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 343 毫秒
1.
The mechanism by which GnRH increases sperm-zona pellucida binding in humans was investigated in this study. We tested whether GnRH increases sperm-zona binding in Ca(2+)-free medium and in the presence of Ca(2+) channel antagonists. We also examined the GnRH effect on the intracellular free Ca(2+) concentration ([Ca(2+)](i)). Sperm treatment with GnRH increased sperm-zona binding 300% but only when Ca(2+) was present in the medium. In Ca(2+)-free medium or in the presence of 400 nM nifedipine, 80 microM diltiazem, or 50 microM verapamil, GnRH did not influence sperm-zona binding. GnRH increased the [Ca(2+)](i) in the sperm in a dose-dependent manner. The maximum effect was reached with 75 nM GnRH. The GnRH-induced increase in [Ca(2+)](i) was fast and transient, from a basal [Ca(2+)](i) of 413 +/- 22 nM to a peak value of 797 +/- 24 nM. The GnRH-induced increase in [Ca(2+)](i) was entirely due to a Ca(2+) influx from the extracellular medium because the increase in [Ca(2+)](i) was blocked by the Ca(2+) chelator EGTA and by the Ca(2+) channel antagonists nifedipine and diltiazem. These antagonists, however, were not able to inhibit the progesterone-activated Ca(2+) influx. On the contrary, T-type calcium channel antagonists pimozide and mibefradil did not affect GnRH-activated Ca(2+) influx but inhibited the progesterone-activated Ca(2+) influx. Finally, the GnRH-induced Ca(2+) influx was blocked by two specific GnRH antagonists, Ac-D-Nal(1)-Cl-D-Phe(2)-3-Pyr-D-Ala(3)-Arg(5)-D-Glu(AA)(6)-GnRH and Ac-(3,4)-dehydro-Pro(1),-p-fluoro-D-Phe(2), D-Trp(3,6)-GnRH. These results suggest that GnRH increases sperm-zona binding via an elevation of [Ca(2+)](i) through T-type, voltage-operated calcium channels.  相似文献   

2.
Capsazepine is thought to be a selective antagonist of vanilloid type 1 receptors; however, its other in vitro effect on different cell types is unclear. In human MG63 osteosarcoma cells, the effect of capsazepine on intracellular Ca(2+) concentrations ([Ca(2+)](i)) and cytotoxicity was explored by using fura-2 and tetrazolium, respectively. Capsazepine caused a rapid rise in [Ca(2+)](i) in a concentration-dependent manner with an EC(50) value of 100 microM. Capsazepine-induced [Ca(2+)](i) rise was partly reduced by removal of extracellular Ca(2+), suggesting that the capsazepine-induced [Ca(2+)](i) rise was composed of extracellular Ca(2+) influx and intracellular Ca(2+). In Ca(2+)-free medium, thapsigargin, an inhibitor of the endoplasmic reticulum Ca(2+)-ATPase, caused a monophasic [Ca(2+)](i) rise, after which the increasing effect of capsazepine on [Ca(2+)](i) was inhibited by 75%. Conversely, pretreatment with capsazepine to deplete intracellular Ca(2+) stores totally prevented thapsigargin from releasing more Ca(2+). U73122, an inhibitor of phospholipase C, abolished histamine (an inositol 1,4,5-trisphosphate-dependent Ca(2+) mobilizer)-induced, but not capsazepine-induced, [Ca(2+)](i) rise. Overnight treatment with 1-100 microM capsazepine inhibited cell proliferation in a concentration-dependent manner. These findings suggest that in human MG63 osteosarcoma cells, capsazepine increases [Ca(2+)](i) by stimulating extracellular Ca(2+) influx and also by causing intracellular Ca(2+) release from the endoplasmic reticulum via a phospholiase C-independent manner. Capsazepine may be mildly cytotoxic.  相似文献   

