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.
We studied the effect of insulin and lavendustin-A (a tyrosine kinase inhibitor) on the short-circuit current (ISC) of primary cultures of fetal distal rat lung epithelium (FDLE). Insulin (2 microM) on the basolateral side of the monolayer increased ISC from 5.76 +/- 0.83 microA/cm2 (SEM, n = 7) to 7.23 +/- 1.00 microA/cm2 (p less than 0.01) under control conditions, and from 1.00 +/- 0.31 microA/cm1 to 1.53 +/- 0.34 microA/cm2 (p less than 0.05, n = 4) when amiloride (10 microM) was present on the apical side of the monolayer. Thus insulin increased both the amiloride-sensitive and insensitive ISC with the insulin-induced increase in ISC in the absence of amiloride (1.47 +/- 0.22 microA/cm2, n = 7) being significantly larger than that in the presence of 10 microM amiloride (0.53 +/- 0.14 microA/cm2, n = 4; p less than 0.025). Insulin's effect reached steady state in 1 hr. Lavendustin-A (10 microM), a tyrosine kinase inhibitor, applied to the apical side of the monolayer attenuated but did not completely block insulin's ability to increase in ISC; i.e., insulin increased ISC in lavendustin-A treated monolayers (0.63 +/- 0.09 microA/cm2, n = 5; p less than 0.0025) but the increase was significantly smaller than that without the pretreatment of lavendustin-A (p less than 0.05). In the presence of amiloride (10 microM) and lavendustin-A (10 microM) insulin was no longer able to increase ISC (change in ISC = 0.04 +/- 0.03 microA/cm2, n = 6), suggesting that lavendustin-A had blocked the insulin's effect on the amiloride-insensitive ISC. Lavendustin-A (10 microM) had no significant effect on the basal ISC in control and amiloride treated monolayers. Our studies demonstrate that insulin increases amiloride-insensitive ISC in FDLE via lavendustin-A sensitive tyrosine kinase and that insulin's action on the amiloride-sensitive ISC of FDLE is mediated through a lavendustin-A insensitive (and presumably tyrosine kinase-independent) pathway.  相似文献   

2.
Modulation of cardiac Na channels by angiotensin II   总被引:2,自引:0,他引:2  
The modulation of Na channels by the vasoactive peptide angiotensin II (AT II) has been studied in isolated ventricular cells of guinea pigs using the patch clamp technique. In cell-attached patches the maximal probability of the channel being open was increased in a concentration range between 0.05 and 1 microM, but decreased at higher concentrations. A maximal increased of 2.5 +/- 0.86 was found at 1 microM AT II. The increase in the probability of the channel being open was due to a decrease in the number of nulls. In all affected cells (n = 17) we observed a delayed inactivation after application of AT II at concentrations between 0.05 and 10 microM. At -30 mV, the time constant of inactivation increased from 1.1 +/- 0.1 ms (controls) to 5.6 +/- 1.6 ms (10 microM AT II). This effect was due to an increased number of openings per sweeps. No significant effect on the mean open time and the first latency were observed. However, due to pronounced bursting, the averaged closed time was significantly increased from 0.8 +/- 0.1 ms to 1.3 +/- 0.1 ms in the presence of 1 microM AT II at -30 mV. An effect of AT II on cardiac Na channels via protein kinase C is discussed.  相似文献   

