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.
Dihydropyridine Receptor-Ryanodine Receptor Uncoupling in Aged Skeletal Muscle   总被引:15,自引:0,他引:15  
The mechanisms underlying skeletal muscle functional impairment and structural changes with advanced age are only partially understood. In the present study, we support and expand our theory about alterations in sarcolemmal excitation-sarcoplasmic reticulum Ca2+ release-contraction uncoupling as a primary skeletal muscle alteration and major determinant of weakness and fatigue in mammalian species including humans. To test the hypothesis that the number of RYR1 (ryanodine receptor) uncoupled to DHPR (dihydropyridine receptor) increases with age, we performed high-affinity ligand binding studies in soleus, extensor digitorum longus (EDL) and in a pool of several skeletal muscles consisting of a mixture of fast- and slow-twitch muscle fibers in middle-aged (14-month) and old (28-months) Fisher 344 Brown Norway F1 hybrids rats. The number of DHPR, RYR1, the coupling between both receptors expressed as the DHPR/RYR1 maximum binding capacity, and their dissociation constant for high-affinity ligands were measured. The DHPR/RYR1 ratio was significantly reduced in the three groups of muscles (pool: 1.03 ± 0.15 and 0.80 ± 0.11, soleus: 0.44 ± 0.12 and 0.26 ± 0.10, and EDL: 0.95 ± 0.14 and 0.68 ± 0.10, for middle-aged and old muscles, respectively). These data support the concept that DHPR-RYR1 uncoupling results in alterations in the voltage-gated sarcoplasmic reticulum Ca2+ release mechanism, decreases in myoplasmic Ca2+ elevation in response to sarcolemmal depolarization, reduced Ca2+ supply to contractile proteins and reduced contraction force with aging. Received: 26 August 1996/Revised: 30 December 1996  相似文献   

2.
The same isoform of ryanodine receptor (RYR1) is expressed in both fast and slow mammalian skeletal muscles. However, differences in contractile activation and calcium release kinetics in intact and skinned fibers have been reported. In this work, intracellular Ca2+ transients were measured in soleus and extensor digitorum longus (EDL) single muscle fibers using mag-fura-2 (K D for Ca2+= 49 μm) as Ca2+ fluorescent indicator. Fibers were voltage-clamped at V h =−90 mV and sarcoplasmic reticulum calcium release was measured at the peak (a) and at the end (b) of 200 msec pulses at +10 mV. Values of a-b and b were assumed to correspond to Ca2+-gated and voltage-gated Ca2+ release, respectively. Ratios (b/a-b) in soleus and EDL fibers were 0.41 ± 0.05 and 1.01 ± 0.13 (n= 12), respectively. This result suggested that the proportion of dihydropyridine receptor (DHPR)-linked and unlinked RYRs is different in soleus and EDL muscle. The number of DHPR and RYR were determined by measuring high-affinity [3H]PN200-110 and [3H]ryanodine binding in soleus and EDL rat muscle homogenates. The B max values corresponded to a PN200-110/ryanodine binding ratio of 0.34 ± 0.05 and 0.92 ± 0.11 for soleus and EDL muscles (n= 4–8), respectively. These data suggest that soleus muscle has a larger calcium-gated calcium release component and a larger proportion of DHPR-unlinked RYRs. Received: 31 August 1995/Revised: 25 January 1996  相似文献   

3.
The microenvironment near the apical membrane of MDCK cells was studied by quantitation of the fluorescence of wheat germ agglutin attached to fluorescein (WGA). WGA was shown to bind to sialic acid residues attached to galactose at the α-2,3 position in the glycocalyx on the apical membrane. Young MDCK cells (5–8 days after splitting) showed a patchy distribution of WGA at stable sites that returned to the same locations after removal of sialic acid residues by neuraminidase treatment. Other lectins also showed stable binding to patches on the apical membrane of young cells. The ratio of WGA fluorescence emission at two excitation wavelengths was used to measure near-membrane pH. The near-membrane pH was markedly acidic to the pH 7.4 bathing solution in both young and older cells (13–21 days after splitting). Patches on the apical membrane of young cells exhibited a range of near-membrane pH values with a mean ±sem of 6.86 ± 0.04 (n= 121) while the near-membrane pH of older cells was 6.61 ± 0.04 (n= 120) with a uniform WGA distribution. We conclude that the distribution of lectin binding sites in young cells reflects the underlying nonrandom location of membrane proteins in the apical membrane and that nonuniformities in the pH of patches may indicate regional differences in membrane acid-base transport as well as in the location of charged sugars in the glycocalyx. Received: 15 December 1999/Revised: 16 March 2000  相似文献   

