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.
The bioactivity of both bFGF and aFGF in the BALB/MK-1 cell line has been compared to that of EGF. Our results indicate that, for that cell type, aFGF was far more potent than bFGF in inducing cell proliferation. In the presence of heparin, aFGF was as potent as EGF. In addition, excess bFGF has an inhibitory effect on the proliferation of MK cells exposed to a saturating concentration of aFGF, therefore acting as a partial agonist of aFGF. Surprisingly, bFGF, although it had low biological activity, was capable of synergizing the effect of EGF. In its presence, cultures exposed to saturating concentration of EGF have a final cell density 3- to 4-fold higher than that of counterpart cultures exposed to EGF alone. TGF beta, which in previous studies has been shown to inhibit the growth of keratinocytes, also inhibited the growth of BALB/MK-1 cells in response to either bFGF or aFGF. These studies suggest a role for FGF in regulating BALB/MK proliferation. aFGF provides positive growth signals which can be negatively modulated by excess bFGF or TGF beta, while bFGF, although a poor mitogen, could act by potentiating the effect of subsaturating concentrations of EGF.  相似文献   

2.
Acidic fibroblast growth factor (aFGF) and basic fibroblast growth factor (bFGF) are present in high levels in most areas of the embryonic rodent brain. To begin to understand the role of these growth factors in brain development, the effects of aFGF and bFGF on dissociated cell cultures prepared from embryonic and neonatal rat brain were studied. Addition of aFGF and heparin or bFGF alone to serum-free cultures of the dissociated Embryonic Day (E) 14.5 mesencephalon stimulates cell proliferation, as judged by [3H]thymidine autoradiography, leading to a maximal 75-fold increase in the total number of cells. This effect is dose-dependent with half-maximal increases at concentrations of about 5-6 ng/ml of aFGF or bFGF and is inhibited by the FGF antagonist HBGF-1U. The effect of aFGF on cell proliferation in cultures prepared from E14.5 mesencephalon is similar to that in cultures prepared from E14.5 cortex. However, in cultures prepared from E14.5 rhombencephalon or diencephalon, the proliferative effect of aFGF is much reduced. In all brain areas studied, the proliferative effect of aFGF declines with increasing age. Immunocytochemical analysis of E14.5 mesencephalic cultures demonstrated that the aFGF-induced increase in cell number is due to the proliferation of A2B5-immunoreactive (IR) glial precursor cells, but not of neuronal precursors, fibroblasts, or microglial cells. Moreover, differentiated glial fibrillary acidic protein-IR astrocytes and 2',3'-cyclic nucleotide 3'-phosphohydrolase-IR oligodendrocytes were not observed in cultures continuously treated with aFGF or bFGF, but were observed in high numbers after removal of the growth factors. These results suggest (1) that aFGF and bFGF are potent mitogens for glial precursor cells in all embryonic brain regions, (2) that the magnitude of the effects of aFGF depends on embryonic age and brain region, and (3) that both growth factors inhibit the differentiation of astrocyte or oligodendrocyte precursors. These observations made in vitro strongly support the hypothesis that FGF plays a critical role in gliogenesis and the timing of glial differentiation in the brain.  相似文献   

3.
Nerve growth factor (NGF) and acidic or basic fibroblast growth factor (aFGF and bFGF, respectively) induce neurite outgrowth from the rat pheochromocytoma cell line, PC12. The neurites induced by these three factors are stable for up to a month in cell culture in the continued presence of any of the above growth factors. bFGF (ED50 = 30 pg/ml) is 800 fold more potent in stimulating neurite outgrowth than aFGF (ED50 = 25 ng/ml) and 260 fold more potent than NGF (ED50 = 8 ng/ml). While the neurotropic activities of aFGF and NGF are potentiated by heparin, that of bFGF is both partially inhibited or stimulated, depending upon the concentration of bFGF. Radioreceptor binding experiments show that aFGF and bFGF bind to a common binding site on the PC12 cell surface. Affinity labeling studies demonstrate a single receptor with an apparent molecular weight of 145,000 daltons, which corresponds to the high molecular weight receptor identified in BHK-21 cells. NGF does not appear to compete with aFGF or bFGF for binding to the receptor. Heparin blocked the binding of bFGF to the receptor but had only a small inhibitory effect on the binding of aFGF to the receptor. Thus, it appears that heparin inhibition of the neurotropic effects of bFGF occurs, at least in part, by impairing the interaction of bFGF with the receptor, while having little effect on that of aFGF. The stimulatory effects of heparin on the neurotropic activity of aFGF, bFGF, and NGF may occur through a site not associated with the respective cellular receptor for the growth factors.  相似文献   

