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 93 毫秒
1.
The two electrode voltage clamp technique was used to investigate the steady-state and presteady-state kinetic properties of the type II Na+/P i cotransporter NaPi-5, cloned from the kidney of winter flounder (Pseudopleuronectes americanus) and expressed in Xenopus laevis oocytes. Steady-state P i -induced currents had a voltage-independent apparent K m for P i of 0.03 mm and a Hill coefficient of 1.0 at neutral pH, when superfusing with 96 mm Na+. The apparent K m for Na+ at 1 mm P i was strongly voltage dependent (increasing from 32 mm at −70 mV to 77 mm at −30 mV) and the Hill coefficient was between 1 and 2, indicating cooperative binding of more than one Na+ ion. The maximum steady-state current was pH dependent, diminishing by 50% or more for a change from pH 7.8 to pH 6.3. Voltage jumps elicited presteady-state relaxations in the presence of 96 mm Na+ which were suppressed at saturating P i (1 mm). Relaxations were absent in non-injected oocytes. Charge was balanced for equal positive and negative steps, saturated at extremes of potential and reversed at the holding potential. Fitting the charge transfer to a Boltzmann relationship typically gave a midpoint voltage (V 0.5) close to zero and an apparent valency of approximately 0.6. The maximum steady-state transport rate correlated linearly with the maximum P i -suppressed charge movement, indicating that the relaxations were NaPi-5-specific. The apparent transporter turnover was estimated as 35 sec−1. The voltage dependence of the relaxations was P i -independent, whereas changes in Na+ shifted V 0.5 to −60 mV at 25 mm Na+. Protons suppressed relaxations but contributed to no detectable charge movement in zero external Na+. The voltage dependent presteady-state behavior of NaPi-5 could be described by a 3 state model in which the partial reactions involving reorientation of the unloaded carrier and binding of Na+ contribute to transmembrane charge movement. Received: 11 March 1997/Revised: 3 June 1997  相似文献   

2.
Ionic (Ii) and gating currents (Ig) from noninactivating Shaker H4 K+ channels were recorded with the cut-open oocyte voltage clamp and macropatch techniques. Steady state and kinetic properties were studied in the temperature range 2–22°C. The time course of Ii elicited by large depolarizations consists of an initial delay followed by an exponential rise with two kinetic components. The main Ii component is highly temperature dependent (Q10 > 4) and mildly voltage dependent, having a valence times the fraction of electric field (z) of 0.2–0.3 eo. The Ig On response obtained between −60 and 20 mV consists of a rising phase followed by a decay with fast and slow kinetic components. The main Ig component of decay is highly temperature dependent (Q10 > 4) and has a z between 1.6 and 2.8 eo in the voltage range from −60 to −10 mV, and ∼0.45 eo at more depolarized potentials. After a pulse to 0 mV, a variable recovery period at −50 mV reactivates the gating charge with a high temperature dependence (Q10 > 4). In contrast, the reactivation occurring between −90 and −50 mV has a Q10 = 1.2. Fluctuation analysis of ionic currents reveals that the open probability decreases 20% between 18 and 8°C and the unitary conductance has a low temperature dependence with a Q10 of 1.44. Plots of conductance and gating charge displacement are displaced to the left along the voltage axis when the temperature is decreased. The temperature data suggests that activation consists of a series of early steps with low enthalpic and negative entropic changes, followed by at least one step with high enthalpic and positive entropic changes, leading to final transition to the open state, which has a negative entropic change.  相似文献   

