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 0 毫秒
1.
The mechanisms leading to dopaminergic neuronal loss in the substantia nigra of patients with Parkinson disease (PD) remain poorly understood. We recently reported that aberrant DNA replication mediated by DNA polymerase-β (DNA pol-β) plays a causal role in the death of postmitotic neurons in an in vitro model of PD. In the present study, we show that both proliferating cell nuclear antigen (PCNA) and DNA pol-β are required for MPP+-induced neuronal death. PCNA binds to the catalytic domain of DNA pol-β in MPP+-treated neurons and in post-mortem brain tissues of PD patients. The PCNA-DNA pol-β complex is loaded into DNA replication forks and mediates DNA replication in postmitotic neurons. The aberrant DNA replication mediated by the PCNA-DNA pol-β complex induces p53-dependent neuronal cell death. Our results indicate that the interaction of PCNA and DNA pol-β contributes to neuronal death in PD.  相似文献   

2.
At clinically relevant doses, chemotherapeutic SN1 DNA methylating agents induce an ATR-mediated checkpoint response in human cells that is dependent on functional MutSα and MutLα. Deficiency of either mismatch repair activity renders cells highly resistant to this class of drug, but the mechanisms linking mismatch repair to checkpoint activation have remained elusive. In this study we have systematically examined the interactions of human MutSα and MutLα with proteins of the ATR-Chk1 pathway using both nuclear extracts and purified proteins. Using nuclear co-immunoprecipitation, we have detected interaction of MutSα with ATR, TopBP1, Claspin, and Chk1 and interaction of MutLα with TopBP1 and Claspin. We were unable to detect interaction of MutSα or MutLα with Rad17, Rad9, or replication protein A in the extract system. Use of purified proteins confirmed direct interaction of MutSα with ATR, TopBP1, and Chk1 and of MutLα with TopBP1. MutSα-Claspin and MutLα-Claspin interactions were not demonstrable with purified proteins, suggesting that extract interactions are indirect or depend on post-translational modification. Use of a modified chromatin immunoprecipitation assay showed that proliferating cell nuclear antigen, ATR, TopBP1, and Chk1 are recruited to chromatin in a MutLα- and MutSα-dependent fashion after N-methyl-N′-nitro-N-nitrosoguanidine treatment. However, chromatin enrichment of replication protein A, Claspin, Rad17-RFC, and Rad9-Rad1-Hus1 was not detected in these experiments. Although our failure to observe enrichment of the latter activities could be due to sensitivity limitations, these observations may indicate a novel mechanism for ATR activation.  相似文献   

3.
The regulated process of protein import into the nucleus of a eukaryotic cell is mediated by specific nuclear localization signals (NLSs) that are recognized by protein-import receptors. In this study, we present fluorescence-based methods to quantitatively address the physicochemical details of NLS recognition by the receptor protein importin α (Impα) in living cells. First, by combining fluorescence recovery after photobleaching measurements and protein-concentration calibration, we quantitatively define nuclear import saturability and afford an affinity value for NLS-Impα binding. Second, by fluorescence lifetime imaging microscopy, we directly monitor the occurrence of NLS-Impα interaction and measure its effective dissociation constant (KD) in the actual cellular environment. Our kinetic and thermodynamic analyses independently indicate that the subsaturation of Impα with the expressed NLS cargo regulates nuclear import rates in living cells, in contrast to what can be predicted on the basis of available in vitro data. Finally, our experiments also provide evidence for the regulation of nuclear import mediated by the intrasteric importin β-binding domain of Impα and yield the first estimate of its autoinhibition energy in living cells.  相似文献   

