首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
This paper presents evidence that a member of the L1 family of ankyrin-binding cell adhesion molecules is a substrate for protein tyrosine kinase(s) and phosphatase(s), identifies the highly conserved FIGQY tyrosine in the cytoplasmic domain as the principal site of phosphorylation, and demonstrates that phosphorylation of the FIGQY tyrosine abolishes ankyrin-binding activity. Neurofascin expressed in neuroblastoma cells is subject to tyrosine phosphorylation after activation of tyrosine kinases by NGF or bFGF or inactivation of tyrosine phosphatases with vanadate or dephostatin. Furthermore, both neurofascin and the related molecule Nr-CAM are tyrosine phosphorylated in a developmentally regulated pattern in rat brain. The FIGQY sequence is present in the cytoplasmic domains of all members of the L1 family of neural cell adhesion molecules. Phosphorylation of the FIGQY tyrosine abolishes ankyrin binding, as determined by coimmunoprecipitation of endogenous ankyrin and in vitro ankyrin-binding assays. Measurements of fluorescence recovery after photobleaching demonstrate that phosphorylation of the FIGQY tyrosine also increases lateral mobility of neurofascin expressed in neuroblastoma cells to the same extent as removal of the cytoplasmic domain. Ankyrin binding, therefore, appears to regulate the dynamic behavior of neurofascin and is the target for regulation by tyrosine phosphorylation in response to external signals. These findings suggest that tyrosine phosphorylation at the FIGQY site represents a highly conserved mechanism, used by the entire class of L1-related cell adhesion molecules, for regulation of ankyrin-dependent connections to the spectrin skeleton.Vertebrate L1, neurofascin, neuroglial cell adhesion molecule (Ng-CAM),1 Ng-CAM–related cell adhesion molecule (Nr-CAM), and Drosophila neuroglian are members of a family of nervous system cell adhesion molecules that possess variable extracellular domains comprised of Ig and fibronectin type III domains and a relatively conserved cytoplasmic domain (Grumet, 1991; Hortsch and Goodman, 1991; Rathgen and Jessel, 1991; Sonderegger and Rathgen, 1992; Hortsch, 1996). Members of this family, including a number of alternatively spliced forms, are abundant in the nervous system during early development as well as in adults. Neurofascin and Nr-CAM, for example, constitute ∼0.5% of the total membrane protein in adult brain (Davis et al., 1993; Davis and Bennett, 1994). Cellular functions attributed to the L1 family include axon fasciculation (Stallcup and Beasley, 1985; Landmesser et al., 1988; Brummendorf and Rathjen, 1993; Bastmeyer et al., 1995; Itoh et al., 1995; Magyar-Lehmann et al., 1995), axonal guidance (van den Pol and Kim, 1993; Liljelund et al., 1994; Brittis and Silver, 1995; Brittis et al., 1995; Lochter et al., 1995; Wong et al., 1996), neurite extension (Chang et al., 1987; Felsenfeld et al., 1994; Hankin and Lagenaur, 1994; Ignelzi et al., 1994; Williams et al., 1994a ,b,c,d; Doherty et al., 1995; Zhao and Siu, 1995), a role in long term potentiation (Luthl et al., 1994), synaptogenesis (Itoh et al., 1995), and myelination (Wood et al., 1990). The potential clinical importance of this group of proteins has been emphasized by the findings that mutations in the L1 gene on the X chromosome are responsible for developmental anomalies including hydrocephalus and mental retardation (Rosenthal et al., 1992; Jouet et al., 1994; Wong et al., 1995).The conserved cytoplasmic domains of L1 family members include a binding site for the membrane skeletal protein ankyrin. This interaction was first described for neurofascin (Davis et. al., 1993) and subsequently has been observed for L1, Nr-CAM (Davis and Bennett, 