPα (12) or reinternalized for endosomal/lysosomal degradation (13, 14). Aβ may also be generated in endosomal/lysosomal compartments (15, 16). In contrast to neurotoxic Aβ peptides, sAPPα possesses neuroprotective potential (17, 18). Thus, the subcellular distribution of APP and proteases that process it directly affect the ratio of sAPPα to Aβ, making delineation of the mechanisms responsible for regulating trafficking of all of these proteins relevant to AD pathogenesis.Presenilin (PS) is a critical component of the γ-secretase. Of the two mammalian PS gene homologues, PS1 and PS2, PS1 encodes the major form (PS1) in active γ-secretase (19, 20). Nascent PSs undergo endoproteolytic cleavage to generate an amino-terminal fragment (NTF) and a carboxyl-terminal fragment (CTF) to form a functional PS heterodimer (21). Based on observations that PSs possess two highly conserved aspartate residues indispensable for γ-secretase activity and that specific transition state analogue γ-secretase inhibitors bind to PS1 NTF/CTF heterodimers (5, 22), PSs are believed to be the catalytic component of the γ-secretase complex. PS assembles with three other components, NCT, APH-1, and PEN-2, to form the functional γ-secretase (5, 6). Strong evidence suggests that PS1/γ-secretase resides principally in the ER, early Golgi, TGN, endocytic and intermediate compartments, most of which (except the TGN) are not major subcellular sites for APP (23, 24). In addition to generating Aβ and cleaving APP to release the APP intracellular domain, PS1/γ-secretase cleaves other substrates such as Notch (25), cadherin (26), ErbB4 (27), and CD44 (28), releasing their respective intracellular domains. Interestingly, PS1/γ-secretase cleavage of different substrates seems to occur at different subcellular compartments; APP is mainly cleaved at the TGN and early endosome domains, whereas Notch is predominantly cleaved at the cell surface (9, 11, 29). Thus, perturbing intracellular trafficking of PS1/γ-secretase may alter interactions between PS1/γ-secretase and APP, contributing to either abnormal Aβ generation and AD pathogenesis or decreased access of PS1/γ-secretase to APP such that Aβ production is reduced. However, mechanisms regulating PS1/γ-secretase trafficking warrant further investigation.In addition to participating in γ-secretase activity, PS1 regulates intracellular trafficking of several membrane proteins, including other γ-secretase components (nicastrin, APH-1, and PEN-2) and the substrate APP (reviewed in Ref. 30). Intracellular APP trafficking is highly regulated and requires other factors such as mint family members and SorLA (2). Moreover, we recently found that phospholipase D1 (PLD1), a phospholipid-modifying enzyme that regulates membrane trafficking events, can interact with PS1, and can regulate budding of APP-containing vesicles from the TGN and delivery of APP to the cell surface (31, 32). Interestingly, Kamal et al. (33) identified an axonal membrane compartment that contains APP, BACE1, and PS1 and showed that fast anterograde axonal transport of this compartment is mediated by APP and kinesin-I, implying a traffic-regulating role for APP. Increased APP expression is also shown to decrease retrograde axonal transport of nerve growth factor (34). However, whether APP indeed regulates intracellular trafficking of proteins including BACE1 and PS1/γ-secretase requires further validation. In the present study we demonstrate that intracellular trafficking of PS1, as well as that of other γ-secretase components, but not BACE1, is regulated by APP. APP deficiency promotes cell surface delivery of PS1/γ-secretase complex and facilitates PS1/γ-secretase-mediated Notch cleavage. In addition, we find that PLD1 also regulates intracellular trafficking of PS1 through a different mechanism and more potently than APP.  相似文献   

13.
