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.
Fibroblast growth factors (FGFs) have been implicated in a number of proliferative lesions, including malignant tumor growth and vascularization. As a result, cytotoxic agents that target cell surface FGF receptors are currently under investigation. Previous reports have shown that conjugation of basic FGF with the ribosome inactivator, saporin, results in a potent cytotoxin specific for cells bearing high-affinity FGF receptors. In this report, we have used this FGF receptor-dependent cytotoxin to study receptor interactions at the surface of embryonal carcinoma cells, which express low numbers of high-affinity FGF receptors. The growth of three embryonal carcinoma cell lines and one embryonic stem cell line was shown to be inhibited by bFGF-saporin, suggesting that these cells are able to bind and internalize FGF through high-affinity FGF receptors. In addition, we determined that the responses of these cells to bFGF-saporin are qualitatively different than the responses of CHO-KI cells, which also exhibit low numbers of high-affinity FGF receptors. Specifically, pretreatment with bFGF-saporin reduces the cloning efficiency of CHO-KI cells 8- to 10-fold, whereas bFGF-saporin has little or no effect on the cloning efficiency of embryonal carcinoma cells. This finding suggests that bFGF-saporin is cytotoxic for CHO-KI cells, but not for embryonal carcinoma cells. Thus, our findings argue strongly that other factors, in addition to high-affinity FGF receptor number, are important in determining sensitivity of cells of bFGF-saporin.  相似文献   

2.
Vascular endothelial cells derived from adult bovine aortic arch can be grown in two ways, either in the presence or absence of fibroblast growth factor. The types of collagen produced by cultures under these two conditions have been compared. In the presence of fibroblast growth factor, cells grow in an orderly fashion, express their normal phenotype and synthesize primarily type III collagen plus collagens types IV and V at a ratio of 10:1:3. Cultures grown in the absence of the factor lose their orderly pattern of growth, lose polarity and normal phenotypic expression. They devote twice the proportion of total protein-synthesizing capacity to collagen, and now synthesize type I in addition to the other collagen types. The ratio of collagen types I:III:IV:V is approximately 30:70:1:13. The kinds of type V collagen chains expressed are also altered. Fibroblast growth factor appears to modulate collagen synthesis, the major component of the extracellular matrix, and indirectly modulates the phenotypic expression of cultured vascular endothelial cells. In atherosclerosis, type I collagen is found in association with the intimal layer. The disorderly growth and the abnormal production of type I collagen by these vascular endothelial cells cultured in the absence of fibroblast growth factor is a model for a number of pathological situations including atherosclerotic plaque formation.  相似文献   

3.
4.
The presence of an efficient high affinity uptake system for L-CSA has been demonstrated in cultured neuronal and glial cells of various types. In neurons and most glial cells L-CSA uptake is inhibited by acidic amino acids,L-glutamate andL-aspartate and requires sodium ions; however the sodium dependence varies from one cell type to the other. The characteristics of the uptake system change during cell maturation, especially in astroblasts. The predominance of CSA uptake in glial cells as compared to neurons, the similarity of the kinetic parameters and of the structural specficity ofL-glutamate uptake suggest that both excitatory amino acids are transported by a common system. In view of accumulating evidence, the present results are in agreement with a role of CSA as a neurotransmitter and as a precursor for taurine biosynthesis in the central nervous system.  相似文献   

5.
Cultured peritubular cells prepared from the testes of 20-day-old rats contained both preproenkephalin (A) mRNA (1.5 kb) and amyloid beta-protein precursor mRNA (3.6 and 2.8 kb). The phorbol ester TPA and forskolin (an adenylate cyclase activator) increased the preproenkephalin mRNA abundance to 9.0 and 5.8 times the control, respectively. TPA alone had no effect on the intracellular cAMP level. A combination of TPA and forskolin elicited a synergistic increase in the ppEnk mRNA abundance over 30-fold. Dexamethasone potentiated the effect of forskolin but not of TPA. These results suggest that TPA regulates the preproenkephalin mRNA abundance through a cAMP-independent pathway. In contrast, TPA, forskolin, and dexamethasone showed little or no effect on the abundance of amyloid beta-protein precursor mRNA.  相似文献   