3.
Inner hair cells (IHCs), the primary sensory receptors of the mammalian cochlea, fire spontaneous Ca2+ action potentials before the onset of hearing. Although this firing activity is mainly sustained by a depolarizing L-type (CaV1.3) Ca2+ current (I Ca), IHCs also transiently express a large Na+ current (I Na). We aimed to investigate the specific contribution of I Na to the action potentials, the nature of the channels carrying the current and whether the biophysical properties of I Na differ between low- and high-frequency IHCs. We show that I Na is highly temperature-dependent and activates at around −60 mV, close to the action potential threshold. Its size was larger in apical than in basal IHCs and between 5% and 20% should be available at around the resting membrane potential (−55 mV/−60 mV). However, in vivo the availability of I Na could potentially increase to >60% during inhibitory postsynaptic potential activity, which transiently hyperpolarize IHCs down to as far as −70 mV. When IHCs were held at −60 mV and I Na elicited using a simulated action potential as a voltage command, we found that I Na contributed to the subthreshold depolarization and upstroke of an action potential. We also found that I Na is likely to be carried by the TTX-sensitive channel subunits NaV1.1 and NaV1.6 in both apical and basal IHCs. The results provide insight into how the biophysical properties of I Na in mammalian cochlear IHCs could contribute to the spontaneous physiological activity during cochlear maturation in vivo.  相似文献   

4.
In the preceding paper (Bevensee, M.O., R.A. Weed, and W.F. Boron. 1997. J. Gen. Physiol. 110: 453–465.), we showed that a Na+-driven influx of HCO3 causes the increase in intracellular pH (pHi) observed when astrocytes cultured from rat hippocampus are exposed to 5% CO2/17 mM HCO3 . In the present study, we used the pH-sensitive fluorescent indicator 2′,7′-biscarboxyethyl-5,6-carboxyfluorescein (BCECF) and the perforated patch-clamp technique to determine whether this transporter is a Na+-driven Cl-HCO3 exchanger, an electrogenic Na/HCO3 cotransporter, or an electroneutral Na/HCO3 cotransporter. To determine if the transporter is a Na+-driven Cl-HCO3 exchanger, we depleted the cells of intracellular Cl by incubating them in a Cl-free solution for an average of ∼11 min. We verified the depletion with the Cl-sensitive dye N-(6-methoxyquinolyl)acetoethyl ester (MQAE). In Cl-depleted cells, the pHi still increases after one or more exposures to CO2/HCO3 . Furthermore, the pHi decrease elicited by external Na+ removal does not require external Cl. Therefore, the transporter cannot be a Na+-driven Cl-HCO3 exchanger. To determine if the transporter is an electrogenic Na/ HCO3 cotransporter, we measured pHi and plasma membrane voltage (Vm) while removing external Na+, in the presence/absence of CO2/HCO3 and in the presence/absence of 400 μM 4,4′-diisothiocyanatostilbene-2,2′-disulphonic acid (DIDS). The CO2/HCO3 solutions contained 20% CO2 and 68 mM HCO3 , pH 7.3, to maximize the HCO3 flux. In pHi experiments, removing external Na+ in the presence of CO2/HCO3 elicited an equivalent HCO3 efflux of 281 μM s−1. The HCO3 influx elicited by returning external Na+ was inhibited 63% by DIDS, so that the predicted DIDS-sensitive Vm change was 3.3 mV. Indeed, we found that removing external Na+ elicited a DIDS-sensitive depolarization that was 2.6 mV larger in the presence than in the absence of CO2/ HCO3 . Thus, the Na/HCO3 cotransporter is electrogenic. Because a cotransporter with a Na+:HCO3 stoichiometry of 1:3 or higher would predict a net HCO3 efflux, rather than the required influx, we conclude that rat hippocampal astrocytes have an electrogenic Na/HCO3 cotransporter with a stoichiometry of 1:2.  相似文献   