3.
The effects of H(2)O(2) on pacemaker activity and underlying membrane currents were studied in isolated rabbit sinoatrial (SA) node cells using perforated patch current- and voltage-clamp methods. Short-term exposure (<10 min) of the nodal cells to H(2)O(2) (200 microM) resulted in an initial shortening of spontaneous action potential cycle length (from 445 +/- 60 to 398 +/- 56 ms; P < 0.05) and a prolongation of action potential duration. H(2)O(2) (100 microM) significantly increased peak L-type Ca(2+) current (I(Ca,L)) from -384 +/- 77 to -439 +/- 84 pA (116 +/- 2%, n = 6). Additionally, the persistent or non-inactivating component of I(Ca,L) was increased from -52 +/- 3 to -88 +/- 14 pA (174 +/- 19%, n = 6). The hyperpolarization-activated current (I(f)) was decreased from -228 +/- 62 to -161 +/- 72 pA after exposure to H(2)O(2) (n = 7). There were no changes in the delayed rectifier K(+) current (I(K)) (n = 7). H(2)O(2)-induced Ca(2+) currents were blocked by 2 microM nicardipine (n = 6), 2 mM Ni(2+) (n = 2), and the protein kinase C (PKC) inhibitor bisindolylmaleimide (10(-7) M; n = 4) but not by 20 microM tetrodotoxin. These results suggest that H(2)O(2) can increase the spontaneous pacing rate in rabbit SA node cells by enhancing I(Ca,L) and that this effect is mediated by a PKC-dependent pathway.  相似文献   

4.
Vasoactive intestinal peptide (VIP) stimulates active Cl- secretion by the intestinal epithelium, a process that depends upon the maintenance of a favorable electrical driving force established by a basolateral membrane K+ conductance. To demonstrate the role of this K- conductance, we measured short-circuit current (I(SC)) across monolayers of the human colonic secretory cell line, T84. The serosal application of VIP (50 nM) increased I(SC) from 3 +/- 0.4 microA/cm2 to 75 +/- 11 microA/cm2 (n = 4), which was reduced to a near zero value by serosal applications of Ba2+ (5 mM). The chromanol, 293B (100 microM), reduced I(SC) by 74%, but charybdotoxin (CTX, 50 nM) had no effect. We used the whole-cell voltage-clamp technique to determine whether the K+ conductance is regulated by cAMP-dependent phosphorylation in isolated cells. VIP (300 nM) activated K+ current (131 +/- 26 pA, n = 15) when membrane potential was held at the Cl- equilibrium potential (E(Cl-) = -2 mV), and activated inward current (179 +/- 28 pA, n = 15) when membrane potential was held at the K+ equilibrium potential (E(K+) = -80 mV); however, when the cAMP-dependent kinase (PKA) inhibitor, PKI (100 nM), was added to patch pipettes, VIP failed to stimulate these currents. Barium (Ba2+ , 5 mM), but not 293B, blocked this K+ conductance in single cells. We used the cell-attached membrane patch under conditions that favor K + current flow to demonstrate the channels that underlie this K+ conductance. VIP activated inwardly rectifying channel currents in this configuration. Additionally, we used fura-2AM to show that VIP does not alter the intracellular Ca2+ concentration, [Ca2 +]i. Caffeine (5 mM), a phosphodiesterase inhibitor, also stimulated K+ current (185 +/- 56 pA, n = 8) without altering [Ca2+]i. These results demonstrate that VIP activates a basolateral membrane K+ conductance in T84 cells that is regulated by cAMP-dependent phosphorylation.  相似文献   

