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 62 毫秒
1.
The regulation of intracellular neuronal pH and pH from the extracellular space was studied in the isolated stomatogastric ganglion of the crab Cancer pagurus. Intracellular neuronal pH was found to be 0.3–0.4 pH units more acidic than the standard bath pH of 7.4 and surprisingly, the extracellular space pH was found to be around 0.1 pH units more alkaline than the bath pH. Extracellular space pH shifts in response to bath pH changes decreased as a function of the depth of the recording site within the ganglion, suggesting the existence of restrictions in the free diffusion of H+. The amplitude of these pHe shifts increased in Na+-free saline or with amiloride, suggesting Na+-dependent regulation of the extracellular space pH. In Na+ free saline or in the presence of amiloride, intracellular pH recovery from an NH4Cl induced acidosis was reduced, and the H+ muffling capacity (cf. Thomas et al. 1991) of the extracellular space was markedly reduced. Changes of bath pH had only small effects on the rhythm generating properties of one of the central pattern generators of the stomatogastric ganglion, while NH4Cl-induced intraganglionic pH changes markedly altered this rhythm.Abbreviations CPG central pattern generator - ECS extracellular space - LP lateral pyloric neuron - NMDG N-methyl-D-glucamine - PD pyloric dilator neuron - pHe extracellular space pH - pHi intracellular pH - pHo bath pH - STG stomatogastric ganglion - Vref reference potential  相似文献   

2.
Cell survival is conditional on the maintenance of a favourable acid–base balance (pH). Owing to intensive respiratory CO2 and lactic acid production, cancer cells are exposed continuously to large acid–base fluxes, which would disturb pH if uncorrected. The large cellular reservoir of H+-binding sites can buffer pH changes but, on its own, is inadequate to regulate intracellular pH. To stabilize intracellular pH at a favourable level, cells control trans-membrane traffic of H+-ions (or their chemical equivalents, e.g. ) using specialized transporter proteins sensitive to pH. In poorly perfused tumours, additional diffusion-reaction mechanisms, involving carbonic anhydrase (CA) enzymes, fine-tune control extracellular pH. The ability of H+-ions to change the ionization state of proteins underlies the exquisite pH sensitivity of cellular behaviour, including key processes in cancer formation and metastasis (proliferation, cell cycle, transformation, migration). Elevated metabolism, weakened cell-to-capillary diffusive coupling, and adaptations involving H+/H+-equivalent transporters and extracellular-facing CAs give cancer cells the means to manipulate micro-environmental acidity, a cancer hallmark. Through genetic instability, the cellular apparatus for regulating and sensing pH is able to adapt to extracellular acidity, driving disease progression. The therapeutic potential of disturbing this sequence by targeting H+/H+-equivalent transporters, buffering or CAs is being investigated, using monoclonal antibodies and small-molecule inhibitors.  相似文献   

3.
Neuronal activity results in release of K+ into the extracellular space of the central nervous system. If the excess K+ is allowed to accumulate, neuronal firing will be compromised by the ensuing neuronal membrane depolarization. The surrounding glial cells are involved in clearing K+ from the extracellular space by molecular mechanism(s), the identity of which have been a matter of controversy for over half a century. Kir4.1-mediated spatial buffering of K+ has been promoted as a major contributor to K+ removal although its quantitative and temporal contribution has remained undefined. We discuss the biophysical and experimental challenges regarding determination of the contribution of Kir4.1 to extracellular K+ management during neuronal activity. It is concluded that 1) the geometry of the experimental preparation is crucial for detection of Kir4.1-mediated spatial buffering and 2) Kir4.1 enacts spatial buffering of K+ during but not after neuronal activity.  相似文献   

4.
Neuronal activity results in release of K+ into the extracellular space of the central nervous system. If the excess K+ is allowed to accumulate, neuronal firing will be compromised by the ensuing neuronal membrane depolarization. The surrounding glial cells are involved in clearing K+ from the extracellular space by molecular mechanism(s), the identity of which have been a matter of controversy for over half a century. Kir4.1-mediated spatial buffering of K+ has been promoted as a major contributor to K+ removal although its quantitative and temporal contribution has remained undefined. We discuss the biophysical and experimental challenges regarding determination of the contribution of Kir4.1 to extracellular K+ management during neuronal activity. It is concluded that 1) the geometry of the experimental preparation is crucial for detection of Kir4.1-mediated spatial buffering and 2) Kir4.1 enacts spatial buffering of K+ during but not after neuronal activity.  相似文献   