Ryanodine receptors (RyR) are Ca2+ channels that mediate Ca2+ release from intracellular stores in response to diverse intracellular signals. In RINm5F insulinoma cells, caffeine, and 4-chloro-m-cresol (4CmC), agonists of RyR, stimulated Ca2+ entry that was independent of store-operated Ca2+ entry, and blocked by prior incubation with a concentration of ryanodine that inactivates RyR. Patch-clamp recording identified small numbers of large-conductance (γK = 169 pS) cation channels that were activated by caffeine, 4CmC or low concentrations of ryanodine. Similar channels were detected in rat pancreatic β-cells. In RINm5F cells, the channels were blocked by cytosolic, but not extracellular, ruthenium red. Subcellular fractionation showed that type 3 IP3 receptors (IP3R3) were expressed predominantly in endoplasmic reticulum, whereas RyR2 were present also in plasma membrane fractions. Using RNAi selectively to reduce expression of RyR1, RyR2, or IP3R3, we showed that RyR2 mediates both the Ca2+ entry and the plasma membrane currents evoked by agonists of RyR. We conclude that small numbers of RyR2 are selectively expressed in the plasma membrane of RINm5F pancreatic β-cells, where they mediate Ca2+ entry.Ryanodine receptors (RyR)3 and inositol 1,4,5-trisphosphate receptors (IP3R) (1, 2) are the archetypal intracellular Ca2+ channels. Both are widely expressed, although RyR are more restricted in their expression than IP3R (3, 4). In common with many cells, pancreatic β-cells and insulin-secreting cell lines express both IP3R (predominantly IP3R3) (5, 6) and RyR (predominantly RyR2) (7). Both RyR and IP3R are expressed mostly within membranes of the endoplasmic (ER), where they mediate release of Ca2+. Functional RyR are also expressed in the secretory vesicles (8, 9) or, and perhaps more likely, in the endosomes of β-cells (10). Despite earlier suggestions (11), IP3R are probably not present in the secretory vesicles of β-cells (8, 12, 13).All three subtypes of IP3R are stimulated by IP3 with Ca2+ (1), and the three subtypes of RyR are each directly regulated by Ca2+. However, RyR differ in whether their most important physiological stimulus is depolarization of the plasma membrane (RyR1), Ca2+ (RyR2) or additional intracellular messengers like cyclic ADP-ribose. The latter stimulates both Ca2+ release and insulin secretion in β-cells (8, 14). The activities of both families of intracellular Ca2+ channels are also modulated by many additional signals that act directly or via phosphorylation (15, 16). Although they commonly mediate release of Ca2+ from the ER, both IP3R and RyR select rather poorly between Ca2+ and other cations (permeability ratio, PCa/PK ∼7) (1, 17). This may allow electrogenic Ca2+ release from the ER to be rapidly compensated by uptake of K+ (18), and where RyR or IP3R are expressed in other membranes it may allow them to affect membrane potential.Both Ca2+ entry and release of Ca2+ from intracellular stores contribute to the oscillatory increases in cytosolic Ca2+ concentration ([Ca2+]i) that stimulate exocytosis of insulin-containing vesicles in pancreatic β-cells (7). Glucose rapidly equilibrates across the plasma membrane (PM) of β-cells and its oxidative metabolism by mitochondria increases the cytosolic ATP/ADP ratio, causing KATP channels to close (19). This allows an unidentified leak current to depolarize the PM (20) and activate voltage-gated Ca2+ channels, predominantly L-type Ca2+ channels (21). The resulting Ca2+ entry is amplified by Ca2+-induced Ca2+ release from intracellular stores (7), triggering exocytotic release of insulin-containing dense-core vesicles (22). The importance of this sequence is clear from the widespread use of sulfonylurea drugs, which close KATP channels, in the treatment of type 2 diabetes. Ca2+ uptake by mitochondria beneath the PM further stimulates ATP production, amplifying the initial response to glucose and perhaps thereby contributing to the sustained phase of insulin release (23). However, neither the increase in [Ca2+]i nor the insulin release evoked by glucose or other nutrients is entirely dependent on Ca2+ entry (7, 24) or closure of KATP channels (25). This