4.
We examined the effect of the2-agonist clenbuterol (50 µM)on depolarization-induced force responses and sarcoplasmic reticulum (SR) function in muscle fibers of the rat (Rattusnorvegicus; killed by halothane overdose) that had beenmechanically skinned, rendering the2-agonist pathway inoperable.Clenbuterol decreased the peak of depolarization-induced forceresponses in the extensor digitorum longus (EDL) and soleus fibers to77.2 ± 9.0 and 55.6 ± 5.4%, respectively, ofcontrols. The soleus fibers did not recover. Clenbuterol significantlyand reversibly reduced SR Ca2+loading in EDL and soleus fibers to 81.5 ± 2.8 and 78.7 ± 4.0%, respectively, of controls. Clenbuterol also producedan ~25% increase in passive leak ofCa2+ from the SR of the EDL andsoleus fibers. These results indicate that clenbuterol has directeffects on fast- and slow-twitch skeletal muscle, in the absence of the2-agonist pathway. Theincreased Ca2+ leak in the triadregion may lead to excitation-contraction coupling damage in the soleusfibers and could also contribute to the anabolic effect of clenbuterolin vivo.

  相似文献   

5.
We have obtained and modeled the electrical characteristics of the plasma membrane of Chara internodal cells: intact, without turgor and perfused with and without ATP. The cells were voltage and space-clamped to obtain the I/V (current-voltage) and G/V (conductance-voltage) profiles of the cell membrane. The intact cells yielded similar I/V characteristics with resting p.d.s of −221 ± 12 mV (cytoplasmic clamp, 5 cells) and −217 ± 12 mV (vacuolar clamp, 5 cells). The cut unperfused cells were depolarized at −169 ± 12 mV (7 cells) compared to the vacuole-clamped intact cells. The cells perfused with ATP fell into three groups: hyperpolarized group with resting p.d. −175 ± 12 mV (4 cells) and I/V profile similar to the intact and cut unperfused cells; depolarized group with resting p.d. of −107 ± 12 mV (6 cells) and I/V profiles close to linear; and excited cells with profiles showing a negative conductance region and resting p.d. at −59 ± 12 mV (5 cells). The cells perfused with medium containing no ATP showed upwardly concave I/V characteristics and resting p.d. at −81 ± 12 mV (6 cells). The I/V curves were modeled employing the ``Two-state' model for the H+ pump (Hansen et al., 1981). The inward and outward rectifiers were fitted to exponential functions and combined with a linear background current. The excitation state in perfused cells was modeled by including an inward current, i excit, with p.d.-dependence described by a combination of hyperbolic tangent functions. An inward current, i no-ATP, with a smaller amplitude, but very similar p.d.-dependence was also included in the simulation of the I/V curves from cells without ATP. This approach avoided I/V curve subtraction. The modeling of the total I/V and G/V characteristics provided more information about the parameters of the ``Two-state' pump model, as well as more quantitative understanding of the interaction of the major transport systems in the plasmalemma in generation of the resting potential under a range of circumstances. ATP had little effect on nonpump currents except the excitation current; depolarization profoundly affected the pump characteristics. Received: 23 January/Revised: 10 October 1995  相似文献   

