首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Early onset generalized dystonia (DYT1) is an autosomal dominant neurological disorder caused by deletion of a single glutamate residue (torsinA ΔE) in the C-terminal region of the AAA+ (ATPases associated with a variety of cellular activities) protein torsinA. The pathogenic mechanism by which torsinA ΔE mutation leads to dystonia remains unknown. Here we report the identification and characterization of a 628-amino acid novel protein, printor, that interacts with torsinA. Printor co-distributes with torsinA in multiple brain regions and co-localizes with torsinA in the endoplasmic reticulum. Interestingly, printor selectively binds to the ATP-free form but not to the ATP-bound form of torsinA, supporting a role for printor as a cofactor rather than a substrate of torsinA. The interaction of printor with torsinA is completely abolished by the dystonia-associated torsinA ΔE mutation. Our findings suggest that printor is a new component of the DYT1 pathogenic pathway and provide a potential molecular target for therapeutic intervention in dystonia.Early onset generalized torsion dystonia (DYT1) is the most common and severe form of hereditary dystonia, a movement disorder characterized by involuntary movements and sustained muscle spasms (1). This autosomal dominant disease has childhood onset and its dystonic symptoms are thought to result from neuronal dysfunction rather than neurodegeneration (2, 3). Most DYT1 cases are caused by deletion of a single glutamate residue at positions 302 or 303 (torsinA ΔE) of the 332-amino acid protein torsinA (4). In addition, a different torsinA mutation that deletes amino acids Phe323–Tyr328 (torsinA Δ323–328) was identified in a single family with dystonia (5), although the pathogenic significance of this torsinA mutation is unclear because these patients contain a concomitant mutation in another dystonia-related protein, ϵ-sarcoglycan (6). Recently, genetic association studies have implicated polymorphisms in the torsinA gene as a genetic risk factor in the development of adult-onset idiopathic dystonia (7, 8).TorsinA contains an N-terminal endoplasmic reticulum (ER)3 signal sequence and a 20-amino acid hydrophobic region followed by a conserved AAA+ (ATPases associated with a variety of cellular activities) domain (9, 10). Because members of the AAA+ family are known to facilitate conformational changes in target proteins (11, 12), it has been proposed that torsinA may function as a molecular chaperone (13, 14). TorsinA is widely expressed in brain and multiple other tissues (15) and is primarily associated with the ER and nuclear envelope (NE) compartments in cells (1620). TorsinA is believed to mainly reside in the lumen of the ER and NE (1719) and has been shown to bind lamina-associated polypeptide 1 (LAP1) (21), lumenal domain-like LAP1 (LULL1) (21), and nesprins (22). In addition, recent evidence indicates that a significant pool of torsinA exhibits a topology in which the AAA+ domain faces the cytoplasm (20). In support of this topology, torsinA is found in the cytoplasm, neuronal processes, and synaptic terminals (2, 3, 15, 2326) and has been shown to bind cytosolic proteins snapin (27) and kinesin light chain 1 (20). TorsinA has been proposed to play a role in several cellular processes, including dopaminergic neurotransmission (2831), NE organization and dynamics (17, 22, 32), and protein trafficking (27, 33). However, the precise biological function of torsinA and its regulation remain unknown.To gain insights into torsinA function, we performed yeast two-hybrid screens to search for torsinA-interacting proteins in the brain. We report here the isolation and characterization of a novel protein named printor (protein interactor of torsinA) that interacts selectively with wild-type (WT) torsinA but not the dystonia-associated torsinA ΔE mutant. Our data suggest that printor may serve as a cofactor of torsinA and provide a new molecular target for understanding and treating dystonia.  相似文献   

2.
3.
4.
TDP-43 is a nuclear protein involved in exon skipping and alternative splicing. Recently, TDP-43 has been identified as the pathological signature protein in frontotemporal lobar degeneration with ubiquitin-positive inclusions and in amyotrophic lateral sclerosis. In addition, TDP-43-positive inclusions are present in Parkinson disease, dementia with Lewy bodies, and 30% of Alzheimer disease cases. Pathological TDP-43 is redistributed from the nucleus to the cytoplasm, where it accumulates. An ∼25-kDa C-terminal fragment of TDP-43 accumulates in affected brain regions, suggesting that it may be involved in the disease pathogenesis. Here, we show that overexpression of the 25-kDa C-terminal fragment is sufficient to cause the mislocalization and cytoplasmic accumulation of endogenous full-length TDP-43 in two different cell lines, thus recapitulating a key biochemical characteristic of TDP-43 proteinopathies. We also found that TDP-43 mislocalization is associated with a reduction in the low molecular mass neurofilament mRNA levels. Notably, we show that the autophagic system plays a role in TDP-43 metabolism. Specifically, we found that autophagy inhibition increases the accumulation of the C-terminal fragments of TDP-43, whereas inhibition of mTOR, a key protein kinase involved in autophagy regulation, reduces the 25-kDa C-terminal fragment accumulation and restores TDP-43 localization. Our results suggest that autophagy induction may be a valid therapeutic target for TDP-43 proteinopathies.TDP-43 (transactive response DNA-binding protein 43) is a conserved and ubiquitously expressed nuclear protein with a theoretical molecular mass of ∼44 kDa. It is encoded by the TARDBP gene on chromosome 1, which is made of six exons that can be alternatively spliced to yield 11 different isoforms, with the mRNA encoding TDP-43 being the major species (1). Functionally, TDP-43 appears to be involved in exon skipping and alternative splicing (2, 3), and it has also been shown to link different types of nuclear bodies (4). Structural studies have confirmed the presence of two RNA recognition motifs (RRM1 and RRM2) and a glycine-rich C-terminal tail, which is thought to mediate protein-protein interaction (5).Recently, TDP-43 has been shown to be the major pathological protein in a wide range of disorders referred to as TDP-43 proteinopathies (68). These include frontotemporal lobar degeneration with ubiquitin-positive inclusions (FTLD-U),2 motor neuron disease, and amyotrophic lateral sclerosis (ALS). These last two disorders have been directly linked to mutations in TDP-43 (9, 10). In addition, TDP-43-positive inclusions are present in Parkinson disease, dementia with Lewy bodies, and 30% of Alzheimer disease cases (1114). Sporadic and familial forms of FTLD-U and ALS are characterized by cytoplasmic accumulation of insoluble, hyperphosphorylated, ubiquitinated, and proteolytically cleaved C-terminal fragments in affected brain and spinal cord regions. The cytoplasmic accumulation of TDP-43 is associated with a depletion of nuclear TDP-43 (8, 1521). These data suggest that some of these TDP-43 proteinopathies may share common mechanisms of pathogenesis.FTLD-U is caused by loss-of-function mutations in the progranulin gene, which lead, by an unknown mechanism, to the accumulation of cytoplasmic TDP-43 inclusions (22, 23). Notably, the TDP-43 inclusions in the ALS and FTLD-U brains are enriched with TDP-43 C-terminal fragments (8, 19). It has been suggested that the C-terminal fragments can be obtained by caspase-dependent cleavage of the full-length protein (24). However, it remains to be established if these fragments play a role in the disease pathogenesis.TDP-43 proteinopathies are characterized by the accumulation of abnormally modified TDP-43, suggesting that dysfunction in the intracellular quality control systems (ubiquitin-proteasome system and the autophagy-lysosome system) may be involved in the disease pathogenesis. The autophagic system is a conserved intracellular system designed for the degradation of long-lived proteins and organelles in lysosomes (25, 26). Three types of autophagy have been described: macroautophagy, microautophagy, and chaperon-mediated autophagy. Whereas macroautophagy and microautophagy involve the “in bulk” degradation of regions of the cytosol (27, 28), chaperon-mediated autophagy is a more selective pathway, and only proteins with a lysosomal targeting sequence are degraded (29). Cumulative evidence has suggested that an age-dependent decrease in the autophagy-lysosome system may account for the accumulation of abnormal proteins during aging (30, 31).Macroautophagy is induced when an isolation membrane is formed surrounding cytosolic components, forming an autophagic vacuole, which will eventually fuse with lysosomes for protein/organelle degradation. Induction of the isolation membrane is negatively regulated by mTOR (mammalian target of rapamycin) (32). It has been shown that increasing autophagy activation by mTOR inhibitors has beneficial effects in neurodegeneration (3335).  相似文献   

5.
