首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Sphingosine, a metabolite of membrane sphingolipids, is generally considered to be cytotoxic for a variety of cell types. However, we have found that sphingosine at low concentrations stimulates DNA synthesis and acts synergistically with known growth factors to induce proliferation of quiescent Swiss 3T3 fibroblasts. Structurally related analogs of sphingosine, such as N-stearoylsphingosine and other long chain aliphatic amines, had no mitogenic effects, suggesting that sphingosine did not induce nonspecific membrane perturbations. Sphingosine, which has been proposed to be a physiological inhibitor of protein kinase C, also markedly potentiates the mitogenic effect of the tumor promoter, 12-O-tetradecanoylphorbol-13-acetate (TPA). Sphingosine still stimulates DNA synthesis in cells made protein kinase C deficient by prolonged treatment with phorbol ester. At mitogenic concentrations, sphingosine does not bind to protein kinase C as shown by its lack of effect on phorbol dibutyrate binding. Only at higher concentrations, in the cytotoxic range, was there a displacement of phorbol dibutyrate from its cellular-binding sites. In contrast to sphingosine, H-7, a known inhibitor of protein kinase C, inhibited the mitogenic response to TPA and the TPA-induced phosphorylation of the 80 kDa cellular substrate of protein kinase C. Our results suggest that sphingosine may play an important role as a positive regulator of cell growth acting in a fundamentally different, protein kinase C-independent pathway.  相似文献   

2.
Sphingosine inhibited protein kinase C activity and phorbol dibutyrate binding. When the mechanism of inhibition of activity and phorbol dibutyrate binding was investigated in vitro using Triton X-100 mixed micellar methods, sphingosine inhibition was subject to surface dilution; 50% inhibition occurred when sphingosine was equimolar with sn-1,2-dioleoylglycerol (diC18:1) or 40% of the phosphatidylserine (PS) present. Sphingosine inhibition was modulated by Ca2+ and by the mole percent of diC18:1 and PS present. Sphingosine was a competitive inhibitor with respect to diC18:1, phorbol dibutyrate, and Ca2+. Increasing levels of PS markedly reduced inhibition by sphingosine. Since protein kinase C activity shows a cooperative dependence on PS, the kinetic analysis of competitive inhibition was only suggestive. Sphingosine inhibited phorbol dibutyrate binding to protein kinase C but did not cause protein kinase C to dissociate from the mixed micelle surface. Sphingosine addition to human platelets blocked thrombin and sn-1,2-dioctanoylglycerol-dependent phosphorylation of the 40-kDa (47 kDa) dalton protein. Moreover, sphingosine was subject to surface dilution in platelets. The mechanism of sphingosine inhibition is discussed in relation to a previously proposed model of protein kinase C activation. The possible physiological role of sphingosine as a negative effector of protein kinase C is suggested and a plausible cycle for its generation is presented. The potential physiological significance of sphingosine inhibition of protein kinase C is further established in accompanying papers on HL-60 cells (Merrill, A. H., Jr., Sereni, A. M., Stevens, V. L., Hannun, Y. A., Bell, R. M., Kinkade, J. M., Jr. (1986) J. Biol. Chem. 261, 12010-12615) and human neutrophils (Wilson, E., Olcott, M. C., Bell, R. M., Merrill, A. H., Jr., and Lambeth, J. D. (1986) J. Biol. Chem. 261, 12616-12623). These results also suggest that sphingosine will be a useful inhibitor for investigating the function of protein kinase C in vitro and in living cells.  相似文献   

