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.
Fibroblast growth factor receptors (FGFRs) are important for many different mechanisms, including cell migration, proliferation, differentiation, and survival. Here, we show a new link between FGFR1 and the cell adhesion molecule neurofascin, which is important for neurite outgrowth. After overexpression in HEK293 cells, embryonal neurofascin isoform NF166 was able to associate with FGFR1, whereas the adult isoform NF186, differing from NF166 in additional extracellular sequences, was deficient. Pharmacological inhibitors and overexpression of dominant negative components of the FGFR signaling pathway pointed to the activation of FGFR1 after association with neurofascin in neurite outgrowth assays in chick tectal neurons and rat PC12-E2 cells. Both extra- and intracellular domains of embryonal neurofascin isoform NF166 were able to form complexes with FGFR1 independently. However, the cytosolic domain was both necessary and sufficient for the activation of FGFR1. Cytosolic serine residues 56 and 100 were shown to be essential for the neurite outgrowth-promoting activity of neurofascin, whereas both amino acid residues were dispensable for FGFR1 association. In conclusion, the data suggest a neurofascin intracellular domain, which activates FGFR1 for neurite outgrowth, whereas the extracellular domain functions as an additional, regulatory FGFR1 interaction domain in the course of development.The four known fibroblast growth factor receptors (FGFRs),2 which are targeted by a large family of 22 fibroblast growth factor ligands, represent a highly diverse signaling system important for migration, proliferation, differentiation, and survival of many different cell types (1, 2). fibroblast growth factor activation of FGFR leads to the activation of mitogen-activated protein kinase (MAPK), phosphatidylinositol 3-kinase (PI3K), and phospholipase Cγ (PLCγ), depending on the cellular system under study. Non-canonical FGFR interactions with NCAM, cadherins, and syndecan via extracellular domains were also described (1). However, the contribution of intracellular interactions of FGFR1 with further membrane co-receptors is poorly understood. Only cytosolic interaction between FGFRs and EphA4 have been described that are involved in mutual transphosphorylation (3).The cell adhesion molecule neurofascin is important for cell-cell communication in the nervous system (4, 5). Neurofascin regulates many different functions in the brain, suggesting that it functions as a key regulator for both developing and differentiated neural cells. Different alternatively spliced neurofascin isoforms are expressed in different cells and at different times of development (6). Embryonal neurofascin NF166 is important for neurite outgrowth and guidance (7, 8). Recently, a role for neurofascin NF166 for early processes of inhibitory synaptogenesis at the axon hillock and for the positioning of inhibitory synapses at the axon initial segment has been proven (9, 10).In the more developed nervous system, NF166 is replaced by NF186, which is inhibitory for neurite outgrowth (11). NF186 is linked to the cortical actin cytoskeleton via ankyrinG (12). Clustering of voltage-gated sodium channels both at axon initial segments and at the nodes of Ranvier is conferred by neurofascin NF186 (13, 14). A further cytosolic interaction partner is the PDZ molecule syntenin-1 (15).Despite the well known functional importance of neurofascin in the nervous system, corresponding signaling pathways have not been investigated. In contrast, signaling by the related molecules NCAM and L1 have been studied with regard to the induction of neurite outgrowth in greater detail (for a review, see Refs. 1618). Both NCAM and L1 induce neurite outgrowth through activation of FGFR1 (1923). NCAM may further undergo lateral interactions with PrP (prion precursor protein) or GFRα, which is part of the glia-derived neurotrophic factor receptor (24, 25). In addition to FGFR1 interaction, both L1 and NCAM are connected to non-receptor tyrosine kinases. However, whereas NCAM employs the non-receptor kinase c-Fyn as an upstream component, L1 is linked to c-Src (26, 27). L1 converges with NCAM signaling upstream of the MAPK pathway at the level of Raf (18, 21, 28, 29). NCAM may induce alternative signaling pathways, including protein kinase A-dependent signaling or G-proteins (18, 30). NCAM signaling to the nucleus may include activation of CREB and c-Fos or NF-κB (29, 31, 32).Here, we elucidate the molecular mechanisms of neurofascin-FGFR1 interaction for neurite outgrowth. We show that both cytosolic and the extracellular domains are important for the association of FGFR1 with neurofascin. Although the cytosolic domain represents a critical determinant for FGFR1 activation, the extracellular sequences of neurofascin act as a regulator for FGFR1-dependent signal transduction in the course of development.  相似文献   

2.
3.
