首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Calcium phospholipid dependent protein kinase C (PKC) is activated by diacylglycerol (DG) and by phorbol esters and is recognized to be the phorbol ester receptor of cells; DG displaces phorbol ester competitively from PKC. A phospholipid, phosphatidylinositol 4,5-bisphosphate (PIP2), can also activate PKC in the presence of phosphatidylserine (PS) and Ca2+ with a KPIP2 of 0.04 mol %. Preliminary experiments have suggested a common binding site for PIP2 and DG on PKC. Here, we investigate the effect of PIP2 on phorbol ester binding to PKC in a mixed micellar assay. In the presence of 20 mol % PS, PIP2 inhibited specific binding of [3H]phorbol 12,13-dibutyrate (PDBu) in a dose-dependent fashion up to 85% at 1 mol %. Inhibition of binding was more pronounced with PIP2 than with DG. Scatchard analysis indicated that the decrease in binding of PDBu in the presence of PIP2 is the result of an altered affinity for the phorbol ester rather than of a change in maximal binding. The plot of apparent dissociation constants (Kd') against PIP2 concentration was linear over a range of 0.01-1 mol % with a Ki of 0.043 mol % and confirmed the competitive nature of inhibition between PDBu and PIP2. Competition between PIP2 and phorbol ester could be demonstrated in a liposomal assay system also. These results indicate that PIP2, DG, and phorbol ester all compete for the same activator-receiving region on the regulatory moiety of protein kinase C, and they lend support to the suggestion that PIP2 is a primary activator of the enzyme.  相似文献   

2.
We have constructed the expression plasmids harboring protein kinase C (PKC) mutant cDNAs with a series of deletions in the PKC coding region. These plasmids were transfected into COS7 cells to characterize the PKC mutants. Immunoblot analysis using the anti-PKC antibody identified proteins with the Mr values expected from the PKC mutant cDNAs in the extracts from COS7 cells. The wild-type PKC, when expressed in COS7 cells, conferred increased phorbol ester binding activity on intact cells; but the PKC mutants with the deletion around the C1 region did not show this activity. The wild-type PKC showed protein kinase activity dependent on phospholipid, Ca2+, and phorbol ester, whereas these PKC mutants exhibited protein kinase activity independent of the activators in a cell-free system. A PKC mutant cDNA with the deletion in the C2 region gave increased phorbol ester binding activity. Protein kinase activity of this mutant was much less dependent on Ca2+ compared with the wild-type PKC. A PKC mutant cDNA with the deletion in the C3 region conferred increased phorbol ester binding activity, but neither activator-dependent nor -independent protein kinase activity. These results indicate that elimination of the C1 region of PKC gives rise to constitutively active PKC independent of phospholipid, Ca2+, and phorbol ester and that the C1-C3 regions play distinct roles in the regulatory and catalytic function of PKC. In another series of experiments, transfection of some PKC mutant cDNAs with the deletions around the C1 region into Chinese hamster ovary and Jurkat cells activated the activator protein-1-binding element or the c-fos gene enhancer linked to the chloramphenicol acetyltransferase reporter gene in the absence of phorbol ester. Microinjection of these constructs into Xenopus oocytes induced initiation of germinal vesicle breakdown, indicating that they stimulated the PKC pathway in vivo. Thus, the phorbol ester-independent PKC mutant cDNAs could be a powerful tool to investigate the transmembrane signaling pathway mediated by PKC.  相似文献   

3.
Resveratrol (1) is a naturally occurring phytoalexin that affects a variety of human disease models, including cardio- and neuroprotection, immune regulation, and cancer chemoprevention. One of the possible mechanisms by which resveratrol affects these disease states is by affecting the cellular signaling network involving protein kinase C (PKC). PKC is the family of serine/threonine kinases, whose activity is inhibited by resveratrol. To develop PKC isotype selective molecules on the resveratrol scaffold, several analogs (2–5) of resveratrol with a long aliphatic chain varying with number of unsaturated doubled bonds have been synthesized, their cytotoxic effects on CHO-K1 cells are measured and their effects on the membrane translocation properties of PKCα and PKCε have been determined. The analogs showed less cytotoxic effects on CHO-K1 cells. Analog 4 with three unsaturated double bonds in its aliphatic chain activated PKCα, but not PKCε. Analog 4 also activated ERK1/2, the downstream proteins in the PKC signaling pathway. Resveratrol analogs 2–5, however, did not show any inhibition of the phorbol ester-induced membrane translocation for either PKCα or PKCε. Molecular docking of 4 into the activator binding site of PKCα revealed that the resveratrol moiety formed hydrogen bonds with the activator binding residues and the aliphatic chain capped the activator binding loops making its surface hydrophobic to facilitate its interaction with the plasma membrane. The present study shows that subtle changes in the resveratrol structure can have profound impact on the translocation properties of PKCs. Therefore, resveratrol scaffold can be used to develop PKC selective modulators for regulating associated disease states.  相似文献   

