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 31 毫秒
1.
Mitragyna ciliata (MYTA) (Rubiaceae) inhibits plasmodia activity. MYTA induces a cardiotonicity of the digitalic type on rat''s isolated heart. In this work we studied the effect of MYTA on microsomal Na+/K+ dependant ATPase (Na+, K+ ATPase) extracted from the heart of a rabbit since digitalics inhibit Na+, K+ ATPase. Our results revealed that the Na+/K+ ATPase has an optimum pH of 7.4 and temperature of 37°C respectively. There is a linear relationship between the organic phosphate formed and the incubation time over 25 mins incubation period. The ATP hydrolysis rate in the presence of MYTA was 0.775 µM/min. LINEWEAVER and BURK plots showed that MYTA did not alter KM (1.31 mM) but decreased VMAX. This study shows that MYTA exerts a non-competitive inhibition on the microsomal Na+/K+ ATPase extracted from rabbit heart with a Ci50 of 48 µg / ml. We conclude that the mechanism of action of MYTA is linked to the inhibition of the Na+/K+ ATPase like cardiotonics of the digitalic type.  相似文献   

2.
Numerous studies have demonstrated heightened Na+/Li+ countertransport (NLCT) activity in erythrocytes of patients with essential hypertension or diabetic nephropathy. The same carrier also contributes to the therapeutic action of lithium salt, widely used in the treatment of psychiatric disorders. However, the molecular origin of NLCT remains unknown. This study examined the role of major ion transporters in NLCT by comparative analysis of its activity and that of ion transporters providing inwardly directed 86Rb, 22Na and 32P fluxes. NLCT was below the detection limit in rat erythrocytes and ∼50-fold higher in rabbits compared to humans. Unlike NLCT, the activities of Na+,K+-ATPase, Na+,K+,2Cl cotransporter and anion exchanger were somewhat similar in the erythrocytes of these species, whereas Na+,Pi cotransport was in 1:2:6 proportion in rats, humans and rabbits, respectively. Loading of erythrocytes with Li+ for NLCT measurement did not affect the activity of Na+,Pi cotransporter. Keeping in mind that NLCT is much higher in rabbits vs humans and rats, we compared the set of membrane proteins in these species using 2-dimensional gel electrophoresis. This approach revealed 174 common spots, whereas 132 proteins were detected only in human and rabbit erythrocyte membranes. Among these proteins, we found 17 spots whose expression was higher by more than 5-fold in rabbit compared to human erythrocytes. Thus, our results argue against the involvement of major ion transporters in NLCT. They also show that comparative proteomics is a potent tool to identify the molecular origin of this carrier.  相似文献   

3.
We studied the effects of Na+ influx on large-conductance Ca2+-activated K+ (BKCa) channels in cultured human umbilical vein endothelial cells (HUVECs) by means of patch clamp and SBFI microfluorescence measurements. In current-clamped HUVECs, extracellular Na+ replacement by NMDG+ or mannitol hyperpolarized cells. In voltage-clamped HUVECs, changing membrane potential from 0 mV to negative potentials increased intracellular Na+ concentration ([Na+]i) and vice versa. In addition, extracellular Na+ depletion decreased [Na+]i. In voltage-clamped cells, BKCa currents were markedly increased by extracellular Na+ depletion. In inside-out patches, increasing [Na+]i from 0 to 20 or 40 mM reduced single channel conductance but not open probability (NPo) of BKCa channels and decreasing intracellular K+ concentration ([K+]i) gradually from 140 to 70 mM reduced both single channel conductance and NPo. Furthermore, increasing [Na+]i gradually from 0 to 70 mM, by replacing K+, markedly reduced single channel conductance and NPo. The Na+–Ca2+ exchange blocker Ni2+ or KB-R7943 decreased [Na+]i and increased BKCa currents simultaneously, and the Na+ ionophore monensin completely inhibited BKCa currents. BKCa currents were significantly augmented by increasing extracellular K+ concentration ([K+]o) from 6 to 12 mM and significantly reduced by decreasing [K+]o from 12 or 6 to 0 mM or applying the Na+–K+ pump inhibitor ouabain. These results suggest that intracellular Na+ inhibit single channel conductance of BKCa channels and that intracellular K+ increases single channel conductance and NPo. GH Liang and MY Kim contributed equally to this publication and therefore share the first authorship.  相似文献   