4.
Gadd45α is a nuclear protein encoded by a DNA damage-inducible gene. Through its interactions with other proteins, Gadd45α participates in the regulation of DNA repair, cell cycle, cell proliferation, and apoptosis. The NMR structure of human Gadd45α has been determined and shows an α/β fold with two long disordered and flexible regions at the N terminus and one of the loops. Human Gadd45α is predominantly monomeric in solution but exists in equilibrium with dimers and other oligomers whose population increases with protein concentration. NMR analysis shows that Aurora A interacts through its N-terminal domain with a region of human Gadd45α encompassing the site of dimerization, suggesting that the oligomerization of Gadd45α could be a regulatory mechanism to modulate its interactions with Aurora A, and possibly with other proteins too. However, Gadd45α appears to interact only weakly with PCNA through its flexible loop, in contrast with previous and contradictory reports.  相似文献   

5.
Kaposi''s sarcoma-associated herpesvirus (KSHV) is tightly linked to at least two lymphoproliferative disorders, primary effusion lymphoma (PEL) and multicentric Castleman''s disease (MCD). However, the development of KSHV-mediated lymphoproliferative disease is not fully understood. Here, we generated two recombinant KSHV viruses deleted for the first RBP-Jκ binding site (RTA1st) and all three RBP-Jκ binding sites (RTAall) within the RTA promoter. Our results showed that RTA1st and RTAall recombinant viruses possess increased viral latency and a decreased capability for lytic replication in HEK 293 cells, enhancing colony formation and proliferation of infected cells. Furthermore, recombinant RTA1st and RTAall viruses showed greater infectivity in human peripheral blood mononuclear cells (PBMCs) relative to wt KSHV. Interestingly, KSHV BAC36 wt, RTA1st and RTAall recombinant viruses infected both T and B cells and all three viruses efficiently infected T and B cells in a time-dependent manner early after infection. Also, the capability of both RTA1st and RTAall recombinant viruses to infect CD19+ B cells was significantly enhanced. Surprisingly, RTA1st and RTAall recombinant viruses showed greater infectivity for CD3+ T cells up to 7 days. Furthermore, studies in Telomerase-immortalized human umbilical vein endothelial (TIVE) cells infected with KSHV corroborated our data that RTA1st and RTAall recombinant viruses have enhanced ability to persist in latently infected cells with increased proliferation. These recombinant viruses now provide a model to explore early stages of primary infection in human PBMCs and development of KSHV-associated lymphoproliferative diseases.  相似文献   

6.
Wnt/β-catenin signaling is initiated at the cell surface by association of secreted Wnt with its receptors Frizzled (Fz) and low density lipoprotein receptor-related protein 5/6 (LRP5/6). The study of these molecular interactions has been a significant technical challenge because the proteins have been inaccessible in sufficient purity and quantity. In this report we describe insect cell expression and purification of soluble mouse Fz8 cysteine-rich domain and human LRP6 extracellular domain and show that they inhibit Wnt/β-catenin signaling in cellular assays. We determine the binding affinities of Wnts and Dickkopf 1 (Dkk1) to the relevant co-receptors and reconstitute in vitro the Fz8 CRD·Wnt3a·LRP6 signaling complex. Using purified fragments of LRP6, we further show that Wnt3a binds to a region including only the third and fourth β-propeller domains of LRP6 (E3E4). Surprisingly, we find that Wnt9b binds to a different part of the LRP6 extracellular domain, E1E2, and we demonstrate that Wnt3a and Wnt9b can bind to LRP6 simultaneously. Dkk1 binds to both E1E2 and E3E4 fragments and competes with both Wnt3a and Wnt9b for binding to LRP6. The existence of multiple, independent Wnt binding sites on the LRP6 co-receptor suggests new possibilities for the architecture of Wnt signaling complexes and a model for broad-spectrum inhibition of Wnt/β-catenin signaling by Dkk1.  相似文献   

