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 8 毫秒
1.
A simple procedure for the isolation of heparan sulfates from pig lung using a poly-L-lysine-Sepharose column is described. Glycosaminoglycans are absorbed on poly-L-lysine-Sepharose at pH 7.5 and eluted with an NaCl linear gradient in the following order: hyaluronic acid (0.32 M NaCl), chondroitin (0.36 M NaCl), keratan sulfate (0.80 M NaCl), chondroitin 4-sulfate (0.86 M NaCl), chondroitin 6-sulfate (0.95 M NaCl), dermatan sulfate (0.91 M NaCl), heparan sulfate (1.2 M NaCl), and heparin (1.35 M NaCl). Based on these observations, isolation of heparan sulfate from pig lung crude heparan sulfate fractions which contain chondroitin sulfates and dermatan sulfate was attempted, using this chromatographic technique.  相似文献   

2.
Glycosaminoglycans complex with constituents of normal human serum, a finding that was exploited to develop a competitive binding assay for these substances. Heparan sulfate was isolated from renal cortex and radiolabeled with tritiated borohydride. The elution pattern of the radiolabeled material on Sephadex G-25, Bio-Gel P-30, and AG- 1X8 resin was identical to that of unlabeled heparan sulfate. The tritiated heparan sulfate formed radiolabeled precipitates when incubated with serum and zinc acetate. Binding was dose dependent and saturable. Heparin, heparan sulfate, and the chondroitin sulfates, but not hyaluronate or keratan sulfate, competed with the radiolabeled heparan sulfate for binding in a dose-dependent manner. The assay is specific for heparin polysaccharides in chondroitinase ABC-treated samples and is sensitive to microgram quantities.  相似文献   

3.
Chen J  Duncan MB  Carrick K  Pope RM  Liu J 《Glycobiology》2003,13(11):785-794
Heparan sulfate 3-O-sulfotransferase transfers sulfate to the 3-OH position of a glucosamine to generate 3-O-sulfated heparan sulfate (HS), which is a rare component in HS from natural sources. We previously reported that 3-O- sulfotransferase isoform 5 (3-OST-5) generates both an antithrombin-binding site to exhibit anticoagulant activity and a binding site for herpes simplex virus 1 glycoprotein D to serve as an entry receptor for herpes simplex virus. In this study, we characterize the substrate specificity of 3-OST-5 using the purified enzyme. The enzyme was expressed in insect cells using the baculovirus expression approach and was purified by using heparin-Sepharose and 3',5'-ADP- agarose chromatographies. As expected, the purified enzyme generates both an antithrombin binding site and a glycoprotein D binding site. We isolated IdoUA-AnMan3S and IdoUA-AnMan3S6S from nitrous acid-degraded 3-OST-5-modified HS (pH 1.5), suggesting that 3-OST-5 enzyme sulfates the glucosamine residue that is linked to an iduronic acid residue at the nonreducing end. We also isolated a disaccharide with a structure of DeltaUA2S-GlcNS3S and a tetrasaccharide with a structure of DeltaUA2S-GlcNS-IdoUA2S-GlcNH23S6S from heparin lyases-digested 3-OST-5-modified HS. Our results suggest that 3-OST-5 enzyme sulfates both N-sulfated glucosamine and N-unsubstituted glucosamine residues. Taken together, the results indicate that 3-OST-5 has broader substrate specificity than those of 3-OST-1 and 3-OST-3. The unique substrate specificity of 3-OST-5 serves as an additional tool to study the mechanism for the biosynthesis of biologically active HS.  相似文献   

4.
Fetal calf serum (FCS) and PMA (phorbol 12-myristate-13-acetate) specifically stimulate the synthesis of heparan sulfate proteoglycan in endothelial cells. Staurosporine and n-butanol, kinase inhibitors, abolish the PMA effect. Forskolin and 8-bromo adenosine 3′:5′-cyclic monophosphate, activators of, respectively, adenylate cyclase and protein kinase A cannot reproduce the PMA effect. The kinetics of cell entry into S phase of the endothelial cells was determined by DNA synthesis ([3H]-thymidine and Br-dU incorporation), and flow cytometry. The mitogenic effect of fetal calf serum is abolished by PMA. Also, PMA pre-treatment inhibits the enhanced synthesis of heparan sulfate proteoglycan after a second PMA exposure. Remarkably, the stimulation of heparan sulfate proteoglycan synthesis by fetal calf serum and PMA seems to be mainly restricted to G1 phase. Therefore fetal calf serum and PMA cause an enhanced synthesis of heparan sulfate proteoglycan, and PMA causes a cell cycle block at G1 phase. J. Cell. Biochem. 70:563–572, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