6.
Tubulin, widely recognized as a GTP/GDP-binding protein, has been isolated in its polymerized state from rat PC12 cells and embryonic chick dorsal root ganglion neurons by Triton X-100 detergent extraction of the cytoskeletal fraction. Perchloric acid extraction and deproteinization of this fraction permitted subsequent analysis of nucleotide identity and content by high performance liquid chromatography. PC12 cells grown in the absence of nerve growth factor (NGF) contained ADP, ATP, GDP, and GTP at levels consistent with the actin and tubulin content of the cytoskeletal fraction. Microtubules from PC12 cells cultured in the presence of NGF contain an additional nucleotide that we have identified as dGTP. Analysis of whole cell nucleotide extracts from PC12 cells grown in the absence or presence of NGF revealed no evidence for the presence of dGTP at 4 and 14 days, respectively. We have determined that embryonic chick dorsal root ganglion neurons also contain this deoxyribonucleotide, and we found virtually no ADP or ATP in the extracted dorsal root ganglion cytoskeletal fraction. On the basis of metabolic labeling studies with [14C] guanine, we have inferred that the presence of dGTP in NGF-treated PC12 cells probably arises either from binding to the nonexchangeable nucleotide site of tubulin undergoing dynamic assembly/disassembly or from binding to the exchangeable site of tubulin subsequently incorporated into highly stabilized microtubules.  相似文献   

7.
8.
Primary cultures of neuronal and glial cells from 1-day-old neonatal rats contain high affinity receptors for insulin-like growth factor I (IGF-I). The IC50 for displacement of 125I-IGF-I binding by unlabeled IGF-I was 3 nM for neuronal cells and 4 nM for glial cells. Unlabeled insulin was 20-50 times less potent. Apparent molecular mass of the alpha subunits of the IGF-I receptor was 125 kDa in neuronal and 135 kDa in glial cells. IGF-I induced autophosphorylation of the IGF-I receptor beta subunit in lectin-purified membrane preparations in a dose-dependent manner. The major phosphoamino acid of the beta subunit in both cell types was tyrosine in the IGF-I-stimulated state and serine in the basal state. Apparent molecular mass of the beta subunits of the IGF-I receptors was 91 kDa for neuronal and 95 kDa for glial cells. Tyrosine kinase activity of the IGF-I receptors was demonstrated by IGF-I-induced phosphorylation of the exogenous substrate poly(Glu, Tyr) 4:1 in both cell types. IGF-I had no effect on 2-deoxyglucose uptake in neuronal cells. In contrast, in glial cells, IGF-I stimulated 2-deoxyglucose uptake at very high doses, presumably acting via the insulin receptor. The effect of IGF-I as a neurotrophic growth factor in both neuronal and glial cells was demonstrated by its stimulation of [3H]thymidine incorporation. These findings suggest the IGF-I is an important growth factor in nervous tissue-derived cells.  相似文献   

9.
We have investigated the effects of bFGF on both the FSH-induced LH receptor expression and cAMP production in cultured rat granulosa cells. Concentrations of pure FGF, from 10(-12) M to 10(-10) M, progressively inhibit the stimulatory actions of FSH with an ED50 of approximately 4 x 10(-12) M for both parameters. Higher FGF concentrations, from 4 x 10(-10) M to 10(-8) M, lead to a gradual reduction of the growth factor inhibitory effect. The effects of FGF are more prominent on the modulation of LH receptors than on the FSH-induced cAMP production. Moreover, FGF impairs the LH receptor formation induced by cholera toxin or 8-Bromo-cAMP, indicating that the growth factor also acts at a step distal to cAMP formation. The inhibitory effect of FGF on LH receptor expression increases during the entire course of granulosa cell differentiation, from 24 to 96 h, and is not due to variations in cell number or viability, but rather to a change in the content of LH receptors with no significant modification of binding affinity (KD congruent to 0.8 x 10(-10) M). These results suggest that bFGF may acutely regulate the capacity of granulosa cells to differentiate upon FSH stimulation and to respond to LH during the ovarian follicular maturation.  相似文献   