5.
Previous studies have suggested that heterotrimeric G(i) proteins, Src tyrosine kinase and phosphatidylinositol-3 kinase (PI3 Kinase) are involved in signaling events induced by lipopolysaccharide (LPS) leading to pro-inflammatory cytokines gene expression. To investigate the involvement of these mediators in Gram-positive bacteria induced pro-inflammatory cytokine expression, LPS (10 ng/ml), heat killed group B Streptococci (GBS 1 microg/ml) and Staphylococcus aureus (SA 10 microg/ml) were used to induce TNFalpha production in the murine J774A.1 macrophage (M?) cell line and human promonocytic THP-1 cell line. Pertussis toxin (PTx, 1 microg/ml), an inhibitor of G(i) protein; pyrazolopyrimidine-2 (PP2, 1 or 25 microM), a Src tyrosine kinase inhibitor; and LY294002 (100 nM), an inhibitor of PI3 Kinase were used to examine the involvement of G(i), Src tyrosine kinase and PI3 Kinase, respectively, in TNFalpha production. In J774A.1 cells, pretreatment with PTx and PP2 attenuated TNFalpha production induced by LPS (60+/-9% and 81+/-11% inhibition, n=3, p<0.05, respectively), GBS (95+/-1% and 80+/-6% inhibition, n=3, p<0.05, respectively) and SA (51+/-18% and 68+/-16% inhibition, n=4, p<0.05, respectively). However, pretreatment with LY 294002 inhibited LPS induced TNFalpha production (82+/-13% inhibition, n=3, p<0.05), but did not inhibit GBS or SA induced TNFalpha production. In THP-1 cells, pretreatment with PTx, PP2 and LY 294002 inhibited TNFalpha production induced by LPS (84+/-3%, 59+/-12% and 84+/-4% inhibition, n=3, p<0.05, respectively) and SA (56+/-7%, 87+/-1% and 35+/-6% inhibition, n=3, p<0.05, respectively). These data support our hypothesis that G(i)-coupled and Src tyrosine kinase-coupled signaling pathways are involved in both Gram-negative and Gram-positive bacteria induced pro-inflammatory cytokine expression. However, unlike LPS, involvement of PI3 Kinase in Gram-positive bacteria induced signaling pathways are species dependent.  相似文献   

6.
1. Conventional microelectrode techniques were used to measure simultaneous changes in membrane potential (Vm) and conductance (Gm) induced by single electrical stimuli in muscles bathed in Cl(-)-free solution containing 40 mM of tetraethylammonium (TEA+). 2. Stimulation induced slow transient depolarizations (slow response) accompanied by increased calcium conductance, while the potassium conductance was first elevated and later reduced. 3. Stepwise elevation of [K+]0 from 2.5 to 5 or 10 mM during the slow response evoked an abrupt repolarization of 42.3 +/- 8.9 mV (n = 4; p less than 0.001), and 24.8 +/- 3.5 +/- mV (n = 5; p less than 0.001), respectively, while Gm was increased to 1.45 +/- 0.25-fold (n = 5; p less than 0.05). Neither the slow response nor K(+)-induced changes in Vm or Gm were sensitive to tetrodotoxin (3 microM), however, nifedipine (10 microM) abolised the slow response. 4. It was concluded that beyond the increase of calcium conductance, the ionic conductance of the inward rectifier K+ channel was reduced during the slow response, which could be restored by the elevation of [K+]0. The results suggest the possible contribution of these mechanisms to the electrical instability of myotonic muscles. Potential therapeutic consequences are discussed.  相似文献   

7.
We examined ion channels derived from a chloroform extract of isolated, dehydrated rat liver mitochondria. The extraction method was previously used to isolate a channel-forming complex containing poly-3-hydroxybutyrate and calcium polyphosphate from Escherichia coli. This complex is also present in eukaryotic membranes, and is located primarily in mitochondria. Reconstituted channels showed multiple subconductance levels and were voltage-dependent, showing an increased probability of higher conductance states at voltages near zero. In symmetric 150 mM KCl, the maximal conductance of the channel ranged from 350 pS to 750 pS. For voltages >+/-60 mV, conductance fluctuated in the range of approximately 50- approximately 200 pS. In the presence of a 1:3 gradient of KCl, at pH = 7.4, selectivity periodically switched between different states ranging from weakly anion-selective (V(rev) approximately -15 mV) to ideally cation-selective (V(rev) approximately +29 mV), without a significant change in its conductance. Overall, the diverse, but highly reproducible, channel activity most closely resembled the behavior of the permeability transition pore channel seen in patch-clamp experiments on native mitoplasts. We suggest that the isolated complex may represent the ion-conducting module from the permeability transition pore.  相似文献   