1994), and Drosophila neuroglian (Dubreuil et al., 1996). The membrane-binding domain of ankyrin contains two distinct sites for neurofascin and has the potential to promote lateral association of neurofascin and presumably other L1 family members (Michaely and Bennett, 1995). Nodes of Ranvier are physiologically relevant axonal sites where ankyrin and L1 family members collaborate, based on findings of colocalization of a specialized isoform of ankyrin with alternatively spliced forms of neurofascin and NrCAM in adults (Davis et al., 1996) as well as in early axonal developmental intermediates (Lambert, S., J. Davis, P. Michael, and V. Bennett. 1995. Mol. Biol. Cell. 6:98a).L1, after homophilic and/or heterophilic binding, participates in signal transduction pathways that ultimately are associated with neurite extension and outgrowth (Ignelzi et al., 1994; Williams et al., 1994a ,b,c,d; Doherty et al., 1995). L1 copurifies with a serine–threonine protein kinase (Sadoul et al., 1989) and is phosphorylated on a serine residue that is not conserved among other family members (Wong et al., 1996). L1 pathway(s) may also involve G proteins, calcium channels, and tyrosine phosphorylation (Williams et al., 1994a ,b,c,d; Doherty et al., 1995). After homophilic interactions, L1 directly activates a tyrosine signaling cascade after a lateral association of its ectodomain with the fibroblast growth factor receptor (Doherty et al., 1995). Antibodies against L1 have also been shown to activate protein tyrosine phosphatase activity in growth cones (Klinz et al., 1995). However, details of the downstream substrates of L1-promoted phosphorylation and dephosphorylation and possible roles of the cytoplasmic domain are not known.Tyrosine phosphorylation is well established to modulate cell–cell and cell–extracellular matrix interactions involving integrins and their associated proteins (Akiyama et al., 1994; Arroyo et al., 1994; Schlaepfer et al., 1994; Law et al., 1996) as well as the cadherins (Balsamo et al., 1996; Krypta et al., 1996; Brady-Kalnay et al., 1995; Shibamoto et al., 1995; Hoschuetzky et al., 1994; Matsuyoshi et al., 1992). For example, the adhesive functions of the calciumdependent cadherin cell adhesion molecule are mediated by a dynamic balance between tyrosine phosphorylation of β-catenin by TrkA and dephosphorylation via the LARtype protein tyrosine phosphatase (Krypta et al., 1996). In this example the regulation of binding among the structural proteins is the result of a coordination between classes of protein kinases and protein phosphatases.This study presents evidence that neurofascin, expressed in a rat neuroblastoma cell line, is a substrate for both tyrosine kinases and protein tyrosine phosphatases at a tyrosine residue conserved among all members of the L1 family. Site-specific tyrosine phosphorylation promoted by both tyrosine kinase activators (NGF and bFGF) and protein tyrosine phosphatase inhibitors (dephostatin and vanadate) is a strong negative regulator of the neurofascin– ankyrin binding interaction and modulates the membrane dynamic behavior of neurofascin. Furthermore, neurofascin and, to a lesser extent Nr-CAM, are also shown here to be tyrosine phosphorylated in developing rat brain, implying a physiological relevance to this phenomenon. These results indicate that neurofascin may be a target for the coordinate control over phosphorylation that is elicited by protein kinases and phosphatases during in vivo tyrosine phosphorylation cascades. The consequent decrease in ankyrin-binding capacity due to phosphorylation of neurofascin could represent a general mechanism among the L1 family members for regulation of membrane–cytoskeletal interactions in both developing and adult nervous systems.  相似文献   

2.
