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.
Fibronectin plays important roles in erythropoiesis through the fibronectin receptors VLA-4 and VLA-5. However, the substantial role of these fibronectin receptors and their functional assignment in erythroid differentiation are not yet fully understood. Here, we investigated the effects of cell adhesion to fibronectin on erythroid differentiation using K562 human erythroid progenitor cells. Erythroid differentiation could be induced in K562 cells in suspension by stimulating with hemin. This hemin-stimulated erythroid differentiation was highly accelerated when cells were induced to adhere to fibronectin by treatment with TNIIIA2, a peptide derived from tenascin-C, which has recently been found to induce β1-integrin activation. Another integrin activator, Mn2+, also accelerated hemin-stimulated erythroid differentiation. Adhesive interaction with fibronectin via VLA-4 as well as VLA-5 was responsible for acceleration of the hemin-stimulated erythroid differentiation in response to TNIIIA2, although K562 cells should have been lacking in VLA-4. Adhesion to fibronectin forced by TNIIIA2 causally induced VLA-4 expression in K562 cells, and this was blocked by the RGD peptide, an antagonist for VLA-5. The resulting adhesive interaction with fibronectin via VLA-4 strongly enhanced the hemin-stimulated activation of p38 mitogen-activated protein kinase, which was shown to serve as a signaling molecule crucial for erythroid differentiation. Suppression of VLA-4 expression by RNA interference abrogated acceleration of hemin-stimulated erythroid differentiation in response to TNIIIA2. Thus, VLA-4 and VLA-5 may contribute to erythropoiesis at different stages of erythroid differentiation.Hematopoietic stem and progenitor cells proliferate and differentiate in the bone marrow and fetal liver (16). Stromal cells of the bone marrow and fetal liver form a hematopoietic microenvironment called a “niche.” This microenvironment niche plays a crucial role in the regulation of the proliferation and differentiation of hematopoietic stem and progenitor cells. Besides humoral factors that include hematopoietic growth factors, adhesive interaction of hematopoietic stem and progenitor cells with stromal cells and/or the extracellular matrix (ECM)2 in the hematopoietic microenvironment is indispensable for hematopoietic development (16). The ECM in the hematopoietic microenvironment is composed of various macromolecules, such as fibronectin (FN), collagens, laminins, and proteoglycans. Among them, FN is one of the most important parts of the microenvironment niche (711). Also, in erythropoiesis, the importance of the adhesion of erythroid progenitors to FN via the FN receptors VLA-4 and VLA-5 has been reported (1116). However, the substantial role of these FN receptors and their functional assignment in erythroid differentiation are not yet fully understood.We previously found that FN, which provides scaffolding for the adhesion of various cell types, has an alternative functional site opposing cell adhesion (17). A 22-mer peptide derived from the 14th FN type III-like (FNIII) repeat of the FN molecule, termed FNIII14, strongly suppresses cell adhesion to FN by inhibiting the activation of β1-integrins including VLA-4 and VLA-5 (18, 19). Conversely, we have recently found that tenascin (TN)-C, which is an anti-adhesive ECM protein (20, 21), has a functional site for stimulating cell adhesion to FN (22). A 22-mer peptide derived from the FNIII repeat A2 in the TN-C molecule, termed TNIIIA2, can induce the conformational change necessary for functional activation of FN receptors through binding with syndecan-4 (22, 23). The active sites of FNIII14 and TNIIIA2 appear to be cryptic in the molecular structures of FN and TN-C but are exposed by conformational change through interaction with other ECM molecules or by processing with matrix metalloproteinase-2 (22, 24). Thus, these functional sites found in FN and TN-C molecules, which act in opposition to their parental ECM proteins, may act as a negative feedback loop for preventing excessive cellular responses to these ECM proteins in biological processes with ECM rearrangement. In any case, FNIII14 and TNIIIA2 enable us to control, either negatively or positively, the adhesion of various cell types to FN.Various hematopoietic progenitor cell lines have been used in in vitro studies of hematopoietic differentiation. However, most hematopoietic progenitor cell lines are nonadherent, because their cell surface β1-integrins, including FN receptors, have impaired ligand-binding activity (25, 26). Therefore, in order to investigate the role of cell adhesion to FN in hematopoietic differentiation, their FN receptors must be activated. Since TNIIIA2 can induce activation of FN receptors in various hematopoietic progenitor cell lines (22), this peptide factor may be useful for investigating the substantial role of cell adhesion to FN in hematopoietic differentiation. Here, we investigate the effects of cell adhesion to FN on erythroid differentiation using TNIIIA2 and Mn2+ as the integrin activator and the human erythroid progenitor cell line K562, which only expresses VLA-5, as the FN receptor (27). As a result, we show that hemin-stimulated erythroid differentiation of K562 cells is strongly enhanced when K562 cells are forced to adhere to FN. Sustained adhesion to FN via VLA-5, which is induced by TNIIIA2 or Mn2+, causes induction of VLA-4 expression. The resulting adhesive interaction with FN via newly expressed VLA-4 then generates a conspicuous increase in the hemin-stimulated phosphorylation/activation of p38 MAP kinase, which is shown to serve as a signaling molecule crucial for erythroid differentiation of K562 cells.  相似文献   

2.
