首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Proline porter II is rapidly activated when nongrowing bacteria are subjected to a hyperosmotic shift (Grothe, S., Krogsrud, R. L., McClellan, D. J., Milner, J. L., and Wood, J. M. (1986) J. Bacteriol. 166, 253-259). Proline porter II was active in membrane vesicles prepared from bacteria grown under optimal conditions, nutritional stress, or osmotic stress. That activity was: (i) dependent on the presence of the energy sources phenazine methosulphate plus ascorbate or D-lactate; (ii) observed only when a hyperosmotic shift accompanied the transport measurement; (iii) inhibited by glycine betaine in a manner analogous to that observed in whole cells; and (iv) eliminated by lesions in proP. Membrane vesicles were able to transport serine but not glutamine and serine transport was reduced by the hyperosmotic shift. In whole cells, proline porter II activity was supported by glucose and by D-lactate in a strain defective for proline porters I and III and the F1F0-ATPase. Glucose energized proline uptake was eliminated by carbonyl cyanide m-chlorophenylhydrazone and KCN as was serine uptake. These results suggested that proline porter II was respiration-dependent and probably ion-linked. Activation of proline porter II in whole cells by sucrose or NaCl was sustained over 30 min, whereas activation by glycerol was transient. Proline porter II was activated by NaCl and sucrose with a half-time of approximately 1 min in both whole cells and membrane vesicles. Thus, activation of proline porter II was reversible. It occurred at a rate comparable to that of K+ influx and much more rapid than the genetic regulatory responses that follow a hyperosmotic shift.  相似文献   

2.
The bactericidal effect of hydrostatic pressure is reduced when bacteria are suspended in media with high osmolarity. To elucidate mechanisms responsible for the baroprotective effect of ionic and nonionic solutes, Lactococcus lactis was treated with pressures ranging from 200 to 600 MPa in a low-osmolarity buffer or with buffer containing 0.5 M sucrose or 4 M NaCl. Pressure-treated cells were characterized in order to determine viability, the transmembrane difference in pH (ΔpH), and multiple-drug-resistance (MDR) transport activity. Furthermore, pressure effects on the intracellular pH and the fluidity of the membrane were determined during pressure treatment. In the presence of external sucrose and NaCl, high intracellular levels of sucrose and lactose, respectively, were accumulated by L. lactis; 4 M NaCl and, to a lesser extent, 0.5 M sucrose provided protection against pressure-induced cell death. The transmembrane ΔpH was reversibly dissipated during pressure treatment in any buffer system. Sucrose but not NaCl prevented the irreversible inactivation of enzymes involved in pH homeostasis and MDR transport activity. In the presence 0.5 M sucrose or 4 M NaCl, the fluidity of the cytoplasmic membrane was maintained even at low temperatures and high pressure. These results indicate that disaccharides protect microorganisms against pressure-induced inactivation of vital cellular components. The protective effect of ionic solutes relies on the intracellular accumulation of compatible solutes as a response to the osmotic stress. Thus, ionic solutes provide only asymmetric protection, and baroprotection with ionic solutes requires higher concentrations of the osmolytes than of disaccharides.  相似文献   

3.
The apparent extracellular space in incubated slices of rat renal cortex, medulla and papilla has been measured using three differently sized marker molecules, mannitol, sucrose and inulin. Cellular volumes have been estimated by following the efflux of 3-O-methyl-D-glucose from equilibrated slices. Sucrose appears to be the most accurate extracellular marker in each of the regions examined, in that the sum of its volume of distribution plus cellular volume approximates most closely to the total slice fluid volume. Inulin has the same volume of distribution as sucrose in cortical slices, but under-penetrates medullary and papillary tissue. Mannitol overestimates the extracellular space in all three regions, although its larger volume of distribution, relative to that of sucrose, was not statistically significant in papillary slices. When cell volume and composition are estimated (a) using sucrose as extracellular marker and (b) making appropriate allowance for the presence of bound tissue electrolytes, it is found that cells in each region have low Na+ and high K+ concentrations and contents. When papillary slices are incubated in medium of very high osmolality (NaCl plus urea, 2000 mosmol/kg H2O) there is a moderate (approx. 23%) decrease in cell volume and an increase in cell fluid Na+ and Cl- concentrations equal to approx. 50% of the increase in the extracellular concentrations. Cell K+ concentrations remain unchanged. The results show that cells in renal slices are able to maintain high K+-to-Na+ ratios when incubated in isosmotic (cortex) or moderately hyperosmotic media (medulla and papilla), and suggest that regulation of papillary cell volume following hyperosmotic shock can only partly be ascribed to uptake of extracellular electrolytes.  相似文献   