4.
When normal human foreskin keratinocytes were cultured in the absence of polypeptide growth factors at densities above 5 x 10(3)/cells cm2, the cells proliferated continuously and the addition of IGF-I, EGF, TGF alpha, bFGF, or aFGF did not significantly alter growth rate. Heparin sulfate, TGF beta, or suramin inhibited keratinocyte growth factor-independent proliferation. The addition of EGF, TGF alpha, or aFGF reversed heparin-induced growth inhibition, while bFGF partially negated this effect. RIA of keratinocyte-derived conditioned medium (CM) indicated the presence of TGF alpha peptide at a concentration of approximately 235 pg/ml. In contrast, clonal growth of keratinocytes required the addition of growth factors to the basal medium. Keratinocyte-derived CM replaced EGF in stimulating keratinocyte clonal growth, and an anti-EGF receptor mAb inhibited CM-induced keratinocyte clonal growth. In addition to its effect on keratinocytes, keratinocyte-derived CM stimulated the incorporation of [3H]thymidine by quiescent cultures of human foreskin fibroblasts, mouse AKR-2B cells, and EGF-receptorless mouse NR6 cells. CM-stimulated [3H]thymidine incorporation into quiescent normal human fibroblasts was partially reduced in the presence of anti-EGF receptor mAb. Heparin sulfate partially inhibited CM-induced keratinocyte clonal growth and [3H]thymidine incorporation into quiescent AKR-2B cells. We hypothesize from these data that autocrine and paracrine-acting factors produced by keratinocytes mediated their effect through the activation of both EGF receptor-dependent and EGF receptor-independent mitogenic pathways and that some of these factors appear to be sensitive to inhibition by heparin.  相似文献   

5.
Fresh bovine, porcine and canine hearts were homogenized and mitogens for mesoderm-derived cells were purified in three different steps. Extraction by two different ammonium sulfate precipitations was followed by cation-exchange chromatography and by heparin-Sepharose affinity chromatography. A heparin-Sepharose fraction from heart (eluted at 1.1 M NaCl) increased mitotic activity in serum-deprived cultures of porcine aortic endothelial and smooth muscle cells, and in human fibroblasts. This mitogenic activity is potentiated by heparin and inhibited by gamma-interferon. The heart mitogenic fraction showed one double peak on HPLC at A215 and one polypeptide band on SDS/PAGE. These peaks and bands were identical to those obtained from bovine brain. The heart acidic fibroblast growth factor (aFGF) showed a positive signal in Western blots using antibodies raised against brain aFGF. Gas-phase amino acid sequencing established that the mitogens were identical to aFGF and the N-terminally truncated aFGF. Extraction in the presence of a protease inhibitor (pepstatin A) produced a higher-molecular mass form of aFGF with a blocked amino terminus. Another mitogen, eluted at 1.6 M NaCl from heparin-Sepharose, reacted with polyclonal antiserum against human recombinant basic fibroblast growth factor (bFGF) and showed a 66% (12 from 18 amino acids determined by gas-phase sequencing) similarity with bFGF. This polypeptide increased the mitotic activity of the same cell lines but was more potent than aFGF.  相似文献   

6.
The effect of various growth factors on the synthesis of hyaluronan in human fibroblasts was investigated. When tested in medium containing 0.5% fetal calf serum, platelet-derived growth factor (PDGF)-BB was found to stimulate hyaluronan synthesis; the maximal response was equal to or higher than that obtained with 10% fetal calf serum. PDGF-AA gave only a limited effect, indicating that the stimulatory effect of PDGF on hyaluronan synthesis was mainly transduced via the B-type PDGF receptor. Epidermal growth factor (EGF), basic fibroblast growth factor (bFGF) and transforming growth factor (TGF)-beta 1 also stimulated hyaluronan synthesis; their effects were less than that of PDGF-BB, but combinations of factors produced potent stimulatory effects on hyaluronan synthesis. All factors stimulated hyaluronan synthesis in sparse as well as dense cultures. The effects of the factors on hyaluronan synthesis did not correlate with their mitogenic activities; PDGF-BB, EGF and bFGF are equipotent mitogens, but PDGF-BB had a much more potent effect on hyaluronan synthesis, and TGF-beta actually inhibits the growth of fibroblasts under the conditions of the assay.  相似文献   