4.
Small (<25 μm in diameter) neurons of the dorsal root ganglion (DRG) express multiple voltage-gated Na+ channel subtypes, two of which being resistant to tetrodotoxin (TTX). Each subtype mediates Na+ current with distinct kinetic property. However, it is not known how each type of Na+ channel contributes to the generation of action potentials in small DRG neurons. Therefore, we investigated the correlation between Na+ currents in voltage-clamp recordings and corresponding action potentials in current-clamp recordings, using wild-type (WT) and NaV1.8 knock-out (KO) mice, to clarify the action potential electrogenesis in small DRG neurons. We classified Na+ currents in small DRG neurons into three categories on the basis of TTX sensitivity and kinetic properties, i.e., TTX-sensitive (TTX-S)/fast Na+ current, TTX-resistant (TTX-R)/slow Na+ current, and TTX-R/persistent Na+ current. Our concurrent voltage- and current-clamp recordings from the same neuron revealed that the action potentials in WT small DRG neurons were mainly dependent on TTX-R/slow Na+ current mediated by NaV1.8. It was surprising that a large portion of TTX-S/fast Na+ current was switched off in WT small DRG neurons due to a hyperpolarizing shift of the steady-state inactivation (h ), whereas in KO small DRG neurons which are devoid of TTX-R/slow Na+ current, the action potentials were generated by TTX-S/fast Na+ current possibly through a compensatory shift of h in the positive direction. We also confirmed that TTX-R/persistent Na+ current mediated by NaV1.9 actually regulates subthreshold excitability in small DRG neurons. In addition, we demon strated that TTX-R/persistent Na+ current can carry an action potential when the amplitude of this current was abnormally increased. Thus, our results indicate that the action potentials in small DRG neurons are generated and regulated with a combination of multiple mechanisms that may give rise to unique functional properties of small DRG neurons.  相似文献   

5.
Cultured vascular smooth muscle cells from porcine aortas incubated in Na+-free medium rapidly release their intracellular Na+ contents (Nai) (23±4% of baseline after 60 min incubation, mean ± SEM of 18 experiments). Total Nai release was inhibited by 35–40% after addition of ouabain and by 60–70% after addition of ouabain + bumetanide. Norepinephrine inhibited ouabain and bumetanide-sensitives Na+ efflux with an IC50 of about 10–9–10–8 M. Addition of the alpha-adrenergic agonist phenylephrine (10 M) to the cells mimicked the inhibitory action of norepinephrine on Nai release. Conversely, the beta-adrenergic agonist isoproterenol was without effect on Nai release. Simultaneous addition of 10 M norepinephrine and the alpha-adrenergic antagonist phentolamine prevented any effect of norepinephrine on the rate of Nai decline. In A-10 cultured vascular smooth muscle cells, the alpha-adrenergic agonist phenylephrine (10 M) inhibited 40.0±8.1% of ouabain-sensitive Rb+ influx and 70.7±6.9% of bumetanide-sensitive Rb+ influx (mean ± SEM of three experiments). 50% inhibition of bumetanide-sensitive Rb+ influx was obtained with about 5×10–7 M of phenylephrine. Our results show that in vascular smooth muscle cells a [Na+, K+, Cl]-cotransport system is able to catalyze outward Na+ movements (in Na+-free media) of a similar order of magnitude to those of the Na+, K+ pump and that alpha-adrenergic stimulation markedly inhibits Na+ efflux (and Rb+ influx) through these two transport systems.  相似文献   

6.
The electrophysiological mechanism underlying afterhyperpolarization induced by the activation of the nicotinic acetylcholine receptor (nAChR) in male rat major pelvic ganglion neurons (MPG) was investigated using a gramicidin-perforated patch clamp and microscopic fluorescence measurement system. Acetylcholine (ACh) induced fast depolarization through the activation of nAChR, followed by a sustained hyperpolarization after the removal of ACh in a dose-dependent manner (10 μM to 1 mM). ACh increased both intracellular Ca2+ ([Ca2+]i) and Na+ concentrations ([Na+]i) in MPG neurons. The recovery of [Na+]i after the removal of ACh was markedly delayed by ouabain (100 μM), an inhibitor of Na+/K+ ATPase. Pretreatment with ouabain blocked ACh-induced hyperpolarization by 67.2 ± 5.4% (n = 7). ACh-induced hyperpolarization was partially attenuated by either the chelation of [Ca2+]i with BAPTA/AM (20 μM) or the blockade of small-conductance Ca2+-activated K+ channels by apamin (500 nM). Taken together, the activation of nAChR increases [Na+]i and [Ca2+]i, which activates Na+/K+ ATPase and Ca2+-activated K+ channels, respectively. Consequently, hyperpolarization occurs after the activation of nAChR in the autonomic pelvic ganglia.  相似文献   

