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.
G protein-coupled receptors (GPCRs) constitute the largest family among mammalian membrane proteins and are capable of initiating numerous essential signaling cascades. Various GPCR-mediated pathways are organized into protein microdomains that can be orchestrated and regulated through scaffolding proteins, such as PSD-95/discs-large/ZO1 (PDZ) domain proteins. However, detailed binding characteristics of PDZ–GPCR interactions remain elusive because these interactions seem to be more complex than previously thought. To address this issue, we analyzed binding modalities using our established model system. This system includes the 13 individual PDZ domains of the multiple PDZ domain protein 1 (MUPP1; the largest PDZ protein), a broad range of murine olfactory receptors (a multifaceted gene cluster within the family of GPCRs), and associated olfactory signaling proteins. These proteins were analyzed in a large-scale peptide microarray approach and continuative interaction studies. As a result, we demonstrate that canonical binding motifs were not overrepresented among the interaction partners of MUPP1. Furthermore, C-terminal phosphorylation and distinct amino acid replacements abolished PDZ binding promiscuity. In addition to the described in vitro experiments, we identified new interaction partners within the murine olfactory epithelium using pull-down-based interactomics and could verify the partners through co-immunoprecipitation. In summary, the present study provides important insight into the complexity of the binding characteristics of PDZ–GPCR interactions based on olfactory signaling proteins, which could identify novel clinical targets for GPCR-associated diseases in the future.PDZ domain proteins comprise one of the largest families among interaction domain scaffolding proteins and are highly abundant in various multicellular eukaryotic species. These proteins fulfill important physiological functions in a broad range of different tissues and cells as they orchestrate complex protein networks. Among putative PDZ interaction partners, one important protein family is the group of GPCRs1, constituting the largest family of membrane proteins in mammals (1). Here, signal efficiency, speed, desensitization, and internalization can be modulated by PDZ proteins (25). Olfactory receptors (ORs) represent a multigene family within this group of seven-transmembrane domain proteins and encompass 2% of the mammalian genome (6). Belonging to class I GPCRs, ORs share many general features of this receptor family, making them an interesting target for interactions involving PDZ proteins. Until recently, an organizing complex builder, such as the inactivation no afterpotential D (InaD) protein in the visual system of Drosophila melanogaster (7, 8), could not be clearly identified for olfactory signaling.The multiple PDZ domain protein 1, with 13 individual PDZ domains, represents the largest of the described PDZ proteins to date (9) and interacts with different GPCRs (1012). One well-described example is its interaction with GABAB receptors, leading to enhanced receptor stability at the plasma membrane and prolonged signaling duration (2). In previous studies, we demonstrated that PDZ domains 1 + 2 can interact with a selected subset of ORs (13). Furthermore, we showed that MUPP1 binds to a specific OR and that most of the described proteins are involved in mammalian olfactory signal transduction in the native system, making MUPP1 a promising candidate for orchestrating the olfactory system (14).Many PDZ–ligand interactions depend on classical binding motifs at the ligand''s C-terminal end, thereby building weak transient protein complexes (15, 16). However, an increasing number of PDZ interactions have emerged that seem to provide more complex binding modalities, differing from the canonical interactions (17, 18). Ligand binding seems not to be exclusively restricted to C-terminal sites, and PDZ domains cannot be distinctly classified but are evenly distributed throughout a selective space (17, 1921). Therefore, it is of great interest to analyze OR–PDZ interactions to characterize the putative binding requirements and to further investigate the role of MUPP1 in olfactory signaling.In the present study, we characterized the binding modalities between the 13 individual PDZ domains of MUPP1 and a broad range of murine olfactory receptors in a large-scale approach, indicating that classical binding motifs were not overrepresented among the evaluated binding partners. In addition, we identified new binding partners from the murine olfactory epithelium using pull-down-based interactomics.  相似文献   

2.
3.