7.
Vaccinia viral envelope protein A27 (110 amino acids) specifically interacts with heparin (HP) or heparan sulfate (HS) proteoglycans for cell surface attachment. To examine the binding mechanism, a truncated soluble form of A27 (sA27-aa; residues 21–84 of A27) with Cys71 and Cys72 mutated to Ala was used as the parent molecule. sA27-aa consists of two structurally distinct domains, a flexible Arg/Lys-rich heparin-binding site (HBS) (residues 21–32; 21STKAAKKPEAKR32) and a rigid coiled-coil domain (residues 43–84), both essential for the specific binding. As shown by surface plasmon resonance (SPR), the binding affinity of sA27-aa for HP (KA = 1.25 × 108 m−1) was approximately 3 orders of magnitude stronger than that for nonspecific binding, such as to chondroitin sulfate (KA = 1.65 × 105 m−1). Using site-directed mutagenesis of HBS and solution NMR, we identified a “KKPE” segment with a turn-like conformation that mediates specific HP binding. In addition, a double mutant T22K/A25K in which the KKPE segment remained intact showed an extremely high affinity for HP (KA = 1.9 × 1011 m−1). Importantly, T22K/A25K retained the binding specificity for HP and HS but not chondroitin sulfate, as shown by in vitro SPR and in vivo cell adhesion and competitive binding assays. Molecular modeling of the HBS was performed by dynamics simulations and provides an explanation of the specific binding mechanism in good agreement with the site-directed mutagenesis and SPR results. We conclude that a turn-like structure introduced by the KKPE segment in vaccinia viral envelope protein A27 is responsible for its specific binding to HP and to HS on cell surfaces.  相似文献   

8.
Endothelin-1 (ET-1) is a potent vasoconstrictor and co-mitogen for vascular smooth muscle and is implicated in pulmonary vascular remodeling and the development of pulmonary arterial hypertension. Vascular smooth muscle is an important source of ET-1. Here we demonstrate synergistic induction of preproET-1 message RNA and release of mature peptide by a combination of tumor necrosis factor α (TNFα) and interferon γ (IFNγ) in primary human pulmonary artery smooth muscle cells. This induction was prevented by pretreatment with the histone acetyltransferase inhibitor anacardic acid. TNFα induced a rapid and prolonged pattern of nuclear factor (NF)-κB p65 subunit activation and binding to the native preproET-1 promoter. In contrast, IFNγ induced a delayed activation of interferon regulatory factor-1 without any effect on NF-κB p65 nuclear localization or consensus DNA binding. However, we found cooperative p65 binding and histone H4 acetylation at distinct κB sites in the preproET-1 promoter after stimulation with both TNFα and IFNγ. This was associated with enhanced recruitment of RNA polymerase II to the ATG start site and read-through of the ET-1 coding region. Understanding such mechanisms is crucial in determining the key control points in ET-1 release. This has particular relevance to developing novel treatments targeted at the inflammatory component of pulmonary vascular remodeling.Endothelin-1 is a 21-amino acid peptide which is known to be both a potent vasoconstrictor and mitogen for vascular smooth muscle (1, 2). It is released as a 38-amino acid precursor (Big ET-12) before cleavage to the mature ET-1 form. As such it has been implicated in the pathogenesis of vascular disease and is particularly associated with pulmonary arterial hypertension (3). Indeed, several endothelin receptor antagonists are now approved for the treatment of pulmonary arterial hypertension (4). However, endothelin receptor antagonists as a class are associated with potentially serious side effects (4), making new treatments aimed at blocking ET-1 synthesis an attractive alternative.Although endothelial cells are thought to be the main source of ET-1 release, several groups including our own have shown that ET-1 can be released from the more numerous vascular smooth muscle cells (510). The vascular pathology observed in pulmonary arterial hypertension is propagated by inflammation, and circulating levels of cytokines including tumor necrosis factor α (TNFα) are elevated in patients with pulmonary arterial hypertension (1115). In many cell types cytokines mediate their biological effects at least in part by the activation of the nuclear factor κB (NF-κB) pathway (16), and a role for NF-κB in pulmonary arterial hypertension has been proposed (17). In addition, we have shown previously that a combination of TNFα and interferon γ (IFNγ) stimulates human pulmonary artery smooth muscle (HPASM) cells to release ET-1 (18). However, the mechanisms underlying this effect are unknown.The preproET-1 promoter region has been shown experimentally to possess binding sites for nuclear factor (NF)-1 and phorbol ester-sensitive c-Fos and c-Jun complexes (19), acute phase reactant regulatory proteins, and binding sites for AP-1 and GATA-2 (2022). In addition, binding sites for interferon regulatory factor-1 (IRF-1) and NF-κB are predicted by Transfac analysis (23). The close proximity of the IRF-1 site and one of the NF-κB sites is characteristic of genes that are regulated by the synergistic action of TNFα and IFNγ, such as interleukin-6 (IL-6) and intercellular adhesion molecule-1 (24, 25), although ET-1 has not previously been recognized in this group.Our aims were, therefore, to investigate the role of NF-κB in ET-1 release by primary HPASM cells. In addition, we were interested in the role of histone acetylation in the epigenetic control of the ET-1 production. Understanding these novel mechanisms will allow a greater understanding of the pathogenesis of vascular remodeling in pulmonary vessels and aid in the development of new treatment strategies aimed at blocking synthesis of ET-1.  相似文献   