3.
The cell-permeable diacylglycerol, sn-1,2-dioctanoylglycerol (DiC8), is shown to mimic the effect of tumor promoting phorbol diesters on epidermal growth factor (EGF) binding and action in intact cells. DiC8 inhibited the binding of [3H]phorbol dibutyrate to A431 cell monolayers indicating that the diacylglycerol interacts with the phorbol diester receptor. At 0.3 microM, DiC8 half-maximally inhibited the high affinity binding of 125I-EGF to A431 human epidermoid carcinoma cells. Scatchard analysis indicated that the inhibition of 125I-EGF binding was very similar to that observed in the presence of 4 beta-phorbol 12 beta-myristate 13 alpha-acetate (PMA). DiC8 also mimicked the action of PMA to increase the phosphorylation state of the EGF receptor in 32P-labeled cells. Phosphoamino acid analysis demonstrated that DiC8 and PMA caused an increase in the level of EGF-receptor phosphoserine and phosphothreonine, whereas EGF caused an increase in the level of phosphoserine, phosphothreonine, and phosphotyrosine. Phosphopeptide mapping of the EGF receptor showed that DiC8 and PMA enhanced the phosphorylation of the same tryptic peptides. DiC8 inhibited the EGF-dependent tyrosine phosphorylation of the EGF receptor in A431 cells in a similar manner to that observed with PMA. In further experiments with quiescent Swiss 3T3 fibroblasts, DiC8 mimicked the ability of PMA to stimulate the incorporation of [methyl-3H]thymidine synergistically with low concentrations of EGF. This result indicates that DiC8 will mimic the long-term effects of PMA to regulate mitogenesis and raises the possibility that it may be active in two stage carcinogenesis. As both DiC8 and PMA stimulate the Ca2+- and phospholipid-dependent protein kinase (C-kinase) in vitro, the results support the hypothesis that the activation of C-kinase is a critical component of phorbol diester action on EGF receptor modulation and cell proliferation.  相似文献   

4.
The role of protein kinase C in activation of the plasma membrane Na+/H+ exchanger was studied in cultured vascular smooth muscle cells. The basic lipid, sphingosine, was used to block enzymatic activity of protein kinase C. Na+/H+ exchange was activated by phorbol 12-myristate 13-acetate (PMA), diacylglycerols, platelet-derived growth factor (PDGF), thrombin, or by osmotically-induced cell shrinkage. Intracellular pH and Na+/H+ exchange activity were measured using the intracellular pH indicator, 2',7'-bis(carboxyethyl)-5(6) carboxyfluorescein. Acting alone, both crude sphingosine and pure, synthetic C18 D-(+)-erythro-sphingosine raised pHi in a dose-dependent manner (from 6.95 +/- 0.02 to 7.19 +/- 0.09 over 10 min for 10 microM sphingosine). This alkalinization was not due to Na+/H+ exchange as it was not altered by t-butylamiloride (50 microM) nor by replacement of the assay medium with a Na(+)-free solution. Sphingosine-induced alkalinization did not require protein kinase C activity, since it was fully intact in protein kinase C-depleted cells. It was also not due to a detergent action of sphingosine on the cell membrane, since both ionic and non-ionic detergents caused cell acidification. Rather, alkalinization induced by sphingosine appeared to be due to cellular uptake of NH3 groups since N-acetylsphingosine showed no alkalinization. After the initial cell alkalinization, cellular uptake of [3H]sphingosine continued slowly for up to 24 h. The ability of PMA or dioctanoylglycerol to activate Na+/H+ exchange fell to 20% of control after 24 h of sphingosine exposure. At all times, C11 and N-acetylsphingosine failed to block PMA-induced activation of the exchanger. Activation of the Na+/H+ exchanger by sucrose, which does not depend on protein kinase C activity, was unaffected by sphingosine. Activation of Na+/H+ exchange by thrombin and PDGF was partially inhibited by 30 and 20%, respectively. These data indicate that both thrombin and PDGF activate Na+/H+ exchange by pathway(s) that are primarily independent of protein kinase C.  相似文献   