suggests that glucose metabolism may also more directly activate RyR (7, 26) and/or IP3R (27) to cause release of Ca2+ from intracellular stores. A change in the ATP/ADP ratio is one means whereby nutrient metabolism may be linked to opening of intracellular Ca2+ channels because both RyR (28) and IP3R (1) are stimulated by ATP.The other major physiological regulators of insulin release are the incretins: glucagon-like peptide-1 and glucose-dependent insulinotropic hormone (29). These hormones, released by cells in the small intestine, stimulate synthesis of cAMP in β-cells and thereby potentiate glucose-evoked insulin release (30). These pathways are also targets of drugs used successfully to treat type 2 diabetes (29). The responses of β-cells to cAMP involve both cAMP-dependent protein kinase and epacs (exchange factors activated by cAMP) (31, 32). The effects of the latter are, at least partly, due to release of Ca2+ from intracellular stores via RyR (3335) and perhaps also via IP3R (36). The interplays between Ca2+ and cAMP signaling generate oscillatory changes in the concentrations of both messengers (37). RyR and IP3R are thus implicated in mediating responses to each of the major physiological regulators of insulin secretion: glucose and incretins.Here we report that in addition to expression in intracellular stores, which probably include both the ER and secretory vesicles and/or endosomes, functional RyR2 are also expressed in small numbers in the PM of RINm5F insulinoma cells and rat pancreatic β-cells.  相似文献   

14.
TANK-binding kinase 1 (TBK1) is a key component of Toll-like receptor-dependent and -independent signaling pathways. In response to microbial components, TBK1 activates interferon regulatory factor 3 (IRF3) and cytokine expression. Here we show that TBK1 is a novel target of the γ134.5 protein, a virulence factor whose expression is regulated in a temporal fashion. Remarkably, the γ134.5 protein is required to inhibit IRF3 phosphorylation, nuclear translocation, and the induction of antiviral genes in infected cells. When expressed in mammalian cells, the γ134.5 protein forms complexes with TBK1 and disrupts the interaction of TBK1 and IRF3, which prevents the induction of interferon and interferon-stimulated gene promoters. Down-regulation of TBK1 requires the amino-terminal domain. In addition, unlike wild type virus, a herpes simplex virus mutant lacking γ134.5 replicates efficiently in TBK1-/- cells but not in TBK1+/+ cells. Addition of exogenous interferon restores the antiviral activity in both TBK1-/- and TBK+/+ cells. Hence, control of TBK1-mediated cell signaling by the γ134.5 protein contributes to herpes simplex virus infection. These results reveal that TBK1 plays a pivotal role in limiting replication of a DNA virus.Herpes simplex virus 1 (HSV-1)3 is a large DNA virus that establishes latent or lytic infection, in which the virus triggers innate immune responses. In HSV-infected cells, a number of antiviral mechanisms operate in a cell type- and time-dependent manner (1). In response to double-stranded RNA (dsRNA), Toll-like receptor 3 (TLR3) recruits an adaptor TIR domain-containing adaptor inducing IFN-β and stimulates cytokine expression (2, 3). In the cytoplasm, RNA helicases, RIG-I (retinoid acid-inducible gene-I), and MDA5 (melanoma differentiation associated gene 5) recognize intracellular viral 5′-triphosphate RNA or dsRNA (2, 4). Furthermore, a DNA-dependent activator of IFN-regulatory factor (DAI) senses double-stranded DNA in the cytoplasm and induces cytokine expression (5). There is also evidence that viral entry induces antiviral programs independent of TLR and RIG-I pathways (6). While recognizing distinct viral components, these innate immune pathways relay signals to the two IKK-related kinases, TANK-binding kinase 1 (TBK1) and inducible IκB kinase (IKKi) (2).The IKK-related kinases function as essential components that phosphorylate IRF3 (interferon regulatory factor 3), as well as the closely related IRF7, which translocates to the nucleus and induces antiviral genes, such as interferon-α/β and ISG56 (interferon-stimulated gene 56) (7, 8). TBK1 is constitutively