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.
Hille C  Walz B 《Cell calcium》2006,39(4):305-311
Stimulation with the neurotransmitter dopamine causes an amplitude-modulated increase in the intracellular Ca(2+) concentration ([Ca(2+)](i)) in epithelial cells of the ducts of cockroach salivary glands. This is completely attributable to a Ca(2+) influx from the extracellular space. Additionally, dopamine induces a massive [Na(+)](i) elevation via the Na(+)K(+)2Cl(-) cotransporter (NKCC). We have reasoned that Ca(2+)-entry is mediated by the Na(+)Ca(2+) exchanger (NCE) operating in the Ca(2+)-entry mode. To test this hypothesis, [Ca(2+)](i) and [Na(+)](i) were measured by using the fluorescent dyes Fura-2, Fluo-3, and SBFI. Inhibition of Na(+)-entry from the extracellular space by removal of extracellular Na(+) or inhibition of the NKCC by 10 microM bumetanide did not influence resting [Ca(2+)](i) but completely abolished the dopamine-induced [Ca(2+)](i) elevation. Simultaneous recordings of [Ca(2+)](i) and [Na(+)](i) revealed that the dopamine-induced [Na(+)](i) elevation preceded the [Ca(2+)](i) elevation. During dopamine stimulation, the generation of an outward Na(+) concentration gradient by removal of extracellular Na(+) boosted the [Ca(2+)](i) elevation. Furthermore, prolonging the dopamine-induced [Na(+)](i) rise by blocking the Na(+)/K(+)-ATPase reduced the recovery from [Ca(2+)](i) elevation. These results indicate that dopamine induces a massive NKCC-mediated elevation in [Na(+)](i), which reverses the NCE activity into the reverse mode causing a graded [Ca(2+)](i) elevation in the duct cells.  相似文献   

2.
We have studied cyclopiazonic acid (CPA)-sensitive store-operated Ca(2+) entry (SOCE) in cultured neurons and astrocytes and examined the effect of 2-[2-[4-(4-nitrobenzyloxy)phenyl]]isothiourea (KB-R7943), which is often used as a selective inhibitor of the Na(+)-Ca(2+) exchanger (NCX), on the SOCE. CPA increased transiently intracellular Ca(2+) concentration ([Ca(2+)](i)) followed by a sustained increase in [Ca(2+)](i) in neurons and astrocytes. The sustained increase in [Ca(2+)](i) depended on the presence of extracellular Ca(2+) and inhibited by SOCE inhibitors, but not by a Ca(2+) channel inhibitor. CPA also caused quenching of fura-2 fluorescence when the cells were incubated in Mn(2+)-containing medium. KB-R7943 at 10 microM inhibited significantly CPA-induced sustained increase in [Ca(2+)](i) in neurons and astrocytes. KB-R7943 also inhibited CPA-induced quenching of fura-2 fluorescence in the presence of extracellular Mn(2+). These results indicate that cultured neurons and astrocytes possess SOCE and that KB-R7943 inhibits not only NCX but also SOCE.  相似文献   

3.
We have investigated a new mechanism by which epidermal growth factor (EGF) increases intracellular Ca(2+) ([Ca(2+)](i)) in Rat-2 fibroblasts. EGF induced a transient increase of [Ca(2+)](i), and sustained Ca(2+) increase disappeared in the absence of extracellular Ca(2+). However, EGF had no effect on the formation of inositol phosphates. Expression of N17Rac or scrape-loading of C3 transferase blocked the elevation of [Ca(2+)](i) by EGF, but not by lysophosphatidic acid (LPA). EGF increased intracellular H(2)O(2), with a maximal increase at 5 min, which was blocked by catalase, scrape-loading of C3 transferase, or expression of N17Rac. H(2)O(2) scavengers, catalase and N-acetyl-L-cysteine, also blocked the Ca(2+) response to EGF, but not to LPA. In the presence of EGTA, preincubation with EGF completely inhibited subsequent Ca(2+) response to extracellular H(2)O(2) and vice versa. Incubation with EGF or phosphatidic acid abolished subsequent elevation of [Ca(2+)](i) by phosphatidic acid or EGF, respectively. Furthermore, preincubation with LPA inhibited the subsequent Ca(2+) response to EGF, but not vice versa. These results suggested that intracellular H(2)O(2) regulated by Rac and RhoA, but not inositol phosphates, was responsible for the EGF-stimulated elevation of [Ca(2+)](i). It was also suggested that EGF cross talked with LPA in the regulation of [Ca(2+)](i) by producing intracellular H(2)O(2).  相似文献   