10.
11.
12.
Endothelial cells line the blood vessel and precursor endothelial cells appear to have a pivotal effect on the organ formation of the heart, the embryonic development of the kidney, and the liver. Several growth factors including the fibroblast growth factors (FGF) seem to be involved in these processes. Ligands such as basic FGF produced and secreted by endothelial cells may also coordinate cellular migration, differentiation, and proliferation under pathological conditions including wound healing, tumorgenesis, and fibrogenesis in the adult. Recently we demonstrated the expression of two secreted FGFs, FGF16, and FGF18, in HUVEC and in rat aortic tissue. In the present report, we confirmed by RT-PCR analysis that FGF18 is wildly expressed in the cardiovascular tissue, while FGF16 showed a more restricted expression pattern. HUVEC clearly demonstrated chemotaxis towards FGF16 and FGF18. Both FGFs also enhanced cell migration in response to mechanical damage. However, recombinant FGF16 and FGF18 failed to induce endothelial cell proliferation or sprouting in a three-dimensional in vitro angiogenesis assay. Fgf18 expression was earlier reported in the liver, and we detected FGF18 expression in liver vascular and liver sinusoidal endothelial cells (LSECs), but not in hepatic parenchymal cells. Recombinant FGF18 stimulated DNA synthesis in primary hepatocytes, suggesting, that endothelial FGF18 might have a paracrine function in promoting growth of the parenchymal tissue. Interestingly, FGF2, which is mitogenic on endothelial cells and hepatocytes stimulates a sustained MAPK activation in both cell types, while FGF18 causes a short transient activation of the MAPK pathway in endothelial cells but a sustained activation in hepatocytes. Therefore, the difference in the time course of MAPK activation by the different FGFs appears to be the cause for the different cellular responses.  相似文献   

13.
Hepatocyte growth factor (HGF) exerts a cytostatic effect on HepG2 and B16-F1 cell lines. To evaluate the possible involvement of the apoptotic process in this effect, we performed studies at cellular and molecular levels. HGF induced apoptosis only in HepG2 hepatoma cells at day 3 in about 20% of the cells undergoing growth inhibition, while hallmarks of apoptosis did not occur in B16-F1 melanoma cells. During the first 24 h after HGF treatment, enhanced expression of the pro-apoptotic genes bax and c-Myc was observed at level of mRNA and protein. Concomitant induction of antizyme (AZ) might lower ornithine decarboxylase (ODC) protein level though a huge increase in ODC mRNA level took place. This was suggested as a signal for apoptosis decisional phase. The levels of the proteins examined except that of AZ fell down thereafter when HepG2 cells underwent apoptosis. In B16-F1 cells, only ODC and AZ protein levels were elevated probably in relation to the initial elevated growth rate and the absence of apoptosis involvement in the following cytostatic effect of HGF in melanoma cells. Consistent with this hypothesis, bax mRNA and protein levels were unchanged or even lower relative to control values.  相似文献   

14.
Summary Recent studies have determined that fibroblast growth factor (FGF) potentiates the soft agar growth responses of NRK-49F cells to several combinations of transforming growth factors (TGFs). In the current study, two other non-transformed cell lines, NR-6 and AKR-2B, which do not spontaneously form colonies in soft agar, were examined for their soft agar growth responses to FGF. Both the acidic form and basic form of FGF were found to induce the soft agar growth of these cells. In the case of NR-6 cells, the effects of TGF-β were also examined. TFG-β potentiated the soft agar growth response of NR-6 cells to FGF, but on its own did not induce soft agar growth. Attempts to identify other factors capable of modulating the response of these cells to FGF, led to the finding that both 12-0-tetra-decanoylphorbol-13-acetate and retinoic acid suppress FGF-induced soft agar growth of NR-6 cells and AKRR-2B cells. The finding that FGF induces the soft agar growth of both non-transformed cell lines, together with the findings of others that both forms of FGF are angiogenic, lends further support to the suggestion that FGF plays a significant role in the in vivo growth of some, and possibly many, tumors. This work was supported by grants from the Nebraska Department of Health (86-11R, 87-38), the National Institute of Child Health and Human Development (HD 19837, HD 21568) and the National Cancer Institute (Laboratory Cancer Research Center Support Grant CA 36727). Editor's Statement The last several years have seen extraordinary advances in the understanding of the biochemistry and physiology of heparin-binding growth factors. Among the activities of these peptides that may be of significance for neoplasia and wound healingin vivo is ability to promote anchorage independent growth of some cell types. In this study the interactions among several stimulatory and inhibitory factors are examined in a soft agar growth assay. An appreciation of these interactions is critical in attempts to relatein vitro effects to those in the intact organism.  相似文献   