The general stress response of Bacillus subtilis can be activated by stimuli such as the addition of salt or ethanol and with blue light. In the latter response, YtvA activates σB through a cascade of Rsb proteins, organized in stressosomes. YtvA is composed of an N-terminal LOV (light, oxygen, and voltage) domain and a C-terminal STAS (sulfate transporter and anti-sigma factor) domain and shows light-modulated GTP binding in vitro. Here, we examine the mechanism of YtvA-mediated activation of σB in vivo with site-directed mutagenesis. Constitutive off and constitutive on mutations have been identified. Disruption of GTP binding in the STAS domain eliminates light activation of σB. In contrast, modification of sites relevant for phosphorylation of STAS domains does not affect the stress response significantly. The data obtained are integrated into a model for the structure of full-length YtvA, which presumably functions as a dimer.LOV2 domains (1), members of the superfamily of PAS domains (2, 3), are abundant in all domains of life and were first identified in plant phototropins (4). These photoreceptors regulate stomatal opening, phototropism, etc. and contain two N-terminal LOV domains that confer light regulation on the C-terminal Ser/Thr kinase domain (4). They also occur in bacteria, in which YtvA from Bacillus subtilis has been best characterized (for a review, see e.g. Ref. 5). Its N-terminal LOV domain binds FMN and shows the typical LOV photochemistry (6, 7): covalent adduct formation between a cysteine and the FMN chromophore. A linker helix, denoted Jα (7), connects the LOV domain to a STAS domain. The latter domain is present in many regulators of the general stress response of B. subtilis (8, 9). Stress via the addition of salt or ethanol (for a review, see Ref. 10) and blue light (11, 12) activates the general stress response via the environmental pathway, which integrates various signals via a large multiprotein complex, called the stressosome (13, 14). YtvA, which mediates light activation of σB (11, 12, 15), co-purifies with other STAS domain proteins in the stressosomes (16).When cells are stressed, STAS domains of several stressosome proteins (e.g. RsbS and RsbR) are phosphorylated by another intrinsic stressosome component, the serine/threonine kinase RsbT (9, 14, 17, 18). Next, RsbT is released from the complex to trigger RsbU, a protein phosphatase, thus (indirectly) activating σB (19). Phosphorylation of YtvA, however, has never been detected. Rather, it has been demonstrated in vitro that YtvA shows light-dependent GTP binding, presumably at its NTP-binding site in the STAS domain (20).Little is known about the mechanism of signal transmission in and by YtvA, except that in the C62A mutant, photochemistry in vitro (12) and light activation of σB in vivo (12, 15) are abolished. More detailed information is available for LOV domains of phototropins. A conserved glutamine, which is in hydrogen-bonding contact with the isoalloxazine ring of FMN, rotates its side chain by 180° upon covalent adduct formation (21). Replacement of this residue by leucine in the LOV2 domain of Phy3 from Adiantum results in a considerable reduction of the light-induced structural change (22). The corresponding mutation in phototropin 1 from Arabidopsis impairs autophosphorylation activity (23). The signal generated in the LOV2 domain is transmitted to the downstream kinase domain of phototropin 1 of Avena sativa through disruption of the interaction between its central β-sheet and the C-terminal linker region, the Jα-helix (24).Here, we study the mechanism of activation of YtvA in vivo, i.e. light-induced activation of the σB response, with site-directed mutagenesis. We focus on three regions of the protein, the flavin-binding pocket, the β-sheet of the LOV domain, and the GTP-binding site, and on potential phosphorylation sites of the STAS domain. We demonstrate that light-activated GTP binding is crucial for functional YtvA. A computational approach was used to model the structure of full-length YtvA. The model suggests that light modulates accessibility of the GTP-binding site of the STAS domain of YtvA.  相似文献   

3.
4.
5.
6.
Helicobacter pylori infection of the human stomach is associated with disease-causing inflammation that elicits DNA damage in both bacterial and host cells. Bacteria must repair their DNA to persist. The H. pylori AddAB helicase-exonuclease is required for DNA repair and efficient stomach colonization. To dissect the role of each activity in DNA repair and infectivity, we altered the AddA and AddB nuclease (NUC) domains and the AddA helicase (HEL) domain by site-directed mutagenesis. Extracts of Escherichia coli expressing H. pylori addANUCB or addABNUC mutants unwound DNA but had approximately half of the exonuclease activity of wild-type AddAB; the addANUCBNUC double mutant lacked detectable nuclease activity but retained helicase activity. Extracts with AddAHELB lacked detectable helicase and nuclease activity. H. pylori with the single nuclease domain mutations were somewhat less sensitive to the DNA-damaging agent ciprofloxacin than the corresponding deletion mutant, suggesting that residual nuclease activity promotes limited DNA repair. The addANUC and addAHEL mutants colonized the stomach less efficiently than the wild type; addBNUC showed partial attenuation. E. coli ΔrecBCD expressing H. pylori addAB was recombination-deficient unless H. pylori recA was also expressed, suggesting a species-specific interaction between AddAB and RecA and also that H. pylori AddAB participates in both DNA repair and recombination. These results support a role for both the AddAB nuclease and helicase in DNA repair and promoting infectivity.Infection of the stomach with Helicobacter pylori causes a variety of diseases including gastritis, peptic ulcers, and gastric cancer (1). A central feature of the pathology of these conditions is the establishment of a chronic inflammatory response that acts both on the host and the infecting bacteria (2). Both epithelial (3, 4) and lymphoid (5, 6) cells in the gastric mucosa of infected individuals release DNA-damaging agents that can introduce double-stranded (ds)2 breaks into the bacterial chromosome (7). The ds breaks must be repaired for the bacteria to survive and establish chronic colonization of the stomach. Homologous recombination is required for the faithful repair of DNA damage and bacterial survival. Alteration of the expression of one of a series of cell surface proteins on H. pylori occurs by an apparent gene conversion of babA, the frequency of which is reduced in repair-deficient strains (8, 9). This change in the cell surface, which may allow H. pylori to evade the host immune response, is a second means by which recombination can promote efficient colonization of the stomach by H. pylori.The initiation or presynaptic steps of recombination at dsDNA breaks in most bacteria involves the coordinated action of nuclease and helicase activities provided by one of two multisubunit enzymes, the AddAB and RecBCD enzymes (10). Escherichia coli recBCD null mutants have reduced cell viability, are hypersensitive to DNA-damaging agents, and are homologous recombination-deficient (1114). Similarly, H. pylori addA and addB null mutants are hypersensitive to DNA-damaging agents, have reduced frequencies of babA gene conversion, and colonize the stomach of mice less efficiently than wild-type strains (8).The activities of RecBCD enzyme from E. coli (1519) and AddAB from H. pylori (8) or Bacillus subtilis (2023) indicate some common general features of the presynaptic steps of DNA repair. In the case of E. coli, repair begins when the RecBCD enzyme binds to a dsDNA end and unwinds the DNA using its ATP-dependent helicase activities (17, 24). Single-stranded (ss) DNA produced during unwinding, with or without accompanying nuclease, is coated with RecA protein (16, 25). This recombinogenic substrate engages in strand exchange with a homologous intact duplex to form a joint molecule. Joint molecules are thought to be converted into intact, recombinant DNA either by replication or by cutting and ligation of exchanged strands (26).Although the AddAB and RecBCD enzymes appear to play similar roles in promoting recombination and DNA repair, they differ in several ways. RecBCD is a heterotrimer, composed of one copy of the RecB, RecC, and RecD gene products (27), whereas AddAB has two subunits, encoded by the addA and addB genes (21, 28). The enzyme subunit(s) responsible for helicase activity can be inferred from the presence of conserved protein domains or the activity of purified proteins. AddA, RecB, and RecD are superfamily I helicases with six highly conserved helicase motifs, including the conserved Walker A box found in many enzymes that bind ATP (2932). A Walker A box is defined by the consensus sequence (G/A)XXGXGKT (X is any amino acid (29). RecBCD enzymes in which the conserved Lys in this motif is changed to Gln have a reduced affinity for ATP binding (33, 34) and altered helicase activity (17, 3537).A nuclease domain with the conserved amino acid sequence LDYK is found in RecB, AddA, AddB, and many other nucleases (38). The conserved Asp plays a role in Mg2+ binding at the active site; Mg2+ is required for nuclease activity (39). The recB1080 mutation, which changes codon 1080 from the conserved Asp in this motif to Ala, eliminates nuclease activity (39).We have recently shown that addA and addB deletion mutants are hypersensitive to DNA-damaging agents and impaired in colonization of the mouse stomach compared with wild-type strains (8). To determine the roles of the individual helicase and nuclease activities of H. pylori AddAB in DNA repair and infectivity, we used site-directed mutagenesis to inactivate the conserved nuclease domains of addA and addB and the conserved ATPase (helicase) domain of AddA. Here, we report that loss of the AddAB helicase is sufficient to impair H. pylori DNA repair and infectivity and, when the genes are expressed in E. coli, homologous recombination. AddAB retains partial activity in biochemical and genetic assays when either of the two nuclease domains is inactivated but loses all detectable nuclease activity when both domains are inactivated. Remarkably, H. pylori AddAB can produce recombinants in E. coli only in the presence of H. pylori RecA, suggesting a species-specific interaction in which AddAB facilitates the production of ssDNA-coated with RecA protein. Our results show that both the helicase and nuclease activities are required for the biological roles of H. pylori AddAB.  相似文献   

7.