9.
10.
Latent transforming growth factor (TGF) β-binding proteins (LTBPs) interact with fibrillin-1. This interaction is important for proper sequestration and extracellular control of TGFβ. Surface plasmon resonance interaction studies show that residues within the first hybrid domain (Hyb1) of fibrillin-1 contribute to interactions with LTBP-1 and LTBP-4. Modulation of binding affinities by fibrillin-1 polypeptides in which residues in the third epidermal growth factor-like domain (EGF3) are mutated demonstrates that the binding sites for LTBP-1 and LTBP-4 are different and suggests that EGF3 may also contribute residues to the binding site for LTBP-4. In addition, fibulin-2, fibulin-4, and fibulin-5 bind to residues contained within EGF3/Hyb1, but mutated polypeptides again indicate differences in their binding sites in fibrillin-1. Results demonstrate that these protein-protein interactions exhibit “exquisite specificities,” a phrase commonly used to describe monoclonal antibody interactions. Despite these differences, interactions between LTBP-1 and fibrillin-1 compete for interactions between fibrillin-1 and these fibulins. All of these proteins have been immunolocalized to microfibrils. However, in fibrillin-1 (Fbn1) null fibroblast cultures, LTBP-1 and LTBP-4 are not incorporated into microfibrils. In contrast, in fibulin-2 (Fbln2) null or fibulin-4 (Fbln4) null cultures, fibrillin-1, LTBP-1, and LTBP-4 are incorporated into microfibrils. These data show for the first time that fibrillin-1, but not fibulin-2 or fibulin-4, is required for appropriate matrix assembly of LTBPs. These studies also suggest that the fibulins may affect matrix sequestration of LTBPs, because in vitro interactions between these proteins are competitive.Fibrillin microfibrils are ubiquitous structural elements in the connective tissue. Fibrillin microfibrils provide organs with tissue-specific architectural frameworks designed to support the mature functional integrity of the particular organ. In addition, fibrillin microfibrils contribute to proper developmental patterning of organs by targeting growth factors to the right location in the extracellular matrix (1, 2).Molecules of fibrillin-1 (3), fibrillin-2 (4, 5), and fibrillin-3 (6) polymerize to form the backbone structure of microfibrils. Latent TGFβ-binding protein (LTBP)3-1 associates with fibrillin microfibrils in the perichondrium and in osteoblast cultures (7, 8), and LTBP-1 and LTBP-4 interact with fibrillin (9). Other proteins associated with fibrillin microfibrils include the fibulins (10, 11), microfibril-associated glycoprotein-1 and -2 (12, 13), decorin (14), biglycan (15), versican (16), and perlecan (17). It is likely that one function of these associated extracellular matrix molecules is to connect the fibrillin microfibril scaffold to other architectural elements in tissue- and organ-specific patterns.In addition to performing architectural functions, fibrillins bind directly to prodomains of bone morphogenetic proteins and growth and differentiation factors (18, 19) and LTBPs bring with them the small latent TGFβ complex (20), suggesting that the microfibril scaffold may position, concentrate, and control growth factor signaling. Studies of fibrillin-1 (Fbn1) and fibrillin-2 (Fbn2) mutant mice demonstrate that loss of fibrillins results in phenotypes associated with dysregulated TGFβ (2123) or bone morphogenetic protein (24) signaling. Microfibril-associated glycoprotein-1 (Magp-1) null mice reveal phenotypes that may also be related to abnormal TGFβ signaling (25).In a previous study (9), we determined that the binding site for LTBP-1 and -4 is contained within a specific four-domain region of fibrillin-1. In this study, we performed additional experiments to more precisely define the LTBP binding site. At the same time, we compared binding of fibulins to fibrillin, because the region in fibrillin-1 that was suggested to contain the fibulin binding site (11) was very close to our region of interest for LTBP binding. Our results demonstrate that LTBPs and fibulins compete for binding to fibrillin-1. However, the proteins tested (LTBP-1, LTBP-4, fibulin-2, fibulin-4, and fibulin-5) displayed “exquisite specificities” in their interactions with fibrillin-1.To test the potential significance of these interactions with fibrillin-1, we investigated matrix incorporation of LTBPs in cell cultures obtained from wild type, Fbn1 null, Fbn2 null, fibulin-2 (Fbln-2) null, and fibulin-4 (Fbln-4) null mice. In addition, we examined the distribution of LTBPs in Fbn1 null and Fbn2 null mice.  相似文献   