LINGO-1 is a component of the tripartite receptor complexes, which act as a convergent mediator of the intracellular signaling in response to myelin-associated inhibitors and lead to collapse of growth cone and inhibition of neurite extension. Although the function of LINGO-1 has been intensively studied, its downstream signaling remains elusive. In the present study, a novel interaction between LINGO-1 and a serine-threonine kinase WNK1 was identified by yeast two-hybrid screen. The interaction was further validated by fluorescence resonance energy transfer and co-immunoprecipitation, and this interaction was intensified by Nogo66 treatment. Morphological evidences showed that WNK1 and LINGO-1 were co-localized in cortical neurons. Furthermore, either suppressing WNK1 expression by RNA interference or overexpression of WNK1-(123–510) attenuated Nogo66-induced inhibition of neurite extension and inhibited the activation of RhoA. Moreover, WNK1 was identified to interact with Rho-GDI1, and this interaction was attenuated by Nogo66 treatment, further indicating its regulatory effect on RhoA activation. Taken together, our results suggest that WNK1 is a novel signaling molecule involved in regulation of LINGO-1 mediated inhibition of neurite extension.Axons of the adult mammalian central nervous system possess an extremely limited ability to regenerate after injury, largely because of inhibitory components of myelin preventing axon growth (1, 2). Several myelin-associated inhibitors have been identified, including myelin-associated glycoprotein (35), chondroitin sulfate proteoglycans (6), oligodendrocyte myelin glycoprotein (7), and Nogo (810). Myelin-associated glycoprotein, oligodendrocyte myelin glycoprotein, and Nogo bind to the Nogo-66 receptor (NgR)3 and exert their actions through a tripartite receptor complex NgR/LINGO-1/p75NTR (11) or NgR/LINGO-1/TROY (12, 13).LINGO-1 is a transmembrane protein that contains a leucine-rich repeat, an immunoglobulin domain, and a short intracellular tail (11). LINGO-1 functions as an essential component of the NgR complexes that mediate the activity of myelin inhibitors to regulate central nervous system axon growth (11, 14). In neurons, the NgR complexes activate RhoA in the presence of myelin inhibitors, which lead to growth cone collapse and neurite extension inhibition (11). Attenuation of LINGO-1 function is able to overcome the myelin inhibitory activity in the spinal cord that prevents axonal regeneration after lesion in rats (15). Besides, it has been reported that LINGO-1 is also expressed in oligodendrocytes, where it negatively regulates oligodendrocyte differentiation and axon myelination (16). Inhibition of LINGO-1 promotes spinal cord remyelination in an experimental model of autoimmune encephalitis (17). Moreover, inhibition of LINGO-1 has been shown to enhance survival, structure, and function of dopaminergic neurons in Parkinson disease models (18). Although the function of LINGO-1 has been intensively studied, much less is known about its downstream signaling.To gain insight into the mechanisms by which LINGO-1 functions, it is of considerable importance to identify new binding partners of LINGO-1. Therefore, using the intracellular domain of LINGO-1 as bait, we employed yeast two-hybrid screening on a brain cDNA library and identified several candidates that interact with LINGO-1, one of which is the protein kinase WNK1.WNKs (with no lysine [K]) are a distinct subfamily of serine-threonine kinases, which are characterized by a unique placement of the lysine that is involved in binding ATP and catalyzing phosphoryl transfer (19). Thus far, WNKs are known composed of four members, WNK1, WNK2, WNK3, and WNK4. Mutations in the serine-threonine kinases WNK1 and WNK4 cause a Mendelian disease PAHII, featuring hypertension and hyperkalemia (20, 21), and their roles in the regulation of electrolyte flux in the kidney have been well established (22). Recently, other important features of WNKs are beginning to be understood. WNKs have also been proposed functioning in a number of non-transport processes, including cell growth, differentiation, and apoptosis (2326). Although WNK1 has been shown to be expressed in brain (27, 28), little is known about its function in the nervous system until recently; mutations of a nervous system-specific exon of the WNK1 gene were found to cause Hereditary sensory and autonomic neuropathy type II (HSANII) (29). In this study WNK1 was demonstrated to interact with LINGO-1 and regulate Nogo-induced inhibition of neurite extension.  相似文献   

4.