8.
三羟异黄酮对豚鼠心室肌细胞L-型钙通道电流的影响   总被引:2,自引:0,他引:2  
Ji ES  Yin JX  Ma HJ  He RR 《生理学报》2004,56(4):466-470
本实验用全细胞膜片钳技术观察三羟异黄酮(genistein,GST)对豚鼠心室肌细胞L-钙通道电流(ICa、L)的影响。结果如下:(1)GST(10、50、100 μmol/L)可浓度依赖性地降低ICa,L(n=6,P<0.01)。GST的非活性结构类似物daidzein(100μmol/L),在同一浓度范围对ICa,L没有影响(n=5,P>0.05)。(2)GST使I-V曲线上移,但对ICa,L的电压依赖特征和最大激活电压无明显影响。(3)GST对ICa,L的激活动力学特性也无影响,但可使钙电流稳态失活曲线左移。V0.5从对照的-28.6±0.6 mV变为-32.8±1.1mV,κ值从对照的5.8±0.5 mV升至6.5±0.9 mV(n=6,P<0.05)。(4)GST明显使复活曲线右移,从而使ICa,L从失活状态下恢复明显减慢(n=7,P<0.01)。(5)酪氨酸磷酸酶抑制剂正钒酸钠(1 mmol/L)显著对抗GST引起的ICa,L抑制效应(n=6,P<0.01)。根据以上结果得出的结论是:GST抑制ICa,L加速钙通道失活和钙通道在失活状态下恢复减慢;GST对ICa,L的这种抑制作用与蛋白酪氨酸激酶(PTK)抑制有关。  相似文献   

9.
The voltage- and Ca2+-dependent gating mechanism of large-conductance Ca2+-activated K+ (BK) channels from cultured rat skeletal muscle was studied using single-channel analysis. Channel open probability (Po) increased with depolarization, as determined by limiting slope measurements (11 mV per e-fold change in Po; effective gating charge, q(eff), of 2.3 +/- 0.6 e(o)). Estimates of q(eff) were little changed for intracellular Ca2+ (Ca2+(i)) ranging from 0.0003 to 1,024 microM. Increasing Ca2+(i) from 0.03 to 1,024 microM shifted the voltage for half maximal activation (V(1/2)) 175 mV in the hyperpolarizing direction. V(1/2) was independent of Ca2+(i) for Ca2+(i) < or = 0.03 microM, indicating that the channel can be activated in the absence of Ca2+(i). Open and closed dwell-time distributions for data obtained at different Ca2+(i) and voltage, but at the same Po, were different, indicating that the major action of voltage is not through concentrating Ca2+ at the binding sites. The voltage dependence of Po arose from a decrease in the mean closing rate with depolarization (q(eff) = -0.5 e(o)) and an increase in the mean opening rate (q(eff) = 1.8 e(o)), consistent with voltage-dependent steps in both the activation and deactivation pathways. A 50-state two-tiered model with separate voltage- and Ca2+-dependent steps was consistent with the major features of the voltage and Ca2+ dependence of the single-channel kinetics over wide ranges of Ca2+(i) (approximately 0 through 1,024 microM), voltage (+80 to -80 mV), and Po (10(-4) to 0.96). In the model, the voltage dependence of the gating arises mainly from voltage-dependent transitions between closed (C-C) and open (O-O) states, with less voltage dependence for transitions between open and closed states (C-O), and with no voltage dependence for Ca2+-binding and unbinding. The two-tiered model can serve as a working hypothesis for the Ca2+- and voltage-dependent gating of the BK channel.  相似文献   