5.
6.
Effects of odorants on voltage-gated ionic channels were investigated in isolated newt olfactory receptor cells by using the whole cell version of the patch–clamp technique. Under voltage clamp, membrane depolarization to voltages between −90 mV and +40 mV from a holding potential (Vh) of −100 mV generated time- and voltage-dependent current responses; a rapidly (< 15 ms) decaying initial inward current and a late outward current. When odorants (1 mM amyl acetate, 1 mM acetophenone, and 1 mM limonene) were applied to the recorded cell, the voltage-gated currents were significantly reduced. The dose-suppression relations of amyl acetate for individual current components (Na+ current: INa, T-type Ca2+ current: ICa,T, L-type Ca2+ current: ICa,L, delayed rectifier K+ current: IKv and Ca2+-activated K+ current: IK(Ca)) could be fitted by the Hill equation. Half-blocking concentrations for each current were 0.11 mM (INa), 0.15 mM (ICa,T), 0.14 mM (ICa,L), 1.7 mM (IKv), and 0.17 mM (IK(Ca)), and Hill coefficient was 1.4 (INa), 1.0 (ICa,T), 1.1 (ICa,L), 1.0 (IKv), and 1.1 (IK(Ca)), suggesting that the inward current is affected more strongly than the outward current. The activation curve of INa was not changed significantly by amyl acetate, while the inactivation curve was shifted to negative voltages; half-activation voltages were −53 mV at control, −66 mV at 0.01 mM, and −84 mV at 0.1 mM. These phenomena are similar to the suppressive effects of local anesthetics (lidocaine and benzocaine) on INa in various preparations, suggesting that both types of suppression are caused by the same mechanism. The nonselective blockage of ionic channels observed here is consistent with the previous notion that the suppression of the transduction current by odorants is due to the direst blockage of transduction channels.  相似文献   

7.

Background

Orthophosphate (Pi) is a central compound in the metabolism of all organisms, including parasites. There are no reports regarding the mechanisms of Pi acquisition by Trypanosoma cruzi.

Methods

32Pi influx was measured in T. cruzi epimastigotes. The expression of Pi transporter genes and the coupling of the uptake to Na+, H+ and K+ fluxes were also investigated. The transport capacities of different evolutive forms were compared.

Results

Epimastigotes grew significantly more slowly in 2 mM than in 50 mM Pi. Influx of Pi into parasites grown under low Pi conditions took place in the absence and presence of Na+. We found that the parasites express TcPho84, a H+:Pi-symporter, and TcPho89, a Na+:Pi-symporter. Both Pi influx mechanisms showed Michaelis–Menten kinetics, with a one-order of magnitude higher affinity for the Na+-dependent system. Collapsing the membrane potential with carbonylcyanide-p-trifluoromethoxyphenylhydrazone strongly impaired the influx of Pi. Valinomycin (K+ ionophore) or SCH28028 (inhibitor of (H+ + K+)ATPase) significantly inhibited Pi uptake, indicating that an inwardly-directed H+ gradient energizes uphill Pi entry and that K+ recycling plays a key role in Pi influx. Furosemide, an inhibitor of the ouabain-insensitive Na+-ATPase, decreased only the Na+-dependent Pi uptake, indicating that this Na+ pump generates the Na+ gradient utilized by the symporter. Trypomastigote forms take up Pi inefficiently.

Conclusions

Pi starvation stimulates membrane potential-sensitive Pi uptake through different pathways coupled to Na+ or H+/K+ fluxes.

General significance

This study unravels the mechanisms of Pi acquisition by T. cruzi, a key process in epimastigote development and differentiation to trypomastigote forms.  相似文献   

8.
9.
Freshly dissociated myocytes from nonpregnant, pregnant, and postpartum rat uteri have been studied with the tight-seal patch-clamp method. The inward current contains both INa and ICa that are vastly different from those in tissue-cultured material. INa is abolished by Na+-free medium and by 1 μM tetrodotoxin. It first appears at ∼−40 mV, reaches maximum at 0 mV, and reverses at 84 mV. It activates with a voltage-dependent τ of 0.2 ms at 20 mV, and inactivates as a single exponential with a τ of 0.4 ms. Na+ conductance is half activated at −21.5 mV, and half inactivated at −59 mV. INa reactivates with a τ of 20 ms. ICa is abolished by Ca2+-free medium, Co2+ (5 mM), or nisoldipine (2 μM), and enhanced in 30 mM Ca2+, Ba2+, or BAY-K 8644. It first appears at ∼−30 mV and reaches maximum at +10 mV. It activates with a voltage-dependent τ of 1.5 ms at 20 mV, and inactivates in two exponential phases, with τ''s of 33 and 133 ms. Ca2+ conductance is half activated at −7.4 mV, and half inactivated at −34 mV. ICa reactivates with τ''s of 27 and 374 ms. INa and ICa are seen in myocytes from nonpregnant estrus uteri and throughout pregnancy, exhibiting complex changes. The ratio of densities of peak INa/ICa changes from 0.5 in the nonpregnant state to 1.6 at term. The enhanced role of INa, with faster kinetics, allows more frequent repetitive spike discharges to facilitate simultaneous excitation of the parturient uterus. In postpartum, both currents decrease markedly, with INa vanishing from most myocytes. Estrogen-enhanced genomic influences may account for the emergence of INa, and increased densities of INa and ICa as pregnancy progresses. Other influences may regulate varied channel expression at different stages of pregnancy.  相似文献   

