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 31 毫秒
1.
Undifferentiated HT29 and differentiated HT29G-human colon carcinoma cells have been used to study the changes in proteoglycan production and structure associated with enterocytic cell differentiation. Differentiated cells incorporate twice as much sulfate than undifferentiated cells when labeled with [35S]sulfate. Both cell lines produce a heparan sulfate proteoglycan which was purified by ion-exchange. The heparan sulfate proteoglycan from differentiated HT29G- cells is larger and more homogeneous in size than that produced by undifferentiated HT29 cells. No differences in the core protein structure were observed. The detailed structural analysis of the heparan sulfate chains revealed that the structure of these chains follows the standard rules for these glycosaminoglycans with N-sulfated domains and N-acetylated domains. The main finding was that differentiated HT29G- cells have a degree of higher sulfation than HT29 cells. These differences were found to affect primarily 6-O-sulfated positions.  相似文献   

2.
Fibroblast growth factors (FGFs) play multiple roles during development and in adult tissues as paracrine regulators of growth and differentiation. FGFs signal through transmembrane receptor tyrosine kinases, but heparan sulfate is also required for signaling by members of the FGF family. In addition, heparan sulfate may be involved in determining tissue distribution of FGFs. Using biotinylated FGF-2 and FGF-7 (KGF) as probes, we have identified specific interactions between FGFs and heparan sulfates in human tissues. Both FGF species bind to tissue mast cells and to epithelial cell membranes. Binding to basement membrane heparan sulfate is tissue source dependent and specific. Although FGF-2 strongly binds to basement membrane heparan sulfate in skin and most other tissue sites examined, FGF-7 fails to bind to basement membrane heparan sulfate in most locations. However, in subendothelial matrix in blood vessels and in the basement membrane of a papillary renal cell carcinoma, strong FGF-7 binding is seen. In summary, distinct and specific affinities of heparan sulfates for different FGFs were identified that may affect growth factor activation and local distribution. Heparan sulfate may have a gatekeeper function to either restrict or permit diffusion of heparin-binding growth factors across the basement membrane.  相似文献   

3.
Stem cell localization, conservation, and differentiation is believed to occur in niches in the marrow stromal microenvironment. Our recent observation that long-term in vitro human hematopoiesis requires a stromal heparan sulfate proteoglycan (HSPG) led us to hypothesize that such HSPG may orchestrate the formation of the stem cell niche. We compared the structure and function of HS from M2-10B4, a hematopoiesis-supportive cell line, with HS from a nonsupportive cell line, FHS-173-We. Long-term culture-initiating cell (LTC-IC) maintenance was enhanced by PG from supportive cells but not by PG from nonsupportive cells (P <.005). The supportive HS were significantly larger and more highly sulfated than the nonsupportive HS. Specifically, supportive HS contained higher 6-O-sulfation on the glucosamine residues. In agreement with these observations, purified 6-O-sulfated heparin and highly 6-O-sulfated bovine kidney HS similarly maintained LTC-IC. In contrast, completely desulfated heparin, N-sulfated heparin, and unmodified heparin did not support LTC-IC maintenance. Moreover, the supportive HS promoted LTC-IC maintenance but not differentiation of CD34(+)/HLA-DR- cells into colony-forming cells (CFCs) and mature blood cells. The supportive HS but not the nonsupportive HS bound both cytokines and matrix components critical for hematopoiesis, including interleukin-3 (IL-3), macrophage inflammatory protein-1 (MIP-1), and thrombospondin (TSP). Significantly more CD34(+) cells adhered directly to immobilized O-sulfated heparin than to N-sulfated or desulfated heparin. Thus, hematopoiesis-supportive stromal HSPG possessing large, highly 6-O-sulfated HS mediate the juxtaposition of hematopoietic progenitors with stromal cells, specific growth-promoting (IL-3) and growth-inhibitory (MIP-1 and platelet factor 4 [PF4]) cytokines, and extracellular matrix (ECM) proteins such as TSP. We conclude that the structural specificity of stromal HSPG that determines the selective colocalization of cytokines and ECM components leads to the formation of discrete niches, thereby orchestrating the controlled growth and differentiation of stem cells. These findings may have important implications for ex vivo expansion of and gene transfer into primitive hematopoietic progenitors.  相似文献   

