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.
Gap junctions between segments of the crayfish septate axon mediate electrotonic transmission of impulses propagating along the length of the nerve cord. We simultaneously measured intracellular pH (pHi) and gap junctional conductance (gj) while axons were exposed to saline equilibrated with CO2, weak acids, and the weak base ammonium chloride. Normal pHi is about 7.1. When pHi is elevated, gj is unaffected. When pHi is reduced, gj declines with an apparent pK of about 6.7 and a Hill coefficient of about 2.7. We also measured effects of pHi on non-junctional conductance (gnj) and on the coupling coefficient, k. Over the pHi range 6.2-8, gnj increases approximately linearly with pHi. Since k is a function of gj and gnj, it reached a maximum at about pHi 7.1, decreasing at higher and lower pHi. The pHi dependence of gj in crayfish septate axon is less steep and has a lower apparent pK than the gj-pHi relation in two vertebrate embryos previously examined. This finding illustrates a difference in gating among analogous and possibly homologous membrane channels.  相似文献   

2.
The purpose of this work was to characterize the gap junctions between cultured glomus cells of the rat carotid body and to assess the effects of acidity and accompanying changes in [Ca2+]i on electric coupling. Dual voltage clamping of coupled glomus cells showed a mean macrojunctional conductance (Gj) of 1.16 nS±0.6 (S.E.), range 0.15–4.86 nS. At normal pHo (7.43), a steady transjunctional voltage (ΔVj=100.1±10.9 mV) showed multiple junction channel activity with a mean microconductance (gj) of 93.98±0.6 pS, range 0.3–324.5 pS. Single-channel conductances, calculated as variance/mean gj, gave a mean value of 16.7±0.2 pS, range 5.13–39.38 pS. Manual measurements of single-channel activity showed a mean gj of 22.03±0.2 pS, range 1.3–160 pS. Computer analysis of the noise spectral density distribution gave a channel mean open time of 12.7±1.5 ms, range 6.37–23.42 ms. The number of junction channels, estimated in each experiment from Gj/single-channel gj, showed a range of 7 to 258 channels (mean, 107.2). Optical measurements of [Ca2+]i gave a mean value of 80.2±4.27 nM at pHo of 7.43. Acidification of the medium with lactic acid (1 mM, pH 6.3) induced: 1) Variable changes in Gj (decreases and increases); 2) A significant decrease in mean gj (to 80.36±0.34 pS) and in single-channel conductance (gj=12.8±0.2 pS in computer analyses and 17.23±0.2 pS when measured by hand); 3) Variable changes in open times, resulting in a similar mean (12.8±1.5 ms) and 4) No change in the number of junction channels. When pHo was lowered to 6.3 [Ca2+]i did not change significantly (there were increases and decreases). However, when pHo was lowered to 4.4, [Ca2+]i increased significantly to 157.1±8.1 nM. It is concluded that saline acidification to pH 6.3 depresses the conductance of junction channels and this effect may be either a direct effect on channel proteins or synergistically enhanced by increases in [Ca2+]i. However, there are no studies correlating changes of [Ca2+]i and intercellular coupling in glomus cells. Stronger acidification (pHo 4.4), producing much larger changes in [Ca2+]i, may enhance this synergism. But, again, there are no studies correlating these effects.  相似文献   