8.
Human concentrative nucleoside transporter 3 (hCNT3) utilizes electrochemical gradients of both Na+ and H+ to accumulate pyrimidine and purine nucleosides within cells. We have employed radioisotope flux and electrophysiological techniques in combination with site-directed mutagenesis and heterologous expression in Xenopus oocytes to identify two conserved pore-lining glutamate residues (Glu-343 and Glu-519) with essential roles in hCNT3 Na+/nucleoside and H+/nucleoside cotransport. Mutation of Glu-343 and Glu-519 to aspartate, glutamine, and cysteine severely compromised hCNT3 transport function, and changes included altered nucleoside and cation activation kinetics (all mutants), loss or impairment of H+ dependence (all mutants), shift in Na+:nucleoside stoichiometry from 2:1 to 1:1 (E519C), complete loss of catalytic activity (E519Q) and, similar to the corresponding mutant in Na+-specific hCNT1, uncoupled Na+ currents (E343Q). Consistent with close-proximity integration of cation/solute-binding sites within a common cation/permeant translocation pore, mutation of Glu-343 and Glu-519 also altered hCNT3 nucleoside transport selectivity. Both residues were accessible to the external medium and inhibited by p-chloromercuribenzene sulfonate when converted to cysteine.Physiologic nucleosides and the majority of synthetic nucleoside analogs with antineoplastic and/or antiviral activity are hydrophilic molecules that require specialized plasma membrane nucleoside transporter (NT)3 proteins for transport into or out of cells (14). NT-mediated transport is required for nucleoside metabolism by salvage pathways and is a critical determinant of the pharmacologic actions of nucleoside drugs (36). By regulating adenosine availability to purinoreceptors, NTs also modulate a diverse array of physiological processes, including neurotransmission, immune responses, platelet aggregation, renal function, and coronary vasodilation (4, 6, 7). Two structurally unrelated NT families of integral membrane proteins exist in human and other mammalian cells and tissues as follows: the SLC28 concentrative nucleoside transporter (CNT) family and the SLC29 equilibrative nucleoside transporter (ENT) family (3, 4, 6, 8, 9). ENTs are normally present in most, possibly all, cell types (4, 6, 8). CNTs, in contrast, are found predominantly in intestinal and renal epithelia and other specialized cell types, where they have important roles in absorption, secretion, distribution, and elimination of nucleosides and nucleoside drugs (13, 5, 6, 9).The CNT protein family in humans is represented by three members, hCNT1, hCNT2, and hCNT3. Belonging to a CNT subfamily phylogenetically distinct from hCNT1/2, hCNT3 utilizes electrochemical gradients of both Na+ and H+ to accumulate a broad range of pyrimidine and purine nucleosides and nucleoside drugs within cells (10, 11). hCNT1 and hCNT2, in contrast, are Na+-specific and transport pyrimidine and purine nucleosides, respectively (1113). Together, hCNT1–3 account for the three major concentrative nucleoside transport processes of human and other mammalian cells. Nonmammalian members of the CNT protein family that have been characterized functionally include hfCNT, a second member of the CNT3 subfamily from the ancient marine prevertebrate the Pacific hagfish Eptatretus stouti (14), CeCNT3 from Caenorhabditis elegans (15), CaCNT from Candida albicans (16), and the bacterial nucleoside transporter NupC from Escherichia coli (17). hfCNT is Na+- but not H+-coupled, whereas CeCNT3, CaCNT, and NupC are exclusively H+-coupled. Na+:nucleoside coupling stoichiometries are 1:1 for hCNT1 and hCNT2 and 2:1 for hCNT3 and hfCNT3 (11, 14). H+:nucleoside coupling ratios for hCNT3 and CaCNT are 1:1 (11, 16).Although much progress has been made in molecular studies of ENT proteins (4, 6, 8), studies of structurally and functionally important regions and residues within the CNT protein family are still at an early stage. Topological investigations suggest that hCNT1–3 and other eukaryote CNT family members have a 13 (or possibly 15)-transmembrane helix (TM) architecture, and multiple alignments reveal strong sequence similarities within the C-terminal half of the proteins (18). Prokaryotic CNTs lack the first three TMs of their eukaryotic counterparts, and functional expression of N-terminally truncated human and rat CNT1 in Xenopus oocytes has established that these three TMs are not required for Na+-dependent uridine transport activity (18). Consistent with this finding, chimeric studies involving hCNT1 and hfCNT (14) and hCNT1 and hCNT3 (19) have demonstrated that residues involved in Na+- and H+-coupling reside in the C-terminal half of the protein. Present in this region of the transporter, but of unknown function, is a highly conserved (G/A)XKX3NEFVA(Y/M/F) motif common to all eukaryote and prokaryote CNTs.By virtue of their negative charge and consequent ability to interact directly with coupling cations and/or participate in cation-induced and other protein conformational transitions, glutamate and aspartate residues play key functional and structural roles in a broad spectrum of mammalian and bacterial cation-coupled transporters (2030). Little, however, is known about their role in CNTs. This study builds upon a recent mutagenesis study of conserved glutamate and aspartate residues in hCNT1 (31) to undertake a parallel in depth investigation of corresponding residues in hCNT3. By employing the multifunctional capability of hCNT3 as a template for these studies, this study provides novel mechanistic insights into the molecular mechanism(s) of CNT-mediated cation/nucleoside cotransport, including the role of the (G/A)XKX3NEFVA(Y/M/F) motif.  相似文献   

9.