5.
Sphingosine is a potent inhibitor of [3H]phorbol dibutyrate binding and protein kinase C activity in vitro and in human platelets (Hannun, Y., Loomis, C., Merrill, A., and Bell, R. (1986) J. Biol. Chem. 261, 12604-12609). Preincubation of platelets with sphingosine resulted in the inhibition of platelet secretion and second phase aggregation in response to ADP, gamma-thrombin, collagen, arachidonic acid, and platelet activating factor. Sphingosine did not affect the initial shape change of platelets or the first phase of aggregation in response to these agonists. Ristocetin-induced platelet agglutination was not affected by sphingosine. Sphingosine inhibition of secondary aggregation (secretion and second phase aggregation) was overcome by phorbol dibutyrate and by the cell-permeable protein kinase C activator, dioctanoylglycerol. Furthermore, platelet secretion and irreversible aggregation were induced by protein kinase C activators in platelets that had been "primed" to undergo initial shape change and first phase aggregation by low concentrations of agonists. These results suggest that protein kinase C activation is a necessary component in the signal transducing pathways that lead to platelet activation. Higher concentrations of agonists, however, induced irreversible aggregation and partial secretion in the presence of sphingosine, suggesting the existence of protein kinase C-independent pathways for platelet activation. These results demonstrate the utility of sphingosine as a pharmacologic tool in probing the role of protein kinase C in signal transduction.  相似文献   

6.
Human neutrophils treated with phorbol 12-myristate 13-acetate (PMA) or dioctanoylglycerol exhibited a large (10-fold), sustained accumulation of the mass of diradylglycerol, beginning 1 min after stimulation and continuing for 30 to 60 min. Phorbol dibutyrate was less potent than PMA in stimulating diradylglycerol accumulation, whereas the 4-alpha analogs of PMA and phorbol dibutyrate were inactive. Submaximal concentrations of PMA (0.5 to 2.5 nM) plus the calcium ionophore, ionomycin (15 to 60 nM), led to synergistic accumulation of diradylglycerols. Chlorpromazine and sphingosine, inhibitors of protein kinase C, blocked PMA-stimulated accumulation of diradylglycerol with IC50 concentrations of 32 and 9 microM, respectively, paralleling their inhibition of PMA-stimulated O2- production. These compounds also inhibited the ionomycin-stimulated accumulation of diradylglycerols. A third protein kinase C inhibitor, H-7, was less effective, inhibiting PMA-stimulated accumulation of diradylglycerol by 25% at 100 microM. Differential sensitivity to alkaline hydrolysis suggests that diradylglycerols that accumulate in response to PMA or ionomycin stimulation are composed of a mixture of two distinct diglyceride species, diacylglycerols and alkylacylglycerols. Whereas diacylglycerol may activate cellular protein kinase C, the importance of the production of alkylacylglycerols is uncertain.  相似文献   

7.
Neutrophils treated with 1,2-dioctanoyl-sn-glycerol (DiC8) are known to release large quantities of superoxide (O2-) and to exhibit an intense phosphorylation of two proteins with molecular masses of approximately 47 and 49 kDa. In this paper, we report that O2- release from guinea pig cells stimulated with a near optimal amount of DiC8 (2.0 microM) is markedly inhibited (greater than or equal to 70%) by antagonists of protein kinase C (i.e. 150 nM staurosporine; 200 microM 1-(5-isoquinolinylsulfonyl)-2-methyl-piperazine (H-7], whereas that from cells stimulated with an optimal amount of DiC8 (7.8 microM) is not (approximately 25% inhibition). However, staurosporine (150 nM) effectively reduced the level of phosphorylation of the 47- and the 49-kDa proteins to that observed in unstimulated cells when either amount of DiC8 (i.e. 2.0 or 7.8 microM) was utilized. Thus, neutrophils stimulated with 7.8 microM DiC8 in the presence of staurosporine release large quantities of O2- without an enhanced phosphorylation of the 47- and the 49-kDa proteins. In contrast, these antagonists of protein kinase C effectively blocked O2- release from neutrophils stimulated with an optimal amount of phorbol 12-myristate 13-acetate (PMA), and the percentage of inhibition was not affected by increasing the concentration of PMA 160-fold. These data show that DiC8 and PMA, both activators of protein kinase C, can have distinct effects on O2- release by neutrophils. Moreover, they suggest that DiC8 (or a metabolite) under certain circumstances may function in a stimulatory pathway for O2- release that is independent of protein kinase C. Differences in the morphology of neutrophils stimulated with PMA and DiC8 are presented. Ancillary data on human neutrophils are also provided.  相似文献   