15.
Fibroblast growth factors (FGFs) interact with heparan sulfate glycosaminoglycans and the extracellular domains of FGF cell surface receptors (FGFRs) to trigger receptor activation and biological responses. FGF homologous factors (FHF1-FHF4; also known as FGF11-FGF14) are related to FGFs by substantial sequence homology, yet their only documented interactions are with an intracellular kinase scaffold protein, islet brain-2 (IB2) and with voltage-gated sodium channels. In this report, we show that recombinant FHFs can bind heparin with high affinity like classical FGFs yet fail to activate any of the seven principal FGFRs. Instead, we demonstrate that FHFs bind IB2 directly, furthering the contention that FHFs and FGFs elicit their biological effects by binding to different protein partners. To understand the molecular basis for this differential target binding specificity, we elucidated the crystal structure of FHF1b to 1.7-A resolution. The FHF1b core domain assumes a beta-trefoil fold consisting of 12 antiparallel beta strands (beta 1 through beta 12). The FHF1b beta-trefoil core is remarkably similar to that of classical FGFs and exhibits an FGF-characteristic heparin-binding surface as attested to by the number of bound sulfate ions. Using molecular modeling and structure-based mutational analysis, we identified two surface residues, Arg52 in the beta 4-beta 5 loop and Val95 in the beta 9 strand of FHF1b that are required for the interaction of FHF1b with IB2. These two residues are unique to FHFs, and mutations of the corresponding residues of FGF1 to Arg and Val diminish the capacity of FGF1 to activate FGFRs, suggesting that these two FHF residues contribute to the inability of FHFs to activate FGFRs. Hence, FHFs and FGFs bear striking structural similarity but have diverged to direct related surfaces toward interaction with distinct protein targets.  相似文献   

16.
Neurons synthesise and secrete many growth and survival factors but it is not usually clear whether they are released locally at the cell body or further afield from axons or axon terminals. Without this information, we cannot predict the site(s) of action or the biological functions of many neuron-derived factors. For example, can neuronal platelet-derived growth factor (PDGF) be secreted from axons and reach glial cells in nerve-fibre (white-matter) tracts? To address this question, we expressed PDGF-A in retinal ganglion neurons in transgenic mice and tested for release of PDGF from cell bodies in the retina and from axons in the optic nerve. In both the retina and optic nerve, there are glial cells that express PDGF receptor alpha (PDGFR alpha) [1] and divide in response to PDGF [2-5], so we could detect functional PDGF indirectly through the mitogenic response of glia at both locations. Expressing PDGF-A in neurons under the control of the neuron-specific enolase promoter (NSE-PDGF-A) resulted in a striking hyperplasia of retinal astrocytes, demonstrating that PDGF is secreted from the cell bodies of neurons in the retina [4]. In contrast, glial proliferation in the optic nerve was unaffected, indicating that PDGF is not released from axons. When PDGF was expressed directly in the optic nerve under the control of an astrocyte-specific promoter (GFAP-PDGF-A), oligodendrocyte progenitors hyperproliferated, resulting in a hypertrophic optic nerve. We conclude that PDGF is constitutively secreted from neuronal cell bodies in vivo, but not from axons in white-matter tracts.  相似文献   

17.
The glial cell line-derived neurotrophic factor (GDNF) is involved in the development and maintenance of neural tissues. Mutations in components of its signaling pathway lead to severe migration deficits of neuronal crest stem cells, tumor formation, or ablation of the urinary system. In animal models of Parkinson's disease, GDNF has been recognized to be neuroprotective and to improve motor function when delivered into the cerebral ventricles or into the substantia nigra. Here, we characterize the network of 43 genes induced by GDNF overproduction of neuronal progenitor cells (ST14A), which mainly regulate migration and differentiation of neuronal progenitor cells. GDNF down-regulates doublecortin, Paf-ah1b (Lis1), dynamin, and alpha-tubulin, which are involved in neocortical lamination and cytoskeletal reorganization. Axonal guidance depends on cell-surface molecules and extracellular matrix proteins. Laminin, Mpl3, Alcam, Bin1, Id1, Id2, Id3, neuregulin1, the ephrinB2-receptor, neuritin, focal adhesion kinase (FAK), Tc10, Pdpk1, clusterin, GTP-cyclooxygenase1, and follistatin are genes up-regulated by GDNF overexpression. Moreover, we found four key enzymes of the cholesterol-synthesis pathway to be down-regulated leading to decreased farnesyl-pyrophospate production. Many proteins are anchored by farnesyl-derivates at the cell membrane. The identification of these GDNF-regulated genes may open new opportunities for directly influencing differentiation and developmental processes of neurons.  相似文献   