3.
Octanol rapidly closes gap junction channels but its mechanism of action is not known. Because intracellular [H+], pHi, also affects the conductance of gap junctions, we studied octanol's effects on pHi in cultured rat astrocytes, which are highly coupled cells. Octanol (1 mM) caused an acid shift in the pHi of 90% of rat hippocampal astrocytes which averaged −0.19 ± 0.09 pH units in magnitude. In 58% of the cells tested, a biphasic change in pHi was seen; octanol produced an initial acidification lasting ∼10 min that was followed by a persistent alkalinization. The related gap junction uncoupling agent, heptanol, had similar effects on pHi. Octanol-induced changes in pHi were similar in nominally HCO3-free and HCO3-containing solutions, although the rate of initial acidification was significantly greater in the presence of HCO3. The initial acidification was inhibited in the presence of the stilbene DIDS, an inhibitor of Na+/HCO3 cotransport, indicating that octanol caused acidification by blocking this powerful acid extruder. The alkalinization was inhibited by amiloride which blocks the Na+/H+ exchanger (NHE), an acid extruder, suggesting that the alkaline shift induced by octanol was caused by stimulation of NHE. As expected, octanol's effects on astrocytic pHi were prevented by removal of external Na+, which blocks both Na+/HCO3 cotransport and NHE. Octanol had only small effects on intracellular Ca2+ (Ca2+i) in astrocytes. Hepatocytes which, like astrocytes, are strongly coupled to one another, showed no change in pHi with octanol application. Fluorescence recovery after photobleaching (FRAP) was used to study the effect of changes in astrocyte pHi on degree of coupling in hippocampal astrocytes. Coupling was decreased by intracellular acid shifts ∼−0.2 pH units in size. Octanol's effects on astrocyte pHi were complex but a prompt initial acidification was nearly always seen and could contribute to the uncoupling action of this drug in astrocytes. Because octanol uncouples hepatocytes without changing their pHi, this compound clearly can influence gap junctional conductance independent of changes in pHi. © 1996 Wiley-Liss, Inc.  相似文献   

4.
Variation of gap junction sensitivity to H ions with time of day   总被引:1,自引:0,他引:1  
A.P. Moreno  F. Ramon  D.C. Spray   《Brain research》1987,400(1):181-184
Conductance of gap junctions between segments of the lateral axons of the crayfish nerve cord is reduced by cytoplasmic acidification. Simultaneous measurements of axoplasmic pH and junctional conductance at 09.00 and 12.00 h revealed a shift of about 0.25 pH units in the apparent pK of the crayfish electrotonic synapse. These findings indicate that gating mechanisms of gap junctions are not fixed properties, but may rather be modulated by other, possibly humoral, factors.  相似文献   

5.
Glutamate is an excitatory receptor agonist in both neurones and glial cells, and, in addition, glutamate is also a substrate for glutamate transporter in glial cells. We have measured intracellular and extracellular pH changes induced by bath application of glutamate, its receptor agonist kainate, and its transporter agonist aspartate, in the giant neuropile glial cell in the central nervous system of the leech Hirudo medicinalis, using double-barrelled pH-sensitive microelectrodes. The giant glial cells responded to glutamate and aspartate (100–500 μM), and kainate (5–20 μM) with a membrane depolarization or an inward current, and with a distinct intracellular acidification. Glutamate and aspartate (both 500 μM) evoked a decrease in intracellular pH (pHi) by 0.187 ± 0.081 (n = 88) and 0.198 ± 0.067 (n = 86) pH units, respectively. With a resting pHi of 7.1 or 80 nM H+, these acidifications correspond to a mean increase of the intracellular H+ activity by 42 nM and 45 nM. Kainate caused a decrease of pHi by 0.1 − 0.35 pH units (n = 15). The glutamate/aspartate-induced decrease in pHi was not significantly affected by the glutamate receptor blockers kynurenic acid (1 mM) and 6-cyano-7-dinitroquinoxaline-2,3-dione (CNQX, 50–100 μM), which greatly reduced the kainate-induced change in pHi. Extracellular alkalinizations produced by glutamate and aspartate were not affected by CNQX. Reduction of the external Na+ concentration gradually decreased the intracellular pH change induced by glutamate/aspartate, indicating half maximal activation of the acidifying process at 5–10 mM external Na+ concentration. When all external Na+ was replaced by NMDG+, the pHiresponses were completely suppressed (glutamate) or reduced to 10% (aspartate). When Na+ was replaced by Li+, the glutamate- and aspartate-evoked pHi responses were reduced to 18% and 14%, respectively. Removal of external Ca2+ reduced the glutamate- and aspartate-induced pHi responses to 93 and 72%, respectively. The glutamate/aspartate-induced intracellular acidifications were not affected by the putative glutamate uptake inhibitor amino-adipidic acid (1 mM). DL-aspartate-β-hydroxamate (1 mM), and dihydrokainate (2 mM), which caused some pHi decrease on its own, reduced the glutamate/aspartate-induced pHi responses by 40 and 69%, respectively. The putative uptake inhibitor DL-threo-β-hydroxyaspartate (THA, 1 mM) induced a prominent intracellular acidification (0.36 ± 0.05 pH units, n = 9), and the pHi change evoked by glutamate or aspartate in the presence of THA was reduced to less than 10%. The results indicate that glutamate, aspartate, and kainate produce substantial intracellular acidifications, which are mediated by at least two independent mechanisms: 1) via activation of non-NMDA glutamate receptors and 2) via uptake of the excitatory amino acids into the leech glial cell. © 1997 Wiley-Liss Inc.  相似文献   