7.
The maintenance of extracellular Na+ and Cl- concentrations in mammals depends, at least in part, on renal function. It has been shown that neural and endocrine mechanisms regulate extracellular fluid volume and transport of electrolytes along nephrons. Studies of sex hormones and renal nerves suggested that sex hormones modulate renal function, although this relationship is not well understood in the kidney. To better understand the role of these hormones on the effects that renal nerves have on Na+ and Cl- reabsorption, we studied the effects of renal denervation and oophorectomy in female rats. Oophorectomized (OVX) rats received 17β-estradiol benzoate (OVE, 2.0 mg·kg-1·day-1, sc) and progesterone (OVP, 1.7 mg·kg-1·day-1, sc). We assessed Na+ and Cl- fractional excretion (FENa+ and FECl-, respectively) and renal and plasma catecholamine release concentrations. FENa+, FECl-, water intake, urinary flow, and renal and plasma catecholamine release levels increased in OVX vs control rats. These effects were reversed by 17β-estradiol benzoate but not by progesterone. Renal denervation did not alter FENa+, FECl-, water intake, or urinary flow values vs controls. However, the renal catecholamine release level was decreased in the OVP (236.6±36.1 ng/g) and denervated rat groups (D: 102.1±15.7; ODE: 108.7±23.2; ODP: 101.1±22.1 ng/g). Furthermore, combining OVX + D (OD: 111.9±25.4) decreased renal catecholamine release levels compared to either treatment alone. OVE normalized and OVP reduced renal catecholamine release levels, and the effects on plasma catecholamine release levels were reversed by ODE and ODP replacement in OD. These data suggest that progesterone may influence catecholamine release levels by renal innervation and that there are complex interactions among renal nerves, estrogen, and progesterone in the modulation of renal function.  相似文献   

8.
We have examined the rapid effect of 1,25-dihydroxyvitamin-D3 [1,25(OH)2D3] on apical Na+/H+ exchange activity in opossum kidney (OK) cells and in MCT cells (a culture of simian-virus-40-immortalized mouse cortical tubule cells) grown on filter support. Addition of 1,25(OH)2D3 (10 nM) for 1 min increased apical Na+/H+ exchange activity [recovery from an acid load; measured by 2,7-bis(2-carboxyethyl)-5(6)-carboxyfluorescein] in OK cells (by 56%) and in MCT cells (by 36%). The cellular mechanisms involved in 1,25(OH)2D3-dependent stimulation of Na+/H+ exchange were analysed in OK cells; stimulation of Na+/ H+ exchange by 1,25(OH)2D3 was not prevented by actinomycin D. Applying parathyroid hormone (PTH) reduced Na+/H+ exchange activity in OK cells (by 34% at 10 nM, 5 min); 1,25(OH)2D3 reversed PTH-induced inhibition, either when PTH was added prior to 1,25(OH)2D3 or when the two agonists were applied together. 1,25(OH)2D3 had no effect on basal OK cell cAMP content or on [Ca2+]i (fura-2). 1,25(OH)2D3 attenuated PTH-induced cAMP accumulation and had no effect on the PTH-dependent increase in [Ca2+]i. These data suggest a regulatory control (stimulation) of proximal tubular brush-border Na+/H+ exchange by 1,25(OH)2D3. This effect is non-genomic and might in part be explained by a release from cAMP-dependent control of transport activity.  相似文献   

9.
Single-channel currents from Na+-dependent K+ channels (KNa) were recorded from cell-attached and inside-out membrane patches of cultured avian trigeminal ganglion neurons by means of the patchclamp technique. Single-channel properties, such as the high elementary conductance and the occurrence of subconductance levels, were unchanged after the patches had been excised from the cells, indicating that they are not under the control of soluble cytoplasmic factors. In cellattached recordings at the cell resting potential the degree of KNa activity, measured as the probability of the channel being open, P o, was low in most cases (around 0.01) and similar to that observed in the inside-out configuration when the bath solution contained concentrations of Na+ around 30 mM and of K+ close to the physiological intracellular levels. However, in some cell-attached patches P o was high (around 0.2) and comparable to the values measured in cell-free recordings with high Na+ concentrations in the bath (100 mM). The excision of a highactivity patch in the presence of 30 mM Na+ resulted in a fall of P o in about 20 s, which is consistent with the wash-out of a soluble cytoplasmic molecule. After the excision all KNa displayed a similar Na+ sensitivity, irrespective of the degree of activation observed in the cellattached mode. In inside-out patches the P o values observed in the presence of either low or high concentrations of Na+ in bath solutions were not modified by internal Ca2+ (0.8–8.5 M). The variable degree of KNa activation observed in cell-attached recordings suggests that either internal Na+ concentrations reach very high levels close to the membrane, or soluble factor(s) are involved in the modulation of KNa activity: under such conditions, the Na+-activated K+ current may contribute to the maintenance of the resting membrane potential and to control neuronal membrane excitability.  相似文献   