A complete understanding of the biological functions of large signaling peptides (>4 kDa) requires comprehensive characterization of their amino acid sequences and post-translational modifications, which presents significant analytical challenges. In the past decade, there has been great success with mass spectrometry-based de novo sequencing of small neuropeptides. However, these approaches are less applicable to larger neuropeptides because of the inefficient fragmentation of peptides larger than 4 kDa and their lower endogenous abundance. The conventional proteomics approach focuses on large-scale determination of protein identities via database searching, lacking the ability for in-depth elucidation of individual amino acid residues. Here, we present a multifaceted MS approach for identification and characterization of large crustacean hyperglycemic hormone (CHH)-family neuropeptides, a class of peptide hormones that play central roles in the regulation of many important physiological processes of crustaceans. Six crustacean CHH-family neuropeptides (8–9.5 kDa), including two novel peptides with extensive disulfide linkages and PTMs, were fully sequenced without reference to genomic databases. High-definition de novo sequencing was achieved by a combination of bottom-up, off-line top-down, and on-line top-down tandem MS methods. Statistical evaluation indicated that these methods provided complementary information for sequence interpretation and increased the local identification confidence of each amino acid. Further investigations by MALDI imaging MS mapped the spatial distribution and colocalization patterns of various CHH-family neuropeptides in the neuroendocrine organs, revealing that two CHH-subfamilies are involved in distinct signaling pathways.Neuropeptides and hormones comprise a diverse class of signaling molecules involved in numerous essential physiological processes, including analgesia, reward, food intake, learning and memory (1). Disorders of the neurosecretory and neuroendocrine systems influence many pathological processes. For example, obesity results from failure of energy homeostasis in association with endocrine alterations (2, 3). Previous work from our lab used crustaceans as model organisms found that multiple neuropeptides were implicated in control of food intake, including RFamides, tachykinin related peptides, RYamides, and pyrokinins (46).Crustacean hyperglycemic hormone (CHH)1 family neuropeptides play a central role in energy homeostasis of crustaceans (717). Hyperglycemic response of the CHHs was first reported after injection of crude eyestalk extract in crustaceans. Based on their preprohormone organization, the CHH family can be grouped into two sub-families: subfamily-I containing CHH, and subfamily-II containing molt-inhibiting hormone (MIH) and mandibular organ-inhibiting hormone (MOIH). The preprohormones of the subfamily-I have a CHH precursor related peptide (CPRP) that is cleaved off during processing; and preprohormones of the subfamily-II lack the CPRP (9). Uncovering their physiological functions will provide new insights into neuroendocrine regulation of energy homeostasis.Characterization of CHH-family neuropeptides is challenging. They are comprised of more than 70 amino acids and often contain multiple post-translational modifications (PTMs) and complex disulfide bridge connections (7). In addition, physiological concentrations of these peptide hormones are typically below picomolar level, and most crustacean species do not have available genome and proteome databases to assist MS-based sequencing.MS-based neuropeptidomics provides a powerful tool for rapid discovery and analysis of a large number of endogenous peptides from the brain and the central nervous system. Our group and others have greatly expanded the peptidomes of many model organisms (3, 1833). For example, we have discovered more than 200 neuropeptides with several neuropeptide families consisting of as many as 20–40 members in a simple crustacean model system (5, 6, 2531, 34). However, a majority of these neuropeptides are small peptides with 5–15 amino acid residues long, leaving a gap of identifying larger signaling peptides from organisms without sequenced genome. The observed lack of larger size peptide hormones can be attributed to the lack of effective de novo sequencing strategies for neuropeptides larger than 4 kDa, which are inherently more difficult to fragment using conventional techniques (3437). Although classical proteomics studies examine larger proteins, these tools are limited to identification based on database searching with one or more peptides matching without complete amino acid sequence coverage (36, 38).Large populations of neuropeptides from 4–10 kDa exist in the nervous systems of both vertebrates and invertebrates (9, 39, 40). Understanding their functional roles requires sufficient molecular knowledge and a unique analytical approach. Therefore, developing effective and reliable methods for de novo sequencing of large neuropeptides at the individual amino acid residue level is an urgent gap to fill in neurobiology. In this study, we present a multifaceted MS strategy aimed at high-definition de novo sequencing and comprehensive characterization of the CHH-family neuropeptides in crustacean central nervous system. The high-definition de novo sequencing was achieved by a combination of three methods: (1) enzymatic digestion and LC-tandem mass spectrometry (MS/MS) bottom-up analysis to generate detailed sequences of proteolytic peptides; (2) off-line LC fractionation and subsequent top-down MS/MS to obtain high-quality fragmentation maps of intact peptides; and (3) on-line LC coupled to top-down MS/MS to allow rapid sequence analysis of low abundance peptides. Combining the three methods overcomes the limitations of each, and thus offers complementary and high-confidence determination of amino acid residues. We report the complete sequence analysis of six CHH-family neuropeptides including the discovery of two novel peptides. With the accurate molecular information, MALDI imaging and ion mobility MS were conducted for the first time to explore their anatomical distribution and biochemical properties.  相似文献   

4.