6.
As alterations in intracellular pH (pHi) tend to exert a profound effect on the properties of cells, this study was undertaken to examine NMDA-induced changes in pHi in rat hippocampal slices using the BCECF fluorescent technique. The ‘resting' pHi in the CA1 pyramidal cell layers was 6.93±0.07 (mean±S.D., n=72 slices) in 25 mM HCO3/5% CO2-buffered solution at 37°C. Exposure of hippocampal slices to NMDA in the range of 10–1000 μM produced a biphasic change in pHi: an initial transient alkaline shift was followed by a long-lasting acid shift. Dizocilpine (10 μM) but not CNQX (40 μM) blocked the NMDA-induced changes in pHi. In 0 Ca medium (0 mM Ca2+ supplemented 1 mM EGTA, referred to as 0 Ca), pHi acid shift caused by NMDA (20 μM) declined by about 11%, whereas the initial alkaline shift almost completely disappeared. In an independent experiment, the NMDA-induced increase in intracellular Ca2+ ([Ca2+]i) was reduced by more than 80% in 0 Ca medium. Glucose substitution using equimolar pyruvate (as an energy-yielding substrate) suppressed this NMDA-induced pHi acid shift by two-thirds, while the NMDA-induced pHi alkaline shift was enhanced. Fluoride (10 mM), a glycolytic inhibitor, abolished NMDA-induced pHi acid shift. Furthermore, the lactate content of hippocampal slices was markedly increased following exposure to NMDA. In conclusion, activation of NMDA receptors in rat hippocampal slices evokes a biphasic change in pHi. The initial alkaline shift is suggested to be associated with calcium influx, and the following acid shift may be caused by an increase in lactate production through the acceleration of glycolysis, as well as the increased [Ca2+]i. The pHi acid shift produced by the increased lactate may contribute to proton modulation of the NMDA receptor and NMDA-induced cell injury or death.  相似文献   

7.
Electrically coupled pairs of cultured rat glomus cells were used. In one group of experiments, both cells were current-clamped. Delivery of positive or negative pulses to Cell 1 elicited appreciable voltage noise in this cell and large action potentials (probably Ca2+ spikes) in about 10% of them. Both passive and active electrical events spread to Cell 2, presumably through the gap junctions between them. The coupling coefficient (Kc) was larger for the spikes than for non-regenerative voltage noise. In another group of experiments, Cell 1 was current-clamped and Cell 2 was voltage-clamped at Cell 1 EM. Pulses of either polarity, delivered to Cell 1, produced current flow through the intercellular junction and allowed direct measurement of junctional currents (Ij) and total conductances (Gj). Ij had a mean value of about 12.5 pA and Gj of 391 pS. Unitary (presumably single channel) conductance (gj) was about 78 pS.  相似文献   