5.
6.
The high pH-maintaining capacity of yeast suspension after glucose-induced acidification, measured as its ability to neutralize added alkali, was found to be due mainly to actively extruded acidity (H+). The buffering action of passively excreted metabolites (CO2, organic acids) and cell surface polyelectrolytes contributed only 15–40% to the overall pH-maintaining capacity which was 10 mmol NaOH/l per pH unit between pH 3 and 4 and 3.5 mmol NaOH/l per pH unit between pH 4 and 7. The buffering capacity of yeast cell-free extract was still higher (up to 4.5-times) than that of glucose-supplied cell suspension; addition of glucose to the extract thus produced considerable titratable acidity but negligible net acidity. The glucose-induced acidification of yeast suspension was stimulated by univalent cations in the sequence K+ >Rb+ >>Li+ ~- Cs+ ~- Na+. The processes participating in the acidification and probably also in the creation of extracellular buffering capacity include excretion of CO2 and organic acids, net extrusion of H+ and K+ (in K+-free media; in K+-containing media this is preceded by an initial rapid K+ uptake), and movements of some anions (phosphate, chlorides). The overall process appears to be electrically silent.  相似文献   

7.
The hypothesis tested was that embryonic metabolism affects the water chemistry in the boundary layer. In addition, embryo crowding would further compound the metabolic effect on the water chemistry in the boundary layer. As development progressed, the magnitude of the boundary layer gradients for O2 and pH, but not for NH, increased. The presence of the egg capsule hindered the diffusion of O2 into and H+ and NH out of the embryo. The magnitude of the O2, pH and NH boundary layer gradient was significantly increased when embryos were surrounded by either sham embryos or live embryos. The majority of this crowding effect on embryo boundary layers was due to changes in water flow rather than due to metabolism directly. These results clearly show that the microenvironment adjacent to the developing rainbow trout Oncorhynchus mykiss embryo becomes more stagnant as development progresses in the presence of the egg capsule and is further intensified with embryo crowding.  相似文献   

8.
Growth yield of the halotolerant bacterium A505 was increased by the supplement of Na+, K+, or Rb+ into the culture media with pH 7.5, and inhibited by Li+ or Cs+. In the presence of less than 0.1 M NaCl or KCl alkaline growth media, pH 9.2 to 9.7, afforded optimal growth of this strain. Intracellular ion content of this microbe changed reflecting on the Na+ or K+ concentration in the media, although it tended to accumulate K+ and extrude Na+ in the media without NaCl supplemented. A 1.2 to 1.4-fold stimulation of in vitro NADH oxidase activity was obtained by supplement of salts, except for LiCl. The rate of NADH oxidation in the absence of salts correlated with the pH and showed clear maxima at pH about 8, irrespective of growth conditions. In the presence of 0.5 M NaCl or KCl, on the other hand, pH dependence was less significant and showed only a flat maximum at pH around 7. Effects of anions on NADH oxidase were realized following the lyotropic series: SO 4 2- >F->CH3COO->Cl->I->SCN-, aside from NO 3 - , which exhibited the largest stimulation on enzyme activity in all the anions examined.Abbreviations HEPES 4-(2-hydroxyethyl)-1-piperazine-ethanesulfonic acid - HQNO 2-heptyl-4-hydroxyquinoline-N-oxide - MES 4-morpholineethanesulfonic acid - Tris tris(hydroxy-methyl)methylamine  相似文献   