5.
Heparan sulfate (HS) is a highly sulfated polysaccharide that plays essential physiological and pathophysiological functions. The biosynthesis of HS involves a series of specialised sulfotransferases, an epimerase and glycosyl transferases. The availability of these enzymes offers a promising method to prepare HS polysaccharides and structurally defined oligosaccharides. Given the fact that chemical synthesis of large HS oligosaccharides is extremely difficult, preparation of HS using a chemoenzymatic approach has gained momentum. This review article summarises recent progress on the development of a chemoenzymatic approach to prepare HS and HS oligosaccharides.  相似文献   

6.
Within cells, dermatan sulfate (DS) and heparan sulfate (HS) are degraded in two steps. The initial endohydrolysis of these polysaccharides is followed by the sequential action of lysosomal exoenzymes to reduce the resulting oligosaccharides to monosaccharides and inorganic sulfate. Mucopolysaccharidosis (MPS) type II is a lysosomal storage disorder caused by a deficiency of the exoenzyme iduronate-2-sulfatase (I2S). Consequently, partially degraded fragments of DS and HS have been shown to accumulate in the lysosomes of affected cells and are excreted in the urine. Di- to hexadecasaccharides, isolated from the urine of a MPS II patient using anion exchange and gel filtration chromatography, were identified using electrospray ionization-tandem mass spectrometry (ESI-MS/MS). These oligosaccharides were shown to have non-reducing terminal iduronate-2-sulfate residues by digestion with recombinant I2S. A pattern of growing oligosaccharide chains composed of alternating uronic acid and N-acetylhexosamine residues was identified and suggested to originate from DS. A series of oligosaccharides consisting of hexosamine/N-acetylhexosamine alternating with uronic acid residues was also identified and on the basis of the presence of unacetylated hexosamine; these oligosaccharides are proposed to derive from HS. The presence of both odd and even-length oligosaccharides suggests both endo-beta-glucuronidase and endo-N-acetylhexosaminidase activities toward both glycosaminoglycans. Furthermore, the putative HS oligosaccharide structures identified indicate that heparanase activities are directed toward regions of both low and high sulfation, while the N-acetylhexosaminidase activity acted only in regions of low sulfation in this polysaccharide.  相似文献   

7.
Evidence suggests that endothelial cell layer heparan sulfate proteoglycans include a variety of different sized molecules which most likely contain different protein cores. In the present report, approximately half of endothelial cell surface associated heparan sulfate proteoglycan is shown to be releasable with soluble heparin. The remaining cell surface heparan sulfate proteoglycan, as well as extracellular matrix heparan sulfate proteoglycan, cannot be removed from the cells with heparin. The heparin nonreleasable cell surface proteoglycan can be released by membrane disrupting agents and is able to intercalate into liposomes. When the heparin releasable and nonreleasable cell surface heparan sulfate proteoglycans are compared, differences in proteoglycan size are also evident. Furthermore, the intact heparin releasable heparan sulfate proteoglycan is closer in size to proteoglycans isolated from the extracellular matrix and from growth medium than to that which is heparin nonreleasable. These data indicate that cultured porcine aortic endothelial cells contain at least two distinct types of cell surface heparan sulfate proteoglycans, one of which appears to be associated with the cells through its glycosaminoglycan chains. The other (which is more tightly associated) is probably linked via a membrane intercalated protein core.Abbreviations ECM extracellular matrix - HSPG heparan sulfate proteoglycan - PAE porcine aortic endothelial - PBS phosphate buffered saline  相似文献   