5.
Given the ease of whole genome sequencing with next-generation sequencers, structural and functional gene annotation is now purely based on automated prediction. However, errors in gene structure are frequent, the correct determination of start codons being one of the main concerns. Here, we combine protein N termini derivatization using (N-Succinimidyloxycarbonylmethyl)tris(2,4,6-trimethoxyphenyl)phosphonium bromide (TMPP Ac-OSu) as a labeling reagent with the COmbined FRActional DIagonal Chromatography (COFRADIC) sorting method to enrich labeled N-terminal peptides for mass spectrometry detection. Protein digestion was performed in parallel with three proteases to obtain a reliable automatic validation of protein N termini. The analysis of these N-terminal enriched fractions by high-resolution tandem mass spectrometry allowed the annotation refinement of 534 proteins of the model marine bacterium Roseobacter denitrificans OCh114. This study is especially efficient regarding mass spectrometry analytical time. From the 534 validated N termini, 480 confirmed existing gene annotations, 41 highlighted erroneous start codon annotations, five revealed totally new mis-annotated genes; the mass spectrometry data also suggested the existence of multiple start sites for eight different genes, a result that challenges the current view of protein translation initiation. Finally, we identified several proteins for which classical genome homology-driven annotation was inconsistent, questioning the validity of automatic annotation pipelines and emphasizing the need for complementary proteomic data. All data have been deposited to the ProteomeXchange with identifier PXD000337.Recent developments in mass spectrometry and bioinformatics have established proteomics as a common and powerful technique for identifying and quantifying proteins at a very broad scale, but also for characterizing their post-translational modifications and interaction networks (1, 2). In addition to the avalanche of proteomic data currently being reported, many genome sequences are established using next-generation sequencing, fostering proteomic investigations of new cellular models. Proteogenomics is a relatively recent field in which high-throughput proteomic data is used to verify coding regions within model genomes to refine the annotation of their sequences (28). Because genome annotation is now fully automated, the need for accurate annotation for model organisms with experimental data is crucial. Many projects related to genome re-annotation of microorganisms with the help of proteomics have been recently reported, such as for Mycoplasma pneumoniae (9), Rhodopseudomonas palustris (10), Shewanella oneidensis (11), Thermococcus gammatolerans (12), Deinococcus deserti (13), Salmonella thyphimurium (14), Mycobacterium tuberculosis (15, 16), Shigella flexneri (17), Ruegeria pomeroyi (18), and Candida glabrata (19), as well as for higher organisms such as Anopheles gambiae (20) and Arabidopsis thaliana (4, 5).The most frequently reported problem in automatic annotation systems is the correct identification of the translational start codon (2123). The error rate depends on the primary annotation system, but also on the organism, as reported for Halobacterium salinarum and Natromonas pharaonis (24), Deinococcus deserti (21), and Ruegeria pomeroyi (18), where the error rate is estimated above 10%. Identification of a correct translational start site is essential for the genetic and biochemical analysis of a protein because errors can seriously impact subsequent biological studies. If the N terminus is not correctly identified, the protein will be considered in either a truncated or extended form, leading to errors in bioinformatic analyses (e.g. during the prediction of its molecular weight, isoelectric point, cellular localization) and major difficulties during its experimental characterization. For example, a truncated protein may be heterologously produced as an unfolded polypeptide recalcitrant to structure determination (25). Moreover, N-terminal modifications, which are poorly documented in annotation databases, may occur (26, 27).Unfortunately, the poor polypeptide sequence coverage obtained for the numerous low abundance proteins in current shotgun MS/MS proteomic studies implies that the overall detection of N-terminal peptides obtained in proteogenomic studies is relatively low. Different methods for establishing the most extensive list of protein N termini, grouped under the so-called “N-terminomics” theme, have been proposed to selectively enrich or improve the detection of these peptides (2, 28, 29). Large N-terminome studies have recently been reported based on resin-assisted enrichment of N-terminal peptides (30) or terminal amine isotopic labeling of substrates (TAILS) coupled to depletion of internal peptides with a water-soluble aldehyde-functionalized polymer (3135). Among the numerous N-terminal-oriented methods (2), specific labeling of the N terminus of intact proteins with N-tris(2,4,6-trimethoxyphenyl)phosphonium acetyl succinamide (TMPP-Ac-OSu)1 has proven reliable (21, 3639). TMPP-derivatized N-terminal peptides have interesting properties for further LC-MS/MS mass spectrometry: (1) an increase in hydrophobicity because of the trimethoxyphenyl moiety added to the peptides, increasing their retention times in reverse phase chromatography, (2) improvement of their ionization because of the introduction of a positively charged group, and (3) a much simpler fragmentation pattern in tandem mass spectrometry. Other reported approaches rely on acetylation, followed by trypsin digestion, and then biotinylation of free amino groups (40); guanidination of lysine lateral chains followed by N-biotinylation of the N termini and trypsin digestion (41); or reductive amination of all free amino groups with formaldehyde preceeding trypsin digestion (42). Recently, we applied the TMPP method to the proteome of the Deinococcus deserti bacterium isolated from upper sand layers of the Sahara desert (13). This method enabled the detection of N-terminal peptides allowing the confirmation of 278 translation initiation codons, the correction of 73 translation starts, and the identification of non-canonical translation initiation codons (21). However, most TMPP-labeled N-terminal peptides are hidden among the more abundant internal peptides generated after proteolysis of a complex proteome, precluding their detection. This results in disproportionately fewer N-terminal validations, that is, 5 and 8% of total polypeptides coded in the theoretical proteomes of Mycobacterium smegmatis (37) and Deinococcus deserti (21) with a total of 342 and 278 validations, respectively.An interesting chromatographic method to fractionate peptide mixtures for gel-free high-throughput proteome analysis has been developed over the last years and applied to various topics (43, 44). This technique, known as COmbined FRActional DIagonal Chromatography (COFRADIC), uses a double chromatographic separation with a chemical reaction in between to change the physico-chemical properties of the extraneous peptides to be resolved from the peptides of interest. Its previous applications include the separation of methionine-containing peptides (43), N-terminal peptide enrichment (45, 46), sulfur amino acid-containing peptides (47), and phosphorylated peptides (48). COFRADIC was identified as the best method for identification of N-terminal peptides of two archaea, resulting in the identification of 240 polypeptides (9% of the theoretical proteome) for Halobacterium salinarum and 220 (8%) for Natronomonas pharaonis (24).Taking advantage of both the specificity of TMPP labeling, the resolving power of COFRADIC for enrichment, and the increase in information through the use of multiple proteases, we performed the proteogenomic analysis of a marine bacterium from the Roseobacter clade, namely Roseobacter denitrificans OCh114. This novel approach allowed us to validate and correct 534 unique proteins (13% of the theoretical proteome) with TMPP-labeled N-terminal signatures obtained using high-resolution tandem mass spectrometry. We corrected 41 annotations and detected five new open reading frames in the R. denitrificans genome. We further identified eight distinct proteins showing direct evidence for multiple start sites.  相似文献   

6.
7.
Most human genes undergo alternative splicing, but aberrant splice forms are hallmarks of many cancers, usually resulting from mutations initiating abnormal exon skipping, intron retention, or the introduction of a new splice sites. We have identified a family of aberrant splice variants of HAS1 (the hyaluronan synthase 1 gene) in some B lineage cancers, characterized by exon skipping and/or partial intron retention events that occur either together or independently in different variants, apparently due to accumulation of inherited and acquired mutations. Cellular, biochemical, and oncogenic properties of full-length HAS1 (HAS1-FL) and HAS1 splice variants Va, Vb, and Vc (HAS1-Vs) are compared and characterized. When co-expressed, the properties of HAS1-Vs are dominant over those of HAS1-FL. HAS1-FL appears to be diffusely expressed in the cell, but HAS1-Vs are concentrated in the cytoplasm and/or Golgi apparatus. HAS1-Vs synthesize detectable de novo HA intracellularly. Each of the HAS1-Vs is able to relocalize HAS1-FL protein from diffuse cytoskeleton-anchored locations to deeper cytoplasmic