expressed, whereas IKKi is engaged as an inducible gene product of innate immune signaling (9, 10). IRF3 activation is attenuated in TBK1-deficient but not in IKKi-deficient cells (11, 12). Its activation is completely abolished in double-deficient cells (12), suggesting a partially redundant function of TBK1 and IKKi. Indeed, IKKi also negatively regulates the STAT-signaling pathway (13). TBK1/IKKi interacts with several proteins, such as TRAF family member-associated NF-κB activator (TANK), NAP1 (NAK-associated protein 1), similar to NAP1TBK1 adaptor (SINTBAD), DNA-dependent activator of IFN-regulatory factors (DAI), and secretory protein 5 (Sec5) in host cells (5, 1418). These interactions are thought to regulate TBK1/IKKi, which delineates innate as well as adaptive immune responses.Upon viral infection, expression of HSV proteins interferes with the induction of antiviral immunity. When treated with UV or cycloheximide, HSV induces an array of antiviral genes in human lung fibroblasts (19, 20). Furthermore, an HSV mutant, with deletion in immediate early protein ICP0, induces ISG56 expression (21). Accordingly, expression of ICP0 inhibits the induction of antiviral programs mediated by IRF3 or IRF7 (2123). However, although ICP0 negatively regulates IFN-β expression, it is not essential for this effect (24). In HSV-infected human macrophages or dendritic cells, an immediate early protein ICP27 is required to suppress cytokine induction involving IRF3 (25). In this context, it is notable that an HSV mutant, lacking a leaky late gene γ134.5, replicates efficiently in cells devoid of IFN-α/β genes (26). Additionally, the γ134.5 null mutant induces differential cytokine expression as compared with wild type virus (27). Thus, HSV modulation of cytokine expression is a complex process that involves multiple viral components. Currently, the molecular mechanism governing this event is unclear. In this study, we show that HSV γ134.5 targets TBK1 and inhibits antiviral signaling. The data herein reveal a previously unrecognized mechanism by which γ134.5 facilitates HSV replication.  相似文献   

15.
16.
The novel class of protein kinase C (nPKC) isoform η is expressed in platelets, but not much is known about its activation and function. In this study, we investigated the mechanism of activation and functional implications of nPKCη using pharmacological and gene knock-out approaches. nPKCη was phosphorylated (at Thr-512) in a time- and concentration-dependent manner by 2MeSADP. Pretreatment of platelets with MRS-2179, a P2Y1 receptor antagonist, or YM-254890, a Gq blocker, abolished 2MeSADP-induced phosphorylation of nPKCη. Similarly, ADP failed to activate nPKCη in platelets isolated from P2Y1 and Gq knock-out mice. However, pretreatment of platelets with P2Y12 receptor antagonist, AR-C69331MX did not interfere with ADP-induced nPKCη phosphorylation. In addition, when platelets were activated with 2MeSADP under stirring conditions, although nPKCη was phosphorylated within 30 s by ADP receptors, it was also dephosphorylated by activated integrin αIIbβ3 mediated outside-in signaling. Moreover, in the presence of SC-57101, a αIIbβ3 receptor antagonist, nPKCη dephosphorylation was inhibited. Furthermore, in murine platelets lacking PP1cγ, a catalytic subunit of serine/threonine phosphatase, αIIbβ3 failed to dephosphorylate nPKCη. Thus, we conclude that ADP activates nPKCη via P2Y1 receptor and is subsequently dephosphorylated by PP1γ phosphatase activated by αIIbβ3 integrin. In addition, pretreatment of platelets with η-RACK antagonistic peptides, a specific inhibitor of nPKCη, inhibited ADP-induced thromboxane generation. However, these peptides had no affect on ADP-induced aggregation when thromboxane generation was blocked. In summary, nPKCη positively regulates agonist-induced thromboxane generation with no effects on platelet aggregation.Platelets are the key cellular components in maintaining hemostasis (1). Vascular injury exposes subendothelial collagen that activates platelets to change shape, secrete contents of granules, generate thromboxane, and finally aggregate via activated αIIbβ3 integrin, to prevent further bleeding (2, 3). ADP is