10.
A voltage-gated K+ conductance resembling that of the human ether-à-go-go-related gene product (HERG) was studied using whole-cell voltage-clamp recording, and found to be the predominant conductance at hyperpolarized potentials in a cell line (MLS-9) derived from primary cultures of rat microglia. Its behavior differed markedly from the classical inward rectifier K+ currents described previously in microglia, but closely resembled HERG currents in cardiac muscle and neuronal tissue. The HERG-like channels opened rapidly on hyperpolarization from 0 mV, and then decayed slowly into an absorbing closed state. The peak K+ conductance–voltage relation was half maximal at −59 mV with a slope factor of 18.6 mV. Availability, assessed by a hyperpolarizing test pulse from different holding potentials, was more steeply voltage dependent, and the midpoint was more positive (−14 vs. −39 mV) when determined by making the holding potential progressively more positive than more negative. The origin of this hysteresis is explored in a companion paper (Pennefather, P.S., W. Zhou, and T.E. DeCoursey. 1998. J. Gen. Physiol. 111:795–805). The pharmacological profile of the current differed from classical inward rectifier but closely resembled HERG. Block by Cs+ or Ba2+ occurred only at millimolar concentrations, La3+ blocked with K i = ∼40 μM, and the HERG-selective blocker, E-4031, blocked with K i = 37 nM. Implications of the presence of HERG-like K+ channels for the ontogeny of microglia are discussed.  相似文献   

11.
We studied the regulation of intracellular pH (pHi) in single cultured astrocytes passaged once from the hippocampus of the rat, using the dye 2′,7′-biscarboxyethyl-5,6-carboxyfluorescein (BCECF) to monitor pHi. Intrinsic buffering power (βI) was 10.5 mM (pH unit)−1 at pHi 7.0, and decreased linearly with pHi; the best-fit line to the data had a slope of −10.0 mM (pH unit)−2. In the absence of HCO3 , pHi recovery from an acid load was mediated predominantly by a Na-H exchanger because the recovery was inhibited 88% by amiloride and 79% by ethylisopropylamiloride (EIPA) at pHi 6.05. The ethylisopropylamiloride-sensitive component of acid extrusion fell linearly with pHi. Acid extrusion was inhibited 68% (pHi 6.23) by substituting Li+ for Na+ in the bath solution. Switching from a CO2/HCO3 -free to a CO2/HCO3 -containing bath solution caused mean steady state pHi to increase from 6.82 to 6.90, due to a Na+-driven HCO3 transporter. The HCO3 -induced pHi increase was unaffected by amiloride, but was inhibited 75% (pHi 6.85) by 400 μM 4,4′-diisothiocyanatostilbene-2,2′-disulfonic acid (DIDS), and 65% (pHi 6.55–6.75) by pretreating astrocytes for up to ∼6.3 h with 400 μM 4-acetamide-4′-isothiocyanatostilbene-2,2′-disulfonic acid (SITS). The CO2/HCO3 -induced pHi increase was blocked when external Na+ was replaced with N-methyl-d-glucammonium (NMDG+). In the presence of HCO3 , the Na+-driven HCO3 transporter contributed to the pHi recovery from an acid load. For example, HCO3 shifted the plot of acid-extrusion rate vs. pHi by 0.15–0.3 pH units in the alkaline direction. Also, with Na-H exchange inhibited by amiloride, HCO3 increased acid extrusion 3.8-fold (pHi 6.20). When astrocytes were acid loaded in amiloride, with Li+ as the major cation, HCO3 failed to elicit a substantial increase in pHi. Thus, Li+ does not appear to substitute well for Na+ on the HCO3 transporter. We conclude that an amiloride-sensitive Na-H exchanger and a Na+-driven HCO3 transporter are the predominant acid extruders in astrocytes.  相似文献   