8.
Sphingosine, sphinganine, and other long-chain (sphingoid) bases inhibit protein kinase C in vitro and block cellular responses to agonists that are thought to act via this enzyme. To gain further insight into the mechanism of this inhibition, a series of long-chain analogues differing in alkyl chain length (11-20 carbon atoms), stereochemistry, and headgroup were examined for (a) inhibition of protein kinase C activity in vitro, (b) the neutrophil respiratory burst in response to phorbol myristate acetate (PMA), (c) the PMA-induced differentiation of HL-60 cells, and (d) the growth of Chinese hamster ovary cells. In every instance, the effects were maximal with the 18-carbon homologues, which are the same length as the predominant naturally occurring long-chain base (sphingosine). The lower potency of the shorter chain homologues was partially due to decreased uptake by cells. Small differences were obtained with the four stereoisomers of sphingosine (i.e., D and L forms of erythro- and threo-sphingosine), with N-methyl derivatives of the different sphingosine homologues, and with simpler alkylamines (e.g., stearylamine). The potency of the different headgroup analogues may be affected by the degree of protonation at the assay pH. The pKa of sphingosine was measured to be 6.7; the pKa varied among the analogues. These findings establish that the major structural features required for inhibition of protein kinase C and cellular processes dependent on this enzyme are the presence of a free amino group and an aliphatic side chain and that other groups have more subtle effects.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
10.
Sphingosine and other protein kinase C inhibitors were tested for their ability to inhibit aldosterone synthesis by bovine adrenal glomerulosa cells. Sphingosine inhibited angiotensin (AII)-stimulated aldosterone synthesis (IC50 of 5 microM). At doses that totally blocked steroidogenesis, sphingosine did not affect protein synthesis or [125I]AII binding to cells. Sphingosine also inhibited dibutyryl cyclic AMP (dbcAMP)-stimulated aldosterone synthesis. Sphingosine inhibited pregnenolone synthesis from cholesterol, but not the conversion of progesterone or 20 alpha-hydroxycholesterol to aldosterone. These results suggest that sphingosine inhibits steroidogenesis at a locus close to that where stimulation occurs by AII and dbcAMP. Other protein kinase C inhibitors were tested. Retinal, 1-(5-isoquinolinesulfonyl)-2-methylpiperazine dihydrochloride (H-7), and staurosporine inhibited aldosterone synthesis stimulated by AII and dbcAMP. Retinal and H-7 also inhibited progesterone conversion to aldosterone, and retinal blocked [125I]AII binding. Staurosporine was more specific, inhibiting AII-stimulated aldosteronogenesis at concentrations which had little effect on conversion of progesterone to aldosterone. Because they inhibited dbcAMP stimulation, none of the inhibitors was sufficiently specific to use as a probe of the role of protein kinase C. The IC50 of sphingosine suggests that this or related products of lipid hydrolysis could act as endogenous regulators of adrenal cell function.  相似文献   

11.
The effects of short-term phorbol ester treatment of CHO cells that stably express 900 fmol of recombinant human serotonin 5-HT1A receptor/mg of protein on coupling to the inhibition of adenylyl cyclase and on phosphorylation of the receptor were studied. Pretreatment of cell monolayers with phorbol 12-myristate 13-acetate (PMA) caused a dose- and time-dependent shift of the half-maximal dose of serotonin (5-HT) required to inhibit membrane adenylyl cyclase (from IC50 approximately 100 nM to approximately 400 nM). This desensitization (shift in IC50) was rapid, occurring with 5 min of pretreatment and being maximal by 10-15 min; it was also dose-dependent, being half-maximal at approximately 300 nM PMA. Desensitization was also induced by sn-dioctanoylglycerol (DiC8) and blocked by the protein kinase C (PKC) inhibitors sphingosine and 1-(5-isoquinolinesulfonyl)-2-methylpiperazine (H-7). In detached permeabilized cells, PMA pretreatment caused a rapid phosphorylation of immunoprecipitated 5-HT1A receptors, with an approximately 3-4-fold increase that was maximal after 15 min and persisted for 90 min. The phosphorylation occurred at a similar dose of PMA as that which induced desensitization (half-maximal at approximately 300 nM, maximal at 500 nM to 1 microM), could be reproduced by pretreatment with the PKC activators DiC8 or phorbol 12,13-dibutyrate (PDBu), and could be blocked by the PKC inhibitors sphingosine or H-7. The stoichiometry of the phosphorylation was approximately 2 mol of [32P]ATP/mol of receptor, suggesting the involvement at least two of three putative PKC sites within the 5-HT1A receptor. The close concordance between the PKC-induced desensitization and phosphorylation suggests a potential causative link between these two effects of PKC on the human 5-HT1A receptor.  相似文献   