Nek2 (for NIMA-related kinase 2) is a mammalian cell cycle–regulated kinase structurally related to the mitotic regulator NIMA of Aspergillus nidulans. In human cells, Nek2 associates with centrosomes, and overexpression of active Nek2 has drastic consequences for centrosome structure. Here, we describe the molecular characterization of a novel human centrosomal protein, C-Nap1 (for centrosomal Nek2-associated protein 1), first identified as a Nek2-interacting protein in a yeast two-hybrid screen. Antibodies raised against recombinant C-Nap1 produced strong labeling of centrosomes by immunofluorescence, and immunoelectron microscopy revealed that C-Nap1 is associated specifically with the proximal ends of both mother and daughter centrioles. On Western blots, anti–C-Nap1 antibodies recognized a large protein (>250 kD) that was highly enriched in centrosome preparations. Sequencing of overlapping cDNAs showed that C-Nap1 has a calculated molecular mass of 281 kD and comprises extended domains of predicted coiled-coil structure. Whereas C-Nap1 was concentrated at centrosomes in all interphase cells, immunoreactivity at mitotic spindle poles was strongly diminished. Finally, the COOH-terminal domain of C-Nap1 could readily be phosphorylated by Nek2 in vitro, as well as after coexpression of the two proteins in vivo. Based on these findings, we propose a model implicating both Nek2 and C-Nap1 in the regulation of centriole–centriole cohesion during the cell cycle.The serine/threonine kinase NIMA of Aspergillus nidulans is considered the founding member of a family of protein kinases with a possible role in cell cycle regulation (for reviews see Fry and Nigg, 1995; Lu and Hunter, 1995a ; Osmani and Ye, 1996). In A. nidulans, NIMA clearly cooperates with the Cdc2 protein kinase to promote progression into mitosis (Osmani et al., 1991), and overexpression of NIMA in a variety of heterologous species promotes a premature onset of chromosome condensation (O''Connell et al., 1994; Lu and Hunter, 1995b ). This has been interpreted to suggest evolutionary conservation of a pathway involving NIMA-related kinases (for review see Lu and Hunter, 1995a ). Indeed, kinases structurally related to NIMA are present in many species (Fry and Nigg, 1997). However, the only bona fide functional homologue of NIMA so far isolated stems from another filamentous fungus, Neurospora crassa (Pu et al., 1995), and the functional relationship between vertebrate NIMA-related kinases and fungal NIMA remains uncertain.The closest known mammalian relative to NIMA is a kinase termed Nek2 (for NIMA-related kinase 2)1 (Fry and Nigg, 1997). This kinase undergoes cell cycle–dependent changes in abundance and activity, reminiscent of NIMA (Schultz et al., 1994; Fry et al., 1995). It is highly expressed in male germ cells (Rhee and Wolgemuth, 1997; Tanaka et al., 1997), and data have been reported consistent with a role for Nek2 in meiotic chromosome condensation (Rhee and Wolgemuth, 1997). However, overexpression of active Nek2 in somatic cells has no obvious effect on chromosome condensation; instead, it induces striking alterations in the structure of the centrosome, the principal microtubule-organizing center of mammalian cells (Fry et al., 1998). Furthermore, immunofluorescence microscopy and subcellular fractionation concur to demonstrate that endogenous Nek2 associates with centrosomes, strongly suggesting that one physiological function of this kinase may relate to the centrosome cycle (Fry et al., 1998).The mammalian centrosome is an organelle of about 1 μm in diameter. It comprises two barrel-shaped centrioles that are made of nine short triplet microtubules and are surrounded by an amorphous matrix known as the pericentriolar material (PCM) (for review see Brinkley, 1985; Vorobjev and Nadehzdina, 1987; Kimble and Kuriyama, 1992; Kalt and Schliwa, 1993; Kellogg et al., 1994; Lange and Gull, 1996). Major progress has recently been made with the demonstration that microtubules are nucleated from γ-tubulin–containing ring complexes (γ-TuRCs), which are concentrated within the PCM (Moritz et al., 1995; Zheng et al., 1995). γ-Tubulin forms complexes with Spc97/98, two evolutionarily conserved proteins first identified in budding yeast spindle pole bodies (Geissler et al., 1996; Knop et al., 1997; Stearns and Winey, 1997), and there is also evidence for an important role of pericentrin and other coiled-coil proteins in organizing γ-TuRCs into higher order lattice structures (Doxsey et al., 1994; Dictenberg et al., 1998). However, in spite of this recent progress, it is clear that the inventory of centrosome components is far from complete.Centrosome structure and function is regulated in a cell cycle–dependent manner (for reviews see Mazia, 1987; Kellogg et al., 1994; Tournier and Bornens, 1994). Once in every cell cycle, and beginning around the G1/S transition, centrioles are duplicated (e.g., Kuriyama and Borisy, 1981a ; Vorobjev and Chentsov, 1982; Kochanski and Borisy, 1990; Chrétien et al., 1997). Late in G2, centrosomes then grow in size (a process referred to as maturation) through the recruitment of additional PCM proteins (Rieder and Borisy, 1982; Kalt and Schliwa, 1993; Lange and Gull, 1995). At the G2/M transition, the duplicated centrosomes separate and migrate to opposite ends of the nucleus. Concomitantly, their microtubule-nucleating activities increase dramatically in preparation for spindle formation (McGill and Brinkley, 1975; Snyder and McIntosh, 1975; Gould and Borisy, 1977; Kuriyama and Borisy, 1981b ; for reviews see Brinkley, 1985; Vorobjev and Nadehzdina, 1987; Karsenti, 1991). By what mechanisms these events are controlled remains largely unknown, but data obtained using phosphoepitope-specific antibodies strongly suggest that phosphorylation of centrosomal proteins plays a major role (Vandré et al., 1984, 1986; Centonze and Borisy, 1990). More direct support for this view stems from the observation that cyclin-dependent kinases (CDKs) enhance the microtubule-nucleation activity of centrosomes at the G2/M transition (Verde et al., 1990, 1992; Buendia et al., 1992) and are involved in promoting centrosome separation (Blangy et al., 1995; Sawin and Mitchison, 1995). Similarly, polo-like kinase 1, a cell cycle regulatory kinase structurally distinct from CDKs, has recently been implicated in centrosome maturation (Lane and Nigg, 1996).The precise role of Nek2 at the centrosome remains to be determined, but it is intriguing that overexpression of this kinase in human cells causes a pronounced splitting of centrosomes. This led us to propose that Nek2-dependent phosphorylation of previously unidentified proteins may cause a loss of centriole–centriole cohesion, and that this event might represent an early step in centrosome separation at the G2/M transition (Fry et al., 1998). With the aim of identifying potential substrates (or regulators) of Nek2, we have now performed a yeast two-hybrid screen, using full-length Nek2 as a bait. We report here the molecular characterization of a novel coiled-coil protein that we call C-Nap1 (for centrosomal Nek2-associated protein 1). C-Nap1 represents a core component of the mammalian centrosome and the first candidate substrate for a member of the NIMA protein kinase family to be identified.  相似文献   

6.
