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.
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   

2.
The ecophysiology of the hypotonic response was studied in the charophyte alga, Lamprothamnium papulosum, which was grown in a marine (SW; 1072 mosmol kg–1) and a brackish (1/2 SW; 536 mosmol kg–1) environment. The cells produced an extracellular mucilage identified by histochemical staining as a mixture of sulphated and carboxylated polysaccharides. The thickness and chemical composition of the mucilage layer was a function of environmental salinity and cell age. Mucilage progressively increased in thickness from the apex (9 SW cells: 12·6 ± 1·8 μm; 15 1/2 SW cells: 4·8 ± 0·7 μm) to the base of the plants (15 SW cells: 44·8 ± 3·3 μm; nine 1/2 SW cells: 23·8 ± 2·5 μm); with a corresponding increase in the sulphated proportion. The mucilage was significantly thicker in SW plants. Hydraulic conductivity (Lp) at the apex of SW plants, measured by transcellular osmosis, was 8·3 × 10–13 m s–1 Pa–1. This was close to Lp of freshwater Chara (8·5 × 10–13 m s–1 Pa–1) which lacked mucilage. Basal SW cells with thicker mucilage had a smaller apparent Lp of 3·5 × 10–13 m s–1 Pa–1. The electrophysiology of the resting state and hypotonic response was compared in cells from the two environments based on current/voltage (I/V) analysis. The resting potential difference (PD) and conductance differed (11 SW cells: – 102·4 ± 10·1 mV, eight SW cells: 18·6 ± 2·4 S m–2; 19 1/2 SW cells: –125·7 ± 5·9 mV, 8·3 ± 0·8 S m–2). The type of cellular response to a hypotonic shock (decrease of 268 mosmol kg–1) also differed. In 1/2 SW plants, only the apical cells with thin mucilage responded classically with depolarization, conductance increase, Ca2+ influx, cessation of cytoplasmic streaming, and K+ and Cl effluxes. Older cells making up the bulk of the plants responded with depolarization, but continued cytoplasmic streaming, and had only a small increase in conductance; or depolarized transiently without altering the I/V profile, conductance or streaming speed. Most cells remained depolarized and in the K+ state 1 h post-shock. Cells treated with the K+ channel blocker tetraethylammonium chloride also depolarized and remained depolarized. The SW cells depolarized but otherwise responded minimally to a 268 mosmol kg–1 drop in osmolarity and required a further 268 mosmol kg–1 down-step to elicit a change in the conductance. A spectrum of responses was measured in successively older and more mucilaginous cells from the same marine plant. We discuss the ecophysiological significance of the mucilage layer which modulates the cellular response to osmotic shock and which can be secreted to different degrees by plants inhabiting environments of different salinity.  相似文献   

3.
Ion transport properties of pulmonary small airway epithelia are poorly understood. To characterize these properties, airways were excised from anesthetized pigs. Transepithelial potential difference (PD) and conductance were measured in five airway regions: trachea (T, 7.9 +/- 0.2 mm diam), mainstem bronchi (MB, 5.5 +/- 0.2 mm diam), large bronchi (LB, 1.69 +/- 0.12 mm diam), small bronchi (SB, 0.70 +/- 0.06 mm diam), and bronchioles (BR, 0.25 +/- 0.05 mm diam). T and MB were mounted in Ussing-type chambers, and LB, SB, and BR were cannulated with pipettes and perfused. PDs of control tissues were -9.7 +/- 0.8 mV (T), -4.0 +/- 0.5 mV (MB), -4.3 +/- 1.0 mV (LB), -4.5 +/- 0.4 mV (SB), and -1.5 +/- 0.4 mV (BR), lumen negative. Amiloride significantly (P < 0.05) inhibited PDs by 25-70% in all airway regions and decreased conductance 17-33% in all regions except LB where a 10% increase was observed. Bumetanide significantly reduced the amiloride-insensitive PD 54-62% in all regions except BR. Bumetanide had little effect on conductance in T, SB, and BR, but conductance was increased in MB and LB. All airways except the smallest BR significantly hyperpolarized when the solution that bathed the lumen was replaced with Cl(-)-free solution. In bronchioles, hyperpolarization by luminal Cl(-)-free solution was inversely related to fractional inhibition of PD with amiloride but directly related to lumen diameter. These results suggest that 1) porcine tracheas, bronchi, and bronchioles actively absorb Na+, and 2) secretion of Cl- may occur in all airway regions except small bronchioles.  相似文献   