7.
8.
Rabbit polyclonal antibodies were raised against ovalbumin conjugates of purified bovine brain acidic fibroblast growth factor (aFGF) and a synthetic peptide containing the N alpha-terminal 1-24 amino acid sequence of bovine basic fibroblast growth factor (bFGF). These antibodies were used to specifically detect 1-ng quantities of aFGF and bFGF by using enzyme-linked immunosorbent assay (ELISA) and Western immunoblot procedures. Antibodies raised against aFGF recognized bovine brain aFGF and bovine recombinant aFGF but very poorly recognized recombinant bFGF or purified porcine or bovine pituitary bFGF with ELISA and Western immunoblot procedures. Antibodies raised against bFGF (1-24) recognized purified bovine, porcine, and recombinant human bFGF but only very poorly recognized aFGF with ELISA and Western immunoblot procedures. In vitro addition of anti-bFGF antibodies was able to partially neutralize bFGF-stimulated 3H-thymidine incorporation by COMMA-D mouse mammary epithelial cells while having no effect on aFGF or epidermal growth factor (EGF) stimulation. In vitro addition of anti-aFGF antibodies had no effect on bFGF- or EGF-stimulated 3H-thymidine incorporation, but surprisingly, had a potentiating effect on aFGF stimulation. Antibodies against aFGF immobilized on protein A-Sepharose were able to specifically and completely remove mitogenic activity from solutions containing aFGF but had no effect on removal of mitogenic activity from control solutions containing bFGF or EGF. Similarly, immobilized anti-bFGF antibodies completely removed mitogenic activity from solutions of bFGF, but not aFGF or EGF controls. These antibodies have been useful for the identification and characterization of growth factors from tissue and recombinant sources.  相似文献   

9.
Keratinocyte growth factor (KGF) is a member of the fibroblast growth factor (FGF) family. KGF exhibits potent mitogenic activity for a variety of epithelial cell types but is distinct from other known FGFs in that it is not mitogenic for fibroblasts or endothelial cells. We report saturable specific binding of 125I-KGF to surface receptors on intact Balb/MK mouse epidermal keratinocytes. 125I-KGF binding was completed efficiently by acidic FGF (aFGF) but with 20-fold lower efficiency by basic FGF (bFGF). The pattern of 125I-acidic FGF binding and competition on Balb/MK keratinocytes and NIH/3T3 fibroblasts suggests that these cell types possess related but distinct FGF receptors. Scatchard analysis of 125I-KGF binding suggested major and minor high affinity receptor components (KD = 400 and 25 pM, respectively) as well as a third high capacity/low affinity heparin-like component. Covalent affinity cross-linking of 125I-KGF to its receptor on Balb/MK cells revealed two species of 115 and 140 kDa. KGF also stimulated the rapid tyrosine phosphorylation of a 90-kDa protein in Balb/MK cells but not in NIH/3T3 fibroblasts. Together these results indicate that Balb/MK keratinocytes possess high affinity KGF receptors to which the FGFs may also bind. However, these receptors are distinct from the receptor(s) for aFGF and bFGF on NIH/3T3 fibroblasts, which fail to interact with KGF.  相似文献   

10.
Polycationic molecules were studied either for their ability to displace the binding of basic fibroblast growth factor (bFGF) to high- and low-affinity membrane interaction sites and/or to modulate bFGF-induced proliferation of fibroblasts. Heparin-binding polypeptides, such as polylysine, protamine, histones, and thrombin-displaced [125I]bFGF bound to bovine brain membrane receptors. The most displacing polypeptides were those with the strongest affinity to heparin. Two of these polypeptides, protamine and polylysine, inhibited (at 5 microM) by more than 90% the mitogenic effect induced by bFGF on Chinese hamster lung fibroblast cells (CCL39). At the same dose, no effect was observed with basic proteins that do not bind to heparin, such as cytochrome C and lysozyme. An interesting observation was that protamine at 1 microM potentiated by 1.5-fold the mitogenic activity of bFGF, while it acted as an inhibitor at higher concentration.  相似文献   