12.
The epithelial sodium channel is a multimeric protein formed by three homologous subunits: α, β, and γ; each subunit contains only two transmembrane domains. The level of expression of each of the subunits is markedly different in various Na+ absorbing epithelia raising the possibility that channels with different subunit composition can function in vivo. We have examined the functional properties of channels formed by the association of α with β and of α with γ in the Xenopus oocyte expression system using two-microelectrode voltage clamp and patch-clamp techniques. We found that αβ channels differ from αγ channels in the following functional properties: (a) αβ channels expressed larger Na+ than Li+ currents (INa+/ILi+ 1.2) whereas αγ channels expressed smaller Na+ than Li+ currents (INa+/ILi+ 0.55); (b) the Michaelis Menten constants (K m) of activation of current by increasing concentrations of external Na+ and Li+ of αβ channels were larger (K m > 180 mM) than those of αγ channels (K m of 35 and 50 mM, respectively); (c) single channel conductances of αβ channels (5.1 pS for Na+ and 4.2 pS for Li+) were smaller than those of αγ channels (6.5 pS for Na+ and 10.8 pS for Li+); (d) the half-inhibition constant (K i) of amiloride was 20-fold larger for αβ channels than for αγ channels whereas the K i of guanidinium was equal for both αβ and αγ. To identify the domains in the channel subunits involved in amiloride binding, we constructed several chimeras that contained the amino terminus of the γ subunit and the carboxy terminus of the β subunit. A stretch of 15 amino acids, immediately before the second transmembrane domain of the β subunit, was identified as the domain conferring lower amiloride affinity to the αβ channels. We provide evidence for the existence of two distinct binding sites for the amiloride molecule: one for the guanidium moiety and another for the pyrazine ring. At least two subunits α with β or γ contribute to these binding sites. Finally, we show that the most likely stoichiometry of αβ and αγ channels is 1α:1β and 1α:1γ, respectively.  相似文献   

13.
(1) The active transport of Na+ across the turtle bladder epithelial cell layer consists of a passive entry step through a Na+-selective path in the apical membrane and an active extrusion step through Na+ pump-containing path in the basal-lateral membrane together with some back-leakage through the paracellular spaces and tight junctions between the epithelial cells. This hypothesis has now been verified qualitatively and to some extent, quantitatively by the use of an intracellularly-located microelectrode in conjunction with a conventional assembly of extracellularly-located macroelectrodes mainly in short-circuited bladders bathed by Na+-rich Ringer media. Under these conditions, the intracellular potential (Vsc) averaged 38.4 mV with the cell electronegative; the fractional resistance of the apical membrane (?Ra) averaged 0.55; while the concomitant transepithelial parameters, short circuiting current (Isc) and electrical conductance (Gt), average 68.6 μA/cm2 and 0.98 mS/cm2, respectively. (2) The relation between these parameters and the transepithelial flow of Na+ (orIsc) is evoked by blocking Na+ entry into the cell (by the mucosal addition of amiloride or removal of mucosal Na+). Amiloride-induced blockade of the Na+ entry step results in a rapid hyperpolarization of the cell interior during which Vsc = —79.1 mV and ?Ra = 0.92. Isc and Gt (equivalent to the shunt conductance under these conditions) averaged 5 μA/cm2 and 0.35 mS/cm2, respectively. The entire process is reversible on re-admission of Na+ entry into the cell. (3) A slow depolarization of the cell interior in the period of blocked transapical Na+ entry is opposite to that expected from an electroneutral Na+-K+ exchanging pump; but instead is the predictable response of an electrogenic Na+ pump in parallel with a passive K+-selective conductance in the basal-lateral membrane. (4) The electrogenicity concept is substantiated after pretreatment of the bladder with serosal ouabain, which changes the response of Vsc to amiloride (from the aforementioned biphasic response) to a step-function response, attributable mainly to the development of a slowly dissipating K+ diffusion potential across the basallateral membrane. (5) Under open-circuit conditions, the electronegativity of cell to mucosa (Va) is a linear inverse function of the electropositivity of serosa to mucosa (Vt). For Vt ? 100 mV, Va is positive; and for Vt between ?30 and 90 mV, Va is negative.  相似文献   