4.
The possible contribution of Ca2+-activated Cl- channel [I(Cl(Ca))] and myosin light-chain kinase (MLCK) to nonadrenergic, noncholinergic slow inhibitory junction potentials (sIJP) was studied using conventional intracellular microelectrode recordings in circular smooth muscle of opossum esophageal body and guinea pig ileum perfused with Krebs solution containing atropine (3 microM), guanethidine (3 microM), and substance P (1 microM). In opossum esophageal circular smooth muscle, resting membrane potential (MP) was -51.9 +/- 0.7 mV (n = 89) with MP fluctuations of 1-3 mV. A single square-wave nerve stimulation of 0.5 ms duration and 80 V induced a sIJP with amplitude of 6.3 +/- 0.2 mV, half-amplitude duration of 635 +/- 19 ms, and rebound depolarization amplitude of 2.4 +/- 0.1 mV (n = 89). 9-Anthroic acid (A-9-C), niflumic acid (NFA), wortmannin, and 1-(5-chloronaphthalene-1-sulfonyl)-1H-hexahydro-1,4-diazepine (ML-9) abolished MP fluctuations, sIJP, and rebound depolarization in a concentration-dependent manner. A-9-C and NFA but not wortmannin and ML-9 hyperpolarized MP. In guinea pig ileal circular smooth muscle, nerve stimulation elicited an IJP composed of both fast (fIJP) and slow (sIJP) components, followed by rebound depolarization. NFA (200 microM) abolished sIJP and rebound depolarization but left the fIJP intact. These data suggest that in the tissues studied, activation of I(Cl(Ca)), which requires MLCK, contributes to resting MP, and that closing of I(Cl(Ca)) is responsible for sIJP.  相似文献   

5.
A channelopathy mutation in segment IIS6 of Ca(V)1.4 (I745T) has been shown to cause severe visual impairment by shifting the activation and inactivation curves to more hyperpolarized voltages and slowing activation and inactivation kinetics. A similar gating phenotype is caused by the corresponding mutation, I781T, in Ca(V)1.2 (midpoint of activation curve (V(0.5)) shifted to -37.7 +/- 1.2 mV). We show here that wild-type gating can partially be restored by a helix stabilizing rescue mutation N785A. V(0.5) of I781T/N785A (V(0.5) = -21.5 +/- 0.6 mV) was shifted back towards wild-type (V(0.5) = -9.9 +/- 1.1 mV). Homology models developed in our group (see accompanying article for details) were used to perform Molecular Dynamics-simulations (MD-simulations) on wild-type and mutant channels. Systematic changes in segment IIIS6 (M1187-F1194) and in helix IIS6 (N785-L786) were studied. The simulated structural changes in S6 segments of I781T/N785A were less pronounced than in I781T. A delicate balance between helix flexibility and stability enabling the formation of hydrophobic seals at the inner channel mouth appears to be important for wild-type Ca(V)1.2 gating. Our study illustrates that effects of mutations in the lower part of IIS6 may not be localized to the residue or even segment being mutated, but may affect conformations of interacting segments.  相似文献   

6.
Hyperpolarization-activated, cyclic nucleotide-gated (HCN) channels underlie spontaneous rhythmic activities in the heart and brain. Sulfhydryl modification of ion channels is a proven approach for studying their structure-function relationships; here we examined the effects of the hydrophilic sulfhydryl-modifying agents methanethiosulfonate ethylammonium (MTSEA(+)) and methanethiosulfonate ethylsulfonate (MTSES(-)) on wild-type (WT) and engineered HCN1 channels. External application of MTSEA(+) to WT channels irreversibly reduced whole-cell currents (I(MTSEA)/I(Control) = 42 +/- 2%), slowed activation and deactivation kinetics ( approximately 7- and approximately 3-fold at -140 and -20 mV, respectively), and produced hyperpolarizing shifts of steady-state activation (V(12)((MTSEA)) = -125.8 +/- 9.0 mV versus V(12)((Control)) = -76.4 +/- 1.6 mV). Sequence inspection revealed the presence of five endogenous cysteines in the transmembrane domains of HCN1: three are putatively close to the extracellular milieu (Cys(303), Cys(318), and Cys(347) in the S5, S5-P, and P segments, respectively), whereas the remaining two are likely to be cytoplasmic or buried. To identify the molecular constituent(s) responsible for the effects of MTSEA(+), we mutated the three "external" cysteines individually to serine. C303S did not yield measurable currents. Whereas C347S channels remained sensitive to MTSEA(+), C318S was not modified (I(MTSEA)/I(Control) = 101 +/- 2%, V(12)((MTSEA)) = -78.4 +/- 1.1 mV, and V(12)((Control)) = -79.8 +/- 2.3 mV). Likewise, WT (but not C318S) channels were sensitive to MTSES(-). Despite their opposite charges, MTSES(-) produced changes directionally similar to those effected by MTSEA(+) (I(MTSES)/I(Control) = 22 +/- 1.6% and V(12)((MTSES)) = -145.9 +/- 4.9 mV). We conclude that S5-P Cys(318) of HCN1 is externally accessible and that the external pore vestibule and activation gating of HCN channels are allosterically coupled.  相似文献   