3.
Lin MC  Jan CR 《Life sciences》2002,71(9):1071-1079
The effect of the anti-anginal drug fendiline on intracellular free Ca(2+) levels ([Ca(2+)](i)) in a rabbit corneal epithelial cell line (SIRC) was explored using fura-2 as a fluorescent Ca(2+) indicator. At a concentration above 1 microM, fendiline increased [Ca(2+)](i) in a concentration-dependent manner with an EC(50) value of 7 microM. The [Ca(2+)](i) response consisted of an immediate rise and an elevated phase. Extracellular Ca(2+) removal decreased half of the [Ca(2+)](i )signal. Fendiline induced quench of fura-2 fluorescence by Mn(2+) (50 microM), suggesting the presence of Ca(2+) influx across the plasma membrane. This Ca(2+) influx was abolished by La(3+) (50 microM), but was insensitive to dihydropyridines, verapamil and diltiazem. Fendiline (10 microM)-induced store Ca(2+) release was largely reduced by pretreatment with thapsigargin (1 microM) (an endoplasmic reticulum Ca(2+) pump inhibitor) to deplete the endoplasmic reticulum Ca(2+). Conversely, pretreatment with 10 microM fendiline abolished thapsigargin-induced Ca(2+) release. Fendiline (10 microM)-induced Ca(2+) release was not altered by inhibiting phospholipase C with 2 microM 1-(6-((17beta-3-methoxyestra-1,3,5(10)-trien-17-yl)amino)hexyl)-1H-pyrrole-2,5-dione (U73122). Cumulatively, this study shows that fendiline induced concentration-dependent [Ca(2+)](i )increases in corneal epithelial cells by releasing the endoplasmic reticulum Ca(2+) in a phospholipase C-independent manner, and by causing Ca(2+) influx.  相似文献   

4.
Chao YY  Jan CR 《Life sciences》2004,74(7):923-933
In canine renal tubular cells, the effect of Y-24180, a presumed specific platelet activating factor (PAF) receptor antagonist, on intracellular Ca(2+) concentration ([Ca(2+)](i)) was measured by using fura-2 as a Ca(2+)-sensitive fluorescent probe. Y-24180 (0.1-10 microM) caused a rapid and sustained [Ca(2+)](i) rise in a concentration-dependent manner. The [Ca(2+)](i) rise was prevented by 30% by removal of extracellular Ca(2+), but was not changed by dihydropyridines, verapamil and diltiazem. Y-24180-induced Ca(2+) influx was confirmed by Mn(2+)-influx induced quench of fura-2 fluorescence. In Ca(2+)-free medium, thapsigargin, an inhibitor of the endoplasmic reticulum Ca(2+)-ATPase, caused a monophasic [Ca(2+)](i) rise, after which the increasing effect of 5 microM Y-24180 on [Ca(2+)](i) was abolished; conversely, depletion of Ca(2+) stores with 5 microM Y-24180 abolished thapsigargin-induced [Ca(2+)](i) rise. U73122, an inhibitor of phoispholipase C, inhibited ATP-, but not Y-24180-induced [Ca(2+)](i) rise. Overnight treatment with Y-24180 did not alter cell proliferation rate. Collectively, these results suggest that Y-24180 acts as a potent, but not cytotoxic, Ca(2+) mobilizer in renal tubular cells, by stimulating both extracellular Ca(2+) influx and intracellular Ca(2+) release. Since alterations in Ca(2+) movement may interfere many cellular signaling processes unrelated to modulation of PAF receptors, caution must be applied in using this chemical as a selective PAF receptor antagonist.  相似文献   

5.
In response to a variety of stimuli, neutrophils release large amount of reactive oxygen species (ROS) generated by NADPH oxidase. This process known as the respiratory burst is dependent on cytosolic free calcium concentration ([Ca(2+)](i)). Proinflammatory cytokines such as interleukin-8 (IL-8) may modulate ROS generation through a priming phenomenon. The aim of this study was to determine the effect of human IL-8 on ROS production in neutrophil-like dimethylsulfoxide-differentiated HL-60 cells (not equalHL-60 cells) and further to examine the role of Ca(2+) mobilization during the priming. IL-8 at 10 nM induced no ROS production but a [Ca(2+)](i) rise (254 +/- 36 nM). IL-8 induced a strongly enhanced (2 fold) ROS release during stimulation with 1 microM of N-formyl-L-methionyl-L-leucyl-L-phenylalanine (fMLF). This potentiation of ROS production is dependent of extracellular Ca(2+) (17.0+/-4.5 arbitrary units (A.U.) in the absence of Ca(2+) versus 56.6 +/- 3.9 A.U. in the presence of 1.25 mM of Ca(2+)). Also, IL-8 enhanced fMLF-stimulated increase in [Ca(2+)](i) (375 +/- 35 versus 245 +/- 21 nM, 0.1 microM of fMLF). IL-8 had no effect on not equalHL-60 cells in response to 1 microM of thapsigargin (472 +/- 66 versus 470 +/- 60 nM). In conclusion, Ca(2+) influx is necessary for a full induction of neutrophil priming by IL-8.  相似文献   

