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 15 毫秒
1.
The transforming growth factor-betas (TGF-betas) are synthesized as precursor proteins that are modified intracellularly prior to secretion. One of the most relevant intracellular modifications is the cleavage of the C-terminal pro-region from the N-terminal portion of the protein. The C-terminal pro-region is referred to as the latency-associated peptide (LAP) while the N-terminal region is called the mature TGF-beta or active TGF-beta. However, with some exceptions the LAP noncovalently associates with the mature TGF-beta prior to secretion. When the mature TGF-beta is associated with the LAP it is called L-TGF-beta and cannot interact with its receptor and has no biological effect. The TGF-betas and their receptors are very ubiquitously expressed, suggesting that the regulation of TGF-beta activity is likely to be complex and multifactorial. However, one of the most important means of controlling the biological effects of TGF-beta is the regulation of converting L-TGF-beta to active TGF-beta. The current literature supports two major mechanisms of activation of L-TGF-beta and suggests that the mechanism of activation of L-TGF-beta may be varied and context-dependent. For TGF-beta to become biologically active the LAP has to be either released from its associations with L-TGF-beta or undergo conformational change such that the LAP is not released from the L-TGF-beta complex but exposes the TGF-beta receptor binding site. Since TGF-beta has been associated with the pathogenesis of numerous diseases, the various mechanisms of activation of L-TGF-beta in context offer the possibility of controlling TGF-beta activity localized to the organ of involvement and to a more specific disease process.  相似文献   

2.
3.
Transforming growth factor beta (TGF-beta) ligands exert their biological effects through type II (TbetaRII) and type I receptors (TbetaRI). Unlike TGF-beta1 and -beta3, TGF-beta2 appears to require the co-receptor betaglycan (type III receptor, TbetaRIII) for high affinity binding and signaling. Recently, the TbetaRIII null mouse was generated and revealed significant non-overlapping phenotypes with the TGF-beta2 null mouse, implying the existence of TbetaRIII independent mechanisms for TGF-beta2 signaling. Because a variant of the type II receptor, the type II-B receptor (TbetaRII-B), has been suggested to mediate TGF-beta2 signaling in the absence of TbetaRIII, we directly tested the ability of TbetaRII-B to bind TGF-beta2. Here we show that the soluble extracellular domain of the type II-B receptor (sTbetaRII-B.Fc) bound TGF-beta1 and TGF-beta3 with high affinity (K(d) values = 31.7 +/- 22.8 and 74.6 +/- 15.8 pm, respectively), but TGF-beta2 binding was undetectable at corresponding doses. Similar results were obtained for the soluble type II receptor (sTbetaRII.Fc). However, sTbetaRII.Fc or sTbetaRII-B.Fc in combination with soluble type I receptor (sTbetaRI.Fc) formed a high affinity complex that bound TGF-beta2, and this complex inhibited TGF-beta2 in a biological inhibition assay. These results show that TGF-beta2 has the potential to signal in the absence of TbetaRIII when sufficient TGF-beta2, TbetaRI, and TbetaRII or TbetaRII-B are present. Our data also support a cooperative model for receptor-ligand interactions, as has been suggested by crystallization studies of TGF-beta receptors and ligands. Our cell-free binding assay system will allow for testing of models of receptor-ligand complexes prior to actual solution of crystal structures.  相似文献   

4.
Germ line mutations in one of two distinct genes, endoglin or ALK-1, cause hereditary hemorrhagic telangiectasia (HHT), an autosomal dominant disorder of localized angiodysplasia. Both genes encode endothelial cell receptors for the transforming growth factor beta (TGF-beta) ligand superfamily. Endoglin has homology to the type III receptor, betaglycan, although its exact role in TGF-beta signaling is unclear. Activin receptor-like kinase 1 (ALK-1) has homology to the type I receptor family, but its ligand and corresponding type II receptor are unknown. In order to identify the ligand and type II receptor for ALK-1 and to investigate the role of endoglin in ALK-1 signaling, we devised a chimeric receptor signaling assay by exchanging the kinase domain of ALK-1 with either the TGF-beta type I receptor or the activin type IB receptor, both of which can activate an inducible PAI-1 promoter. We show that TGF-beta1 and TGF-beta3, as well as a third unknown ligand present in serum, can activate chimeric ALK-1. HHT-associated missense mutations in the ALK-1 extracellular domain abrogate signaling. The ALK-1/ligand interaction is mediated by the type II TGF-beta receptor for TGF-beta and most likely through the activin type II or type IIB receptors for the serum ligand. Endoglin is a bifunctional receptor partner since it can bind to ALK-1 as well as to type I TGF-beta receptor. These data suggest that HHT pathogenesis involves disruption of a complex network of positive and negative angiogenic factors, involving TGF-beta, a new unknown ligand, and their corresponding receptors.  相似文献   