4.
Heparan sulfate is a molecule that possesses a large structural variability and which has been shown to inhibit the proliferation of fibroblasts in vitro. The aim of this study was to determine whether the anti-proliferative effects of heparan sulfate were exerted by regulation of the activity of the platelet-derived growth factor and/or of the platelet-derived growth factor receptors. Both l-iduronate-rich, anti-proliferative and the l-iduronate-poor, non-anti-proliferative heparan sulfate species, were incubated with confluent human embryonic lung fibroblasts for 24 h. The mRNA levels for PDGF-AA, PDGF-BB, and their receptors were measured. Binding studies were performed with [125I]-PDGF-BB and [125I]-EGF for 2 h at 4 degreesC in cultures preincubated with both types of heparan sulfate for 24 h. In separate experiments, cultures were incubated together with heparan sulfate and [125I]-PDGF-BB for 2 h at 4 degreesC. Increases of two- to threefold in the mRNA levels for both the alpha- and the beta-receptors of PDGF was obtained after treatment with both types of heparan sulfate, whereas the mRNA levels of both the PDGF-AA and the PDGF-BB were essentially unaffected. A sixfold increase in binding was only noted for [125I]-PDGF-BB in cultures pre-treated with the anti-proliferative heparan sulfate for 24 h, whereas no effect was noted with use of the non-anti-proliferative heparan sulfate. Incubating the [125I]-PDGF-BB and the anti-proliferative heparan sulfate together for 2 h resulted in a smaller, threefold increase in binding. This indicates that the anti-proliferative heparan sulfate both stabilizes and increases expression of the PDGF receptors. To investigate whether the increased number of PDGF receptors could affect cell activity, cells were preincubated with anti-proliferative heparan sulfate and then treated with PDGF-BB. This resulted in an increase in mitogenicity compared to cells treated only with PDGF-BB. Neither an increase in binding for [125I-EGF] nor an increase in the mitogenic response of EGF could be observed in cultures pre-treated with the anti-proliferative heparan sulfate. The results indicate that the extracellular matrix itself may regulate important biological phenomena such as cell proliferation and matrix production through affecting the expression of receptors of PDGF, which initiate both stimulatory and inhibitory signals.  相似文献   

5.
Heparan sulfate is thought to regulate the biological activities of several proteins implicated in the pathogenesis of atherosclerosis. While the interactions of heparan sulfate with lipoprotein lipase and various growth factors have been actively studied, little is known of the cellular regulation of heparan sulfate biosynthesis in response to lipid accumulation. We have investigated heparan sulfate biosynthesis during conversion of murine J774 macrophages into lipid-laden foam cells. Such conversion is shown to accelerate the rate of glycosaminoglycan synthesis and the transport of newly synthesized proteoglycans into the medium. Moreover, the structure of heparan sulfate is specifically altered due to an approximately 30% increase in the 6-O-sulfation of glucosamine residues within the N-sulfated heparan sulfate domains, whereas the sulfation of chondroitin sulfate remains unaffected. These results suggest a selective effect of foam cell conversion on the biosynthesis of heparan sulfate.  相似文献   

6.
Heparan sulfate interacts with growth factors, matrix components, effectors and modulators of enzymatic catalysis as well as with microbial proteins via sulfated oligosaccharide domains. Although a number of such domains have been characterized, little is known about the regulation of their formation in vivo. Here we show that the structure of human aorta heparan sulfate is gradually modulated during aging in a manner that gives rise to markedly enhanced binding to isoforms of platelet-derived growth factor A and B chains containing polybasic cell retention sequences. By contrast, the binding to fibroblast growth factor 2 is affected to a much lesser extent. The enhanced binding of aorta heparan sulfate to platelet-derived growth factor is suggested to be due to an age-dependent increase of GlcN 6-O-sulfation, resulting in increased abundance of the trisulfated L-iduronic acid (2-OSO3)-GlcNSO3(6-OSO3) disaccharide unit. Such units have been shown to hallmark the platelet-derived growth factor A chain-binding site in heparan sulfate.  相似文献   