8.
Interregional differences in intracellular pH (pHi) in brain tissue, and its regulation following 1 and 5 h of respiratory alkalosis (with and without hypoxemia) were determined in N2O anesthetized dogs. Two techniques for pHi estimation were used (Tco2 and14C-DMO) and included corrections for measured extracellular fluid (35SO42−) space (ECS). Cortical pHi by the two techniques agreed closely in control and in 3 of the 4 experimental conditions, suggesting: (a) our estimation of extracellular fluid (ECF) [HCO3] from measured CSF [HCO3] was a valid ssumption; and (b) our method had sufficient resolution to determine the magnitude of brain pHi regulation during respiratory acid-base disturbances.When moderate normoxic respiratory alkalosis (PaCO2 ≈ 25 mm Hg) was imposed for 5 h, pHi (in most brain regions) was well regulated and always exceeded the incomplete regulation noted in bulk CSF. When moderate hypoxemia (PaO2 ≈ 45 mm Hg) accompanied hypocapnia, pHi was more closely regulated during the early phase (1 h) of respiratory alkalosis.Increased levels of metabolic acids (especially lactic acid) were critical to brain pHi regulation during the initial hour of respiratory alkalosis and accounted for much of the independent effect of hypoxemia on pHi regulation. However, these metabolic acids remained unchanged as pHi was more completely regulated between 1 and 5 h of continued hypocapnia or hypoxic hypocapnia. This time-dependent tregulation of pHi may involve some regulatory role for changed transmembrane fluxes of H+ and/or HCO3.Significant interregional differences were observed in both pHi and in ECS; with tendencies toward more alkaline pHi and lower ECS in brain stem and white matter. With respiratory alkalosis ECS fell and intracellular fluid increased in both cortex and caudate nucleus, possibly reflecting an osmotic effect of increased metabolic acid levels or reduction in cell membrane ion pumping.  相似文献   

9.
The objective of this study was to assess the influence of Ca2+ influx on intracellular pH (pHi) of neocortical neurons in primary culture. Neurons were exposed to glutamate (100–500 μM) or KCl (50 mM), and pHi was recorded with microspectroflurometric techniques. Additional experiments were carried out in which calcium influx was triggered by ionomycin (2 μM) or the calcium ionophore 4-Br-A23187 (2 μM). Glutamate exposure either caused no, or only a small decrease in pHi (ΔpH ≈ 0.06 units). When a decrease was observed, a rebound rise in pHi above control was observed upon termination of glutamate exposure. In about 20% of the cells, the acidification was more pronounced (ΔpH ≈ 0.20 units), but all these cells had high control pHi values, and showed gradual acidification. Exposure of cells to 50 mM KCl consistently increased pHi. Since this increase was similar in the presence and nominal absence of HCO3, it probably did not reflect influx of HCO3 via a Na+-HCO3 symporter. Furthermore, since it occurred in the absence of external Ca2+ (or a measurable rise in Cai2+) it seemed independent of Ca2+ influx. It is tentatively concluded that the rise in pHi was due to reduced passive influx of H+ along the electrochemical gradient, which is reduced by depolarization. In Ca2+-containing solutions, depolarization led to a rebound increase in pHi above control. This, and the rebound found after glutamate transients, may reflect Ca2+-triggered phosphorylation and upregulation of the Na+/H+ antiporter which extrudes H+ from the cell. Ionomycin and 4-Br-A23187 gave rise to a large rise in Cai2+ and to alkalinization of the cell (ΔpH ≈ 0.5). Since amiloride or removal of Na+ from the external solution did not alter the rise in pHi, it was probably not due to accelerated H+ extrusion. However, removal of Ca2+ from extracellular fluid prevented the rise, suggesting that it was secondary to Ca2+/2H+ exchange across plasma membranes.  相似文献   