Insulin plays a central role in the regulation of vertebrate metabolism. The hormone, the post-translational product of a single-chain precursor, is a globular protein containing two chains, A (21 residues) and B (30 residues). Recent advances in human genetics have identified dominant mutations in the insulin gene causing permanent neonatal-onset DM2 (14). The mutations are predicted to block folding of the precursor in the ER of pancreatic β-cells. Although expression of the wild-type allele would in other circumstances be sufficient to maintain homeostasis, studies of a corresponding mouse model (57) suggest that the misfolded variant perturbs wild-type biosynthesis (8, 9). Impaired β-cell secretion is associated with ER stress, distorted organelle architecture, and cell death (10). These findings have renewed interest in insulin biosynthesis (1113) and the structural basis of disulfide pairing (1419). Protein evolution is constrained not only by structure and function but also by susceptibility to toxic misfolding.Insulin plays a central role in the regulation of vertebrate metabolism. The hormone, the post-translational product of a single-chain precursor, is a globular protein containing two chains, A (21 residues) and B (30 residues). Recent advances in human genetics have identified dominant mutations in the insulin gene causing permanent neonatal-onset DM2 (14). The mutations are predicted to block folding of the precursor in the ER of pancreatic β-cells. Although expression of the wild-type allele would in other circumstances be sufficient to maintain homeostasis, studies of a corresponding mouse model (57) suggest that the misfolded variant perturbs wild-type biosynthesis (8, 9). Impaired β-cell secretion is associated with ER stress, distorted organelle architecture, and cell death (10). These findings have renewed interest in insulin biosynthesis (1113) and the structural basis of disulfide pairing (1419). Protein evolution is constrained not only by structure and function but also by susceptibility to toxic misfolding.  相似文献   

10.
11.
Butyrophilin 1A1 (BTN1A1) and xanthine oxidoreductase (XOR) are highly expressed in the lactating mammary gland and are secreted into milk associated with the milk fat globule membrane (MFGM). Ablation of the genes encoding either protein causes severe defects in the secretion of milk lipid droplets, suggesting that the two proteins may function in the same pathway. Therefore, we determined whether BTN1A1 and XOR directly interact using protein binding assays, surface plasmon resonance analysis, and gel filtration. Bovine XOR bound with high affinity in a pH- and salt-sensitive manner (KD = 101 ± 31 nm in 10 mm HEPES, 150 mm NaCl, pH 7.4) to the PRY/SPRY/B30.2 domain in the cytoplasmic region of bovine BTN1A1. Binding was stoichiometric, with one XOR dimer binding to either two BTN1A1 monomers or one dimer. XOR bound to BTN1A1 orthologs from mice, humans, or cows but not to the cytoplasmic domains of the closely related human paralogs, BTN2A1 or BTN3A1, or to the B30.2 domain of human RoRet (TRIM 38), a protein in the TRIM family. Analysis of the protein composition of the MFGM of wild type and BTN1A1 null mice showed that most of the XOR in mice lacking BTN1A1 was released from the MFGM in a soluble form when the milk lipid droplets were disrupted to prepare membrane, compared with wild-type mice, in which most of the XOR remained membrane-bound. Thus BTN1A1 functions in vivo to stabilize the association of XOR with the MFGM by direct interactions through the PRY/SPRY/B30.2 domain. The potential significance of BTN1A1/XOR interactions in the mammary gland and other tissues is discussed.Members of the butyrophilin (BTN)3 gene family are attracting increasing attention because they may play multifunctional roles in diverse physiologies, including lactation (1, 2), selection and regulation of T-cells in the immune system (36), and modulation of autoimmune disease (79). BTN proteins have the canonical structures of cell surface receptors, which, after an N-terminal signal sequence, generally comprise two exoplasmic Ig folds (10, 11), a membrane anchor and a cytoplasmic domain consisting of a stem region, a PRY/SPRY/B30.2 domain (12, 13), and a cytoplasmic tail at the C terminus (14).The eponymous BTN1A1 protein has been linked to the secretion of milk lipid droplets because it is highly expressed in the mammary epithelium during lactation and is incorporated into the surface membrane coat surrounding cytoplasmic lipid droplets (the milk fat globule membrane (MFGM)) as they bud into milk from the apical surface (15). Furthermore, ablation of the Btn1a1 gene disrupts lipid secretion, causing the accumulation of large pools of triacylglycerol in the cytoplasm of Btn1a1 null mice (1). In a different context, dietary exposure to BTN1A1 in dairy products has been associated with modulation of the autoimmune disease multiple sclerosis because of structural similarities between the IgI fold of BTN1A1 (16) and the IgV fold of myelin oligodendrocyte glycoprotein (MOG) (17) an antigen on the myelin nerve sheath that is a target for autoantibodies in multiple sclerosis patients (810).Potential interactions between the exoplasmic Ig folds of several BTN proteins, and putative receptors on immune cells are postulated to regulate positive selection of epidermal γδ-T cells in the case of Skint1 (6) and suppress T-cell activation in the case of BTNL2 (4, 5). In addition, BTN2A1 binds to the C-type lectin, DC-SIGN, on immature dendritic cells (18), and proteins in the BTN3A1–3 subfamily bind to an unidentified ligand on various immune cells (19).Interactions between the cytoplasmic domain of BTN and intracellular proteins have not been investigated in any detail. The intracellular region of most BTNs is dominated by the B30.2 or the PRY/SPRY domain, which comprises two sheets of antiparallel β-strands folded into a β- sandwich, which in some proteins is contiguous at the N terminus with one or two α-helices (2024) (for a discussion of the relationship between PRY, SPRY, and B30.2 domains, see Ref. (25)). This domain (here abbreviated as B30.2),4 is postulated to serve as a protein-binding module, by which proteins interact through the extended surface loops that adjoin individual β-strands (22).One protein that may bind to the cytoplasmic region of BTN proteins (and the B30.2 domain) is the redox enzyme, xanthine oxidoreductase (XOR),5 because it was shown to bind to the cytoplasmic domain of mouse BTN1A1 in an in vitro binding assay (26). Furthermore, one XOR-deficient mouse strain (Xdh+/−) (27) displayed a lactation phenotype similar to that of Btn1a1 null mice (1), suggesting that the two proteins may be functionally linked by direct interaction. These conclusions, however, have been challenged, because XOR does not co-localize with BTN1A1 in immunolabeled freeze-fractured replicates of secreted milk lipid droplets (28), and a second mouse strain deficient in XOR does not appear to have an altered lactation phenotype (29).In this paper, we devise in vivo and in vitro assays to show that the cytoplasmic domain of BTN1A1 binds to XOR via the B30.2 domain and that BTN1A1 is required for the stable association of XOR with the MFGM in vivo. Furthermore, interaction with XOR appears to be limited to BTN1A1 orthologs. These results are discussed in the context of potential functions of BTN1A1 in the mammary gland and other tissues and the relationship of BTN1A1 to other BTN family members.  相似文献   

12.
13.