We purified from rat brain a novel actin filament (F-actin)–binding protein of ∼180 kD (p180), which was specifically expressed in neural tissue. We named p180 neurabin (neural tissue–specific F-actin– binding protein). We moreover cloned the cDNA of neurabin from a rat brain cDNA library and characterized native and recombinant proteins. Neurabin was a protein of 1,095 amino acids with a calculated molecular mass of 122,729. Neurabin had one F-actin–binding domain at the NH2-terminal region, one PSD-95, DlgA, ZO-1–like domain at the middle region, a domain known to interact with transmembrane proteins, and domains predicted to form coiled-coil structures at the COOH-terminal region. Neurabin bound along the sides of F-actin and showed F-actin–cross-linking activity. Immunofluorescence microscopic analysis revealed that neurabin was highly concentrated in the synapse of the developed neurons. Neurabin was also concentrated in the lamellipodia of the growth cone during the development of neurons. Moreover, a study on suppression of endogenous neurabin in primary cultured rat hippocampal neurons by treatment with an antisense oligonucleotide showed that neurabin was involved in the neurite formation. Neurabin is a candidate for key molecules in the synapse formation and function.During the development of the nervous system, the distal tip of the elongating axon—the growth cone—actively migrates toward its target cell in response to the combined actions of attractive and repulsive guidance molecules in the extracellular environment (Garrity and Zipursky, 1995; Keynes and Cook, 1995; Chiba and Keshishian, 1996; Culotti and Kolodkin, 1996; Friedman and O''Leary, 1996; Tessier-Lavigne and Goodman, 1996). When the growth cone contacts with the target cell, it is transformed into the functional presynaptic terminal (Garrity and Zipursky, 1995; Chiba and Kishishian, 1996). The actin cytoskeleton has been shown to play crucial roles in these processes of the synapse formation (Mitchison and Kirschner, 1988; Smith, 1988; Bentley and O''Connor, 1994; Lin et al., 1994; Mackay et al., 1995; Tanaka and Sabry, 1995).In the developing nervous system, the actin cytoskeleton is prominent in two structural domains of the growth cone, filopodia and lamellipodia (Mitchison and Kirschner, 1988; Smith, 1988; Bentley and O''Connor, 1994; Lin et al., 1994; Mackay et al., 1995; Tanaka and Sabry, 1995). In these domains, actin filament (F-actin)1 assembled at the leading edge are transported into the center of the growth cone and disassembled there. It has been suggested that this retrograde flow of F-actin is crucial for the growth cone motility. Drugs that disrupt F-actin have also been shown to cause the lamellipodial and filopodial collapse and block the ability of neurons to extend the growth cone in the correct direction (Marsh and Letourneau, 1984; Forscher and Smith, 1988; Bentley and Toroian-Raymond, 1986; Chien et al., 1993). These results suggest that the actin cytoskeleton regulates not only the growth cone motility but also the growth cone directionality. Recently, a variety of guidance molecules and their receptors have been identified (Garrity and Zipursky, 1995; Keynes and Cook, 1995; Chiba and Keshishian, 1996; Culotti and Kolodkin, 1996; Friedman and O''Leary, 1996; Tessier-Lavigne and Goodman, 1996). However, which molecules of the actin cytoskeleton are essential for the growth cone motility and directionality is not well understood.When the growth cone contacts with the target cell, the target cell regulates the development of the presynaptic nerve terminal and the formation of the functional synapse (Bowe and Fallon, 1995; Chiba and Keshishian, 1996). In the established nervous system, the presynaptic and postsynaptic membranes get aligned in space and constitute the synaptic junction (Burns and Augustine, 1995; Garner and Kindler, 1996). Electron microscopic studies have revealed the ultrastructural features of the synaptic junction (Burns and Augustine, 1995; Garner and Kindler, 1996). The presynaptic cytoplasm is characterized by synaptic vesicles (SVs). SVs are not distributed uniformly; SVs cluster together in the vicinity of the presynaptic plasma membrane, where F-actin forms a network and is associated with the presynaptic plasma membrane (Hirokawa et al., 1989). Most SVs within the cluster are linked through thin strands to each other, to F-actin, or to both (Hirokawa et al., 1989). A subset of SVs within the cluster are attached by fine filamentous threads to neurotransmitter release zone at the presynaptic plasma membrane (Hirokawa et al., 1989). The presynaptic submembranous cytoskeleton is assumed to be involved in recruiting Ca2+ channels and the components of the SV fusion complex, delivering SVs to the neurotransmitter release zone, and keeping them in place (Burns and Augustine, 1995; Garner and Kindler, 1996). At the inner surface of the post-synaptic plasma membrane, there is an electron dense thickening, called postsynaptic density. The postsynaptic density is assumed to be involved in the selective targeting and accumulation of ion channels and receptors (Burns and Augustine, 1995; Garner and Kindler, 1996). It is also assumed that the presynaptic and postsynaptic submembranous cytoskeleton elements are linked to cell adhesion molecules to regulate the synaptic stabilization and plasticity (Fields and Itoh, 1996; Garner and Kindler, 1996). The presynaptic and postsynaptic submembranous cytoskeleton elements are thought to be composed of spectrin/fodrin, ankyrin, α-adducin, and protein 4.1 isoforms and to be linked to F-actin through these cytoskeleton proteins (Garner and Kindler, 1996). However, little is known about which molecules of the submembranous cytoskeleton are essential for the synaptic transmission and/or the synaptic stabilization.To understand the regulation of the actin cytoskeleton during and after the development of the nervous system, it is of crucial importance to identify F-actin–binding proteins implicated in the synapse formation and function. Therefore, we attempted here to isolate neural tissue–specific F-actin–binding proteins. We isolated a novel neural tissue–specific F-actin–binding protein from rat brain, which may be involved in neurite formation, and named it neurabin (neural tissue–specific F-actin–binding protein).  相似文献   

7.
8.
Although PTIP is implicated in the DNA damage response, through interactions with 53BP1, the function of PTIP in the DNA damage response remain elusive. Here, we show that RNF8 controls DNA damage-induced nuclear foci formation of PTIP, which in turn regulates 53BP1 localization to the DNA damage sites. In addition, SMC1, a substrate of ATM, could not be phosphorylated at the DNA damage sites in the absence of PTIP. The PTIP-dependent pathway is important for DNA double strand breaks repair and DNA damage-induced intra-S phase checkpoint activation. Taken together, these results suggest that the role of PTIP in the DNA damage response is downstream of RNF8 and upstream of 53BP1. Thus, PTIP regulates 53BP1-dependent signaling pathway following DNA damage.The DNA damage response pathways are signal transduction pathways with DNA damage sensors, mediators, and effectors, which are essential for maintaining genomic stability (13). Following DNA double strand breaks, histone H2AX at the DNA damage sites is rapidly phosphorylated by ATM/ATR/DNAPK (410), a family homologous to phosphoinositide 3-kinases (11, 12). Subsequently, phospho-H2AX (γH2AX) provides the platform for accumulation of a larger group of DNA damage response factors, such as MDC1, BRCA1, 53BP1, and the MRE11·RAD50·NBS1 complex (13, 14), at the DNA damage sites. Translocalization of these proteins to the DNA double strand breaks (DSBs)3 facilitates DNA damage checkpoint activation and enhances the efficiency of DNA damage repair (14, 15).Recently, PTIP (Pax2 transactivation domain-interacting protein, or Paxip) has been identified as a DNA damage response protein and is required for cell survival when exposed to ionizing radiation (IR) (1, 1618). PTIP is a 1069-amino acid nuclear protein and has been originally identified in a yeast two-hybrid screening as a partner of Pax2 (19). Genetic deletion of the PTIP gene in mice leads to early embryonic lethality at embryonic day 8.5, suggesting that PTIP is essential for early embryonic development (20). Structurally, PTIP contains six tandem BRCT (BRCA1 carboxyl-terminal) domains (1618, 21). The BRCT domain is a phospho-group binding domain that mediates protein-protein interactions (17, 22, 23). Interestingly, the BRCT domain has been found in a large number of proteins involved in the cellular response to DNA damages, such as BRCA1, MDC1, and 53BP1 (7, 2429). Like other BRCT domain-containing proteins, upon exposure to IR, PTIP forms nuclear foci at the DSBs, which is dependent on its BRCT domains (1618). By protein affinity purification, PTIP has been found in two large complexes. One includes the histone H3K4 methyltransferase ALR and its associated cofactors, the other contains DNA damage response proteins, including 53BP1 and SMC1 (30, 31). Further experiments have revealed that DNA damage enhances the interaction between PTIP and 53BP1 (18, 31).To elucidate the DNA damage response pathways, we have examined the upstream and downstream partners of PTIP. Here, we report that PTIP is downstream of RNF8 and upstream of 53BP1 in response to DNA damage. Moreover, PTIP and 53BP1 are required for the phospho-ATM association with the chromatin, which phosphorylates SMC1 at the DSBs. This PTIP-dependent pathway is involved in DSBs repair.  相似文献   

9.