spaces. This HAS1-Vs-mediated relocalization occurs through strong molecular interactions, which also serve to protect HAS1-FL from its otherwise high turnover kinetics. In co-transfected cells, HAS1-FL and HAS1-Vs interact with themselves and with each other to form heteromeric multiprotein assemblies. HAS1-Vc was found to be transforming in vitro and tumorigenic in vivo when introduced as a single oncogene to untransformed cells. The altered distribution and half-life of HAS1-FL, coupled with the characteristics of the HAS1-Vs suggest possible mechanisms whereby the aberrant splicing observed in human cancer may contribute to oncogenesis and disease progression.About 70–80% of human genes undergo alternative splicing, contributing to proteomic diversity and regulatory complexities in normal development (1). About 10% of mutations listed so far in the Human Gene Mutation Database (HGMD) of “gene lesions responsible for human inherited disease” were found to be located within splice sites. Furthermore, it is becoming increasingly apparent that aberrant splice variants, generated mostly due to splicing defects, play a key role in cancer. Germ line or acquired genomic changes (mutations) in/around splicing elements (24) promote aberrant splicing and aberrant protein isoforms.Hyaluronan (HA)3 is synthesized by three different plasma membrane-bound hyaluronan synthases (1, 2, and 3). HAS1 undergoes alternative and aberrant intronic splicing in multiple myeloma, producing truncated variants termed Va, Vb, and Vc (5, 6), which predicted for poor survival in a cohort of multiple myeloma patients (5). Our work suggests that this aberrant splicing arises due to inherited predispositions and acquired mutations in the HAS1 gene (7). Cancer-related, defective mRNA splicing caused by polymorphisms and/or mutations in splicing elements often results in inactivation of tumor suppressor activity (e.g. HRPT2 (8, 9), PTEN (10), MLHI (1114), and ATR (15)) or generation of dominant negative inhibitors (e.g. CHEK2 (16) and VWOX (17)). In breast cancer, aberrantly spliced forms of progesterone and estrogen receptors are found (reviewed in Ref. 3). Intronic mutations inactivate p53 through aberrant splicing and intron retention (18). Somatic mutations with the potential to alter splicing are frequent in some cancers (1925). Single nucleotide polymorphisms in the cyclin D1 proto-oncogene predispose to aberrant splicing and the cyclin D1b intronic splice variant (2629). Cyclin D1b confers anchorage independence, is tumorogenic in vivo, and is detectable in human tumors (30), but as yet no clinical studies have confirmed an impact on outcome. On the other hand, aberrant splicing of HAS1 shows an association between aberrant splice variants and malignancy, suggesting that such variants may be potential therapeutic targets and diagnostic indicators (19, 3133). Increased HA expression has been associated with malignant progression of multiple tumor types, including breast, prostate, colon, glioma, mesothelioma, and multiple myeloma (34). The three mammalian HA synthase (HAS) isoenzymes synthesize HA and are integral transmembrane proteins with a probable porelike structural assembly (3539). Although in humans, the three HAS genes are located on different chromosomes (hCh19, hCh8, and hCh16, respectively) (40), they share a high degree of sequence homology (41, 42). HAS isoenzymes synthesize a different size range of HA molecules, which exhibit different functions (43, 44). HASs contribute to a variety of cancers (4555). Overexpression of HASs promotes growth and/or metastatic development in fibrosarcoma, prostate, and mammary carcinoma, and the removal of the HA matrix from a migratory cell membrane inhibits cell movement (45, 53). HAS2 confers anchorage independence (56). Our work has shown aberrant HAS1 splicing in multiple myeloma (5) and Waldenstrom''s macroglobulinemia (6). HAS1 is overexpressed in colon (57), ovarian (58), endometrial (59), mesothelioma (60), and bladder cancers (61). A HAS1 splice variant is detected in bladder cancer (61).Here, we characterize molecular and biochemical characteristics of HAS1 variants (HAS1-Vs) (5), generated by aberrant splicing. Using transient transfectants and tagged HAS1 family constructs, we show that HAS1-Vs differ in cellular localization, de novo HA localization, and turnover kinetics, as compared with HAS1-FL, and dominantly influence HAS1-FL when co-expressed. HAS1-Vs proteins form intra- and intermolecular associations among themselves and with HAS1-FL, including covalent interactions and multimer formation. HAS1-Vc supports vigorous cellular transformation of NIH3T3 cells in vitro, and HAS1-Vc-transformed NIH3T3 cells are tumorogenic in vivo.  相似文献   

8.
9.