6.
The structural determinants of mibefradil inhibition were analyzed using wild-type and inactivation-modified CaV1.2 (α1C) and CaV2.3 (α1E) channels. Mibefradil inhibition of peak Ba2+ currents was dose- and voltage-dependent. An increase of holding potentials from −80 to −100 mV significantly shifted dose-response curves toward higher mibefradil concentrations, namely from a concentration of 108 ± 21 μm (n= 7) to 288 ± 17 μm (n= 3) for inhibition of half of the Cav1.2 currents (IC 50) and from IC 50= 8 ± 2 μm (n= 9) to 33 ± 7 μm (n= 4) for CaV2.3 currents. In the presence of mibefradil, CaV1.2 and CaV2.3 experienced significant use-dependent inhibition (0.1 to 1 Hz) and slower recovery from inactivation suggesting mibefradil could promote transition(s) to an absorbing inactivated state. In order to investigate the relationship between inactivation and drug sensitivity, mibefradil inhibition was studied in inactivation-altered CaV1.2 and CaV2.3 mutants. Mibefradil significantly delayed the onset of channel recovery from inactivation in CEEE (Repeat I + part of the I–II linker from CaV1.2 in the CaV2.3 host channel), in EC(AID)EEE (part of the I–II linker from CaV1.2 in the CaV2.3 host channel) as well as in CaV1.2 E462R, and CaV2.3 R378E (point mutation in the β-subunit binding motif) channels. Mibefradil inhibited the faster inactivating chimera EC(IS1-6)EEE with an IC 50= 7 ± 1 μm (n= 3), whereas the slower inactivating chimeras EC(AID)EEE and CEEE were, respectively, inhibited with IC 50= 41 ± 5 μm (n= 4) and IC 50= 68 ± 9 μm (n= 5). Dose-response curves were superimposable for the faster EC(IS1-6)EEE and CaV2.3, whereas intermediate-inactivating channel kinetics (CEEE, CaV1.2 E462R, and CaV1.2 E462K) were inhibited by similar concentrations of mibefradil with IC 50≈ 55–75 μm. The slower CaV1.2 wild-type and CaV1.2 Q473K channels responded to higher doses of mibefradil with IC 50≈ 100–120 μm. Mibefradil was also found to significantly speed up the inactivation kinetics of slower channels (CaV1.2, CEEE) with little effect on the inactivation kinetics of faster-inactivating channels (CaV2.3). A open-channel block model for mibefradil interaction with high-voltage-activated Ca2+ channels is discussed and shown to qualitatively account for our observations. Hence, our data agree reasonably well with a ``receptor guarded mechanism' where fast inactivation kinetics efficiently trap mibefradil into the channel. Received: 14 March 2001/Revised: 25 June 2001  相似文献   

7.
Passive membrane properties, steady-state Na/K pump current (I p) and modulation of I p by adrenergic agonists were studied with patch-clamp techniques in adult rat ventricular myocytes that were freshly isolated or maintained in culture for 1–4 days. Freshly isolated (day 0) myocytes had a 1.7–1.8 times smaller specific membrane resistance compared with that of cells on any day in culture. From day 0 to 4 there was a progressive decrease in cell capacitance (−17.6 ± 0.8 pF/day) without a parallel decline in cell dimensions. The pump current density (1.55 pA/pF) was stable over the 0–4 days in culture. In rod-shaped myocytes norepinephrine (NE) and isoproterenol (ISO) stimulated I p in a dose-dependent manner, with an apparent affinity of 36 ± 8 and 1.5 ± 0.4 nm, and maximum stimulation of 0.65 ± 0.02 and 0.57 ± 0.02 pA/pF, respectively. Nadolol suppressed this effect, suggesting that it was mediated by β-adrenergic receptor activation. An inverse relationship was found between steady-state I p and the stimulation of I p by NE. In contrast to what was shown in guinea pig cardiac myocytes, in rat myocytes isoproterenol stimulation of I p was not increased by intracellular [Ca] and it did not change the I p -membrane potential relation. These results show that in adult rat cardiac myocytes NE and ISO are potent stimulators of Na/K pump activity, and this effect may be studied using rat myocytes maintained in short-term culture. Received: 12 November 1997/Revised: 5 March 1998  相似文献   

8.
MDCK cells display several acid-base transport systems found in intercalated cells, such as Na+-H+ exchange, H+–K+ ATPase and Cl/HCO 3 exchange. In this work we studied the functional activity of a vacuolar H+-ATPase in MDCK cells and its chloride dependence. We measured intracellular pH (pHi) in monolayers grown on glass cover slips utilizing the pH sensitive probe BCECF. To analyze the functional activity of the H+ transporters we observed the intracellular alkalinization in response to an acute acid load due to a 20 mm NH+ 4 pulse, and calculated the initial rate of pHi recovery (dpHi/dt). The cells have a basal pHi of 7.17 ± 0.01 (n= 23) and control dpHi/dt of 0.121 ± 0.006 (n= 23) pHi units/min. This pHi recovery rate is markedly decreased when Na+ was removed, to 0.069 ± 0.004 (n= 16). It was further reduced to 0.042 ± 0.005 (n= 12) when concanamycin 4.6 × 10−8 m (a specific inhibitor of the vacuolar H+-ATPase) was added to the zero Na+ solution. When using a solution with zero Na+, low K+ (0.5 mm) plus concanamycin, pHi recovery fell again, significantly, to 0.023 ± 0.006 (n= 14) as expected in the presence of a H+–K+-ATPase. This result was confirmed by the use of 5 × 10−5 m Schering 28080. The Na+ independent pHi recovery was significantly reduced from 0.069 ± 0.004 to 0.042 ± 0.004 (n= 12) when NPPB 10−5 m (a specific blocker of Cl channels in renal tubules) was utilized. When the cells were preincubated in 0 Cl/normal Na+ solution for 8 min. before the ammonium pulse, the pHi recovery fell from 0.069 ± 0.004 to 0.041 ± 0.007 (n= 12) in a Na+ and Cl free solution. From these results we conclude that: (i) MDCK cells have two Na+-independent mechanisms of pHi recovery, a concanamycin sensitive H+-ATPase and a K+ dependent, Schering 28080 sensitive H+–K+ ATPase; and, (ii) pHi recovery in Na+-free medium depends on the presence of a chloride current which can be blocked by NPPB and impaired by preincubation in Cl–free medium. This finding supports a role for chloride in the function of the H+ ATPase, which might be electrical shunting or a biochemical interaction. Received: 24 October 1997/Revised: 19 February 1998  相似文献   