9.
Zinc activates a specific Zn2+-sensing receptor, ZnR/GPR39, and thereby triggers cellular signaling leading to epithelial cell proliferation and survival. Epithelial cells that express ZnR, particularly colonocytes, face frequent changes in extracellular pH that are of physiological and pathological implication. Here we show that the ZnR/GPR39-dependent Ca2+ responses in HT29 colonocytes were maximal at pH 7.4 but were reduced by about 50% at pH 7.7 and by about 62% at pH 7.1 and were completely abolished at pH 6.5. Intracellular acidification did not attenuate ZnR/GPR39 activity, indicating that the pH sensor of this protein is located on an extracellular domain. ZnR/GPR39-dependent activation of extracellular-regulated kinase (ERK)1/2 or AKT pathways was abolished at acidic extracellular pH of 6.5. A similar inhibitory effect was monitored for the ZnR/GPR39-dependent up-regulation of Na+/H+ exchange activity at pH 6.5. Focusing on residues putatively facing the extracellular domain, we sought to identify the pH sensor of ZnR/GPR39. Replacing the histidine residues forming the Zn2+ binding site, His17 or His19, or other extracellular-facing histidines to alanine residues did not abolish the pH dependence of ZnR/GPR39. In contrast, replacing Asp313 with alanine resulted in similar Ca2+ responses triggered by ZnR/GPR39 at pH 7.4 or 6.5. This mutant also showed similar activation of ERK1/2 and AKT pathways, and ZnR-dependent up-regulation of Na+/H+ exchange at pH 7.4 and pH 6.5. Substitution of Asp313 to His or Glu residues restored pH sensitivity of the receptor. This indicates that Asp313, which was shown to modulate Zn2+ binding, is an essential residue of the pH sensor of GPR39. In conclusion, ZnR/GPR39 is tuned to sense physiologically relevant changes in extracellular pH that thus regulate ZnR-dependent signaling and ion transport activity.  相似文献   

10.
The effect of HCO 3 - on ion absorption by young corn roots was studied in conditions allowing the independent control of both the pH of uptake solution and the CO2 partial pressure in air bubbled through the solution. The surface pH shift in the vicinity of the outer surface of the plasmalemma induced by active H+ excretion was estimated using the initial uptake rate of acetic acid as a pH probe (Sentenac and Grignon (1987) Plant Physiol. 84, 1367). Acetic acid and orthophosphate uptake rates and NO 3 - accumulation were slowed down, while 86Rb+ uptake and K+ accumulation rates were increased by HCO 3 - . These effects were similar to those induced by 4-(2-hydroxyethyl)-1-piperazineethane sulfonic acid/2-amino-2-(hydroxymethyl)-1,3-propanediol (Hepes-Tris). They were more pronounced when the H+ excretion was strong, were rapidly reversible and were not additive to those of Hepes-Tris. The hypothesis is advanced that the buffering system CO2/H2CO3/HCO 3 - accelerated the diffusion of equivalent H+ inside the cell wall towards the medium. This attenuated the surface pH shift in the vicinity the plasma membrane and affected the coupling between the proton pump and cotransport systems.Abbreviations FW fresh weight - Hepes 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid - Jaa acetic acid influx - JK + K+ influx - JPi orthophosphate influx - Mes 2-(N-morpholino)ethanesulfonic acid - pCO2 CO2 partial pressure - Tris 2-amino-2-(hydroxymethyl)-1,3-propanediol  相似文献   

11.
A vibrating probe was used to measure extracellular electrical currents around developing somatic embryos in two lines (RCC27, RCC48) of cultured cells of Daucus carota L. at the heart and torpedo stages. At pH 5.5, an inward current of 1.2±0.1 A·cm-2 (n=23) was detected at the cotyledon, and an outward current of 1.0±0.1 A·cm-2 (n=22) was found at the radicle in torpedostage embryos from the RCC27 line. At a pH of 5.75 the currents increased by 0.2–0.3 A·cm-2 (n=60–62). In a few cases an additional small inward current was detected at the tip of the radicle in toepedo-stage embryos from RCC27 line. Such an inward current at the radicle seemed to appear earlier, some time after the heart stage, in embryos from the RCC48 line.Both extracellular pH measurements (using microelectrodes filled with ion-sensitive resin) and ion-substitution studies were carried out in order to ascertain the ionic composition of the currents in torpedo-stage embryos from the RCC27 line. Regions adjacent to the cotyledon and radicle, at the points of current entry and exit, were found to be more acidic by 0.02±0.01 (n=14) and 0.07±0.01 (n=12) pH units, respectively, than the bulk medium. Removal of K+ from the medium reversibly reduced the currents to about 25% of their original value at both cotyledon and radicle. Deletion of Cl- decreased the currents slightly. Removal of Ca2+ resulted in a rapid doubling of currents. Addition of either N,N-dicyclohexylcarbodiimide or tetraethyl ammonium chloride substantially reduced overall currents, and their removal resulted in partial recovery of the currents. It is suggested that the inward current at the cotyledon is comprised largely of K+ influx and the outward current at the radicle is mainly the result of active H+ efflux.Abbreviations 2,4-D 2,4-dichlorophenoxyacetic acid - MS Murashige and Skoog  相似文献   