4.
We constructed mutant protein kinase C (PKC) cDNAs which expressed PKC activity in vivo in the absence of phorbol ester activation. A hybrid PKC gene, PKAC, was constructed by substituting the coding region for the N-terminal 253 amino acids of PKC alpha with the N-terminal 17 amino acids of the cyclic AMP-dependent protein kinase catalytic subunit (PKA). A truncated PKC gene, delta PKC beta, lacking the coding region for amino acid positions 6 to 159 of PKC beta was also constructed. These mutant kinase genes expressed under the control of the SR alpha promoter activated the c-fos gene enhancer in Jurkat cells and initiated maturation of Xenopus laevis oocytes. Phorbol ester binding activity was absent in both constructs but was preserved in another hybrid gene, PKCA, which was composed of the coding region for 1 to 253 amino acids of PKC alpha at the N-terminal side and the coding region for 18 to 350 amino acids of PKA at the C-terminal side. These results indicate that elimination of the regulatory domain of PKC produces constitutively active PKC that can bypass activation by the phorbol ester. delta PKC beta, in synergy with a calcium ionophore, was capable of activating the interleukin 2 promoter, indicating that cooperation of PKC-dependent and calcium-dependent pathways is necessary for activation of the interleukin 2 gene.  相似文献   

5.
In order to examine the role of phosphatidylinositol bisphosphate (PIP2) hydrolysis in B cell activation, we studied the effect of various classes of protein kinase C (PKC) activators on anti-Ig-mediated B cell stimulation. Anti-Ig-stimulated PIP2 hydrolysis, elevations in [Ca2+]i, and induction of DNA synthesis were inhibited by PMA (a phorbol ester) as previously reported. In contrast, indolactam (an alkaloid PKC activator) inhibited PIP2 hydrolysis and elevations in [Ca2+]i, but stimulated rather than inhibited cellular proliferation. In order to examine whether the binding avidity of the PKC activators to PKC played a role in determining their activity to stimulate or inhibit B cell activation, we studied two other PKC activators, bryostatin and mezerein. Again, both inhibited anti-Ig mediated PIP2 hydrolysis and elevations in [Ca2+]i, whereas only the former inhibited induction of DNA synthesis. These data suggest that a) high levels of PIP2 hydrolysis and elevations in [Ca2+]i are not essential for anti-Ig-mediated induction of B cell DNA synthesis and b) activation of PKC may induce both stimulatory and inhibitory pathways of B cell activation, and whether stimulation or inhibition of cell activation is observed may depend on the combined intensity of these two signals.  相似文献   

6.
The immunologic effects of bryostatin (Bryo), a PKC activator with antineoplastic activity, were assessed and compared to PMA. Bryo induced IL-2R expression on CD4+ and CD8+ human T lymphocytes with a dose response comparable to PMA. However, Bryo induced only a marginal proliferative response as compared with the vigorous response induced by PMA. Bryo mediated functional receptor expression because the proliferative response was enhanced by addition of rIL-2. Furthermore, the proliferative response was inhibited by the relatively specific Ca+, phospholipid-dependent protein kinase (PKC) inhibitor, H-7, indicating a role of PKC in Bryo-induced activation. Addition of the calcium ionophore, ionomycin, to Bryo-stimulated lymphocytes resulted in the production and secretion of IL-2 with a concomitant proliferative response. This effect of the calcium ionophore could be inhibited by cyclosporine with identical results obtained in PMA-stimulated cultures. A most intriguing finding was that Bryo could effectively antagonize PMA-induced T cell proliferation. Although this mechanism of inhibition is unclear, a discussion with respect to differential effects on potential intracellular PKC isoforms is provided. These studies indicated that Bryo has potent immunopotentiating properties that share some similar effects of the phorbol ester, PMA, but offers the additional property of modulating other phorbol ester effects on proliferation.  相似文献   