4.
In Alzheimer's disease, amyloid beta (Abeta) peptide is deposited in neuritic plaques in the brain. The Abeta peptide 1-42 or the fragment 25-35 are neurotoxic. We here review our recent explorations of the mechanisms of Abeta toxicity in hippocampal cultures. Abeta had no effect on intracellular calcium in neurons but caused striking changes in nearby astrocytes. The [Ca(2+)](c) signals started approximately 5-15 min after Abeta application and consisted of sporadic [Ca(2+)](c) pulses. These were entirely dependent on extracellular Ca(2+), independent of ER Ca(2+) stores and resulted from Ca(2+) influx, probably through Abeta-induced membrane channels. The Ca(2+) signals were closely associated with transient, episodic acidification which may reflect displacement of protons from binding sites or Ca(2+)/2H(+) exchange. Abeta caused an increased rate of generation of reactive oxygen species (ROS), also seen in astrocytes and not in neurons. The increased ROS generation was blocked by inhibitors of the NADPH oxidase, strongly suggesting that this enzyme, normally associated with immune cells, is expressed in astrocytes. ROS generation was also Ca(2+)-dependent, suggesting that Abeta activation of the enzyme may be secondary to the increase in [Ca(2+)](c). Abeta caused delayed neuronal death despite the fact that all responses were seen only in astrocytes. Neurons could not be protected by glutamate receptor antagonists, but were rescued by inhibition of the NADPH oxidase, by antioxidants and by increasing glutathione. These data suggest that Abeta causes Ca(2+)-dependent oxidative stress by activating an astrocytic NADPH oxidase, and that neuronal death follows through a failure of antioxidant support.  相似文献   

5.
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   

6.
Recently, a new system of astrocyte-neurone glutamatergic signalling has been identified. It is started in astrocytes by ectocellular, CD38-catalysed conversion of NAD(+) to the calcium mobilizer cyclic ADP-ribose (cADPR). This is then pumped by CD38 itself into the cytosol where the resulting free intracellular Ca(2+) concentration [Ca(2+)](i) transients elicit an increased release of glutamate, which can induce an enhanced Ca(2+) response in neighbouring neurones. Here, we demonstrate that co-culture of either cortical or hippocampal astrocytes with neurones results in a significant overexpression of astrocyte CD38 both on the plasma membrane and intracellularly. The causal role of neurone-released glutamate in inducing overexpression of astrocyte CD38 is demonstrated by two observations: first, in the absence of neurones, induction of CD38 in pure astrocyte cultures can be obtained with glutamate and second, it can be prevented in co-cultures by glutamate receptor antagonists. The neuronal glutamate-mediated effect of neurones on astrocyte CD38 expression is paralleled by increased intracellular cADPR and [Ca(2+)](i) levels, both findings indicating functionality of overexpressed CD38. These results reveal a new neurone-to-astrocyte glutamatergic signalling based on the CD38/cADPR system, which affects the [Ca(2+)](i) in both cell types, adding further complexity to the bi-directional patterns of communication between astrocytes and neurones.  相似文献   