10.
In the present study we used the Na+-sensitive fluorescent dye SBFI and optical measurement of endpiece volume to investigate the transport of Na+ in sheep parotid secretory cells. Sheep parotid endpiece cells bathed in a HCO 3 -free Cl-rich solution had a resting intracellular Na+ concentration ([Na+]i) of 17±2 mmol/l (n=39). Exposure of the cells to a 2-min pulse of acetylcholine (ACh) (3×10–7 mol/l) in a HCO 3 -free bathing solution produced no change in [Na+]i or in cell volume. Changing from a Cl-containing HCO 3 -free bath solution to a Cl solution containing 25 mmol/l HCO 3 caused the endpieces to swell by 8±2 % (n=11) and the [Na+]i to increase by 10±2 mmol/l (n=14). Subsequent exposure of the cells to ACh led to shrinkage of the cells by 12±2 % from the volume in the HCO 3 -containing solution prior to ACh exposure, with the maximum decrease occurring after 29±7 s (n=9). This shrinkage was accompanied by a rapid and transient increase in [Na+]i, the [Na+]i reaching a peak at 70±5 mmol/l above the unstimulated level (n=9). Substitution of gluconate for Cl did not significantly alter the effects of HCO 3 on unstimulated [Na+]i or endpiece volume, nor did it significantly inhibit the effects of ACh on these two parameters when HCO 3 was present. Addition of 200 mol/l dihydrogen-4,4-diisothiocyanatostilbene-2,2-disulfonic acid (H2-DIDS) to the gluconate/HCO 3 solution significantly reduced the peak of the ACh-induced increase in [Na+]i to 34±10 mmol/l (n=4), but did not have any significant effect on the magnitude of the ACh-induced shrinkage. At 500 mol/l, H2-DIDS abolished the ACh-induced increase in [Na+]i and also significantly reduced the shrinkage due to ACh. Finally, we found that the rate of endpiece shrinkage following ACh stimulation did not depend on the presence of Cl.We interpret these results as indicating that sheep parotid secretory cells do not contain significant Na+-K+-2Cl co-transport activity and do not actively accumulate Cl. Rather, the mechanism of spontaneous basal secretion by these cells, in the presence of extracellular HCO 3 , is based on the accumulation of HCO 3 by the Na+-H+ exchanger. During ACh stimulation, the concentration of HCO 3 in the cytosol is also maintained by the operation of a H2-DIDS-sensitive Na+-HCO 3 co-transporter. HCO 3 efflux across the apical membrane occurs via a HCO 3 conductance pathway rather than by the coupled operation of a Cl channel and a Cl-HCO 3 exchanger.  相似文献   