11.
This study reports on the effects of heparin, basic and acidic fibroblast growth factors (bFGF and aFGF, respectively), and transforming growth factor type-e (TGFe) on the growth of a human adrenocortical carcinoma cell line, SW-13. Heparin has previously been shown to inhibit growth in several cell types, including smooth muscle cells, certain fibroblasts, and epithelial cells, and to modulate the effects of fibroblast growth factors. Whereas bFGF and aFGF bind tightly to heparin and elute from a heparin-Sepharose column with 2 M NaCl and 1.6 M NaCl, respectively, TGFe binds to heparin with lower affinity and can be eluted from heparin-Sepharose column with 0.5 M NaCl. TGFe is a polypeptide unrelated to FGF, is present in neoplastic and nonneoplastic tissues, and stimulates the growth of certain epithelial cells and fibroblasts in soft agar and monolayer. Since the growth of SW-13 cells is stimulated by TGFe and by bFGF, we hypothesized that heparin would inhibit the growth of SW-13 cells by binding to these growth factors and that the effects of heparin could be overcome with the addition of either growth factor. Our experiments confirmed that heparin inhibits the growth of SW-13 cells. A dose-dependent growth inhibition was observed in both monolayer and soft agar. The inhibition in monolayer was partially reversed upon heparin withdrawal. The effects of heparin in both monolayer and soft agar were at least partially overcome by TGFe and by basic or acidic FGF. Overall protein synthesis does not appear to be affected by heparin as measured by [35S]methionine uptake. In contrast, epidermal growth factor (EGF) and insulin-like growth factor I (IGF-I) were unable to overcome heparin-induced inhibition both in monolayer and in soft agar. Heparin also inhibited [3H]thymidine incorporation in AKR-2B and partially inhibited AKR-2B cell stimulation by TGFe; however, it further potentiated the already potent stimulation by bFGF. We propose that heparin, TGFe, bFGF, and aFGF modulate the growth of SW-13 cells and possibly of other epithelial cells in complex ways and that heparin-like substances present in the extracellular matrix play an important role in the control of epithelial growth.  相似文献   

12.
This article sumarizes the structural and biological properties of the family of fibroblast growth factors (FGF). Basic FGF (bFGF) and acidic FGF (aFGF) are the best characterized members of this family. bFGF and aFGF are potent modulators of cell proliferation, motility and differentiation. They are also potent angiogenesis factors in vivo. Some of the important biological characteristics of bFGF and aFGF discussed in the review include the affinity of bFGF and aFGF for heparin, their lack of secretion in culture and their association with extracellular matrix. Recently, several oncogenes, 40–50% homologous in sequence to bFGF and aFGF have been identified. These include int-2, hst, K-fgf and FGF-5. The structural and biological properties of these FGF-related oncogenes are also discussed.  相似文献   

13.
Summary The hormonal control of tissue inhibitor of metalloproteinase-1 (TIMP-1) gene expression and production by growth factors, gonadotrophins, and serum factors in cultured bovine granulosa cells (BGC) were investigated. Confluent cultures of BGC were exposed to various factors in a defined medium and levels of TIMP-1 in the conditioned medium were determined by enzyme immunoassay. Basic fibroblast growth factor (bFGF) and acidic fibroblast growth factor (aFGF) showed potent stimulation of cell proliferation and TIMP-1 production by BGC, while insulin stimulated growth but not TIMP-1 production. Basic FGF stimulated TIMP-1 production and BGC cell proliferation in a dose-dependent manner. A time course of TIMP-1 production showed substantially increased levels between 18 and 24 h in both control and bFGF-stimulated BGC cultures with bFGF-stimulated cultures having markedly higher TIMP-1 production at all time points. Consistent with the TIMP-1 production data, bFGF and aFGF increased the expression of TIMP-1 mRNA as determined by northern blot analysis, while insulin, inhibited TIMP-1 mRNA levels. These results indicate that FGF-induced TIMP-1 production by BGC may support bovine embryo development in vitro.  相似文献   