There are a large number of Rho guanine nucleotide exchange factors, most of which have no known functions. Here, we carried out a short hairpin RNA-based functional screen of Rho-GEFs for their roles in leukocyte chemotaxis and identified Arhgef5 as an important factor in chemotaxis of a macrophage phage-like RAW264.7 cell line. Arhgef5 can strongly activate RhoA and RhoB and weakly RhoC and RhoG, but not Rac1, RhoQ, RhoD, or RhoV, in transfected human embryonic kidney 293 cells. In addition, Gβγ interacts with Arhgef5 and can stimulate Arhgef5-mediated activation of RhoA in an in vitro assay. In vivo roles of Arhgef5 were investigated using an Arhgef-5-null mouse line. Arhgef5 deficiency did not affect chemotaxis of mouse macrophages, T and B lymphocytes, and bone marrow-derived mature dendritic cells (DC), but it abrogated MIP1α-induced chemotaxis of immature DCs and impaired migration of DCs from the skin to lymph node. In addition, Arhgef5 deficiency attenuated allergic airway inflammation. Therefore, this study provides new insights into signaling mechanisms for DC migration regulation.Leukocyte chemotaxis underlies leukocyte migration, infiltration, trafficking, and homing that are not only important for normal leukocyte functions, but also have a important role in inflammation-related diseases. Leukocyte chemotaxis is regulated by leukocyte chemoattractants that include bacterial by-products such as formylmethionylleucylphenylalanine, complement proteolytic fragments such as C5a, and the superfamily of chemotactic cytokines, chemokines. These chemoattractants bind to their specific cell G protein-coupled receptors and are primarily coupled to the Gi family of G proteins to regulate leukocyte chemotaxis. Previous studies have established that the Rho family of small GTPases regulates leukocyte migration (1, 2). Rac, Cdc42, and RhoA are the three best studied Rho small GTPases. In myeloid cells, Cdc42 regulates directionality by directing where F-actin and lamellipodia are formed, and Rac regulates F-actin formation in the lamellipodia, which provides a driving force for cell motility (36). On the other hand, RhoA regulates the formation and contractility of the actomyosin structure at the back that provides a pushing force (5, 7). Rho guanine nucleotide exchange factors (GEF)3 are key regulators for the activity of these small GTPases. GEFs activate small GTPases by promoting the loading of GTP to the small GTPases, a rate-limiting step in GTPase regulation (811). Previous biochemical and genetic studies have revealed how Cdc42 and Rac may be regulated by chemokine receptors in leukocytes. Chemokine receptors can regulate Cdc42 via a Rho-GEF PIXα, which is regulated by Gβγ from the Gi proteins via the interactions between Gβγ and Pak1 and between Pak1 and PIXα in myeloid cells 12. On the other hand, in neutrophils chemokine receptors regulate Rac2 via another Rho-GEF P-Rex1, which is directly regulated by Gβγ (1315). Two Rho-GEFs have been implicated in regulation of RhoA in neutrophils. GEF115 was found in the leading edges of polarized mouse neutrophils, whereas PDZ Rho-GEF was found in the uropods of differentiated HL-60 cells. Both Rho-GEFs were believed to mediate pertussis toxin-resistant activation of RhoA in these cells. However, a significant portion of RhoA activity in leukocytes are pertussis toxin-sensitive, which is presumably regulated by the α and/or βγ subunits from the Gi proteins. The signaling mechanism for this pertussis toxin-sensitive RhoA regulation by chemokine receptors remains largely elusive.Molecular cloning and genomic sequencing have identified more than 70 Rho-GEFs in mammals (1620). Many of these Rho-GEFs have been shown to activate RhoA in in vitro and overexpression assays (1620). However, it is not known if any of them regulate RhoA in vivo, we have found that PIXα is a specific GEF for Cdcd42 in neutrophils (12) despite its potent activity on Rac in in vitro and overexpression assays (21, 22). Therefore, we used a siRNA-based loss of function screen in an attempt to identify the GEFs that regulate myeloid cell migration and RhoA activity. One of the candidates, Arhgef5, was found to be directly activated by Gβγ to regulate RhoA and has an important role in immature DC migration. In addition, Arhgef5 deficiency attenuated allergic airway inflammation in a mouse model.  相似文献   

5.
6.