8.
We examined the bactericidal activity of two proteins that are abundant in the cytoplasmic granules of human eosinophils, major basic protein (MBP) and eosinophil cationic protein (ECP). Unlike the human neutrophil's peptide defensins, both MBP and ECP killed stationary phase Staphylococcus aureus 502A in a simple nutrient-free buffer solution. Although MBP also killed Escherichia coli ML-35 with considerable efficacy under these experimental conditions, the in vitro activity of ECP against E. coli was considerably enhanced if mid-logarithmic phase bacteria replaced stationary phase organisms or if the assay medium was enriched with trypticase soy broth. The antibacterial activity of both eosinophil proteins was modulated by incubation time, protein concentration, temperature and pH. A pBR322-transformed derivative of E. coli ML-35 was used to examine the effects of ECP and MBP on integrity of the bacterial inner membrane (IM) and outer membrane. Although both MBP and ECP caused outer and inner membrane permeabilization when nutrients were present, only MBP was effective under nutrient-free conditions. Two proton ionophores (DNP and carbonyl cyanide m-chlorophenyl hydrazone) protected E. coli from the bactericidal effects of ECP but not from MBP. These findings establish that MBP and ECP have bactericidal properties and suggest that these proteins kill E. coli by similar but nonidentical mechanisms marked by an attack on the target cell's membranes. In view of evidence that high concentrations of ECP and MBP exist in cytoplasmic granules whose contents are translocated to phagocytic vacuoles, we suggest that MBP and ECP contribute to the eosinophil's ability to kill ingested bacteria.  相似文献   

9.
Expression of the basement membrane heparan sulfate proteoglycan (HSPG), perlecan (Pln), mRNA, and protein has been examined during murine development. Both Pln mRNA and protein are highly expressed in cartilaginous regions of developing mouse embryos, but not in areas of membranous bone formation. Initially detected at low levels in precartilaginous areas of d 12.5 embryos, Pln protein accumulates in these regions through d 15.5 at which time high levels are detected in the cartilage primordia. Laminin and collagen type IV, other basal lamina proteins commonly found colocalized with Pln, are absent from the cartilage primordia. Accumulation of Pln mRNA, detected by in situ hybridization, was increased in d 14.5 embryos. Cartilage primordia expression decreased to levels similar to that of the surrounding tissue at d 15.5. Pln accumulation in developing cartilage is preceded by that of collagen type II. To gain insight into Pln function in chondrogenesis, an assay was developed to assess the potential inductive activity of Pln using multipotential 10T1/2 murine embryonic fibroblast cells. Culture on Pln, but not on a variety of other matrices, stimulated extensive formation of dense nodules reminiscent of embryonic cartilaginous condensations. These nodules stained intensely with Alcian blue and collagen type II antibodies. mRNA encoding chondrocyte markers including collagen type II, aggrecan, and Pln was elevated in 10T1/2 cells cultured on Pln. Human chondrocytes that otherwise rapidly dedifferentiate during in vitro culture also formed nodules and expressed high levels of chondrocytic marker proteins when cultured on Pln. Collectively, these studies demonstrate that Pln is not only a marker of chondrogenesis, but also strongly potentiates chondrogenic differentiation in vitro.  相似文献   

10.
Heparan sulfate (HS) interacts with numerous growth factors, morphogens, receptors, and extracellular matrix proteins. Disruption of HS synthetic enzymes causes perturbation of growth factor signaling and malformation in vertebrate and invertebrate development. Our previous studies show that the O‐sulfation patterns of HS are essential for the specific binding of growth factors to HS chains, and that depletion of O‐sulfotransferases results in remarkable developmental defects in Drosophila, zebrafish, chick, and mouse. Here, we show that inhibition of chick HS‐6‐O‐sulfotransferases (HS6ST‐1 and HS6ST‐2) in the prospective limb region by RNA interference (RNAi) resulted in the truncation of limb buds and reduced Fgf‐8 and Fgf‐10 expressions in the apical ectodermal ridge and in the underlying mesenchyme, respectively. HS6ST‐2 RNAi resulted in a higher frequency of limb truncation and a more marked change in both Fgf‐8 and Fgf‐10 expressions than that achieved with HS6ST‐1 RNAi. HS6ST‐1 RNAi and HS6ST‐2 RNAi caused a significant but distinct reduction in the levels of different 6‐O‐sulfation in HS, possibly as a result of their different substrate specificities. Our data support a model where proper levels and patterns of 6‐O‐sulfation of HS play essential roles in chick limb bud development.  相似文献   