4.
The bactericidal effect of hydrostatic pressure is reduced when bacteria are suspended in media with high osmolarity. To elucidate mechanisms responsible for the baroprotective effect of ionic and nonionic solutes, Lactococcus lactis was treated with pressures ranging from 200 to 600 MPa in a low-osmolarity buffer or with buffer containing 0.5 M sucrose or 4 M NaCl. Pressure-treated cells were characterized in order to determine viability, the transmembrane difference in pH (DeltapH), and multiple-drug-resistance (MDR) transport activity. Furthermore, pressure effects on the intracellular pH and the fluidity of the membrane were determined during pressure treatment. In the presence of external sucrose and NaCl, high intracellular levels of sucrose and lactose, respectively, were accumulated by L. lactis; 4 M NaCl and, to a lesser extent, 0.5 M sucrose provided protection against pressure-induced cell death. The transmembrane DeltapH was reversibly dissipated during pressure treatment in any buffer system. Sucrose but not NaCl prevented the irreversible inactivation of enzymes involved in pH homeostasis and MDR transport activity. In the presence 0.5 M sucrose or 4 M NaCl, the fluidity of the cytoplasmic membrane was maintained even at low temperatures and high pressure. These results indicate that disaccharides protect microorganisms against pressure-induced inactivation of vital cellular components. The protective effect of ionic solutes relies on the intracellular accumulation of compatible solutes as a response to the osmotic stress. Thus, ionic solutes provide only asymmetric protection, and baroprotection with ionic solutes requires higher concentrations of the osmolytes than of disaccharides.  相似文献   

5.
Hepatocyte transmembrane potential (Vm) behaves as an osmometer and varies with changes in extracellular osmotic pressure created by altering the NaCl concentration in the external medium (Howard, L.D. and Wondergem, R. (1987) J. Membr. Biol. 100, 53). We now have demonstrated similar effects on Vm by increasing external osmolality with added sucrose and not altering ionic strength. We also have demonstrated that hyperosmotic stress-induced depolarization of Vm results from changes in membrane K+ conductance, gK, rather than from changes in the K+ equilibrium potential. Vm and aKi of hepatocytes in liver slices were measured by conventional and ion-sensitive microelectrodes, respectively. Cell water vols. were estimated by differences in wet and dry weights of liver slices after 10-min incubations. Effect of hyperosmotic medium on membrane transference number for K+, tK, was measured by effects on Vm of step-changes in external [K+]. Hepatocyte Vm decreased 34, 52 and 54% when tissue was superfused with medium made hyperosmotic with added sucrose (50, 100 and 150 mM). Correspondingly, aKi increased 10, 18 and 29% with this hyperosmotic stress of added sucrose. Tissue water of 2.92 +/- 0.10 kg H2O/kg dry weight in control solution decreased to 2.60 +/- 0.05, 2.25 +/- 0.06 and 2.22 +/- 0.05 kg H2O/kg dry weight with additions to medium of 50, 100 and 150 mM sucrose, respectively. Adding 50 mM sucrose to medium decreased tK from 0.20 +/- 0.01 to 0.05 +/- 0.01. Depolarization by 50% with hyperosmotic stress (100 mM sucrose) also occurred in Cl-free medium where Cl- was substituted with gluconate. We conclude that hepatocytes shrink during hyperosmotic stress, and the aKi increases. The accompanying decrease in Vm is opposite to that expected by an increase in aKi, and at least in part results from a concomitant decrease in gK. Changes in membrane Cl- conductance most likely do not contribute to osmotic stress-induced depolarization, since equivalent decreases in Vm occurred with added sucrose in cells depleted of Cl- by superfusing tissue with Cl-free medium.  相似文献   