Fibronectin Binds and Enhances the Activity of Bone Morphogenetic Protein 1   总被引:1,自引:0,他引:1  
Bone morphogenetic protein-1-like proteinases play key roles in formation of the extracellular matrix (ECM) in vertebrates via biosynthetic processing of precursors into mature functional proteins involved in ECM assembly. Such processing includes proteolytic activation of the zymogen for lysyl oxidase. Fibronectin (FN) is an abundant protein component of the ECM that is capable of regulating manifold cellular functions through its interactions with various ECM and cell surface proteins. It was previously shown that proteolytic activation of lysyl oxidase is much reduced in cultures of FN-null mouse embryo fibroblasts (MEFs). Here we demonstrate that cellular fibronectin, the form produced by fibroblasts and various other tissue cell types, and plasma fibronectin bind BMP1 with dissociation constants (KD) of ∼100 nm, consistent with a physiological role. Also consistent with such a role, cellular fibronectin FN is shown to positively regulate BMP1 processing activity against Chordin, probiglycan, and type I procollagen in vitro. Endogenous FN and BMP1 are demonstrated to co-localize in cell layers and to form complexes in culture medium. In addition, processing of endogenous BMP1 substrates Chordin, probiglycan, and procollagen is demonstrated to be strikingly reduced in cultures of FN−/− MEFs compared with FN+/− MEF cultures despite similar levels of endogenous BMP1. These data support the conclusion that FN binds BMP1-like proteinases in vivo and that FN is an important determinant of the in vivo activity levels of BMP1-like proteinases.Fibronectin (FN)3 is a noncollagenous extracellular matrix (ECM) glycoprotein of relatively high abundance that regulates a wide variety of cellular functions, including adhesion, migration, proliferation, differentiation, and apoptosis (14). FN is secreted as a disulfide-bonded dimer, and each subunit comprises 12 type I, 2 type II, and 15–17 type III FN modules as well as a “variable” (V) region that lacks homology to other protein domains (3). FN is found as two different major forms, plasma fibronectin (pFN), a soluble form synthesized by hepatocytes, and cellular fibronectin (cFN), which is locally expressed by many other cell types in various tissues (5). Both forms can be assembled into a fibrillar ECM by cultured fibroblasts (6). Differences between cFN and pFN arise from alternative RNA splicing in three regions; two type III repeats (designated EDA and EDB) and the V region. EDA and EDB are present in cFN but absent from pFN, whereas although only one subunit of the pFN dimer contains the V region, almost all cFN subunits contain this region (7). These differences in domain structure contribute to distinct functions for pFN and cFN; cFN plays roles in the dynamic tissue modeling of early embryogenesis and wound healing (8), whereas pFN subserves roles in hemostasis and thrombosis and immune responses (3, 911) and provides a reservoir for deposition in tissue (12).BMP1-like proteinases are evolutionary conserved extracellular metalloproteinases that play multiple roles in fostering ECM formation and activating TGFβ-like growth factors (13). These proteinases biosynthetically convert a variety of precursors into mature functional proteins with roles in ECM formation, including collagen types I-III, V, VII, and XI, laminin 332, and the small leucine-rich proteoglycans biglycan and osteoglycin. One important example is the zymogen for lysyl oxidase (LOX), an enzyme essential to formation of the covalent cross-links responsible for providing collagen and elastic fibers with much of their tensile strength (14). Recently, FN was reported to bind LOX in vitro (15). It was also suggested to positively regulate the proteolytic activation of LOX, as dramatically decreased processing of the zymogen for LOX was observed in FN−/− mouse embryo fibroblast (MEF) cultures compared with FN+/− MEF cultures even though equal amounts of BMP1 proteinase were produced by MEFs of the two different genotypes (15). These observations prompted the present study to determine whether FN might be involved in modulating the activities of BMP1-like proteinases. Herein, we provide evidence for direct interaction between FN and BMP1. BMP1 is shown to bind multiple FN sites via its non-protease domains, with affinities in the ∼100 nm range for cFN and pFN. This is a range congruent with KD values (30–800 nm) previously estimated for binding of FN to its integrin receptors (16, 17) and is, thus, consistent with the likelihood of in vivo FN-BMP1 interactions. Moreover, cFN is shown to positively regulate BMP1 processing activity against a number of substrates in vitro. Consistent with the in vitro evidence of FN-BMP1 interactions, we demonstrate FN-BMP1 co-localization and the existence of FN-BMP1 complexes in cell cultures. Also demonstrated is a striking decrease in the processing of various endogenous BMP1 substrates in cultures of FN−/− MEFs compared with FN+/− MEF cultures. Implications of the data, which support the conclusion that FN positively regulates BMP1 activities in vivo, are discussed.  相似文献   

14.
15.
STIM1 and STIM2 are dynamic transmembrane endoplasmic reticulum Ca2+ sensors, coupling directly to activate plasma membrane Orai Ca2+ entry channels. Despite extensive sequence homology, the STIM proteins are functionally distinct. We reveal that the short variable N-terminal random coil sequences of STIM1 and STIM2 confer profoundly different activation properties. Using Orai1-expressing HEK293 cells, chimeric replacement of the 43-amino-acid STIM1 N terminus with that of STIM2 attenuates Orai1-mediated Ca2+ entry and drastically slows store-induced Orai1 channel activation. Conversely, the 55-amino-acid STIM2 terminus substituted within STIM1 strikingly enhances both Orai1-mediated Ca2+ entry and constitutive coupling to activate Orai1 channels. Hence, STIM N termini are powerful coupling modifiers, functioning in STIM2 to “brake” the otherwise constitutive activation of Orai1 channels afforded by its high sensitivity to luminal Ca2+.The transmembrane ER4 proteins STIM1 and STIM2 function as sensors of Ca2+ within ER stores (1, 2). Depletion of luminal Ca2+ within the ER triggers aggregation and translocation of STIMs into junctions closely associated with the plasma membrane, where they activate the highly Ca2+-selective Orai family of store-operated channels (SOCs) via conformational coupling (38). Recent investigations of the cytoplasmic portion of STIM1 revealed that it alone is sufficient to activate Orai (912) via a short (∼100 amino acids) region centered around the second coiled-coil domain (see Fig. 1) (1315). However, although activation of Orai1 is mediated entirely within the C-terminal portion of STIM, physiological control of STIM1 and STIM2 is exerted via their N-terminal ER-luminal Ca2+-sensing domains. The extent to which structural differences between these domains in STIM1 and STIM2 contribute to their distinct properties (1619) remains poorly understood. Although STIM2 has the capacity to sense ER Ca2+ and activate SOCs (16, 17, 19), overexpressed STIM2 inhibits endogenous SOCs (18). Moreover, the kinetics of SOC activation by STIM2 are much slower than STIM1 (17). STIM2 was recently revealed to have a decreased Ca2+-sensing affinity when compared with STIM1 by virtue of three amino acid substitutions in the Ca2+-binding EF-hand domain (16). Although the lower affinity of the STIM2 EF-hand accounts for differences in the activation thresholds of STIM1 and STIM2 (16, 20, 21), it does not explain the slow kinetics of STIM2 nor its dominance over endogenous SOC activation. However, recent investigations reveal similar abilities of the cytosolic portions of STIM1 and STIM2 to activate Orai1 (12). Hence, although activation of Orai1 is mediated entirely within the C-terminal portion of STIM, physiological control of STIM1 and STIM2 is exerted via their N-terminal ER-luminal Ca2+-sensing domains.Open in a separate windowFIGURE 1.Schematic diagram depicting the domain structure of STIM1, STIM2, and STIM chimeras. The currently defined domains of STIM1 and STIM2 are depicted as canonical (cEF) and hidden (hEF) EF-hands, SAM domains, transmembrane domains (TM), coiled-coil structures, a proline-rich domain (P), and a polybasic tail (K). The sequences of the STIM1 and STIM2 N-terminal domains were aligned using the lalign program and depicted with red indicating identical amino acids and blue indicating similarity.The initial triggering events for STIM1 and STIM2 proteins involve the unfolding and aggregation of the N-terminal domains resulting from dissociation of Ca2+ from the luminal EF-hand Ca2+ binding domains (2023). Recent evidence reveals that this unfolding is much slower for the N terminus of STIM2 than for STIM1 (21). Although most of the N termini of STIM1 and STIM2 are highly homologous, significant variability exists in the first 60 N-terminal amino acids upstream from the EF-hands, comprising a flexible random coil domain (21). Intriguingly, these upstream sequences appear to markedly modify the stability of the N-terminal domains of STIM1 and STIM2 (21). We reveal here that these sequences confer profound distinctions between STIM1 and STIM2 in their coupling to activate SOCs. In STIM2, this domain acts as a powerful “brake” to restrict constitutive activation of SOCs, occurring as a result of its high sensitivity to luminal Ca2+.  相似文献   

16.