11.
Oocyte maturation-preventing factor (OMPF) was extracted from bovine granulosa cells with a buffer containing 1 M urea and 5 mM EDTA. OMPF was partially purified by gel filtration on Sephadex G25 and by reversed-phase high performance liquid chromatography. The maturation-preventing activity of purified fractions was determined by measuring their capacity to block the spontaneous dissolution of the germinal vesicle (GVBD) of isolated cumulus-enclosed mouse oocytes. Hyaluronic acid, chondroitin sulfate, heparin, heparan sulfate, keratan sulfate, and dextran sulfate at concentrations of 500 μg/ml did not affect the frequency of GVBD of isolated mouse oocytes. Heparin and heparan sulfate, however, blocked the inhibitory effect of OMPF, whereas the inhibition of GVBD induced with dibutyryl cAMP, forskolin, W7 (calmodulin antagonist), and 3-isobutyl-l-methylxanthine (phosphodiesterase inhibitor) was not blocked. OMPF was eluted in the adsorbed fraction when chromatographed on heparin-Agarose, showing interaction of OMPF with heparin. The present results suggest that the glycosaminoglycan matrix may influence OMPF action on oocytes.  相似文献   

12.
Mummery RS  Mulloy B  Rider CC 《Glycobiology》2007,17(10):1094-1103
Recombinant human betacellulin binds strongly to heparin, requiring of the order of 0.8 M NaCl for its elution from a heparin affinity matrix. This is in complete contrast to the prototypic member of its cytokine superfamily, epidermal growth factor, which fails to bind to the column at physiological pH and strength. We used a well-established heparin binding ELISA to demonstrate that fucoidan and a highly sulfated variant of heparan sulfate compete strongly for heparin binding. Low sulfated heparan sulfates and also chondroitin sulfates are weaker competitors. Moreover, although competitive activity is reduced by selective desulfation, residual binding to extensively desulfated heparin remains. Even carboxyl reduction followed by extensive desulfation does not completely remove activity. We further demonstrate that both hyaluronic acid and the E. coli capsular polysaccharide K5, both of which are unsulfated polysaccharides with unbranched chains of alternating N-acetylglucosamine linked beta(1-4) to glucuronic acid, are also capable of a limited degree of competition with heparin. Heparin protects betacellulin from proteolysis by LysC, but K5 polysaccharide does not. Betacellulin possesses a prominent cluster of basic residues, which is likely to constitute a binding site for sulfated polysaccharides, but the binding of nonsulfated polysaccharides may take place at a different site.  相似文献   

13.
The glycosaminoglycan heparan sulfate (HS), present at the surface of most cells and ubiquitous in extracellular matrix, binds many soluble extracellular signalling molecules such as chemokines and growth factors, and regulates their transport and effector functions. It is, however, unknown whether upon binding HS these proteins can affect the long-range structure of HS. To test this idea, we interrogated a supramolecular model system, in which HS chains grafted to streptavidin-functionalized oligoethylene glycol monolayers or supported lipid bilayers mimic the HS-rich pericellular or extracellular matrix, with the biophysical techniques quartz crystal microbalance (QCM-D) and fluorescence recovery after photobleaching (FRAP). We were able to control and characterize the supramolecular presentation of HS chains—their local density, orientation, conformation and lateral mobility—and their interaction with proteins. The chemokine CXCL12α (or SDF-1α) rigidified the HS film, and this effect was due to protein-mediated cross-linking of HS chains. Complementary measurements with CXCL12α mutants and the CXCL12γ isoform provided insight into the molecular mechanism underlying cross-linking. Fibroblast growth factor 2 (FGF-2), which has three HS binding sites, was also found to cross-link HS, but FGF-9, which has just one binding site, did not. Based on these data, we propose that the ability to cross-link HS is a generic feature of many cytokines and growth factors, which depends on the architecture of their HS binding sites. The ability to change matrix organization and physico-chemical properties (e.g. permeability and rigidification) implies that the functions of cytokines and growth factors may not simply be confined to the activation of cognate cellular receptors.  相似文献   

14.
Heparin and heparan sulfate: biosynthesis, structure and function   总被引:7,自引:0,他引:7  
Heparin and heparan sulfate glycosaminoglycans are acidic complex polysaccharides found on the cell surface and in the extracellular matrix. Recent progress has uncovered a virtual explosion of important roles of these biopolymers in fundamental biological processes. Advances in the understanding of biosynthesis and structure and the development of novel analytical methods for composition and sequence analysis have provided remarkable insights into structure/function relationships of these complex and once elusive polysaccharides.  相似文献   