12.
Sphingosine inhibited [3H]methylhistidine-thyrotropin-releasing hormone (MeTRH) binding to intact GH3 cells and to GH3 membranes. This inhibition was dependent on the concentration of sphingosine and on the ratio of sphingosine to cell number (or membrane protein) and was partly reversed by washing. In intact cells, the IC50 was 63 microM (1.8 X 10(6) cells/ml; 2 nM MeTRH), and 100 microM sphingosine was found, by Scatchard analysis, to increase the apparent dissociation constant (Kd) from 1.1 +/- 0.3 to 6.5 +/- 2.3 nM and to decrease the maximal binding capacity (Bmax) to 41 +/- 9.5% of control. Kinetic analysis showed that the major effect of sphingosine on Kd was due to a marked decrease in the apparent association rate constant for MeTRH from 2.5 +/- 0.4 X 10(5) M-1 s-1 to 0.10 +/- 0.015 X 10(5) M-1 s-1. At 100 microM, sterylamine was as effective as sphingosine in inhibiting MeTRH binding, whereas sphinganine was less effective, and psychosine and steroylsphingosine were without effect. The following observations show that sphingosine inhibition of MeTRH binding did not involve protein kinase C. The IC50 for sphingosine inhibition of MeTRH binding was the same in GH3 cells that had been incubated with 1 microM phorbol 12-myristate 13-acetate for 16 h, to "down-regulate" protein kinase C, as in control cells. Sphingosine inhibited MeTRH binding to membranes isolated from GH3 cells that contain very little protein kinase C activity. In GH3 membranes, 100 microM sphingosine increased the Kd for MeTRH from 3.4 +/- 0.1 to 13 +/- 3.1 nM but did not significantly decrease Bmax (12 +/- 5.0% of control, p greater than 0.05). And, 1-(5-isoquinolinesulfonyl)-2-methylpiperazine dihydrochloride, an inhibitor of protein kinase C, failed to decrease MeTRH binding to intact GH3 cells or to membranes, and did not interfere with the effects of sphingosine. These data show that sphingosine and its analogs have complex actions to inhibit MeTRH binding to GH3 cells, at least some of which are independent of protein kinase C, and thereby demonstrate that sphingolipids cannot be used as specific inhibitors of protein kinase C.  相似文献   

13.
Human neutrophils synthesize platelet-activating factor (PAF) and leukotriene B4 (LTB4) when stimulated with the Ca2+ ionophore A23187. These processes are enhanced to a variable extent by phorbol 12-myristate 13-acetate (PMA), a direct activator of protein kinase C. The long chain amines sphingosine, stearylamine (Hannun, Y.A., Loomis, C.R., Merrill, A.H., Jr., and Bell, R.M. (1986) J. Biol. Chem. 261, 12604-12609), and palmitoylcarnitine competitively inhibit activation of purified protein kinase C in vitro and inhibit protein kinase C-mediated activation of the respiratory burst in human neutrophils (Wilson, E., Olcott, M.C., Bell, R.M., Merrill, A.H., Jr., and Lambeth, J.D. (1986) J. Biol. Chem. 261, 12616-12623). These amines were found to inhibit A23187-induced PAF and LTB4 synthesis. Inhibition of PAF and LTB4 synthesis occurred in parallel; half-maximal inhibition by sphingosine occurred at 7 microM, with complete inhibition at 15 microM. PMA by itself did not induce the synthesis of PAF or LTB4, although it did enhance PAF and LTB4 synthesis at suboptimal concentrations of A23187. PMA reversed long chain amine inhibition of PAF and LTB4 accumulation. Reversal of the inhibition of PAF and LTB4 accumulation occurred in parallel, was concentration-dependent, and was complete by approximately 3 x 10(-8) M PMA. The inactive 4 alpha-phorbol didecanoate ester did not reverse inhibition at these concentrations. Sphingosine completely prevented the A23187-induced release of [3H]arachidonate and its various metabolites from [3H]arachidonate-labeled cells. PMA, but not 4 alpha-phorbol didecanoate, restored arachidonate release and its metabolism. Therefore, while activation of protein kinase C is not sufficient to induce PAF and LTB4 synthesis, its action appears to be required to couple a rise in intracellular Ca2+ to their synthesis. This coupling occurs at the level of the initial reaction in the production of lipid mediators, a phospholipase A2-like activity that mobilizes the two substrates 1-O-alkyl-sn-glycero-3-phosphocholine and arachidonic acid from complex lipids.  相似文献   