5.
A low molecular weight inhibitor of TGF-beta 1 binding was detected in partially purified human platelet extracts by using Hep 3B hepatoma cells in the binding assays. The inhibitory protein was purified to homogeneity and was identified as platelet factor 4 on the basis of its amino acid sequence. TGF-beta 1 binding to Hep 3B cells was almost completely inhibited by 100 nM concentrations of platelet factor 4, but TGF-beta 1 binding to NRK 49F fibroblasts was inhibited only slightly. Affinity cross-linking experiments revealed that these differences in the inhibition of TGF-beta 1 binding by platelet factor 4 were due to differences in the complements of TGF-beta 1 binding proteins present on these two cell types. In Hep 3B cells the majority of bound TGF-beta 1 was cross-linked to a complex which had an apparent molecular weight of 70 kDa. TGF-beta 1 binding to this protein was the most sensitive to inhibition by platelet factor 4. Based on its size and TGF-beta 1 binding properties, we believe this protein is the type I TGF-beta 1 receptor. Hep 3B cells also had a high-affinity TGF-beta 1 binding protein which appeared as an 80 kDa complex, and which we believe to be the type II TGF-beta 1 receptor. TGF-beta 1 binding to this protein was not inhibited by platelet factor 4. TGF-beta 1 was also cross-linked to complexes of higher molecular weights in Hep 3B cells, but it was not clear whether any of them represented the type III TGF-beta 1 receptor. In NRK 49F cells, the majority of bound TGF-beta 1 was cross-linked to a high molecular weight complex which probably represented the type III TGF-beta 1 receptor. NRK 49F cells also had type I TGF-beta 1 receptors and platelet factor 4 inhibited binding to these receptors in the NRK cells. Since the type I receptor contributed only a small percentage of total TGF-beta 1 binding, however, the overall effects of platelet factor 4 on TGF-beta 1 binding to NRK 49F cells were negligible. We were unable to demonstrate specific or saturable binding of platelet factor 4 to Hep 3B cells using either direct binding or affinity cross-linking assays. Thus, it is not clear whether platelet factor 4 inhibits TGF-beta 1 binding by competition for binding to the type I receptor. Modest concentrations of TGF-beta 1 reduced the adherence of Hep 3B cells to tissue culture dishes.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

6.
7.
Transforming growth factor beta (TGF-beta) signals through three high affinity cell surface receptors, TGF-beta type I, type II, and type III receptors. The type III receptor, also known as betaglycan, binds to the type II receptor and is thought to act solely by "presenting" the TGF-beta ligand to the type II receptor. The short cytoplasmic domain of the type III receptor is thought to have no role in TGF-beta signaling because deletion of this domain has no effect on association with the type II receptor, or with the presentation role of the type III receptor. Here we demonstrate that the cytoplasmic domains of the type III and type II receptors interact specifically in a manner dependent on the kinase activity of the type II receptor and the ability of the type II receptor to autophosphorylate. This interaction results in the phosphorylation of the cytoplasmic domain of the type III receptor by the type II receptor. The type III receptor with the cytoplasmic domain deleted is able to bind TGF-beta, to bind the type II receptor, and to enhance TGF-beta binding to the type II receptor but is unable to enhance TGF-beta2 signaling, determining that the cytoplasmic domain is essential for some functions of the type III receptor. The type III receptor functions by selectively binding the autophosphorylated type II receptor via its cytoplasmic domain, thus promoting the preferential formation of a complex between the autophosphorylated type II receptor and the type I receptor and then dissociating from this active signaling complex. These studies, for the first time, elucidate important functional roles of the cytoplasmic domain of the type III receptor and demonstrate that these roles are essential for regulating TGF-beta signaling.  相似文献   