7.
Substance P (SP) plays an important role in pain transmission through the stimulation of the neurokinin (NK) receptors expressed in neurons of the spinal cord, and the subsequent increase in the intracellular Ca(2+) concentration ([Ca(2+)](i)) as a result of this stimulation. Recent studies suggest that spinal astrocytes also contribute to SP-related pain transmission through the activation of NK receptors. However, the mechanisms involved in the SP-stimulated [Ca(2+)](i) increase by spinal astrocytes are unclear. We therefore examined whether (and how) the activation of NK receptors evoked increase in [Ca(2+)](i) in rat cultured spinal astrocytes using a Ca(2+) imaging assay. Both SP and GR73632 (a selective agonist of the NK1 receptor) induced both transient and sustained increases in [Ca(2+)](i) in a dose-dependent manner. The SP-induced increase in [Ca(2+)](i) was significantly attenuated by CP-96345 (an NK1 receptor antagonist). The GR73632-induced increase in [Ca(2+)](i) was completely inhibited by pretreatment with U73122 (a phospholipase C inhibitor) or xestospongin C (an inositol 1,4,5-triphosphate (IP(3)) receptor inhibitor). In the absence of extracellular Ca(2+), GR73632 induced only a transient increase in [Ca(2+)](i). In addition, H89, an inhibitor of protein kinase A (PKA), decreased the GR73632-mediated Ca(2+) release from intracellular Ca(2+) stores, while bisindolylmaleimide I, an inhibitor of protein kinase C (PKC), enhanced the GR73632-induced influx of extracellular Ca(2+). RT-PCR assays revealed that canonical transient receptor potential (TRPC) 1, 2, 3, 4 and 6 mRNA were expressed in spinal astrocytes. Moreover, BTP2 (a general TRPC channel inhibitor) or Pyr3 (a TRPC3 inhibitor) markedly blocked the GR73632-induced sustained increase in [Ca(2+)](i). These findings suggest that the stimulation of the NK-1 receptor in spinal astrocytes induces Ca(2+) release from IP(3-)sensitive intracellular Ca(2+) stores, which is positively modulated by PKA, and subsequent Ca(2+) influx through TRPC3, which is negatively regulated by PKC.  相似文献   

8.
We analyzed the role of inositol 1,4,5-trisphosphate-induced Ca(2+) release from the endoplasmic reticulum (ER) (i) in powering mitochondrial Ca(2+) uptake and (ii) in maintaining a sustained elevation of cytosolic Ca(2+) concentration ([Ca(2+)](c)). For this purpose, we expressed in HeLa cells aequorin-based Ca(2+)-sensitive probes targeted to different intracellular compartments and studied the effect of two agonists: histamine, acting on endogenous H(1) receptors, and glutamate, acting on co-transfected metabotropic glutamate receptor (mGluR1a), which rapidly inactivates through protein kinase C-dependent phosphorylation and thus causes transient inositol 1,4,5-trisphosphate production. Glutamate induced a transient [Ca(2+)](c) rise and drop in ER luminal [Ca(2+)] ([Ca(2+)](er)), and then the ER refilled with [Ca(2+)](c) at resting values. With histamine, [Ca(2+)](c) after the initial peak stabilized at a sustained plateau, and [Ca(2+)](er) decreased to a low steady-state value. In mitochondria, histamine evoked a much larger mitochondrial Ca(2+) response than glutamate ( approximately 15 versus approximately 65 microm). Protein kinase C inhibition, partly relieving mGluR1a desensitization, reestablished both the [Ca(2+)](c) plateau and the sustained ER Ca(2+) release and markedly increased the mitochondrial Ca(2+) response. Conversely, mitochondrial Ca(2+) uptake evoked by histamine was drastically reduced by very transient ( approximately 2-s) agonist applications. These data indicate that efficient mitochondrial Ca(2+) uptake depends on the preservation of high Ca(2+) microdomains at the mouth of ER Ca(2+) release sites close to mitochondria. This in turn depends on continuous Ca(2+) release balanced by Ca(2+) reuptake into the ER and maintained by Ca(2+) influx from the extracellular space.  相似文献   