7.
The thiazolidinediones troglitazone and BRL 49653 improve insulin sensitivity in humans and animals with insulin resistance. Adipose tissue lipoprotein lipase is an insulin-sensitive enzyme. We examined the effects of thiazolidinediones on lipoprotein lipase expression in adipocytes. When added to 3T3-F442A, 3T3-L1, and rat adipocytes in culture, troglitazone and BRL 49653 inhibited lipoprotein lipase activity. This inhibition was observed at concentrations as low as 0.1 microM and within 2 h after addition of the drug. Lipoprotein lipase activity was inhibited in differentiated adipocytes as well as the differentiating cells. Despite this decrease in enzyme activity, these drugs increased mRNA levels in undifferentiated 3T3-F442A and 3T3-L1 cells and had no effect on mRNA expression or synthesis of lipoprotein lipase in differentiated cells. Western blot analysis showed that these drugs did not affect the mass of the enzyme protein. Lipoprotein lipase activity in cultured Chinese hamster ovary cells was not inhibited by troglitazone. Glucose transport, biosynthesis of lipids from glucose or the biosynthesis of proteins were unaffected by thiazolidinediones in differentiated cells, whereas glucose transport and lipid biosynthesis were increased when these drugs were added during differentiation. These results show that troglitazone and BRL 49653 have a specific, post-translational inhibitory effect on lipoprotein lipase in adipocytes, yet they promote lipid accumulation and differentiation in preadipocytes.  相似文献   

8.
9.
Fibroblast growth factors are important heparin binding, mitogenic proteins. The binding site in heparin and heparan sulfate for fibroblast growth factor-2 (basic fibroblast growth factor) has been described as rich in glucosamine-2-sulfate 1-->4 linked to iduronic acid-2-sulfate. The glucosamine residue in the heparin binding site is also 6-sulfated. A new glycosaminoglycan, acharan sulfate, has been chemically modified to prepare a polysaccharide, N-sulfoacharan sulfate, consisting of glucosamine-2-sulfate 1-->4 linked to iduronic acid-2-sulfate. Acharan sulfate binds very weakly to fibroblast growth factor-2 while N-sulfoacharan sulfate binds with nearly the same affinity as heparin. Mitogenicity studies were performed using heparan sulfate-free cells stably transfected with fibroblast growth factor receptor-1. Acharan sulfate inhibits heparin's enhancement of fibroblast growth factor-2 mitogenic activity, without affecting cell viability, while N-sulfoacharan sulfate shows heparin-like activity but at a greatly reduced level. These results suggest additional mechanisms not requiring high affinity glycosaminoglycan binding to fibroblast growth factor-2 may be important in its mitogenic activity.  相似文献   

10.
Platelet-derived growth factors (PDGFs) are homo- or heterodimers of two related polypeptides, known as A and B chains. The A chain exists as two splice variants due to the alternative usage of exons 6 (PDGF-AL, longer) and 7 (PDGF-AS, shorter). Exon 6 encodes an 18-amino acid sequence rich in basic amino acid residues, which has been implicated as a cell retention signal. Several lines of evidence indicate that the retention is due to binding of PDGF-AL to glycosaminoglycans, especially to heparan sulfate. We have analyzed the saccharide domains of smooth muscle cell-derived heparan sulfate involved in this interaction. Furthermore, we have employed selectively modified heparin oligosaccharides to elucidate the dependence of the binding on different sulfate groups and on fragment length. The shortest PDGF-AL binding domain consists of 6-8 monosaccharide units. Studies using selectively desulfated heparins and heparin fragments suggest that N-, 2-O-, and 6-O-sulfate groups all contribute to the interaction. Structural comparison of heparan sulfate oligosaccharides separated by affinity chromatography on immobilized PDGF-AL showed that the bound pool was enriched in -IdceA(2-OSO3)-GlcNSO3(6-OSO3)- disaccharide units. Furthermore, analogous separation of a partially O-desulfated heparin decamer preparation, using a highly selective nitrocellulose filter-trapping system, yielded a PDGF-AL-bound fraction in which more than half of the disaccharide units had the structure -IdceA(2-OSO3)-GlcNSO3(6-OSO3)-. Our results suggest that the interaction between PDGF-AL and heparin/heparan sulfate is mediated via N-sulfated saccharide domains containing both 2-O- and 6-O-sulfate groups.  相似文献   