12.
We have studied the mechanism of the response to iron deficiency in rape (Brassica napus L.), taking into account our previous results: net H+ extrusion maintains a pH shift between the root apoplast and the solution, and the magnitude of the pH shift decreases as the buffering power in the solution increases. The ferric stress increased the ability of roots to reduce Fe[III]EDTA. Buffering the bulk solution (without change in pH) inhibited Fe[III]EDTA reduction. At constant bulk pH, the inhibition (ratio of the Fe[III]EDTA-reduction rates measured in the presence and in the absence of buffer) increased with the rate of H+ extrusion (modulated by the length of a pretreatment in 0.2 mM CaSO4). These results support the hypothesis that the apoplastic pH shift caused by H+ excretion stimulated Fe[III] reduction. The shape of the curves describing the pH-dependency of Fe[III]EDTA reduction in the presence and in the absence of a buffer fitted this hypothesis. When compared to the titration curves of Fe[III]citrate and of Fe[III]EDTA, the curves describing the dependency of the reduction rate of these chelates on pH indicated that the stimulation of Fe[III] reduction by the apoplastic pH shift due to H+ excretion could result from changes in electrostatic interactions between the chelates and the fixed chargers of the cell wall and-or plasmalemma. Blocking H+ excretion by vanadate resulted in complete inhibiton of Fe[III] reduction, even in an acidic medium in which there was neither a pH shift nor an inhibitory effect of a buffer. This indicates that the apoplastic pH shift resulting from H+ pumping is not the only mechanism which is involved in the coupling of Fe[III] reduction to H+ transport. Our results shed light on the way by which the strong buffering effect of HCO 3 - in some soils may be involved in iron deficiency encountered by some of the plants which grow in them.  相似文献   

13.
Summary Suspensions of LLC-PK1 cells (a continuous epitheliod cell line with renal characteristics) are examined for mechanisms of intracellular pH regulation using the fluorescent probe BCECF. Initial experiments determine suitable calibration procedures for use of the BCECF fluorescent signal. They also determine that the cell suspension contains cells which (after 4 hr in suspension) have Na+ and K+ gradients comparable to those of cells in monolayer culture. The steady-state intracellular pH (7.05±0.01,n=5) of cells which have recovered in (pH 7.4) Na+-containing medium is not affected over several minutes by addition of 100 M amiloride or removal of extracellular Na+ (Na o + /H i + and Na i + /H o + exchange reactions are functionally inactive (compared to cellular buffering capacity). In contrast, Na o + /H i + exchange is activated by an increased cellular acid load. This activation may be observed directly either as a stimulation of net H+ efflux or net Na+ influx with decreasing intracellular pH. The extrapolation of this latter data suggests a set point of Na+/H+ exchange of approximately pH 7.0, consistent with the observed resting intracellular pH of approximately 7.05.  相似文献   

14.
A stopped flow rapid reaction apparatus capable of following changes of ±0.02 pH unit in 0.1 ml of solution in less than 0.005 sec has been developed, utilizing a commercially available pH-sensitive glass electrode. Using this instrument, extracellular pH at 37°C was followed from less than 0.025 sec to 300 sec after mixing equal volumes of the following CO2-free solutions: (A) normal human red cells, washed three times and resuspended in 150 mM NaCl at pH 7.2 with a hematocrit of 18%; and, (B) 150 mM NaCl adjusted with HCl or NaOH to pH 2.1 to pH 10.3. A minimum of 2 ml of mixture had to flow through the electrode chamber to ensure complete washout. The mixing process produced a step change in the pH of the extracellular fluid, after which exchanges across the red cell membrane and buffering by intracellular hemoglobin caused it to return toward pH 7.2 with an approximately exponential time course. Under the assumption that pH changes after mixing represent exchanges of hydroxyl for chloride ions across the cell membrane, hydroxyl ion permeabilities (P OH - in cm/sec) were calculated and found to vary from 2 x 10-4 at pH 9 to 4 x 10-1 at pH 4 according to the empirical relationship P OH - = 170 exp (-1.51 pH). The form of the dependence of P OH - on extracellular pH does not appear compatible with a simple fixed charge theory of membrane permselectivity.  相似文献   