7.
The protein kinase C (PKC) family of serine/threonine kinases is an attractive drug target because of its involvement in the regulation of various cellular functions, including cell growth, differentiation, metabolism, and apoptosis. The endogenous PKC activator diacylglycerol contains two long carbon chains, which are attached to the glycerol moiety via ester linkage. Natural product curcumin (1), the active constituent of Curcuma L., contains two carbonyl and two hydroxyl groups. It modulates PKC activity and binds to the activator binding site (Majhi et al., Bioorg. Med. Chem.2010, 18, 1591). To investigate the role of the carbonyl and hydroxyl groups of curcumin in PKC binding and to develop curcumin derivatives as effective PKC modulators, we synthesized several isoxazole and pyrazole derivatives of curcumin (2-6), characterized their absorption and fluorescence properties, and studied their interaction with the activator-binding second cysteine-rich C1B subdomain of PKCδ, PKCε and PKCθ. The EC(50)s of the curcumin derivatives for protein fluorescence quenching varied in the range of 3-25 μM. All the derivatives showed higher binding with the PKCθC1B compared with PKCδC1B and PKCεC1B. Fluorescence emission maxima of 2-5 were blue shifted in the presence of the C1B domains, confirming their binding to the protein. Molecular docking revealed that hydroxyl, carbonyl and pyrazole ring of curcumin (1), pyrazole (2), and isoxazole (4) derivatives form hydrogen bonds with the protein residues. The present result shows that isoxazole and pyrazole derivatives bind to the activator binding site of novel PKCs and both carbonyl and hydroxy groups of curcumin play roles in the binding process, depending on the nature of curcumin derivative and the PKC isotype used.  相似文献   

8.
The effects of a synthetic protein kinase C (PKC) activator, (-)-indolactam V (ILV), were studied in SH-SY5Y human neuroblastoma cells. (-)-ILV induced a translocation of PKC from cytosol to plasma membrane and displaced 3H-phorbol dibutyrate binding in the micromolar range. In addition, (-)-ILV caused a decreased sensitivity of cells to muscarinic agonist-induced Ca2+ mobilization measured with quin-2 and induced a down-regulation of cell surface muscarinic receptors. All the changes induced by (-)-ILV were similar in magnitude to those seen with the phorbol ester tetradecanoyl phorbol acetate (TPA). The results suggest that (-)-ILV is a full activator of PKC and a promising alternative to phorbol esters in studies on mechanism of actions of PKC.  相似文献   

9.
Calcium/phosphatidylserine-dependent protein kinase C (PKC) is activated by phosphatidylinositol 4,5-bisphosphate (PIP2), as well as by diacylglycerol (DG) and phorbol esters. Here we report that PIP2, like DG, increases the affinity of PKC for Ca2+, and causes Ca(2+)-dependent translocation of the enzyme from the soluble to a particulate fraction (liposomes). Phosphatidylinositol 4-phosphate (PIP) also displaces phorbol ester from PKC and causes Ca(2+)-dependent translocation of the enzyme to liposomes, but is much less efficient than PIP2, and a much weaker activator, with a histone phosphorylation v(PIP)/v(PIP2) of approximately 0.15. Scatchard analysis indicates competitive inhibition between PIP and phorbol ester with Ki(PIP) = 0.26 mol% as compared with Ki(PIP2) = 0.043 mol%. No effect of phosphatidylinositol (PI) on phorbol ester binding to PKC, translocation of PKC, or activation of PKC was observed. These results suggest that both PIP and PIP2 can complex with PKC, but full activation of the enzyme takes place only when PIP is converted to PIP2. We suggest that an inositide interconversion shuttle has a role in the regulation of protein phosphorylation.  相似文献   

10.
In vitro linoleic acid activation of protein kinase C   总被引:2,自引:0,他引:2  
The importance of membrane fluidity in the activation of protein kinase C (PKC) was examined using the membrane fluidizer, linoleic acid, in a well-defined model membrane system. Biochemical and biophysical properties of the system were monitored. Linoleic acid activated PKC to a level of 50% of that observed for diacylglycerol. In contrast, linoleic acid did not directly interact with the phorbol ester binding site as did diacylglycerol. This was determined by the lack of involvement of the ionizable group of the fatty acid with activity and the enhancement of phorbol ester binding by linoleic acid and its ester analogs. The membrane fluidity of this model membrane system in the presence of linoleic acid was increased as determined by fluorescence polarization. This increased the availability of phospholipids, thus, explaining the linoleic acid-induced enhancement of phorbol ester binding. The PKC conformation as determined from intrinsic tryptophan fluorescence spectra was different for lipid mixtures containing linoleic acid or diacylglycerol correlating with the difference in biochemical activation properties. This study provides evidence that membrane fluidization is not the predominant function of the lipid activator in PKC activation, but may play a role in obtaining the preferred membrane state for maximal activation.  相似文献   