6.
The effect of salt stress (NaCl 85.7 or 110 mmol/L) was investigated in the triticale T300 and its parental species, Triticum dicoccum farrum (Triticum df) and Secale cereale cv. Petkus. Triticum df and T300 were more salt-tolerant than the rye (110 mmol/L NaCl was the highest concentration allowing rye growth to the three-leaf stage). Na+, K+ and Cl- ions accounted for almost half of the osmotic adjustment in Triticum df and T300, and up to 90% in rye. Salinity decreased the net photosynthesis and transpiration rates of the three cereals as compared to control plants, but induced no significant change in chlorophyll a fluorescence parameters. Water-use efficiency (WUE) increased with salinity. In the presence of 110 mmol/L NaCl, the K+/Na+ ratio decreased markedly in rye as compared to the other two cereals. Proline concentration, which increased in Triticum df and T300, could have protected membrane selectivity in favour of K+. Proline content remained low in rye, and increasing soluble sugar content did not appear to prevent competition between Na+ and K+. The salt sensitivity of rye could be due to low K+ uptake in the presence of a high NaCl concentration.  相似文献   

7.
Changes in the oxidation-reduction potential of medium during the growth of Escherichia coli bacteria under anaerobic conditions in the pH range 5.5-7.5 were studied. It was shown that dithiothreitol, a reducer of thiol groups, at a concentration of 3 mM decreased the initial value of oxidation-reduction potential to negative values of -60 to -220 mV, increased the duration of the lag growth phase, decreased the growth rate, and lowered the acidification of the medium and accumulation of potassium ions independently of pH. The value of the potential in the presence of 3 mM dithiothreitol decreased by 15-20 mV both at alkaline and acidic pH. Dithiothreitol at a concentration of 10 mM stopped molecular hydrogen production at pH 5.5 only. The effects of dithiothreitol on the oxidation-reduction potential and bacterial growth were probably associated with changes in the state of thiol groups in membrane proteins, which affects the generation of membrane potential, proton secretion from cells, K+ accumulation in bacteria, and the activity of membrane-associated enzymes.  相似文献   

8.
Rhodococcus sp. RB1 was able to thrive in media with up to 0.9 M NaCl or KCl and in the presence of high concentrations of nitrate (up to 0.9 M) and nitrite (up to 60 mM), but only under oxic conditions. An adaptation period was not required for salt tolerance, but a rapid extrusion of K+ and intake of Na+ was observed after addition of 0.5 M NaCl. Nitrate assimilation was limited by the carbon supply, but nitrite was not accumulated in the culture medium, even at nitrate concentrations as high as 0.8 M, thus suggesting that nitrite reduction does not limit nitrate assimilation. The presence of NaCl or KCl did not affect nitrate or nitrite uptake, which were completely inhibited by ammonium or glutamine. Rhodococcus sp. RB1 nitrate reductase had an apparent molecular mass of 142 kDa and used NADH and reduced bromophenol blue or viologens as electron donors, independently of the presence of salt. The enzyme was associated with an NADH-diaphorase activity and was induced by nitrate and repressed by ammonium or glutamine, thus showing typical biochemical and regulatory properties of bacterial assimilatory NADH-nitrate reductases. The enzyme was active in vitro in the presence of 3 M NaCl or KCI, but the maximal activity was observed at 0.5 M salt. Addition of 2 M NaCl increased the optimal temperature of the enzyme from 12 to 32 degrees C, but the optimal pH (10.3) was unaffected.  相似文献   

9.
Escherichia coli is able to grow at increased NaCl concentrations that provides an increase in medium osmolarity and cellular Na+ content. The addition of 0.5 M NaCl to the growth medium led to a substantial decrease in growth rate during anaerobic fermentation on glucose at pH of 7.3 or 9.0. This inhibitory effect of 0.5 M NaCl was at least threefold stronger than that seen under aerobic conditions, and stronger than equivalent concentrations of sucrose, KCl, or potassium glutamate under anaerobic conditions. Further, proline was found to stimulate the growth rate at high NaCl concentration under anaerobic and to a lesser extent, under aerobic conditions. Wild-type cells and mutants having a functional NhaA or ChaA alone grown under anaerobic conditions at pH 9.0 and subsequently loaded with Na+ were shown to extrude Na+ at a rate that were lower than the extrusion rate reported for appropriate aerobically grown bacteria (Sakuma et al. [1998] Biochim Biophys Acta 1363:231–237). The growth rate and Na+ extrusion activity of a mutant having a functional NhaA were similar to that of the wild type and higher than that of a mutant with an active ChaA. A mutant defective for both NhaA and ChaA was unable to grow under anaerobic conditions at pH 9.0 in the presence of 0.15 M Na+. It is suggested that the observed strong inhibition in the growth of E. coli during fermentation under anaerobic conditions in the presence of increased NaCl concentration could be due to a decrease in Na+ extrusion activity. Received: 18 September 1998 / Accepted: 2 April 1999  相似文献   