a physiological agonist of platelets secreted from dense granules and is involved in feedback activation of platelets and hemostatic plug stabilization (4). It activates two distinct G-protein-coupled receptors (GPCRs) on platelets, P2Y1 and P2Y12, which couple to Gq and Gi, respectively (58). Gq activates phospholipase Cβ (PLCβ), which leads to diacyl glycerol (DAG)2 generation and calcium mobilization (9, 10). On the other hand, Gi is involved in inhibition of cAMP levels and PI 3-kinase activation (4, 6). Synergistic activation of Gq and Gi proteins leads to the activation of the fibrinogen receptor integrin αIIbβ3. Fibrinogen bound to activated integrin αIIbβ3 further initiates feed back signaling (outside-in signaling) in platelets that contributes to the formation of a stable platelet plug (11).Protein kinase Cs (PKCs) are serine/threonine kinases known to regulate various platelet functional responses such as dense granule secretion and integrin αIIbβ3 activation (12, 13). Based on their structure and cofactor requirements, PKCs are divided in to three classes: classical (cofactors: DAG, Ca2+), novel (cofactors: DAG) and atypical (cofactors: PIP3) PKC isoforms (14). All the members of the novel class of PKC isoforms (nPKC), viz. nPKC isoforms δ, θ, η, and ε, are expressed in platelets (15), and they require DAG for activation. Among all the nPKCs, PKCδ (15, 16) and PKCθ (1719) are fairly studied in platelets. Whereas nPKCδ is reported to regulate protease-activated receptor (PAR)-mediated dense granule secretion (15, 20), nPKCθ is activated by outside-in signaling and contributes to platelet spreading on fibrinogen (18). On the other hand, the mechanism of activation and functional role of nPKCη is not addressed as yet.PKCs are cytoplasmic enzymes. The enzyme activity of PKCs is modulated via three mechanisms (14, 21): 1) cofactor binding: upon cell stimulus, cytoplasmic PKCs mobilize to membrane, bind cofactors such as DAG, Ca2+, or PIP3, release autoinhibition, and attain an active conformation exposing catalytic domain of the enzyme. 2) phosphorylations: 3-phosphoinositide-dependent kinase 1 (PDK1) on the membrane phosphorylates conserved threonine residues on activation loop of catalytic domain; this is followed by autophosphorylations of serine/threonine residues on turn motif and hydrophobic region. These series of phosphorylations maintain an active conformation of the enzyme. 3) RACK binding: PKCs in active conformation bind receptors for activated C kinases (RACKs) and are lead to various subcellular locations to access the substrates (22, 23). Although various leading laboratories have elucidated the activation of PKCs, the mechanism of down-regulation of PKCs is not completely understood.The premise of dynamic cell signaling, which involves protein phosphorylations by kinases and dephosphorylations by phosphatases has gained immense attention over recent years. PP1, PP2A, PP2B, PHLPP are a few of the serine/threonine phosphatases reported to date. Among them PP1 and PP2 phosphatases are known to regulate various platelet functional responses (24, 25). Furthermore, PP1c, is the catalytic unit of PP1 known to constitutively associate with αIIb and is activated upon integrin engagement with fibrinogen and subsequent outside-in signaling (26). Among various PP1 isoforms, recently PP1γ is shown to positively regulate platelet functional responses (27). Thus, in this study we investigated if the above-mentioned phosphatases are involved in down-regulation of nPKCη. Furthermore, reports from other cell systems suggest that nPKCη regulates ERK/JNK pathways (28). In platelets ERK is known to regulate agonist induced thromboxane generation (29, 30). Thus, we also investigated if nPKCη regulates ERK phosphorylation and thereby agonist-induced platelet functional responses.In this study, we evaluated the activation of nPKCη downstream of ADP receptors and its inactivation by an integrin-associated phosphatase PP1γ. We also studied if nPKCη regulates functional responses in platelets and found that this isoform regulates ADP-induced thromboxane generation, but not fibrinogen receptor activation in platelets.  相似文献   

17.
