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.
3.
Members of the CLC gene family either function as chloride channels or as anion/proton exchangers. The plant AtClC-a uses the pH gradient across the vacuolar membrane to accumulate the nutrient in this organelle. When AtClC-a was expressed in Xenopus oocytes, it mediated exchange and less efficiently mediated Cl/H+ exchange. Mutating the “gating glutamate” Glu-203 to alanine resulted in an uncoupled anion conductance that was larger for Cl than . Replacing the “proton glutamate” Glu-270 by alanine abolished currents. These could be restored by the uncoupling E203A mutation. Whereas mammalian endosomal ClC-4 and ClC-5 mediate stoichiometrically coupled 2Cl/H+ exchange, their transport is largely uncoupled from protons. By contrast, the AtClC-a-mediated accumulation in plant vacuoles requires tight coupling. Comparison of AtClC-a and ClC-5 sequences identified a proline in AtClC-a that is replaced by serine in all mammalian CLC isoforms. When this proline was mutated to serine (P160S), Cl/H+ exchange of AtClC-a proceeded as efficiently as exchange, suggesting a role of this residue in exchange. Indeed, when the corresponding serine of ClC-5 was replaced by proline, this Cl/H+ exchanger gained efficient coupling. When inserted into the model Torpedo chloride channel ClC-0, the equivalent mutation increased nitrate relative to chloride conductance. Hence, proline in the CLC pore signature sequence is important for exchange and conductance both in plants and mammals. Gating and proton glutamates play similar roles in bacterial, plant, and mammalian CLC anion/proton exchangers.CLC proteins are found in all phyla from bacteria to humans and either mediate electrogenic anion/proton exchange or function as chloride channels (1). In mammals, the roles of plasma membrane CLC Cl channels include transepithelial transport (25) and control of muscle excitability (6), whereas vesicular CLC exchangers may facilitate endocytosis (7) and lysosomal function (810) by electrically shunting vesicular proton pump currents (11). In the plant Arabidopsis thaliana, there are seven CLC isoforms (AtClC-a–AtClC-g)2 (1215), which may mostly reside in intracellular membranes. AtClC-a uses the pH gradient across the vacuolar membrane to transport the nutrient nitrate into that organelle (16). This secondary active transport requires a tightly coupled exchange. Astonishingly, however, mammalian ClC-4 and -5 and bacterial EcClC-1 (one of the two CLC isoforms in Escherichia coli) display tightly coupled Cl/H+ exchange, but anion flux is largely uncoupled from H+ when is transported (1721). The lack of appropriate expression systems for plant CLC transporters (12) has so far impeded structure-function analysis that may shed light on the ability of AtClC-a to perform efficient exchange. This dearth of data contrasts with the extensive mutagenesis work performed with CLC proteins from animals and bacteria.The crystal structure of bacterial CLC homologues (22, 23) and the investigation of mutants (17, 1921, 2429) have yielded important insights into their structure and function. CLC proteins form dimers with two largely independent permeation pathways (22, 25, 30, 31). Each of the monomers displays two anion binding sites (22). A third binding site is observed when a certain key glutamate residue, which is located halfway in the permeation pathway of almost all CLC proteins, is mutated to alanine (23). Mutating this gating glutamate in CLC Cl channels strongly affects or even completely suppresses single pore gating (23), whereas CLC exchangers are transformed by such mutations into pure anion conductances that are not coupled to proton transport (17, 19, 20). Another key glutamate, located at the cytoplasmic surface of the CLC monomer, seems to be a hallmark of CLC anion/proton exchangers. Mutating this proton glutamate to nontitratable amino acids uncouples anion transport from protons in the bacterial EcClC-1 protein (27) but seems to abolish transport altogether in mammalian ClC-4 and -5 (21). In those latter proteins, anion transport could be restored by additionally introducing an uncoupling mutation at the gating glutamate (21).The functional complementation by AtClC-c and -d (12, 32) of growth phenotypes of a yeast strain deleted for the single yeast CLC Gef1 (33) suggested that these plant CLC proteins function in anion transport but could not reveal details of their biophysical properties. We report here the first functional expression of a plant CLC in animal cells. Expression of wild-type (WT) and mutant AtClC-a in Xenopus oocytes indicate a general role of gating and proton glutamate residues in anion/proton coupling across different isoforms and species. We identified a proline in the CLC signature sequence of AtClC-a that plays a crucial role in exchange. Mutating it to serine, the residue present in mammalian CLC proteins at this position, rendered AtClC-a Cl/H+ exchange as efficient as exchange. Conversely, changing the corresponding serine of ClC-5 to proline converted it into an efficient exchanger. When proline replaced the critical serine in Torpedo ClC-0, the relative conductance of this model Cl channel was drastically increased, and “fast” protopore gating was slowed.  相似文献   