10.
Addition of 10(-5) M amphotericin B to the tear solution of an in vitro preparation of the frog cornea increased the transepithelial conductance, gt, and decreased the apical membrane fractional resistance, f(R0), in the presence or absence of tear Na+ and Cl-. In the presence of tear Na+ and Cl-, amphotericin B increased the short-circuit current, Isc, from 3.9 to 8.8 microA.cm-2 and changed the intracellular potential, V0, from -48.5 to -17.9 mV probably due to a higher increase in the Na+ than in the K+ conductance. In the absence of tear Na+ and Cl-, amphotericin B decreased Isc from 5.5 to about 0 microA.cm-2 due to K+ (and possibly Na+) flux from cell to tear and changed V0 from -35.4 to -63.6 mV due to the increase in conductance of both ions. Increase in the tear K+ from 4 to 79 mM (in exchange for choline), in the presence of amphotericin B and absence of tear Na+ and Cl-, decreased f(R0) from 0.09 to 0.06, increased gt from 0.23 to 0.31 mS, increased Isc from 0.63 to 7.3 microA.cm-2, and changed V0 from -65.5 to -17.3 mV due to the change in EK in the presence of a high conductance in the tear membrane. Similar effects were observed with an increase of tear Na+. Results support the concept that the Na+ conductance opened by amphotericin B in the apical membrane is greater than the K+ conductance. Previously observed transepithelial effects of the ionophore may be explained mostly on the basis of its effect on the apical membrane.  相似文献   

11.
Epidermal growth factor (EGF) and transforming growth factor alpha bind to a common receptor at the cell surface. Both the affinity and the tyrosine protein kinase activity of the receptor are regulated by exogenous factors, such as platelet-derived growth factor. A protein kinase C-dependent (Ca2+/phospholipid-dependent enzyme) and independent regulatory mechanism have been described. The protein kinase C-dependent mechanism results in the inhibition of the affinity and tyrosine kinase activity of the EGF receptor. We describe in this report an alternative mechanism of regulation of the receptor that is mediated by sphingosine. Treatment of WI-38 human fetal lung fibroblasts with 5 microM sphingosine for 2 min at 37 degrees C caused a marked increase in the affinity of the EGF receptor. Similar results were obtained when isolated plasma membranes prepared from these cells were incubated with sphingosine. A stimulation of the EGF receptor tyrosine protein kinase activity was also observed after sphingosine-treatment of plasma membranes. Sphingosine caused a decrease in the Km for ATP and an increase in the Vmax for the tyrosine phosphorylation of a synthetic peptide substrate. Control experiments demonstrated that these actions of sphingosine were not secondary to the inhibition of protein kinase C. These data indicate that sphingosine causes the functional conversion of the EGF receptor into an activated state that expresses both a high affinity for EGF and an increased tyrosine kinase activity. We conclude that sphingosine is a bioactive molecule in human fibroblasts.  相似文献   

12.
The Kv1.5 K(+) channel is functionally altered by coassembly with the Kvbeta1.3 subunit, which induces fast inactivation and a hyperpolarizing shift in the activation curve. Here we examine kinase regulation of Kv1.5/Kvbeta1.3 interaction after coexpression in human embryonic kidney 293 cells. The protein kinase C inhibitor calphostin C (3 microM) removed the fast inactivation (66 +/- 1.9 versus 11 +/- 0.25%, steady state/peak current) and the beta-induced hyperpolarizing voltage shift in the activation midpoint (V(1/2)) (-21.9 +/- 1.4 versus -4.3 +/- 2.0 mV). Calphostin C had no effect on Kv1.5 alone with respect to inactivation kinetics and V(1/2). Okadaic acid, but not the inactive derivative, blunted both calphostin C effects (V(1/2) = -17.6 +/- 2.2 mV, 38 +/- 1.8% inactivation), consistent with dephosphorylation being required for calphostin C action. Calphostin C also removed the fast inactivation (57 +/- 2.6 versus 16 +/- 0.6%) and the shift in V(1/2) (-22.1 +/- 1.4 versus -2.1 +/- 2.0 mV) conferred onto Kv1.5 by the Kvbeta1.2 subunit, which shares only C terminus sequence identity with Kvbeta1. 3. In contrast, modulation of Kv1.5 by the Kvbeta2.1 subunit was unaffected by calphostin C. These data suggest that Kvbeta1.2 and Kvbeta1.3 subunit modification of Kv1.5 inactivation and voltage sensitivity require phosphorylation by protein kinase C or a related kinase.  相似文献   