18.
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   

19.
Neuropathologic and genetics studies as well as transgenic animal models have provided strong evidence linking misfolding and aggregation of α-synuclein to the progression of Parkinson disease (PD) and other related disorders. A growing body of evidence implicates various oligomeric forms of α-synuclein as the toxic species responsible for neurodegeneration and neuronal cell death. Although numerous different oligomeric forms of α-synuclein have been identified in vitro, it is not known which forms are involved in PD or how, when, and where different forms contribute to the progression of PD. Reagents that can interact with specific aggregate forms of α-synuclein would be very useful not only as tools to study how different aggregate forms affect cell function, but also as potential diagnostic and therapeutic agents for PD. Here we show that a single chain antibody fragment (syn-10H scFv) isolated from a phage display antibody library binds to a larger, later stage oligomeric form of α-synuclein than a previously reported oligomeric specific scFv isolated in our laboratory. The scFv described here inhibits aggregation of α-synuclein in vitro, blocks extracellular α-synuclein-induced toxicity in both undifferentiated and differentiated human neuroblastoma cell lines (SH-SY5Y), and specifically recognizes naturally occurring aggregates in PD but not in healthy human brain tissue.Parkinson disease (PD)2 is the second most common neurodegenerative disorder of the elderly, affecting more than 500,000 people in the United States (1), with 50,000 new cases reported each year at an annual cost estimated at 10 billion dollars per year. Pathologically, PD is characterized by the progressive loss of dopaminergic neurons in the substantia nigra and formation of fibrillar cytoplasmic inclusions known as Lewy bodies and Lewy neurites (2, 3). The protein α-synuclein has been strongly linked to PD (4, 5) and other related neurodegenerative disorders (6, 7) by several lines of evidence. 1) It is the major component of the hallmark Lewy body aggregates associated with PD. 2) Mutations (A53T, A30P, and E46K, where A30P is human A30P α-synuclein; A53T is human A53T α-synuclein; E46K is human E46K α-synuclein) or multiplication in the α-synuclein gene have been linked to familial PD (810). 3) Overexpression of α-synuclein in transgenic mice and Drosophila has been shown to induce the formation of PD-like pathological phenotypes and behavior, although the animal models do not in general replicate neuronal loss patterns (11, 12).α-Synuclein is a small protein (14 kDa) expressed mainly in brain tissues and is primarily localized at the presynaptic terminals of neurons (13). The primary structure of α-synuclein consists of three distinct regions. The N-terminal region of α-synuclein includes the mutation sites associated with familial PD (A53T, A30P, and E46K) and contains six imperfectly conserved repeats (KTKEGV) that may facilitate protein-protein binding. This repeat section is predicted to form amphipathic α-helices, typical of the lipid-binding domain of apolipoproteins (14). The central region, non-amyloid component, is extremely hydrophobic and includes a 12-residue stretch (VTGVTAVAQKTV) that is essential for aggregation (15). The C-terminal region is enriched with acidic glutamate and aspartate residues and is responsible for the chaperone function of α-synuclein (16).α-Synuclein normally exists as an unfolded protein, but it can adopt several different folded conformations depending on the environment, including small aggregates or oligomers, spherical and linear protofibrils, as well as the fibrillar structure found in Lewy bodies (14, 15). A growing body of evidence implicates the oligomeric forms of α-synuclein as the toxic species responsible for neurodegeneration and neuronal cell death (1618). Several different oligomeric forms of α-synuclein including spherical, annular (19), pore-like (20), and dopamine-stabilized structures have been identified in vitro (21).