10.
The effect of different neurotransmitters on the intracellular pH (pHi) and intracellular calcium (Ca2+i) was studied in cultured astrocytes from neonatal rat cerebellum, using the fluorescent dyes 2,7′-bis(carboxyethyl)-5,6-carboxy-fluorescein (BCECF) and Fura-2. Application of glutamate or kainate (100 μM) in a HEPES-buffered, CO2/HCO3?-free saline induced a decrease in pHi and an increase in Ca2+i. Amplitude and time course of the pHi and Ca2+i transients were different. Glutamate and kainate evoked a mean acidification of 0.22 ± 0.05 (n = 29) and 0.20 ± 0.04 (n = 12) pH units, respectively. The changes in pHi and Ca2+i induced by kainate, but not by glutamate, were inhibited by 6-cyano-7-dinitroquinozalin-2,3-dion (CNQX; 50 μM). In order to elucidate the mechanism of the agonist-induced acidification, whether the pHi changes were secondary to the Ca2+ rises was tested. In the absence of extracellular Ca2+, the kainate-induced Ca2+i transient was suppressed, while the intracellular acidification was only reduced by 13%. Removal of extracellular Ca2+ reduced the glutamate-induced pHi change by 8%, while the second component of the Ca2+i transient was abolished. Application of trans-(±)-1-amino-(1S,3R)-cyclopentadicarboxylic acid (t-ACPD, 100 μM), a metabotropic glutamate receptor agonist, and of noradrenaline (20 μM) evoked a Ca2+i increase, but no change of pHi. D-aspartate, which has a low affinity to glutamate receptors, but is known to be transported by the glutamate uptake system in some astrocytes, evoked an intracellular acidification, similar to that induced by glutamate, but no Ca2+i transient. The results suggest that the kainate-induced acidification is only partly due to the concomitant Ca2+i rise, while the glutamate/aspartate-induced acidification is mainly due to the activation of the glutamate uptake system. © 1995 Wiley-Liss, Inc.  相似文献   

11.
To test the hypothesis whether CO2–HCO3 buffer is essential for the expression of chemoreception and to distinguish between pHi and pHo interaction with pCO in the carotid chemosensory response, we superfused-perfused in vitro cat carotid bodies using HEPES-Tyrode's solution with and without CO2–HCO3, and compared the responses at the same pHo in the absence and presence of light. In the absence of light, pCO (>138 Torr) stimulated the carotid body chemoreceptors in CO2–HCO3 buffer at pHo of 7.40, whereas pCO (69–550 Torr) did not stimulate the neural discharge in HEPES buffer at the pHo of 7.4–7.1 but did so below pHo 7.1. In the presence of light, all the responses were diminished proportionately.  相似文献   

12.
The clearance of extracellular glutamate is mainly mediated by pH‐ and sodium‐dependent transport into astrocytes. During hepatic encephalopathy (HE), however, elevated extracellular glutamate concentrations are observed. The primary candidate responsible for the toxic effects observed during HE is ammonium (NH4+/NH3). Here, we examined the effects of NH4+/NH3 on steady‐state intracellular pH (pHi) and sodium concentration ([Na+]i) in cultured astrocytes in two different age groups. Moreover, we assessed the influence of NH4+/NH3 on glutamate transporter activity by measuring D ‐aspartate‐induced pHi and [Na+]i transients. In 20–34 days in vitro (DIV) astrocytes, NH4+/NH3 decreased steady‐state pHi by 0.19 pH units and increased [Na+]i by 21 mM. D ‐Aspartate‐induced pHi and [Na+]i transients were reduced by 80–90% in the presence of NH4+/NH3, indicating a dramatic reduction of glutamate uptake activity. In 9–16 DIV astrocytes, in contrast, pHi and [Na+]i were minimally affected by NH4+/NH3, and D ‐aspartate‐induced pHi and [Na+]i transients were reduced by only 30–40%. Next we determined the contribution of Na+, K+, Cl?‐cotransport (NKCC). Immunocytochemical stainings indicated an increased expression of NKCC1 in 20–34 DIV astrocytes. Moreover, inhibition of NKCC with bumetanide prevented NH4+/NH3‐evoked changes in steady‐state pHi and [Na+]i and attenuated the reduction of D ‐aspartate‐induced pHi and [Na+]i transients by NH4+/NH3 to 30% in 20–34 DIV astrocytes. Our results suggest that NH4+/NH3 decreases steady‐state pHi and increases steady‐state [Na+]i in astrocytes by an age‐dependent activation of NKCC. These NH4+/NH3‐evoked changes in the transmembrane pH and sodium gradients directly reduce glutamate transport activity, and may, thus, contribute to elevated extracellular glutamate levels observed during HE. © 2008 Wiley‐Liss, Inc.  相似文献   