13.
Dimerization of epidermal growth factor receptor (EGFR) leads to the activation of its tyrosine kinase. To elucidate whether dimerization is responsible for activation of the intracellular tyrosine kinase domain or just plays a role in the stabilization of the active form, the activated status of wild-type EGFR moiety in the heterodimer with kinase activity-deficient mutant receptors was investigated. The kinase activity of the wild-type EGFR was partially activated by EGF in the heterodimer with intracellular domain deletion (sEGFR) or ATP binding-deficient mutant (K721A) EGFRs, while the wild-type EGFR in the heterodimer of wild-type and phosphate transfer activity-deficient mutant receptor D813N could be fully activated. After treatment with EGF, the ATP binding affinity and the V(max) of the wild-type EGFR increased. In the presence of sEGFR, a similar increase in the affinity for ATP was observed, but V(max) did not change. A two-step activation mechanism for EGFR was proposed: upon binding of EGF, the affinity for ATP increased and then, as a result of interaction between the neighboring tyrosine kinase domain, V(max) increased.  相似文献   

14.
Treatment of A431 human epidermoid cells with epidermal growth factor (EGF; 20 nM) results in decreased proliferation. This is associated with blockage of the cells in the S and/or G2 phases of the cell cycle. We found that tyrphostin, a putative tyrosine kinase inhibitor, in the range of 50 to 100 microM, partially reversed the growth-inhibitory and cell cycle changes induced by EGF. By using high-pressure liquid chromatography with electrochemical detection, we found that tyrphostin was readily incorporated into A431 cells, reaching maximal levels within 1 h. Although tyrphostin (50 to 100 microM) had no effect on high-affinity binding of EGF to its receptor in A431 cells for up to 24 h, the compound partially inhibited EGF-stimulated EGF receptor tyrosine kinase activity. However, this effect was evident only after prolonged treatment of the cells (4 to 24 h) with the drug. When the peak intracellular concentration of tyrphostin occurred (1 h), no inhibition of tyrosine kinase activity was observed. After both 1 and 24 h, tyrphostin was a less effective inhibitor of tyrosine kinase activity than the potent tumor promoter 12-O-tetradecanoyl phorbol-13-acetate, which almost completely blocked EGF receptor autophosphorylation. On the basis of our data, we hypothesize that tyrphostin is not a competitive inhibitor of the EGF receptor tyrosine kinase in intact cells and that it functions by an indirect mechanism.  相似文献   

15.
Epidermal growth factor and insulin share many biological activities, including stimulation of cell proliferation, ion flux, glycolysis, fatty acid and glycogen synthesis, and activation of receptor-linked tyrosine kinase activity. In the kidney, insulin has been shown to regulate transport processes and inhibit gluconeogenesis in the proximal tubule. Since the kidney represents a major source of EGF, the present studies investigated whether proximal tubule contained EGF receptors, whether EGF receptors were localized to apical or basolateral membranes, and whether EGF receptor activation participated in the regulation of an important proximal tubule function, gluconeogenesis. Specific EGF receptors were demonstrated in the basolateral membrane of proximal tubule. Following incubation with 125I EGF, basolateral membranes demonstrated equilibrium binding at 4 degrees C and 23 degrees C. There was 78 +/- 2% specific binding (n = 13). The dissociation constant (Kd) was 1.5 x 10(-9) M and maximal binding was 44 fmol/mg protein. There was ninefold more specific binding to proximal tubule basolateral membrane than to brush border membrane. In basolateral, but not brush border membranes, EGF induced phosphorylation of the tyrosine residues of intrinsic membrane proteins, including a 170 kDa protein, corresponding to the EGF receptor. In the presence of the gluconeogenic substrates, alanine, lactate, and succinate, proximal tubule suspensions synthesized glucose. EGF inhibited glucose production in a concentration-dependent manner over a concentration range of 3 x 10(-11) to 3 x 10(-9) M. In addition, EGF inhibited angiotensin II-stimulated glucose production in the proximal tubule suspensions. EGF did not significantly increase net glucose metabolism nor decrease cellular ATP concentrations. Therefore, these studies demonstrated that rat proximal tubule contained specific receptors for EGF that were localized to the basolateral membrane and linked to tyrosine kinase activity. EGF significantly inhibited proximal tubule glucose production without significantly increasing net glucose consumption.  相似文献   

