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 used isothermal titration calorimetry in the temperature range 21-25 degrees C to investigate the effect of pH on the calorimetric enthalpy (delta H(cal)) for sequence specific DNA-binding of the glucocorticoid receptor DNA-binding domain (GR DBD). Titrations were carried out in solutions containing 100 mM NaCl, 1 mM dithiothreitol, 5% glycerol by volume, and 20 mM Tris, Hepes, Mops, or sodium phosphate buffers at pH 7.5. A strong dependence of delta H(cal) on the buffer ionization enthalpy is observed, demonstrating that the DNA binding of the GR DBD is linked to proton uptake at these conditions. The apparent increase in the pK(a) for an amino acid side chain upon DNA binding is supported by the results of complementary titrations, where delta H(cal) shows a characteristic dependence on the solution pH. delta H(cal) is also a function of the NaCl concentration, with opposite dependencies in Tris and Hepes buffers, respectively, such that a similar delta H(cal) value is approached at 300 mM NaCl. This behavior shows that the DNA-binding induced protonation is inhibited by increased concentrations of NaCl. A comparison with structural data suggests that the protonation involves a histidine (His451) in the GR DBD, because in the complex this residue is located close to a DNA phosphate at an orientation that is consistent with a charged-charged hydrogen bond in the protonated state. NMR spectra show that His451 is not protonated in the unbound protein at pH 7.5. The pH dependence in delta H(cal) can be quantitatively described by a shift of the pK(a) of His451 from approximately 6 in the unbound state to close to 8 when bound to DNA at low salt concentration conditions. A simple model involving a binding competition between a proton and a Na(+) counterion to the GR DBD-DNA complex reproduces the qualitative features of the salt dependence.  相似文献   

2.
The sign of B2, the micro-second component of the photocurrent from oriented purple membrane, is that of positive charge moving away from the purple membrane in the direction of proton release. B2 could be due to internal dipole or proton movement, proton release, or metal cation release. We found that the waveform of B2 is virtually insensitive to changes in the salt concentration as long as it is >40 mM KCl, >5 mM CaCl2, or >0.5 mM LaCl3. However, below these limits, B2's apparent rate of decay increases as the salt concentration decreases without any change in the initial amplitude. This salt dependence suggests that B2 is due to a positive charge, either a metal cation or a proton, moving from the membrane into the solution. That the positive charge is not a metal cation is suggested by the waveform of B2 remaining unchanged upon replacing the cations both in solution and in the binding sites of the purple membrane. Direct evidence that the positive charge movement is due to protons was obtained by examining the correlation of B2 with the proton dependent processes of bacteriorhodopsin in buffers and dyes. Based on these observations, we suggest that most, if not all, of the intrinsic B2 component of the photocurrent at moderate salt concentration is due to proton release.

The photocurrents from purple membranes whose surface potential has been reduced by delipidation or chemical modification of carboxyl groups with methyl esters were found to be only modestly changed. This suggests that the salt effect is not through its modulation of the surface potential. Rather, we propose that in low salt B2 represents the sum of a proton release from the surface of the purple membrane and a second current component, due to cations moving back towards the membrane, which is only important in low salt. The cation counter current is induced by proton release which creates a transient uncompensated negative charge on the membrane.

  相似文献   

3.
Buffers change the electric signals of light-excited bacteriorhodopsin molecules in purple membrane if their concentration and the pH of the low-salt solution are properly selected. "Positive" buffers produce a positive component, and "negative" buffers a negative component in addition to the signals due to proton pumping. Measurement of the buffer effects in the presence of glycyl-glycine or bis-tris propane revealed an increase of approximately 2 and a change of sign and a decrease to approximately -0.5 in the translocated charge in these cases, respectively. These factors do not depend on temperature. The Arrhenius parameters established from the evaluation of the kinetics indicate activation enthalpies of 35-40 kJ/mol and negative activation entropies for the additional signals. These values agree with those found by surface-bound pH-sensitive probes in the search of the timing of proton release and uptake. The electric signals were also measured in the case of D(2)O solutions with similar results, except for the increased lifetimes. We offer a unified explanation for the data obtained with surface-bound probes and electric signals based on the clusters at extracellular and cytoplasmic sites of bacteriorhodopsin participating in proton release and uptake.  相似文献   