7.
Nasal potential difference (PD) measurements have been used to demonstrate defective CFTR function in cystic fibrosis (CF) and to evaluate potential CF therapies. We used the selective thiazolidinone CFTR inhibitor CFTR(inh)-172 to define the involvement of CFTR in nasal PD changes in mice and pigs. In normal mice infused intranasally with a physiological saline solution containing amiloride, nasal PD was -4.7 +/- 0.7 mV, hyperpolarizing by 15 +/- 1 mV after a low-Cl- solution, and a further 3.9 +/- 0.5 mV after forskolin. CFTR(inh)-172 produced 1.1 +/- 0.9- and 4.3 +/- 0.7-mV depolarizations when added after low Cl- and forskolin, respectively. Systemically administered CFTR(inh)-172 reduced the forskolin-induced hyperpolarization from 4.7 +/- 0.4 to 0.9 +/- 0.1 mV but did not reduce the low Cl(-)-induced hyperpolarization. Nasal PD was -12 +/- 1 mV in CF mice after amiloride, changing by <0.5 mV after low Cl- or forskolin. In pigs, nasal PD was -14 +/- 3 mV after amiloride, hyperpolarizing by 13 +/- 2 mV after low Cl- and a further 9 +/- 1 mV after forskolin. CFTR(inh)-172 and glibenclamide did not affect nasal PD in pigs. Our results suggest that cAMP-dependent nasal PDs in mice primarily involve CFTR-mediated Cl- conductance, whereas cAMP-independent PDs are produced by a different, but CFTR-dependent, Cl- channel. In pigs, CFTR may not be responsible for Cl- channel-dependent nasal PDs. These results have important implications for interpreting nasal PDs in terms of CFTR function in animal models of CFTR activation and inhibition.  相似文献   

8.
The heterogeneity of arginases in rat tissues.   总被引:11,自引:0,他引:11       下载免费PDF全文
1. The mid-point reduction potentials of the various groups in xanthine oxidase from bovine milk were determined by potentiometric titration with dithionite in the presence of dye mediators, removing samples for quantification of the reduced species by e.p.r. (electron-paramagnetic-resonance) spectroscopy. The values obtained for the functional enzyme in pyrophosphate buffer, pH8.2, are: Fe/S centre I, -343 +/- 15mV; Fe/S II, -303 +/- 15mV; FAD/FADH-; -351 +/- 20mV; FADH/FADH2, -236 +/-mV; Mo(VI)/Mo(V) (Rapid), -355 +/- 20mV; Mo(V) (Rapid)/Mo(IV), -355 +/- 20mV. 2. Behaviour of the functional enzyme is essentially ideal in Tris but less so in pyrophosphate. In Tris, the potential for Mo(VI)/Mo(V) (Rapid) is lowered relative to that in pyrophosphate, but the potential for Fe/S II is raised. The influence of buffer on the potentials was investigated by partial-reduction experiments with six other buffers. 3. Conversion of the enzyme with cyanide into the non-functional form, which gives the Slow molybdenum signal, or alkylation of FAD, has little effect on the mid-point potentials of the other centres. The potentials associated with the Slow signal are: Mo(VI)/Mo(V) (Slow), -440 +/- 25mV; Mo(V) (Slow)/Mo(IV), -480 +/- 25 mV. This signal exhibits very sluggish equilibration with the mediator system. 4. The deviations from ideal behaviour are discussed in terms of possible binding of buffer ions or anti-co-operative interactions amongst the redox centres.  相似文献   