8.
9.
R A Ignotz  J Massagué 《Cell》1987,51(2):189-197
Transforming growth factor-beta (TGF-beta) increases the incorporation of fibronectin and type I collagen into the extracellular matrix of fibroblasts and epithelial cells and enhances the attachment of thymocytes onto a fibronectin substratum. Investigation of the molecular basis for these effects showed that TGF-beta elevates specifically the expression of cell adhesion protein receptors. Treatment of cells with either form of TGF-beta, TGF-beta 1, or TGF-beta 2, increases the rate of receptor synthesis and the level of receptors on the cell surface. TGF-beta acts via two complementary mechanisms, elevation of receptor mRNA and faster kinetics of receptor beta subunit precursor to product conversion. The results show that the expression of cell adhesion receptors is susceptible to pretranslational and posttranslational regulation by factors that control cell morphology, proliferation, and differentiation such as TGF-beta.  相似文献   

10.
Type beta transforming growth factor (TGF-beta) is found in large amounts in bone tissue, and is a potent mitogen for osteoblast-enriched cell cultures obtained from fetal rat parietal bone. Because other local and systemic factors may be presented to bone cells simultaneously with TGF-beta, it is important to understand the effects of this complex growth regulator in such circumstances. Unlike the effects observed in many tissue systems, TGF-beta does not invariably inhibit the mitogenic response of bone cells to other growth promoters. In contrast, other factors such as epidermal growth factor (EGF), basic fibroblast growth factor (bFGF), and type alpha tumor necrosis factor (TNF-alpha) limit the response of osteoblastic bone cells to TGF-beta. TGF-beta is a much weaker mitogen for fibroblastic cells obtained from fetal rat bone, whereas fetal bovine serum, EGF, bFGF, and TNF-alpha are more potent stimulators. In addition, TGF-beta does not significantly impair the response of the fibroblastic bone cells to the other tested agents. These findings reinforce a role of TGF-beta as an anabolic bone growth regulator, and suggest that its function may be modified by other local or systemic agents that can also affect bone cells.  相似文献   

11.
12.
Transforming growth factor beta type 1 (TGF-beta 1) was reacted with NHS-biotin to yield a derivative of TGF-beta 1 which was biotinylated on lysine residues. The biotinylated form of TGF-beta 1 was separated from the unreacted material by reverse phase chromatography. In three separate bioassays, the derivatized peptide was as active as the starting material. The use of FITC-avidin in conjunction with flow cytometry demonstrated that the binding of biotinylated TGF-beta 1 to its receptor is saturable, competable, and specific. A 100-fold molar excess of underivatized TGF-beta 1 gave 85% inhibition of binding of the biotinylated peptide to the mink lung cell line CCL-64, while TGF-beta 2 showed no inhibition of binding, nor did insulin, calcitonin, or TGF-alpha. Both CCL-64 cells and human umbilical vein endothelial cells showed a density-dependent down-regulation of receptor expression in culture. Several factors were examined that might mediate this effect. The down-regulation was shown not to be due to the secretion of an active form of TGF-beta 1. The extracellular matrix from high-density cells did not decrease expression of the receptor. Fibronectin, collagen, and gelatin were also unable to signal changes in receptor expression, even though in other systems such matrix components can regulate the responsiveness of cells to TGF-beta 1. Lastly, staining simultaneously for DNA content and TGF-beta 1 receptor expression showed that there was no correlation between cell cycle and receptor levels.  相似文献   