6.
Transient influx of Ca(2+) constitutes an early element of signaling cascades triggering pathogen defense responses in plant cells. Treatment with the Phytophthora sojae-derived oligopeptide elicitor, Pep-13, of parsley cells stably expressing apoaequorin revealed a rapid increase in cytoplasmic free calcium ([Ca(2+)](cyt)), which peaked at approximately 1 microM and subsequently declined to sustained values of 300 nM. Activation of this biphasic [Ca(2+)](cyt) signature was achieved by elicitor concentrations sufficient to stimulate Ca(2+) influx across the plasma membrane, oxidative burst, and phytoalexin production. Sustained concentrations of [Ca(2+)](cyt) but not the rapidly induced [Ca(2+)](cyt) transient peak are required for activation of defense-associated responses. Modulation by pharmacological effectors of Ca(2+) influx across the plasma membrane or of Ca(2+) release from internal stores suggests that the elicitor-induced sustained increase of [Ca(2+)](cyt) predominantly results from the influx of extracellular Ca(2+). Identical structural features of Pep-13 were found to be essential for receptor binding, increases in [Ca(2+)](cyt), and activation of defense-associated responses. Thus, a receptor-mediated increase in [Ca(2+)](cyt) is causally involved in signaling the activation of pathogen defense in parsley.  相似文献   

7.
Wang GJ  Lin LC  Chen CF  Cheng JS  Lo YK  Chou KJ  Lee KC  Liu CP  Wu YY  Su W  Chen WC  Jan CR 《Life sciences》2002,71(9):1081-1090
The effects of timosaponin A-III (TA-III), from Rhizoma Anemarrhenae, on Ca(2+) mobilization in vascular endothelial cells and smooth muscle cells and on vascular tension have been explored. TA-III increased intracellular Ca(2+) concentrations ([Ca(2+)](i)) in endothelials cells at a concentration larger than 5 microM with an EC(50) of 15 microM, and increased [Ca(2+)](i) in smooth muscle cells at a concentration larger than 1 microM with an EC(50) of 8 microM. Within 5 min, the [Ca(2+)](i) signal was composed of a gradual rise, and the speed of rising depended on the concentration of TA-III. The [Ca(2+)](i) signal was abolished by removing extracellular Ca(2+) and was recovered after reintroduction of Ca(2+). The TA-III-induced [Ca(2+)](i) increases in smooth muscle cells were partly inhibited by 10 microM nifedipine or 50 microM La(3+), but was insensitive to 10 microM verapamil and diltiazem. TA-III (10-100 microM) inhibited 0.3 microM phenylephrine-induced vascular contraction, which was abolished by pretreatment with 100 microM N(omega)-nitro-L-arginine (L-NNA) or by denuding the aorta. TA-III also increased [Ca(2+)](i) in renal tubular cells with an EC(50) of 8 microM. Collectively, the results show for the first time that TA-III causes [Ca(2+)](i) increases in the vascular system. TA-III acted by causing Ca(2+) influx without releasing intracellular Ca(2+). TA-III induced relaxation of phenylephrine-induced vascular contraction via inducing release of nitric oxide from endothelial cells.  相似文献   