14.
We sought to investigate the mechanisms by which the calcium ionophore A23187 triggers arachidonic acid release in bovine pulmonary endothelial cells and to test the hypothesis that protein kinase C is involved in this process. Our results indicate that the mechanism by which A23187 increases phospholipase A2 activity and arachidonic acid release in bovine pulmonary arterial endothelial cells depends upon the concentration studied. At concentrations of 1 microM and 2.5 microM, A23187 increases phospholipase A2 activity and arachidonic acid release without stimulating protein kinase C. At concentrations of 5-12.5 microM, A23187 increases arachidonic acid release and phospholipase A2 activity in conjunction with a dose-dependent activation of membrane-bound protein kinase C. To test the hypothesis that these doses of A23187 increase phospholipase A2 activity by stimulating protein kinase C, we studied the effect of prior treatment with the protein kinase C inhibitor sphingosine. Sphingosine inhibits the increase in phospholipase A2 activity and arachidonic acid release caused by A23187 over the range 5-12.5 microM. To investigate further the potential role of protein kinase C, we studied the effects of the inactive phorbol ester 4 alpha-phorbol 12 beta-myristate 13 alpha-acetate (4 alpha-PMA) and an active phorbol ester 4 beta-phorbol 12 beta-myristate 13 alpha-acetate (4 beta PMA). Neither 4 alpha-PMA nor 4 beta-PMA affected basal arachidonic acid release. 4 alpha-PMA also did not augment the effects of A23187. In contrast, 4 beta-PMA significantly augments the increase in phospholipase A2 activity and arachidonic acid release caused by lower doses of A23187. Under these conditions, sphingosine completely inhibits the stimulatory effects of 4 beta-PMA on protein kinase C translocation, phospholipase A2 and arachidonic acid release. Thus, at low doses (1 microM and 2.5 microM) A23187 increases phospholipase A2 activity and arachidonic acid release by a mechanism that does not involve protein kinase C. At these A23187 doses, activating membrane-bound protein kinase C with 4 beta-PMA causes a synergistic increase in phospholipase A2 activity and arachidonic acid release. At higher doses (5-12.5 microM), A23187 acts in large part by stimulating protein kinase C translocation. Overall, our results indicate that activating membrane-bound protein kinase C by itself is an insufficient stimulus to increase phospholipase A2 activity and arachidonic acid release in pulmonary endothelial cells, but activating protein kinase C can substantially augment the increase in phospholipase A2 activity and arachidonic acid caused by a small increase in intracellular calcium.  相似文献   