Many biological processes involve the mechanistic/mammalian target of rapamycin complex 1 (mTORC1). Thus, the challenge of deciphering mTORC1-mediated functions during normal and pathological states in the central nervous system is challenging. Because mTORC1 is at the core of translation, we have investigated mTORC1 function in global and regional protein expression. Activation of mTORC1 has been generally regarded to promote translation. Few but recent works have shown that suppression of mTORC1 can also promote local protein synthesis. Moreover, excessive mTORC1 activation during diseased states represses basal and activity-induced protein synthesis. To determine the role of mTORC1 activation in protein expression, we have used an unbiased, large-scale proteomic approach. We provide evidence that a brief repression of mTORC1 activity in vivo by rapamycin has little effect globally, yet leads to a significant remodeling of synaptic proteins, in particular those proteins that reside in the postsynaptic density. We have also found that curtailing the activity of mTORC1 bidirectionally alters the expression of proteins associated with epilepsy, Alzheimer''s disease, and autism spectrum disorder—neurological disorders that exhibit elevated mTORC1 activity. Through a protein–protein interaction network analysis, we have identified common proteins shared among these mTORC1-related diseases. One such protein is Parkinson protein 7, which has been implicated in Parkinson''s disease, yet not associated with epilepsy, Alzheimers disease, or autism spectrum disorder. To verify our finding, we provide evidence that the protein expression of Parkinson protein 7, including new protein synthesis, is sensitive to mTORC1 inhibition. Using a mouse model of tuberous sclerosis complex, a disease that displays both epilepsy and autism spectrum disorder phenotypes and has overactive mTORC1 signaling, we show that Parkinson protein 7 protein is elevated in the dendrites and colocalizes with the postsynaptic marker postsynaptic density-95. Our work offers a comprehensive view of mTORC1 and its role in regulating regional protein expression in normal and diseased states.The mechanistic/mammalian target of rapamycin complex 1 (mTORC1)1 is a serine/threonine protein kinase that is highly expressed in many cell types (1). In the brain, mTORC1 tightly coordinates different synaptic plasticities — long-term potentiation (LTP) and long-term depression (LTD) — the molecular correlates of learning and memory (25). Because mTORC1 is at the core of many synaptic signaling pathways downstream of glutamate and neurotrophin receptors, many hypothesize that dysregulated mTORC1 signaling underlies cognitive deficits observed in several neurodegenerative diseases (3, 617). For example, mTORC1 and its downstream targets are hyperactive in human brains diagnosed with Alzheimer''s disease (AD) (1820). Additionally in animal models of autism spectrum disorder (ASD), altered mTORC1 signaling contributes to the observed synaptic dysfunction and aberrant network connectivity (13, 15, 2127). Furthermore, epilepsy, which is common in AD and ASD, has enhanced mTORC1 activity (2832).Phosphorylation of mTORC1, considered the active form, is generally regarded to promote protein synthesis (33). Thus, many theorize that diseases with overactive mTORC1 arise from excessive protein synthesis (14). Emerging data, however, show that suppressing mTORC1 activation can trigger local translation in neurons (34, 35). Pharmacological antagonism of N-methyl-d-aspartate (NMDA) receptors, a subtype of glutamate receptors that lies upstream of mTOR activation, promotes the synthesis of the voltage-gated potassium channel, Kv1.1, in dendrites (34, 35). Consistent with these results, in models of temporal lobe epilepsy there is a reduction in the expression of voltage-gated ion channels including Kv1.1 (30, 31, 36). Interestingly in a model of focal neocortical epilepsy, overexpression of Kv1.1 blocked seizure activity (37). Because both active and inactive mTORC1 permit protein synthesis, we sought to determine the proteins whose expression is altered when mTORC1 phosphorylation is reduced in vivo.Rapamycin is an FDA-approved, immunosuppressive drug that inhibits mTORC1 activity (38). We capitalized on the ability of rapamycin to reduce mTORC1 activity in vivo and the unbiased approach of mass spectrometry to identify changes in protein expression. Herein, we provide evidence that mTORC1 activation bidirectionally regulates protein expression, especially in the PSD where roughly an equal distribution of proteins dynamically appear and disappear. Remarkably, using protein–protein interaction networks facilitated the novel discovery that PARK7, a protein thus far only implicated in Parkinson''s disease, (1) is up-regulated by increased mTORC1 activity, (2) resides in the PSD only when mTORC1 is active, and (3) is aberrantly expressed in a rodent model of TSC, an mTORC1-related disease that has symptoms of epilepsy and autism. Collectively, these data provide the first comprehensive list of proteins whose abundance or subcellular distributions are altered with acute changes in mTORC1 activity in vivo.  相似文献   

10.
11.
12.
13.