8.
Histamine, through H(2) receptors, triggers a prominent rise in intracellular free Ca(2+) concentration ([Ca(2+)](i)) in addition to an elevation of cAMP level in HL-60 promyelocytes. Here we show that the histamine-induced [Ca(2+)](i) rise was due to influx of Ca(2+) from the extracellular space, probably through nonselective cation channels, as incubation of the cells with SKF 96365 abolished the histamine-induced [Ca(2+)](i) rise, Na(+) influx, and membrane depolarization. The Ca(2+) influx was specifically inhibited by pretreatment of the cells with PMA or extracellular ATP with 50% inhibitory concentrations of 0.12 +/- 0.03 nM and 185 +/- 17 microM, respectively. Western blot analysis of protein kinase C (PKC) isoforms revealed that PMA (< or =1 nM) and ATP (300 microM) caused selective translocation of PKC-delta to the particulate/membrane fraction. Costimulation of the cells with histamine and SKF 96365 partially reduced histamine-induced granulocytic differentiation, which was evaluated by looking at the extent of fMet-Leu-Phe-induced [Ca(2+)](i) rise and superoxide generation. In conclusion, nonselective cation channels are opened by stimulation of the H(2) receptor, and the channels are at least in part involved in the induction of histamine-mediated differentiation processes. Both effects of histamine were selectively inhibited probably by the delta isoform of PKC in HL-60 cells.  相似文献   

9.
Endothelin-1 (ET-1) increases intracellular Ca(2+) concentration ([Ca(2+)](i)) in pulmonary arterial smooth muscle cells (PASMCs); however, the mechanisms for Ca(2+) mobilization are not clear. We determined the contributions of extracellular influx and intracellular release to the ET-1-induced Ca(2+) response using Indo 1 fluorescence and electrophysiological techniques. Application of ET-1 (10(-10) to 10(-8) M) to transiently (24-48 h) cultured rat PASMCs caused concentration-dependent increases in [Ca(2+)](i). At 10(-8) M, ET-1 caused a large, transient increase in [Ca(2+)](i) (>1 microM) followed by a sustained elevation in [Ca(2+)](i) (<200 nM). The ET-1-induced increase in [Ca(2+)](i) was attenuated (<80%) by extracellular Ca(2+) removal; by verapamil, a voltage-gated Ca(2+)-channel antagonist; and by ryanodine, an inhibitor of Ca(2+) release from caffeine-sensitive stores. Depleting intracellular stores with thapsigargin abolished the peak in [Ca(2+)](i), but the sustained phase was unaffected. Simultaneously measuring membrane potential and [Ca(2+)](i) indicated that depolarization preceded the rise in [Ca(2+)](i). These results suggest that ET-1 initiates depolarization in PASMCs, leading to Ca(2+) influx through voltage-gated Ca(2+) channels and Ca(2+) release from ryanodine- and inositol 1,4,5-trisphosphate-sensitive stores.  相似文献   

10.
Melittin, a peptide from bee venom, is thought to be a phospholipase A(2) activator and Ca(2+) influx inducer that can evoke cell death in different cell types. However, the effect of melittin on cytosolic free Ca(2+) concentration ([Ca(2+)](i)) and viability has not been explored in human osteoblast-like cells. This study examined whether melittin altered [Ca(2+)](i) and killed cells in MG63 human osteosarcoma cells. [Ca(2+)](i) changes and cell viability were measured by using the fluorescent dyes fura-2 and WST-1, respectively. Melittin at concentrations above 0.075 microM increased [Ca(2+)](i) in a concentration-dependent manner. The Ca(2+) signal was abolished by removing extracellular Ca(2+). Melittin-induced Ca(2+) entry was confirmed by Mn(2+) quenching of fura-2 fluorescence at 360 nm excitation wavelength which was Ca(2+)-insensitive. The melittin-induced Ca(2+) influx was unchanged by modulation of protein kinase-C activity with phorbol 12-myristate 13-acetate (PMA) and GF 109203X, or inhibition of phospholipase A(2) with AACOCF(3) and aristolochic acid; but was substantially inhibited by blocking L-type Ca(2+) channels. At concentrations of 0.5 microM and 1 microM, melittin killed 33% and 45% of cells, respectively, via inducing apoptosis. Lower concentrations of melittin failed to kill cells. The cytotoxic effect of 1 microM melittin was completely reversed by pre-chelating cytosolic Ca(2+) with BAPTA. Taken together, these data showed that in MG63 cells, melittin induced a [Ca(2+)](i) increase by causing Ca(2+) entry through L-type Ca(2+) channels in a manner independent of protein kinase-C and phospholipase A(2) activity; and this [Ca(2+)](i) increase subsequently caused apoptosis.  相似文献   