14.
We previously demonstrated a transmural gradient in Na/K pump current (I P) and [Na+] i , with the highest maximum I P and lowest [Na+] i in epicardium. The present study examines the relationship between the transmural gradient in I P and Na/Ca exchange (NCX). Myocytes were isolated from canine left ventricle. Whole-cell patch clamp was used to measure current generated by NCX (I NCX) and inward background calcium current (I ibCa), defined as inward current through Ca2+ channels less outward current through Ca2+-ATPase. When resting myocytes from endocardium (Endo), midmyocardium (Mid) or epicardium (Epi) were studied in the same conditions, I NCX was the same and I ibCa was zero. Moreover, Western blots were consistent with NCX protein being uniform across the wall. However, the gradient in [Na+] i , with I ibCa = 0, should create a gradient in [Ca2+] i . To test this hypothesis, we measured resting [Ca2+] i using two methods, based on either transport or the Ca2+-sensitive dye Fura2. Both methods demonstrated a significant transmural gradient in resting [Ca2+] i , with Endo > Mid > Epi. This gradient was eliminated by exposing Epi to sufficient ouabain to partially inhibit Na/K pumps, thus increasing [Na+] i to values similar to those in Endo. These data support the existence of a transmural gradient for Ca2+ removal by NCX. This gradient is not due to differences in expression of NCX; rather, it is generated by a transmural gradient in [Na+] i , which is due to a transmural gradient in plasma membrane expression of the Na/K pump.  相似文献   

15.
16.
Summary Photoreceptor cells of the drone,Apismellifera , have a voltage-gated Na+ membrane conductance that can be blocked by tetrodotoxin (TTX) and generates an action potential on abrupt depolarization: an action potential is triggered by the rising phase of a receptor potential evoked by an intense light flash (Autrum and von Zwehl 1964; Baumann 1968). We measured the intracellular voltage response to a small (9%), brief (30 ms) decrease in light intensity from a background, and found that its amplitude was decreased by 1M TTX. The response amplitude was maximal when the background intensity depolarized the cell to –38 mV. With intensities depolarizing the cell membrane to –45 to –33 mV the average response amplitude was decreased by TTX from 1.2mV to 0.5mV. TTX is also known to decrease the voltage noise during steady illumination (Ferraro et al. 1983) but, despite this, the ratio of peak-to-peak signal to noise was, on average, decreased by TTX. The results suggest that drone photoreceptors use voltage-gated Na+ channels for graded amplification of responses to small, rapid changes in light intensity.Abbreviations TTX tetrodotoxin - V i intracellular potential with respect to the bath - V o extracellular potential - V m,V i-V o approximate transmembrane potential - S amplitude of the voltage response to an 8.9% decrease in light intensity - N voltage noise, usually measured as root mean square voltage deviation as described in Methods  相似文献   

17.
Extracellular glucoamylase of Colletotrichum sp. KCP1 produced through solid state fermentation was purified by two steps purification process comprising ammonium sulphate precipitation followed by gel permeation chromatography (GPC). The Recovery of glucoamylase after GPC was 50.40 % with 19.3-fold increase in specific activity. The molecular weight of enzyme was found to be 162.18 kDa by native-PAGE and was dimeric protein of two sub-units with molecular weight of 94.62 and 67.60 kDa as determined by SDS-PAGE. Activation energy for starch hydrolysis was 26.45 kJ mol−1 while temperature quotient (Q10) was found to be 1.9. The enzyme was found to be stable over wide pH range and thermally stable at 40–50 °C up to 120 min while exhibited maximum activity at 50 °C with pH 5.0. The pKa1 and pKa2 of ionisable groups of active site controlling Vmax were 3.5 and 6.8, respectively. Vmax, Km and Kcat for starch hydrolysis were found to be 58.82 U ml−1, 1.17 mg (starch) ml−1 and 449 s−1, respectively. Activation energy for irreversible inactivation (Ea(d)) of glucoamylase was 74.85 kJ mol−1. Thermodynamic parameters of irreversible inactivation of glucoamylase and starch hydrolysis were also determined.  相似文献   