9.
Application of the current-clamp technique in rainbow trout atrial myocytes has yielded resting membrane potentials that are incompatible with normal atrial function. To investigate this paradox, we recorded the whole membrane current (I(m)) and compared membrane potentials recorded in isolated cardiac myocytes and multicellular preparations. Atrial tissue and ventricular myocytes had stable resting potentials of -87 +/- 2 mV and -83.9 +/- 0.4 mV, respectively. In contrast, 50 out of 59 atrial myocytes had unstable depolarized membrane potentials that were sensitive to the holding current. We hypothesized that this is at least partly due to a small slope conductance of I(m) around the resting membrane potential in atrial myocytes. In accordance with this hypothesis, the slope conductance of I(m) was about sevenfold smaller in atrial than in ventricular myocytes. Interestingly, ACh increased I(m) at -120 mV from 4.3 pA/pF to 27 pA/pF with an EC(50) of 45 nM in atrial myocytes. Moreover, 3 nM ACh increased the slope conductance of I(m) fourfold, shifted its reversal potential from -78 +/- 3 to -84 +/- 3 mV, and stabilized the resting membrane potential at -92 +/- 4 mV. ACh also shortened the action potential in both atrial myocytes and tissue, and this effect was antagonized by atropine. When applied alone, atropine prolonged the action potential in atrial tissue but had no effect on membrane potential, action potential, or I(m) in isolated atrial myocytes. This suggests that ACh-mediated activation of an inwardly rectifying K(+) current can modulate the membrane potential in the trout atrial myocytes and stabilize the resting membrane potential.  相似文献   

10.
The dependence of the Ca++-activated Cl- channels on potential difference (PD) was extracted from current-voltage (I/V) profiles recorded at the time of hypotonic regulation while the large conductance (G) K+ channels were blocked by tetraethylammonium (TEA). The total clamp current (I) was dominated by the Cl- I, i(Cl), with small contribution from the background I (i(background)). The i(Cl) was fitted by the Goldman-Hodgkin-Katz (GHK) model with enhanced PD dependence simulated by Boltzmann probability distributions. The i(background) was modelled by an empirical equation. The i(Cl) responded to PD changes within tens of milliseconds. The G maxima were located between -20 and -150 mV. The Cl- channel number and channel permeability parameter, N(Cl)P(Cl), decreased as a function of time in a hypotonic medium (from 0.45 x 10(-7) to 0.17 x 10(-7) ms(-1) in 19 min), with the positive half activation PD, V50+, shifting from +35 to -65 mV, and the negative half activation PD, V50-, shifting from -134 to -310 mV. The fitted Cl- concentration [Cl-]cyt at the time of hypotonic regulation indicated rapid equalization of vacuolar and cytoplasmic concentrations. Excellent data obtained under similar experimental conditions in a previous study enabled us to infer [Ca++]cyt influences on the Cl- channel characteristics. Thick sulphated polysaccharide mucilage, found on Lamprothamnium cells acclimated to more saline media, eliminated the activation of the i(Cl) at the time of the hypotonic regulation. This effect was reversed by the application of the enzyme heparinase. The characteristics of the i(Cl) were found to be consistent with a component of the excitation Is at the time of the action potential (AP). The short duration of the excitation transients was contrasted with that of the hypotonic regulation. The mechanisms for Cl- channel activation (and hence the Ca++ channel activation) were considered.  相似文献   