α-Synuclein is considered a cytosolic protein, and consequently its pathogenic effect was assumed to be limited to the cytoplasm of single cells. However, recent studies have suggested that α-synuclein also has extracellular pathogenic effects (2225). α-Synuclein was detected in blood plasma and cerebrospinal fluid in both monomeric and oligomeric forms (2225), and the presence of significantly elevated levels of oligomeric species of α-synuclein has been reported extracellularly in plasma and cerebrospinal fluid samples from patients with PD (23). Furthermore, various studies have shown that aggregated α-synuclein added extracellularly to the culture medium is cytotoxic (2632).Despite all these studies, it is still not clear how the various aggregate forms of α-synuclein are involved in the progression of PD. Therefore, reagents that can interact with specific aggregate forms of α-synuclein would be very useful not only for fundamental studies of how α-synuclein aggregates affect cell function but also as potential diagnostic and therapeutic agents for PD.Recently, we reported inhibition of both aggregation and extracellular toxicity of α-synuclein in vitro by a single chain variable domain antibody fragment (scFv) that specifically recognized an oligomeric form of α-synuclein (32). In this study, we describe a second scFv (syn-10H) that binds a larger later stage oligomeric form of α-synuclein than the previously reported scFv. The syn-10H scFv neutralizes α-synuclein-induced toxicity in both undifferentiated and differentiated SH-SY5Y human neuroblastoma cell line and inhibits α-synuclein aggregation in vitro. The syn-10H scFv reacts specifically with homogenized PD brain tissue but does not cross-react with similarly treated samples taken from Alzheimer disease (AD) or healthy brain samples. Such scFvs therefore have potential value as diagnostic reagents to identify the presence of specific oligomeric species in PD tissue and fluid samples. The scFvs also have value as therapeutic agents as they can be used either extracellularly or expressed intracellularly (intrabodies) to prevent formation of toxic aggregates in vivo whether inside or outside of cells. Intrabodies have been used efficiently to neutralize toxic effects of different pathogenic agents, including α-synuclein (3336). Moreover, immunization studies in mouse models of PD have shown that extracellular antibodies can reduce accumulation of intracellular aggregates of α-synuclein (37), thereby providing precedent for the use of scFvs in potential passive vaccination strategies for treating PD.  相似文献   

20.
Photoaffinity labeling of γ-aminobutyric acid type A (GABAA)-receptors (GABAAR) with an etomidate analog and mutational analyses of direct activation of GABAAR by neurosteroids have each led to the proposal that these structurally distinct general anesthetics bind to sites in GABAARs in the transmembrane domain at the interface between the β and α subunits. We tested whether the two ligand binding sites might overlap by examining whether neuroactive steroids inhibited etomidate analog photolabeling. We previously identified (Li, G. D., Chiara, D. C., Sawyer, G. W., Husain, S. S., Olsen, R. W., and Cohen, J. B. (2006) J. Neurosci. 26, 11599–11605) azietomidate photolabeling of GABAAR α1Met-236 and βMet-286 (in αM1 and βM3). Positioning these two photolabeled amino acids in a single type of binding site at the interface of β and α subunits (two copies per pentamer) is consistent with a GABAAR homology model based upon the structure of the nicotinic acetylcholine receptor and with recent αM1 to βM3 cross-linking data. Biologically active neurosteroids enhance rather than inhibit azietomidate photolabeling, as assayed at the level of GABAAR subunits on analytical SDS-PAGE, and protein microsequencing establishes that the GABAAR-modulating neurosteroids do not inhibit photolabeling of GABAAR α1Met-236 or βMet-286 but enhance labeling of α1Met-236. Thus modulatory steroids do not bind at the same site as etomidate, and neither of the amino acids identified as neurosteroid activation determinants (Hosie, A. M., Wilkins, M. E., da Silva, H. M., and Smart, T. G. (2006) Nature 444, 486–489) are located at the subunit interface defined by our etomidate site model.GABAA3 receptors (GABAAR) are major mediators of brain inhibitory neurotransmission and participate in most circuits and behavioral pathways relevant to normal and pathological function (1). GABAAR are subject to modulation by endogenous neurosteroids, as well as myriad clinically important central nervous system drugs including general anesthetics, benzodiazepines, and possibly ethanol (1, 2). The mechanism of GABAAR modulation by these different