18.
The ability of cadmium-bound metallothionein(Cd-MT) to induce apoptosis was investigated in vivo and in vitro. Administration of purified Cd-MT (0.15 mg MT bound Cd per kg body weight) to the rat induces DNA fragmentation, a biochemical characteristic of apoptosis in the kidney at 16 h, which was detectable by ethidium bromide staining on an agarose gel. It was still detected 24 h after administration. Induction of apoptosis by Cd-MT was specific to kidney; it was not observed in cerebrum, cerebellum, heart, lung, liver, testis, dorsolateral prostate, and ventral prostate. In contrast, addition of Cd-MT (0.01-100 microM) to the cultured porcine kidney LLC-PK1 cells failed to induce apoptosis under the condition where cadmium chloride (10 microM) did. There was no additivity of induction of apoptosis by CdCl2 (10 microM) in the presence of Cd-MT (0.01-100 microM). To examine the effect of intracellular MT on cadmium-induced apoptosis in cultured cells, new cell lines were established, which constitutively produce MT, being termed as Cd(r)-LLC-PK1 cells since Cd-MT exogenously added had much less permeability to the cultured cells. Followed by exposure of wild-type LLC-PK1 cells to 50 microM CdCl2 for 24 h, the surviving cells(Cd(r)-LLC-PK1 cells) induce MT at the level of 1.9 microg/2 x 10(6) cells. In Cd(r)-LLC-PK1 cells, 10 microM CdCl2 failed to induce apoptosis, but 60 microM CdCl2 could exert the apoptotic response, indicating that intracellular MT which was induced by CdCl2 did not facilitate CdCl2-elicited apoptosis. Furthermore, chromatin in rat kidneys was condensed by Cd-MT, but not that in LLC-PK1 cells. Thus, Cd-MT induces apoptosis in rat kidneys, but not in the cultured renal cells, suggesting that the ionic form of cadmium was required for programmed cell death.  相似文献   

19.
R Siman  J P Card  R B Nelson  L G Davis 《Neuron》1989,3(3):275-285
Although the beta-amyloid peptide is an established core component of neuritic plaques that accumulate in Alzheimer's disease, the mechanisms responsible for its deposition are not well understood. We now report that lesions of rat hippocampal neurons cause a time-dependent, long-lasting elevation of immunoreactivity for the beta-amyloid precursor protein (APP) in neighboring astrocytes, a cell type not normally containing the protein. The increase represents astroglial expression of the protein rather than a scavenging of APP released by damaged neurons. Immunoelectron microscopy confirmed that APP-containing cells are reactive astroglia, both surrounding capillaries and within the neuropil. These results demonstrate that neuronal damage stimulates APP expression in adult brain and suggest that reactive astrocytes may be a source of the beta-amyloid that forms neuropathological plaques in Alzheimer's disease.  相似文献   

20.
The mechanism of epidermal growth factor receptor (EGF-R) autophosphorylation in intact A431 cells was studied. We detected epidermal growth factor (EGF) induced tyrosine phosphorylation of EGF-R not occupied with ligand. Cell monolayers were subjected to irradiation after incubation with photoreactive derivative of EGF and uncoupled EGF was extracted by acidic treatment. Subsequent immunoprecipitation with antiphosphotyrosine antibodies resulted in precipitation of both EGF-R complexes with EGF and EGF-R with unoccupied ligand-binding site. The fact of precipitation of EGF-R with unoccupied ligand-binding site in conjunction with our finding of rapid dephosphorylation of EGF-R after EGF extraction by acidic treatment, strongly supports the interpretation that cross-phosphorylation of EGF-R may take place in intact cells.  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