首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
ATP-sensitive potassium (K(ATP)) channels in pancreatic beta-cells comprise sulfonylurea receptor (SUR) 1 and inwardly-rectifying potassium channel (Kir) 6.2 subunits. We have evaluated the effect of intracellular taurine on K(ATP) channel activity in rat pancreatic beta-cells using the patch-clamp technique. The mechanism of taurine action was also examined using recombinant K(ATP) channels. The islets and single beta-cells from male Sprague-Dawley rats were collected by collagenase digestion technique. Single K(ATP) channel currents were recorded by the inside-out mode at a membrane potential of -60mV. Cytosolic free-Ca(2+) concentration ([Ca(2+)](c)) and insulin secretory capacity were measured by the dual-excitation fluorimetry and radioimmunoassay, respectively. The native beta-cell K(ATP) channel was directly inhibited by taurine in a dose-dependent manner. Taurine did not influence ATP-mediated inhibition or MgADP-induced activation of the channel activity. The sensitivity of the K(ATP) channel to glybenclamide, but not gliclazide, was enhanced by taurine. Glybenclamide elicited a greater increase in [Ca(2+)](c) and increased insulin secretion in the beta-cells when pretreated with taurine. Taurine did not inhibit Kir6.2DeltaC36 currents, a truncated form of Kir6.2, expressed in Xenopus oocytes without SUR. These results demonstrate that taurine inhibits the K(ATP) channel activity in the beta-cells, interacting with a benzamido-binding site on SUR1, but not Kir6.2.  相似文献   

2.
Taurine has been found to inhibit ATP-sensitive K+ (KATP) channels in rat pancreatic beta-cells [Park et al., Biochem Pharmacol 2004;67:1089-1096] which could be due to its interaction with a benzamido-binding site on SUR1. In present study, we further evaluated the mechanism of taurine action on the KATP-channel inhibition, using cloned KATP-channels with different types of SUR subunit expressed in Xenopus laevis oocytes. The oocytes were coinjected with Kir6.2 mRNA, and mRNA encoding SUR1, SUR2A or SUR2B. Macroscopic currents were recorded from giant excised inside-out patches. The binding of glibenclamide to SUR1 was assessed by using a glibenclamide-fluorescent probe. Intracellular taurine inhibited all three types of KATP-channels to a similar extent. They were fit to the Hill equation, showing IC50 of 11.0 mM for Kir6.2/SUR1, 10.9 mM for Kir6.2/SUR2A, and 9.0 mM for Kir6.2/SUR2B currents. Taurine at the concentration of 10 mM enhanced the high-affinity bindings of glibenclamide and repaglinide on all types of SUR, whereas the low-affinity binding on Kir6.2 was not affected. The intensity of glibenclamide fluorescence was higher in the plasma membrane of taurine-pretreated oocytes. The high-affinity binding of tolbutamide or gliclazide on SUR was not modified by taurine. These results suggest that the taurine inhibition of KATP-channels is mediated by an interaction with the site on SUR where the benzamido group is bound. Therefore, intracellular concentrations of taurine in different tissues may be more important in determining taurine modulation of the KATP-channel rather than distinct types of SUR subunit.  相似文献   