The c-Jun N-terminal kinase (JNK) is an important mediator of physiological and pathophysiological processes in the central nervous system. Importantly, JNK not only is involved in neuronal cell death, but also plays a significant role in neuronal differentiation and regeneration. For example, nerve growth factor induces JNK-dependent neuronal differentiation in several model systems. The mechanism by which JNK mediates neuronal differentiation is not well understood. Here, we employed a proteomic strategy to better characterize the function of JNK during neuronal differentiation. We used SILAC-based quantitative proteomics to identify proteins that interact with JNK in PC12 cells in a nerve growth factor–dependent manner. Intriguingly, we found that JNK interacted with neuronal transport granule proteins such as Sfpq and Nono upon NGF treatment. We validated the specificity of these interactions by showing that they were disrupted by a specific peptide inhibitor that blocks the interaction of JNK with its substrates. Immunoprecipitation and Western blotting experiments confirmed the interaction of JNK1 with Sfpq/Nono and demonstrated that it was RNA dependent. Confocal microscopy indicated that JNK1 associated with neuronal granule proteins in the cytosol of PC12 cells, primary cortical neurons, and P19 neuronal cells. Finally, siRNA experiments confirmed that Sfpq was necessary for neurite outgrowth in PC12 cells and that it most likely acted in the same pathway as JNK. In summary, our data indicate that the interaction of JNK1 with transport granule proteins in the cytosol of differentiating neurons plays an important role during neuronal development.The members of the c-Jun N-terminal kinase (JNK)1 family are important mediators of a broad range of biological processes in the brain (1, 2). JNK belongs to the family of mitogen-activated protein kinases and is encoded by three different genes: jnk1, jnk2, and jnk3. Alternative splicing combined with alternative exon usage of the three genes leads to at least 10 different JNK isoforms (3). JNK is induced by various stimuli such as cytokines, ligands of Toll-like receptors, or growth factors (4).JNK is well known to induce neuronal cell death. However, it has become clear within the past decade that JNK plays an important role in neuronal regeneration, migration, and differentiation (58). JNK controls dendritic microtubule assembly and disassembly in sympathetic neurons (2, 9) and is responsible for dendritic elongation during brain development in mice (10). Furthermore, several cytoskeleton-regulating proteins such as doublecortin, superior cervical ganglion-10 protein, microtubule-associated protein 1B, microtubule-associated protein 2, and MARCKS-like protein 1 have been identified as JNK substrates (1116). Collectively, these observations have established that JNKs not only are involved in neuronal death, but also play an important role during differentiation and regeneration. Although stress-induced pro-apoptotic JNK signaling is well characterized, the mechanisms involved in JNK-dependent neuronal differentiation remain enigmatic.Quantitative mass-spectrometry-based proteomics has emerged as a powerful technology for investigating mammalian signaling pathways (17, 18). Specifically, stable isotope labeling by amino acids in cell culture (SILAC) has been shown to be a powerful approach for studying dynamic changes in protein–protein interactions during cell signaling. We reasoned that an unbiased proteomic analysis of JNK interaction partners during neuronal differentiation could provide novel insights into the mechanisms involved. For that purpose, we used PC12 cells as a well-characterized classical model for neuronal differentiation (19). Stimulation of PC12 cells with nerve growth factor (NGF) induces a major shift in phenotype, from proliferating tumor cells to non-dividing neurons showing characteristics of sympathetic neurons such as the growth of long neurites and electrical excitability (19, 20). In PC12 cells, NGF binds to neurotrophic tyrosine kinase receptor 1, low-affinity neurotrophin receptor, and fibroblast growth factor receptors (21, 22). This leads to a coordinated activation of downstream signaling cascades such as the Ras/Raf/Erk1–2, PI3K, and JNK pathways, resulting in increased expression of genes that are involved in neuronal differentiation (23, 24). Furthermore, G-proteins including Ras, Rap, and Cdc42 have been shown to link receptor activity to downstream kinase activation (2527).Here, we used quantitative interaction proteomics to analyze dynamic changes in JNK interaction partners during NGF-induced differentiation of PC12 cells. Our results show that JNK dynamically interacts with G-proteins, cytoskeletal proteins, and RNA binding proteins with distinct kinetic patterns. Intriguingly, several of the identified RNA binding proteins are known components of transport granules involved in mRNA localization and localized translation in neurons. Western blotting and co-localization experiments using confocal microscopy and proximity ligation assays validated the NGF-induced association of JNK1 with two RNA binding proteins, Sfpq and Nono. Furthermore, Sfpq knockdown decreased NGF-induced neurite outgrowth in PC12 cells, supporting the hypothesis that the interaction of JNK and Sfpq may contribute to neuronal differentiation.  相似文献   

14.
15.
16.
17.
18.