4.
The acoustic absorption of protein solutions in the presence of phosphate and other buffering ions has been studied in the physiological pH range. Buffers containing hydroxyl residues as titratable groups cause a pronounced increase of protein sound absorption, which is attributed to relaxation processes of proton transfer reactions between buffer ions and accessible imidazole and -amino groups of the protein surface. Amino group based buffers like Good's buffers do not induce additional sound absorption. Measurement of the ultrasonic absorption as a function of pH and of buffer concentration, and corresponding parameter fitting of the equation describing proton transfer relaxation processes has been used to evaluate equilibrium parameters. For the imidazole group of the amino acid histidine a pK value of 6.22 and for the imidazole group of the protein lysozyme a pK value of 5.71 have been determined. In hemoglobin the ligand-linked pK changes have been monitored by recording ultrasonic titration curves.  相似文献   

5.
The binding of aminoglycoside antibiotics to a broad range of macromolecular targets is coupled to protonation of one or more of the amino groups that typify this class of drugs. Determining how and to what extent this linkage influences the energetics of the aminoglycoside-macromolecule binding reaction requires a detailed understanding of the thermodynamics associated with the protonation equilibria of the aminoglycoside amino groups. In recognition of this need, a calorimetric- and NMR-based approach for obtaining the requisite thermodynamic information is presented using paromomycin as the model aminoglycoside. Temperature- and pH-dependent 15N NMR studies provide pK(a) values for the five paromomycin amino groups, as well as the temperature dependence of these pK(a) values. These studies also indicate that the observed pK(a) values associated with the free base form of paromomycin are lower in magnitude than the corresponding values associated with the sulfate salt form of the drug. This difference in pK(a) is due to drug interactions with the sulfate counterions at the high drug concentrations (> or = 812 mM) used in the 15N NMR studies. Isothermal titration calorimetry studies conducted at drug concentrations < or = 45 microM reveal that the extent of paromomycin protonation linked to the binding of the drug to its pharmacologically relevant target, the 16 S rRNA A-site, is consistent with the pK(a) values of the free base and not the sulfate salt form of the drug. Temperature- and pH-dependent isothermal titration calorimetry studies yield exothermic enthalpy changes (deltaH) for protonation of the five paromomycin amino groups, as well as positive heat capacity changes (deltaC(p)) for three of the five amino groups. Regarded as a whole, the results presented here represent an important first step toward establishing a thermodynamic database that can be used to predict how aminoglycoside-macromolecule binding energetics will be influenced by conditions such as temperature, pH, and ionic strength. Such a predictive capability is a critical component of any drug design strategy.  相似文献   

6.
The role of fixed and mobile buffers in the kinetics of proton movement   总被引:9,自引:0,他引:9  
We derive a simple expression for the effective diffusion coefficient of protons in Fick's second law, Deff, when both spatially fixed, HF, and mobile, HM, buffers are present. These buffers are present at moderately high concentrations ([Ftot], [Mtot] greater than 1 mM) in most biological systems. We consider only the case where the protonation reactions remain at equilibrium during the diffusion process. When the pH is to the alkaline side of the pK values of the fixed and mobile buffers ([H+] less than KF, KM), the effective diffusion coefficient of protons in Ficks second law is: (Formula: see text) where DH is the diffusion coefficient of the protons free in the aqueous phase and DHM is the diffusion coefficient of the mobile buffer. The equation illustrates three features of diffusion in a buffered system. Firstly, the effective diffusion coefficient of protons is always lower than the diffusion coefficient of free protons. Secondly, increasing the concentration of fixed buffers always decreases Deff. Thirdly, increasing the concentration of mobile buffer can increase Deff when fixed buffers are present.  相似文献   

7.
The interaction of DNA with various neutral pH, amine-based buffers has been analyzed by free solution capillary electrophoresis, using a mixture of a plasmid-sized DNA molecule and a small DNA oligonucleotide as the reporter system. The two DNAs migrate as separate, nearly Gaussian-shaped peaks in 20-80 mM TAE (TAE, Tris-acetate-EDTA; Tris, tris[hydroxymethyl]aminomethane) buffer. The separation between the peaks gradually increases with increasing TAE buffer concentration because of differences in solvent friction between large and small DNA molecules. The two DNAs form complexes with the borate ions in TBE (Tris-borate-EDTA) buffer, with mobilities that depend on the DNA/borate ratio. In 45 mM TBE buffer, the two DNAs comigrate as a single sharp peak, with a mobility that is faster than either of the constituent DNAs in the same buffer. Hence, the mixed DNA-borate complex is stabilized by the binding of additional borate ions, possibly forming bridges between the different DNAs. The mixed DNA-borate complex is gradually dissociated into its component DNAs by increasing the TBE concentration, possibly because the borate binding sites become saturated at high buffer concentrations. Other neutral pH, amine-based buffers, such as Mops (3-[N-morpholino]propanesulfonic acid), Hepes (N-[2-hydroxyethyl]piperazine-N'-[2-ethanesulfonic acid]), Bes (N,N-bis[2-hydroxyethyl]-2-aminoethanesulfonic acid), Tes (N-tris[hydroxymethyl]methyl-2-aminoethanesulfonic acid), and tricine (N-tris[hydroxymethyl]methylglycine) also form complexes with DNA, giving distorted peaks in the electropherograms. The combined results indicate that borate buffers and most neutral pH, amine-based buffers interact with DNA.  相似文献   