Human FGF1 (fibroblast growth factor 1) is a powerful signaling molecule with a short half-life in vivo and a denaturation temperature close to physiological. Binding to heparin increases the stability of FGF1 and is believed to be important in the formation of FGF1·fibroblast growth factor receptor (FGFR) active complex. In order to reveal the function of heparin in FGF1·FGFR complex formation and signaling, we constructed several FGF1 variants with reduced affinity for heparin and with diverse stability. We determined their biophysical properties and biological activities as well as their ability to translocate across cellular membranes. Our study showed that increased thermodynamic stability of FGF1 nicely compensates for decreased binding of heparin in FGFR activation, induction of DNA synthesis, and cell proliferation. By stepwise introduction of stabilizing mutations into the K118E (K132E) FGF1 variant that shows reduced affinity for heparin and is inactive in stimulation of DNA synthesis, we were able to restore the full mitogenic activity of this mutant. Our results indicate that the main role of heparin in FGF-induced signaling is to protect this naturally unstable protein against heat and/or proteolytic degradation and that heparin is not essential for a direct FGF1-FGFR interaction and receptor activation.FGF1 (fibroblast growth factor 1) belongs to a family of polypeptide growth factors comprising in humans 22 structurally related proteins (1, 2). The signaling induced by the growth factor leads to a wide range of cellular responses during development as well as in adult life, such as growth regulation, differentiation, survival, stress response, migration, and proliferation of different cell types (3). The biological activity of FGF1 is exerted through binding to four high affinity cell surface receptors (FGFR1–4), resulting in receptor dimerization and transphosphorylation in its tyrosine kinase domain (4, 5). The activated FGFR3 induces cellular response by initiating several signaling cascades, including mitogen-activated protein kinase (MAPK), phosphoinositide 3-kinase/Akt, and phospholipase C-γ (PLC-γ) pathways (6).In addition to FGFRs, FGF1 binds to heparan sulfates (HS) associated with proteoglycans at the cell surface and in the extracellular matrix (7). Among the physiological sugars, the highest affinity for FGF1 is shown by heparin, a widely used linear, highly sulfated polysaccharide composed of 2-O-sulfated iduronic acid and 6-O-sulfated, N-sulfated glucosamine units (8).Despite many years of research, there is still controversy regarding the molecular role of heparin/HS in FGF1- and FGF2-induced signaling. Thus, the question of whether or not the linkage of two molecules of the growth factor by heparin/HS is an absolute prerequisite for induction of FGFR dimerization is still open. Numerous studies have concluded that the presence of heparin/HS is obligatory for FGF signaling. It is widely believed that heparin/HS is directly involved in receptor dimerization and is critical for mitogenic response stimulated by the growth factor (4, 6, 810).On the other hand, several authors working on FGF1 and FGF2 have suggested that there is no mandatory requirement for heparin for the assembly and activation of the FGF·FGFR complex. They imply that heparin only plays a role in association of two molecules of the growth factor and therefore facilitates their binding to FGFR (11). It has been reported that FGF1 and FGF2 can interact with the FGFR and trigger phosphorylation of p42/44 MAPK and activation of other signaling pathways even in the absence of HS (1216).The accepted role of heparin/HS in FGF1 signaling is to prevent the degradation of the growth factor (17). The interaction with heparin or HS protects FGF1 against heat, acidic pH, and proteases (18, 19). HS also seems to regulate the activity of different FGFs by creating their local reservoir and generating a concentration gradient of the growth factor (6, 17).The binding of FGF1 to heparin/HS is mediated by specific residues forming a positively charged patch on the protein surface (20, 21). The major contribution is made by Lys118 (Lys132 in the full-length numbering system), which was identified by Harper and Lobb (22), and Lys112 and Arg122 (23, 24). Additional residues of FGF1 involved in the interaction with heparin are the positively charged Lys113, Arg119, and Lys128 and the polar Asn18, Asn114, and Gln127 (20, 21). Site-directed mutagenesis and other studies have revealed the importance of Lys118 not only in heparin binding but also for the biological function of FGF1 (22, 25, 26). It was shown that the K118E (K132E) mutant is inactive in stimulation of DNA synthesis, although its affinity for FGFR and the ability to activate signaling cascades is not reduced (27, 28). Despite extensive research, the reason for the lack of mitogenic potential of K118E FGF1 is still not clear.In this paper, we verified the function of heparin in FGF1·FGFR complex formation and signaling by constructing several FGF1 mutants with reduced affinity for heparin. To recover the stability of these variants, which could no longer be stabilized by heparin, we supplemented them stepwise with stabilizing mutations (29). We analyzed thoroughly their biological activity and their ability to translocate across cellular membranes (3034). Interestingly, the full mitogenic activity of the K118E FGF1 variant was restored by the introduced stabilizing mutations.Our results indicate that the main role of heparin in FGF-induced signaling is to protect this naturally unstable protein against heat denaturation and proteolytic degradation and that the increased stability of the growth factor can compensate for reduced heparin binding.  相似文献   

3.