Fungal Zymosan and Mannan Activate the Cryopyrin Inflammasome   总被引:1,自引:0,他引:1  
Some fungal species are opportunistic pathogens that can cause infection in people with compromised immune systems. Activation of caspase-1 and the subsequent secretion of mature interleukin (IL)-1β is a major signaling pathway of the innate immune system, but how yeasts induce caspase-1 activation is unknown. We show here that stimulation of macrophages and dendritic cells with heat-killed Saccharomyces cerevisiae or the purified cell wall components zymosan and mannan induced caspase-1 activation and IL-1β secretion when combined with ATP. Macrophages deficient for the inflammasome adaptor ASC were defective in caspase-1 activation and IL-1β secretion, suggesting involvement of an ASC-dependent inflammasome. Indeed, caspase-1 activation was abrogated in macrophages lacking the NOD-like (NLR) protein Cryopyrin/Nalp3 and in wild type macrophages pretreated with the pannexin-1 inhibitor probenecid. IL-1β secretion further required the Toll-like receptor (TLR) adaptors MyD88 and TRIF, and partially relied on TLR2. We previously showed that bacterial molecules such as lipopolysaccharide (LPS) and peptidoglycan induce activation of caspase-7 through the Cryopyrin inflammasome. Similarly, Cryopyrin and ASC were required for activation of caspase-7 in macrophages stimulated with zymosan or mannan and ATP. These results demonstrate that the conserved fungal components zymosan and mannan require ASC and Cryopyrin for caspase-1 activation and IL-1β secretion and suggest an important role for the Cryopyrin inflammasome during fungal infections.Pathogen recognition by the innate immune system relies on a limited number of fixed germline-encoded receptors, which have evolved to identify so-called pathogen-associated molecular patterns (PAMPs),2 conserved microbial structures not shared by the host and essential for their survival (1). Examples of PAMPs are LPS from Gram-negative bacteria, peptidoglycan (PGN) from Gram-positive bacteria, and zymosan and mannan from fungi. Several structurally and functionally diverse classes of pattern-recognition receptors (PRRs) exist that induce various host defense pathways, including the Toll-like receptors (TLRs) located in the plasma membrane and intracellular organelles and the more recently identified intracellular family of NOD-like receptors (NLRs) (2).Previous studies have shown that gain-of-function mutations within the NLR protein Cryopyrin/NALP3 are associated with three autoinflammatory disorders characterized by skin rashes and prolonged episodes of fever in the absence of any apparent infection (3, 4). These hereditary periodic fever syndromes are Muckle-Wells syndrome (MWS), familial cold autoinflammatory syndrome (FACS), and neonatal-onset multisystem inflammatory disease (NOMID), and they are collectively referred to as the Cryopyrin/NALP3-associated periodic syndromes (CAPS). Subsequent studies revealed that the Cryopyrin/Nalp3 plays a crucial role in the assembly of a large (700 kDa) cytosolic protein complex termed the “inflammasome” (57). The bipartite adaptor protein ASC bridges the interaction between Cryopyrin/Nalp3 and caspase-1 in the inflammasome; thus allowing the recruitment and autoproteolytic activation of the cysteine protease (2). Activated caspase-1 subsequently mediates the maturation and secretion of the pro-inflammatory cytokines interleukin (IL)-1β and IL-18 (810). Interestingly, the Cryopyrin/Nalp3 inflammasome mediates caspase-1 activation in response to a variety of bacterial PAMPs such as LPS and PGN when combined with a second stimulus such as the P2X7 receptor ligand ATP (1114). Cryopyrin/Nalp3 also mediates caspase-1 activation and IL-1β secretion in macrophages stimulated with viral RNA and ATP (15) or exposed to crystalline substances including uric acid, silica and asbestos (1618). In contrast, the related NLR protein Ipaf is required for caspase-1 activation in macrophages infected with the intracellular pathogens Salmonella, Legionella, and Shigella (1921).Although the roles of specific inflammasomes in response to bacterial and viral PAMPs have been described, the inflammasome complexes that recognize fungal PAMPs to induce caspase-1 activation and IL-1β secretion are unknown. Here we show that heat-killed Saccharomyces cerevisiae and the purified cell wall components zymosan and mannan induced caspase-1 activation and IL-1β secretion from macrophages and dendritic cells upon co-stimulation with ATP. Macrophages deficient for the inflammasome adaptor ASC or the NLR protein Cryopyrin/Nalp3 were defective in zymosan- and mannan-induced caspase-1 activation and IL-1β secretion, whereas TNF-α secretion remained unaffected. Although macrophages lacking the TLR adaptors MyD88 or TRIF still activated caspase-1, zymosan- and mannan-induced secretion of IL-1β was significantly hampered. These results demonstrate that the conserved fungal cell wall components zymosan and mannan require ASC and Cryopyrin for caspase-1 activation and IL-1β secretion and suggest an important role for the Cryopyrin inflammasome during fungal infections.  相似文献   

10.
Detection of endogenous ubiquitination sites by mass spectrometry has dramatically improved with the commercialization of anti-di-glycine remnant (K-ε-GG) antibodies. Here, we describe a number of improvements to the K-ε-GG enrichment workflow, including optimized antibody and peptide input requirements, antibody cross-linking, and improved off-line fractionation prior to enrichment. This refined and practical workflow enables routine identification and quantification of ∼20,000 distinct endogenous ubiquitination sites in a single SILAC experiment using moderate amounts of protein input.The commercialization of antibodies that recognize lysine residues modified with a di-glycine remnant (K-ε-GG)1 has significantly transformed the detection of endogenous protein ubiquitination sites by mass spectrometry (15). Prior to the development of these highly specific reagents, proteomics experiments were limited to identification of up to only several hundred ubiquitination sites, which severely limited the scope of global ubiquitination studies (6). Recent proteomic studies employing anti-K-ε-GG antibodies have enhanced our understanding of ubiquitin biology through the identification of thousands of ubiquitination sites and the analysis of the change in relative abundance of these sites after chemical or biological perturbation (13, 5, 7). Use of stable isotope labeling by amino acids in cell culture (SILAC) for quantification has enabled researchers to better understand the extent of ubiquitin regulation upon proteasome inhibition and precisely identify those protein classes, such as newly synthesized proteins or chromatin-related proteins, that see overt changes in their ubiquitination levels upon drug treatment (2, 3, 5). Emanuel et al. (1) have combined genetic and proteomics assays implementing the anti-K-ε-GG antibody to identify hundreds of known and putative Cullin-RING ligase substrates, which has clearly demonstrated the extensive role of Cullin-RING ligase ubiquitination on cellular protein regulation.Despite the successes recently achieved with the use of the anti-K-ε-GG antibody, increased sample input (up to ∼35 mg) and/or the completion of numerous experimental replicates have been necessary to achieve large numbers of K-ε-GG sites (>5,000) in a single SILAC-based experiment (13, 5). For example, it has been recently shown that detection of more than 20,000 unique ubiquitination sites is possible from the analysis of five different murine tissues (8). However, as the authors indicate, only a few thousands sites are detected in any single analysis of an individual tissue sample (8). It is recognized that there is need for further improvements in global ubiquitin technology to increase the depth-of-coverage attainable in quantitative proteomic experiments using moderate amounts of protein input (9). Through systematic study and optimization of key pre-analytical variables in the preparation and use of the anti-K-ε-GG antibody as well as the proteomic workflow, we have now achieved, for the first time, routine quantification of ∼20,000 nonredundant K-ε-GG sites in a single SILAC triple encoded experiment starting with 5 mg of protein per SILAC channel. This represents a 10-fold improvement over our previously published method (3).  相似文献   

11.