15.
Summary The intracellular pH (pH i ) of tissue-cultured bovine lens epithelial cells was measured in small groups of 6 to 10 cells using the trapped fluorescent dye 2,7-bis-(2-,carboxyethyl)-5 (and 6)carboxyfluorescein (BCECF). When perifused at 35°C with artificial aqueous humour solution (AAH) containing 16 mM HCO 3 - and 5% CO2, pH 7.25, pH i was 7.19±0.02 (sem, n = 95). On removing HCO 3 - and CO2 there was an initial transient alkalinization followed by a fall in pH to a steady value of 6.97±0.03 (sem, n = 54). Addition of 0.25 mM 4,4-diisothiocyanatostilbene2, 2-disulfonic acid (DIDS) to AAH containing HCO 3 - and CO2 led to a rapid and pronounced fall in pH. Exposure to Na+-free AAH again led to a marked fall in pH i , but in this case the addition of DIDS did not produce a further fall. Substitution of the impermeant anion gluconate for Cl in the presence of HCO 3 - led to a rise in pH i , while substitution in the absence of HCO 3 - led to a fall in pH i . The above data indicate a significant role for a sodium-dependent Cl-HCO 3 - exchange mechanism in the regulation of pH i . Addition of 1 mM amiloride to control AAH in both the presence and absence of HCO 3 - led to a marked fall in pH i , indicating that a Na+/H+ exchange mechanism also has a significant role in the regulation of pH i . There is evidence for a lactic acid transport mechanism in bovine lens cells, as addition of lactate to the external medium produced a rapid fall in pH i . Larger changes in pH i were observed in control compared to HCO 3 - -free AAH and in the latter case a pronounced alkalinizing overshoot was obtained on removing external lactate. Tissue-cultured bovine lens cells thus possess at least three membrane transport mechanisms that are involved in pH regulation. The buffering capacity of the lens cells was measured by perturbing pH i with either NH 4 + or procaine. The values obtained were similar in both cases and the intrinsic buffering capacity measured in the absence of external HCO 3 - was 5 mm/pH unit (procaine). However, in the presence of HCO 3 - and CO2 the buffer capacity increases approximately fourfold, indicating that HCO 3 - is the principal intracellular buffer.We acknowledge financial support from the Wellcome Trust and the Humane Research Trust for this project. M.R. Williams was in receipt of a Science & Engineering Research Council studentship.  相似文献   

16.
The ectoderm of the one-day chick embryo generates dorsoventrally oriented short-circuit current (I sc) entirely dependent on extracellular sodium.At the dorsal cell membrane, the I sc was modified reversibly and in a concentration-dependent manner by: amiloride (60% decrease at 1 mm, with 2 apparent IC50s: 0.13 and 48 m), phlorizin (0.1 mm) or removal of glucose (30% decrease, additive to that of amiloride), SITS (1 mm, 13% decrease). Acidification or alkalinization of the dorsal (but not ventral) superfusate produced, respectively, decrease or increase of I sc with a pH50 of 7.64.Ba2+ (0.1–1 mm) from either side of the ectoderm decreased the I sc by 30%. Anthracene-9-carboxylic acid, furosemide and inducers of cAMP had no effect on electrophysiological properties of the blastoderm.The chick ectoderm is therefore a highly polarized epithelium containing, at the dorsal membrane, the high and low affinity amiloride-sensitive Na+ channels, Na+-glucose cotransporter, K+ channels and pH sensitivity, and, at the ventral membrane, the Na+, K+-ATPase and K+ channels. The Na+ transport reacts to pH, but lacks the cAMP regulatory system, well known in many epithelia.The active Na+ transport drives glucose and fluid into the intraembryonic space, across and around the blastoderm which, in the absence of blood circulation, could secure renewal of extracellular fluid and disposal of wastes and thus maintain the cell homeostasis.This work was supported by the Swiss National Research Foundation (grant 3.418-0.86 to P.K.), by the Roche Research Foundation (grant to U.K.), the Fond du 450ème anniversaire de l'Université de Lausanne and the Société Académique Vaudoise (grants to H.A.). We thank C. Bareyre, G. de Torrenté and R. Ksontini for excellent technical assistance and Drs. E. Raddatz, Y. de Ribaupierre and B. Prod'hom for helpful discussions.  相似文献   