9.
The modulation of the calmodulin-induced inhibition of the calcium release channel (ryanodine receptor) by two sulfhydryl oxidizing compounds, 4-(chloromercuri)phenyl–sulfonic acid (4-CMPS) and 4,4′-dithiodipyridine (4,4′-DTDP) was determined by single channel current recordings with the purified and reconstituted calcium release channel from rabbit skeletal muscle sarcoplasmic reticulum (HSR) and [3H]ryanodine binding to HSR vesicles. 0.1 μm CaM reduced the open probability (P o ) of the calcium release channel at maximally activating calcium concentrations (50–100 μm) from 0.502 ± 0.02 to 0.137 ± 0.022 (n= 28), with no effect on unitary conductance. 4-CMPS (10–40 μm) and 4,4′-DTDP (0.1–0.3 mm) induced a concentration dependent increase in P o (> 0.9) and caused the appearance of longer open states. CaM shifted the activation of the calcium release channel by 4-CMPS or 4,4′-DTDP to higher concentrations in single channel recordings and [3H]ryanodine binding. 40 μm 4-CMPS induced a near maximal (P o > 0.9) and 0.3 mm 4,4′-DTDP a submaximal (P o = 0.74) channel opening in the presence of CaM, which was reversed by the specific sulfhydryl reducing agent DTT. Neither 4-CMPS nor 4,4′-DTDP affected Ca-[125I]calmodulin binding to HSR. 1 mm MgCl2 reduced P o from 0.53 to 0.075 and 20–40 μm 4-CMPS induced a near maximal channel activation (P o > 0.9). These results demonstrate that the inhibitory effect of CaM or magnesium in a physiological concentration is diminished or abolished at high concentrations of 4-CMPS or 4,4′-DTDP through oxidation of activating sulfhydryls on cysteine residues of the calcium release channel. Received: 22 July 1999/Revised: 15 November 1999  相似文献   

10.
Alpha-sarcoglycan (Sgca) is a transmembrane glycoprotein of the dystrophin complex located at skeletal and cardiac muscle sarcolemma. Defects in the alpha-sarcoglycan gene (Sgca) cause the severe human-type 2D limb girdle muscular dystrophy. Because Sgca-null mice develop progressive muscular dystrophy similar to human disorder they are a valuable animal model for investigating the physiopathology of the disorder. In this study, biochemical and functional properties of fast-twitch extensor digitorum longus (EDL) and slow-twitch soleus muscles of the Sgca-null mice were analyzed. EDL muscle of Sgca-null mice showed twitch and tetanic kinetics comparable with those of wild-type controls. In contrast, soleus muscle showed reduction of twitch half-relaxation time, prolongation of tetanic half-relaxation time, and increase of maximal rate of rise of tetanus. EDL muscle of Sgca-null mice demonstrated a marked reduction of specific twitch and tetanic tensions and a higher resistance to fatigue compared with controls, changes that were not evident in dystrophic soleus. Contrary to EDL fibers, soleus muscle fibers of Sgca-null mice distinctively showed right shift of the pCa-tension (pCa is the negative log of Ca2+ concentration) relationships and reduced sensitivity to caffeine of sarcoplasmic reticulum. Both EDL and soleus muscles showed striking changes in myosin heavy-chain (MHC) isoform composition, whereas EDL showed a larger number of hybrid fibers than soleus. In contrast to the EDL, soleus muscle of Sgca-null mice contained a higher number of regenerating fibers and thus higher levels of embryonic MHC. In conclusion, this study revealed profound distinctive biochemical and physiological modifications in fast- and slow-twitch muscles resulting from alpha-sarcoglycan deficiency.  相似文献   