A reversibly glycosylated polypeptide from pea (Pisum sativum) is thought to have a role in the biosynthesis of hemicellulosic polysaccharides. We have investigated this hypothesis by isolating a cDNA clone encoding a homolog of Arabidopsis thaliana, Reversibly Glycosylated Polypeptide-1 (AtRGP1), and preparing antibodies against the protein encoded by this gene. Polyclonal antibodies detect homologs in both dicot and monocot species. The patterns of expression and intracellular localization of the protein were examined. AtRGP1 protein and RNA concentration are highest in roots and suspension-cultured cells. Localization of the protein shows it to be mostly soluble but also peripherally associated with membranes. We confirmed that AtRGP1 produced in Escherichia coli could be reversibly glycosylated using UDP-glucose and UDP-galactose as substrates. Possible sites for UDP-sugar binding and glycosylation are discussed. Our results are consistent with a role for this reversibly glycosylated polypeptide in cell wall biosynthesis, although its precise role is still unknown.The primary cell wall of dicot plants is laid down by young cells prior to the cessation of elongation and secondary wall deposition. Making up to 90% of the cell''s dry weight, the extracellular matrix is important for many processes, including morphogenesis, growth, disease resistance, recognition, signaling, digestibility, nutrition, and decay. The composition of the cell wall has been extensively described (Bacic et al., 1988; Levy and Staehelin, 1992; Zablackis et al., 1995), and yet many questions remain unanswered regarding the synthesis and interaction of these components to provide cells with a functional wall (Carpita and Gibeaut, 1993; Carpita et al., 1996).Heteropolysaccharide biosynthesis can be divided into four steps: (a) chain or backbone initiation, (b) elongation, (c) side-chain addition, and (d) termination and extracellular deposition (Waldron and Brett, 1985). The similarity between various polysaccharide backbones leads to the prediction that the synthesizing machinery would be conserved between them. For example, the backbone of xyloglucan polymers, β-1,4 glucan, can be synthesized independently of or concurrently with side-chain addition (Campbell et al., 1988; White et al., 1993), and this polymer and the chains that make up cellulose are identical. The later addition of side chains to xyloglucan are catalyzed by specific transferases (Kleene and Berger, 1993) such as xylosyltransferase (Campbell et al., 1988), galactosyltransferase, and fucosyltransferase (Faïk et al., 1997), all of which are localized to the Golgi compartment (Brummell et al., 1990; Driouich et al., 1993; Staehelin and Moore, 1995).The enzymes involved in wall biosynthesis have been recalcitrant to isolation (Carpita et al., 1996; Albersheim et al., 1997). Only recently has the first gene encoding putative cellulose biosynthetic enzymes, celA, been isolated from cotton (Gossypium hirsutum) and rice (Oryza sativa; Pear et al., 1996).During studies of polysaccharide synthesis in pea (Pisum sativum) Golgi membranes, Dhugga et al. (1991) identified a 41-kD protein doublet that they suggested was involved in polysaccharide synthesis. The authors showed that this protein could be glycosylated by radiolabeled UDP-Glc but that this labeling could be reversibly competed with by unlabeled UDP-Glc, UDP-Xyl, and UDP-Gal, the sugars that make up xyloglucan (Hayashi, 1989). The 41-kD protein was named PsRGP1 (P. sativum Reversibly Glycosylated Polypeptide-1; Dhugga et al., 1997). Furthermore, the conditions that stimulate or inhibit Golgi-localized β-glucan synthase activity are the same conditions that stimulate or inhibit the glycosylation of PsRGP1 (Dhugga et al., 1991). To address the role of this protein in polysaccharide synthesis, the authors purified the polypeptides and obtained the sequences from tryptic peptides (Dhugga and Ray, 1994). Antibodies raised against PsRGP1 showed that it is soluble and localized to the plasma membrane (Dhugga et al., 1991) and Golgi compartment (Dhugga et al., 1997). In addition to its Golgi localization, the steady-state glycosylation of PsRGP1 is approximately 10:7:3 (UDP-Glc:-Xyl:-Gal), which is similar to the typical sugar composition of xyloglucan (1.0:0.75:0.25; Dhugga et al., 1997).We were interested in studying various aspects of cell wall metabolism, including the synthesis of polysaccharides and their delivery to the cell wall. Studies in pea have shown that a 41-kD protein may be involved in cell wall polysaccharide synthesis, possibly that of xyloglucan (Dhugga et al., 1997). Here we report the characterization of AtRGP1 (Arabidopsis thaliana Reversibly Glycosylated Polypeptide-1), a soluble protein that can also be found weakly associated with membrane fractions, most likely the Golgi fraction. The reversible nature of the glycosylation of this Arabidopsis homolog by the substrates used to make polysaccharides (nucleotide sugars) suggests a possible role for AtRGP1 in polysaccharide biosynthesis.  相似文献   

12.
13.
14.
15.
We outline several principles that we believe define the gating of two bacterial mechanosensitive channels, MscL and MscS. Serving as turgor regulators in bacteria and other walled cells, these molecules are tangible models for studying conformational transitions in membrane proteins driven directly by membrane tension. MscL, a compact pentamer, reversibly opens a gigantic 30-Å pore at near-lytic tensions. MscS, a heptameric complex, exhibits transient activation of a smaller pore at moderate tensions, thereby entering a tension-insensitive inactivated state. By comparing the structures and predicted transitions in these channels, we concluded that opening is commonly achieved through tilting and outward motion of the pore-lining helices, which is kinetically limited by hydration of the pore. The intricate adaptive behavior in MscS appears to depend on specific interhelical associations and the flexibility of the pore-lining helices. We discuss physical factors that may direct the transitions and stabilize main functional states in these channels.Osmotic forces are strong, which necessitated development of osmoregulation along with the first semipermeable membrane delineating the early cell. A simple estimation shows that a 1-μm cell behaving as an ideal osmometer would sustain a downshock no stronger than 20 mm, after which membrane tension would exceed the lytic limit of 10–12 dynes/cm. Thus, a cell without a reinforcing envelope or protective valves is very vulnerable. Free-living and enteric microorganisms cycling through the soil and experiencing drastic environmental changes developed robust mechanisms to maintain volume and integrity (1). The mechanosensitive channels MscS and MscL (mechanosensitive channels of small and large conductance, respectively) have been identified as primary osmolyte release valves limiting the turgor pressure under acute osmotic shock (24).Without mscS and mscL genes, Escherichia coli survives a 300 mosm osmotic downshock (2), its resistance attributed to the peptidoglycan layer partially restraining swelling. However, expression of either MscS or MscL allows cells to withstand a 700–800 mosm downshock through release of small osmolytes (2). Purification and reconstitution proved that MscL and MscS respond directly to tension in the lipid bilayer (57). Both channels reside in the inner (cytoplasmic) membrane (8), with MscL localized at the cell poles, bearing high curvature (9).As primary components of the turgor regulation system, E. coli MscS and MscL became convenient models for studies of tension-driven conformational transitions in membrane proteins (10). The crystal structures of closed-state Mycobacterium tuberculosis MscL (11) and E. coli MscS in two distinct conformations (12, 13) provided invaluable initial points to explore their gating mechanisms, in which computational methods play increasingly important roles.  相似文献   

16.