13.
Mature transforming growth factor-beta (TGF-beta) is proteolytically derived from the C terminus of a precursor protein. Latency-associated protein (LAP), the N-terminal remnant of the TGF-beta precursor, is able to bind and neutralize TGF-beta. Mature TGF-beta exerts its activity by binding and complexing members of two subfamilies of receptors, the type I and II receptors. In addition to these signaling receptors, TGF-beta can also interact with an accessory receptor termed the type III receptor. Using a surface plasmon resonance-based biosensor (BIAcore), we determined the mechanisms of interaction of four binding proteins (LAP, the type II and III receptor ectodomains (EDs), and a type II receptor ED/Fc chimera) with three TGF-beta isoforms, and we quantified their related kinetic parameters. Using global fitting based on a numerical integration data analysis method, we demonstrated that LAP and the type II receptor/Fc chimera interacted with the TGF-beta isoforms with a 1:1 stoichiometry. In contrast, the type II ED interactions with TGF-beta were best fit by a kinetic model assuming the presence of two independent binding sites on the ligand molecule. We also showed that the type III ED bound two TGF-beta molecules. Further experiments revealed that LAP was able to block the interactions of TGF-beta with the two EDs, but that the two EDs did not compete or cooperate with each other. Together, these results strongly support the existence of a cell-surface complex consisting of one type III receptor, two TGF-beta molecules, and four type II receptors, prior to the recruitment of the type I receptor for signal transduction. Additionally, our results indicate that the apparent dissociation rate constants are more predictive of the neutralizing potency of these TGF-beta-binding proteins (LAP, the type II and III receptor EDs, and the type II receptor/Fc chimera) than the apparent equilibrium constants.  相似文献   

14.
Transforming growth factor-beta (TGF-beta) is a bimodal regulator of cellular growth. The cellular effects of TGF-beta depend on the intensity of signals emanating from TGF-beta receptors. Low levels of receptor activity are sufficient to stimulate cell proliferation, while higher degrees of receptor activation are associated with growth inhibition. To study the mechanisms of these effects, a tetracycline-inducible expression system was used to overexpress type II TGF-beta receptors in NIH 3T3 fibroblasts. Overexpressed type II TGF-beta receptors suppressed fibroblast proliferation elicited by TGF-beta1, fibroblast growth factor (FGF) or platelet-derived growth factor (PDGF). Accompanying these anti-proliferative effects, increases in extracellular-signal regulated kinase (ERK) and c-Jun N-terminal kinase (JNK) activity were detected. Furthermore, PDGF alpha-, but not PDGF beta-receptor protein levels were reduced by type II TGF-beta receptor overexpression. In conclusion, our system is an excellent tool to study the molecular mechanisms of growth inhibition by TGF-beta in fibroblasts. Activation of JNK and ERK, or modulation of PDGF receptor expression may be involved in this process.  相似文献   

15.
TGF-betas and TGF-beta receptors in atherosclerosis   总被引:4,自引:0,他引:4  
  相似文献   

16.
Loss of growth regulation by transforming growth factor-beta (TGF-beta) may be an important step in carcinogenesis. We have used a cell fusion system to show that inhibition of growth by TGF-beta can be restored to carcinoma cell lines that are unresponsive to the inhibitory effects of TGF-beta. In a previous study, the EJ bladder carcinoma line was fused to the SW480 colon adenocarcinoma line and found to produce nontumorigenic hybrid cells along with one hybrid cell clone of low tumorigenicity. Here we show that the capacity of the nontumorigenic hybrid cells to respond to either TGF-beta 1 or TGF-beta 2 has been restored, while the parental or tumorigenic hybrid cells show little or no inhibition of growth following TGF-beta treatment. Cross-linking analyses with labeled TGF-beta 1 demonstrated much higher levels of the type II (85 kDa) receptor in the hybrid cells compared with the parental tumor lines. Both the parental and tumorigenic hybrid cell lines were capable of responding to TGF-beta as evidenced by increased levels of mRNA for fibronectin, type IV collagenase, and plasminogen activator inhibitor after treatment with TGF-beta 1. These results suggest that the type II receptor is necessary for mediating the effects of TGF-beta on inhibition of growth but not on gene activation of the hybrid cells.  相似文献   