Activation of fibroblast growth factor (FGF) signaling is initiated by a multiprotein complex formation between FGF, FGF receptor (FGFR), and heparan sulfate proteoglycan on the cell membrane. Cross-talk with other factors could affect this complex assembly and modulate the biological response of cells to FGF. We have previously demonstrated that anosmin-1, a glycosylated extracellular matrix protein, interacts with the FGFR1 signaling complex and enhances its activity in an IIIc isoform-specific and HS-dependent manner. The molecular mechanism of anosmin-1 action on FGFR1 signaling, however, remains unknown. Here, we show that anosmin-1 directly binds to FGFR1 with high affinity. This interaction involves domains in the N terminus of anosmin-1 (cysteine-rich region, whey acidic protein-like domain and the first fibronectin type III domain) and the D2–D3 extracellular domains of FGFR1. In contrast, anosmin-1 binds to FGFR2IIIc with much lower affinity and displays negligible binding to FGFR3IIIc. We also show that FGFR1-bound anosmin-1, although capable of binding to FGF2 alone, cannot bind to a FGF2·heparin complex, thus preventing FGFR1·FGF2·heparin complex formation. By contrast, heparin-bound anosmin-1 binds to pre-formed FGF2·FGFR1 complex, generating an anosmin-1·FGFR1·FGF2·heparin complex. Furthermore, a functional interaction between anosmin-1 and the FGFR1 signaling complex is demonstrated by immunofluorescence co-localization and Transwell migration assays where anosmin-1 was shown to induce opposing effects during chemotaxis of human neuronal cells. Our study provides molecular and cellular evidence for a modulatory action of anosmin-1 on FGFR1 signaling, whereby binding of anosmin-1 to FGFR1 and heparin can play a dual role in assembly and activity of the ternary FGFR1·FGF2·heparin complex.FGF5 signaling plays an important role in a wide range of fundamental biological responses (13). Both FGF and FGFR bind to heparan sulfate (HS) and heparin, a highly sulfated type of HS produced in connective tissue mast cells. Heparan sulfate proteoglycans (HSPG) are the cell surface co-receptors essential for the formation of functional FGF·FGFR signaling complex (4, 5). There are four structurally related FGFRs (FGFR1–4), which consist of an extracellular ligand-binding region containing three immunoglobulin (Ig)-like domains (D1–D3), a single transmembrane domain, and a cytoplasmic domain with protein-tyrosine kinase catalytic activity. The 22 members of the FGF family bind to the interface formed by the D2/D3 domains and the linker between these domains (6, 7), whereas a conserved positively charged region in D2 serves as the HS binding site (8). An unusual stretch of seven to eight acidic residues designated as the “acid box” is present in the linker connecting D1 and D2. Alternative splicing events occur to generate various isoforms, including a truncated receptor lacking D1 and the D1–D2 linker or a full-length receptor that differs in the second half of D3, designated as IIIb and IIIc isoforms (5). Two crystal structures have been proposed to demonstrate how the FGF·FGFR·heparin complex is assembled (9, 10). Recent evidence suggests that both may be biologically relevant (11, 12).The diversity of FGF signaling pathways and consequent biological functions require that activation of FGFR should be tightly regulated. Such regulation can occur either at the level of the extracellular receptor-ligand complex assembly or via intracellular modulation of downstream effectors (13). Extracellular regulation mainly involves the interaction between each component of the FGF·FGFR·HS signaling complex. For example, FGF8 is shown to bind mostly to the FGFR IIIc isoforms, whereas FGF7 acts as the preferential ligand for the FGFR2 IIIb isoform (13, 14). Sequence specificity, length, and sulfation patterns of HS are also important regulators of the FGF·FGFR interaction (15, 16).Cell surface proteins other than FGFs and HSPGs participate in FGFR signaling regulation. FLRT3 (a member of the fibronectin-leucine-rich transmembrane protein family) promotes FGF signaling and interacts with FGFR1 and FGFR4 via its extracellular fibronectin type III (FnIII) domain (17). Sef (similar expression to fgf genes) functions as an antagonist of FGF signaling in zebrafish. The two FnIII regions of Sef are essential for its function and interaction with FGFR1 and FGFR2 (18). Neuronal cell adhesion molecule (NCAM), N-cadherin, and L1 have also been identified as functionally relevant in FGFR-mediated neurite outgrowth (1922). The FnIII domains of NCAM bind to the D2 and D3 domains of FGFR1 (19) and FGFR2 (23) to induce ligand-independent receptor phosphorylation.Anosmin-1, an extracellular matrix-associated glycosylated protein, appears to be a novel member of the extracellular FGFR signaling modulators (24, 25). Loss-of-function mutations of anosmin-1 and FGFR1 are associated with Kallmann syndrome (KS), underlying X-linked, and autosomal dominant/recessive inheritance mode, respectively (2628). KS is a human developmental genetic disorder characterized by loss of sense of smell (anosmia) caused by abnormal olfactory bulb development and delayed, even arrested puberty caused by disrupted migration of the gonadotropin-releasing hormone (GnRH)-secreting neuron. We previously reported that anosmin-1 acts as an FGFR1IIIc isoform-specific co-ligand, which enhances signaling activity. In human embryonic GnRH olfactory neuroblast FNC-B4 cells, anosmin-1 induced neurite outgrowth and cytoskeletal rearrangements through FGFR1-dependent mechanisms involving p42/44 and p38 mitogen-activated protein kinases and Cdc42/Rac1 activation (25). A functional interaction is also demonstrable between anosmin-1 and FGFR1 in optic nerve oligodendrocyte precursor development (24). Structurally, anosmin-1 comprises an N-terminal cysteine-rich domain (CR) and a whey acidic protein-like (WAP) domain, followed by four tandem FnIII repeats and a C-terminal histidine rich region (Fig. 1a). Current evidence suggests that anosmin-1 functions by affecting FGF2-induced activation of FGFR1 signaling rather than by directly stimulating the receptor. However, the precise molecular mechanism of this interaction remains unclear.Open in a separate windowFIGURE 1.Generation of recombinant anosmin-1, anosmin-1 mutants, FGFR1D1D3, and FGFR1D2D3 proteins. a, the schematic structures of recombinant proteins of anosmin-1 and FGFR1. Each domain in the wild type (PIWF4), point mutants (mPIWF4N267K, mPIWF4E514K, and mPIWF4F517L), and truncated (PIWF1, PIWF2, and PIF4) anosmin-1 protein analogues are represented by a shaded rectangle. V5 and 6His epitopes at the C terminus are represented by a clear rectangle. Each immunoglobulin-like domain in the full ectodomain (FGFR1D1D3) and truncated form (FGFR1D2D3) of FGFR1 is represented by a half circle. The acid box (AB) is represented by a filled rectangle. H, histidine-rich region. b, 0.5–1 μg of purified recombinant proteins are loaded in each lane and visualized by colloidal blue staining. Molecular mass markers in kilodaltons are shown on the left.We now report for the first time that anosmin-1 directly binds to FGFR1 using surface plasmon resonance (SPR), chemical cross-linking, and immunofluorescence co-localization studies in living cells. This interaction occurs between the N-terminal CR, WAP, and the first FnIII domain of anosmin-1 and D2 and D3 ectodomains of FGFR1. Moreover, SPR studies using sequential injections and Transwell migration assays in immortalized FNC-B4-hTERT cells suggest that anosmin-1 can have opposing effects in the formation and activation of the FGF2·FGFR1·heparin complex depending on the order of their binding interactions with anosmin-1.  相似文献   

4.
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   

5.
6.
7.