15.
The glucosaminoglycans isolated from the skin of control and streptozotocin-diabetic rats were fractionated on ion-exchange chromatography into a heparan sulfate (HS)-like and a heparin-like species. In addition, a low sulfated fraction was isolated from the diabetics. The HS and heparin-like fractions isolated from the diabetics (in contrast to the low sulfated fractions) retained high affinity for the acidic (FGF-1) and basic (FGF-2) fibroblast growth factors. In culture, the fractions purified from the control rats and the heparin-like material isolated from the diabetics mediated the biological activity of both FGFs in a dose-dependent manner. By contrast, the diabetic HS-like fractions promoted the biological activity of FGF-2 but not of FGF-1. The results support the idea that the structural motives in HS required for FGF-1 and FGF-2 mediated receptor signalling are different. They may be relevant to the impaired wound healing observed in the disease. This revised version was published online in November 2006 with corrections to the Cover Date.  相似文献   

16.
In eukaryotic cells, mitochondrial dysfunction is associated with a variety of human diseases. Delivery of exogenous functional mitochondria into damaged cells has been proposed as a mechanism of cell transplant and physiological repair for damaged tissue. We here demonstrated that isolated mitochondria can be transferred into homogeneic and xenogeneic cells by simple co‐incubation using genetically labelled mitochondria, and elucidated the mechanism and the effect of direct mitochondrial transfer. Intracellular localization of exogenous mitochondria was confirmed by PCR, real‐time PCR, live fluorescence imaging, three‐dimensional reconstruction imaging, continuous time‐lapse microscopic observation, flow cytometric analysis and immunoelectron microscopy. Isolated homogeneic mitochondria were transferred into human uterine endometrial gland‐derived mesenchymal cells in a dose‐dependent manner. Moreover, mitochondrial transfer rescued the mitochondrial respiratory function and improved the cellular viability in mitochondrial DNA‐depleted cells and these effects lasted several days. Finally, we discovered that mitochondrial internalization involves macropinocytosis. In conclusion, these data support direct transfer of exogenous mitochondria as a promising approach for the treatment of various diseases.  相似文献   

17.
Nitric oxide (NO), a reactive nitrogen species, plays an important role in inflammatory lung damage. In the present study, we investigated the role of NO in DNA-binding activity of NF-B in macrophages stimulated with silica or other inflammatory stimulants. Treatment of mouse macrophages (RAW264.7 cells) with a selective inhibitor of inducible nitric oxide synthase (iNOS), L-N6-(1-iminoethyl) lysine (L-NIL), or a nonselective iNOS inhibitor, N-nitro-L-arginine methylester (L-NAME), resulted in inhibition of silica-induced nitric oxide production as well as silica-induced NF-B activation. L-NIL also effectively inhibited NF-B activation induced by other inflammatory stimulants, such as lipopolysaccharide (LPS) or muramyl dipeptide (MDP). These inhibitory effects of L-NIL and L-NAME on silica- or LPS-induced NF-B activation were also observed in primary rat alveolar macrophages. Furthermore, NO generating compounds, such as sodium nitroprusside (SNP) and 3-morpholinosydnonimine (SIN-1), caused a dose-dependent increase in NF-B activation, which was positively correlated with the level of NO production. Specific inhibitors of protein tyrosine kinase, such as genistein and AG494, prevented NF-B activation in SNP- or SIN-1 treated cells, suggesting involvement of tyrosine kinase in the NO signaling pathway leading to NF-B activation. In contrast, inhibitors of protein kinase C or A, such as staurosporine or H89, had no inhibitory effect on SIN-1 induced NF-B activation. Metalloporphyrins, such as tetrakis (N-methyl-4-pyridyl) porphyrinato iron (III) (Fe-TMPyP) and Zn-TMPyP which are known to alter NO-dependent activity, markedly inhibited silica- and LPS-induced NF-B activation. The results suggest that NF-B activation in macrophages can be induced under certain conditions by nitric oxide and that nitric oxide produced by phagocytes exposed to inflammatory agents may up-regulate the activation of NF-B.  相似文献   