13.
The intracellular Ca2+ (Ca2+i) and the intracellular pH (pHi) were measured in identified neuropile glial cells in the central nervous system of the leech Hirudo medicinalis, using the fluorescent dye fura-2, and double-barrelled, neutral carrier, pH-sensitive microelectrodes. Different stimuli were used to elicit Ca2+i and/or pHi changes, such as application of ammonium, high external K+-concentration, and low external pH. Ammonium (20 mM) and high external K+ (20 mM), which depolarized the glial membrane by 20–30 mV, evoked rapid and large rises of Ca2+i. In contrast to the Ca2i changes, amplitude and direction of the pHi changes were dependent on the presence of CO2/HCO3? in the saline. The addition of CO2/HCO3?, and the subsequent reduction of external pH from 7.4 to 7.0, had no effect on Ca2+i, but caused significant changes of pHi. The results suggest that the ammonium- and K+-induced Ca2+i rises were due to the membrane depolarization leading to a Ca2+ influx through voltage-gated Ca2+ channels in the glial membrane, while the pHi changes resulted from movements of ammonia and from the activation or inhibition of the Na+-HCO3? cotransporter. This indicates that changes of intracellular Ca2+ and pH can occur independently of each other, suggesting that the homeostasis of these ions is not necessarily interrelated in these glial cells.  相似文献   

14.
Clustered and isolated glomus cells, cultured from rat carotid bodies, were exposed to hypoxia (pO2 2–30 torr) induced by applications of sodium-dithionite (Na2S2O4). Hypoxia decreased or increased intracellular pH (pHi) of clustered cells about equally, but lowered it in most isolated cells. The levels of intracellular acidification were similar in both groups whereas alkalinization was more pronounced in the clusters. The H+ equilibrium potential (EH) and its changes during hypoxia (ΔEH), were determined almost exclusively by pHi. Seventy-five percent of clustered cells became depolarized whereas 80% of isolated cells underwent hyperpolarization. In both groups, changes in the resting potential (ΔEM) were directly and significantly correlated with ΔEH, thus ΔpHi. These observations support the view that clustered and isolated rat glomus cells behave differently. This difference may occur because of the presence (in the clusters) or absence (in isolated cells) of enveloping sustentacular cell processes.  相似文献   

15.
Conduction in inhomogeneous axons may be blocked by several mechanisms. Conduction in demyelinated axons may fail since normal internodal membrane is inexcitable, because values of sodium conductance are too low to support impulse conduction. In addition, focal loss of myelin causes increased current leakage which slows or blocks invasion of impulses into the demyelinated zone due to inadequate current density. Similar considerations apply to the invasion of non-myelinated preterminal axons from myelinated parent fibers, where conduction can be blocked as a result of inadequate current density. A cable model of an axon is presented which allows myelinated regions, regions without myelin, and variable length transition zones of redistributed channel densities, to be studied. Action potentials and membrane currents were studied. Computer simulations using this model show that the safety factor for invasion is dependent on temperature. These studies also show that small changes in axon membrane properties, at the transition region between the myelinated zone and the region without myelin, may promote invasion of the region without myelin. In particular, increasing sodium conductance (gNa) or decreasing potassium conductance (gK) promotes invasion. Because of the non-linear behavior of excitable membranes the spatial distribution of channels is shown also to have significant effects on invasion. Thus, relatively small degrees of membrane reorganization may lead to functional changes with respect to the invasion of demyelinated axon regions. Similarly, the properties of the heminode at the distal part of the parent myelinated fiber may determine the invasion characteristics of non-myelinated terminal axons.  相似文献   