3.
ATP-sensitive K+ channels in the kidney   总被引:5,自引:0,他引:5  
ATP-sensitive K+ channels (KATP channels) form a link between the metabolic state of the cell and the permeability of the cell membrane for K+ which, in turn, is a major determinant of cell membrane potential. KATP channels are found in many different cell types. Their regulation by ATP and other nucleotides and their modulation by other cellular factors such as pH and kinase activity varies widely and is fine-tuned for the function that these channels have to fulfill. In most excitable tissues they are closed and open when cell metabolism is impaired; thereby the cell is clamped in the resting state which saves ATP and helps to preserve the structural integrity of the cell. There are, however, notable exceptions from this rule; in pancreatic -cells, certain neurons and some vascular beds, these channels are open during the normal functioning of the cell.In the renal tubular system, KATP channels are found in the proximal tubule, the thick ascending limb of Henle's loop and the cortical collecting duct. Under physiological conditions, these channels have a high open probability and play an important role in the reabsorption of electrolytes and solutes as well as in K+ homeostasis. The physiological role of their nucleotide sensitivity is not entirely clear; one consequence is the coupling of channel activity to the activity of the Na-K-ATPase (pump-leak coupling), resulting in coordinated vectorial transport. In ischemia, however, the reduced ATP/ADP ratio would increase the open probability of the KATP channels independently from pump activity; this is particularly dangerous in the proximal tubule, where 60 to 70% of the glomerular ultrafiltrate is reabsorbed.The pharmacology of KATP channels is well developed including the sulphonylureas as standard blockers and the structurally heterogeneous family of channel openers. Blockers and openers, exemplified by glibenclamide and levcromakalim, show a wide spectrum of affinities towards the different types of KATP channels. Recent cloning efforts have solved the mystery about the structure of the channel: the KATP channels in the pancreatic -cell and in the principal cell of the renal cortical collecting duct are heteromultimers, composed of an inwardly rectifying K+ channel and sulphonylurea binding subunit(s) with unknown stoichiometry. The proteins making up the KATP channel in these two cell types are different (though homologous), explaining the physiological and pharmacological differences between these channel subtypes.  相似文献   

4.
5.
The substituted benzo[c]quinolizinium compounds MPB-07 and MPB-91 are novel activators of the cystic fibrosis transmembrane conductance regulator (CFTR) chloride channel. High homologies between CFTR and the sulfonylurea receptor (SUR), which associates with the potassium channel Kir6.2 to form the ATP-sensitive K(+) (K(ATP)) channel, prompted us to examine possible effects of these compounds on K(ATP) channels using electrophysiological recordings and binding assays. Activity of recombinant K(ATP) channels expressed in Xenopus oocytes was recorded in the inside-out configuration of the patch-clamp technique. Channels were practically unaffected by MPB-07 but were fully blocked by MPB-91 with half-inhibition achieved at approximately 20 microM MPB-91. These effects were similar on channels formed by Kir6.2, and either the SUR1 or SUR2A isoforms were independent of the presence of nucleotides. They were not influenced by SUR mutations known to interfere with its nucleotide-binding capacity. MPB-91, but not MPB-07, was able to displace binding of glibenclamide to HEK cells expressing recombinant SUR1/Kir6.2 channels. Glibenclamide binding to native channels from pancreatic MIN6 cells was also displaced by MPB-91. A Kir6.2 mutant able to form channels without SUR was also blocked by MPB-91, but not by MPB-07. These observations demonstrate that neither MPB-07 nor MPB-91 interact with SUR, in spite of its high homology with CFTR, and that MPB-91 blocks K(ATP) channels by binding to the Kir6.2 subunit. Thus, caution should be exercised when planning to use MPB compounds in cystic fibrosis therapy, specially MPB-91 which could nonetheless find interesting applications as the precursor of a new class of K channel blockers.  相似文献   