11.
The giant marine alga Valonia utricularis is a classical model system for studying the electrophysiology and water relations of plant cells by using microelectrode and pressure probe techniques. The recent finding that protoplasts can be prepared from the giant ``mother cells' (Wang, J., Sukhorukov, V.L., Djuzenova, C.S., Zimmermann, U., Müller, T., Fuhr, G., 1997, Protoplasma 196:123–134) allowed the use of the patch-clamp technique to examine ion channel activity in the plasmalemma of this species. Outside-out and cell-attached experiments displayed three different types of voltage-gated Cl channels (VAC1, VAC2, VAC3, Valonia Anion Channel 1,2,3), one voltage-gated K+ channel (VKC1, Valonia K + Channel 1) as well as stretch-activated channels. In symmetrical 150 mm Cl media, VAC1 was most frequently observed and had a single channel conductance of 36 ± 7 pS (n= 4) in the outside-out and 33 ± 5 pS (n= 10) in the cell-attached configuration. The reversal potential of the corresponding current-voltage curves was within 0 ± 4 mV (n= 4, outside-out) and 9 ± 7 mV (n= 10, cell-attached) close to the Nernst potential of Cl and shifted towards more negative values when cell-attached experiments were performed in asymmetrical 50:150 mm Cl media (bath/pipette; E Cl− −20 ± 7 mV (n= 4); Nernst potential −28 mV). Consistent with a selectivity for Cl, VAC1 was inhibited by 100 μM DIDS (4,4′-diisothiocyanatostilbene-2,2′-disulfonic acid). VAC1 was activated by a hyperpolarization of the patch. Boltzmann fits of the channel activity under symmetrical 150 mm Cl conditions yielded a midpoint potential of −12 ± 5 mV (n= 4, outside-out) and −3 ± 6 mV (n= 9, cell-attached) and corresponding apparent minimum gating charges of 15 ± 3 (n= 4) and 18 ± 5 (n= 9). The midpoint potential shifted to more negative values in the presence of a Cl gradient. VAC2 was activated by voltages more negative than E Cl− and was always observed together with VAC1, but less frequently. It showed a ``flickering' gating. The single channel conductance was 99 ± 10 pS (n= 6). VAC3 was activated by membrane depolarization and frequently exhibited several subconductance states. The single channel conductance of the main conductance state was 36 ± 5 pS (n= 5). VKC1 was also activated by positive clamped voltages. Up to three conductance states occurred whereby the main conductance state had a single channel conductance of 124 ± 27 pS (n= 6). In the light of the above results it seems to be likely that VAC1 contributes mainly to the Cl conductance of the plasmalemma of the turgescent ``mother cells' and that this channel (as well as VAC2) can operate in the physiological membrane potential range. The physiological significance of VAC3 and VKC1 is unknown, but may be related (as the stretch-activated channels) to processes involved in turgor regulation. Received: 24 June 1999/Revised: 2 September 1999  相似文献   

12.
The fumarate transport system of the bacterium Helicobacter pylori was investigated employing radioactive tracer analysis. The transport of fumarate at micromolar concentrations was saturable with a K M of 220 ± 21 μm and V max of 54 ± 2 nmole/min/mg protein at 20°C, depended on temperature between 4 and 40°C, and was susceptible to inhibitors, suggesting the presence of one or more fumarate carriers. The release of fumarate from cells was also saturable with a K M of 464 ± 71 μm and V max of 22 ± 2 nmol/min/mg protein at 20°C. The rates of fumarate influx at millomolar concentrations increased linearly with permeant concentration, and depended on the age of the cells. The transport system was specific for dicarboxylic acids suggesting that fumarate is taken up via dicarboxylate transporters. Succinate and fumarate appeared to form an antiport system. The properties of fumarate transport were elucidated by investigating the effects of amino acids, monovalent cations, pH and potential inhibitors. The results provided evidence that influx and efflux of fumarate at low concentrations from H. pylori cells was a carrier-mediated secondary transport with the driving force supplied by the chemical gradient of the anion. The anaerobic C4-dicarboxylate transport protein identified in the genome of the bacterium appeared to be a good candidate for the fumarate transporter. Received: 11 December 1997/Revised: 7 May 1998  相似文献   