14.
Cultured endothelial cells have been shown to possess two mechanisms of intercellular adhesion: Ca2(+)-dependent and Ca2(+)-independent. We report here that growth of bovine aortic endothelial cells (BAEC) in complete medium containing purified basic fibroblast growth factor (bFGF, 6 ng/ml) results in loss of Ca2(+)-dependent intercellular adhesion. In the presence of heparin (90 micrograms/ml), this effect is reproduced upon treatment with acidic fibroblast growth factor (aFGF, 6 ng/ml) or endothelial cell growth supplement (ECGS, 100 micrograms/ml), in both human umbilical vein endothelial cells (HUVEC) and BAEC. Treatment at these doses with aFGF in the absence of heparin or with heparin alone is without significant effect. Loss of Ca2(+)-dependent adhesion following treatment of cells with heparin-binding growth factors (HBGFs) is prevented by pre-treatment of cell layers with cycloheximide. The Ca2(+)-independent adhesion mechanism is unaffected by HBGF treatment. Exposure of endothelial cells to HBGFs, moreover, prevents the eventual establishment of quiescence in growing cultures and restimulates replication in confluent cultures that have reached a final density-inhibited state. Addition of bFGF alone or aFGF + heparin at these doses results in a 4-fold increase in DNA synthesis over untreated control cultures at saturation density as reflected by thymidine index. A single addition of bFGF (6 ng/ml) to untreated quiescent confluent BAEC monolayers results in an increase in 3H-TdR incorporation reaching a peak at 22 hours with a parallel loss of Ca2(+)-dependent adhesiveness. Fluorescent staining with rhodamine-phalloidin demonstrates an altered distribution of polymerized F-actin in the bFGF-treated monolayers, marked by disruption of the dense peripheral microfilament bands retained by untreated confluent monolayers. Together, these results indicate that the mitogenic effect of HBGFs in cultured endothelial cells is associated with a "morphogenic" set of responses, perhaps dependent on breakdown of calcium-dependent cell-cell contacts.  相似文献   

15.
Acidic and basic fibroblast growth factors (aFGF and bFGF) belong to a family of structurally related polypeptides characterized by a high affinity for heparin. a and bFGF display mitogenic activity for many cell types. Biological activity is strongly potentiated by heparin which stabilizes their molecular conformation by preventing physicochemical or enzymatic degradation. In our previous study we have shown that a water-soluble derivatized dextran named DDE, containing 82.2% methyl carboxylic acid groups, 6.1% benzylamide, and 5.6% sulfonate with a specific anticoagulant activity equivalent to heparin of 0.5 IU/mg could potentiate the mitogenic activity of aFGF on CCL39 cells. Optimal concentrations for maximal potentiation of 400 micrograms/ml and 20 micrograms/ml were obtained respectively for DDE and heparin. In the present report, we have uncovered the fact that several carboxymethyl benzylamide sulfonate dextrans differing in degree and positioning of the substituent groups can mimic heparin in regard to the protection, stabilization, and potentiating effects with aFGF or bFGF. Our data establishes that the dextran derivatives studied can act as potentiating agents for FGFs. Native dextran (DDA) had no effect. Dextran derivatives can also protect aFGF and bFGF from heat as well as from pH denaturation, and against trypsic and chymotrypsic degradation. The dextran derivative DDI (82% methylcarboxylic acid, 23% benzylamide, 13% sulfonate) was studied in greater detail and exhibited a greater protection for bFGF and a lesser protecting effect for aFGF than heparin. Derivatized dextrans which have very weak anticoagulant activity are of great interest as alternatives to heparin for use as stabilizers, potentiators, protectants, and slow-release matrices for FGFs in pharmaceutical formulations.  相似文献   