Previous studies have shown that protein-protein interactions among splicing factors may play an important role in pre-mRNA splicing. We report here identification and functional characterization of a new splicing factor, Sip1 (SC35-interacting protein 1). Sip1 was initially identified by virtue of its interaction with SC35, a splicing factor of the SR family. Sip1 interacts with not only several SR proteins but also with U1-70K and U2AF65, proteins associated with 5′ and 3′ splice sites, respectively. The predicted Sip1 sequence contains an arginine-serine-rich (RS) domain but does not have any known RNA-binding motifs, indicating that it is not a member of the SR family. Sip1 also contains a region with weak sequence similarity to the Drosophila splicing regulator suppressor of white apricot (SWAP). An essential role for Sip1 in pre-mRNA splicing was suggested by the observation that anti-Sip1 antibodies depleted splicing activity from HeLa nuclear extract. Purified recombinant Sip1 protein, but not other RS domain-containing proteins such as SC35, ASF/SF2, and U2AF65, restored the splicing activity of the Sip1-immunodepleted extract. Addition of U2AF65 protein further enhanced the splicing reconstitution by the Sip1 protein. Deficiency in the formation of both A and B splicing complexes in the Sip1-depleted nuclear extract indicates an important role of Sip1 in spliceosome assembly. Together, these results demonstrate that Sip1 is a novel RS domain-containing protein required for pre-mRNA splicing and that the functional role of Sip1 in splicing is distinct from those of known RS domain-containing splicing factors.Pre-mRNA splicing takes place in spliceosomes, the large RNA-protein complexes containing pre-mRNA, U1, U2, U4/6, and U5 small nuclear ribonucleoprotein particles (snRNPs), and a large number of accessory protein factors (for reviews, see references 21, 22, 37, 44, and 48). It is increasingly clear that the protein factors are important for pre-mRNA splicing and that studies of these factors are essential for further understanding of molecular mechanisms of pre-mRNA splicing.Most mammalian splicing factors have been identified by biochemical fractionation and purification (3, 15, 19, 3136, 45, 6971, 73), by using antibodies recognizing splicing factors (8, 9, 16, 17, 61, 66, 67, 74), and by sequence homology (25, 52, 74).Splicing factors containing arginine-serine-rich (RS) domains have emerged as important players in pre-mRNA splicing. These include members of the SR family, both subunits of U2 auxiliary factor (U2AF), and the U1 snRNP protein U1-70K (for reviews, see references 18, 41, and 59). Drosophila alternative splicing regulators transformer (Tra), transformer 2 (Tra2), and suppressor of white apricot (SWAP) also contain RS domains (20, 40, 42). RS domains in these proteins play important roles in pre-mRNA splicing (7, 71, 75), in nuclear localization of these splicing proteins (23, 40), and in protein-RNA interactions (56, 60, 64). Previous studies by us and others have demonstrated that one mechanism whereby SR proteins function in splicing is to mediate specific protein-protein interactions among spliceosomal components and between general splicing factors and alternative splicing regulators (1, 1a, 6, 10, 27, 63, 74, 77). Such protein-protein interactions may play critical roles in splice site recognition and association (for reviews, see references 4, 18, 37, 41, 47 and 59). Specific interactions among the splicing factors also suggest that it is possible to identify new splicing factors by their interactions with known splicing factors.Here we report identification of a new splicing factor, Sip1, by its interaction with the essential splicing factor SC35. The predicted Sip1 protein sequence contains an RS domain and a region with sequence similarity to the Drosophila splicing regulator, SWAP. We have expressed and purified recombinant Sip1 protein and raised polyclonal antibodies against the recombinant Sip1 protein. The anti-Sip1 antibodies specifically recognize a protein migrating at a molecular mass of approximately 210 kDa in HeLa nuclear extract. The anti-Sip1 antibodies sufficiently deplete Sip1 protein from the nuclear extract, and the Sip1-depleted extract is inactive in pre-mRNA splicing. Addition of recombinant Sip1 protein can partially restore splicing activity to the Sip1-depleted nuclear extract, indicating an essential role of Sip1 in pre-mRNA splicing. Other RS domain-containing proteins, including SC35, ASF/SF2, and U2AF65, cannot substitute for Sip1 in reconstituting splicing activity of the Sip1-depleted nuclear extract. However, addition of U2AF65 further increases splicing activity of Sip1-reconstituted nuclear extract, suggesting that there may be a functional interaction between Sip1 and U2AF65 in nuclear extract.  相似文献   

7.
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   

8.
A decoding algorithm is tested that mechanistically models the progressive alignments that arise as the mRNA moves past the rRNA tail during translation elongation. Each of these alignments provides an opportunity for hybridization between the single-stranded, -terminal nucleotides of the 16S rRNA and the spatially accessible window of mRNA sequence, from which a free energy value can be calculated. Using this algorithm we show that a periodic, energetic pattern of frequency 1/3 is revealed. This periodic signal exists in the majority of coding regions of eubacterial genes, but not in the non-coding regions encoding the 16S and 23S rRNAs. Signal analysis reveals that the population of coding regions of each bacterial species has a mean phase that is correlated in a statistically significant way with species () content. These results suggest that the periodic signal could function as a synchronization signal for the maintenance of reading frame and that codon usage provides a mechanism for manipulation of signal phase.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]  相似文献   

9.