16.
We used the fluorescent pH-sensitive dye 2′,7′-bis(carboxyethyl)-5,6-carboxyfluorescein (BCECF) to monitor intracellular pH (pHi) in single astrocytes cultured from the forebrain of neonatal rats. When exposed to a nominally CO2/HCO3? -free medium buffered to pH 7.40 with HEPES at 37°C, the cells had a mean pHi of 6.89. Switching to a medium buffered to pH 7.40 with 5% CO2 and 25 mM HCO3? caused the steady-state pHi to increase by an average of 0.35, suggesting the presence of a HCO3? -dependent acid-extrusion mechanism. The sustained alkalinization was sometimes preceded by a small transient acidification. In experiments in which astrocytes were exposed to nominally HCO3?-free (HEPES-buffered) solutions, the application and withdrawal of 20 mM extracellular NH4+ caused pHi to fall to a value substantially below the initial one. pHi spontaneously recovered from this acid load, stabilizing at a value ~ 0.1 higher than the one prevailing before the application of NH4+. In other experiments conducted on cells bathed in HEPES-buffered solutions, removing extracellular Na+ caused pHi to decrease rapidly by 0.5. Returning the Na+ caused pHi to increase rapidly, indicating the presence of an Na+-dependent/HCO3?-independent acid-extrusion mechanism; the final pHi after returning Na+ was ~ 0.08 higher than the initial value. This pHi recovery elicited by returning Na+ was not substantially affected by 50 μM ethylisopropylamiloride (EIPA), but was speeded up by 50 μM 4,4′-diisothiocyanostilbene-2,2′-disulfonate (DIDS). Increasing [K+]? from 5 to 25 mM caused pHi to increase reversibly by ~ 0.2 in nominally CO2/HCO3?-free solutions, and by ~ 0.1 in CO2/HCO3?-containing solutions, although the initial pHi was ~ 0.17 higher in the presence of CO2/HCO3-. These results suggest the presence of a depolarization-induced alkalinization. Our results suggest the presence of both HCO3? dependent and -independent acid-base transport systems in cultured mammalian astrocytes, and indicate that astrocyte pHi is sensitive to changes in either membrane voltage or [K+]0 per se. © 1993 Wiley-Liss, Inc.  相似文献   

17.
We examined H+ and HCO3? transport mechanisms that are involved in the regulation of intracellular pH of Schwann cells. Primary cultures of Schwann cells were prepared from the sciatic nerves of 1–3-day-old rats. pHi of single cells attached to cover slips was continuously monitored by measuring the absorbance spectra of the pH-sensitive dye dimethylcarboxyfluorescein incorporated intracellularly. The average pHi of neonatal Schwann cells bathed in HEPES mammalian solution was 7.17 ± 0.02 (n = 32). In the nominal absence of HCO3?, pHi spontaneously recovered from an acute acid load induced by exposing the Schwann cells to 20 mM NH4+ (NH4+ prepulse). This pHi recovery from the acute acid load was totally inhibited in the absence of external Na+ or in the presence of 1 mM amiloride. In both cases, the pHi recovery was readily restored upon readdition of external Na+ or removal of amiloride. In the steady-state, addition of amiloride caused a small and slow decrease in pHi which was readily reversed upon removal of amiloride. In the presence of HCO3?, removal of external Cl- caused pHi to rapidly and reversibly increase by 0.23 = 0.03 (n = 15) and the initial rate of alkalinization was 20.6 ± 2.7 × 10-4 pH/sec. In the absence of external Na+, removal of bath Cl? still caused pHi to increase by 0.15 ± 0.02 and the initial rate of pHi increase was not significantly altered. In the nominal absence of HCO3?, removal of bath Cl- caused pHi to increase very slightly (0.05 ± 0.01) with an initial dpHi/dt of only 4.4 ± 0.2 × 10?4 pH/sec (n = 4). Addition of 100 μM DIDS did not inhibit the pHi increase caused by removal of bath Cl?. These data indicate that (1) Rat Schwann cells regulate their pHi via an Na-H exchange mechanism which is moderately active at steady-state pHi. (2) In the presence of HCO3?, there is a Na-independent Cl-HCO3 (base) exchanger which also contributes to regulation of intracellular pH in Schwann cells. © 1994 Wiley-Liss, Inc.  相似文献   