11.
Ubiquitination and deubiquitination of receptor-interacting protein 1 (RIP1) play an important role in the positive and negative regulation of the tumor necrosis factor α (TNFα)-induced nuclear factor κB (NF-κB) activation. Using a combination of functional genomic and proteomic approaches, we have identified ubiquitin-specific peptidase 21 (USP21) as a deubiquitinase for RIP1. USP21 is constitutively associated with RIP1 and deubiquitinates RIP1 in vitro and in vivo. Notably, knockdown of USP21 in HeLa cells enhances TNFα-induced RIP1 ubiquitination, IκB kinase β (IKKβ), and NF-κB phosphorylation, inhibitor of NF-κB α (IκBα) phosphorylation and ubiquitination, as well as NF-κB-dependent gene expression. Therefore, our results demonstrate that USP21 plays an important role in the down-regulation of TNFα-induced NF-κB activation through deubiquitinating RIP1.  相似文献   

12.
13.
14.
The Auxin-Binding Protein 1 (ABP1) was identified over 30 years ago thanks to it''s high affinity for active auxins. ABP1 plays an essential role in plant life yet to this day, its function remains ‘enigmatic.’ A recent study by our laboratory shows that ABP1 is critical for regulation of the cell cycle, acting both in G1 and at the G2/M transition. We showed that ABP1 is likely to mediate the permissive auxin signal for entry into the cell cycle. These data were obtained by studying a conditional functional knock-out of ABP1 generated by cellular immunization in the model tobacco cell line, Bright Yellow 2.Key Words: auxin responses, auxin-binding protein 1, immunomodulation, cellular immunisation  相似文献   