11.
Summary We measured the ouabain- and bumetanide-resistant Na+ efflux in Mg2+-sucrose medium (passive Na+ leak) in erythrocytes from 30 normotensive controls and 72 essential hypertensive patients. The mean values (±SEM) of the rate constant of Na+ leak (kpNa) were not significantly different between normotensives and hypertensives. Nevertheless, using the 95% confidence limits of the kpNa (in 10–3.h–1) in the normotensive group as a cut-off point, 7 (9.7%) essential hypertensives exhibited increased values (58.96±10.12) when compared with the other 65 patients (23.86±0.74). revealing increased passive Na+ permeability in the former (leak + hypertensives). Na+ fluxes depending on the Na+-K+ pump, outward Na+-K+ cotransport, and Na+-Li+ countertransport were also measured in fresh erythrocytes from the same 72 patients. Three of them (4.2%) exhibited decreased values of ouabain-sensitive Na+ efflux and 6 (8.3%) of bumetanide-sensitive Na+ efflux, while 8 patients (11.1%) showed increased values of Li+-stimulated Na+ efflux and, finally, 48 patients (59.7%) did not present any evident abnormality in these Na+ transport systems. No differences were observed between leak + hypertensives and the remaining 65 patients when both basal erythrocyte Na+ content and clinical parameters of hypertension were compared. However, Na+ efflux depending on the outward Na+-K+ cotransport was significantly higher in the leak + hypertensive subset (299.43±43.18 vs 181.52±10.76 µmol.(l cells.h)–1;P=0.0078), suggesting a compensatory phenomenon. Enhancement of Na+ permeability detected in 3% to 16% of essential hypertensives may be implicated in the pathogenesis of primary hypertension.Abbreviations ATPase adenosine triphosphatase - Dcat difference between the external Na+ concentration after incubation at 37° C and at zero time - kpNa rate constant of passive Na+ leak - Leak + hypertensive essential hypertensive patient with abnormally high erythrocyte Na+ leak - MOPS 4-morpholinopropanesulfonic acid - OBR ouabain- and bumetanide-resistant - PRA plasma renin activity - sPRA plasma renin activity stimulated after furosemide infusion - SEM standard error of the mean Supported in part by Grant 87/1078 of the Fondo de Investigaciones Sanitarias de la Seguridad Social and Grant PA85/0168 of the Comisión Asesora de Investigación Científica y Técnica  相似文献   

12.
Kinetic properties of the Na+-H+ antiport in the acinar cells of the isolated, superfused mouse lacrimal gland were studied by measuring intracellular pH (pHi) and Na+ activity (aNai) with the aid of double-barreled H+- and Na+-selective microelectrodes, respectively. Bicarbonate-free solutions were used throughout. Under untreated control conditions, pHi was 7.12±0.01 and aNai was 6.7±0.6 mmol/l. The cells were acid-loaded by exposure to an NH 4 + solution followed by an Na+-free N-methyl-d-glucamine (NMDG+) solution. Intracellular Na+ and H+ concentrations were manipulated by changing the duration of exposure to the above solutions. Subsequent addition of the standard Na+ solution rapidly increased pHi. This Na+-induced increase in pHi was almost completely inhibited by 0.5 mmol/l amiloride and was associated with a rapid, amiloride-sensitive increase in aNai. The rate of pHi recovery induced by the standard Na+ solution increased in a saturable manner as pHi decreased, and was negligible at pHi 7.2–7.3, indicating an inactivation of the Na+-H+ antiport. The apparent K m for intracellular H+ concentration was 105 nmol/l (pH 6.98). The rate of acid extrusion from the acid-loaded cells increased proportionally to the increase in extracellular pH. Depletion of aNai to less than 1 mmol/l by prolonged exposure to NMDG+ solution significantly increased the rate of Na+-dependent acid extrusion. The rate of acid extrusion increased as the extracellular Na+ concentration increased following Michaelis-Menten kinetics (V max was 0.55 pH/min and the apparent K m was 75 mmol/l at pHi 6.88). The results clearly showed that the Na+-H+ antiport activity is dependent on the chemical potential gradient of both Na+ and H+ ions across the basolateral membrane, and that the antiporter is asymmetric with respect to the substrate affinity of the transport site. The data agree with the current model of activation and inactivation of the antiporter by an intracellular site through changes in the intracellular Na+ and H+ concentrations.  相似文献   

13.
The current studies examine the presence of the Na+-HCO3 cotransporter in chicken enterocytes and its role in cytosolic pH (pHi) regulation. The pH-sensitive dye 2,7-bis(carboxyethyl)-5,6-carboxy-fluorescein (BCECF) was used to monitor pHi. Under resting conditions, pHi was 7.25 in solutions buffered with bis(2-hydroxyethyl)-1-piperazine ethanesulphonic acid (HEPES) and 7.17 in those buffered with HCO3 . Removal of external Na+ decreased pHi and readdition of Na+ rapidly increased pHi towards the control values. These Na+-dependent changes were greater in HCO 3 than in HEPES-buffered solutions. In HCO 3 - free solutions the Na+-dependent changes in pHi were prevented by 5-(N-ethyl-N-isopropyl)-amiloride (EIPA) and unaffected by 4,4-diisothiocyanatostilbene disulphonic acid (H2-DIDS). In the presence of HCO 3 , the Na+-induced changes in pHi were sensitive to both EIPA and H2-DIDS. In the presence of EIPA, cells partially recovered from a moderate acid load only when both Na+ and HCO 3 were present. This pHi recovery, which was EIPA resistant, and dependent on Na+ and HCO 3 , was inhibited by H2-DIDS and occurred at equal rates in both Cl-containing and Cl-free solutions. Kinetic analysis of the rate of HCO 3 - and Na+- dependent pHi recovery from an acid load as a function of the Na+ concentration revealed first-order kinetics with a Michaelis constant, K m, of 11 mmol/l Na+. It is concluded that in HCO3 /– buffered solutions both the Na+/H+ exchanger and the Na+-HCO3 cotransporter participate in setting the resting pHi in isolated chicken enterocytes and help the recovery from acid loads.  相似文献   