13.
In order to define potential interaction sites of SGLT1 with the transport inhibitor phlorizin, mutagenesis studies were performed in a hydrophobic region of loop 13 (aa 604–610), located extracellularly, close to the C-terminus. COS 7 cells were transiently transfected with the mutants and the kinetic parameters of α-methyl-d-glucopyranoside (AMG) uptake into the cells were investigated. Replacement of the respective amino acids with lysine reduced the maximal uptake rate: Y604K showed 2.2%, L606K 48.4%, F607K 15.1%, C608K 13.1%, G609K 14.1%, and L610K 17.2% of control. In all mutants the apparent K i for phlorizin increased at least by a factor of 5 compared to the wild-type K i of 4.6 ± 0.7 μmol/l; most striking changes were observed for Y604K (K i = 75.3 ± 19.0 μmol/l) and C608K (K i = 83.6 ± 13.9 μmol/l). Replacement of these amino acids with a nonpolar amino acid instead of lysine such as in Y604F, Y604G and C608A showed markedly higher affinities for phlorizin. In cells expressing the mutants the apparent affinity of AMG uptake for the sugar was not statistically different from that of the wild type (K m = 0.8 ± 0.2 mmol/l). These studies suggest that the region between amino acids 604 and 610 is involved in the interaction between SGLT1 and phlorizin, probably by providing a hydrophobic pocket for one of the aromatic rings of the aglucone moiety of the glycoside. Received: 29 March 2001/Revised: 15 June 2001  相似文献   

14.
A detailed temperature dependence study of a well-defined plant ion channel, the Ca2+-activated K+ channel of Chara corallina, was performed over the temperature range of their habitats, 5–36°C, at 1°C resolution. The temperature dependence of the channel unitary conductance at 50 mV shows discontinuities at 15 and 30°C. These temperatures limit the range within which ion diffusion is characterized by the lowest activation energy (E a = 8.0 ± 1.6 kJ/mol) as compared to the regions below 15°C and above 30°C. Upon reversing membrane voltage polarity from 50 to −50 mV the pattern of temperature dependence switched from discontinuous to linear with E a = 13.6 ± 0.5 kJ/mol. The temperature dependence of the effective number of open channels at 50 mV showed a decrease with increasing temperature, with a local minimum at 28°C. The mean open time exhibited a similar behavior. Changing the sign of membrane potential from 50 to −50 mV abolished the minima in both temperature dependencies. These data are discussed in the light of higher order phase transitions of the Characean membrane lipids and corresponding change in the lipid-protein interaction, and their modulation by transmembrane voltage. Received: 14 June 2000/Revised: 20 September 2000  相似文献   

15.
Adenosine 3′,5′-cyclic monophosphate (cAMP) is known to stimulate exogenous IsK channel current in the Xenopus oocyte expression system. The present study was performed to determine whether elevation of cytosolic cAMP in a native mammalian epithelium known to secrete K+ through endogenously expressed IsK channels would stimulate K+ secretion through these channels. The equivalent short circuit current (I sc ) across vestibular dark cell epithelium in gerbil was measured in a micro-Ussing chamber and the apical membrane current (I IsK ) and conductance (g IsK ) of IsK channels was recorded with both the on-cell macro-patch and nystatin-perforated whole-cell patch-clamp techniques. It has previously been shown that I sc can be accounted for by transepithelial K+ secretion and that the apical IsK channels constitute a significant pathway for K+ secretion. The identification of the voltage-dependent whole-cell currents in vestibular dark cells was strengthened by the finding that a potent blocker of IsK channels, chromanol 293B, strongly reduced I IsK from 646 ± 200 to 154 ± 22 pA (71%) and g IsK from 7.5 ± 2.6 to 2.8 ± 0.4 nS (53%). Cytoplasmic cAMP was elevated by applying dibutyryl cyclic AMP (dbcAMP), or the phosphodiesterase inhibitors 3-isobutyl-1-methylxanthine (IBMX) and Ro-20-1724. dbcAMP (1 mm) increased I sc and I IsK from 410 ± 38 to 534 ± 40 μA/cm2 and from 4.3 ± 0.8 to 11.4 ± 2.2 pA, respectively. IBMX (1 mm) caused transient increases of I sc from 415 ± 30 to 469 ± 38 μA/cm2 and Ro-20-1724 (0.1 mm) from 565 ± 43 to 773 ± 58 μA/cm2. IBMX increased I IsK from 5.5 ± 1.5 to 16.9 ± 5.8 pA in on-cell experiments and from 191 ± 31 to 426 ± 53 pA in whole-cell experiments. The leak conductance due to all non-IsK channel sources did not change during dbcAMP and IBMX while 293B in the presence of dbcAMP reduced I IsK by 84% and g IsK by 62%, similar to unstimulated conditions. These results demonstrate that the cAMP pathway is constitutively active in vestibular dark cells and that the cAMP pathway stimulates transepithelial K+ secretion by increasing IsK channel current rather than by altering another transport pathway. Received: 9 June 1995/Revised: 17 October 1996  相似文献   