4.
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   

5.
6.
7.
8.
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   

9.
10.
The presence of extensive reciprocal conformational freedom between the catalytic and the hemopexin-like domains of full-length matrix metalloproteinase-1 (MMP-1) is demonstrated by NMR and small angle x-ray scattering experiments. This finding is discussed in relation to the essentiality of the hemopexin-like domain for the collagenolytic activity of MMP-1. The conformational freedom experienced by the present system, having the shortest linker between the two domains, when compared with similar findings on MMP-12 and MMP-9 having longer and the longest linker within the family, respectively, suggests this type of conformational freedom to be a general property of all MMPs.Matrix metalloproteinases (MMP)2 are extracellular hydrolytic enzymes involved in a variety of processes including connective tissue cleavage and remodeling (13). All 23 members of the family are able to cleave simple peptides derived from connective tissue components such as collagen, gelatin, elastin, etc. A subset of MMPs is able to hydrolyze more resistant polymeric substrates, such as cross-linked elastin, and partially degraded collagen forms, such as gelatin and type IV collagens (4). Intact triple helical type I–III collagen is only attacked by collagenases MMP-1, MMP-8, and MMP-13 and by MMP-2 and MMP-14 (512). Although the detailed mechanism of cleavage of single chain peptides by MMP has been largely elucidated (1319), little is known about the process of hydrolysis of triple helical collagen. In fact, triple helical collagen cannot be accommodated in the substrate-binding groove of the catalytic site of MMPs (9).All MMPs (but MMP-7) in their active form are constituted by a catalytic domain (CAT) and a hemopexin-like domain (HPX) (2022). The CAT domain contains two zinc ions and one to three calcium ions. One zinc ion is at the catalytic site and is responsible for the activity, whereas the other metal ions have structural roles. The isolated CAT domains retain full catalytic activity toward simple peptides and single chain polymeric substrates such as elastin, whereas hydrolysis of triple helical collagen also requires the presence of the HPX domain (9, 2325). It has been shown that the isolated CAT domain regains a small fraction of the activity of the full-length (FL) protein when high amounts of either inactivated full-length proteins or isolated HPX domains are added to the assay solution (9). Finally, it has been shown that the presence of the HPX domain alone alters the CD spectrum of triple helical collagen in a way that suggests its partial unwinding (26, 27). It is tempting to speculate that full-length collagenases attack collagen by first locally unwinding the triple helical structure with the help of the HPX domain and then cleaving the resulting, exposed, single filaments (9, 28).Until 2007, three-dimensional structures of full-length MMPs had been reported only for collagenase MMP-1 (2931) and gelatinase MMP-2 (32). The structures of the two proteins are very similar and show a compact arrangement of the two domains, which are connected by a short linker (14 and 20 amino acids, respectively). It is difficult to envisage that rigid and compact molecules of this type can interact with triple helical collagen in a way that can lead to first unwinding and then cleavage of individual filaments. It has been recently suggested that such concerted action could occur much more easily if the two domains could enjoy at least a partial conformational independence (9). Slight differences in the reciprocal orientation of the CAT and HPX domains of MMP-1 in the presence (29) and absence (30, 31) of the prodomain were indeed taken as a hint that the two domains could experience relative mobility (29).Two recent solution studies have shown that conformational independence is indeed occurring in gelatinase MMP-9 (33) and elastase MMP-12 (34), whereas the x-ray structure of the latter (34) is only slightly less compact than those of MMP-1 (2931) and MMP-2 (32). Among MMPs, MMP-9 features an exceptionally long linker (68 amino acid) (33, 35), which in fact constitutes a small domain by itself (the O-glycosylated domain) (33), and therefore, this inspiring observation can hardly be taken as evidence that conformational freedom is a general characteristic of the two-domain MMPs. MMP-12 features a much more normal 16-amino acid linker, thereby making more probable a general functional role for this conformational freedom (34). However, both MMP-9 and MMP-12 retain their full catalytic activity against their substrates even when deprived of the HPX domain (9). Therefore, the question remains of whether conformational freedom is also a required characteristic for those MMPs that are only active as full-length proteins, i.e. collagenases. Interestingly, the three collagenases (MMP-1, MMP-8, and MMP-13) have the shortest linker (14 amino acids) among all MMPs. Demonstrating or negating the presence of conformational freedom in one of these collagenases would therefore constitute a significant step forward to formulate mechanistic hypotheses on their collagenolytic activity.Our recent studies on MMP-12 in solution (34) have shown that a combination of NMR relaxation studies and small angle x-ray scattering (SAXS) is enough to show the presence and the extent of the relative conformational freedom of the two domains of MMPs. Here we apply the same strategy to full-length MMP-1 and show that sizable conformational freedom is indeed experienced even by this prototypical collagenase, although somewhat less pronounced than that observed for MMP-12.  相似文献   