10.
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   

11.
12.
13.
A Boolean network is a model used to study the interactions between different genes in genetic regulatory networks. In this paper, we present several algorithms using gene ordering and feedback vertex sets to identify singleton attractors and small attractors in Boolean networks. We analyze the average case time complexities of some of the proposed algorithms. For instance, it is shown that the outdegree-based ordering algorithm for finding singleton attractors works in time for , which is much faster than the naive time algorithm, where is the number of genes and is the maximum indegree. We performed extensive computational experiments on these algorithms, which resulted in good agreement with theoretical results. In contrast, we give a simple and complete proof for showing that finding an attractor with the shortest period is NP-hard.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]  相似文献   

14.
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   

15.
16.
The mechanisms of free fatty acid-induced lipoapoptosis are incompletely understood. Here we demonstrate that Mcl-1, an anti-apoptotic member of the Bcl-2 family, was rapidly degraded in hepatocytes in response to palmitate and stearate by a proteasome-dependent pathway. Overexpression of a ubiquitin-resistant Mcl-1 mutant in Huh-7 cells attenuated palmitate-mediated Mcl-1 loss and lipoapoptosis; conversely, short hairpin RNA-targeted knockdown of Mcl-1 sensitized these cells to lipoapoptosis. Palmitate-induced Mcl-1 degradation was attenuated by the novel protein kinase C (PKC) inhibitor rottlerin. Of the two human novel PKC isozymes, PKCδ and PKCθ, only activation of PKCθ was observed by phospho-immunoblot analysis. As compared with Jurkat cells, a smaller PKCθ polypeptide and mRNA were expressed in hepatocytes consistent with an alternative splice variant. Short hairpin RNA-mediated knockdown of PKCθ reduced Mcl-1 degradation and lipoapoptosis. Likewise, genetic deletion of Pkcθ also attenuated Mcl-1 degradation and cytotoxicity by palmitate in primary hepatocytes. During treatment with palmitate, rottlerin inhibited phosphorylation of Mcl-1 at Ser159, a phosphorylation site previously implicated in Mcl-1 turnover. Consistent with these results, an Mcl-1 S159A mutant was resistant to degradation and improved cell survival during palmitate treatment. Collectively, these results implicate PKCθ-dependent destabilization of Mcl-1 as a mechanism contributing to hepatocyte lipoapoptosis.Current evidence suggests that hepatic steatosis is present in up to 30% of the American population (1). A subset of these individuals develop severe hepatic lipotoxicity, a syndrome referred to as NASH2 (2), which can progress to cirrhosis and its chronic sequela (3, 4). A major risk factor for hepatic lipotoxicity is insulin resistance (57), resulting in excessive lipolysis within peripheral adipose tissue with release of high levels of free fatty acids (FFA) to the circulation. Circulating FFA are taken up by the liver via fatty acid transporter 5 and CD36 (810), and the bulk of hepatic neutral fat is derived from re-esterification of circulating FFA (8). Current concepts indicate that FFA, and not their esterified product (triglyceride), mediate hepatic lipotoxicity (11, 12). Elevated serum FFA correlate with liver disease severity (1315), and therapies that enhance insulin sensitivity ameliorate hepatic lipotoxicity, in part, by decreasing plasma FFA (16). Hepatic FFA also accumulate in experimental steatohepatitis, further supporting a role for these nutrients in hepatic lipotoxicity (17). Saturated FFA are more strongly implicated in hepatic lipotoxicity than unsaturated FFA (18, 19). Saturated FFA induce hepatocyte apoptosis (20, 21), a cardinal feature of nonalcoholic fatty liver disease (22), and serum biomarkers of apoptosis are useful for identifying hepatic lipotoxicity (23). Thus, FFA-mediated lipotoxicity occurs, in part, by apoptosis.Apoptosis is regulated by members of the Bcl-2 protein family (24). These proteins can be categorized into three subsets as follows: the guardians or anti-apoptotic members of this family, which include Bcl-2, A1, Mcl-1, Bcl-xL, and Bcl-w; the multidomain executioners or proapoptotic members of this family, which include Bax and Bak; and the messengers or biosensors of cell death, which