Proteolytic processing modifies the pleiotropic functions of many large, complex, and modular proteins and can generate cleavage products with new biological activity. The identification of exact proteolytic cleavage sites in the extracellular matrix laminins, fibronectin, and other extracellular matrix proteins is not only important for understanding protein turnover but is needed for the identification of new bioactive cleavage products. Several such products have recently been recognized that are suggested to play important cellular regulatory roles in processes, including angiogenesis. However, identifying multiple cleavage sites in extracellular matrix proteins and other large proteins is challenging as N-terminal Edman sequencing of multiple and often closely spaced cleavage fragments on SDS-PAGE gels is difficult, thus limiting throughput and coverage. We developed a new liquid chromatography-mass spectrometry approach we call amino-terminal oriented mass spectrometry of substrates (ATOMS) for the N-terminal identification of protein cleavage fragments in solution. ATOMS utilizes efficient and low cost dimethylation isotopic labeling of original N-terminal and proteolytically generated N termini of protein cleavage fragments followed by quantitative tandem mass spectrometry analysis. Being a peptide-centric approach, ATOMS is not dependent on the SDS-PAGE resolution limits for protein fragments of similar mass. We demonstrate that ATOMS reliably identifies multiple proteolytic sites per reaction in complex proteins. Fifty-five neutrophil elastase cleavage sites were identified in laminin-1 and fibronectin-1 with 34 more identified by matrix metalloproteinase cleavage. Hence, our degradomics approach offers a complimentary alternative to Edman sequencing with broad applicability in identifying N termini such as cleavage sites in complex high molecular weight extracellular matrix proteins after in vitro cleavage assays. ATOMS can therefore be useful in identifying new cleavage products of extracellular matrix proteins cleaved by proteases in pathology for bioactivity screening.Recently, considerable efforts have been deployed to develop high throughput proteomic screens to identify protease substrates in complex biological samples (18). Validation of substrates identified by these approaches or identification of cleavage sites by in vitro incubation of candidate substrates with the protease of interest is generally performed by SDS-PAGE analysis and Edman degradation and sequencing. However, the complexity of large modular proteins renders Edman sequencing of proteolytic fragments difficult to apply because each of the numerous proteolytic fragments should be analyzed separately, and high coverage of cleavage sites is rarely attained (9). Cleavage site identification after protein degradation is also very difficult for small peptide products less than 4 kDa. Consequently, the precise cleavage sites in complex extracellular matrix proteins such as laminin and fibronectin by important tissue and inflammatory cell proteases such as the matrix metalloproteinases (MMPs)1 and neutrophil elastase are mostly unknown.These limitations of Edman sequencing are problematic in the study of tissue remodeling and proteolysis in pathology. Neutrophil elastase and several MMPs such as MMP2, MMP8, and MMP9 play key roles in inflammation (10, 11), tissue healing (12, 13), and carcinogenesis (14, 15) and are well known for degrading extracellular matrix proteins (16). More recently, signaling functions for MMPs are increasingly recognized as one of their most important roles by the precise processing of cytokines or their binding proteins (17). In addition, several important examples are now known of cryptic binding sites being exposed after precise protein cleavage or new proteins termed neoproteins (18) being released upon limited cleavage of extracellular matrix proteins and having completely different functions compared with their parent molecule, including several with importance in angiogenesis (1925). Many such sites or neoproteins are generated by inflammatory proteases or proteases of the coagulation and fibrinolysis systems (24, 25), and this is a burgeoning field of discovery that is often hampered by difficulties in their N-terminal sequencing.In light of this limitation, we developed, validated, and used a new method for targeted and simultaneous N-terminal sequencing of one or a small number of protein N termini or cleavage products we call amino-terminal oriented mass spectrometry of substrates (ATOMS). We applied ATOMS for the analysis of cleavage sites generated in laminin-1 and fibronectin-1 by neutrophil elastase and neutrophil and tissue MMPs. Laminin-1 (LM-111), a trimeric glycoprotein composed of the α1, β1, and γ1 chains, is ubiquitously expressed in epithelium and endothelium. Proteolytic processing of laminins greatly affects cellular behavior and is also implicated in cancer cell migration (20, 2629). Another important extracellular matrix protein is plasma fibronectin (also known as fibronectin isoform 1) and its cellular isoforms, which are homodimers linked by a disulfide bridge at the C terminus (30) that are important for cell adhesion and intracellular signaling (3134). Fibronectin is susceptible to proteolysis (35, 36), which affects its biological functions (3739). However, the cleavage sites within these two molecules by inflammatory MMPs and neutrophil elastase are largely unknown. Here we identified a total of 55 neutrophil elastase cleavage sites in LM-111 and fibronectin-1 and 34 MMP cleavage sites, demonstrating the capacity of ATOMS to identify multiple N-terminal sequences in solution. ATOMS also outperformed N-terminal Edman sequencing with 50% more cleavage sites identified by ATOMS, representing a significant advance in N-terminal sequencing technology. The utility of the method is broadly applicable for the analysis of multiple cleavages in other very large molecules and so offers great potential to accurately identify and rapidly sequence multiple cryptic bioactive protein fragments liberated following proteolytic processing.  相似文献   

17.
CCN3 (NOV), a putative ligand for integrin receptors, is tightly associated with the extracellular matrix and mediates diverse cellular functions, including cell adhesion and proliferation. CCN3 has been shown to negatively regulate growth although it promotes migration in a cell type-specific manner. In this study, overexpression of CCN3 reduces growth and increases intercellular adhesion of breast cancer cells. Interestingly, CCN3 overexpression also led to the formation of multiple pseudopodia that are enriched in actin, CCN3, and vinculin. Breast cancer cells preincubated with exogenous CCN3 protein also induced the same phenotype, indicating that secreted CCN3 is sufficient to induce changes in cell morphology. Surprisingly, extracellular CCN3 is internalized to the early endosomes but not to the membrane protrusions, suggesting pseudopodia-enriched CCN3 may derive from a different source. The presence of an intracellular variant of CCN3 will be consistent with our finding that the cytoplasmic tail of the gap junction protein connexin43 (Cx43) associates with CCN3. Cx43 is a channel protein permitting intercellular communication to occur. However, neither the channel properties nor the protein levels of Cx43 are affected by the CCN3 protein. In contrast, CCN3 proteins are down-regulated in the absence of Cx43. Finally, we showed that overexpression of CCN3 increases the activity of the small GTPase Rac1, thereby revealing a pathway that links Cx43 directly to actin reorganization.The CCN (CYR61/Connective Tissue Growth Factor/Nephroblastoma Overexpressed) family of multimodular proteins mediates diverse cellular functions, including cell adhesion, migration, and proliferation (13). Overexpression of CCN3, one of the founding members of the family, inhibits proliferation in most types of tumors such as glioblastoma and Ewing sarcoma (4, 5). Similarly, down-regulation of CCN3 has been suggested to promote melanoma progression (6). On the other hand, CCN3 can also promote migration in sarcoma and glioblastoma (4, 7), although a separate study shows that it decreases the invasion of melanoma (6). Therefore, in contrast to its role in growth suppression, the role of CCN3 signaling in cell motility is less clear.Most evidence suggests CCN3 mediates its effects by binding to the integrin proteins, such as the αVβ3 receptors (8, 9), and that CCN3 alters cell adhesion in an integrin-dependent fashion (4, 10). In melanocytes, the discoidin domain receptor 1 mediates CCN3-dependent adhesion (11). CCN3 has also been observed to associate with Notch1 (12), fibulin 1C (13), S100A4 (14), and the gap junction protein Cx433 (15, 16), suggesting that CCN3 may also modulate cell growth via non-integrin signaling pathways.Gap junction proteins are best known for forming channels between cells, contributing to intercellular communication by allowing the exchange of small ions and molecules (17, 18). Consequently, attenuated intercellular communication has been implicated in promoting carcinogenesis (19, 20). Recent evidence has indicated that connexins can mediate channel-independent growth control through interaction of their C-terminal cytoplasmic tail with various intracellular signaling molecules (2123). In addition, many Cx43-interacting proteins, including ZO-1 (zonula occludens-1) (24), Drebrin (25), and N-cadherin (26) associate with F-actin, thus placing Cx43 in close proximity to the actin cytoskeleton.In this study, we show for the first time that CCN3 reorganizes the actin cytoskeleton of the breast cancer cells MDA-MB-231 with the formation of multiple cell protrusions, possibly by activating the small GTPase Rac1. Our results also suggest an alternative route by which Cx43 may be functionally linked to actin cytoskeletal signaling via CCN3.  相似文献   

18.