18.
Specific-ion effects are ubiquitous in nature; however, their underlying mechanisms remain elusive. Although Hofmeister-ion effects on proteins are observed at higher (>0.3M) salt concentrations, in dilute (<0.1M) salt solutions nonspecific electrostatic screening is considered to be dominant. Here, using effective charge (Q*) measurements of hen-egg white lysozyme (HEWL) as a direct and differential measure of ion-association, we experimentally show that anions selectively and preferentially accumulate at the protein surface even at low (<100 mM) salt concentrations. At a given ion normality (50 mN), the HEWL Q* was dependent on anion, but not cation (Li+, Na+, K+, Rb+, Cs+, GdnH+, and Ca2+), identity. The Q* decreased in the order F > Cl > Br > NO3 ∼ I > SCN > ClO4 ≫ SO42−, demonstrating progressively greater binding of the monovalent anions to HEWL and also show that the SO42− anion, despite being strongly hydrated, interacts directly with the HEWL surface. Under our experimental conditions, we observe a remarkable asymmetry between anions and cations in their interactions with the HEWL surface.  相似文献   

19.
Postinhibitory rebound (PIR) is an intrinsic property often exhibited by neurons involved in generating rhythmic motor behaviors. Cell DE-3, a dorsal excitatory motor neuron in the medicinal leech exhibits PIR responses that persist for several seconds following the offset of hyperpolarizing stimuli and are suppressed in reduced Na+ solutions or by Ca2+ channel blockers. The long duration and Na+ dependence of PIR suggest a possible role for persistent Na+ current (I NaP). In vertebrate neurons, the neuroprotective agent riluzole can produce a selective block of I NaP. This study demonstrates that riluzole inhibits cell DE-3 PIR in a concentration- and Ca2+-dependent manner. In 1.8 mM Ca2+ solution, 50–100 µM riluzole selectively blocked the late phase of PIR, an effect similar to that of the neuromodulator serotonin. However, 200 µM riluzole blocked both the early and late phases of PIR. Increasing extracellular Ca2+ to 10 mM strengthened PIR, but high riluzole concentrations continued to suppress both phases of PIR. These results indicate that riluzole may suppress PIR via a nonspecific inhibition of Ca2+ conductances and suggest that a Ca2+-activated nonspecific current (I CAN), rather than I NaP, may underlie the Na+-dependent component of PIR.  相似文献   

20.
Although the role of Na+ in several aspects of Ca2+ regulation has already been shown, the exact mechanism of intracellular Ca2+ concentration ([Ca2+]i) increase resulting from an enhancement in the persistent, non‐inactivating Na+ current (INa,P), a decisive factor in certain forms of epilepsy, has yet to be resolved. Persistent Na+ current, evoked by veratridine, induced bursts of action potentials and sustained membrane depolarization with monophasic intracellular Na+ concentration ([Na+]i) and biphasic [Ca2+]i increase in CA1 pyramidal cells in acute hippocampal slices. The Ca2+ response was tetrodotoxin‐ and extracellular Ca2+‐dependent and ionotropic glutamate receptor‐independent. The first phase of [Ca2+]i rise was the net result of Ca2+ influx through voltage‐gated Ca2+ channels and mitochondrial Ca2+ sequestration. The robust second phase in addition involved reverse operation of the Na+–Ca2+ exchanger and mitochondrial Ca2+ release. We excluded contribution of the endoplasmic reticulum. These results demonstrate a complex interaction between persistent, non‐inactivating Na+ current and [Ca2+]i regulation in CA1 pyramidal cells. The described cellular mechanisms are most likely part of the pathomechanism of certain forms of epilepsy that are associated with INa,P. Describing the magnitude, temporal pattern and sources of Ca2+ increase induced by INa,P may provide novel targets for antiepileptic drug therapy.  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