8.
Using neuronal nuclei (N1) isolated from cerebral cortices of 15-day-old rabbits the incorporation of [3H]arachidonate into N1 lipids was followed in vitro. Arachidonate was principally incorporated into triacylglycerol and phosphatidylinositol. When low concentrations (32 mM) of Tris-HC1 (pH 7.4) were used, rates of total arachidonate incorporation were small and phosphatidylinositol received the bulk (greater than 84%) of the arachidonate. When the concentration of Tris-HC1 (pH 7.4) or, in certain cases, the concentration of arachidonate was increased, there was a rise in total arachidonate incorporation into N1, with an increasing proportion of radioactivity entering triacylglycerol until it was the predominantly labelled lipid. Using other buffers (phosphate, imidazole, HEPES, pH 7.4), the shift from phosphatidylinositol to triacylglycerol as principal labelled lipid, with buffer concentration, was not as marked as with Tris-HC1 (pH 7.4). When the buffer concentration was maintained at 107 mM and the pH was lowered to 6.5, the three amine-containing buffers showed a sizeable decline in arachidonate incorporation into N1 lipids and a corresponding decrease in triacylglycerol labelling. The proportion of the total radioactivity in N1 phosphatidylinositol rose as the pH declined. Of the buffers used, Tris-HC1 showed the greatest changes over the pH range. Based upon pK values for the amine buffers, it is suggested that an increased proportion of the protonated amine may be inhibitory to arachidonate incorporation in N1. Studies of acyl-CoA synthetase in N1 indicated this enzyme as the site of the inhibition.  相似文献   

9.
Cells of Staphylococcus aureus heated at 52 degrees C in magnesium-chelating buffers [pH 7.2, 50 mM potassium phosphate or 50 mM tris(hydroxymethyl)-aminomethane containing 1 mM ethylenediaminetetraacetic acid] leaked 260-nm absorbing material, shown to be RNA, and suffered destruction of their ribosomes. These cells did not regain their salt tolerance when repair was carried out in the presence of actinomycin D (5 microgram/ml). Cells similarly heated in magnesium-conserving buffers [pH 7.2, 50 mM tris(hydroxymethyl)aminomethane containing 10 mM MgCl2 or piperazine buffer] did not leak RNA, suffered no ribosomal damage when heated for 15 min, and recovered, at least partially, in the presence of actinomycin D. Ribosomal damage, is therefore, a consequence of Mg2+ loss and is not an effect of heat per se. Cells suspended in either Mg2+-chelating or Mg2+-conserving buffers lost salt tolerance to about the same extent during heating at 52 degrees C. Therefore, sublethal heat injury can not be attributed to ribosomal damage.  相似文献   

10.
We have investigated the effects of extracellular and intracellular pH on single channel and macroscopic (macropatches) currents through volume-regulated anion channels (VRAC) in endothelial cells. Protonation of extracellular binding sites with an apparent pK of 4.6 increased voltage independent of the single-channel amplitude. Cytosolic acidification had a dual effect on VRAC currents: on the one hand, it increased single channel conductance by ∼20% due to protonation of a group with an apparent pK of 6.5 and a Hill coefficient of 2. On the other hand, it reduced channel activity due to protonation of a group with an apparent pK of 6.3 and a Hill coefficient of 2.1. This dual effect enhances the macroscopic current at a slightly acidic pH but inhibits it at more acidic pH. Cytosolic alkalization also reduced channel activity with a pK of 8.4 and a Hill coefficient of 1.9, but apparently did not affect single-channel conductance. These data show that VRAC channels are maintained in an active state in a narrow pH range around the normal physiological pH and shut down outside this range. They also show that HEPES-buffered pipette solutions do not effectively buffer pH in the vicinity of the VRAC channels. Received: 31 January 2000/Revised: 21 April 2000  相似文献   