Leptospira interrogans is a pathogenic spirochete that causes disease in both humans and animals. LigB (Leptospiral immunoglobulin-like protein B) contributes to the binding of Leptospira to extracellular matrix proteins such as fibronectin (Fn), fibrinogen, laminin, and collagen. A high affinity Fn-binding region of LigB has been recently localized to LigBCen2, which contains the partial eleventh and full twelfth immunoglobulin-like repeats (LigBCen2R) and 47 amino acids of the non-repeat region (LigBCen2NR) of LigB. In this study, LigBCen2NR was shown to bind to the N-terminal domain (NTD) of Fn (KD = 379 nm) by an enzyme-linked immunosorbent assay and isothermal titration calorimetry. Interestingly, this sequence was not observed to adopt secondary structure by far UV circular dichroism or by differential scanning calorimetry, in agreement with computer-based secondary structure predictions. A low partition coefficient (Kav) measured with gel permeation chromatography, a high hydrodynamic radius (Rh) measured with dynamic light scattering, and the insensitivity of the intrinsic viscosity to guanidine hydrochloride treatment all suggest that LigBCen2NR possesses an extended and disordered structure. Two-dimensional 15N-1H HSQC NMR spectra of intact LigBCen2 in the absence and presence of NTD are consistent with these observations, suggesting the presence of both a β-rich region and an unstructured region in LigBCen2 and that the latter of these selectively interacts with NTD. Upon binding to NTD, LigBCen2NR was observed by CD to adopt a β-strand-rich structure, suggestive of the known β-zipper mode of NTD binding.Leptospira interrogans is a pathogenic spirochete that causes leptospirosis throughout the world, especially in developing countries but also in regions of the United States where it has reemerged (1). Weil''s syndrome, a severe form of this disease, is an acute febrile illness associated with multiorgan damage, including liver failure (jaundice), renal failure (nephritis), pulmonary hemorrhage, and meningitis (1), and has a 15% mortality rate if not treated (2). The molecular pathogenesis of leptospirosis is poorly understood, and the bacterial virulence factors involved are largely unknown. Recently, several potential Leptospira virulence factors have been described, including sphingomyelinases, serine proteases, zinc-dependent proteases, and collagenase (3); LipL32 (4); lipopolysaccharide (5); a novel factor H, laminin, and Fn-binding protein (Lsa24 or Len) (68); Loa 22 (9); and Lig (Leptospiral immunoglobulin-like) proteins (1012).Lig proteins, including LigA, LigB, and LigC, contain multiple immunoglobulin-like repeat domains (13 in LigA, 12 in LigB and LigC) (1012). Interestingly, the first 630 residues, from the N terminus to the first half of the seventh immunoglobulin-like domain, are conserved between LigA and LigB, but the rest of the immunoglobulin-like domains are variable (1012) between the two proteins. Also, a non-immunoglobulin-like repeat region found on the C-terminal tail of LigB is not found in LigA (1012). Lig proteins are categorized as microbial surface components recognizing adhesive matrix molecules (MSCRAMMs)2 due to their ability to bind to eukaryotic cells (13) through their interactions with extracellular matrix components, including fibronectin (Fn), laminin, collagens, elastin, and tropoelastin (13, 14, 45). Previously, a high affinity Fn-binding region was localized to LigBCen2, which includes the partial eleventh and complete twelfth immunoglobulin-like repeat region and the first 47 amino acids of the non-repeat regions of LigB (15). LigBCen2 was shown to bind to both the N-terminal domain (NTD) and the gelatin binding domain (GBD) of Fn. The addition of calcium induces a conformational change in LigBCen2 and enhances binding between LigBCen2 and the NTD of Fn (15).The first step in the process of bacterial infection is cellular adhesion, mediated by bacterial adhesins interacting with various components of the extracellular matrix (16). Known interaction modes between Fn and bacterial Fn-binding proteins include the β-zipper (17, 18) and the cationic cradle (19). It was recently discovered that the Fn-binding domains in certain Fn-binding proteins are disordered and extended but gain structure upon binding to the NTD of Fn (2022).We have performed a fine-mapping study of the NTD-binding site on LigBCen2 and identified this site as LigBCen2NR, a portion of the non-repeat region (amino acids 1119–1165). The addition of NTD promotes the folding of LigBCen2NR from a disordered and extended structure to a folded structure. This finding is notable, since LigBCen2NR is located in the non-immunoglobulin-like region of LigB, as compared with other Fn-binding proteins, such as Staphylococcus aureus FnbpA and FnbpB (23), Streptococcus dysgalactiae FnBB (17), and Streptococcus pyogenes SfbI and SfbII (24). Thus, the binding mode appears to be similar to the known β-zipper mechanism but unique in sequence-specific interactions. This finding provides the fundamental groundwork for the development of a therapeutic agent to target this interaction in order to prevent or treat Leptospira infection.  相似文献   

17.