11.
Interleukin-8, a member of the CXC chemokine family, has been shown to bind to glycosaminoglycans. It has been suggested that heparan sulfate on cell surfaces could provide specific ligand sites on endothelial cells to retain the highly diffusible inflammatory chemokine for presentation to leukocytes. By using selectively modified heparin and heparan sulfate fragments in a nitrocellulose filter trapping system, we have analyzed sequence requirements for interleukin-8 binding to heparin/heparan sulfate. We demonstrate that the affinity of a monomeric interleukin-8 molecule for heparin/heparan sulfate is too weak to allow binding at physiological ionic strength, whereas the dimeric form of the protein mediates binding to two sulfated domains of heparan sulfate. These domains, each an N-sulfated block of approximately 6 monosaccharide units, are contained within an approximately 22-24-mer sequence and are separated by a region of 相似文献   

12.
Interleukin-5 (IL-5) is the major cytokine regulating eosinophil production. In allergic disease tissue damage is primarily caused by eosinophils. Heparan sulfate proteoglycans are components of the bone marrow stroma, which supports hemopoietic cell differentiation and proliferation. We show that at low IL-5 concentrations heparan sulfate enhances the proliferation of an IL-5-dependent cell line. To investigate a mechanism for this effect we used an artificial proteoglycan to establish an enzyme-linked immunosorbent assay for the binding of heparin to proteins. Using this assay we demonstrate that IL-5 binds to heparin. The IL-5/heparin interaction is inhibited by ethylenediaminetetraacetate and enhanced by low concentrations of zinc ions. IL-5 interacts with iduronic acid containing glycosaminoglycans, and heparan sulfate preparations that have numerous N-sulfated domains per chain are especially efficient at inhibiting heparin binding. Both IL-5/heparin binding and the synergistic effect of IL-5 and heparan sulfate on cell proliferation were inhibited by an anti-IL-5 monoclonal antibody. These data suggest that the binding of IL-5 to heparan sulfate modulates IL-5 activity.  相似文献   

13.
The distribution and localization of collagen types were studied immunohistochemically in resected tissues obtained from gastric cancer patients. The expression of transforming growth factor (TGF) -alpha, TGF-beta 1 and TGF-beta 2 on cancer cells as well as the aggregation of T lymphocytes in the cancer tissue were also studied, in order to determine the differences between differentiated and undifferentiated type cancer. The interstitial tissues of differentiated type cancer showed intense staining for types I and III collagen, while those of undifferentiated type cancer showed intense staining for types I and III collagen, in addition to the stronger staining for types IV, V and VI collagen. Characteristically, type IV collagen was intensely stained in the interstitium in 18 of 20 undifferentiated type cancer (90%), but was stained in only one of 15 differentiated type cancer (6%). CD 3+ T lymphocytes were aggregated in the interstitial tissue of both the tumors, where the density of CD 4+ cells and the ratio of CD 4 to CD 8 were significantly higher in undifferentiated type cancer than in differentiated type cancer. TGF-alpha was detected in cancer cells in 80% of the differentiated cases and in 45% of the undifferentiated cases. The staining of TGF-beta 1 was also detected in 80% of the undifferentiated cases, which was significantly higher than 47% in differentiated cases. There were no differences in the incidences of staining for TGF-beta 2 between differentiated (33%) and undifferentiated type cancer (40%). These results suggest that there exist different mechanisms in the regulation of collagen production between differentiated and undifferentiated types of gastric cancer.  相似文献   

14.
To assess the autocrine function of insulin-like growth factor II (IGF-II) in the balance of proliferation and differentiation in HT29-D4 human colonic cancer cells, we studied the expression of IGF-I receptors (IGF-IR) and insulin receptors (IR) in relation to the state of cell differentiation. IGF-IR and IR were expressed in both undifferentiated and enterocyte-like differentiated HT29-D4 cells. IGF-IR had two isoforms with a 97-kDa and a 102-kDa beta-subunit. In addition, HT29-D4 cells expressed hybrid receptors (HR) formed by the association of two alphabeta heterodimers from both IR and IGF-IR. HR were evidenced through 1) inhibition of IGF-I binding by the B6 anti-IR antibody and 2) immunoprecipitation with the alpha-IR3 anti-IGF-IR antibody, which revealed an additional 95-kDa IR beta-subunit that disappeared when the heterotetrameric receptor was dissociated by disulfide reduction into alphabeta heterodimers before immunoprecipitation. Like IGF-IR, HR had a high affinity for IGF-I (Kd, approximately 1.5 nM), but did not bind insulin significantly; the latter interacted with the native IR only (Kd, approximately 4 nM). In the differentiated HT29-D4 cell monolayer, all receptor species were strongly polarized (>97%) toward the basolateral membrane. Moreover, HT29-D4 cell differentiation was accompanied by an approximately 2-fold increase in the number of IR, whereas the number of IGF-I-binding sites was unaltered. However, in differentiated HT29-D4 cells, approximately 55% of the latter were involved in HR vs. approximately 20% in undifferentiated HT29-D4 cells. Thus, HT29-D4 cell differentiation is characterized by an up-regulation (approximately 3-fold) of the level of HR coupled to a down-regulation (approximately 40%) of the level of native tetrameric IGF-IR. Alterations were induced early during the cell differentiation process, i.e. 5 days postconfluence, and remained unchanged for at least 21 days. Taken together, these results suggest that the IGF-II autocrine loop in HT29-D4 cells may trigger distinct signaling pathways if it activates native IGF-IR, which predominate in undifferentiated cells, or if it activates HR, which are up-regulated in differentiated cells.  相似文献   