17.
The initial rate of (2-14C)acetic acid (AA) uptake by corn roots was used for probing the dependency of the root cell surface pH on H+ excretion. AA influx was linearly related to AA concentration, dependent on the concentration of the undissociated form (AH), unaffected by variations of the membrane potential, and was thus assumed to result mainly from the free diffusion of AH across the membrane. Various treatments (vanadate, dicyclohexylcarbodiimide, hypoxia, nitrate, root ageing, fusicoccin) were used to vary H+ flux while the medium pH was maintained constant. There was a positive relation between AA influx and the net H+ efflux. This relation disappeared when the proton buffering strength of the absorption medium was increased. These results indicate that the pH at the membrane surface was lowered by H+ excretion, even in situations where the bulk (pH) was unaffected. The depressive effect of vanadate on AA influx was counteracted by acidifying the medium in order to estimate this pH shift: −1.2 pH units in 12.5 millimolar K2SO4 (pH 6.8). Substituting AA by butyric acid showed that this estimation was not dependent of the probe used.  相似文献   

18.
Summary The loop diuretic bumetanide binds specifically to the Na/K/2Cl cotransporter of many cell types including duck erythrocytes. Membranes isolated from these erythrocytes retain the ability to bind bumetanide when cells are exposed to cotransport activity stimuli prior to membrane isolation. An extensive study of the effects of ions on specific [3H]bumetanide binding to such membranes is presented here and compared to the activity of these ions in supporting transport function in intact cells. Both Na+ and K+ enhanced bumetanide binding in a saturable manner consistent with a single-site interaction. The K m for each ion was dependent on the concentration of the other cation suggesting heterotropic cooperative interactions between the Na+ and K+ binding sites. Na+ and K+ were partially replaceable, with the selectivity of the Na+ site being Na+ > Li+ > NH 4 + ; N-methyl-d-glucamine+, choline+ and tetramethylammonium+ also supported a small amount of specific binding when substituted for Na+. The selectivity of the K+ site was K+ Rb+ > NH 4 + > Cs+; N-methyl-d-glucamine+, choline+ and tetramethylammonium+ were inactive at this site. The results of transport experiments revealed a slightly different pattern. Li+ could partially substitute for Na+ in supporting coteansport, but other monovalent cations were completely inactive. The order of potency at the K+ site was NH 4 + > K+ Rb+ > Cs+ other monovalent cations. The effect of Cl- on bumetanide binding was biphasic, being stimulatory at low [Cl-] but inhibitory at high [Cl-]. As this implies the existence of two Cl- binding sites (termed Cl H and Cl L for the high- and low- affinity sites, respectively) each phase was examined individually. Cl- binding to Cl H could be described by a rectangular hyperbola with a K m of 2.5 mm, while kinetic analysis of the inhibition of bumetanide binding at high [Cl-] revealed that it was of a noncompetitive type (K i = 112.9 mm). The selectivity of anion binding to the two sites was distinct. Cl H was highly selective with Cl- > SCN- > Br-; F-, NO 3 - , ClO 4 - , MeSO 4 - , gluconate- and SO 4 2- were inactive. The efficacy of anion inhibition of binding to Cl L was ClO 4 - > I- > SCN- > NO3 > Cl-; F-, MeSO 4 - , gluconate-, and SO 4 2- were inactive. Thus, Cl H is much more selective than Cl L and largely accounts for the specificity of the system with respect to anion transport. SO 4 - , NO 3 - , I-, SCN- and ClO 4 - did not support cotransport when bound to Cl L and the latter three anions were inhibitory. Mg2+ was found to stimulate binding at a narrowly defined peak around 1.5 mm, but was inhibitory at higher concentrations. Other divalent cations caused a similar inhibition of bumetanide binding but did not exert a stimulatory effect at 1.5 mm. Divalent cations have little effect on cotransport in intact cells at concentrations up to 20 mm, suggesting that their effects on diuretic binding reflect interactions at internally disposed sites. Bumetanide binding was optimal at a pH of 7.8–8.1 and declined sharply as the pH was lowered towards 6. The titration curve correlated well with the effect of pH on cotransport in intact cells; the inhibitory effect of low pH suggests that protonation of the cotransporter may inhibit its function.We thank Drs. Brad Pewitt, John Westley and Mrinalini Rao for discussion, Sara Leung and Artelia Watson for their excellent technical assistance, and Dr. R.J. Turner for his gift of [3H] bumetanide. This work was supported in part by Cystic Fibrosis Center grant #CF RO11 7-04.  相似文献   