17.
Transforming growth factor-beta 1 (TGF-beta 1) is a potent mito-inhibiting substance that is thought to play an important function in regulating hepatocyte proliferation during liver regeneration. In this investigation, we have shown by immunohistochemistry that hepatocytes containing significant intracellular concentrations of TGF-beta 1 12 h after a two-thirds partial hepatectomy. This increase in hepatocyte TGF-beta 1 concentration was initially confined to those cells that resided in the periportal region of the liver. The elevation of intracellular TGF-beta 1 was, however, transient, and within 36 h, the hepatocytes positive for TGF-beta 1 had changed in a wavelike fashion from the periportal to the pericentral region of the liver lobules. By 48 h, most hepatocytes no longer contained TGF-beta 1. Interestingly, this temporary increase in TGF-beta 1 always preceded the onset of hepatocyte replication by approximately 3-6 h. Since TGF-beta 1 mRNA has been shown to be absent from hepatocytes normally and throughout liver regeneration, these results imply that the increase in intracellular TGF-beta 1 resulted from an augmented uptake. We have further shown that the insulin-like growth factor-II/mannose 6-phosphate (IGF-II/Man-6-P) receptors were up-regulated during liver regeneration and that the increased expression of this receptor co-localized in those hepatocytes containing elevated concentrations of TGF-beta 1. The latent TGF-beta 1 phosphomannosyl glycoprotein complex has been shown to bind to the IGF-II/Man-6-P receptor. Therefore, our data are consistent with the hypothesis that this latent complex is internalized through the IGF-II/Man-6-P receptor to the intracellular acidic prelysosomal/endosomal compartments where the mature TGF-beta 1 molecule could be activated by dissociation from the latent complex.  相似文献   

18.
19.
Mast cell activation through the high affinity IgE receptor (FcepsilonRI) is a critical component of atopic inflammation. The cytokine TGF-beta1 has been shown to inhibit IgE-dependent mast cell activation, possibly serving to dampen mast cell-mediated inflammatory responses. We present proof that TGF-beta1 inhibits mast cell FcepsilonRI expression through a reversible pathway that diminishes protein, but not mRNA, expression of the FcepsilonRI subunit proteins alpha, beta, and gamma. The stability of the expressed proteins and the assembled cell surface complex was unaltered by TGF-beta1 treatment. However, TGF-beta1 decreased the rate of FcepsilonRI beta-chain synthesis, arguing that this inhibitory cytokine exerts its effects at the level of mRNA translation. TGF-beta1 consistently diminished FcepsilonRI expression on cultured human or mouse mast cells as well as freshly isolated peritoneal mast cells. The related cytokines, TGF-beta2 and TGF-beta3, had similar effects. We propose that TGF-beta1 acts as a negative regulator of mast cell function, in part by decreasing FcepsilonRI expression.  相似文献   

20.
We have previously demonstrated that activation of the Ras/Mapk pathways is required for transforming growth factor beta (TGF-beta) induction of TGF-beta(1) expression. Here we examined the role of the Ras/Mapk pathways in TGF-beta induction of urokinase-type plasminogen activator receptor (uPAR) expression in untransformed intestinal epithelial cells (IECs). TGF-beta activated the stress-activated protein kinases (Sapk)/c-Jun N-terminal kinases (JNKs) within 5-10 min, an effect that preceeded TGF-beta induction of uPAR expression in these cells. TGF-beta induction of both JNK1 activity and JunD phosphorylation was blocked by expression of a dominant-negative mutant of the type II TGF-beta receptor (DN TbetaRII), a dominant-negative mutant of MKK4 (DN MKK4), or a dominant-negative mutant of Ras (RasN17), or by the addition of the JNK inhibitor SP600125. TGF-beta also induced AP-1 complex formation at the distal AP-1 site (-184 to -178) of the uPAR promoter within 2 h of TGF-beta addition, consistent with the time-dependent up-regulation of uPAR expression. The primary components present in the TGF-beta-stimulated AP-1 complex bound to the uPAR promoter were Jun D and Fra-2. Moreover, addition of SP600125, or expression of DN MKK4 or DN TbetaRII, blocked TGF-beta up-regulation of uPAR in IECs. Accordingly, our results indicate that TGF-beta activates the Ras/MKK4/JNK1 signaling cascade, leading to induction of AP-1 activity, which, in turn, up-regulates uPAR expression. Our results also indicate that the type II TGF-beta receptor (RII) is required for TGF-beta activation of JNK1 and the resulting up-regulation of uPAR expression.  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