6.
Hypoglycemic sulfonylureas (e.g. tolbutamide, glibenclamide) exert their stimulatory effects on pancreatic beta-cells by closure of ATP-sensitive K(+) (K(ATP)) channels. Pancreatic K(ATP) channels are composed of two subunits, a pore-forming inwardly rectifying K(+) channel (Kir6.2) subunit and a regulatory subunit (the sulfonylurea receptor of subtype 1 (SUR1)) in a (SUR1/Kir6.2)(4) stoichiometry. The aim of the present study was to characterize the interaction of green-fluorescent 3-[3-(4,4 difluoro-5,7-dimethyl-4-bora-3a,4a-diaza-S-indacen-3-yl)propanamido] glibenclamide (Bodipy-glibenclamide) with pancreatic beta-cell K(ATP) channels using patch-clamp and fluorescence microscopy techniques. Bodipy-glibenclamide inhibited K(ATP) currents from the clonal insulinoma cell line RINm5F half-maximally at a concentration of 0.6nM. Using laser-scanning confocal microscopy Bodipy-glibenclamide was shown to induce a diffuse fluorescence across the RINm5F cell, but only about 17% of total Bodipy-glibenclamide-induced fluorescence intensity in RINm5F cells was due to specific binding to SUR1. Using fluorescence correlation spectroscopy, it could be demonstrated that the fluorescence label contributes to the protein binding and, therefore, possibly also to the non-specific binding of Bodipy-glibenclamide observed in RINm5F cells. Specific binding of Bodipy-glibenclamide to SUR1 in RINm5F cells might be localized to different intracellular structures (nuclear envelope, endoplasmic reticulum, Golgi compartment, insulin secretory granules) as well as to the plasma membrane. In conclusion, Bodipy-glibenclamide is a high-affinity blocker of pancreatic beta-cell K(ATP) currents and can be used for visualizing SUR1 in intact pancreatic beta-cells, although non-specific binding must be taken into account in confocal microscopy experiments on intact beta-cells.  相似文献   

7.
Antimicrobial fluoroquinolones induce, with strongly varying frequency, life-threatening hypoglycemias, which is explained by their ability to block KATP channels in pancreatic B-cells and thus to initiate insulin secretion. In apparent contradiction to this, we observed that none of the fluoroquinolones in this study (gatifloxacin, moxifloxacin, ciprofloxacin, and a number of fluorophenyl-substituted compounds) initiated insulin secretion of perifused mouse islets when the glucose concentration was basal (5 mM). Only when the glucose concentration was stimulatory by itself (10 mM), the fluoroquinolones enhanced secretion. The fluoroquinolones were ineffective on SUR1 Ko islets, which do not have functional KATP channels. All of these fluoroquinolones depolarized the membrane potential of mouse B-cells (patch-clamping in the whole-cell mode). Using metabolically intact B-cells (perforated-patch mode) however, 100 μM of gatifloxacin, ciprofloxacin or moxifloxacin were unable to depolarize when the glucose concentration was 5 mM, whereas other KATP channel blockers (tolbutamide and efaroxan) remained effective. Only at a very high concentration (500 μM) gatifloxacin and moxifloxacin, but not ciprofloxacin induced repetitive depolarizations which could be antagonized by diazoxide. In the presence of 10 mM glucose all fluoroquinolones which enhanced secretion markedly elevated cytosolic calcium concentration ([Ca2+]i). In the presence of 5 mM glucose gatifloxacin and moxifloxacin at 500 μM but not at 100 μM elevated [Ca2+]i. It is concluded that fluoroquinolones in the clinically relevant concentration range are not initiators, but rather enhancers of glucose-induced insulin secretion. The block of KATP channels appears necessary but not sufficient to explain the hypoglycemic effect of fluoroquinolones.  相似文献   

8.

BACKGROUND AND PURPOSE

Rosiglitazone is an anti-diabetic drug acting as an insulin sensitizer. We recently found that rosiglitazone also inhibits the vascular isoform of ATP-sensitive K+ channels and compromises vasodilatory effects of β-adrenoceptor activation and pinacidil. As its potency for the channel inhibition is in the micromolar range, rosiglitazone may be used as an effective KATP channel inhibitor for research and therapeutic purposes. Therefore, we performed experiments to determine whether other isoforms of KATP channels are also sensitive to rosiglitazone and what their sensitivities are.

EXPERIMENTAL APPROACH

KIR6.1/SUR2B, KIR6.2/SUR1, KIR6.2/SUR2A, KIR6.2/SUR2B and KIR6.2ΔC36 channels were expressed in HEK293 cells and were studied using patch-clamp techniques.