15.
Employing clonal cell lines derived from rat embryonic hippocampal cells, we detected neuropeptide Y (NPY) mRNA in three progenitor subcloned cell lines. These cell lines upon differentiation express markers indicative of commitment to either neuronal (H19-7; NF +, GFAP -), glial (H19-5; GFAP +, NF -), or bipotential (H583-5; NF +, GFAP + ) lineages. Induction of differentiation was associated with the persistence of the NPY mRNA, however, in the differentiated H19-7 cells a 20-fold increase in NPY mRNA levels was observed (P<0.05). NPY immunoreactivity was observed only in cells with a differentiated neuronal phenotype. The cellular radioimmunoassayable NPY peptide levels increased twelve-fold without a change in extracellular NPY peptide levels by multi-factorially induced neuronal or glial cell differentiation. The differentiated H19-5 cells expressed lower levels of NPY that could not be immunocytochemically detected. The peripheral sympathetic PC-12 neuronal cells examined in the undifferentiated and nerve growth factor-driven differentiated states expressed NPY only upon differentiation. We conclude that NPY is expressed by the cultured undifferentiated and differentiated rat hippocampal clonal cell lines, while the peripheral sympathetic PC-12 neuronal cell line only expresses the NPY gene upon differentiation. These immortalized embryonic neural cell line(s) will provide a hippocampal cell line(s) to conduct future in-vitro investigations targeted at determining the cellular and molecular mechanisms governing NPY gene expression.  相似文献   

16.
We have analyzed the regulation of fibroblast growth factor receptors (FGFRs) during retinoic acid (RA) induced differentiation of Tera-2 human embryonal carcinoma cells. Undifferentiated Tera-2 cells expressed mRNAs for all four known FGFRs. Their differentiation led to loss of FGFR-4 mRNA expression and mRNA levels for FGFR-2 and FGFR-3 were considerably downregulated, whereas the mRNA levels for FGFR-1 remained unaltered. A substantial decrease in binding of K-FGF was found to occur upon RA-induced differentiation of the cells. In undifferentiated Tera-2 cells FGF stimulation caused an increase of c-fos mRNA, and c-jun mRNAs, but no increase of junB mRNA, whereas in the differentiated cells, FGFs strongly stimulated the expression of all three genes. Thus differentiation of the Tera-2 cells leads to marked changes in FGFR gene expression as well as to complex alterations in their responses to exogenous FGFs.  相似文献   

17.
Heparan sulphates, the N-sulphated polysaccharides components of proteoglycans, are common constituents of cell surfaces and the extracellular matrix. The heparan sulphate polysaccharide chain has a unique molecular design in the which the clusters of N- and O-sulphated sugar residues, separated by regions of low sulphation, determine specific protein binding properties. The heparan sulphate chains are attached to various protein cores, which determine the location of the proteoglycan in the cell membrane and extracellular matrix. The diverse functions of heparan sulphate, which range from the control of blood coagulation to the regulation of cell growth and adhesion, depend on the capacity of the chains to activate protein ligands, such as antithrombin III and members of the fibroblast growth factor family. These properties are currently being exploited in the development of synthetic heparan sulphates as anticoagulants and promoters of wound healing. Conversely organic mimics of growth factor activating saccharides could possibly be designed to suppress tumour growth and prevent restenosis after coronary vessel angioplasty.  相似文献   