15.
Bradykinin is known to activate phospholipase D in PC12 cells. Because bradykinin may also activate protein kinase C in these cells, the possible role of this kinase in mediating the action of bradykinin was investigated. Phospholipase D activity in PC12 cells was assayed by measuring the formation of [3H]phosphatidylethanol in cells prelabeled with [3H]palmitic acid and incubated in the presence of ethanol. The phorbol ester phorbol dibutyrate mimicked the effect of bradykinin on [3H]phosphatidylethanol formation. The protein kinase C inhibitor staurosporine (1 microM) significantly attenuated the effect of phorbol dibutyrate (35-70%) but did not block bradykinin-stimulated [3H]phosphatidylethanol formation. In addition, the effect of phorbol dibutyrate was additive with that of bradykinin. Prolonged treatment of PC12 cells with phorbol dibutyrate (24 h), which depletes cells of protein kinase C, greatly attenuated bradykinin-stimulated [3H]phosphatidylethanol accumulation in intact cells. This treatment caused a 55% decrease in both fluoride-stimulated [3H]phosphatidylethanol production in the intact cell and phospholipase D activity as assessed by an in vitro assay using an exogenous substrate. Therefore, the effect of prolonged phorbol dibutyrate pretreatment on bradykinin-stimulated [3H]phosphatidylethanol production could not be attributed exclusively to the depletion of protein kinase C. Thus, although the data with phorbol ester suggest that activation of protein kinase C leads to an increase in phospholipase D activity, this kinase probably does not play a role in mediating the effect of bradykinin. Finally, although pretreatment with phorbol dibutyrate completely blocked bradykinin-stimulated [3H]phosphatidylethanol production in the intact cell, it only partially (approximately 50%) inhibited bradykinin-stimulated [3H]diacylglycerol formation.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
Tissue factor is a lipoprotein, expressed on the surface of cells, which binds coagulation Factor VII or VIIa, leading to activation of Factors X and IX with subsequent fibrin generation. Cellular tissue factor activity is important in pathophysiologic processes such as inflammation and disseminated intravascular coagulation. In this study, the long-chain base sphingosine inhibited coagulation initiated by lipopolysaccharide-stimulated intact human monocytes. Sphingosine (5-100 microM) also profoundly inhibited thromboplastin-initiated coagulation (greater than 90% decrease in thromboplastin activity). This inhibition was dose- and time-dependent. Sphingosine inhibited neither the intrinsic pathway of coagulation nor thrombin generation of fibrin. The sphingosine analogues sphingomyelin, ceramide, or N-acetylsphingosine did not affect thromboplastin activity, suggesting that the polar head of sphingosine was necessary for interaction of the molecule with the coagulation system. Investigation of the biochemical mechanism revealed that sphingosine (5-50 microM), but neither sphingomyelin nor ceramide, inhibited specific binding of radiolabeled Factor VII to lipopolysaccharide-stimulated intact monocytes. The results suggest that sphingosine may regulate monocyte tissue factor-initiated coagulation by modulating Factor VII binding to tissue factor. Sphingosine may represent a new class of inhibitors of hemostasis.  相似文献   

17.
The regulation of protein phosphorylation by sphingosine in A431 human epidermoid carcinoma cells was examined. Sphingosine is a competitive inhibitor of phorbol ester binding to protein kinase C (Ca2+/phospholipid-dependent enzyme) and potently inhibits phosphotransferase activity in vitro. Addition of sphingosine to intact A431 cells caused an inhibition of the phorbol ester-stimulated phosphorylation of two protein kinase C substrates, epidermal growth factor (EGF) receptor threonine 654 and transferrin receptor serine 24. We conclude that sphingosine inhibits the activity of protein kinase C in intact A431 cells. However, further experiments demonstrated that sphingosine-treatment of A431 cells resulted in the regulation of the EGF receptor by a mechanism that was independent of protein kinase C. First, sphingosine caused an increase in the threonine phosphorylation of the EGF receptor on a unique tryptic peptide. Second, sphingosine caused an increase in the affinity of the EGF receptor in A431 and in Chinese hamster ovary cells expressing wild-type (Thr654) and mutated (Ala654) EGF receptors. Sphingosine was also observed to cause an increase in the number of EGF-binding sites expressed at the surface of A431 cells. Examination of the time course of sphingosine action demonstrated that the effects on EGF binding were rapid (maximal at 2 mins) and were observed prior to the stimulation of receptor phosphorylation (maximal at 20 mins). We conclude that sphingosine is a potently bioactive molecule that modulates cellular functions by: 1) inhibiting protein kinase C; 2) stimulating a protein kinase C-independent pathway of protein phosphorylation; and 3) increasing the affinity and number of cell surface EGF receptors.  相似文献   