The polysialyltransferases ST8Sia II and ST8Sia IV polysialylate the glycans of a small subset of mammalian proteins. Their most abundant substrate is the neural cell adhesion molecule (NCAM). An acidic surface patch and a novel α-helix in the first fibronectin type III repeat of NCAM are required for the polysialylation of N-glycans on the adjacent immunoglobulin domain. Inspection of ST8Sia IV sequences revealed two conserved polybasic regions that might interact with the NCAM acidic patch or the growing polysialic acid chain. One is the previously identified polysialyltransferase domain (Nakata, D., Zhang, L., and Troy, F. A. (2006) Glycoconj. J. 23, 423–436). The second is a 35-amino acid polybasic region that contains seven basic residues and is equidistant from the large sialyl motif in both polysialyltransferases. We replaced these basic residues to evaluate their role in enzyme autopolysialylation and NCAM-specific polysialylation. We found that replacement of Arg276/Arg277 or Arg265 in the polysialyltransferase domain of ST8Sia IV decreased both NCAM polysialylation and autopolysialylation in parallel, suggesting that these residues are important for catalytic activity. In contrast, replacing Arg82/Arg93 in ST8Sia IV with alanine substantially decreased NCAM-specific polysialylation while only partially impacting autopolysialylation, suggesting that these residues may be particularly important for NCAM polysialylation. Two conserved negatively charged residues, Glu92 and Asp94, surround Arg93. Replacement of these residues with alanine largely inactivated ST8Sia IV, whereas reversing these residues enhanced enzyme autopolysialylation but significantly reduced NCAM polysialylation. In sum, we have identified selected amino acids in this conserved polysialyltransferase polybasic region that are critical for the protein-specific polysialylation of NCAM.Polysialic acid is a linear homopolymer of α2,8-linked sialic acid that is added to a small subset of mammalian glycoproteins by the polysialyltransferases (polySTs)3 ST8Sia II (STX) and ST8Sia IV (PST) (14). Substrates for the polySTs include the neural cell adhesion molecule (NCAM) (5, 6), the α-subunit of the voltage-dependent sodium channel (7, 8), CD36, a scavenger receptor found in milk (9), neuropilin-2 expressed by dendritic cells (10), and the polySTs themselves, which can polysialylate their own N-glycans in a process called autopolysialylation (11, 12). This small number of polysialylated proteins and other evidence from our laboratory (1315) suggest that polysialylation is a protein-specific modification that requires an initial protein-protein interaction between the polySTs and their glycoprotein substrates.The most abundant polysialylated protein is NCAM. The three major NCAM isoforms consist of five Ig domains, two fibronectin type III repeats, and a transmembrane domain and cytoplasmic tail (NCAM140 and NCAM180) or a glycosylphosphatidylinositol anchor (NCAM120) (16). Polysialylation takes place primarily on two N-linked glycans in the Ig5 domain (17). We have previously shown that a truncated NCAM140 protein consisting of Ig5, the first fibronectin type III repeat (FN1), the transmembrane region, and cytoplasmic tail is fully polysialylated (13). However, a protein consisting of Ig5, the transmembrane region, and cytoplasmic tail is not polysialylated (13). This suggests that the polySTs recognize and bind the FN1 domain to polysialylate N-glycans on the adjacent Ig5 domain. We subsequently identified an acidic patch unique to NCAM FN1, consisting of Asp497, Asp511, Glu512, and Glu514 (15).4 When three of these residues (Asp511, Glu512, and Glu514) are mutated to alanine or arginine, NCAM polysialylation is reduced or abolished, suggesting that the acidic patch is part of a larger recognition region. We anticipate that within this putative recognition region there will be amino acids required for mediating polyST-NCAM binding, and those that do not mediate binding per se but instead are required for correct positioning of the enzyme-substrate complex for polysialylation. For example, we have identified a novel α-helix in the FN1 domain that when replaced leads to polysialylation of O-glycans found on the FN1 domain rather than N-glycans on the Ig5 domain (14). This helix may mediate an interdomain interaction that positions the Ig5 N-glycans for polysialylation by an enzyme bound to the FN1 domain (14). Alternatively, the helix could act as a secondary interaction site that positions the polyST properly on the substrate.The expression of the polySTs is developmentally regulated with high levels of STX and moderate levels of PST expressed throughout the developing embryo (2, 18, 19). STX levels decline after birth, although PST expression persists in specific regions of the adult brain where polysialylated NCAM is involved in neuronal regeneration and synaptic plasticity (1823). The large size and negative charge of polysialic acid disrupt NCAM-dependent and NCAM-independent