11.
12.
A protein kinase C alpha (PKC alpha) cDNA confers increased phorbol ester binding activity to intact cells when transiently expressed in COS cells or expressed stably in transfected rat 3Y1 fibroblasts. A point mutant (PKC alpha K----R) of PKC alpha, where Lys368 at the putative ATP-binding site is replaced with Arg, confers enhanced phorbol ester binding activity to both transiently and stably expressed COS and 3Y1 cells, respectively. Like endogenous and exogenously expressed wild type PKC alpha, the mutant PKC alpha K----R is translocated from the cytosol to the particulate fraction when cells are treated with a phorbol ester, 12-O-tetradecanoylphorbol-13-acetate (TPA). On the other hand, the mutant PKC alpha K----R is not degraded when cells are treated with TPA, making a clear contrast to wild type PKC alpha; i.e. the mutant is resistant to TPA-mediated down-regulation. The mutant lacks kinase activity as expected, as judged by autophosphorylation and by a kinase assay using a peptide substrate, although the phorbol ester binding activity remains intact. These results suggest a link between the kinase activity of PKC alpha and the sensitivity to TPA-mediated proteolytic degradation. We propose that autophosphorylation of PKC alpha is a prerequisite for proteolytic cleavage associated with the down-regulation of PKC alpha.  相似文献   

13.
Calcium-calmodulin (CaM) binding to the epidermal growth factor receptor (EGFR) has been shown to both inhibit and stimulate receptor activity. CaM binds to the intracellular juxtamembrane (JM) domain (Met645-Phe688) of EGFR. Protein kinase C (PKC) mediated phosphorylation of Thr654 occurs within this domain. CaM binding to the JM domain inhibits PKC phosphorylation and conversely PKC mediated phosphorylation of Thr654 or Glu substitution of Thr654 inhibits CaM binding. A second threonine residue (Thr669) within the JM domain is phosphorylated by the mitogen-activated protein kinase (MAPK). Previous results have shown that CaM interferes with EGFR-induced MAPK activation. If and how phosphorylation of Thr669 affects CaM-EGFR interaction is however not known.In the present study we have used surface plasmon resonance (BIAcore) to study the influence of Thr669 phosphorylation on real time interactions between the intracellular juxtamembrane (JM) domain of EGFR and CaM. The EGFR-JM was expressed as GST fusion proteins in Escherichia coli and phosphorylation was mimicked by generating Glu substitutions of either Thr654 or Thr669. Purified proteins were coupled to immobilized anti-GST antibodies at the sensor surface and increasing concentration of CaM was applied. When mutating Thr654 to Glu654 no specific CaM binding could be detected. However, neither single substitutions of Thr669 (Gly669 or Glu669) nor double mutants Gly654/Gly669 or Gly654/Glu669 influenced the binding of CaM to the EGFR-JM. This clearly shows that PKC may regulate EGF-mediated CaM signalling through phosphorylation of Thr654 whereas phosphorylation of Thr669 seems to play a CaM independent regulatory role. The role of both residues in the EGFR-calmodulin interaction was also studied in silico. Our modelling work supports a scenario where Thr654 from the JM domain interacts with Glu120 in the calmodulin molecule. Phosphorylation of Thr654 or Glu654 substitution creates a repulsive electrostatic force that would diminish CaM binding to the JM domain. These results are in line with the Biacore experiments showing a weak binding of the CaM to the JM domain with Thr654 mutated to Glu. Furthermore, these results provide a hypothesis to how CaM binding to EGFR might both positively and negatively interfere with EGFR-activity.  相似文献   