Genomic stability requires a functional Fanconi anemia (FA) pathway composed of an upstream “core complex” (FA proteins A/B/C/E/F/G/L/M) that mediates monoubiquitination of the downstream targets FANCD2 and FANCI. Unique among FA core complex members, FANCM has processing activities toward replication-associated DNA structures, suggesting a vital role for FANCM during replication. Using Xenopus egg extracts, we analyzed the functions of FANCM in replication and the DNA damage response. xFANCM binds chromatin in a replication-dependent manner and is phosphorylated in response to DNA damage structures. Chromatin binding and DNA damage-induced phosphorylation of xFANCM are mediated in part by the downstream FA pathway protein FANCD2. Moreover, phosphorylation and chromatin recruitment of FANCM is regulated by two mayor players in the DNA damage response: the cell cycle checkpoint kinases ATR and ATM. Our results indicate that functions of FANCM are controlled by FA- and non-FA pathways in the DNA damage response.Fanconi anemia is a genetic disease characterized by genomic instability and cancer predisposition. Cells from FA3 patients show hypersensitivity to DNA interstrand cross-links and have highly elevated chromosomal breakage rates, indicating a role for FA proteins in the cellular DNA damage response. The FA pathway consists of an upstream FA core complex containing at least eight proteins (FANCA, -B, -C, -E, -F, -G, -L, and -M) that is required for the DNA damage-induced monoubiquitination of two downstream proteins, FANCD2 and FANCI. Although the molecular function of the FA pathway is unknown, the identification of additional FA genes FANCD1 (BRCA2), FANCN (PALB2), and the DNA helicase FANCJ (BRIP1) as breast cancer (BRCA) susceptibility genes suggests convergence of the FA/BRCA pathway with a larger network of proteins involved in DNA repair (reviewed in Ref. 1).In addition to monoubiquitination by the FA core complex, FANCD2 and FANCI are phosphorylated by the two major cell cycle checkpoint kinases, ATM (ataxia telangiectasia mutated) and ATR (ATM and Rad3-related),y in response to DNA damage (26). ATM-dependent phosphorylation of FANCD2 occurs following ionizing irradiation and is required for activation of the ionizing irradiation-induced intra-S phase checkpoint (4). ATR-dependent phosphorylation of FANCD2 is triggered by various types of DNA damage, including replication stress, and is required for the interstrand cross-link-induced intra-S phase checkpoint response (2, 3). Moreover, phosphorylation by ATR is required for efficient FANCD2 monoubiquitination in response to DNA damage, suggesting that the FA pathway might participate in ATR-dependent coordination of the S phase of the cell cycle (3, 7).The recent identification of a highly conserved FA core complex member, FANCM (8, 9), indicates a direct role of FA pathway proteins in repair steps at sites of DNA damage. FANCM is a homolog of the archaebacterial Hef protein (helicase-associated endonuclease for fork-structured DNA) and contains two DNA processing domains: a DEAH box helicase domain and an XPF/ERCC4-like endonuclease domain. FANCM has ATP-dependent DNA translocase activity and can dissociate DNA triple helices in vitro (8). Moreover, FANCM binds Holliday junctions and DNA replication fork structures in vitro and promotes ATP-dependent branch point migration, suggesting that FANCM might be involved in DNA processing at stalled replication forks (10, 11). In human cells, FANCM localizes to chromatin and is required for chromatin recruitment of other FA core complex proteins (8, 12). FANCM is phosphorylated during both the M and S phases and in response to DNA-damaging agents (8, 12, 13). Interestingly, DNA damage-induced phosphorylation of FANCM is independent of the FA core complex (8), suggesting that FANCM is controlled by other, as yet unknown upstream components of the DNA damage response. Here, we used cell-free Xenopus egg extracts to investigate the role of FANCM during replication and in the DNA damage response. We show that Xenopus FANCM (xFANCM) binds chromatin in a replication-dependent manner and is phosphorylated during unperturbed replication as well as in response to various DNA damage structures. Both chromatin recruitment and phosphorylation of xFANCM are partially controlled by xFANCD2, suggesting feedback signaling from xFANCD2 to the upstream xFA core complex via regulation of xFANCM. In addition, chromatin recruitment during unperturbed replication and activation of xFANCM in response to DNA damage are controlled by the xATR and xATM cell cycle kinases.  相似文献   

19.
Group I metabotropic glutamate receptors (mGluRs) are coupled via phospholipase Cβ to the hydrolysis of phosphoinositides and function to modulate neuronal excitability and synaptic transmission at glutamatergic synapses. The desensitization of Group I mGluR signaling is thought to be mediated primarily via second messenger-dependent protein kinases and G protein-coupled receptor kinases. We show here that both mGluR1 and mGluR5 interact with the calcineurin inhibitor protein (CAIN). CAIN is co-immunoprecipitated in a complex with Group I mGluRs from both HEK 293 cells and mouse cortical brain lysates. Purified CAIN and its C-terminal domain specifically interact with glutathione S-transferase fusion proteins corresponding to the second intracellular loop and the distal C-terminal tail domains of mGluR1. The interaction of CAIN with mGluR1 could also be blocked using a Tat-tagged peptide corresponding to the mGluR1 second intracellular loop domain. Overexpression of full-length CAIN attenuates the agonist-stimulated endocytosis of both mGluR1a and mGluR5a in HEK 293 cells, but expression of the CAIN C-terminal domain does not alter mGluR5a internalization. In contrast, overexpression of either full-length CAIN or the CAIN C-terminal domain impairs agonist-stimulated inositol phosphate formation in HEK 293 cells expressing mGluR1a. This CAIN-mediated antagonism of mGluR1a signaling appears to involve the disruption of receptor-Gαq/11 complexes. Taken together, these observations suggest that the association of CAIN with intracellular domains involved in mGluR/G protein coupling provides an additional mechanism by which Group I mGluR endocytosis and signaling are regulated.Metabotropic glutamate receptors (mGluRs)2 play an essential role in regulating neuronal plasticity, development, and neurotoxicity and belong to the G protein-coupled receptor superfamily of integral membrane proteins (14). The mGluR family can be subclassified into three groups based on sequence homology, G protein specificity, and pharmacology. Group I mGluRs (mGluR1 and mGluR5) couple via the heterotrimeric Gαq/11 proteins to the activation of phospholipase Cβ, resulting in the formation of inositol 1,4,5-triphosphate and diacylglycerol, the release of Ca2+ from intracellular stores, and the activation of protein kinase C (PKC) (46).The regulation of mGluR signal transduction involves numerous proteins that function to regulate signaling at both the level of the heterotrimeric G protein and the receptor (68). At the level of the receptor, Group I mGluR activity is regulated by a process termed desensitization, which protects against both acute and chronic receptor overstimulation (9, 10). The attenuation of Group I mGluR signaling can be mediated by both phosphorylation-dependent and phosphorylation-independent processes (11). The phosphorylation-independent attenuation of Group I mGluR signaling is mediated by GRK2 (G protein-coupled receptor kinase 2), which is composed of three functional domains: an N-terminal RGS (regulator of G protein signaling) homology domain, a central catalytic domain, and a C-terminal Gβγ-binding pleckstrin homology domain (12). GRK2-mediated desensitization of Group I mGluRs does not require catalytic activity but rather requires the interaction of the GRK2 RGS homology domain with both the second intracellular loop domain of mGluR1 and the α-subunit of Gαq/11, thereby attenuating heterotrimeric G protein coupling (1315). Phosphorylation-independent desensitization of mGluR1 signaling is also mediated by optineurin, an effect that is enhanced by the expression of mutant huntingtin (16). Phosphorylation-dependent desensitization of Group I mGluR responsiveness involves the phosphorylation of PKC consensus sequence localized within the intracellular loop and C-terminal tail domains of mGluR1 and mGluR5 by PKC (17, 18). It is proposed that calcineurin and mGluR5 may exist in a signaling complex in the brain and that calcineurin may function to modulate mGluR5 signaling by directly dephosphorylating the receptor at a PKC consensus site that contributes to mGluR5 desensitization (19). Calcineurin is also linked to the regulation of endocytosis via its interaction with dynamin-1 (20).On the basis of the observation that calcineurin may form a complex with Group I mGluRs, we hypothesized that CAIN (calcineurin inhibitor protein) might also interact with Group I mGluRs and modulate their endocytosis and signaling. CAIN, also known as Cabin1 (calcineurin-binding protein), was first identified as a protein that binds to calcineurin and was shown to inhibit calcineurin catalytic activity (2123). Previous studies also demonstrated that CAIN may interact with amphiphysin-1, dynamin-1, and α-adaptin and led to the suggestion that CAIN functions as a component of synaptic endocytic complexes (24). Consistent with this hypothesis, the overexpression of CAIN in human embryonic kidney (HEK 293) cells resulted in attenuated transferrin receptor endocytosis.We show here that CAIN interacts with the second intracellular loop and C-terminal tail domains of Group I mGluRs, inhibits Group I mGluR internalization, and attenuates mGluR1a signaling by disrupting receptor-Gαq/11 complexes. Taken together, these results describe an additional mechanism by which Group I mGluR activity may be regulated.  相似文献   

20.