19.
Summary The novel application of a two-substrate model (Florini and Vestling 1957) from enzymology to transport kinetics at the gills of freshwater trout indicated that Na+/acidic equivalent and Cl-/basic equivalent flux rates are normally limited by the availability of the internal acidic and basic counterions, as well as by external Na+ and Cl- levels. Adult rainbow trout fitted with dorsal aortic and bladder catheters were chronically infused (10–16 h) with isosmotic HCl to induce a persistent metabolic acidosis. Acid-base neutral infusions of isosmotic NaCl and non-infused controls were also performed. Results were compared to previous data on metabolic alkalosis in trout induced by either isosmotic NaHCO3 infusion or recovery from environmental hyperoxia (Goss and Wood 1990a, b). Metabolic acidosis resulted in a marked stimulation of Na+ influx, no change in Cl- influx, positive Na+ balance, negative Cl- balance, and net H+ excretion at the gills. Metabolic alkalosis caused a marked inhibition of Na+ influx and stimulation of Cl- influx, negative Na+ balance, positive Cl- balance, and net H+ uptake (=base excretion). Mean gill intracellular pH qualitatively followed extracellular pH. Classical one-substrate Michaelis-Menten analysis of kinetic data indicated that changes in Na+ and Cl- transport during acid-base disturbance are achieved by large increases and decreases in Jmax, and by increases in Km. However, one-substrate analysis considers only external substrate concentration and cannot account for transport limitations by the internal substrate. The kinetic data were fitted successfully to a two-substrate model, using extracellular acid-base data as a measure of internal HCO 3 - and H+ availability. This analysis indicated that true Jmax values for Na+/acidic equivalent and Cl-/basic equivalent transport are 4–5 times higher than apparent Jmax values by one-substrate analysis. Flux rates are limited by the availability of the internal counterions; transport Km values for HCO 3 - and H+ are far above their normal internal concentrations. Therefore, small changes in acid-base status will have large effects on transport rates, and on apparent Jmax values, without alterations in the number of transport sites. This system provides an automatic, negative feedback control for clearance or retention of acidic/basic equivalents when acid-base status is changing.Abbreviations Amm total ammonia in water - DMO 55-dimethyl-24-oxyzolidine-dione - Jin unidirectional inward ion movement across the gill - Jout unidirectional outward ion movement across the gill - Jnet net transfer of ions (sum of Jin and Jout) across the gill - Jmax maximal transport rate for ion - Km inverse of affinity of transporter for ion - PIO2 partial pressure of oxygen in inspired water - PaCO2 partial pressure of carbon dixide in arterial blood - TAlk titratable alkalinity of the water - PEG polyethylene glycol - NEN New England Nuclear  相似文献   

20.
A pH-sensitive glass electrode was used in a temperature-controlled stopped-flow rapid reaction apparatus to determine rates of pH equilibration in red cell suspensions. The apparatus requires less than 2 ml of reactants. The electrode is insensitive to pressure and flow variations, and has a response time of < 5 ms. A 20% suspension of washed fresh human erythrocytes in saline at pH 7.7 containing NaHCO3 and extracellular carbonic anhydrase is mixed with an equal volume of 30 mM phosphate buffer at pH 6.7. Within a few milliseconds after mixing, extracellular HCO3- reacts with H+ to form CO2, which enters the red cells and rehydrates to form HCO3-, producing an electrochemical potential gradient for HCO3- from inside to outside the cells. HCO3- then leaves the cells in exchange for Cl-, and extracellular pH increases as the HCO3- flowing out of the cells reacts with H+. Flux of HCO3- is calculated from the dpH/dt during HCO3--Cl- exchange, and a velocity constant is computed from the flux and the calculated intracellular and extracellular [HCO3-]. The activation energy for the exchange process is 18.6 kcal/mol between 5°C and 17°C (transition temperature), and 11.4 kcal/mol from 17°C to 40°C. The activation energies and transition temperature are not significantly altered in the presence of a potent anion exchange inhibitor (SITS), although the fluxes are markedly decreased. These findings suggest that the rate-limiting step in red cell anion exchange changes at 17°C, either because of an alteration in the nature of the transport site or because of a transition in the physical state of membrane lipids affecting protein-lipid interactions.  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