KEY RESULTS

Rosiglitazone inhibited all isoforms of KATP channels in excised patches and in the whole-cell configuration. Its IC50 was 10 µmol·L−1 for the KIR6.1/SUR2B channel and ∼45 µmol·L−1 for KIR6.2/SURx channels. Rosiglitazone also inhibited KIR6.2ΔC36 channels in the absence of the sulphonylurea receptor (SUR) subunit, with potency (IC50= 45 µmol·L−1) almost identical to that for KIR6.2/SURx channels. Single-channel kinetic analysis showed that the channel inhibition was mediated by augmentation of the long-lasting closures without affecting the channel open state and unitary conductance. In contrast, rosiglitazone had no effect on KIR1.1, KIR2.1 and KIR4.1 channels, suggesting that the channel inhibitory effect is selective for KIR6.x channels.

CONCLUSIONS AND IMPLICATIONS

These results suggest a novel KATP channel inhibitor that acts on the pore-forming KIR6.x subunit, affecting the channel gating.

LINKED ARTICLE

This article is commented on by Dart, pp. 23–25 of this issue. To view this commentary visit http://dx.doi.org/10.1111/j.1476-5381.2012.01990.x  相似文献   

9.
ATP-sensitive potassium (KATP) channels are weak, inward rectifiers that couple metabolic status to cell membrane electrical activity, thus modulating many cellular functions. An increase in the ADP/ATP ratio opens KATP channels, leading to membrane hyperpolarization. KATP channels are ubiquitously expressed in neurons located in different regions of the brain, including the hippocampus and cortex. Brief hypoxia triggers membrane hyperpolarization in these central neurons. In vivo animal studies confirmed that knocking out the Kir6.2 subunit of the KATP channels increases ischemic infarction, and overexpression of the Kir6.2 subunit reduces neuronal injury from ischemic insults. These findings provide the basis for a practical strategy whereby activation of endogenous KATP channels reduces cellular damage resulting from cerebral ischemic stroke. KATP channel modulators may prove to be clinically useful as part of a combination therapy for stroke management in the future.  相似文献   

10.
A prolonged exposure of isolated pancreatic islets to insulin secretagogues, the imidazolines phentolamine, alinidine and idazoxan (100microM each), the sulfonylurea tolbutamide (500microM), or the alkaloid quinine (100microM) resulted in morphological damage of 4-18% of beta-cells compared to less than 2% in controls. Thus, the question arose whether K(ATP) channel-blocking insulin secretagogues are beta-cell toxic as has already been suggested for sulfonylureas. The concentration- and time-dependency of the secretagogue-associated toxicity was documented by viability assays in insulin-secreting HIT T15 cells. Treatment for 24h with idazoxan reduced MTT conversion by 50% at 100microM and by 98% at 1000microM. Phentolamine and quinine reduced viability comparably at 1000microM, but were less toxic at 100microM. On the other hand, the imidazoline alinidine and the sulfonylurea tolbutamide were only moderately toxic (less than 40% viability loss at 1000microM). The imidazoline efaroxan appeared even to be non-toxic. Apoptotic DNA fragmentation and DEVD-caspase activation was observed at 100microM of idazoxan and phentolamine, whereas at 1000microM signs of necrosis predominated. Alinidine, tolbutamide and quinine treatment did not increase markers of apoptotic cell death. Blocking Ca(2+) influx by D600 did not diminish secretagogue-associated toxicity. Electron microscopy confirmed the validity of these observations for beta-cells in intact mouse islets. In summary, beta-cell toxicity of the tested insulin secretagogues varied widely and did not depend on a prolonged Ca(2+) influx via L-type Ca(2+) channels. Thus, secretagogue-mediated closure of K(ATP) channels is apparently not per se beta-cell toxic.  相似文献   