16.
Extracellular nucleotides modulate renal ion transport. Our previous results in M-1 cortical collecting duct cells indicate that luminal and basolateral ATP via P2Y2 receptors stimulate luminal Ca2+-activated Cl channels and inhibit Na+ transport. Here we address the mechanism of ATP-mediated inhibition of Na+ transport. M-1 cells had a transepithelial voltage (V te ) of −31.4 ± 1.3 mV and a transepithelial resistance (R te ) of 1151 ± 28 Ωcm2. The amiloride-sensitive short circuit current (I sc ) was −28.0 ± 1.1 μA/cm2. The ATP-mediated activation of Cl channels was inhibited when cytosolic Ca2+ increases were blocked with cyclopiazonic acid (CPA). Without CPA the ATP-induced [Ca2+]i increase was paralleled by a rapid and transient R te decrease (297 ± 51 Ωcm2). In the presence of CPA, basolateral ATP led to an R te increase by 144 ± 17 Ωcm2 and decreased V te from −31 ± 2.6 to −26.6 ± 2.5 mV. I sc dropped from −28.6 ± 2.4 to −21.6 ± 1.9 μA/cm2. Similar effects were observed with luminal ATP. In the presence of amiloride, ATP was without effect. This reflects ATP-mediated inhibition of Na+ absorption. Lowering [Ca2+]i by removal of extracellular Ca2+ did not alter the ATP effect. PKC inhibition or activation were without effect. Na+ absorption was activated by pHi alkalinization and inhibited by pHi acidification. ATP slightly acidified M-1 cells by 0.05 ± 0.005 pH units, quantitatively not explaining the ATP-induced effect. In summary this indicates that extracellular ATP via luminal and basolateral P2Y2 receptors inhibits Na+ absorption. This effect is not mediated via [Ca2+]i, does not involve PKC and is to a small part mediated via intracellular acidification. Received: 9 February 2001/Revised: 17 May 2001  相似文献   

17.
In this study, we investigated whether the previously established differences between fast- and slow-twitch single skeletal muscle fibers of the rat, in terms of myosin heavy chain (MHC) isoform composition and contractile function, are also detectable in excitation-contraction (E-C) coupling. We compared the contractile responsiveness of electrophoretically typed, mechanically skinned single fibers from the soleus (Sol), the extensor digitorum longus (EDL), and the white region of the sternomastoid (SM) muscle to t-system depolarization-induced activation. The quantitative parameters assessed were the amplitude of the maximum depolarization-induced force response (DIFR(max); normalized to the maximum Ca(2+)-activated force in that fiber) and the number of responses elicited until the force declined by 75% of DIFR(max) (R-D(75%)). The mean DIFR(max) values for type IIB EDL and type IIB SM fibers were not statistically different, and both were greater than the mean DIFR(max) for type I Sol fibers. The mean R-D(75%) for type IIB EDL fibers was greater than that for type I Sol fibers as well as type IIB SM fibers. These data suggest that E-C coupling characteristics of mechanically skinned rat single muscle fibers are related to MHC-based fiber type and the muscle of origin.  相似文献   

18.
We investigated the effects of exclusive and sustained transgenic overexpression of insulin-like growth factor (IGF)-I in the central nervous system (CNS) on the age-dependent decline in muscle strength, excitation-contraction coupling, muscle innervation and neuromuscular junction postterminal architecture. We found that (1) transgenic IGF-I overexpression in the CNS does not modify the decline in extensor digitorum longus (EDL) and soleus muscle weight with aging and (2) strength significantly decreases in transgenic (Tg) compared to wild-type mice. The latter finding is consistent with (3) the decreased absolute and specific force measured in the EDL muscle in vitro and (4) the decreased charge movement and peak intracellular Ca2+ mobilization in individual muscle fibers from old IGF-I Tg mice compared to young wild-type mice, which also is associated with (5) decreased dihydropyridine receptor α1-subunit expression in old compared to young IGF-I Tg mice. (6) Tg IGF-I prevents a change in muscle fiber type that is associated with (7) improved muscle innervation and postterminal neuromuscular structure. (8) IGF-I is expressed extensively across the spinal cord gray matter and the lateral motor column. Our results raise questions about the timing and cell location of CNS IGF-I overexpression necessary to prevent or to ameliorate age-dependent alterations in the structure and function of skeletal muscle. Ramón Jiménez Moreno and María Laura Messi contributed equally to this work.  相似文献   