18.
Both intracellular pH (pHi) and synaptic cleft pH change during neuronal activity yet little is known about how these pH shifts might affect synaptic transmission by influencing vesicle fusion. To address this we imaged pH‐ and Ca2+‐sensitive fluorescent indicators (HPTS, Oregon green) in boutons at neuromuscular junctions. Electrical stimulation of motor nerves evoked presynaptic Ca2+i rises and pHi falls (~0.1 pH units) followed by recovery of both Ca2+i and pHi. The plasma‐membrane calcium ATPase (PMCA) inhibitor, 5(6)‐carboxyeosin diacetate, slowed both the calcium recovery and the acidification. To investigate a possible calcium‐independent role for the pHi shifts in modulating vesicle fusion we recorded post‐synaptic miniature end‐plate potential (mEPP) and current (mEPC) frequency in Ca2+‐free solution. Acidification by propionate superfusion, NH4+ withdrawal, or the inhibition of acid extrusion on the Na+/H+ exchanger (NHE) induced a rise in miniature frequency. Furthermore, the inhibition of acid extrusion enhanced the rise induced by propionate addition and NH4+ removal. In the presence of NH4+, 10 out of 23 cells showed, after a delay, one or more rises in miniature frequency. These findings suggest that Ca2+‐dependent pHi shifts, caused by the PMCA and regulated by NHE, may stimulate vesicle release. Furthermore, in the presence of membrane permeant buffers, exocytosed acid or its equivalents may enhance release through positive feedback. This hitherto neglected pH signalling, and the potential feedback role of vesicular acid, could explain some important neuronal excitability changes associated with altered pH and its buffering. Synapse 67:729–740, 2013 . © 2013 Wiley Periodicals, Inc.  相似文献   

19.
Brain ischemia is accompanied by lowering of intra- and extracellular pH. Stroke often leads to irreversible damage of synaptic transmission by unknown mechanism. We investigated an influence of lowering of pHi and pHo on free radical formation in synaptosomes. Three models of acidosis were used: (1) pHo 6.0 corresponding to pHi decrease down to 6.04; (2) pHo 7.0 corresponding to the lowering of pHi down to 6.92: (3) 1 mM amiloride corresponding to pHi decrease down to 6.65. We have shown that both types of extracellular acidification, but not intracellular acidification, increase 2′,7′-dichlorodihydrofluorescein diacetate fluorescence that reflects free radical formation. These three treatments induce the rise of the dihydroethidium fluorescence that reports synthesis of superoxide anion. However, the impact of amiloride on superoxide anion synthesis was less than that induced by moderate extracellular acidification. Superoxide anion synthesis at pHo 7.0 was almost completely eliminated by mitochondrial uncoupler carbonyl cyanide 4-(trifluoromethoxy)phenylhydrazone. Furthermore, using fluorescent dyes JC-1 and rhodamine-123, we confirmed that pHo lowering, but not intracellular acidification, led to depolarization of intrasynaptosomal mitochondria. We have shown that pHo but not pHi lowering led to oxidative stress in neuronal presynaptic endings that might underlie the long-term irreversible changing in synaptic transmission.  相似文献   

20.
Membrane cable properties of skeletal muscle fibers of dystrophic mice (Rej-129) and their littermate controls were examined using a conventional two-microelectrode recording technique. Fibers from dystrophic mice had a decreased membrane resistivity (Rm) compared with those from normal mice (517 ± 27 vs 642 ± 34 Ω ? cm2), while the internal resistivities (Ri) did not differ significantly. The increase in membrane specific conductance was due to an increased Cl? conductance (gCl) (2304 vs 1346 μS/cm2 for normal fibers), although the K+ conductance (gK) was actually decreased (234 vs 369 μS/cm2 for normal fibers). With changes in pH, membrane conductances of normal and dystrophic skeletal muscle fibers varied differently, mainly due to differences in effects on the Cl? conductance. This contrast may be due to altered regulation of internal pH in dystrophic muscle.  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