15.
Synapses are asymmetric intercellular junctions that mediate neuronal communication. The number, type, and connectivity patterns of synapses determine the formation, maintenance, and function of neural circuitries. The complexity and specificity of synaptogenesis relies upon modulation of adhesive properties, which regulate contact initiation, synapse formation, maturation, and functional plasticity. Disruption of adhesion may result in structural and functional imbalance that may lead to neurodevelopmental diseases, such as autism, or neurodegeneration, such as Alzheimer''s disease. Therefore, understanding the roles of different adhesion protein families in synapse formation is crucial for unraveling the biology of neuronal circuit formation, as well as the pathogenesis of some brain disorders. The present review summarizes some of the knowledge that has been acquired in vertebrate and invertebrate genetic model organisms.Synapses are asymmetric, intercellular junctions that are the basic structural units of neuronal transmission. The correct development of synaptic specializations and the establishment of appropriate connectivity patterns are crucial for the assembly of functional neuronal circuits. Improper synapse formation and function may cause neurodevelopmental disorders, such as mental retardation (MsR) and autism spectrum disorders (ASD) (McAllister 2007; Sudhof 2008), and likely play a role in neurodegenerative disorders, such as Alzheimer''s disease (AD) (Haass and Selkoe 2007).At chemical synapses (reviewed in Sudhof 2004; Zhai and Bellen 2004; Waites et al. 2005; McAllister 2007; Jin and Garner 2008), the presynaptic compartment contains synaptic vesicles (SV), organized in functionally distinct subcellular pools. A subset of SVs docks to the presynaptic membrane around protein-dense release sites, named active zones (AZ). Upon the arrival of an action potential at the terminal, the docked and “primed” SVs fuse with the plasma membrane and release neurotransmitter molecules into the synaptic cleft. Depending on the type of synapse (i.e., excitatory vs. inhibitory synapses), neurotransmitters ultimately activate an appropriate set of postsynaptic receptors that are accurately apposed to the AZ.Synapse formation occurs in several steps (Fig. 1) (reviewed in Eaton and Davis 2003; Goda and Davis 2003; Waites et al. 2005; Garner et al. 2006; Gerrow and El-Husseini 2006; McAllister 2007). Spatiotemporal signals guide axons through heterogeneous cellular environments to contact appropriate postsynaptic targets. At their destination, axonal growth cones initiate synaptogenesis through adhesive interactions with target cells. In the mammalian central nervous system (CNS), immature postsynaptic dendritic spines initially protrude as thin, actin-rich filopodia on the surface of dendrites. Similarly, at the Drosophila neuromuscular junction (NMJ), myopodia develop from the muscles (Ritzenthaler et al. 2000). The stabilization of intercellular contacts and their elaboration into mature, functional synapses involves cytoskeletal arrangements and recruitment of pre- and postsynaptic components to contact sites in spines and boutons. Conversely, retraction of contacts results in synaptic elimination. Both stabilization and retraction sculpt a functional neuronal circuitry.Open in a separate windowFigure 1.(A–C) Different stages of synapse formation. (A) Target selection, (B) Synapse assembly, (C) Synapse maturation and stabilization. (D–F) The role of cell adhesion molecules in synapse formation is exemplified by the paradigm of N-cadherin and catenins in regulation of the morphology and strength of dendritic spine heads. (D) At an early stage the dendritic spines are elongated from motile structures “seeking” their synaptic partners. (E) The contacts between the presynaptic and postsynaptic compartments are stabilized by recruitment of additional cell adhesion molecules. Adhesional interactions activate downstream pathways that remodel the cytoskeleton and organize pre- and postsynaptic apparatuses. (F) Cell adhesion complexes, stabilized by increased synaptic activity, promote the expansion of the dendritic spine head and the maturation/ stabilization of the synapse. Retraction and expansion is dependent on synaptic plasticity.In addition to the plastic nature of synapse formation, the vast heterogeneity of synapses (in terms of target selection, morphology, and type of neurotransmitter released) greatly enhances the complexity of synaptogenesis (reviewed in Craig and Boudin 2001; Craig et al. 2006; Gerrow and El-Husseini 2006). The complexity and specificity of synaptogenesis relies upon the modulation of adhesion between the pre- and postsynaptic components (reviewed in Craig et al. 2006; Gerrow and El-Husseini 2006; Piechotta et al. 2006; Dalva et al. 2007; Shapiro et al. 2007; Yamada and Nelson 2007; Gottmann 2008). Cell adhesive interactions enable cell–cell recognition via extracellular domains and also mediate intracellular signaling cascades that affect synapse morphology and organize scaffolding complexes. Thus, cell adhesion molecules (CAMs) coordinate multiple synaptogenic steps.However, in vitro and in vivo studies of vertebrate CAMs are often at odds with each other. Indeed, there are no examples of mutants for synaptic CAMs that exhibit prominent defects in synapse formation. This apparent “resilience” of synapses is probably caused by functional redundancy or compensatory effects among different CAMs (Piechotta et al. 2006). Hence, studies using simpler organisms less riddled by redundancy, such as Caenorhabditis elegans and Drosophila, have aided in our understanding of the role that these molecules play in organizing synapses.In this survey, we discuss the roles of the best characterized CAM families of proteins involved in synaptogenesis. Our focus is to highlight the complex principles that govern the molecular basis of synapse formation and function from a comparative perspective. We will present results from cell culture studies as well as in vivo analyses in vertebrate systems and refer to invertebrate studies, mainly performed in Drosophila and C. elegans, when they have provided important insights into the role of particular CAM protein families. However, we do not discuss secreted factors, for which we refer the reader to numerous excellent reviews (as for example Washbourne et al. 2004; Salinas 2005; Piechotta et al. 2006; Shapiro et al. 2006; Dalva 2007; Yamada and Nelson 2007; Biederer and Stagi 2008; Salinas and Zou 2008).  相似文献   