11.
I(f), encoded by the hyperpolarization-activated cyclic nucleotide-modulated (HCN) channel family, is a key player in cardiac and neuronal pacing. Although HCN channels structurally resemble voltage-gated K(+) (Kv) channels, their structure-function correlation is much less clear. Here we probed the functional importance of the HCN1 S3-S4 linker by multiple substitutions of its residues. Neutralizing Glu(235), an acidic S3-S4 linker residue conserved in all hyperpolarization-activated channels, by Ala substitution produced a depolarizing activation shift (V(12) = -65.0 +/- 0.7 versus -70.6 +/- 0.7 mV for wild-type HCN1); the charge-reversed mutation E235R shifted activation even more positively (-56.2 +/- 0.5 mV). Increasing external Mg(2+) mimicked the progressive rightward shifts of E235A and E235R by gradually shifting activation (V(12) = 1 < 3 < 10 < 30 mm); Delta V(12) induced by 30 mm Mg(2+) was significantly attenuated for E235A (+7.9 +/- 1.2 versus +11.3 +/- 0.9 mV for wild-type HCN1) and E235R (+3.3 +/- 1.4 mV) channels, as if surface charges were already shielded. Consistent with an electrostatic role, the energetic changes associated with Delta V(12) resulting from various Glu(235) substitutions (i.e. Asp, Ala, Pro, His, Lys, and Arg) displayed a strong correlation with their charges (Delta Delta G = -2.1 +/- 0.3 kcal/mol/charge; r = 0.94). In contrast, D233E, D233A, D233G, and D233R did not alter activation gating. D233C (in C318S background) was also not externally accessible when probed with methanethiosulfonate ethylammonium (MTSEA). We conclude that the S3-S4 linker residue Glu(235) influences activation gating, probably by acting as a surface charge.  相似文献   

12.
The effects of the agonist enantiomer S(-)Bay K 8644 on gating charge of L-type Ca channels were studied in single ventricular myocytes. From a holding potential (Vh) of -40 mV, saturating (250 nm) S(-)Bay K shifted the half-distribution voltage of the activation charge (Q1) vs. V curve -7.5 +/- 0.8 mV, almost identical to the shift produced in the Ba conductance vs. V curve (-7.7 +/- 2 mV). The maximum Q1 was reduced by 1.7 +/- 0.2 nC/microF, whereas Q2 (charge moved in inactivated channels) was increased in a similar amount (1.4 +/- 0.4 nC/microF). The steady-state availability curves for Q1, Q2, and Ba current showed almost identical negative shifts of -14.8 +/- 1.7 mV, -18.6 +/- 5.8 mV, and -15.2 +/- 2.7 mV, respectively. The effects of the antagonist enantiomer R(+)BayK 8644 were also studied, the Q1 vs. V curve was not significantly shifted, but Q1max (Vh = -40 mV) was reduced and the Q1 availability curve shifted by -24.6 +/- 1.2 mV. We concluded that: a) the left shift in the Q1 vs. V activation curve produced by S(-)BayK is a purely agonistic effect; b) S(-)BayK induced a significantly larger negative shift in the availability curve than in the Q1 vs. V relation, consistent with a direct promotion of inactivation; c) as expected for a more potent antagonist, R(+)Bay K induced a significantly larger negative shift in the availability curve than did S(-)Bay K.  相似文献   

13.
Activation of a two-dimensional sheet network (5 parallel chains of 5 cells each) of simulated intestinal smooth muscle cells (SMCs) by one interstitial cell of Cajal (ICC) was modeled by PSpice simulation. The network of 25 cells was not interconnected by gap-junction channels; instead, excitation was transmitted by the electric field that develops in the junctional clefts (JC) when the prejunctional membrane fires an action potential (AP). Transverse propagation between the parallel chains occurs similarly. The ICC cell was connected to cell E5 of the network [5th cell of the 5th (E) chain] via a high-resistance junction. The stimulating current, applied to the ICC cell interior, was made to resemble the endogenous undershooting slow wave (I(SW)). An I(SW) of 2.4 nA (over a rise time of 4 ms) took the ICC cell from a resting potential (RP) of -80 mV to a membrane potential of -41 mV. The slow wave produced a large negative cleft potential in the JC (V(JC); ICC-E5). The V(jc) brought the postjunctional membrane of E5 to threshold, causing this cell to fire an AP. This, in turn, propagated throughout the SMC network. If the ICC cell was given an RP of -55 mV (like SMC) and a slow wave of 40 mV amplitude (I(SW) of 1.8 nA), it still activated the SMC network. This was also true when the ICC cell was made excitable (developing an overshooting, fast-rising AP). In summary, one ICC cell displaying a slow wave was capable of activating a network of SMC in the absence of gap junctions.  相似文献   