9.
A rise in intracellular free Ca(2+) concentration ([Ca(2+)](i)) is required to activate sperm of all organisms studied. Such elevation of [Ca(2+)](i) can occur either by influx of extracellular Ca(2+) or by release of Ca(2+) from intracellular stores. We have examined these sources of Ca(2+) in sperm from the sea squirt Ascidia ceratodes using mitochondrial translocation to evaluate activation and the Ca(2+)-sensitive dye fura-2 to monitor [Ca(2+)](i) by bulk spectrofluorometry. Sperm activation artificially evoked by incubation in high-pH seawater was inhibited by reducing seawater [Ca(2+)], as well as by the presence of high [K(+)](o) or the Ca channel blockers pimozide, penfluridol, or Ni(2+), but not nifedipine or Co(2+). The accompanying rise in [Ca(2+)](i) was also blocked by pimozide or penfluridol. These results indicate that activation produced by alkaline incubation involves opening of plasmalemmal voltage-dependent Ca channels and Ca(2+) entry to initiate mitochondrial translocation. Incubation in thimerosal or thapsigargin, but not ryanodine (even if combined with caffeine pretreatment), evoked sperm activation. Activation by thimerosal was insensitive to reduced external calcium and to Ca channel blockers. Sperm [Ca(2+)](i) increased upon incubation in high-pH or thimerosal-containing seawater, but only the high-pH-dependent elevation in [Ca(2+)](i) could be inhibited by pimozide or penfluridol. Treatment with the protonophore CCCP indicated that only a small percentage of sperm could release enough Ca(2+) from mitochondria to cause activation. Inositol 1,4,5-trisphosphate (IP(3)) delivered by liposomes or by permeabilization increased sperm activation. Both of these effects were blocked by heparin. We conclude that high external pH induces intracellular alkalization that directly or indirectly activates plasma membrane voltage-dependent Ca channels allowing entry of external Ca(2+) and that thimerosal stimulates release of Ca(2+) from IP(3)-sensitive intracellular stores.  相似文献   

10.
Glucose increases insulin secretion by raising cytoplasmic Ca(2+) ([Ca(2+)](i)) in beta-cells (triggering pathway) and augmenting the efficacy of Ca(2+) on exocytosis (amplifying pathway). It has been suggested that glutamate formed from alpha-ketoglutarate is a messenger of the amplifying pathway (Maechler, P., and Wollheim, C. B. (1999) Nature 402, 685-689). This hypothesis was tested with mouse islets depolarized with 30 mm KCl (+ diazoxide) or with a saturating concentration of sulfonylurea. Because [Ca(2+)](i) was elevated under these conditions, insulin secretion was stimulated already in 0 mm glucose. The amplification of secretion produced by glucose was accompanied by an increase in islet glutamate. However, glutamine (0.5-2 mm) markedly augmented islet glutamate without affecting insulin secretion, whereas glucose augmented secretion without influencing glutamate levels when these were elevated by glutamine. Allosteric activation of glutamate dehydrogenase by BCH (2-amino 2-norbornane carboxylic acid) lowered islet glutamate but increased insulin secretion. Similar insulin secretion thus occurred at very different cellular glutamate levels. Glutamine did not affect islet [Ca(2+)](i) and pH(i), whereas glucose and BCH slightly raised pH(i) and either slightly decreased (30 mm KCl) or increased (tolbutamide) [Ca(2+)](i). The general dissociation between changes in islet glutamate and insulin secretion refutes a role of beta-cell glutamate in the amplification of insulin secretion by glucose.  相似文献   

11.
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   

12.
The reverse-mode of the Na(+)/Ca(2+)-exchanger (NCX) mediates Ca(2+)-entry in agonist-stimulated vascular smooth muscle (VSM) and plays a central role in salt-sensitive hypertension. We investigated buffering of Ca(2+)-entry by peripheral mitochondria upon NCX reversal in rat aortic smooth muscle cells (RASMC). [Ca(2+)] was measured in mitochondria ([Ca(2+)](MT)) and the sub-plasmalemmal space ([Ca(2+)](subPM)) with targeted aequorins and in the bulk cytosol ([Ca(2+)](i)) with fura-2. Substitution of extracellular Na(+) by N-methyl-d-glucamine transiently increased [Ca(2+)](MT) ( approximately 2microM) and [Ca(2+)](subPM) ( approximately 1.3microM), which then decreased to sustained plateaus. In contrast, Na(+)-substitution caused a delayed and tonic increase in [Ca(2+)](i) (<100nM). Inhibition of Ca(2+)-uptake by the sarcoplasmic reticulum (SR) (30microM cyclopiazonic acid) or mitochondria (2microM FCCP or 2microM ruthenium red) enhanced the elevation of [Ca(2+)](subPM). These treatments also abolished the delay in the [Ca(2+)](i) response to 0Na(+) and increased its amplitude. Extracellular ATP (1mM) caused a peak and plateau in [Ca(2+)](i), and only the plateau was inhibited by KB-R7943 (10microM), a selective blocker of reverse-mode NCX. Evidence for ATP-mediated NCX-reversal was also found in changes in [Na(+)](i). Mitochondria normally exhibited a transient elevation of [Ca(2+)] in response to ATP, but inhibiting the mitochondrial NCX with CGP-37157 (10microM) unmasked an agonist-induced increase in mitochondrial Ca(2+)-flux. This flux was blocked by KB-R7943. In summary, mitochondria and the sarcoplasmic reticulum co-operate to buffer changes in [Ca(2+)](i) due to agonist-induced NCX reversal.  相似文献   