classes of drugs is of major interest, including identification of the receptor amino acid residues involved in binding and action of the drugs.In the absence of high resolution crystal structures of drug-receptor complexes, the locations of anesthetic binding sites in GABAARs have been predicted based upon analyses of functional properties of point mutant receptors, which identified residues in the α and β subunit M1–M4 transmembrane helices important for modulation by volatile anesthetics (primarily α subunit) and by intravenous agents, including etomidate and propofol (β subunit) (35). Position βM2–15, numbered relative to the N terminus of the helix, functions as a major determinant of etomidate and propofol potency as GABA modulators in vitro and in vivo (68). By contrast, this residue is not implicated for modulation by the neurosteroids, potent endogenous modulators of GABAAR (9).Photoaffinity labeling, which allows the identification of residues in proximity to drug binding sites (10, 11), has been used to identify two GABAAR amino acids covalently modified by the etomidate analog [3H]azietomidate (12): α1Met-236 within αM1 and βMet-286 within βM3. Photolabeling of these residues was inhibited equally by nonradioactive etomidate and enhanced proportionately by GABA present in the assay, consistent with the presence of these two residues in a common drug binding pocket that would be located at the interface between the β and α subunits in the transmembrane domain (12). Mutational analyses identify these positions as etomidate and propofol sensitivity determinants (1315).A recent mutagenesis study (16) identified two other residues in GABAAR αM1 and βM3 as critical for direct activation by neurosteroids, αThr-236 (rat numbering, corresponding to α1Thr-237, bovine numbering used here and by Li et al. (12))4 and βTyr-284. These residues were also proposed to contribute to a neurosteroid binding pocket in the transmembrane domain at the interface between β and α subunits, based upon their location in an alternative GABAAR structural model that positioned those amino acids, and not α1Met-236 or βMet-286, at the subunit interface. For GABAARs and other members of the Cys-loop superfamily of neurotransmitter-gated ion channels, the transmembrane domain of each subunit is made up of a loose bundle of four α helices (M1–M4), with M2 from each subunit contributing to the lumen of the ion channel and M4 positioned peripherally in greatest contact with lipid, as seen in the structures of the Torpedo nicotinic acetylcholine receptor (nAChR) (17) and in distantly related prokaryote homologs (18). However, uncertainties in the alignment of GABAAR subunit sequences relative to those of the nAChR result in alternative GABAAR homology models (12, 19, 20) that differ in the location of amino acids in the M3 and M4 membrane-spanning helices and in the M1 helix in some models (16, 21).If etomidate and neurosteroids both bind at the same β/α interface in the GABAAR transmembrane domain, the limited space available for ligand binding suggests that their binding pockets might overlap and that ligand binding would be mutually exclusive. To address this question, we photolabeled purified bovine brain GABAAR with [3H]azietomidate in the presence of different neuroactive steroids and determined by protein microsequencing whether active neurosteroids inhibited labeling of α1Met-236 and βMet-286, as expected for mutually exclusive binding, or resulted in [3H]azietomidate photolabeling of other amino acids, a possible consequence of allosteric interactions. Active steroids failed to inhibit labeling and enhanced labeling of α1Met-236, clearly indicating that the neurosteroid and the etomidate sites are distinct. Our GABAAR homology model that positions α1Met-236 and βMet-286 at the β/α interface, but not that of Hosie et al. (16), is also consistent with cysteine substitution cross-linking studies (20, 22), which define the proximity relations between amino acids in the αM1, αM2, αM3, and βM3 helices, and these results support the interpretation that the two residues photolabeled by [3H]azietomidate are part of a single subunit interface binding pocket, whereas the steroid sensitivity determinants identified by mutagenesis neither are at the β/α subunit interface nor are contributors to a common binding pocket.  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