14.
Summary Short-circuit current (SCC) techniques were used to monitor the effects of various diuretic agents on Na+ transport in isolated frog skin, a model for the late distal tubule and the collecting duct of the mammalian kidney. Acctazolamide, hydrochlorothiazide, torasemide, and ethacrynic acid did not affect sodium transport (as indicated by the SCC) or transepithelial electrical resistance when added either to the apical (outer) or to the inner (basolateral, corial) bathing solution of the tissue. However, Na+ transport was sensitive to amiloride, the triamterene derivate dimethylamino-hydroxypropoxytriamterene (RPH 2823), and to furosemide. Whereas apical amiloride, and RPH 2823 induced a dose-dependent decrease in SCC and increase in transepithelial electrical resistance, apical furosemide resulted in a dose-dependent increase in SCC and a decrease in electrical resistance. None of the three diuretic agents caused a significant change in SCC when applied to the inner bathing Ringer's solution. The small furosemide-induced decrease in resistance compared with the huge increase in SCC suggests that furosemide affects Cl permeability as well as Na+ permeability. Evidence for this notion was achieved by the following findings: (a) The decrease in resistance after furosemide was more pronounced in tissues bathed in Cl-free solutions compared with Cl-containing solutions. (b) In contrast, SCC stimulation by apical furosemide is Cl-ion independent, but strongly Na+-ion dependent. (c) SCC stimulation by furosemide is amiloride-sensitive. With respect to the onset, locus, and reversibility of action, it seems reasonable to assume that amiloride, RPH 2823, and furosemide all influence transepithelial Na+ transport by interacting with the Na+ channel or a regulator site of it within the apical membrane. The stoichiometry of the amiloride (RPH 2823)-receptor site interaction revealed Hill-coefficient(s) of less than 1, indicating a negative cooperativity among the receptor sites. The interaction between Na+ ions and amiloride or RPH 2823 displayed mixed competitive-noncompetitive inhibition. Taken together, these results support the hypothesis that amiloride and Na+ as well as RPH 2823 and Na+ may act at different loci on the apical entry mechanism inRana esculenta skin.Abbreviations R Transepithelial electrical resistance - R x /R0 Transepithelial electrical resistance at timex (the time after manipulation) divided by the value at time 0 (the time before manipulation) - SCC Short-circuit current - SCC x /SCC0 Short-circuit current at timex (the time after manipulation) divided by the value at time 0 (the time before manipulation) Dedicated to Prof. Dr. F. Krück on the occasion of his 65th birthday  相似文献   

15.
Coenzyme Q10 (CoQ10) exerts neuroprotective effects in several in vivo and in vitro models of neurodegenerative disorders. However, the mechanisms of action are not fully understood. The aim in this study was to investigate whether oral administration of CoQ10 could inhibit cytochrome c (cyt c) release from mitochondria induced by 1-methyl-4-phenylpyridinium ion (MPP+), which causes dopaminergic cell death by selective inhibition of complex I of the electron transport chain, in mouse brain synaptosomes. An increase of cyt c was detected in the cytosolic fraction from mouse brain synaptosomes treated with MPP+. Oral administration of CoQ10 prevented the mitochondrial cyt c release in the MPP+-treated synaptosomes. In addition, CoQ10 did not affect the MPP+-induced decrease in mitochondrial oxidation–reduction activity and membrane potential in brain synaptosomes. Our findings demonstrate that MPP+-induced mitochondrial cyt c release in brain synaptosomes is prevented by oral administration of CoQ10 independently of mitochondrial dysfunction prior to the cyt c release.  相似文献   