11.
Inhibition by polyvalent cations is a defining characteristic of voltage-gated proton channels. The mechanism of this inhibition was studied in rat alveolar epithelial cells using tight-seal voltage clamp techniques. Metal concentrations were corrected for measured binding to buffers. Externally applied ZnCl(2) reduced the H(+) current, shifted the voltage-activation curve toward positive potentials, and slowed the turn-on of H(+) current upon depolarization more than could be accounted for by a simple voltage shift, with minimal effects on the closing rate. The effects of Zn(2+) were inconsistent with classical voltage-dependent block in which Zn(2+) binds within the membrane voltage field. Instead, Zn(2+) binds to superficial sites on the channel and modulates gating. The effects of extracellular Zn(2+) were strongly pH(o) dependent but were insensitive to pH(i), suggesting that protons and Zn(2+) compete for external sites on H(+) channels. The apparent potency of Zn(2+) in slowing activation was approximately 10x greater at pH(o) 7 than at pH(o) 6, and approximately 100x greater at pH(o) 6 than at pH(o) 5. The pH(o) dependence suggests that Zn(2+), not ZnOH(+), is the active species. Evidently, the Zn(2+) receptor is formed by multiple groups, protonation of any of which inhibits Zn(2+) binding. The external receptor bound H(+) and Zn(2+) with pK(a) 6.2-6.6 and pK(M) 6.5, as described by several models. Zn(2+) effects on the proton chord conductance-voltage (g(H)-V) relationship indicated higher affinities, pK(a) 7 and pK(M) 8. CdCl(2) had similar effects as ZnCl(2) and competed with H(+), but had lower affinity. Zn(2+) applied internally via the pipette solution or to inside-out patches had comparatively small effects, but at high concentrations reduced H(+) currents and slowed channel closing. Thus, external and internal zinc-binding sites are different. The external Zn(2+) receptor may be the same modulatory protonation site(s) at which pH(o) regulates H(+) channel gating.  相似文献   

12.
Aspects of a possible involvement of hydrogen ions in the electrophysiological responses to light of Limulus ventral photoreceptors were investigated. A 1 M solution of either a zwitter-ionic pH buffer or a weakly-buffering control substance was pressure injected through a micropipette into a ventral photoreceptor cell. To estimate the amount injected, 35SO4 was included in the solution. Membrane currents induced by light flashes were measured by a voltage-clamp technique. The buffer-filled micropipette passed current and a 3M KCl filled micropipette monitored membrane voltage. The sensitivity (peak light-induced current/stimulus energy) was measured, after dark adaptation, before and after injection. Injections of buffers, pH 6.3-7.2, to intracellular concentrations of at least 40-200 mM produced only a small mean decrease in sensitivity, approximately equal to that caused by injections of control substances. Excitation, therefore, apparently is not mediated by a change in intracellular pH. Buffers with pH values 5.4-8.4 were also injected. The time to peak of the response depended on pH, being shortened by up to 20% at pH values below 7.7 and lengthened at higher pH values. The time to peak of the response appeared to be shortened by an increase in intracellular pH-buffering capacity even when there was no change in intracellular pH.  相似文献   

13.
Aspects of a possible involvement of hydrogen ions in the electrophysiological responses to light of Limulus ventral photoreceptors were investigated. A I M solution of either a zwitter-ionic pH buffer or a weakly-buffering control substance was pressure injected through a micropipette into a ventral photoreceptor cell. To estimate the amount injected, 35SO4 was included in the solution. Membrane currents induced by light flashes were measured by a voltage-clamp technique. The buffer-filled micropipette passed current and a 3 M KCl filled micropipette monitored membrane voltage. The sensitivity (peak light-induced current/stimulus energy) was measured, after dark adaptation, before and after injection. Injections of buffers, pH 6.3–7.2, to intracellular concentrations of at least 40–200 mM produced only a small mean decrease in sensitivity, approximately equal to that caused by injections of control substances. Excitation, therefore, apparently is not mediated by a change in intracellular pH. Buffers with pH values 5.4–8.4 were also injected. The time to peak of the response depended on pH, being shortened by up to 20% at pH values below 7.7 and lengthened at higher pH values. The time to peak of the response appeared to be shortened by an increase in intracellular pH-buffering capacity even when there was no change in intracellular pH.  相似文献   