Transglutaminase type 2 (TG2) is both a protein cross-linking enzyme and a cell adhesion molecule with an elusive unconventional secretion pathway. In normal conditions, TG2-mediated modification of the extracellular matrix modulates cell motility, proliferation and tissue repair, but under continuous cell insult, higher expression and elevated extracellular trafficking of TG2 contribute to the pathogenesis of tissue scarring. In search of TG2 ligands that could contribute to its regulation, we characterized the affinity of TG2 for heparan sulfate (HS) and heparin, an analogue of the chains of HS proteoglycans (HSPGs). By using heparin/HS solid-binding assays and surface plasmon resonance we showed that purified TG2 has high affinity for heparin/HS, comparable to that for fibronectin, and that cell-surface TG2 interacts with heparin/HS. We demonstrated that cell-surface TG2 directly associates with the HS chains of syndecan-4 without the mediation of fibronectin, which has affinity for both syndecan-4 and TG2. Functional inhibition of the cell-surface HS chains of wild-type and syndecan-4-null fibroblasts revealed that the extracellular cross-linking activity of TG2 depends on the HS of HSPG and that syndecan-4 plays a major but not exclusive role. We found that heparin binding did not alter TG2 activity per se. Conversely, fibroblasts deprived of syndecan-4 were unable to effectively externalize TG2, resulting in its cytosolic accumulation. We propose that the membrane trafficking of TG2, and hence its extracellular activity, is linked to TG2 binding to cell-surface HSPG.Transglutaminase type 2 (TG2,2 EC 2.3.2.13) is the most widespread member of a large family of enzymes that catalyze the Ca2+-dependent post-translational modification of proteins leading to intra- or intermolecular Nϵ(γ-glutamyl)lysine bonds (1, 2). Unlike other family members, TG2 is uniquely exported through a yet to be elucidated non-conventional pathway. Once secreted, TG2 finds in the extracellular compartment the ideal conditions of high Ca2+ and low GTP concentration for the activation of its intrinsic transamidation activity (cross-linking) (2, 3). Intracellularly, GTP binding suppresses the Ca2+-dependent cross-linking activity and determines the additional GTPase activity of TG2 (4, 5), which is responsible for signal transduction (6). Once externalized, TG2 remains tightly bound to the cell surface and to the extracellular matrix (ECM) (7, 8), and it is rarely found free in the conditioned medium, unless overexpressed by cell transfection (9).Extracellular TG2 activity is involved in the cross-linking of the ECM, conferring resistance to matrix metalloproteinase and promoting cell-matrix interactions via cross-linking of fibronectin (FN) and collagen (1, 7, 11, 12). TG2 has an additional non-enzymatic role in the matrix as an integrin-β1 co-receptor (8) by supporting RGD-independent cell adhesion to FN (8, 13, 14).Extracellular cross-linking and TG2-mediated adhesion facilitate the repair process in many tissue compartments (1, 2, 15, 16). On the other hand, uncontrolled cross-linking as a consequence of chronic cell insult and secretion of TG2 has been implicated in a number of pathological conditions, including kidney, liver, and pulmonary fibrosis (1720).Understanding how TG2 is exported and targeted to the cell surface is critical for limiting its cellular secretion and extracellular action. Although a key trigger for TG2 export is cell stress (2, 21, 22), TG2 is not unspecifically released, because extracellular trafficking occurs in the absence of leakage of intracellular components and cells remain viable (23). We know that TG2 requires the tertiary structure of its active site region to be secreted (9); moreover, TG2 is acetylated on the N terminus (24), a process reported to affect membrane targeting of non-conventional secreted proteins (25). Two main binding partners for TG2, FN and integrin-β1, have both been attributed a possible role in the transport of TG2 to the cell surface (8, 26). FN was shown to co-localize with TG2 once released (26), and integrin-β1 to co-associate with TG2 in cells induced to differentiate (8).TG2 has also long been known to have some affinity for heparin (27, 28), a highly sulfated analogue of heparan sulfate (HS) glycosaminoglycan chains, which are abundant constituents of the cell surface/ECM. HS chains are linear polysaccharides consisting of alternating N-acetylated or N-sulfated glucosamine units (GlcNAc or GlcNS), and uronic acids (glucuronic acid GlcA or iduronic acid IdoA residues) (29), which only exist covalently bound to the core protein of cell-surface proteoglycans (syndecans and glypicans) and secreted proteoglycans (29). Heparin binding is a property common to many ECM proteins (29), but the level of affinity has never been established for TG2, which makes it difficult to estimate the real biological significance of this interaction. Heparan sulfate proteoglycans (HSPG) bind ECM ligands through the HS chains, influencing their biological activity, trafficking, and secretion. Among the HSPG subfamilies, the syndecans act as co-receptors for both ECM components and soluble ligands (30), and syndecan-4 has overlapping roles with extracellular TG2 in wound healing and fibrosis (31, 32). In this study, we show that TG2 has a surprisingly high affinity for heparin and HS, raising the hypothesis that HSPG are involved in its biological activity. We demonstrate that HSPGs are essential for the transamidating activity of TG2 at the cell surface and that syndecan-4 acts as a receptor for TG2, which is involved in the trafficking and cell-surface localization, and thus activity of TG2.  相似文献   

18.
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   

19.