share only the third Bcl-2 homology domain and are referred to as BH3-only proteins. This last group of proteins includes Bid, Bim, Bmf, Puma, Noxa, Hrk, Bad, and Bik. We have previously reported that cytotoxic FFA induce Bim expression by a FoxO3a-dependent mechanism that contributes, in part, to lipoapoptosis by activating Bax (20, 21). However, Bax activation can be held in check by anti-apoptotic members of the Bcl-2 family suggesting their function may also be dysregulated during FFA-mediated cytotoxicity.Bcl-2 is not expressed in hepatocytes at the protein level (25), whereas Bcl-w and Bfl-1/A1 knock-out mice have no liver phenotype (2628). However, both potent anti-apoptotic proteins Bcl-xL and Mcl-1 are expressed by hepatocytes and exhibit a liver phenotype in knock-out mice (29, 30), whereas up-regulation of Mcl-1 renders hepatocytes resistant to apoptosis (3133). It has also been posited that cellular elimination of Mcl-1 is a critical step in certain proapoptotic cascades (34, 35). Mcl-1 is unique among Bcl-2 proteins in that it has a short half-life, 30–120 min in most cell types, due to the presence of two sequences rich in proline, glutamic acid, serine, and threonine, which target the protein for rapid degradation by the proteasome (36). Proteasomal degradation of Mcl-1 is promoted by ubiquitination, which in turn is regulated by various kinase cascades (36). Despite its potential importance, a role for Mcl-1 in regulating hepatocyte FFA-mediated lipoapoptosis remains unexplored.Given that FFA induce insulin resistance (37), the kinases potentially regulating lipoapoptosis are likely those also identified in insulin resistance syndromes, especially the novel PKC isoforms PKCδ and PKCθ (38). The novel PKC isoforms are activated by diacylglycerol, which rises in the presence of FFA (3941), and diacylglycerol levels are significantly increased in NASH (42). A role for PKCδ in apoptosis has not been described. PKCθ has recently been shown to be activated by endoplasmic reticulum stress in liver cells (43) and lipids in vivo (44, 45). Furthermore, PKCθ has also been implicated in apoptosis of Jurkat cells, neuroblastoma cells, and myeloid leukemia cells (46, 47). However, neither its role in mediating lipoapoptosis nor modulating levels/activity of Bcl-2 proteins has been examined.This study addresses the role of Mcl-1 and PKCθ in FFA-induced lipoapoptosis. We identify a pathway that involves PKCθ-dependent proteasomal degradation of Mcl-1. Using inhibitors of various steps along this pathway, along with Mcl-1 mutants that are resistant to proteasomal degradation or Ser159 phosphorylation, our studies implicate Mcl-1 degradation via a PKCθ-dependent process as a critical step in lipoapoptosis.  相似文献   

17.
18.
Dishevelled (Dvl) is an essential protein in the Wnt signaling pathways; it uses its PDZ domain to transduce the Wnt signals from the membrane receptor Frizzled to downstream components. Here, we report identifying a drug-like small molecule compound through structure-based ligand screening and NMR spectroscopy and show the compound to interact at low micromolar affinity with the PDZ domain of Dvl. In a Xenopus testing system, the compound could permeate the cell membrane and block the Wnt signaling pathways. In addition, the compound inhibited Wnt signaling and reduced the levels of apoptosis in the hyaloid vessels of eye. Moreover, this compound also suppressed the growth of prostate cancer PC-3 cells. These biological effects suggest that by blocking the PDZ domain of Dvl, the compound identified in our studies effectively inhibits the Wnt signaling and thus provides a useful tool for studies dissecting the Wnt signaling pathways.The Wnt signaling pathways are regulated by a family of secreted Wnt glycoproteins. The canonical Wnt pathway, which is highly conserved, is best understood. In this pathway, Wnt molecules interact with the seven-transmembrane Frizzled (Fz)2 proteins (1) by binding to an N-terminal cysteine-rich-domain (2). The signal is then transduced into the cell through an internal sequence of Fz, C-terminal to the seventh transmembrane domain, which binds directly to the PDZ (postsynaptic density-95/discs large/zonula occludens-1) domain of the cytoplasmic protein Dishevelled (Dvl) (3). Dvl then transduces the Wnt signals to downstream components (4). Three Dvl homologs (Dvl-1, -2, and -3) have been identified in