11.
ATP-sensitive K(+) channels (K(ATP)) are major component of preventing ischemia-reperfusion injury. However, there is little information regarding to the expressional difference of K(ATP) and its function between left and right ventricles. In this study, we measured the lactate dehydrogenase release of rabbit heart slices in vitro and determined the difference of the K(ATP) expression at the both ventricles by measuring the level of K(ATP)-forming Kir6.2 (OcKir6.2) mRNA using in situ hybridization. The hearts were preconditioned with 15 min hypoxia and reoxygenated for 15 min before a hypoxic period of 60 min, followed by reoxygenation for 180 min. With hypoxic preconditioning (100% N(2)) with 15 min, left ventricles (LV) showed higher release of LDH comparing with right ventricles (RV). Adding K(ATP) blocker glibenclamide (10 μM) prior to a hypoxic period of 60 min, hypoxic preconditioning effect of RV was more abolished than LV. With in situ hybridization, the optical density of OcKir6.2 was higher in RV. Therefore, we suggest that different K(ATP) expression between LV and RV is responsible for the different response to hypoxia and hypoxic preconditioning of rabbit hearts.  相似文献   

12.
BACKGROUND AND PURPOSE: ATP-sensitive K+ channels (K(ATP) channels) play important roles in regulating the resting membrane potential of detrusor smooth muscle. Actions of ZD0947, a novel KATP channel opener, on both carbachol (CCh)-induced detrusor contractions and membrane currents in human urinary bladder myocytes were investigated. EXPERIMENTAL APPROACH: Tension measurements and patch-clamp techniques were utilized to study the effects of ZD0947 in segments of human urinary bladder. Immunohistochemistry was also performed to detect the expression of the sulphonylurea receptor 1 (SUR1) and the SUR2B antigens in human detrusor muscle. KEY RESULTS: ZD0947 (> or = 0.1 microM) caused a concentration-dependent relaxation of the CCh-induced contraction of human detrusor, which was reversed by glibenclamide. The rank order of the potency to relax the CCh-induced contraction was pinacidil > ZD0947 > diazoxide. In conventional whole-cell configuration, ZD0947 (> or = 1 microM) caused a concentration-dependent inward K+ current which was suppressed by glibenclamide at -60 mV. When 1 mM ATP was included in the pipette solution, application of pinacidil or ZD0947 caused no inward K+ current at -60 mV. Gliclazide (< or =1 microM), a selective SUR1 blocker, inhibited the ZD0947-induced currents (Ki = 4.0 microM) and the diazoxide-induced currents (high-affinity site, Ki1 = 42.4 nM; low-affinity site, Ki2 = 84.5 microM) at -60 mV. Immunohistochemical studies indicated the presence of SUR1 and SUR2B proteins, which are constituents of KATP channels, in the bundles of human detrusor smooth muscle. CONCLUSIONS AND IMPLICATIONS: These results suggest that ZD0947 caused a glibenclamide-sensitive detrusor relaxation through activation of glibenclamide-sensitive KATP channels in human urinary bladder.  相似文献   

13.
A peptide was isolated from the venom of the scorpion Tityus trivittatus. It is an isoform of the toxin TsTX-IV earlier described [Toxicon 37 (1999) 651] and identical to butantoxin [Arch. Biochem. Biophys. 379 (2000) 18], both isolated from the Brazilian scorpion Tityus serrulatus. This newly characterized peptide contains 40 amino acid residues with a molecular mass of [M+H(+)] 4507.0, cross-linked by four disulfide bridges, made between the cysteine pairs: Cys2-Cys5, Cys10-Cys31, Cys16-Cys36 and Cys20-Cys38. It blocks in a completely reversible manner the Shaker B K(+)-channels, with a K(d) around 660nM. It belongs to the sub-family 12 and it is now being classified as alpha-KTx 12.2.  相似文献   

14.

BACKGROUND AND PURPOSE

Analogues of the endogenous diacylglycerols have been used extensively as pharmacological activators of protein kinase C (PKC). Several reports show that some of these compounds have additional effects that are independent of PKC activation, including direct block of K+ and Ca2+ channels. We investigated whether dioctanoyl-sn-glycerol (DiC8), a commonly used diacylglycerol analogue, blocks K+ currents of rat mesenteric arterial smooth muscle in a PKC-independent manner.

EXPERIMENTAL APPROACH

Conventional whole-cell and inside-out patch clamp was used to measure the inhibition of K+ currents of rat isolated mesenteric smooth muscle cells by DiC8 in the absence and presence of PKC inhibitor peptide.