13.
Simultaneous somatic patch-pipette recording of a single astrocyte to evoke voltage-gated calcium currents, and Ca(2+) imaging, were used to study the spatial and temporal profiles of depolarization-induced changes in intracellular Ca(2+) ([Ca(2+)](i)) in the processes of cultured rat cortical astrocytes existing as pairs. Transient Ca(2+) changes locked to depolarization were observed as microdomains in the processes of the astrocyte pairs, and the responses were more pronounced in the adjoining astrocyte. Considering the functional significance of higher concentrations of glutamate observed in certain pathological conditions, Ca(2+) transients were recorded following pretreatment of cells with glutamate (500 microM for 20 min). This showed distance-dependent incremental scaling and attenuation in the presence of the metabotropic glutamate receptor (mGluR) antagonist, alpha-methyl(4-carboxy-phenyl) glycine (MCPG). Estimation of local Ca(2+) diffusion coefficients in the astrocytic processes indicated higher values in the adjoining astrocyte of the glutamate pretreated group. Intracellular heparin introduced into the depolarized astrocyte did not affect the Ca(2+) transients in the heparin-loaded astrocyte but attenuated the [Ca(2+)](i) responses in the adjoining astrocyte, suggesting that inositol 1,4,5 triphosphate (IP(3)) may be the transfer signal. The uncoupling agent, 1-octanol, attenuated the [Ca(2+)](i) responses in both the control and glutamate pretreated astrocytes, indicating the role of gap junctional communication. Our studies indicate that individual astrocytes have distinct functional domains, and that the glutamate-induced alterations in Ca(2+) signaling involve a sequence of intra- and intercellular steps in which phospholipase C (PLC), IP(3), internal Ca(2+) stores, VGCC and gap junction channels appear to play an important role.  相似文献   

14.
Liberation of zinc from intracellular stores contributes to oxidant-induced neuronal injury. However, little is known regarding how endogenous oxidant systems regulate intracellular free zinc ([Zn(2+)](i)). Here we simultaneously imaged [Ca(2+)](i) and [Zn(2+)](i) to study acute [Zn(2+)](i) changes in cultured rat forebrain neurons after glutamate receptor activation. Neurons were loaded with fura-2FF and FluoZin-3 to follow [Ca(2+)](i) and [Zn(2+)](i), respectively. Neurons treated with glutamate (100 microM) for 10 min gave large Ca(2+) responses that did not recover after termination of the glutamate stimulus. Glutamate also increased [Zn(2+)](i), however glutamate-induced [Zn(2+)](i) changes were completely dependent on Ca(2+) entry, appeared to arise entirely from internal stores, and were substantially reduced by co-application of the membrane-permeant chelator TPEN during the glutamate treatment. Pharmacological maneuvers revealed that a number of endogenous oxidant producing systems, including nitric oxide synthase, phospholipase A(2), and mitochondria all contributed to glutamate-induced [Zn(2+)](i) changes. We found no evidence that mitochondria buffered [Zn(2+)](i) during acute glutamate receptor activation. We conclude that glutamate-induced [Zn(2+)](i) transients are caused in part by [Ca(2+)](i)-induced reactive oxygen species that arises from both cytosolic and mitochondrial sources.  相似文献   