11.
12.
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   

13.
Hepatocellular carcinoma (HCC) is one of the most common and aggressive human malignancies. Recombinant tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) is a promising anti-tumor agent. However, many HCC cells show resistance to TRAIL-induced apoptosis. In this study, we showed that bortezomib, a proteasome inhibitor, overcame TRAIL resistance in HCC cells, including Huh-7, Hep3B, and Sk-Hep1. The combination of bortezomib and TRAIL restored the sensitivity of HCC cells to TRAIL-induced apoptosis. Comparing the molecular change in HCC cells treated with these agents, we found that down-regulation of phospho-Akt (P-Akt) played a key role in mediating TRAIL sensitization of bortezomib. The first evidence was that bortezomib down-regulated P-Akt in a dose- and time-dependent manner in TRAIL-treated HCC cells. Second, LY294002, a PI3K inhibitor, also sensitized resistant HCC cells to TRAIL-induced apoptosis. Third, knocking down Akt1 by small interference RNA also enhanced TRAIL-induced apoptosis in Huh-7 cells. Finally, ectopic expression of mutant Akt (constitutive active) in HCC cells abolished TRAIL sensitization effect of bortezomib. Moreover, okadaic acid, a protein phosphatase 2A (PP2A) inhibitor, reversed down-regulation of P-Akt in bortezomib-treated cells, and PP2A knockdown by small interference RNA also reduced apoptosis induced by the combination of TRAIL and bortezomib, indicating that PP2A may be important in mediating the effect of bortezomib on TRAIL sensitization. Together, bortezomib overcame TRAIL resistance at clinically achievable concentrations in hepatocellular carcinoma cells, and this effect is mediated at least partly via inhibition of the PI3K/Akt pathway.Hepatocellular carcinoma (HCC)2 is currently the fifth most common solid tumor worldwide and the fourth leading cause of cancer-related death. To date, surgery is still the only curative treatment but is only feasible in a small portion of patients (1). Drug treatment is the major therapy for patients with advanced stage disease. Unfortunately, the response rate to traditional chemotherapy for HCC patients is unsatisfactory (1). Novel pharmacological therapy is urgently needed for patients with advanced HCC. In this regard, the approval of sorafenib might open a new era of molecularly targeted therapy in the treatment of HCC patients.Tumor necrosis factor (TNF)-related apoptosis-inducing ligand (TRAIL), a type II transmembrane protein and a member of the TNF family, is a promising anti-tumor agent under clinical investigation (2). TRAIL functions by engaging its receptors expressed on the surface of target cells. Five receptors specific for TRAIL have been identified, including DR4/TRAIL-R1, DR5/TRAIL-R2, DcR1, DcR2, and osteoprotegerin. Among TRAIL receptors, only DR4 and DR5 contain an effective death domain that is essential to formation of death-inducing signaling complex (DISC), a critical step