KEY RESULTS

Mesenteric artery smooth muscle Kv currents inactivated very slowly with a time constant of about 2 s following pulses from −65 to +40 mV. Application of 1 µM DiC8 produced an approximate 40-fold increase in the apparent rate of inactivation. Pretreatment of the cells with PKC inhibitor peptide had a minimal effect on the action of DiC8, and substantial inactivation still occurred, indicating that this effect was mainly independent of PKC. We also found that DiC8 blocked BK and KATP currents, and again a significant proportion of these blocks occurred independently of PKC activation.

CONCLUSIONS AND IMPLICATIONS

These results show that DiC8 has a direct effect on arterial smooth muscle K+ channels, and this precludes its use as a PKC activator when investigating PKC-mediated effects on vascular K+ channels.  相似文献   

15.
Summary Effects of a vasodilator, nicorandil (2-nicotinamidoethyl nitrate) on four kinds for cardiac K+ channels were investigated in guinea pig ventricular and atrial cells using inside-out patch recording combined with oilgate concentration jump method.Nicorandil of 300 mol/l failed to affect the inward-rectifier K+ channel and the Na+-activated K+ channel. The open probability of the muscarinic K+ channel, when activated by the application of GTP, was not changed by the drug. Nicorandil selectively increased the open probability of the ATP-sensitive K+ channel that was partly suppressed by intracellular ATP. The median effective concentration (EC50) of nicorandil was 74 mol/l and Hill coefficient was 1.32 in the concentration-open probability relationship. The closing rate of the K+ channel by ATP was markedly delayed by the drug, whereas the open rate on removal of ATP was scarcely affected. Nicorandil had only little effect on this channel after run-down. It was concluded that nicorandil selectively activates the ATP-sensitive K+ channel mainly by modulating the ATP-dependent gate.Send offprint requests to M. Takano at the above address  相似文献   

16.
We investigated the role of nitric oxide (NO) in pacemaker activity and signal mechanisms in cultured interstitial cells of Cajal (ICC) of the mouse small intestine using whole cell patch-clamp techniques at 30°C. ICC generated pacemaker potential in the current clamp mode and pacemaker currents at a holding potential of –70 mV. (±)-S-nitroso-N-acetylpenicillamine (SNAP; a NO donor) produced membrane hyperpolarization and inhibited the amplitude and frequency of the pacemaker currents, and increased resting currents in the outward direction. These effects were blocked by the use of glibenclamide (an ATP-sensitive K+ channel blocker), but not by the use of 5-hydroxydecanoic acid (a mitochondrial ATP-sensitive K+ channel blocker). Pretreatment with ODQ (a guanylate cyclase inhibitor) almost blocked the NO-induced effects. The use of cell-permeable 8-bromo-cyclic GMP also mimicked the action of SNAP. However, the use of KT-5823 (a protein kinase G inhibitor) did not block the NO-induced effects. Spontaneous [Ca2+]i oscillations in ICC were inhibited by the treatment of SNAP, as seen in recordings of intracellular Ca2+ ([Ca2+]i). These results suggest that NO inhibits pacemaker activity by the activation of ATP-sensitive K+ channels via a cyclic GMP dependent mechanism in ICC, and the activation of ATP-sensitive K+ channels mediates the inhibition of spontaneous [Ca2+]i oscillations.  相似文献   