14.
The stoichiometry and voltage dependence of the Na/K pump were studied in internally dialyzed, voltage-clamped squid giant axons by simultaneously measuring, at various membrane potentials, the changes in Na efflux (delta phi Na) and holding current (delta I) induced by dihydrodigitoxigenin (H2DTG). H2DTG stops the Na/K pump without directly affecting other current pathways: (a) it causes no delta I when the pump lacks Na, K, Mg, or ATP, and (b) ouabain causes no delta I or delta phi Na in the presence of saturating H2DTG. External K (Ko) activates Na efflux with Michaelis-Menten kinetics (Km = 0.45 +/- 0.06 mM [SEM]) in Na-free seawater (SW), but with sigmoid kinetics in approximately 400 mM Na SW (Hill coefficient = 1.53 +/- 0.08, K1/2 = 3.92 +/- 0.29 mM). H2DTG inhibits less strongly (Ki = 6.1 +/- 0.3 microM) in 1 or 10 mM K Na-free SW than in 10 mM K, 390 mM Na SW (1.8 +/- 0.2 microM). Dialysis with 5 mM each ATP, phosphoenolpyruvate, and phosphoarginine reduced Na/Na exchange to at most 2% of the H2DTG-sensitive Na efflux. H2DTG sensitive but nonpump current caused by periaxonal K accumulation upon stopping the pump, was minimized by the K channel blockers 3,4-diaminopyridine (1 mM), tetraethylammonium (approximately 200 mM), and phenylpropyltriethylammonium (20-25 mM) whose adequacy was tested by varying [K]o (0-10 mM) with H2DTG present. Two ancillary clamp circuits suppressed stray current from the axon ends. Current and flux measured from the center pool derive from the same membrane area since, over the voltage range -60 to +20 mV, tetrodotoxin-sensitive current and Na efflux into Na-free SW, under K-free conditions, were equal. The stoichiometry and voltage dependence of pump Na/K exchange were examined at near-saturating [ATP], [K]o and [Na]i in both Na-free and 390 mM Na SW. The H2DTG-sensitive F delta phi Na/delta I ratio (F is Faraday's constant) of paired measurements corrected for membrane area match, was 2.86 +/- 0.09 (n = 8) at 0 mV and 3.05 +/- 0.13 (n = 6) at -60 to -90 mV in Na-free SW, and 2.72 +/- 0.09 (n = 7) at 0 mV and 2.91 +/- 0.21 (n = 4) at -60 mV in 390 mM Na SW. Its overall mean value was 2.87 +/- 0.07 (n = 25), which was not significantly different from the 3.0 expected of a 3 Na/2 K pump.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

15.
Low conductance sodium channels in canine cardiac Purkinje cells.   总被引:4,自引:0,他引:4       下载免费PDF全文
Low conductance sodium (Na) channels have been observed in nerve, skeletal muscle, and cardiac cells. In cardiac tissues the higher amplitude, more commonly observed Na channel was first investigated in detail by Cachelin et al. (Cachelin, A.B., J.E. de Peyer, S. Kokubun, and H. Reuter, 1983, J. Physiol. (Lond.), 340:389-402). They also reported low amplitude Na channel events. We have studied this low conductance Na channel in single canine cardiac Purkinje cells using cell-attached patches. Patch pipette solutions contained either 140 or 280 mM NaCl, and cells were bathed in a solution of 150 mM KCl to bring their resting potential close to zero. In 140 mM Na+, during steps to -50 mV, the lower and higher openings had amplitudes of 0.57 +/- 0.2 and 1.2 +/- 0.2 pA (means +/- SD of Gaussian fits). In 280 mM Na+ at -50 mV, amplitudes were 0.72 +/- 0.2 and 1.55 +/- 0.2 pA. Over a substantial voltage range, the lower events had amplitudes of about one-third that of the higher events. The frequency of the low conductance openings varied in different patches from zero to 22% of total openings. Histograms of open durations and latencies at several voltages suggested no difference in kinetics between the two channel events. The behavior of the low conductance channels was more consistent with a second population of channels rather than a second open state.  相似文献   