11.
The regulation of cytosolic Ca(2+) homeostasis is essential for cells, including vascular smooth muscle cells. Arterial tone, which underlies the maintenance of peripheral resistance in the circulation, is a major contributor to the control of blood pressure. Confocal microscopy was employed to study the alteration in intracellular calcium ion concentration ([Ca(2+)](i)) in arterioles (external diameters <100 microm) with respect to selected modifying reagents. 5-Hydroxytryptamine (1 microM), ATP (10 microM), and endothelin 1-3 (5 nM) elicited an increase in [Ca(2+)](i) in most arteriole smooth muscle cells. The [Ca(2+)](i) increase sometimes propagated in an intercellular manner. When noradrenaline (10 microM) was used as a stimulant, [Ca(2+)](i) increase was observed only in a portion of the smooth muscle cells. It was also noted that the reaction of these cells with respect to ATP is different between testis and brain arterioles; the [Ca(2+)](i) increase in testicular arterioles is dependent on Ca(2+) influx from extracellular space, whereas in cerebral arterioles it plays a role in both the influx of extracellular Ca(2+) and the release of Ca(2+) from intracellular stores (i.e., sarco/endoplasmic reticulum). These results indicate that arterioles in different tissues may differ greatly in their responses. Real-time confocal microscopy was found to be a useful tool for investigating the structural and functional changes in living tissues.  相似文献   

12.
Bleb formation is an early event of cellular damage observed in a variety of cell types upon hypoxia. Although we previously found that the [Ca(2+)](i) rise before bleb formation only at the same loci of HUVECs upon hypoxia (localized [Ca(2+)](i) rise), the mode of the [Ca(2+)](i) rise remains ill-defined. In order to clarify the mechanisms causing the localized [Ca(2+)](i) rise in hypoxia challenged HUVECs, we studied the effects of several Ca(2+) channel blockers or a Ca(2+) chelator, EGTA, which reduces extracellular Ca(2+) concentration on the hypoxia-induced localized [Ca(2+)](i) rise and bleb formation by employing a confocal laser scanning microscopy (CLSM). After the initiation of hypoxia, [Ca(2+)](i) rose gradually in a localized fashion up to 15 min, which was associated with bleb formation at the same loci. The maximal [Ca(2+)](i) rise was 435 +/- 84 nM at the loci of bleb formation. Ca(2+) channel blockers including Ni(2+) (non-specific, 1 mM), nifedipine (L type, 10 microM), nicardipine (L + T type, 10 microM), and cilnidipine (L + N type, 10 microM) did not inhibit either the localized [Ca(2+)](i) rise or bleb formation. Although both the localized [Ca(2+)](i) rise and bleb formation were inhibited by lowering extracellular Ca(2+) concentration below 100 nM, a diffuse [Ca(2+)](i) rise through the cytoplasm remained without bleb formation, which was inhibited by a phospholipase C (PLC) inhibitor, U73122. In conclusion, hypoxia causes both the Ca(2+) mobilization and the Ca(2+) influx in HUVECs and the Ca(2+) influx through unknown Ca(2+) channels is responsible for the localized [Ca(2+)](i) rise integral to bleb formation.  相似文献   

13.
BACKGROUND: The effect of bradykinin on intracellular free Ca(2+) levels ([Ca(2+)](i)) in MG63 human osteosarcoma cells was explored using fura-2 as a Ca(2+) dye. METHODS/RESULTS: Bradykinin (0.1 nM-1 microM) increased [Ca(2+)](i) in a concentration-dependent manner with an EC(50) value of 0.5 nM. The [Ca(2+)](i) signal comprised an initial peak and a fast decay which returned to baseline in 2 min. Extracellular Ca(2+) removal inhibited the peak [Ca(2+)](i )signals by 35 +/- 3%. Bradykinin (1 nM) failed to increase [Ca(2+)](i) in the absence of extracellular Ca(2+ )after cells were pretreated with thapsigargin (an endoplasmic reticulum Ca(2+) pump inhibitor; 1 microM). Bradykinin (1 nM)-induced intracellular Ca(2+) release was nearly abolished by inhibiting phospholipase C with 2 microM 1-(6-((17 beta-3-methoxyestra-1,3,5(10)-trien-17-yl)amino)hexyl)-1H-pyrrole-2,5-dione (U73122). The [Ca(2+)](i )increase induced by 1 nM bradykinin in Ca(2+)- free medium was abolished by 1 nM HOE 140 (a B2 bradykinin receptor antagonist) but was not altered by 100 nM Des-Arg-HOE 140 (a B1 bradykinin receptor antagonist). Pretreatment with 1 pM pertussis toxin for 5 h in Ca(2+) medium inhibited 30 +/- 3% of 1 nM bradykinin-induced peak [Ca(2+)](i) increase. CONCLUSIONS: Together, this study shows that bradykinin induced [Ca(2+)](i) increases in a concentration-dependent manner, by stimulating B2 bradykinin receptors leading to mobilization of Ca(2+) from the thapsigargin-sensitive stores in a manner dependent on inositol-1,4,5-trisphosphate, and also by inducing extracellular Ca(2+) influx. The bradykinin response was partly coupled to a pertussis toxin-sensitive G protein pathway.  相似文献   