A 135-kD actin-bundling protein was purified from pollen tubes of lily (Lilium longiflorum) using its affinity to F-actin. From a crude extract of the pollen tubes, this protein was coprecipitated with exogenously added F-actin and then dissociated from F-actin by treating it with high-ionic-strength solution. The protein was further purified sequentially by chromatography on a hydroxylapatite column, a gel-filtration column, and a diethylaminoethyl-cellulose ion-exchange column. In the present study, this protein is tentatively referred to as P-135-ABP (Plant 135-kD Actin-Bundling Protein). By the elution position from a gel-filtration column, we estimated the native molecular mass of purified P-135-ABP to be 260 kD, indicating that it existed in a dimeric form under physiological conditions. This protein bound to and bundled F-actin prepared from chicken breast muscle in a Ca2+-independent manner. The binding of 135-P-ABP to actin was saturated at an approximate stoichiometry of 26 actin monomers to 1 dimer of P-135-ABP. By transmission electron microscopy of thin sections, we observed cross-bridges between F-actins with a longitudinal periodicity of 31 nm. Immunofluorescence microscopy using rhodamine-phalloidin and antibodies against the 135-kD polypeptide showed that P-135-ABP was colocalized with bundles of actin filaments in lily pollen tubes, leading us to conclude that it is the factor responsible for bundling the filaments.Actin filaments, one of the major components of the cytoskeleton, are organized into a highly ordered architecture and are involved in various kinds of cell motility. Their architecture is regulated by several kinds of actin-binding proteins, including cross-linking proteins, severing proteins, end-capping proteins, and monomer-sequestering proteins in animal, protozoan, and yeast cells (Stossel et al., 1985; Pollard and Cooper, 1986; Vandekerckhove and Vancompernolle, 1992). In plant cells the organization of the actin cytoskeleton also changes remarkably during the cell cycle or during developmental processes, and it is suggested that actin-binding proteins are involved in their dynamic change. However, little is known about actin-binding proteins in plant cells.Only a low-Mr actin-binding and -depolymerizing protein, profilin, in white birch (Betula verrucosa; Valenta et al., 1991), maize (Zea mays; Staiger et al., 1993; Ruhlandt et al., 1994), bean (Phaseolus vulgaris; Vidali et al., 1995), tobacco (Nicotiana tabacum; Mittermann et al., 1995), tomato (Lycopersicon esculentum; Darnowski et al., 1996), Arabidopsis (Arabidopsis thaliana; Huang et al., 1996), and lily (Lilium longiflorum; Vidali and Hepler, 1997), and an ADF in lily (Kim et al., 1993), rapeseed (Brassica napus; Kim et al., 1993), and maize (Rozycka et al., 1995; Lopez et al., 1996), have been identified by biochemical or molecular biological means.The native and recombinant forms of these proteins are capable of binding to animal or plant actin (Valenta et al., 1993; Giehl et al., 1994; Ruhlandt et al., 1994; Lopez et al., 1996; Perelroizen et al., 1996; Carlier et al., 1997). Plant profilin expressed in mammalian BHK-21 cells (Rothkegel et al., 1996) or profilin-deficient Dictyostelium discoideum cells (Karakesisoglou et al., 1996) was able to functionally substitute for endogenous profilin in these cells. The introduction of plant profilin into living stamen hair cells by microinjection caused the rapid reduction of the number of actin filaments (Staiger et al., 1994; Karakesisoglou et al., 1996; Ren et al., 1997). These results indicate that plant profilin and ADF share many functional similarities with other eukaryote profilins and ADFs.It is well known that the actin cytoskeleton undergoes dynamic changes in organization during hydration and activation of the vegetative cells of pollen grains (Pierson and Cresti, 1992). Before hydration actin filaments exist as fusiform or spiculate structures (a storage form), but they are rearranged to form a network upon hydration (Heslop-Harrison et al., 1986; Tiwari and Polito, 1988). In the growing pollen tube there are strands or bundles of actin filaments parallel to the long axis (Perdue et al., 1985; Pierson et al., 1986; Miller et al., 1996) that are involved in cytoplasmic streaming (Franke et al., 1972; Mascarenhas and Lafountain, 1972) and transport of vegetative nuclei and generative cells to the growing tip (Heslop-Harrison et al., 1988; Heslop-Harrison and Heslop-Harrison, 1989). Characterization of the function of actin-binding proteins is essential to understanding the regulation of actin organization during the developmental process of pollen. Since only a small number of vacuoles containing proteases develop in pollen grains and pollen tubes at a younger stage, pollen tubes are suitable materials for isolating and biochemically studying actin-binding proteins responsible for organizing actin filaments into various forms.In a previous paper we reported that several components in a crude extract prepared from lily pollen tubes, including a 170-kD myosin heavy chain and 175-, 135-, and 110-kD polypeptides, could be coprecipitated with exogenously added F-actin (Yokota and Shimmen, 1994). We also found that rhodamine-labeled F-actin was tightly bound to the glass surface treated with the fraction containing the 135- and 110-kD polypeptides (Yokota and Shimmen, 1994). These results suggested that either one or both of the 135- and 110-kD polypeptides possesses an F-actin-binding activity. In the present study, we purified the 135-kD polypeptide from lily pollen tubes by biochemical procedures and then characterized its F-actin-binding properties and distribution in the pollen tubes. This protein was able to bundle F-actin isolated from chicken breast muscle and colocalized with actin-filament bundles in pollen tubes. We refer to this protein as P-135-ABP (Plant 135-kD Actin-Bundling Protein).  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