16.
Eleven structural analogues of human basic fibroblast growth factor (bFGF) have been prepared by site-directed mutagenesis of a synthetic bFGF gene to examine the effect of amino acid substitutions in the three putative heparin-binding domains on FGF's biological activity. After expression in Escherichia coli, the mutant proteins were purified to homogeneity by use of heparin-Sepharose chromatography and analyzed for their ability to stimulate DNA synthesis in human foreskin fibroblasts. Recombinant human bFGF 1-146 and [Ala69,Ser87]bFGF, an analogue where two of the four cysteines had been replaced by alanine and serine, were equipotent to standard bovine basic fibroblast growth factor. Substitution of aspartic acid-19 by arginine in the first heparin-binding domain yielded a molecule that stimulated a higher total mitogenic response in fibroblasts as compared to bFGF. In addition, replacement of either arginine-107 in the second domain or glutamine-123 in the third domain with glutamic acid resulted in compounds that were 2 and 4 times more potent than bFGF. In contrast, substitution of arginine-107 with isoleucine reduced the activity of the molecule by 100-fold. Combination of domain substitutions to generate the [Glu107,123]bFGF and [Arg19,Lys123,126]bFGF mutants did not show any additivity of the mutations on biological activity. Alterations in the biological activity of the analogues was dependent on both the site of and the type of modification. Increased positive charge in the first domain and increased negative charge in the second and third domains enhanced biological potency. The altered activities of the derivatives appear to be due in part to changes in the affinity of the analogues for heparin. We conclude that changes in all three of the putative heparin-binding domains result in altered mitogenic activity and heparin interaction of basic fibroblast growth factor.  相似文献   

17.
Heparin was previously reported to potentiate the mitogenic activity of endothelial cell mitogens in a crude extract of bovine hypothalami (Thornton, S. C., Mueller, S. N., and Levine, E. M. (1983) Science 222, 623-625). We and others (Gospodarowicz, D., and Cheng, J. (1986) J. Cell. Physiol. 128, 475-484) have reported that the growth stimulatory effects of acidic fibroblast growth factor (aFGF) are potentiated in a similar manner. We have used these observations as the basis of an assay to characterize the importance of size, sulfation, and anticoagulant activity of heparin in mediating this effect. Partial nitrous acid depolymerization of heparin from porcine intestinal mucosa resulted in a mixture of heparin fragments, containing oligosaccharides ranging from disaccharides to polysaccharides of about 40 monosaccharides in length. This mixture was fractionated by ion exchange chromatography and gel permeation chromatography to obtain size-homogeneous oligosaccharides with different degrees of sulfation. Assay of these heparin-derived saccharides in the presence of a suboptimal concentration of aFGF revealed that a minimum chain length and a certain degree of sulfation is required in order to potentiate the action of aFGF. Low sulfate oligosaccharides (4-16 units) were unable to potentiate aFGF, whereas medium sulfate fractions of octadecasaccharides and larger were able to moderately potentiate aFGF. The potentiation of aFGF by the high sulfate fraction correlated with the saccharide size: 12 or more monosaccharide units were necessary to achieve potentiation equivalent to whole heparin, octa- and decasaccharides were mildly stimulatory, and hexasaccharides were without effect. In the absence of aFGF, intact heparin as well as all the oligosaccharides examined, inhibited the proliferation of capillary endothelial cells to approximately the same degree, between 20 and 50% inhibition. When a tetradecasaccharide was separated into a binding and a nonbinding fraction on matrix-bound antithrombin III, no difference was seen for these fractions in the endothelial cell proliferation assay. These results indicate that both size and sulfation of a heparin-derived oligosaccharide contribute to its ability to interact with aFGF and/or endothelial cells and that this interaction is independent of anticoagulant activity. In addition, our findings suggest that the inhibitory and potentiating effects of heparin on capillary endothelial cells have different structural requirements.  相似文献   