14.
Zhang W  Segura BJ  Mulholland MW 《Peptides》2002,23(10):1793-1801
The responsiveness of cultured myenteric neurons to cholecystokinin (CCK-8) was examined using fura-2-based digital microfluorimetric measurement of intracellular calcium ([Ca(2+)](i)). CCK-8 (10(-10)-10(-6)M) evoked concentration-dependent increases in percentage of neurons responding (8-52%) and delta[Ca(2+)](i) (76-169 nM). Gastrin (1 microM) also induced an increase in [Ca(2+)](i) in 29+/-6% of neurons (delta[Ca(2+)](i): 71+/-3 nM). L-364,718, an antagonist for the CCK-A receptor, blocked [Ca(2+)](i) response to CCK-8. Removal of extracellular calcium eliminated CCK-induced [Ca(2+)](i) increments, as did the addition of the calcium channel inhibitors nickel (1mM) and lanthanum (5mM). Nifedipine (1-50 microM) dose-dependently attenuated CCK-caused [Ca(2+)](i) responses. CCK evokes [Ca(2+)](i) signaling in myenteric neurons by the influx of extracellular calcium, likely through L-type calcium channels.  相似文献   

15.
Ca(2+) influx triggered by depletion of sarcoplasmic reticulum (SR) Ca(2+) stores [mediated via store-operated Ca(2+) channels (SOCC)] was characterized in enzymatically dissociated porcine airway smooth muscle (ASM) cells. When SR Ca(2+) was depleted by either 5 microM cyclopiazonic acid or 5 mM caffeine in the absence of extracellular Ca(2+), subsequent introduction of extracellular Ca(2+) further elevated [Ca(2+)](i). SOCC was insensitive to 1 microM nifedipine- or KCl-induced changes in membrane potential. However, preexposure of cells to 100 nM-1 mM La(3+) or Ni(2+) inhibited SOCC. Exposure to ACh increased Ca(2+) influx both in the presence and absence of a depleted SR. Inhibition of inositol 1,4,5-trisphosphate (IP)-induced SR Ca(2+) release by 20 microM xestospongin D inhibited SOCC, whereas ACh-induced IP(3) production by 5 microM U-73122 had no effect. Inhibition of Ca(2+) release through ryanodine receptors (RyR) by 100 microM ryanodine also prevented Ca(2+) influx via SOCC. Qualitatively similar characteristics of SOCC-mediated Ca(2+) influx were observed with cyclopiazonic acid- vs. caffeine-induced SR Ca(2+) depletion. These data demonstrate that a Ni(2+)/La(3+)-sensitive Ca(2+) influx via SOCC in porcine ASM cells involves SR Ca(2+) release through both IP(3) and RyR channels. Additional regulation of Ca(2+) influx by agonist may be related to a receptor-operated, noncapacitative mechanism.  相似文献   