for TRAIL-induced apoptosis. Notably, the trimerization of the death domains recruits an adaptor molecule, Fas-associated protein with death domain (FADD), which subsequently recruits and activates caspase-8. In type I cells, activation of caspase-8 is sufficient to activate caspase-3 to induce apoptosis; however, in another type of cells (type II), the intrinsic mitochondrial pathway is essential for apoptosis characterized by cleavage of Bid and release of cytochrome c from mitochondria, which subsequently activates caspase-9 and caspase-3 (3).Although TRAIL induces apoptosis in malignant cells but sparing normal cells, some tumor cells are resistant to TRAIL-induced apoptosis. Mechanisms responsible for the resistance include receptors and intracellular resistance. Although the cell surface expression of DR4 or DR5 is absolutely required for TRAIL-induced apoptosis, tumor cells expressing these death receptors are not always sensitive to TRAIL due to intracellular mechanisms. For example, the cellular FLICE-inhibitory protein (c-FLIP), a homologue to caspase-8 but without protease activity, has been linked to TRAIL resistance in several studies (4, 5). In addition, inactivation of Bax, a proapoptotic Bcl-2 family protein, resulted in resistance to TRAIL in MMR-deficient tumors (6, 7), and reintroduction of Bax into Bax-deficient cells restored TRAIL sensitivity (8), indicating that the Bcl-2 family plays a critical role in intracellular mechanisms for resistance of TRAIL.Bortezomib, a proteasome inhibitor approved clinically for multiple myeloma and mantle cell lymphoma, has been investigated intensively for many types of cancer (9). Accumulating studies indicate that the combination of bortezomib and TRAIL overcomes the resistance to TRAIL in various types of cancer, including acute myeloid leukemia (4), lymphoma (1013), prostate (1417), colon (15, 18, 19), bladder (14, 16), renal cell carcinoma (20), thyroid (21), ovary (22), non-small cell lung (23, 24), sarcoma (25), and HCC (26, 27). Molecular targets responsible for the sensitizing effect of bortezomib on TRAIL-induced cell death include DR4 (14, 27), DR5 (14, 20, 2223, 28), c-FLIP (4, 11, 2123, 29), NF-κB (12, 24, 30), p21 (16, 21, 25), and p27 (25). In addition, Bcl-2 family also plays a role in the combinational effect of bortezomib and TRAIL, including Bcl-2 (10, 21), Bax (13, 22), Bak (27), Bcl-xL (21), Bik (18), and Bim (15).Recently, we have reported that Akt signaling is a major molecular determinant in bortezomib-induced apoptosis in HCC cells (31). In this study, we demonstrated that bortezomib overcame TRAIL resistance in HCC cells through inhibition of the PI3K/Akt pathway.  相似文献   

14.
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   

15.
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   

16.
17.
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   

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.
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   