16.
Endothelium-derived hyperpolarizing factor (EDHF) is released in response to agonists such as ACh and bradykinin and regulates vascular smooth muscle tone. Several studies have indicated that ouabain blocks agonist-induced, endothelium-dependent hyperpolarization of smooth muscle. We have demonstrated that epoxyeicosatrienoic acids (EETs), cytochrome P-450 metabolites of arachidonic acid, function as EDHFs. To further test the hypothesis that EETs represent EDHFs, we have examined the effects of ouabain on the electrical and mechanical effects of 14,15- and 11,12-EET in bovine coronary arteries. These arteries are relaxed in a concentration-dependent manner to 14,15- and 11,12-EET (EC(50) = 6 x 10(-7) M), bradykinin (EC(50) = 1 x 10(-9) M), sodium nitroprusside (SNP; EC(50) = 2 x 10(-7) M), and bimakalim (BMK; EC(50) = 1 x 10(-7) M). 11,12-EET-induced relaxations were identical in vessels with and without an endothelium. Potassium chloride (1-15 x 10(-3) M) inhibited [(3)H]ouabain binding to smooth muscle cells but failed to relax the arteries. Ouabain (10(-5) to 10(-4) M) increased basal tone and inhibited the relaxations to bradykinin, 11,12-EET, and 14,15-EET, but not to SNP or BMK. Barium (3 x 10(-5) M) did not alter EET-induced relaxations and ouabain plus barium was similar to ouabain alone. Resting membrane potential (E(m)) of isolated smooth muscle cells was -50.2 +/- 0.5 mV. Ouabain (3 x 10(-5) and 1 x 10(-4) M) decreased E(m) (-48.4 +/- 0.2 mV), whereas 11,12-EET (10(-7) M) increased E(m) (-59.2 +/- 2.2 mV). Ouabain inhibited the 11,12-EET-induced increase in E(m). In cell-attached patch clamp studies, 11,12-EET significantly increased the open-state probability (NP(o)) of a calcium-activated potassium channel compared with control cells (0.26 +/- 0.06 vs. 0.02 +/- 0.01). Ouabain did not change NP(o) but blocked the 14,15-EET-induced increase in NP(o). These results indicate that: 1) EETs relax coronary arteries in an endothelium-independent manner, 2) unlike EETs, potassium chloride does not relax the coronary artery, and 3) ouabain inhibits bradykinin- and EET-induced relaxations as has been reported for EDHF. These findings provide further evidence that EETs are EDHFs.  相似文献   

17.
The 6 cpm omnipresent slow waves recorded in the circular muscle (CM) layer of canine colon are generated at the submucosal surface of the CM layer. After removal of the submucosal network of interstitial cells of Cajal (ICC), 66% of the CM preparations (25 of 38) were quiescent in Krebs solution. In the presence of carbachol, seven of nine of these spontaneously quiescent CM preparations demonstrated slow wave-like activity with mean frequency, duration and amplitude of 5.9 +/- 0.4 cpm, 2.8 +/- 0.5 s, and 0.8 +/- 0.2 mV, respectively. Similar slow wave-like activities were induced by TEA (seven out of eight quiescent CM preparations) with frequency, duration and amplitude of 6.1 +/- 0.2 cpm, 2.7 +/- 0.5 s, and 1.0 +/- 0.2 mV, respectively, and by BaCl2 (eight of eight quiescent CM preparations) with frequency, duration, and amplitude of 6.3 +/- 0.3 cpm, 1.8 +/- 0.2 s, and 0.5 +/- 0.1 mV, respectively. All the induced activities were abolished in the presence of 1 microM D600. CM preparations with the submucosal ICC network intact (ICC-CM) showed slow wave activity in Krebs solution at a frequency of 6.2 +/- 0.2 cpm, a duration of 3.6 +/- 0.2 s, and an amplitude of 1.0 +/- 0.1 mV (n = 22). When ICC-CM preparations were stimulated by BaCl2, carbachol, or TEA, the slow wave frequency did not change significantly, but the duration increased as well as the amplitude. In the presence of D600, the upstroke of slow waves remained and the frequency was not affected.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
The mid-point potentials of the Fe protein components (Ac2 and Ac2* respectively) of the Mo nitrogenase and V nitrogenase from Azotobacter chroococcum were determined in the presence of MgADP to be -450 mV (NHE) [Ac2(MgADP)2-Ac2*ox.(MgADP)2 couple] and -463 mV (NHE) [Ac2* (MgADP)2-Ac2*ox.(ADP)2 couple] at 23 degrees C at pH 7.2. These values are consistent with a flavodoxin characterized by Deistung & Thorneley [(1986) Biochem. J. 239, 69-75] with Em = -522 mV (NHE) being an effective electron donor to both the Mo nitrogenase and the V nitrogenase in vivo. Ac2*ox.(MgADP)2 and Ac2*ox.(MgADP)2 were reduced by SO2.- (formed by the predissociation of dithionite ion, S2O4(2-)) at similar rates, k = 4.7 X 10(6) +/- 0.5 X 10(6) M-1.s-1 and 3.2 X 10(6) +/- 0.2 X 10(6) M-1.s-1 respectively, indicating structural homology at the electron-transfer site associated with the [4Fe-4S] centre in these proteins.  相似文献   