16.
The prostaglandin PGE(2), a metabolite of the cyclooxygenase pathway, activates Ca(2+)-permeable cation channels in erythrocyte cell membranes leading to entry of Ca(2+) with subsequent eryptosis, i.e. cell shrinkage, breakdown of phosphatidylserine (PS) asymmetry and membrane blebbing, all features typical for apoptosis in nucleated cells. PS exposing cells are recognized by macrophages, engulfed, degraded and thus cleared from circulating blood. The present study explored whether the specific lipoxygenase inhibitor Bay-Y5884 influences eryptosis. As determined by competitive ELISA, Bay-Y5884 (20 microM) enhanced the release of PGE(2) from human erythrocytes. According to whole-cell patch-clamp, Bay-Y5884 (20 microM) activated nonselective cation channels. The effect of Bay-Y5884 on cation channels was abolished by the cyclooxygenase inhibitor diclophenac (10 microM). Bay-Y5884 (30-40 microM) significantly increased erythrocyte free Ca(2+) concentration and PS exposure as analyzed in flow cytometry by Fluo3 fluorescence and annexin-V binding, respectively. PS exposure triggered by 20 microM (but not by 40 microM) Bay-Y5884 was blunted by cyclooxygenase inhibitors acetylsalicylic acid (50 microM) and diclophenac (10 microM). In conclusion, the lipoxygenase inhibitor Bay-Y5884 enhances erythrocyte PGE(2) formation with subsequent activation of non-selective cation channels, Ca(2+) entry and phospholipid scrambling.  相似文献   

17.
ATP-elicited oscillations of the concentration of free intracellular Ca(2+) ([Ca(2+)](i)) in rat brain astrocytes were abolished by simultaneous arachidonic acid (AA) addition, whereas the tetraenoic analogue 5,8,11,14-eicosatetraynoic acid (ETYA) was ineffective. Inhibition of oscillations is due to suppression by AA of intracellular Ca(2+) store refilling. Short-term application of AA, but not ETYA, blocked Ca(2+) influx, which was evoked by depletion of stores with cyclopiazonic acid (CPA) or thapsigargin (Tg). Addition of AA after ATP blocked ongoing [Ca(2+)](i) oscillations. Prolonged AA application without or with agonist could evoke a delayed [Ca(2+)](i) increase. This AA-induced [Ca(2+)](i) rise developed slowly, reached a plateau after 5 min, could be reversed by addition of bovine serum albumin (BSA), that scavenges AA, and was blocked by 1 microM Gd(3+), indicative for the influx of extracellular Ca(2+). Specificity for AA as active agent was demonstrated by ineffectiveness of C16:0, C18:0, C20:0, C18:2, and ETYA. Moreover, the action of AA was not affected by inhibitors of oxidative metabolism of AA (ibuprofen, MK886, SKF525A). Thus, AA exerted a dual effect on astrocytic [Ca(2+)](i), firstly, a rapid reduction of capacitative Ca(2+) entry thereby suppressing [Ca(2+)](i) oscillations, and secondly inducing a delayed activation of Ca(2+) entry, also sensitive to low Gd(3+) concentration.  相似文献   

18.
Recently, glycine has been shown to prevent liver injury after endotoxin treatment in vivo. We demonstrated that ethanol and endotoxin stimulated Kupffer cells to release PGE(2), which elevated oxygen consumption in parenchymal cells. Because glycine has been reported to protect renal tubular cells, isolated hepatocytes, and perfused livers against hypoxic injury, the purpose of this study was to determine whether glycine prevents increases in intracellular free Ca(2+) concentration ([Ca(2+)](i)) in hepatic parenchymal cells by agonists released during stress, such as with PGE(2) and adrenergic hormones. Liver parenchymal cells isolated from female Sprague-Dawley rats were cultured for 4 h in DMEM/F12 medium, and [Ca(2+)](i) in individual cells was assessed fluorometrically using the fluorescent calcium indicator fura 2. PGE(2) caused a dose-dependent increase in [Ca(2+)](i) from basal values of 130 +/- 10 to maximal levels of 434 +/- 55 nM. EGTA partially prevented this increase, indicating that either extracellular calcium or agonist binding is Ca(2+) dependent. 8-(Diethylamino)octyl 3,4,5-trimethoxybenzoate (TMB-8), an agent that prevents the release of Ca(2+) from intracellular stores, also partially blocked the increase in [Ca(2+)](i) caused by PGE(2), suggesting that intracellular Ca(2+) pools are involved. Together, these results are consistent with the hypothesis that both the intracellular and extracellular Ca(2+) pools are involved in the increase in [Ca(2+)](i) caused by PGE(2). Interestingly, glycine, which activates anion (i.e., chloride) channels, blocked the increase in [Ca(2+)](i) due to PGE(2) in a dose-dependent manner. Low-dose strychnine, an antagonist of glycine-gated chloride channel in the central nervous system, partially reversed the inhibition by glycine. When extracellular Cl(-) was omitted, glycine was much less effective in preventing the increase in [Ca(2+)](i) due to PGE(2). Phenylephrine, an alpha(1)-type adrenergic receptor agonist, also increased [Ca(2+)](i), as expected, from 159 +/- 20 to 432 +/- 43 nM. Glycine also blocked the increase in [Ca(2+)](i) due to phenylephrine, and the effect was also reversed by low-dose strychnine. Together, these data indicate that glycine rapidly blocks the increase in [Ca(2+)](i) in hepatic parenchymal cells due to agonists released during stress, most likely by actions on a glycine-sensitive anion channel and that this may be a major aspect of glycine-induced hepatoprotection.  相似文献   