16.
Maspin is a serpin that has multiple effects on cell behavior, including inhibition of migration. How maspin mediates these diverse effects remains unclear, as it is devoid of protease inhibitory activity. We have previously shown that maspin rapidly inhibits the migration of vascular smooth muscle cells (VSMC), suggesting the involvement of direct interactions with cell surface proteins. Here, using immunofluorescence microscopy, we demonstrate that maspin binds specifically to the surface of VSMC in the dedifferentiated, but not the differentiated, phenotype. Ligand blotting of VSMC lysates revealed the presence of several maspin-binding proteins, with a protein of 150 kDa differentially expressed between the two VSMC phenotypes. Western blotting suggested that this protein was the β1 integrin subunit, and subsequently both α3β1 and α5β1, but not αvβ3, were shown to associate with maspin by coimmunoprecipitation. Specific binding of these integrins was also observed using maspin-affinity chromatography, using HT1080 cell lysates. Direct binding of maspin to α5β1 was confirmed using a recombinant α5β1-Fc fusion protein. Using conformation-dependent anti-β1 antibodies, maspin binding to VSMC was found to lead to a decrease in the activation status of the integrin. The functional involvement of α5β1 in mediating the effect of maspin was established by the inhibition of migration of CHO cells overexpressing human α5 integrin, but not those lacking α5 expression. Our observations suggest that maspin engages in specific interactions with a limited number of integrins on VSMC, leading to their inactivation, and that these interactions are responsible for the effects of maspin in the pericellular environment.Maspin is a member of the serpin family of serine protease inhibitors (SERPINB5).2 It was originally identified as a gene down-regulated in invasive breast cancer and proposed as a class II tumor suppressor (1), and has since been shown to have many effects on cellular behavior that are consistent with this activity. It has been shown to decrease the proliferation, migration, and metastasis of tumor cells in vivo (1, 2) and their invasion in vitro (3, 4), and to increase apoptosis of endothelial cells (5) and inhibit angiogenesis (6). However, the cellular effects of maspin are not restricted to tumor cells, and we have demonstrated that maspin can inhibit the migration of vascular smooth muscle cells (7).VSMC migration is a key event in the development of atherosclerosis (8), and contributes significantly to restenosis after angioplasty (9) and transplant arteriosclerosis (10). VSMC are not terminally differentiated and acquire migratory capacity as part of a phenotypic switch from a contractile, quiescent state to a dedifferentiated phenotype, characterized by proliferation and increased extracellular matrix synthesis, in addition to motility (11). This allows VSMC to respond to environmental cues following vascular injury. The phenotypic plasticity of VSMC is regulated by an array of signals, among which integrin-mediated association with surrounding extracellular matrix and changes in the expression of matrix-degrading proteases are prominent (1214).How maspin mediates its various cellular effects is unclear. Maspin has been reported to be an inhibitor of plasminogen activation (3, 15, 16), but we have shown that maspin is unable to inhibit either uPA- or tPA-catalyzed plasminogen activation under conditions in which the serpin PAI-1 was completely inhibitory (7). The anti-proteolytic inhibitory mechanism of serpins is dependent on characteristics of the reactive center loop (RCL) allowing it to adopt the necessary canonical conformation and rearrangements subsequent to protease binding (17). The RCL of maspin does not have the required characteristics (7, 18), and the conclusion that maspin is a non-inhibitory serpin is fully supported by its crystal structure (19, 20).Another confounding factor in understanding the mechanisms underlying the cellular effects of maspin is that, in common with the serpin PAI-2, it lacks an authentic secretion signal sequence. Nevertheless it has been shown to enter secretory vesicles (21) and is found extracellularly, in the cytoplasm and also in the nucleus (21, 22). Cytoplasmic and nuclear binding proteins for maspin have been identified (2325), and may be responsible for its effects on proliferation and apoptosis. How secreted, extracellular maspin exerts its effects is unclear, but a function as a cell signaling ligand has been proposed (2628). However, the characteristics of the maspin inhibitory effect on VSMC migration point to a more direct effect of maspin.To determine the mechanism of the maspin effect on VSMC migration, we have now attempted to identify maspin-binding proteins on the surface of these cells. In this report we provide biochemical, cellular, and functional evidence that the effect of maspin on cell migration is mediated by specific binding to cell adhesion receptors of the integrin family. We find that maspin binds specifically to β1 integrins on the surface of dedifferentiated VSMC, which leads to a reduction in the activation status of the integrin, and that the binding of maspin to α5β1 is sufficient for its inhibitory effects on cell migration and may represent a more general mechanism underlying its diverse biological effects.  相似文献   