Clinically, amniotic membrane (AM) suppresses inflammation, scarring, and angiogenesis. AM contains abundant hyaluronan (HA) but its function in exerting these therapeutic actions remains unclear. Herein, AM was extracted sequentially with buffers A, B, and C, or separately by phosphate-buffered saline (PBS) alone. Agarose gel electrophoresis showed that high molecular weight (HMW) HA (an average of ∼3000 kDa) was predominantly extracted in isotonic Extract A (70.1 ± 6.0%) and PBS (37.7 ± 3.2%). Western blot analysis of these extracts with hyaluronidase digestion or NaOH treatment revealed that HMW HA was covalently linked with the heavy chains (HCs) of inter-α-inhibitor (IαI) via a NaOH-sensitive bond, likely transferred by the tumor necrosis factor-α stimulated gene-6 protein (TSG-6). This HC·HA complex (nHC·HA) could be purified from Extract PBS by two rounds of CsCl/guanidine HCl ultracentrifugation as well as in vitro reconstituted (rcHC·HA) by mixing HMW HA, serum IαI, and recombinant TSG-6. Consistent with previous reports, Extract PBS suppressed transforming growth factor-β1 promoter activation in corneal fibroblasts and induced mac ro phage apo pto sis. However, these effects were abolished by hyaluronidase digestion or heat treatment. More importantly, the effects were retained in the nHC·HA or rcHC·HA. These data collectively suggest that the HC·HA complex is the active component in AM responsible in part for clinically observed anti-inflammatory and anti-scarring actions.Hyaluronan (HA)4 is widely distributed in extracellular matrices, tissues, body fluids, and even in intracellular compartments (reviewed in Refs. 1 and 2). The molecular weight of HA ranges from 200 to 10,000 kDa depending on the source (3), but can also exist as smaller fragments and oligosaccharides under certain physiological or pathological conditions (1). Investigations over the last 15 years have suggested that low Mr HA can induce the gene expression of proinflammatory mediators and proangiogenesis, whereas high molecular weight (HMW) HA inhibits these processes (47).Several proteins have been shown to bind to HA (8) such as aggrecan (9), cartilage link protein (10), versican (11), CD44 (12, 13), inter-α-inhibitor (IαI) (14, 15), and tumor necrosis factor-α stimulated gene-6 protein (TSG-6) (16, 17). IαI consists of two heavy chains (HCs) (HC1 and HC2), both of which are linked through ester bonds to a chondroitin sulfate chain that is attached to the light chain, i.e. bikunin. Among all HA-binding proteins, only the HCs of IαI have been clearly demonstrated to be covalently coupled to HA (14, 18). However, TSG-6 has also been reported to form stable, possibly covalent, complexes with HA, either alone (19, 20) or when associated with HC (21).The formation of covalent bonds between HCs and HA is mediated by TSG-6 (2224) where its expression is often induced by inflammatory mediators such as tumor necrosis factor-α and interleukin-1 (25, 26). TSG-6 is also expressed in inflammatory-like processes, such as ovulation (21, 27, 28) and cervical ripening (29). TSG-6 interacts with both HA (17) and IαI (21, 24, 3033), and is essential for covalently transferring HCs on to HA (2224). The TSG-6-mediated formation of the HC·HA complex has been demonstrated to play a crucial role in female fertility in mice. The HC·HA complex is an integral part of an expanded extracellular “cumulus” matrix around the oocyte, which plays a critical role in successful ovulation and fertilization in vivo (22, 34). HC·HA complexes have also been found at sites of inflammation (3538) where its pro- or anti-inflammatory role remain arguable (39, 40).Immunostaining reveals abundant HA in the avascular stromal matrix of the AM (41, 42).5 In ophthalmology, cryopreserved AM has been widely used as a surgical graft for ocular surface reconstruction and exerts clinically observable actions to promote epithelial wound healing and to suppress inflammation, scarring, and angiogenesis (for reviews see Refs. 4345). However, it is not clear whether HA in AM forms HC·HA complex, and if so whether such an HC·HA complex exerts any of the above therapeutic actions. To address these questions, we extracted AM with buffers of increasing salt concentration. Because HMW HA was found to form the HC·HA complex and was mainly extractable by isotonic solutions, we further purified it from the isotonic AM extract and reconstituted it in vitro from three defined components, i.e. HMW HA, serum IαI, and recombinant TSG-6. Our results showed that the HC·HA complex is an active component in AM responsible for the suppression of TGF-β1 promoter activity, linkable to the scarring process noted before by AM (4648) and by the AM soluble extract (49), as well as for the promotion of macrophage death, linkable to the inflammatory process noted by AM (50) and the AM soluble extract (51).  相似文献   

20.
Collapsin response mediator protein 2 (CRMP2) is an intracellular protein that mediates signaling of Semaphorin3A (Sema3A), a repulsive axon guidance molecule. Fyn, a Src-type tyrosine kinase, is involved in the Sema3A signaling. However, the relationship between CRMP2 and Fyn in this signaling pathway is still unknown. In our research, we demonstrated that Fyn phosphorylated CRMP2 at Tyr32 residues in HEK293T cells. Immunohistochemical analysis using a phospho-specific antibody at Tyr32 of CRMP showed that Tyr32-phosphorylated CRMP was abundant in the nervous system, including dorsal root ganglion neurons, the molecular and Purkinje cell layer of adult cerebellum, and hippocampal fimbria. Overexpression of a nonphosphorylated mutant (Tyr32 to Phe32) of CRMP2 in dorsal root ganglion neurons interfered with Sema3A-induced growth cone collapse response. These results suggest that Fyn-dependent phosphorylation of CRMP2 at Tyr32 is involved in Sema3A signaling.Collapsin response mediator proteins (CRMPs)4 have been identified as intracellular proteins that mediate Semaphorin3A (Sema3A) signaling in the nervous system (1). CRMP2 is one of the five members of the CRMP family. CRMPs also mediate signal transduction of NT3, Ephrin, and Reelin (24). CRMPs interact with several intracellular molecules, including tubulin, Numb, kinesin1, and Sra1 (58). CRMPs are involved in axon guidance, axonal elongation, cell migration, synapse maturation, and the generation of neuronal polarity (1, 2, 4, 5).CRMP family proteins are known to be the major phosphoproteins in the developing brain (1, 9). CRMP2 is phosphorylated by several Ser/Thr kinases, such as Rho kinase, cyclin-dependent kinase 5 (Cdk5), and glycogen synthase kinase 3β (GSK3β) (2, 1013). The phosphorylation sites of CRMP2 by these kinases are clustered in the C terminus and have already been identified. Rho kinase phosphorylates CRMP2 at Thr555 (10). Cdk5 phosphorylates CRMP2 at Ser522, and this phosphorylation is essential for sequential phosphorylations by GSK3β at Ser518, Thr514, and Thr509 (2, 1113). These phosphorylations disrupt the interaction of CRMP2 with tubulin or Numb (2, 3, 13). The sequential phosphorylation of CRMP2 by Cdk5 and GSK3β is an essential step in Sema3A signaling (11, 13). Furthermore, the neurofibrillary tangles in the brains of people with Alzheimer disease contain hyperphosphorylated CRMP2 at Thr509, Ser518, and Ser522 (14, 15).CRMPs are also substrates of several tyrosine kinases. The phosphorylation of CRMP2 by Fes/Fps and Fer has been shown to be involved in Sema3A signaling (16, 17). Phosphorylation of CRMP2 at Tyr479 by a Src family tyrosine kinase Yes regulates CXCL12-induced T lymphocyte migration (18). We reported previously that Fyn is involved in Sema3A signaling (19). Fyn associates with PlexinA2, one of the components of the Sema3A receptor complex. Fyn also activates Cdk5 through the phosphorylation at Tyr15 of Cdk5 (19). In dorsal root ganglion (DRG) neurons from fyn-deficient mice, Sema3A-induced growth cone collapse response is attenuated compared with control mice (19). Furthermore, we recently found that Fyn phosphorylates CRMP1 and that this phosphorylation is involved in Reelin signaling (4). Although it has been shown that CRMP2 is involved in Sema3A signaling (1, 11, 13), the relationship between Fyn and CRMP2 in Sema3A signaling and the tyrosine phosphorylation site(s) of CRMPs remain unknown.Here, we show that Fyn phosphorylates CRMP2 at Tyr32. Using a phospho-specific antibody against Tyr32, we determined that the residue is phosphorylated in vivo. A nonphosphorylated mutant CRMP2Y32F inhibits Sema3A-induced growth cone collapse. These results indicate that tyrosine phosphorylation by Fyn at Tyr32 is involved in Sema3A signaling.  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