19.
Mechanism of reaction of myeloperoxidase with nitrite   总被引:10,自引:0,他引:10  
Myeloperoxidase (MPO) is a major neutrophil protein and may be involved in the nitration of tyrosine residues observed in a wide range of inflammatory diseases that involve neutrophils and macrophage activation. In order to clarify if nitrite could be a physiological substrate of myeloperoxidase, we investigated the reactions of the ferric enzyme and its redox intermediates, compound I and compound II, with nitrite under pre-steady state conditions by using sequential mixing stopped-flow analysis in the pH range 4-8. At 15 degrees C the rate of formation of the low spin MPO-nitrite complex is (2.5 +/- 0.2) x 10(4) m(-1) s(-1) at pH 7 and (2.2 +/- 0.7) x 10(6) m(-1) s(-1) at pH 5. The dissociation constant of nitrite bound to the native enzyme is 2.3 +/- 0.1 mm at pH 7 and 31.3 +/- 0.5 micrometer at pH 5. Nitrite is oxidized by two one-electron steps in the MPO peroxidase cycle. The second-order rate constant of reduction of compound I to compound II at 15 degrees C is (2.0 +/- 0.2) x 10(6) m(-1) s(-1) at pH 7 and (1.1 +/- 0.2) x 10(7) m(-1) s(-1) at pH 5. The rate constant of reduction of compound II to the ferric native enzyme at 15 degrees C is (5.5 +/- 0.1) x 10(2) m(-1) s(-1) at pH 7 and (8.9 +/- 1.6) x 10(4) m(-1) s(-1) at pH 5. pH dependence studies suggest that both complex formation between the ferric enzyme and nitrite and nitrite oxidation by compounds I and II are controlled by a residue with a pK(a) of (4.3 +/- 0.3). Protonation of this group (which is most likely the distal histidine) is necessary for optimum nitrite binding and oxidation.  相似文献   

20.
Transepithelial potential (V(T)), conductance (G(T)), and water flow (J(V)) were measured simultaneously with good time resolution (min) in isolated toad (Bufo bufo) skin epithelium with Ringer on both sides. Inside application of 5 microM isoproterenol resulted in the fast increase in G(T) from 1.2+/-0.3 to 2.4+/-0.4 mS x cm(-2) and slower increases in equivalent short circuit current, I(SC)(Eqv) = -G(T) x V(T), from 12.7+/-3.2 to 33.1+/-6.8 microA cm(-2), and J(V) from 0.72+/-0.17 to 3.01+/-0.49 nL cm(-2) s(-1). Amiloride in the outside solution abolished I(SC)(Eqv) (-1.6+/-0.1 microA cm(-2)) while J(V) decreased to 0.50+/-0.15 nL cm(-2) x s(-1), which is significantly different from zero. Isoproterenol decreased the osmotic concentration of the transported fluid, C(osm) approximately 2 x I(SC)(Eqv)/J(V), from 351+/-72 to 227+/-28 mOsm (Ringer's solution: 252.8 mOsm). J(V) depicted a saturating function of [Na+]out in agreement with Na+ self-inhibition of ENaC. Ouabain on the inside decreased I(SC)(Eqv) from 60+/-10 to 6.1+/-1.7 microA cm(-2), and J(V) from 3.34+/-0.47 to 1.40+/-0.24 nL cm(-2) x s(-1). Short-circuited preparations exhibited a linear relationship between short-circuit current and J(V) with a [Na+] of the transported fluid of 130+/-24 mM ([Na+]Ringer's solution = 117.4 mM). Addition of bumetanide to the inside solution reduced J(V). Water was transported uphill and J(V) reversed at an excess outside osmotic concentration, deltaC(S,rev) = 28.9+/-3.9 mOsm, amiloride decreased deltaC(S,rev) to 7.5+/-1.5 mOsm. It is concluded that water uptake is accomplished by osmotic coupling in the lateral intercellular space (lis), and hypothesized that a small fraction of the Na+ flux pumped into lis is recirculated via basolateral NKCC transporters.  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