20.
Aggregation of the Ure2 protein is at the origin of the [URE3] prion trait in the yeast Saccharomyces cerevisiae. The N-terminal region of Ure2p is necessary and sufficient to induce the [URE3] phenotype in vivo and to polymerize into amyloid-like fibrils in vitro. However, as the N-terminal region is poorly ordered in the native state, making it difficult to detect structural changes in this region by spectroscopic methods, detailed information about the fibril assembly process is therefore lacking. Short fibril-forming peptide regions (4–7 residues) have been identified in a number of prion and other amyloid-related proteins, but such short regions have not yet been identified in Ure2p. In this study, we identify a unique cysteine mutant (R17C) that can greatly accelerate the fibril assembly kinetics of Ure2p under oxidizing conditions. We found that the segment QVNI, corresponding to residues 18–21 in Ure2p, plays a critical role in the fast assembly properties of R17C, suggesting that this segment represents a potential amyloid-forming region. A series of peptides containing the QVNI segment were found to form fibrils in vitro. Furthermore, the peptide fibrils could seed fibril formation for wild-type Ure2p. Preceding the QVNI segment with a cysteine or a hydrophobic residue, instead of a charged residue, caused the rate of assembly into fibrils to increase greatly for both peptides and full-length Ure2p. Our results indicate that the potential amyloid stretch and its preceding residue can modulate the fibril assembly of Ure2p to control the initiation of prion formation.The [URE3] phenotype of Saccharomyces cerevisiae arises because of conversion of the Ure2 protein to an aggregated propagatable prion state (1, 2). Ure2p contains two regions: a poorly structured N-terminal region and a compactly folded C-terminal region (3, 4). The N-terminal region is rich in Asn and Gln residues, is highly flexible, and is without any detectable ordered secondary structure (46). This region is necessary and sufficient for prion behavior in vivo (2) and amyloid-forming capacity in vitro (5, 7), so it is referred to as the prion domain (PrD).2 The C-terminal region has a fold similar to the glutathione S-transferase superfamily (8, 9) and possesses glutathione-dependent peroxidase activity (10). Upon fibril formation, the N-terminal region undergoes a significant conformational change from an unfolded to a thermally resistant conformation (11), whereas the glutathione S-transferase-like C-terminal domain retains its enzymatic activity, suggesting that little conformational change occurs (10, 12). Ure2p fibrils show various morphologies, including variations in thickness and the presence or absence of a periodic twist (1316). The overall structure of the fibrils imaged by cryoelectron microscopy suggests that the intact fibrils contain a 4-nm amyloid filament backbone surrounded by C-terminal globular domains (17).It is widely accepted that disulfide bonds play a critical role in maintaining protein stability (1821) and also affect the process of protein folding by influencing the folding pathway (2225). A recent study shows that the presence of a disulfide bond in a protein can markedly accelerate the folding process (26). Therefore, a disulfide bond is a useful tool to study protein folding. In the study of prion and other amyloid-related proteins, cysteine scanning has been widely used to study the structure of amyloid fibrils, the driving force of amyloid formation, and the plasticity of amyloid fibrils (13, 2731).Short segments from amyloid-related proteins, including IAPP (islet amyloid polypeptide), β2-microglobulin, insulin, and the amyloid-β peptide, show amyloid-forming capacity (3234). Hence, the amyloid stretch hypothesis has been proposed, which suggests that a short amino acid stretch bearing a highly amyloidogenic motif might supply most of the driving force needed to trigger the self-catalytic assembly process of a protein to form fibrils (35, 36). In support of this hypothesis, it was found that the insertion of an amyloidogenic stretch into a non-amyloid-related protein can trigger the amyloidosis of the protein (36). At the same time, the structural information obtained from microcrystals formed by amyloidogenic stretches and bearing cross-β-structure has contributed significantly to our understanding of the structure of intact fibrils at the atomic level (34, 37). However, no amyloidogenic stretches <10 amino acids have so far been identified in the yeast prion protein Ure2.In this study, we performed a cysteine scan within the N-terminal PrD of Ure2p and found a unique cysteine mutant (R17C) that eliminates the lag phase of the Ure2p fibril assembly reaction upon the addition of oxidizing agents. Furthermore, we identified a 4-residue region adjacent to Arg17 as a potential amyloid stretch in Ure2p.  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