18.
Aging poses one of the largest risk factors for the development of cardiovascular disease. The increased propensity toward vascular pathology with advancing age maybe explained, in part, by a reduction in the ability of circulating endothelial progenitor cells to contribute to vascular repair and regeneration. Although there is evidence to suggest that colony forming unit‐Hill cells and circulating angiogenic cells are subject to age‐associated changes that impair their function, the impact of aging on human outgrowth endothelial cell (OEC) function has been less studied. We demonstrate that OECs isolated from cord blood or peripheral blood samples from young and old individuals exhibit different characteristics in terms of their migratory capacity. In addition, age‐related structural changes were discovered in OEC heparan sulfate (HS), a glycocalyx component that is essential in many signalling pathways. An age‐associated decline in the migratory response of OECs toward a gradient of VEGF significantly correlated with a reduction in the relative percentage of the trisulfated disaccharide, 2‐O‐sulfated‐uronic acid, N, 6‐O‐sulfated‐glucosamine (UA[2S]‐GlcNS[6S]), within OEC cell surface HS polysaccharide chains. Furthermore, disruption of cell surface HS reduced the migratory response of peripheral blood‐derived OECs isolated from young subjects to levels similar to that observed for OECs from older individuals. Together these findings suggest that aging is associated with alterations in the fine structure of HS on the cell surface of OECs. Such changes may modulate the migration, homing, and engraftment capacity of these repair cells, thereby contributing to the progression of endothelial dysfunction and age‐related vascular pathologies.  相似文献   

19.
The heparan sulfate proteoglycans that bind and activate antithrombin III (aHSPGs) are synthesized by endothelial cells as well as other nonvascular cells. We determined the amounts of cell surface–associated and soluble aHSPGs generated by the rat fat pad endothelial (RFP) cell line and the fibroblast (LTA) cell line. The RFP cells exhibit higher levels of cell surface–associated aHSPGs as compared to LTA cells, whereas LTA cells release larger amounts of soluble aHSPGs as compared to RFP cells. After confluence RFP cells show an increase in both cell surface–associated and soluble aHSPGs. In contrast, postconfluent LTA cells maintain a constant level of cell surface–associated and soluble aHSPGs. These observations indicate that different cells types can preferentially accumulate aHSPGs as cell surface–associated or soluble forms which could reflect alternate biological functions.  相似文献   

20.
Understanding the process of wound healing will provide valuable insight for the development of new strategies to treat diseases associated with improper regeneration, such as blindness induced by corneal scarring. Heparan sulfate proteoglycans (HSPG) are not normally expressed in the corneal stroma, but their presence at sites of injury suggests their involvement in the wound healing response. Primary cultured corneal stromal fibroblasts constitutively express HSPG and represent an injured phenotype. Recently, nuclear localization of HSPG was shown to increase in corneal stromal fibroblasts plated on fibronectin (FN), an extracellular matrix protein whose appearance in the corneal stroma correlates with injury. One possible role for the nuclear localization of HSPG is to function as a shuttle for the nuclear transport of heparin-binding growth factors, such as basic fibroblast growth factor (FGF-2). Once in the nucleus, these growth factors might directly modulate cellular activities. To investigate this hypothesis, cells were treated with (125)I-labelled FGF-2 under various conditions and fractionated. Our results show that nuclear localization of FGF-2 was increased in cells plated on FN compared to those on collagen type I (CO). Interestingly, FGF-2-stimulated proliferation was increased in cells plated on FN compared to CO and this effect was absent in the presence of heparinase III. Furthermore, pre-treatment with heparinase III decreased nuclear FGF-2, and CHO cells defective in the ability to properly synthesize heparan sulfate chains showed reduced nuclear FGF-2 indicating that the heparan sulfate chains of HSPG are critical for this process. HSPG signaling, particularly through the cytoplasmic tails of syndecans, was investigated as a potential mechanism for the nuclear localization of FGF-2. Treatment with phorbol 12-myristate-13-acetate (PMA), under conditions that caused downregulation of protein kinase Calpha (PKCalpha), decreased nuclear FGF-2. Using pharmacological inhibitors of specific PKC isozymes, we elucidated a potential mode of regulation whereby PKCalpha mediates the nuclear localization of FGF-2 and PKCdelta inhibits it. Our studies suggest a novel mechanism in which FGF-2 translocates to the nucleus in response to injury.  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