16.
Alterations in EGF receptor (EGFR) signaling occur in intestinal disorders associated with dysregulated epithelial transport. In the present study, we investigated a role for the EGFR in the chronic regulation of intestinal epithelial secretory function. Epithelial Cl(-) secretion was measured as changes in short-circuit current (Isc) across voltage-clamped monolayers of T84 cells in Ussing chambers. Acute treatment of T84 cells with EGF (100 ng/ml, 15 min) chronically enhanced Isc responses to a broad range of secretagogues. This effect was apparent within 3 h, maximal by 6 h, and sustained for 24 h after treatment with EGF. The Na+/K+/2Cl(-) cotransporter (NKCC1) inhibitor bumetanide (100 microM) abolished the effect of EGF, indicating increased responses are due to potentiated Cl(-) secretion. Neither basal nor agonist-stimulated levels of intracellular Ca2+ or PKA activity were altered by EGF, implying that the effects of the growth factor are not due to chronic alterations in levels of second messengers. EGF increased the expression of NKCC1 with a time course similar to that of its effects on Cl(-) secretion. This effect of EGF was maximal after 6 h, at which time NKCC1 expression in EGF-treated cells was 199.9 +/- 21.9% of that in control cells (n = 21, P < 0.005). EGF-induced NKCC1 expression was abolished by actinomycin D, and RT-PCR analysis demonstrated EGF increased expression of NKCC1 mRNA. These data increase our understanding of mechanisms regulating intestinal fluid and electrolyte transport and reveal a novel role for the EGFR in the chronic regulation of epithelial secretory capacity through upregulation of NKCC1 expression.  相似文献   

17.
Rabbit urethral smooth muscle cells were studied at 37 degrees C by using the amphotericin B perforated-patch configuration of the patch-clamp technique, using Cs(+)-rich pipette solutions. Two components of current, with electrophysiological and pharmacological properties typical of T- and L-type Ca(2+) currents, were recorded. Fitting steady-state inactivation curves for the L current with a Boltzmann equation yielded a V(1/2) of -41 +/- 3 mV. In contrast, the T current inactivated with a V(1/2) of -76 +/- 2 mV. The L currents were reduced by nifedipine (IC(50) = 225 +/- 84 nM), Ni(2+) (IC(50) = 324 +/- 74 microM), and mibefradil (IC(50) = 2.6 +/- 1.1 microM) but were enhanced when external Ca(2+) was substituted with Ba(2+). The T current was little affected by nifedipine at concentrations <300 nM but was increased in amplitude when external Ca(2+) was substituted with Ba(2+). Both Ni(2+) and mibefradil reduced the T current with an IC(50) = 7 +/- 1 microM and approximately 40 nM, respectively. Spontaneous electrical activity recorded with intracellular electrodes from strips of rabbit urethra consisted of complexes comprising a series of spikes superimposed on a slow spontaneous depolarization (SD). Inhibition of T current reduced the frequency of these SDs but had no effect on either the number of spikes per complex or the amplitude of the spikes. In contrast, application of nifedipine failed to significantly alter the frequency of the SD but reduced the number and amplitude of the spikes in each complex.  相似文献   