15.
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   

16.
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   

17.
Stimulation of Dictyostelium discoideum with cAMP evokes a change of the cytosolic free Ca(2+) concentration ([Ca(2+)](i)). We analyzed the role of the filling state of Ca(2+) stores for the [Ca(2+)] transient. Parameters tested were the height of the [Ca(2+)](i) elevation and the percentage of responding amoebae. After loading stores with Ca(2+), cAMP induced a [Ca(2+)](i) transient in many cells. Without prior loading, cAMP evoked a [Ca(2+)](i) change in a few cells only. This indicates that the [Ca(2+)](i) elevation is not mediated exclusively by Ca(2+) influx but also by Ca(2+) release from stores. Reducing the Ca(2+) content of the stores by EGTA preincubation led to a cAMP-activated [Ca(2+)](i) increase even at low extracellular [Ca(2+)]. Moreover, the addition of Ca(2+) itself elicited a capacitative [Ca(2+)](i) elevation. This effect was not observed when stores were emptied by the standard technique of inhibiting internal Ca(2+) pumps with 2,5-di-(t-butyl)-1,4-hydroquinone. Therefore, in Dictyostelium, an active internal Ca(2+)-ATPase is absolutely required to allow for Ca(2+) entry. No influence of the filling state of stores on Ca(2+) influx characteristics was found by the Mn(2+)-quenching technique, which monitors the rate of Ca(2+) entry. Both basal and cAMP-activated Mn(2+) influx rates were similar in control cells and cells with empty stores. By contrast, determination of extracellular free Ca(2+) concentration ([Ca(2+)](e)) changes, which represent the sum of Ca(2+) influx and efflux, revealed a higher rate of [Ca(2+)](e) decrease in EGTA-treated than in control amoebae. We conclude that emptying of Ca(2+) stores does not change the rate of Ca(2+) entry but results in inhibition of the plasma membrane Ca(2+)-ATPase. Furthermore, the activities of the Ca(2+) transport ATPases of the stores are of crucial importance for the regulation of [Ca(2+)](i) changes.  相似文献   

18.
In this study, we examined the acute effects of thyroid hormones (TH) T(3) and T(4), leading to improvement of myocardial function through activation of Ca(2+) extrusion mechanisms and, consequently, prevention of intracellular calcium overload. Extracellular calcium elevation from 1.8 to 3.8 mM caused immediate increase in intracellular calcium level ([Ca(2+)](i)) in newborn cardiomyocyte cultures. Administration of 10 or 100 nM T(3) or T(4) rapidly (within 10 sec) decreased [Ca(2+)](i) to its control level. Similar results were obtained when [Ca(2+)](i) was elevated by decreasing extracellular Na(+) concentration, causing backward influx of Ca(2+) through Na(+)/Ca(2+) exchanger, or by administration of caffeine, releasing Ca(2+) from the sarcoplasmic reticulum (SR). Under these conditions, T(3) or T(4) decreased [Ca(2+)](i). T(3) and T(4) also exhibited protective effects during ischemia. T(3) or T(4) presence during hypoxia for 120 min in culture medium restricted the increase of [Ca(2+)](i) and prevented the pathological effects of its overload. An inhibitor of SR Ca(2+)-ATPase (SERCA2a), thapsigargin, increases [Ca(2+)](i) and in its presence neither T(3) nor T(4) had any effect on the [Ca(2+)](i) level. The reduction of [Ca(2+)](i) level by T(3) and T(4) was also blocked in the presence of H-89 (a PKA inhibitor), and by calmodulin inhibitors. The effect of TH on the reduction of [Ca(2+)](i) was prevented by propranolol, indicating that the hormones exert their effect through interaction with adrenergic receptors. These results support our hypothesis that TH prevent calcium overload in newborn rat cardiomyocytes, most likely by a direct, acute, and nongenomic effect on Ca(2+) transport into the SR.  相似文献   