17.
The mTOR (mammalian target of rapamycin) promotes growth in response to nutrients and growth factors and is deregulated in numerous pathologies, including cancer. The mechanisms by which mTOR senses and regulates energy metabolism and cell growth are relatively well understood, whereas the molecular events underlining how it mediates survival and proliferation remain to be elucidated. Here, we describe the existence of the mTOR splicing isoform, TORβ, which, in contrast to the full-length protein (mTORα), has the potential to regulate the G1 phase of the cell cycle and to stimulate cell proliferation. mTORβ is an active protein kinase that mediates downstream signaling through complexing with Rictor and Raptor proteins. Remarkably, overexpression of mTORβ transforms immortal cells and is tumorigenic in nude mice and therefore could be a proto-oncogene.  相似文献   

18.
Our previous study demonstrated that transforming growth factor (TGF)-β activates β-catenin signaling through Smad3 interaction with β-catenin in chondrocytes. In the present studies, we further investigated the detailed molecular mechanism of the cross-talk between TGF-β/Smad3 and Wnt/β-catenin signaling pathways. We found that C-terminal Smad3 interacted with both the N-terminal region and the middle region of β-catenin protein in a TGF-β-dependent manner. Both Smad3 and Smad4 were required for the interaction with β-catenin and protected β-catenin from an ubiquitin-proteasome-dependent degradation. In addition, the formation of the Smad3-Smad4-β-catenin protein complex also mediated β-catenin nuclear translocation. This Smad3-mediated regulatory mechanism of β-catenin protein stability enhanced the activity of β-catenin to activate downstream target genes during chondrogenesis. Our findings demonstrate a novel mechanism between TGF-β and Wnt/β-catenin signaling pathways during chondrocyte development.  相似文献   

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