18.
Primary cultures of the medroxyprogesterone acetate-induced mouse mammary tumor line C4-HD are stimulated by medroxyprogesterone acetate (MPA) or progesterone. Serum obtained from ovariectomized, MPA-treated animals (OVX-MPA) exerts a stimulatory effect that is significantly higher than that induced by serum obtained from OVX mice with the exogenous addition of MPA, suggesting the involvement of MPA-induced serum factors potentiating the proliferative effect of MPA. The object of this paper is to further explore the stimulatory effect of mouse serum and to investigate the role of aFGF and bFGF on cell proliferation. The role of PR as possible mediators was tested using two different antiprogestins and antisense oligodeoxynucleotides of PR A isoform. Serum was obtained from OVX untreated or MPA-treated mice and was charcoalized and/or heat-inactivated. The effect of MPA or mifepristone at 10 nM concentrations was tested. Charcoalization and heat inactivation exerted a stimulatory effect (P<0.01) when OVX-serum was used. This effect was potentiated by MPA. Charcoalized OVX-MPA serum induced a significant inhibition of cell proliferation that was restored by the exogenous addition of MPA or by heat inactivation. Mifepristone induced an inhibition of 3H-thymidine uptake when OVX-MPA serum was used. These results suggest that serum factors activated by different manipulations may replace the stimulatory effect of MPA. When charcoalized fetal calf serum (chFCS) was used, a higher proliferative activity was obtained using higher serum concentrations. Mifepristone and onapristone 10 nM also inhibited this effect. aFGF and bFGF 100 ng/ml were both able to stimulate 3H-thymidine uptake. MPA exerted an additive effect. Mifepristone 10 nM inhibited bFGF and MPA+bFGF induced cell proliferation. Antisense oligodeoxynucleotides of PR (ASPR) were used to further confirm the participation of PR in the proliferative pathway of these cells. They inhibited serum and bFGF-induced cell proliferation in a specific dose-dependent manner. Our results suggest that PR play a central role in proliferation and suggest the existence of a cross-talk between steroid and growth factor signaling pathways.  相似文献   

19.
Polyclonal antibodies were prepared against recombinant basic fibroblast growth factor (bFGF) that reacted only with bFGF but not acidic FGF. These antibodies were able to inhibit various biological activities of bFGF such as the ability of bFGF to stimulate DNA synthesis in 3T3 cells, proliferation and migration of bovine capillary endothelial cells (BCEC), and neurite extension in pheochromocytoma (PC12) cells. The anti-bFGF antibodies also inhibited the mitogenic activity of subendothelial cell extracellular matrix for BCEC, demonstrating that the growth factor component in extracellular matrix required for supporting BCEC proliferation was bFGF. Anti-bFGF antibodies inhibited the cross-linking of bFGF to its high affinity receptor on BCEC cells. However, these antibodies did not inhibit the binding of bFGF to heparin-Sepharose or to the low affinity receptors of BCEC which have been demonstrated to be heparin-like molecules. These results suggest that bFGF has distinct domains for binding to high affinity cellular receptors and for binding to heparin.  相似文献   

20.
Confluent cultures of human endothelial cells deposit into extracellular matrix (ECM) distinct heparan sulfate proteoglycans (HSPG) which modulate acidic fibroblast growth factor's (aFGF) ability to stimulate human endothelial cell mitogenic capacity. Extracellular matrix 35S-HSPG were isolated from cultures metabolically labelled with Na235SO4 by DEAE-Sepharose, Sepharose CL-4B, and aFGF-Affi-Gel 15 column chromatography and identified by resistance to chondroitinase ABC and sensitivity to nitrous acid. Fifty to sixty percent of the 35S-HSPG deposited into ECM do not bind aFGF. The bound 35S-HSGP (40-50% of the total counts applied) eluted from the aFGF-Affi-Gel column after the addition of buffer containing 2 M NaCl. aFGF-binding and aFGF-nonbinding 35S-HSPG were individually pooled and further purified by Sepharose CL-4B column chromatography. 35S-HSPG which bind aFGF, designated HSPGP, were 100-fold superior to heparin in augmenting the mitogenic efficacy of aFGF in sparse proliferating cultures. In contrast, however, 35S-HSPG, which did not bind aFGF, designated HSPG1, inhibited aFGF-stimulated proliferation in both sparse and subconfluent endothelial cell cultures. The majority of the biological activity of both aFGF-potentiating HSPGP and aFGF-inhibitory HSPG1 was contained in the glycosaminoglycan chains released by alkaline borohydride treatment of intact HSPGP or HSPG1, respectively. 3H-Core protein derived from HSPGP or HSPG1 contained only minor biological activity. The ability of heparitinase or heparinase (Flavobacterium heparinum) to abolish biological activity differed, depending upon the HSPG tested, also suggested that these are two distinct HSPGs.  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