18.
We have investigated the expression patterns and subcellular localization in nervous tissue of glypican, a major glycosylphosphatidylinositol-anchored heparan sulfate proteoglycan that is predominantly synthesized by neurons, and of biglycan, a small, leucine-rich chondroitin sulfate proteoglycan. By laser scanning confocal microscopy of rat central nervous tissue and C6 glioma cells, we found that a significant portion of the glypican and biglycan immunoreactivity colocalized with nuclear staining by propidium iodide and was also seen in isolated nuclei. In certain regions, staining was selective, insofar as glypican and biglycan immunoreactivity in the nucleus was seen predominantly in a subpopulation of large spinal cord neurons. The amino acid sequences of both proteoglycans contain potential nuclear localization signals, and these were demonstrated to be functional based on their ability to target beta-galactosidase fusion proteins to the nuclei of transfected 293 cells. Nuclear localization of glypican beta-galactosidase or Fc fusion proteins in transfected 293 cells and C6 glioma cells was greatly reduced or abolished after mutation of the basic amino acids or deletion of the sequence containing the nuclear localization signal, and no nuclear staining was seen in the case of heparan sulfate and chondroitin sulfate proteoglycans that do not possess a nuclear localization signal, such as syndecan-3 or decorin (which is closely related in structure to biglycan). Transfection of COS-1 cells with an epitope-tagged glypican cDNA demonstrated transport of the full-length proteoglycan to the nucleus, and there are also dynamic changes in the pattern of glypican immunoreactivity in the nucleus of C6 cells both during cell division and correlated with different phases of the cell cycle. Our data therefore suggest that in certain cells and central nervous system regions, glypican and biglycan may be involved in the regulation of cell division and survival by directly participating in nuclear processes.  相似文献   

19.
Epithelial cells are important components of the thymus microenvironment and are involved in thymocyte differentiation. The production and secretion of sulfated glycosaminoglycans by these cells grown in culture were investigated using labeling with radioactive 35S-Na2SO4 and 3H-glucosamine. The major glycosaminoglycans synthesized by these cells are heparan sulfate and hyaluronic acid. The structure of the heparan sulfate was investigated by the pattern of degradation products formed by deaminative cleavage with nitrous acid. The ratio 35S-sulfate/ H-glucosamine is high in the segments of the heparan sulfate released during the deaminative cleavage with nitrous acid but low in the resistant portion of the molecule. Thus, the heparan sulfate synthesized by the thymic epithelial cells contains a highly sulfated region. Digestion with heparitinase reveals that this highly sulfated region is a heparin-like segment of the molecule. The heparan sulfate is rapidly incorporated into the cell surface but its secretion to the extracellular medium requires a longer incubation period. Finally, heparin was used to mimic the possible effect of this heparan sulfate with a highly sulfated region, as ascertained by its ability to modulate thymocyte adhesion to thymic epithelial cells. Since heparin actually enhanced thymocyte adhesion, it is suggested that the heparan sulfate described herein, secreted by the thymic epithelium, may play a role upon intrathymic heterotypic cellular interactions.  相似文献   

20.
Our previous in vivo and in vitro studies showed that 5-azacytidine (5Az), a cytidine analog, induced apoptosis in developing neuronal cells in mice. To develop a system in which the precise molecular mechanism of 5Az-induced apoptosis in developing neuronal cells could be elucidated, we carried out the present study with PC12 cells. These cells are derived from a rat pheochromocytoma and extrude neurites in response to nerve growth factor (NGF). Light microscopy showed dose-dependent pyknotic and karyorrhectic changes in undifferentiated PC12 cells. Although they were less sensitive to 5Az, NGF-treated differentiated cells showed the same changes. Analysis by the TUNEL method (an in situ method for the detection of apoptosis) showed positive signals in the pyknotic and fragmented nuclei of these cells. Transmission electron microscopy revealed margination, segmentation, and condensation of nuclear chromatin, cell body shrinkage, and cytoplasmic vacuolization. Scanning electron microscopy demonstrated bleb formation on the cell surface. These pathomorphological changes are typical of apoptosis. 5Az seemed to affect cells that were in the proliferative stage; when the cells were terminally differentiated, their sensitivity to 5Az appeared to decline. PC12 cells could be used as a pathomorphological and biochemical model for studies of 5Az-induced neuronal cell apoptosis at the molecular and genetic level.  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