18.
19.
The neutrophil oxidative burst is characterized by increased cellular O2 consumption due to the activation of a membrane-associated superoxide-generating NADPH-oxidase. The response is triggered by a variety of stimuli, including opsonized zymosan, formylmethionylleucinephenylalanine (FMLP), arachidonate, short-chain diacylglycerols, and phorbol myristate acetate (PMA). We herein demonstrate that incubation of cells with sphinganine or sphingosine blocks or reverses activation by these agonists. The inhibition is reversible, does not affect cell viability, and does not affect another complex cell function, phagocytosis. Inhibitory concentrations of sphinganine did not significantly affect cytoplasmic calcium levels or FMLP-generated calcium transients. Structural requirements for inhibition of the oxidative burst include a long aliphatic chain and an amino-containing head-group, and there is modest specificity for the native (erythro) isomer of sphinganine. Inhibition involves stimulus-induced activation mechanisms rather than a direct effect on the NADPH oxidase, since sphinganine did not inhibit NADPH-dependent superoxide generation in isolated membranes containing the active enzyme. Activation by FMLP, diacylglycerol, PMA, opsonized zymosan, and arachidonate was blocked by the same concentrations of sphinganine, indicating that these agonists share a common inhibited step. Three lines of evidence indicate that this step involves protein kinase C. First, in a micelle system and in platelets, long-chain bases are inhibitors of this enzyme (Hannun, Y., Loomis, C., Merrill, A., and Bell, R. M. (1986) J. Biol. Chem. 261, 12604-12609). Second, sphinganine blocks PMA-stimulated incorporation of 32PO4 into neutrophil proteins. Third, sphinganine inhibits the binding of [3H]phorbol dibutyrate to its cellular receptor, known to be protein kinase C. We suggest that long-chain bases function as physiologic modulators of cellular regulatory pathways involving protein kinase C.  相似文献   

20.
The combination of phorbol 12-myristate 13-acetate (PMA) and ionomycin produces a dramatic increase in the incorporation of [2-3H]mannose into Glc3Man9GlcNAc2-P-P-dolichol and glycoprotein, and the induction of RNA and DNA synthesis in murine splenic B lymphocytes (B cells). The kinetics of the induction processes and the concentrations of PMA and ionomycin required for the optimal response have been defined. While the levels of induction of RNA and DNA synthesis by PMA + ionomycin were similar to the mitogenic response to bacterial lipopolysaccharide, activation by PMA and the calcium ionophore resulted in a threefold higher stimulation in dolichol-linked oligosaccharide biosynthesis and protein N-glycosylation. These results indicate that all signalling mechanisms that trigger RNA and DNA synthesis may not be sufficient to produce maximal induction of the N-glycosylation apparatus. 1-(5-Isoquinolinesulfonyl)-2-methylpiperazine (H-7), a potent protein kinase C inhibitor, prevented the induction of protein N-glycosylation activity (IC50 = 11 microM), as well as RNA (IC50 = 18 microM) and DNA synthesis (IC50 = 12 microM), two common indices of B cell activation. N-[2-(Methylamino)ethyl]-5-isoquinolinesulfonamide (H-8) also inhibited the induction of oligosaccharide-lipid intermediate, glycoprotein, RNA, and DNA synthesis, but required higher concentrations than H-7 for 50% inhibition. N-(2-Guanidinoethyl)-5-isoquinolinesulfonamide (HA1004), a potent inhibitor of cyclic nucleotide-dependent protein kinases, had little effect on the activation of the B cell metabolic processes. The H-7-sensitive reactions involved in the induction of RNA and DNA synthesis occurred within 4 h, but induction of lipid intermediate and glycoprotein biosynthesis remained sensitive to H-7 for 10 h after exposure to PMA and ionomycin. Direct in vitro assays in the presence of 0.6% Brij 58 reveal that a cytosolic, phospholipid-dependent protein kinase activity is translocated to a membrane site(s) after treatment with PMA and ionomycin, and the translocated protein kinase is sensitive to H-7. The relative order of potency of the protein kinase inhibitors on the metabolic processes strongly supports the hypothesis that protein kinase C, acting synergistically with Ca2+ mobilization, plays a key regulatory role in the early stages of B cell activation. The synthesis of oligosaccharide-lipid intermediates and protein N-glycosylation are also shown to be induced in B cells activated by PMA + ionomycin.  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