16.
The specific inhibitor of the -aminobutyric acid (GABA) carrier, NNC-711, {1-[(2-diphenylmethylene) amino]oxyethyl}-1,2,5,6-tetrahydro-3-pyridinecarboxylic acid hydrochloride, blocks the Ca2+-independent release of [3H]GABA from rat brain synaptosomes induced by 50 mM K+ depolarization. Thus, in the presence of this inhibitor, it was possible to study the Ca2+-dependent release of [3H]GABA in the total absence of carrier-mediated release. Reversal of the Na+/Ca2+ exchanger was used to increase the intracellular free Ca2+ concentration ([Ca2+]i) to test whether an increase in [Ca2+]i alone is sufficient to induce exocytosis in the absence of depolarization. We found that the [Ca2+]i may rise to values above 400 nM, as a result of Na+/Ca2+ exchange, without inducing release of [3H]GABA, but subsequent K+ depolarization immediately induced [3H]GABA release. Thus, a rise of only a few nanomolar Ca2+ in the cytoplasm induced by 50 mM K+ depolarization, after loading the synaptosomes with Ca2+ by Na+/Ca2+ exchange, induced exocytotic [3H]GABA release, whereas the rise in cytoplasmic [Ca2+] caused by reversal of the Na+/Ca2+ exchanger was insufficient to induce exocytosis, although the value for [Ca2+]i attained was higher than that required for exocytosis induced by K+ depolarization. The voltage-dependent Ca2+ entry due to K+ depolarization, after maximal Ca2+ loading of the synaptosomes by Na+/Ca2+ exchange, and the consequent [3H]GABA release could be blocked by 50 M verapamil. Although preloading the synaptosomes with Ca2+ by Na+/Ca2+ exchange did not cause [3H]GABA release under any conditions studied, the rise in cytoplasmic [Ca2+] due to Na+/Ca2+ exchange increased the sensitivity to external Ca2+ of the exocytotic release of [3H]GABA induced by subsequent K+ depolarization. Thus, our results show that the vesicular release of [3H]GABA is rather insensitive to bulk cytoplasmic [Ca2+] and are compatible with the view that GABA exocytosis is triggered very effectively by Ca2+ entry through Ca2+ channels near the active zones.  相似文献   

17.
Luminal P2 receptors are ubiquitously expressed in transporting epithelia. In steroid-sensitive epithelia (e.g., lung, distal nephron) epithelial Na+ channel (ENaC)-mediated Na+ absorption is inhibited via luminal P2 receptors. In distal mouse colon, we have identified that both, a luminal P2Y2 and a luminal P2Y4 receptor, stimulate K+ secretion. In this study, we investigate the effect of luminal adenosine triphosphate/uridine triphosphate (ATP/UTP) on electrogenic Na+ absorption in distal colonic mucosa of mice treated on a low Na+ diet for more than 2 weeks. Transepithelial electrical parameters were recorded in an Ussing chamber. Baseline parameters: transepithelial voltage (V te): −13.7 ± 1.9 mV (lumen negative), transepithelial resistance (R te): 24.1 ± 1.8 Ω cm2, equivalent short circuit current (I sc): −563.9 ± 63.8 μA/cm2 (n = 21). Amiloride completely inhibited I sc to −0.5 ± 8.5 μA/cm2. Luminal ATP induced a slowly on-setting and persistent inhibition of the amiloride-sensitive I sc by 160.7 ± 29.7 μA/cm2 (n = 12, NMRI mice). Luminal ATP and UTP were almost equipotent with IC50 values of 10 μM and 3 μM respectively. In P2Y2 knock-out (KO) mice, the effect of luminal UTP on amiloride-sensitve Na+ absorption was absent. In contrast, in P2Y4 KO mice the inhibitory effect of luminal UTP on Na+ absorption remained present. Semiquantitative polymerase chain reaction did not indicate regulation of the P2Y receptors under low Na+ diet, but it revealed a pronounced axial expression of both receptors with highest abundance in surface epithelia. Thus, luminal P2Y2 and P2Y4 receptors and ENaC channels co-localize in surface epithelium. Intriguingly, only the stimulation of the P2Y2 receptor mediates inhibition of electrogenic Na+ absorption.  相似文献   