humans; all are expressed in both embryonic and adult tissues, including brain, heart, lung, kidney, skeletal muscle, and others (4). Up-regulation and overexpression of Dvl proteins have been reported in many cancers, including those of breast, colon, prostate, mesothelium, and lung (non-small cell) (58).The Dvl protein is made up of three conserved domains: an N-terminal DIX domain, a central PDZ domain, and a C-terminal DEP domain (9). The central PDZ domain is of particular interest because of its interaction with Fz and other Wnt pathway proteins (3, 10). The direct interaction between the PDZ domain and Fz peptides is relatively weak, and other factors may play a role to ensure the communication between the two molecules (3). For example, several studies suggest that the DEP domain of Dvl has a membrane-targeting function that may facilitate PDZ-Fz interaction (1114). However, the weak PDZ-Fz interaction provides an opportunity to block Wnt signaling at the Dvl level by using a small molecule inhibitor. An earlier study in our laboratories used an NMR-assisted virtual ligand screening approach to identify a peptide mimic that can bind to the Dvl PDZ domain (15). We have now used an improved algorithm to conduct an additional structure-based virtual screen of the PDZ domain of Dvl and have discovered a group of drug-like compounds that bind to the PDZ domain with moderate to low micromolar affinity. One of these compounds effectively blocked Wnt signaling in vivo and reduced the growth rate of a prostate cancer cell line.  相似文献   

19.
20.
The kinetochore, which consists of centromere DNA and structural proteins, is essential for proper chromosome segregation in eukaryotes. In budding yeast, Sgt1 and Hsp90 are required for the binding of Skp1 to Ctf13 (a component of the core kinetochore complex CBF3) and therefore for the assembly of CBF3. We have previously shown that Sgt1 dimerization is important for this kinetochore assembly mechanism. In this study, we report that protein kinase CK2 phosphorylates Ser361 on Sgt1, and this phosphorylation inhibits Sgt1 dimerization.The kinetochore is a structural protein complex located in the centromeric region of the chromosome coupled to spindle microtubules (1, 2). The kinetochore generates a signal to arrest cells during mitosis when it is not properly attached to microtubules, thereby preventing chromosome missegregation, which can lead to aneuploidy (3, 4). The molecular structure of the kinetochore complex of the budding yeast Saccharomyces cerevisiae has been well characterized; it is composed of more than 70 proteins, many of which are conserved in mammals (2).The centromere DNA in the budding yeast is a 125-bp region that contains three conserved regions, CDEI, CDEII, and CDEIII (5, 6). CDEIII (25 bp) is essential for centromere function (7) and is bound to a key component of the centromere, the CBF3 complex. The CBF3 complex contains four proteins, Ndc10, Cep3, Ctf13 (815), and Skp1 (14, 15), all essential for viability. Mutations in any of the CBF3 proteins abolish the ability of CDEIII to bind to CBF3 (16, 17). All of the kinetochore proteins, except the CDEI-binding Cbf1 (1820), localize to the kinetochores in a CBF3-dependent manner (2). Thus, CBF3 is a fundamental kinetochore complex, and its mechanism of assembly is of great interest.We have previously found that Sgt1 and Skp1 activate Ctf13; thus, they are required for assembly of the CBF3 complex (21). The molecular chaperone Hsp90 is also required to form the active Ctf13-Skp1 complex (22). Sgt1 has two highly conserved motifs that are required for protein-protein interaction: the tetratricopeptide repeat (21) and the CHORD protein and Sgt1-specific motif. We and others have found that both domains are important for the interaction of Sgt1 with Hsp90 (2326), which is required for assembly of the core kinetochore complex. This interaction is an initial step in kinetochore activation (24, 26, 27), which is conserved between yeast and humans (28, 29).We have recently shown that Sgt1 dimerization is important for Sgt1-Skp1 binding and therefore for kinetochore assembly (30). In this study, we have found that protein kinase CK2 phosphorylates Sgt1 at Ser361, and this phosphorylation inhibits Sgt1 dimerization. Therefore, CK2 appears to regulate kinetochore assembly negatively in budding yeast.  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