14.
To further investigate the effect of single amino acid substitution on the structure and function of the gramicidin channel, an analogue of gramicidin A (GA) has been synthesized in which Trp(15) is replaced by Gly in the critical aqueous interface and cation binding region. The structure of Gly(15)-GA incorporated into SDS micelles has been determined using a combination of 2D-NMR spectroscopy and molecular modeling. Like the parent GA, Gly(15)-GA forms a dimeric channel composed of two single-stranded, right-handed beta(6.3)-helices joined by hydrogen bonds between their N-termini. The replacement of Trp(15) by Gly does not have a significant effect on backbone structure or side chain conformations with the exception of Trp(11) in which the indole ring is rotated away from the channel axis. Measurement of the equilibrium binding constants and Delta G for the binding of monovalent cations to GA and Gly(15)-GA channels incorporated into PC vesicles using (205)Tl NMR spectroscopy shows that monovalent cations bind much more weakly to the Gly(15)-GA channel entrance than to GA channels. Utilizing the magnetization inversion transfer NMR technique, the transport of Na(+) ions through GA and Gly(15)-GA channels incorporated into PC/PG vesicles has been investigated. The Gly(15) substitution produces an increase in the activation enthalpy of transport and thus a significant decrease in the transport rate of the Na(+) ion is observed. The single-channel appearances show that the conducting channels have a single, well-defined structure. Consistent with the NMR results, the single-channel conductances are reduced by 30% and the lifetimes by 70%. It is concluded that the decrease in cation binding, transport, and conductance in Gly(15)-GA results from the removal of the Trp(15) dipole and, to a lesser extent, the change in orientation of Trp(11).  相似文献   

15.
Protein kinase C (PKC) comprises a family of distinct isoenzymes that are involved in signal transduction pathways linking the cell to triggers perceived via membrane receptors. These isoenzymes differ in their tissue distribution, activation requirements, and substrate specificity. One common denominator among different PKC subspecies is their activation by phorbol esters. We have developed a sensitive method permitting the measurement of phorbol ester binding sites, their quantitation, as well as their dissociation kinetics, by performing cytofluorometric analyses on intact cells or on isolated PKC associated to phosphatidylserine vesicles incubated in the presence of fluorochrome-labeled phorbol ester. Both PKC isozymes beta I/beta II and alpha from brain and spleen after incorporation into phosphatidylserine vesicles, display affinities with apparent Kd of 120 and 50 nM, respectively; although PKC gamma from brain exhibits a Kd of 210 nM. In addition to these receptors, on PKC isozymes from spleen, an intermediate affinity phorbol ester receptor (Kd of 3 nM) and an additional high affinity phorbol ester binding site with a Kd of 0.1 to 0.5 nM were also detected. This latter receptor comigrates with high m.w. PKC isoforms. In different cell lines, the phorbol ester binding patterns, as well as the expression of individual PKC isoenzymes, could be positively correlated.  相似文献   

16.
17.
Gonadotropin-releasing hormone (GnRH) regulates pituitary gonadotropin release by a Ca2+-dependent mechanism involving receptor-mediated phosphoinositide hydrolysis. Previous studies indicate that activation of pituitary protein kinase C (PKC), while not required for acute gonadotropin release in response to GnRH, is likely involved in the chronic regulation of gonadotrope responsiveness. Studies from our laboratory have shown that activation of PKC by phorbol esters produces both the uncoupling of GnRH-stimulated phosphoinositide hydrolysis and the selective enhancement of GnRH agonist binding in pituitary cell cultures. In the present work, we have examined the possibility that these processes are related in mechanism. Dissociation of bound agonist radioligand at 23 degrees C was found to be reduced in the presence of phorbol esters, and ligand bound in the presence of phorbol ester was resistant to displacement by competing ligands at 4 degrees C. However, agonist bound in the presence of phorbol ester was dissociable by subsequently washing cells at pH 3. Receptor photoaffinity labeling studies confirmed that agonist association with membrane component(s) identified as the GnRH receptor was increased in the presence of phorbol ester. These results suggest that, in the presence of a phorbol ester PKC activator, agonist-occupied GnRH receptors remain at the cell surface, but are sequestered in some manner. In other experiments, cell preloaded with [3H]inositol were treated with GnRH agonist ligand and phorbol ester at 4 degrees C to form a pool of sequestered, agonist-occupied receptors, and then displaceable (nonsequestered) agonist was removed by incubation with antagonist ligand. After addition of LiCl and warming to 37 degrees C, [3H]inositol phosphate production (an index of phosphoinositide hydrolysis) in phorbol ester-treated cells was reduced to 67% of vehicle control, although residual specific agonist binding had been increased to greater than 300% of control. The appearance of sequestered receptors and inhibition of [3H]inositol phosphate production had similar phorbol ester concentration dependencies. These results suggest that the same agonist-occupied GnRH receptors sequestered as a result of PKC activation also are preferentially uncoupled from phosphoinositide hydrolysis.  相似文献   