18.
Intracellular Na+ concentration ([Na+]i) rises in the heart during ischemia, and on reperfusion, there is a transient rise followed by a return toward control. These changes in [Na+]i contribute to ischemic and reperfusion damage through their effects on Ca2+ overload. Part of the rise of [Na+]i during ischemia may be caused by increased activity of the cardiac Na+/H+ exchanger (NHE1), activated by the ischemic rise in [H+]i. In support of this view, NHE1 inhibitors reduce the [Na+]i rise during ischemia. Another possibility is that the rise of [Na+]i during ischemia is caused by Na+ influx through channels. We have reexamined these issues by use of two different NHE1 inhibitors, amiloride, and zoniporide, in addition to tetrodotoxin (TTX), which blocks voltage-sensitive Na+ channels. All three drugs produced cardioprotection after ischemia, but amiloride (100 μM) and TTX (300 nM) prevented the rise in [Na+]i during ischemia, whereas zoniporide (100 nM) did not. Both amiloride and zoniporide prevented the rise of [Na+]i on reperfusion, whereas TTX was without effect. In an attempt to explain these differences, we measured the ability of the three drugs to block Na+ currents. At the concentrations used, TTX reduced the transient Na+ current (I Na) by 11 ± 2% while amiloride and zoniporide were without effect. In contrast, TTX largely eliminated the persistent Na+ current (I Na,P) and amiloride was equally effective, whereas zoniporide had a substantially smaller effect reducing I Na,P to 41 ± 8%. These results suggest that part of the effect of NHE1 inhibitors on the [Na+]i during ischemia is by blockade of I Na,P. The fact that a low concentration of TTX eliminated the rise of [Na+]i during ischemia suggests that I Na,P is a major source of Na+ influx in this model of ischemia.  相似文献   

19.
The epithelial Na+ conductance was expressed in Xenopus laevis oocytes by injection of size-fractionated mRNA of bovine tracheal epithelium. Fractionation was achieved by sucrose density gradient centrifugation. Successful expression was analysed by recording current/voltage (I/V) curves in the presence and absence of amiloride (10 mol/l). The newly expressed conductance was half-maximally inhibited by 44 nmol/l amiloride and exhibited a selectivity for Na+ over K+ of 1401. I/V curves obtained at different extracellular Na+ concentrations ([Na+]o) were subjected to a Goldman-fit analysis to obtain the relation between Na+ permeability (P Na) and [Na+]o. The data show that decreasing [Na+]o from 85 mmol/l to 0.85 mmol/l increased P Na by more than threefold, which is thought to reflect Na+ channel inhibition by increasing [Na+]o. This effect clearly exceeded what can be attributed to concentration saturation of single Na+ channel conductance (Palmer and Frindt (1986) Proc Natl Acad Sci USA 83:2767). No correlation of inhibition with intracellular Na+ concentration was observed. Preservation of the [Na+]o-dependent self-inhibition by the newly expressed Na+ conductance suggests that it is an intrinsic property of the Na+ channel protein, probably mediated by an extracellular Na+ binding site.  相似文献   

20.
The pH regulation in HT29 colon carcinoma cells has been investigated using the pH-sensitive fluorescent indicator 2,7-biscarboxyethyl-5(6)-carboxyfluorescein (BCECF). Under control conditions, intracellular pH (pHi) was 7.21±0.07 (n=22) in HCO 3 -containing and 7.21±0.09 (n=12) in HCO 3 -free solution. HOE-694 (10 mol/l), a potent inhibitor of the Na+/H+ exchanger, did not affect control pHi. As a means to acidify cells we used the NH 4 + /NH3 (20 mmol/l) prepulse technique. The mean peak acidification was 0.37±0.07 pH units (n=6). In HCC 3 -free solutions recovery from acid load was completely blocked by HOE-694 (1 mol/l), whereas in HCO3 3 -containing solutions a combination of HOE-694 and 4,4-diisothiocyanatostilbene-2, 2-disulphonate (DIDS, 0.5 mmol/l) was necessary to show the same effect. Recovery from acid load was Na+-dependent in HCO 3 -containing and HCO 3 -free solutions. Removal of external Cl caused a rapid, DIDS-blockable alkalinization of 0.33±0.03 pH units (n=15) and of 0.20±0.006 pH units (n=5), when external Na+ was removed together with Cl. This alkalinization was faster in HCO 3 -containing than in HCO 3 -free solutions. The present observations demonstrate three distinct mechanisms of pH regulation in HT29 cells: (a) a Na+/H+ exchanger, (b) a HCO 3 /Cl exchanger and (c) a Na+-dependent HCC 3 transporter, probably the Na+-HCO 3 /Cl antiporter. Under HCO 3 — free conditions the Na+/H+ exchanger fully accounts for recovery from acid load, whereas in HCO 3 -containing solutions this is accomplished by the Na+/H+ exchanger and a Na+-dependent mechanism, which imports HCO 3 . Recovery from alkaline load is caused by the HCO 3 /Cl exchanger.This study was supported by DFG Gr 480/10  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