19.
Gastrin-releasing peptide (GRP) and its amphibian homolog, bombesin, are potent secretogogues in mammals. We determined the roles of intracellular free Ca(2+) ([Ca(2+)](i)), protein kinase C (PKC), and mitogen-activated protein kinases (MAPK) in GRP receptor (GRP-R)-regulated secretion. Bombesin induced either [Ca(2+)](i) oscillations or a biphasic elevation in [Ca(2+)](i). The biphasic response was associated with peptide secretion. Receptor-activated secretion was blocked by removal of extracellular Ca(2+), by chelation of [Ca(2+)](i), and by treatment with inhibitors of phospholipase C, conventional PKC isozymes, and MAPK kinase (MEK). Agonist-induced increases in [Ca(2+)](i) were also inhibited by dominant negative MEK-1 and the MEK inhibitor, PD89059, but not by an inhibitor of PKC. Direct activation of PKC by a phorbol ester activated MAPK and stimulated peptide secretion without a concomitant increase in [Ca(2+)](i). Inhibition of MEK blocked both bombesin- and phorbol 12-myristate 13-acetate-induced secretion. GRP-R-regulated secretion is initiated by an increase in [Ca(2+)](i); however, elevated [Ca(2+)](i) is insufficient to stimulate secretion in the absence of activation of PKC and the downstream MEK/MAPK pathways. We demonstrated that the activity of MEK is important for maintaining elevated [Ca(2+)](i) levels induced by GRP-R activation, suggesting that MEK may affect receptor-regulated secretion by modulating the activity of Ca(2+)-sensitive PKC.  相似文献   

20.
Active neurons communicate to intracerebral arterioles in part through an elevation of cytosolic Ca(2+) concentration ([Ca(2+)](i)) in astrocytes, leading to the generation of vasoactive signals involved in neurovascular coupling. In particular, [Ca(2+)](i) increases in astrocytic processes ("endfeet"), which encase cerebral arterioles, have been shown to result in vasodilation of arterioles in vivo. However, the spatial and temporal properties of endfoot [Ca(2+)](i) signals have not been characterized, and information regarding the mechanism by which these signals arise is lacking. [Ca(2+)](i) signaling in astrocytic endfeet was measured with high spatiotemporal resolution in cortical brain slices, using a fluorescent Ca(2+) indicator and confocal microscopy. Increases in endfoot [Ca(2+)](i) preceded vasodilation of arterioles within cortical slices, as detected by simultaneous measurement of endfoot [Ca(2+)](i) and vascular diameter. Neuronal activity-evoked elevation of endfoot [Ca(2+)](i) was reduced by inhibition of inositol 1,4,5-trisphosphate (InsP(3)) receptor Ca(2+) release channels and almost completely abolished by inhibition of endoplasmic reticulum Ca(2+) uptake. To probe the Ca(2+) release mechanisms present within endfeet, spatially restricted flash photolysis of caged InsP(3) was utilized to liberate InsP(3) directly within endfeet. This maneuver generated large amplitude [Ca(2+)](i) increases within endfeet that were spatially restricted to this region of the astrocyte. These InsP(3)-induced [Ca(2+)](i) increases were sensitive to depletion of the intracellular Ca(2+) store, but not to ryanodine, suggesting that Ca(2+)-induced Ca(2+) release from ryanodine receptors does not contribute to the generation of endfoot [Ca(2+)](i) signals. Neuronally evoked increases in astrocytic [Ca(2+)](i) propagated through perivascular astrocytic processes and endfeet as multiple, distinct [Ca(2+)](i) waves and exhibited a high degree of spatial heterogeneity. Regenerative Ca(2+) release processes within the endfeet were evident, as were localized regions of Ca(2+) release, and treatment of slices with the vasoactive neuropeptides somatostatin and vasoactive intestinal peptide was capable of inducing endfoot [Ca(2+)](i) increases, suggesting the potential for signaling between local interneurons and astrocytic endfeet in the cortex. Furthermore, photorelease of InsP(3) within individual endfeet resulted in a local vasodilation of adjacent arterioles, supporting the concept that astrocytic endfeet function as local "vasoregulatory units" by translating information from active neurons into complex InsP(3)-mediated Ca(2+) release signals that modulate arteriolar diameter.  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