17.
The effects of imipramine on A-type delayed rectifier K+ currents and ATP-sensitive K+ (KATP) currents were studied in isolated murine proximal colonic myocytes using the whole-cell patch-clamp technique. Depolarizing test pulses between -80 mV and +30 mV with 10 mV increments from the holding potential of -80 mV activated voltage-dependent outward K+ currents that peaked within 50 ms followed by slow decreasing sustained currents. Early peak currents were inhibited by the application of 4-aminopyridine, whereas sustained currents were inhibited by the application of TEA. The peak amplitude of A-type delayed rectifier K+ currents was reduced by external application of imipramine. The half-inactivation potential and the half-recovery time of A-type delayed rectifier K+ currents were not changed by imipramine. With 0.1 mM ATP and 140 mM K+ in the pipette and 90 mM K+ in the bath solution and a holding potential of -80 mV, pinacidil activated inward currents; this effect was blocked by glibenclamide. Imipramine also inhibited KATP currents. The inhibitory effects of imipramine in A-type delayed rectifier K+ currents and KATP currents were not changed by guanosine 5-O-(2-thiodiphosphate) (GDPbetaS) and chelerythrine, a protein kinase C inhibitor. These results suggest that imipramine inhibits A-type delayed rectifier K+ currents and KATP currents in a manner independent of G-protein and protein kinase C.  相似文献   

18.
19.
The present study was designed to further evaluate the biological effects and tissue selectivity of new 3-alkylaminobenzo- and 3-alkylaminopyridothiadiazine 1,1-dioxides bearing identical branched alkylamino chains at the 3-position. These original compounds were compared with their parent molecules; namely the K(ATP) channel openers diazoxide and pinacidil. All tested 3-alkylamino-substituted derivatives provoked a marked, concentration-dependent inhibition of the glucose-induced insulin secretion from rat pancreatic islets. The 3-alkylaminopyridothiadiazine 1,1-dioxides evoked a weak vasorelaxant activity whilst their 7-halo-substituted benzothiadiazine counterparts elicited pronounced, concentration-dependent, relaxations of rat aortic rings. The myorelaxant effect of the different drugs was tightly correlated with their capacity to increase (86)Rb outflow ((42)K substitute) from prelabelled and perifused rat aortic rings. EC(50)/IC(50) ratios (vascular/pancreatic) revealed a pronounced selectivity of the 3-alkylaminopyridothiadiazine 1,1-dioxides towards the pancreatic endocrine tissue. Structure-activity relationships suggest that, besides the requirement of an electronegative pole at the 7-position of the heterocycle, a minimal steric hindrance confers an optimal insulin-secreting cell selectivity. Lastly, radioisotopic, electrophysiological and pharmacological investigations indicate that the marked vasorelaxant properties of the 3-alkylaminobenzothiadiazine 1,1-dioxides are related to the activation of smooth muscle K(ATP) channels.  相似文献   

20.
The ATP-sensitive K(+) (K(ATP)) channels are composed of sulfonylurea receptor and inwardly rectifying K(+) channel (Kir6.2) subunit. These channels are regulated by intracellular ADP/ATP ratio and play a role in cellular metabolism. Diethyl pyrocarbonate (DEPC), a histidine-specific alkylating reagent, is known to modify the histidine residues of the structure of proteins. The objective of this study was to determine whether DEPC modifies K(ATP)-channel activity in pituitary GH(3) cells. Steady-state fluctuation analyses of macroscopic K(+) current at -120 mV produced power spectra that could be fitted with a single Lorentzian curve in these cells. The time constants in the presence of DEPC were increased. Consistent with fluctuation analyses, the mean open time of K(ATP)-channels was significantly increased during exposure to DEPC. However, DEPC produced no change in single-channel conductance, despite the ability of this compound to enhance K(ATP)-channel activity in a concentration-dependent manner with an EC(50) value of 16 microM. DEPC-stimulated K(ATP)-channel activity was attenuated by pretreatment with glibenclamide. In current-clamp configuration, DEPC decreased the firing of action potentials in GH(3) cells. A further application of glibenclamide reversed DEPC-induced inhibition of spontaneous action potentials. Intracellullar Ca(2+) measurements revealed the ability of DEPC to decrease Ca(2+) oscillations in GH(3) cells. Simulation studies also demonstrated that the increased conductance of K(ATP)-channels used to mimic DEPC actions reduced the frequency of spontaneous action potentials and fluctuation of intracellular Ca(2+). The results indicate that chemical modification with DEPC enhances K(ATP)-channel activity and influences functional activities of pituitary GH(3) cells.  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