19.
Cytoplasmic Ca(2+) ([Ca(2+)](i)) and membrane potential changes were measured in clonal pancreatic beta cells using a fluorimetric imaging plate reader (FLIPR). KCl (30 mM) produced a fast membrane depolarization immediately followed by increase of [Ca(2+)](i) in BRIN-BD11 cells. l-Alanine (10 mM) but not l-arginine (10 mM) mimicked the KCl profile and also produced a fast membrane depolarization and elevation of [Ca(2+)](i). Conversely, a rise in glucose from 5.6 mM to 11.1 or 16.7 mM induced rapid membrane depolarization, followed by a slower and delayed increase of [Ca(2+)](i). GLP-1 (20 nM) did not affect membrane potential or [Ca(2+)](i). In contrast, acetylcholine (ACh, 100 microM) induced fast membrane depolarization immediately followed by a modest [Ca(2+)](i) increase. When extracellular Ca(2+) was buffered with EGTA, ACh mobilized intracellular calcium stores and the [Ca(2+)](i) increase was reduced by 2-aminoethoxydiphenyl borate but not by dantrolene, indicating the involvement of inositol triphosphate receptors (InsP(3)R). It is concluded that membrane depolarization of beta cells by glucose stimulation is not immediately followed by elevation of [Ca(2+)](i) and other metabolic events are involved in glucose induced stimulus-secretion coupling. It is also suggested that ACh mobilizes intracellular Ca(2+) through store operated InsP(3)R.  相似文献   

20.
Little is known about the regulation of cytosolic calcium Ca(2+) levels ([Ca(2+)](i)) in breast cancer cells. We investigated the existence of capacitative calcium entry (CCE) in the tumorigenic cell line MCF-7 and its responsiveness to ATP. MCF-7 cells express purinergic receptors as well as estrogen receptors (ER). Depletion of calcium stores with thapsigargin (TG, 500 nM) or ATP (10 microM) in the absence of extracellular Ca(2+), resulted in a rapid and transient elevation in [Ca(2+)](i). After recovery of basal levels, Ca(2+) readmission (1.5 mM) to the medium increased Ca(2+) influx (twofold over basal), reflecting pre-activation of a CCE pathway. Cells pretreated with TG were unable to respond to ATP, thus indicating that the same Ca(2+) store is involved in their response. Moreover, IP(3)-dependent ATP-induced calcium mobilization and CCE were completely blocked using compound U-73122, an inhibitor of phospholipase C. Compound 2-APB (75 microM) and Gd(3+) (10 microM), antagonists of the CCE pathway, completely prevented ATP-stimulated capacitative Ca(2+) entry. CCE in MCF-7 cells was highly permeable to Mn(2+) and to the Ca(2+) surrogate Sr(2+). Mn(2+) entry sensitivity to Gd(3+) matched that of the Ca(2+) entry pathway. 17Beta-estradiol blocked ATP-induced CCE, but was without effect on TG-induced CCE. Besides, the estrogen blockade of the ATP-induced CCE was completely abolished by preincubation of the cells with an ER monoclonal antibody. ER alpha immunoreactivity could also be detected in a purified plasma membrane fraction of MCF-7 cells. These results represent the first evidence on the operation of a ATP-responsive CCE pathway in MCF-7 cells and also indicate that 17beta-estradiol interferes with this mechanism by acting at the cell surface 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