14.
A theoretical development in the evaluation of proton linkage in protein binding reactions by isothermal titration calorimetry (ITC) is presented. For a system in which binding is linked to protonation of an ionizable group on a protein, we show that by performing experiments as a function of pH in buffers with varying ionization enthalpy, one can determine the pK(a)'s of the group responsible for the proton linkage in the free and the liganded states, the protonation enthalpy for this group in these states, as well as the intrinsic energetics for ligand binding (delta H(o), delta S(o), and delta C(p)). Determination of intrinsic energetics in this fashion allows for comparison with energetics calculated empirically from structural information. It is shown that in addition to variation of the ligand binding constant with pH, the observed binding enthalpy and heat capacity change can undergo extreme deviations from their intrinsic values, depending upon pH and buffer conditions.  相似文献   

15.
Effects of ph and ionic strength on phosphatidylserine/phosphatidylcholine mixed membranes prepared on Millipore filter pore surfaces have been studied using spin-labeled phosphatidylcholine. Lowering pH at constant ionic strength and lowering ionic strength at constant pH caused a lateral reorganization of the membrane. The trigger was protonation of the serine carboxyl group which caused solidification of phosphatidylserine molecules in the membrane, leaving a fluid phase consisting mainly of phosphatidylcholine. The appearent pK for the proton-induced phase separation was measured in a wide range of salt concentrations. The ionic strength dependence was satisfactorily explained based on the electrostatic free energy of proton in the field of membrane surface potential. The Gouy-Chapman theory gave a good approximation for the surface potential. The surface pK of phosphatidylserine and phosphatidic acid vesicles was directly measured in various salt concentrations by 31P-NMR and the results confirmed validity of the Gouy-Chapman-type analysis. The lateral reorganization was triggered by electrostatic interaction but the bulk of the stabilization energy for the structural changes would be the gains in intermolecular van der Waals energy due to closer packing of phosphatidylserine on solidification.  相似文献   

16.
We studied the reaction between vinyl butyrate and 2-phenyl-1-propanol in acetonitrile catalyzed by Fusarium solani pisi cutinase immobilized on zeolites NaA and NaY and on Accurel PA-6. The choice of 2-phenyl-1-propanol was based on modeling studies that suggested moderate cutinase enantioselectivity towards this substrate. With all the supports, initial rates of transesterification were higher at a water activity (a(w)) of 0.2 than at a(w) = 0.7, and the reverse was true for initial rates of hydrolysis. By providing acid-base control in the medium through the use of solid-state buffers that control the parameter pH-pNa, which we monitored using an organo-soluble chromoionophoric indicator, we were able, in some cases, to completely eliminate dissolved butyric acid. However, none of the buffers used were able to improve the rates of transesterification relative to the blanks (no added buffer) when the enzyme was immobilized at an optimum pH of 8.5. When the enzyme was immobilized at pH 5 and exhibited only marginal activity, however, even a relatively acidic buffer with a pK(a) of 4.3 was able to restore catalytic activity to about 20% of that displayed for a pH of immobilization of 8.5, at otherwise identical conditions. As a(w) was increased from 0.2 to 0.7, rates of transesterification first increased slightly and then decreased. Rates of hydrolysis showed a steady increase in that a(w) range, and so did total initial reaction rates. The presence or absence of the buffers did not impact on the competition between transesterification and hydrolysis, regardless of whether the butyric acid formed remained as such in the reaction medium or was eliminated from the microenvironment of the enzyme through conversion into an insoluble salt. Cutinase enantioselectivity towards 2-phenyl-1-propanol was indeed low and was not affected by differences in immobilization support, enzyme protonation state, or a(w).  相似文献   

17.
Drosophila serves as a playground for examining the effects of genetic mutations on development, physiological function and behavior. Many physiological measures that address the effects of mutations require semi-intact or cultured preparations. To obtain consistent physiological recordings, cellular function needs to remain viable. Numerous physiological salines have been developed for fly preparations, with emphasis on nervous system viability. The commonly used saline drifts in pH and will cause an alteration in the heart rate. We identify a saline that maintains a stable pH and physiological function in the larval heart, skeletal neuromuscular junction, and ventral nerve cord preparations. Using these common assays, we screened various pH buffers of differing concentrations to identify optimum conditions. Buffers at 25 mM produce a stable heart rate with minimal variation in pH. Excitatory junction potentials evoked directly on larval muscles or through sensory-CNS-motor circuits were unaffected by at buffers at 25 mM. The salines examined did not impede the modulatory effect of serotonin on heart rate or neural activity. Together, our results indicate that the higher buffer concentrations needed to stabilize pH in HL3 hemolymph-like saline do not interfere with the acute function of neurons or cardiac myocytes.  相似文献   