18.
Intramembrane charge movement was measured on skeletal muscle fibers of the frog in a single Vaseline-gap voltage clamp. Charge movements determined both under polarized conditions (holding potential, VH = -100 mV; Qmax = 30.4 +/- 4.7 nC/micro(F), V = -44.4 mV, k = 14.1 mV; charge 1) and in depolarized states (VH = 0 mV; Qmax = 50.0 +/- 6.7 nC/micro(F), V = -109.1 mV, k = 26.6 mV; charge 2) had properties as reported earlier. Linear capacitance (LC) of the polarized fibers was increased by 8.8 +/- 4.0% compared with that of the depolarized fibers. Using control pulses measured under depolarized conditions to calculate charge 1, a minor change in the voltage dependence (to V = -44.6 mV and k = 14.5 mV) and a small increase in the maximal charge (to Qmax = 31.4 +/- 5.5 nC/micro(F] were observed. While in most cases charge 1 transients seemed to decay with a single exponential time course, charge 2 currents showed a characteristic biexponential behavior at membrane potentials between -90 and -180 mV. The voltage dependence of the rate constant of the slower component was fitted with a simple constant field diffusion model (alpha m = 28.7 s-1, V = -124.0 mV, and k = 15.6 mV). The midpoint voltage (V) was similar to that obtained from the Q-V fit of charge 2, while the steepness factor (k) resembled that of charge 1. This slow component could also be isolated using a stepped OFF protocol; that is, by hyperpolarizing the membrane to -190 mV for 200 ms and then coming back to 0 mV in two steps. The faster component was identified as an ionic current insensitive to 20 mM Co2+ but blocked by large hyperpolarizing pulses. These findings are consistent with the model implying that charge 1 and the slower component of charge 2 interconvert when the holding potential is changed. They also explain the difference previously found when comparing the steepness factors of the voltage dependence of charge 1 and charge 2.  相似文献   

19.
The aim of this study was to explore the effect of prostaglandin E1 (PGE1) on the membrane current of whole-cell voltage-clamped NG108-15 neuroblastoma x glioma hybrid cells. Perforated patch was used. The membrane current at -70 mV (leakage current) and the current-voltage curve produced by ramp pulses from -70 to 0 mV were recorded; from the I-V curve, the conductance of the leakage current and its reversal potential were determined. Bath application of PGE1 (22 nM-3 microM) produced an inward current accompanied by a reversible conductance increase. The PGE1 effect varied greatly from cell to cell. In a group of 11 differentiated cells, the inward current induced by 0.2 microM PGE1 was on average 171.1 +/- 49.8 pA, the conductance increased 2.66 +/- 0.50-fold and the reversal potential shifted by + 13.2 +/- 4.0 mV. The average values observed with 22 nM and 3 microM PGE1 were similar. The cell-permeable cAMP analog CPT-cAMP (0.5 mM) acted like PGE1. In 9 out of 16 cells, the PGE1 effect did not disappear and was not even noticeably reduced when the NaCl in the bath was replaced by N-methyl-D-glucamine (NMDG). The PGE1 effect was also seen in Ca2(+)-free NMDG bath but vanished when NMDG was replaced by glucose. It is concluded that PGE1, probably acting via intracellular cAMP, opens non-selective cation channels with large pore diameters which allow the passage of big organic cations.  相似文献   

20.
(1) Na+ currents and Na+-current fluctuations were measured in myelinated frog nerve fibres at 15 degrees C during 7.7 ms depolarizations to V = 40, 60 and 80 mV. (2) The conductance gamma of a single Na+ channel and the number No of channels per node were calculated from ensemble average values of the mean Na+ current and the variance of Na+-current fluctuations. (3) For a hyperpolarizing holding potential of VH = -28 mV the mean values of the channel conductance and number were gamma = 9.8 pS and No = 74000. (4) After changing the holding potential to the resting potential (VH = 0) the conductance gamma increased by a factor of 1.37 whereas the number No decreased by a factor of 0.60. (5) Addition of 8 nM tetrodotoxin at a holding potential of VH = -28 mV increased gamma by a factor of 1.55 and reduced No by a factor of 0.25. (6) The increase of the channel conductance at reduced channel numbers suggests negative cooperativity between Na+ channels in the nodal membrane.  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