SPA2 encodes a yeast protein that is one of the first proteins to localize to sites of polarized growth, such as the shmoo tip and the incipient bud. The dynamics and requirements for Spa2p localization in living cells are examined using Spa2p green fluorescent protein fusions. Spa2p localizes to one edge of unbudded cells and subsequently is observable in the bud tip. Finally, during cytokinesis Spa2p is present as a ring at the mother–daughter bud neck. The bud emergence mutants bem1 and bem2 and mutants defective in the septins do not affect Spa2p localization to the bud tip. Strikingly, a small domain of Spa2p comprised of 150 amino acids is necessary and sufficient for localization to sites of polarized growth. This localization domain and the amino terminus of Spa2p are essential for its function in mating. Searching the yeast genome database revealed a previously uncharacterized protein which we name, Sph1p (Spa2p homolog), with significant homology to the localization domain and amino terminus of Spa2p. This protein also localizes to sites of polarized growth in budding and mating cells. SPH1, which is similar to SPA2, is required for bipolar budding and plays a role in shmoo formation. Overexpression of either Spa2p or Sph1p can block the localization of either protein fused to green fluorescent protein, suggesting that both Spa2p and Sph1p bind to and are localized by the same component. The identification of a 150–amino acid domain necessary and sufficient for localization of Spa2p to sites of polarized growth and the existence of this domain in another yeast protein Sph1p suggest that the early localization of these proteins may be mediated by a receptor that recognizes this small domain.Polarized cell growth and division are essential cellular processes that play a crucial role in the development of eukaryotic organisms. Cell fate can be determined by cell asymmetry during cell division (Horvitz and Herskowitz, 1992; Cohen and Hyman, 1994; Rhyu and Knoblich, 1995). Consequently, the molecules involved in the generation and maintenance of cell asymmetry are important in the process of cell fate determination. Polarized growth can occur in response to external signals such as growth towards a nutrient (Rodriguez-Boulan and Nelson, 1989; Eaton and Simons, 1995) or hormone (Jackson and Hartwell, 1990a , b ; Segall, 1993; Keynes and Cook, 1995) and in response to internal signals as in Caenorhabditis elegans (Goldstein et al., 1993; Kimble, 1994; Priess, 1994) and Drosophila melanogaster (St Johnston and Nusslein-Volhard, 1992; Anderson, 1995) early development. Saccharomyces cerevisiae undergo polarized growth towards an external cue during mating and to an internal cue during budding. Polarization towards a mating partner (shmoo formation) and towards a new bud site requires a number of proteins (Chenevert, 1994; Chant, 1996; Drubin and Nelson, 1996). Many of these proteins are necessary for both processes and are localized to sites of polarized growth, identified by the insertion of new cell wall material (Tkacz and Lampen, 1972; Farkas et al., 1974; Lew and Reed, 1993) to the shmoo tip, bud tip, and mother–daughter bud neck. In yeast, proteins localized to growth sites include cytoskeletal proteins (Adams and Pringle, 1984; Kilmartin and Adams, 1984; Ford, S.K., and J.R. Pringle. 1986. Yeast. 2:S114; Drubin et al., 1988; Snyder, 1989; Snyder et al., 1991; Amatruda and Cooper, 1992; Lew and Reed, 1993; Waddle et al., 1996), neck filament components (septins) (Byers and Goetsch, 1976; Kim et al., 1991; Ford and Pringle, 1991; Haarer and Pringle, 1987; Longtine et al., 1996), motor proteins (Lillie and Brown, 1994), G-proteins (Ziman, 1993; Yamochi et al., 1994; Qadota et al., 1996), and two membrane proteins (Halme et al., 1996; Roemer et al., 1996; Qadota et al., 1996). Septins, actin, and actin-associated proteins localize early in the cell cycle, before a bud or shmoo tip is recognizable. How this group of proteins is localized to and maintained at sites of cell growth remains unclear.Spa2p is one of the first proteins involved in bud formation to localize to the incipient bud site before a bud is recognizable (Snyder, 1989; Snyder et al., 1991; Chant, 1996). Spa2p has been localized to where a new bud will form at approximately the same time as actin patches concentrate at this region (Snyder et al., 1991). An understanding of how Spa2p localizes to incipient bud sites will shed light on the very early stages of cell polarization. Later in the cell cycle, Spa2p is also found at the mother–daughter bud neck in cells undergoing cytokinesis. Spa2p, a nonessential protein, has been shown to be involved in bud site selection (Snyder, 1989; Zahner et al., 1996), shmoo formation (Gehrung and Snyder, 1990), and mating (Gehrung and Snyder, 1990; Chenevert et al., 1994; Yorihuzi and Ohsumi, 1994; Dorer et al., 1995). Genetic studies also suggest that Spa2p has a role in cytokinesis (Flescher et al., 1993), yet little is known about how this protein is localized to sites of polarized growth.We have used Spa2p green fluorescent protein (GFP)1 fusions to investigate the early localization of Spa2p to sites of polarized growth in living cells. Our results demonstrate that a small domain of ∼150 amino acids of this large 1,466-residue protein is sufficient for targeting to sites of polarized growth and is necessary for Spa2p function. Furthermore, we have identified and characterized a novel yeast protein, Sph1p, which has homology to both the Spa2p amino terminus and the Spa2p localization domain. Sph1p localizes to similar regions of polarized growth and sph1 mutants have similar phenotypes as spa2 mutants.  相似文献   

19.
20.
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   

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