18.
Although activation of protein kinase C (PKC) is known to promote cell survival and protect against cell death, the PKC targets and pathways that serve this function have remained elusive. Here we demonstrate that two potent activators of PKC, 12-O-tetradecanoylphorbol-13-acetate and bryostatin, both stimulate phosphorylation of Bad at Ser(112), a site known to regulate apoptotic cell death by interleukin-3. PKC inhibitors but not PI 3-kinase/Akt inhibitors block 12-O-tetradecanoylphorbol-13-acetate-stimulated Bad phosphorylation. PKC isoforms tested in vitro were unable to phosphorylate Bad at Ser(112), suggesting that PKC acts indirectly to activate a downstream Bad kinase. p90(RSK) and family members RSK-2 and RSK-3 are activated by phorbol ester and phosphorylate Bad at Ser(112) both in vitro and in vivo. p90(RSK) stimulates binding of Bad to 14-3-3 and blocks Bad-mediated cell death in a Ser(112)-dependent manner. These findings suggest that p90(RSK) can function in a PKC-dependent pathway to promote cell survival via phosphorylation and inactivation of Bad-mediated cell death.  相似文献   

19.
Y Madrona  S Tripathi  H Li  TL Poulos 《Biochemistry》2012,51(33):6623-6631
The crystal structure of the P450cin substrate-bound nitric oxide complex and the substrate-free form have been determined revealing a substrate-free structure that adopts an open conformation relative to the substrate-bound structure. The region of the I helix that forms part of the O(2) binding pocket shifts from an α helix in the substrate-free form to a π helix in the substrate-bound form. Unique to P450cin is an active site residue, Asn242, in the I helix that H-bonds with the substrate. In most other P450s this residue is a Thr and plays an important role in O(2) activation by participating in an H-bonding network required for O(2) activation. The π/α I helix transition results in the carbonyl O atom of Gly238 moving in to form an H-bond with the water/hydroxide ligand in the substrate-free form. The corresponding residue, Gly248, in the substrate-free P450cam structure experiences a similar motion. Most significantly, in the oxy-P450cam complex Gly248 adopts a position midway between the substrate-free and -bound states. A comparison between these P450cam and the new P450cin structures provides insights into differences in how the two P450s activate O(2). The structure of P450cin complexed with nitric oxide, a close mimic of the O(2) complex, shows that Gly238 is likely to form tighter interactions with ligands than the corresponding Gly248 in P450cam. Having a close interaction between an H-bond acceptor, the Gly238 carbonyl O atom, and the distal oxygen atom of O(2) will promote protonation and hence further reduction of the oxy complex to the hydroperoxy intermediate resulting in heterolytic cleavage of the peroxide O-O bond and formation of the active ferryl intermediate required for substrate hydroxylation.  相似文献   

20.
Cell-cell adhesion mediated by cadherins is believed to play an essential role in the control of cell differentiation and tissue formation. Our recent studies indicate that N-cadherin is involved in human osteoblast differentiation. However, the signalling molecules that regulate cadherins in osteoblasts are not known. We tested the possibility that N-cadherin expression and function may be regulated by direct activation of protein kinase C (PKC) in human osteoblasts. Treatment of immortalized human neonatal calvaria (IHNC) cells with phorbol 12,13-dibutyrate (100 nM) transiently increased PKC activity. RT-PCR analysis showed that transient treatment with phorbol ester transiently increased N-cadherin mRNA levels at 4-12 h. Western blot analysis showed that N-cadherin protein levels were increased by phorbol ester at 24-48 h, and this was confirmed by immunocytochemical analysis. In contrast, E-cadherin expression was not affected. Transient treatment of IHNC cells with phorbol ester increased cell-cell aggregation, which was suppressed by neutralizing N-cadherin antibody, showing that the increased N-cadherin induced by phorbol ester was functional. Finally, phorbol ester dose-dependently increased alkaline phosphatase activity, an early marker of osteoblast differentiation. This effect was comparable to the promoting effect of BMP-2, a potent activator of osteoblast differentiation. These data show that direct activation of PKC by phorbol ester increases N-cadherin expression and function, and promotes ALP activity in human calvaria osteoblasts, which provides a signaling mechanism by which N-cadherin is regulated and suggests a role for PKC in N-cadherin-mediated control of human osteoblast differentiation.  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