19.
Collapsed proximal convoluted tubules (PCT) shrink to reach a volume 20% lower than control and do not exhibit regulatory volume increase when submitted to abrupt 150 mOsm/kg hypertonic shock. The shrinking is accompanied by a rapid depolarization of the basolateral membrane potential (V BL) of 8.4 ± 0.5 mV, with respect to a control value of −54.5 ± 1.9 mV (n= 15). After a small and transient hyperpolarization, V BL further depolarizes to reach a steady depolarization of 19.5 ± 1.5 mV (n= 15) with respect to control. In the post-control period, V BL returns to −55.8 ± 1.5 mV. The basolateral partial conductance to K+ (t K ) which is 0.17 ± 0.01 (n= 5) in control condition, decreases rapidly to nonmeasurable values during the hypertonic shock and returns to 0.23 ± 0.03 in the post-control period. The basolateral partial conductance to Cl (t Cl), which is 0.05 ± 0.02 (n= 5) in control, also decreases in hypertonicity to a nonmeasurable value and returns to 0.03 ± 0.01 in post control. The partial conductance mediated by the Na-HCO3 cotransporter (t NaHCO3), which is 0.48 ± 0.06 (n= 5) in control condition, remains the same at 0.44 ± 0.05 (n= 5) during the hypertonic period. Similarly, the membrane absolute conductance mediated by the Na-HCO3 cotransporter (G Na-HCO3) does not vary appreciably. Concomitant with cell shrinkage, intracellular pH (pH i ) decreases from a control value of 7.26 ± 0.01 to 7.13 ± 0.02 (n= 12) and then remains constant. Return to control solution brings back pH i to 7.28 ± 0.03. From these results, we conclude that in collapsed PCT, a sustained decrease in cellular volume leads to cell acidification and to inhibition of K+ and Cl conductances. Received: 6 February 1996/Revised: 10 October 1996  相似文献   

20.
In frog red blood cells, K-Cl cotransport (i.e., the difference between ouabain-resistant K fluxes in Cl and NO3) has been shown to mediate a large fraction of the total K+ transport. In the present study, Cl-dependent and Cl-independent K+ fluxes via frog erythrocyte membranes were investigated as a function of external and internal K+ ([K+] e and [K+] i ) concentration. The dependence of ouabain-resistant Cl-dependent K+ (86Rb) influx on [K+] e over the range 0–20 mm fitted the Michaelis-Menten equation, with an apparent affinity (K m ) of 8.2 ± 1.3 mm and maximal velocity (V max ) of 10.4 ± 1.6 mmol/l cells/hr under isotonic conditions. Hypotonic stimulation of the Cl-dependent K+ influx increased both K m (12.8 ± 1.7 mm, P < 0.05) and V max (20.2 ± 2.9 mmol/l/hr, P < 0.001). Raising [K+] e above 20 mm in isotonic media significantly reduced the Cl-dependent K+ influx due to a reciprocal decrease of the external Na+ ([Na+] e ) concentration below 50 mm. Replacing [Na+] e by NMDG+ markedly decreased V max (3.2 ± 0.7 mmol/l/hr, P < 0.001) and increased K m (15.7 ± 2.1 mm, P < 0.03) of Cl-dependent K+ influx. Moreover, NMDG+ Cl substitution for NaCl in isotonic and hypotonic media containing 10 mm RbCl significantly reduced both Rb+ uptake and K+ loss from red cells. Cell swelling did not affect the Na+-dependent changes in Rb+ uptake and K+ loss. In a nominally K+(Rb+)-free medium, net K+ loss was reduced after lowering [Na+] e below 50 mm. These results indicate that over 50 mm [Na+] e is required for complete activation of the K-Cl cotransporter. In nystatin-pretreated cells with various intracellular K+, Cl-dependent K+ loss in K+-free media was a linear function of [K+] i , with a rate constant of 0.11 ± 0.01 and 0.18 ± 0.008 hr−1 (P < 0.001) in isotonic and hypotonic media, respectively. Thus K-Cl cotransport in frog erythrocytes exhibits a strong asymmetry with respect to transported K+ ions. The residual, ouabain-resistant K+ fluxes in NO3 were only 5–10% of the total and were well fitted to linear regressions. The rate constants for the residual influxes were not different from those for K+ effluxes in isotonic (∼0.014 hr−1) and hypotonic (∼0.022 hr−1) media, but cell swelling resulted in a significant increase in the rate constants. Received: 19 November 1998/Revised: 23 August 1999  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