18.
The pH dependence of the PHM-catalyzed monooxygenation of dansyl-YVG was studied in two different buffer systems in the pH range of 4-10. The pH-activity profile measured in a sulfonic acid buffer exhibited a maximum at pH 5.8 and became inactive at pH >9. The data could be fit to a model that assumed a protonated unreactive species A, a major reactive species B, and a less reactive species C. B formed in a deprotonation step with pK(a) of 4.6, while C formed and decayed with pK(a)s of 6.8 and 8.2, respectively. The pH dependence was found to be dominated by k(cat), with K(m)(dansyl-YVG) remaining pH-independent over the pH range of 5-8. Acetate-containing buffers shifted the pH maximum to 7.0, and the activity-pH profile could be simulated by formation and decay of a single active species with pK(a)s of 5.8 and 8.3, respectively. The pH-dependent changes in activity could be correlated with a change in the Debye-Waller factor for the Cu-S(met) (M314) component of the X-ray absorption spectrum which underwent a transition from a tightly bound inactive "met-on" form to a conformationally mobile active "met-off" form with a pK(a) which tracked the formation of the active species in both sulfonic acid and acetate-containing buffer systems. The data suggested that the conformational mobility of the bound substrate relative to the copper-superoxo active species is critical to catalysis and further suggested the presence of an accessible vibrational mode coupling Cu-S motion to the H tunneling probability along the Cu-O...H...C coordinate.  相似文献   

19.
The problem of predicting the kinetics of proton efflux and the decay of the internal proton concentration for vesicles containing one or more buffers for which the internal proton concentration is initially higher than that of the surrounding medium is examined. An analytical solution is derived that describes the time course of the proton efflux from vesicles and the decay of the internal proton concentration under conditions of zero transmembrane electric potential. The effect of the internal buffers is to increase the time required for the proton concentration gradient to equilibrate across the membrane. To simplify the analysis we assume that the equilibration of the internal and external proton activity is due primarily to proton diffusion through the membrane, and not to hydroxyl ion flux. For a vesicle containing a single buffer the solution requires six independent physical parameters: the initial internal proton concentration, the external proton concentration, the ratio of the vesicle surface area to the internal volume, the permeability coefficient of the membrane for protons, the total concentration of the internal buffer, and the equilibrium constant for the dissociation of the internal buffer. Determination of these physical values is sufficient to predict the time dependence of the internal proton concentration and of the proton efflux. Over a pH range that is below or near the pK of the internal buffer the solution is complex. However, if the initial pH is one unit or more higher than the pK of the internal buffer the kinetics of the internal proton concentration and proton efflux can be described by a pseudo first order reaction. In this case the apparent rate constant depends linearly on the permeability coefficient and is dominated by the total internal buffer concentration and its pK. For example, increasing the internal buffer concentration inside a vesicle by 10-fold results in an approximately 10-fold increase in the half-time of the proton efflux kinetics. The theoretical analysis is applied to thylakiod vesicles using experimentally determined values for the physical parameters. The predictions of the analysis are compared to experimentally observed kinetics.  相似文献   

20.
Alveolar epithelial cells isolated from rats and maintained in primary culture were studied using the whole-cell configuration of the "patch-clamp" technique. After other ionic conductances were eliminated by replacing permeant ions with N-methyl-D-glucamine methanesulfonate, large voltage-activated hydrogen-selective currents were observed. Like H+ currents in snail neurons and axolotl oocytes, those in alveolar epithelium are activated by depolarization, deactivate upon repolarization, and are inhibited by Cd2+ and Zn2+. Activation of H+ currents is slower in alveolar epithelium than in other tissues, and often has a sigmoid time course. Activation occurs at more positive potentials when external pH is decreased. Saturation of the currents suggests that diffusion limitation may occur; increasing the pipette buffer concentration from 5 to 120 mM at a constant pH of 5.5 increased the maximum current density from 8.7 to 27.3 pA/pF, indicating that the current amplitude can be limited in 5 mM buffer solutions by the rate at which buffer molecules can supply H+ to the membrane. These data indicate that voltage-dependent H+ currents exist in mammalian cells.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号