interactions, thereby negatively modulating cell adhesion (2426). Simultaneous disruption of both PST and STX in mice results in severe neuronal defects and death usually within 4 weeks after birth (27). Interestingly, when NCAM expression is also eliminated in these mice, they have a nearly normal phenotype, suggesting the main function of polysialic acid is to modulate NCAM-mediated cell adhesion during development (27). In addition, re-expression of highly polysialylated NCAM has been associated with several cancers, including neuroblastomas, gliomas, small cell lung carcinomas, and Wilms tumor. The presence of polysialic acid is thought to promote cancer cell growth and invasiveness (2835).Sialyltransferases, including the polySTs, have three motifs required for catalytic activity (3638) (see Fig. 1A). Sialyl motif Large (SML) is thought to bind the donor substrate CMP-sialic acid (39), whereas sialyl motif Small (SMS) is believed to bind both donor and carbohydrate acceptor substrates (40). The sialyl motif Very Small (SMVS) has a conserved His residue that is required for catalytic activity (38, 41). However, the precise function of this motif is unknown. An additional 4-amino acid motif, motif III, is conserved in the sialyltransferases (4244). It was suggested that this motif, and particularly His and Tyr residues within its sequence, may be required for optimal activity and acceptor recognition (42).Open in a separate windowFIGURE 1.PST and STX polybasic regions and mutants generated for this study. A, representation of the polySTs and their polybasic regions and sialyl motifs. The PBR is a 35-amino acid region present in both PST and STX, equidistant from the SML of each enzyme and rich in conserved positively charged amino acids. The PSTD is a region identified by Nakata et al. (47) that is 32 amino acids in length, rich in basic residues, and contiguous with the SMS of the enzymes. The sialyl motifs (SML, SMS, SMVS, and motif III) are regions of homology found in all sialyltransferases that are believed to be involved in substrate and donor interactions. B, PSTD of PST and the mutants made in this region that are used in this study. C, PBR of PST and STX and the mutants made in this region that are used in this study.Angata et al. (45) used chimeric enzymes to identify regions within the polySTs required for catalytic activity and NCAM polysialylation. Sequences from PST, STX, and ST8Sia III were used to construct the chimeric proteins. ST8Sia III is an α2,8-sialyltransferase that typically adds one or two sialic acid residues to glycoprotein or glycolipid substrates, can autopolysialylate its own glycans, but cannot polysialylate NCAM (46). Deletion analysis showed that amino acids 62–356 are required for PST catalytic activity. Replacement of segments of this region with corresponding STX or ST8Sia III sequences led to the suggestion that amino acids 62–127 and possibly 194–267 of PST may be required for NCAM recognition (45).Recently, Troy and co-workers (47, 48) identified a stretch of basic residues, termed the polysialyltransferase domain (PSTD), which is only observed in the two polySTs and not in other sialyltransferases. The PSTD is contiguous with SMS and extends from amino acids 246–277 in PST and 261–292 in STX. Mutation analysis demonstrated that the overall positive charge of this motif is important for activity and identified specific residues required for NCAM polysialylation (Arg252, Ile275, Lys276, and Arg277) (47).In this study, we have scanned the critical polyST regions identified by the work of Angata et al. (45) for sequences that may be involved in protein-protein recognition and NCAM polysialylation. We identified a second polybasic motif that we named the polybasic region (PBR). The PBR is conserved in PST and STX and is located equidistant from the SML of each enzyme. It consists of 35 amino acids of which 7 are the basic amino acids Arg and Lys. We found that the replacement of two specific residues within the PBR (Arg82 and Arg93 of PST and Arg97 and Lys108 of STX) have a greater negative effect on NCAM polysialylation than on autopolysialylation. Replacement of acidic residues surrounding PST Arg93 led to a similar disparate effect on these processes. Comparison of the critical residues in both the PSTD and PBR demonstrated that the replacement of PSTD residues had an equally negative impact on both NCAM polysialylation and enzyme autopolysialylation, whereas replacement of selected PBR residues more severely impacted NCAM polysialylation, suggesting that the PBR residues may play important roles in NCAM-specific polysialylation.  相似文献   

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

9.
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   

10.
11.
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   

12.
13.
14.
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.
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   

18.
19.
20.
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