10.
Bacteria respond to osmotic stress by a substantial increase in the intracellular osmolality, adjusting their cell turgor for altered growth conditions. Using Escherichia coli as a model organism we demonstrate here that bacterial responses to hyperosmotic stress specifically depend on the nature of osmoticum used. We show that increasing acute hyperosmotic NaCl stress above ∼1.0 Os kg−1 causes a dose-dependent K+ leak from the cell, resulting in a substantial decrease in cytosolic K+ content and a concurrent accumulation of Na+ in the cell. At the same time, isotonic sucrose or mannitol treatment (non-ionic osmotica) results in a gradual increase of the net K+ uptake. Ion flux data are consistent with growth experiments showing that bacterial growth is impaired by NaCl at the concentration resulting in a switch from net K+ uptake to efflux. Microarray experiments reveal that about 40% of upregulated genes shared no similarity in their responses to NaCl and sucrose treatment, further suggesting specificity of osmotic adjustment in E. coli to ionic and non-ionic osmotica. The observed differences are explained by the specificity of the stress-induced changes in the membrane potential of bacterial cells highlighting the importance of voltage-gated K+ transporters for bacterial adaptation to hyperosmotic stress.  相似文献   

11.
Isolated caprine early-staged follicles were submitted to osmotic tolerance tests in the presence of sucrose, ethylene glycol (EG), or NaCl solutions and were exposed to and cryopreserved (by slow or rapid cooling) in MEM alone or MEM supplemented with sucrose, EG (1.0 or 4.0 M), or both. When follicles were exposed to 1.5 M NaCl, only 2% of the follicles were viable, whereas 87% of the follicles were viable after exposure to 4.0 M EG. Regarding exposure time, the highest percentage of viable follicles was obtained when follicles were exposed for 10 min to 1.0 M EG + 0.5 M sucrose; exposure for 60 s to 4.0 M EG + 0.5 M sucrose also maintained high percentage viability in follicles. Slow cooling in the presence of 1.0 M EG + 0.5 M sucrose (75%) or rapid cooling in the presence of 4.0 M EG + 0.5 M sucrose (71%) resulted in a significantly higher proportion of viable follicles than all other treatments (P < 0.05). A 24-h culture of frozen-thawed follicles was used to assess survival; only slow-frozen follicles showed viability rates similar to control follicles (64% vs. 69% respectively; P > 0.05). Interestingly, the percentage of viable rapid-cooled follicles (59%) was similar to that obtained after in vitro culture of conventional slow-cooled follicles but was significantly lower than that in controls. Thus, in addition to determining improved procedures for the exposure of follicles to EG and sucrose before and after freezing of caprine early-staged follicles, we report the development of rapid- and slow-cooling protocols.  相似文献   

12.
The effects of salt stress on carbohydrate metabolism in Microcoleus vaginatus Gom., a cyanobacterium isolated from desert algal crusts, were investigated in the present study. Extracellular total carbohydrates and exopolysaccharides (EPS) in the culture medium produced by M. vaginatus increased significantly during the growth phase and reached a maximum during the stationary phase. The production of extracellular carbohydrates also significantly increased under higher salt concentrations, which was attributed to an increase in low molecular weight carbohydrates. In the presence of NaCI, the production of cellular total carbohydrates decreased and photosynthetic activity was impaired, whereas cellular reducing sugars, water-soluble sugars and sucrose content and sucrose phosphate synthase activity increased, reaching a maximum in the presence of 200 mmol/L NaCI. These parameters were restored to original levels when the algae were transferred to a non-saline medium. Sodium and K+ concentrations of stressed cells decreased significantly and H+-ATPase activity increased after the addition of exogenous sucrose or EPS. The results suggest that EPS and sucrose are synthesized to maintain the cellular osmOtic equilibrium between the intra-and extracellular environment, thus protecting algal cells from osmotic damage, which was attributed to the selective exclusion of cellular Na+ and K+ by H+-ATPase.  相似文献   

13.
Cultured chick cardiac cells possess a Na+K+Cl-co-transport system that is inhibited by the "loop diuretics" benzmetanide (IC50 = 0.3 microM), bumetanide (IC50 = 0.6 microM), piretanide (IC50 = 1.5 microM) and furosemide (IC50 = 5 microM). The K0.5 values for Cl- and Na+ activation of the bumetanide-sensitive 86Rb+ uptake are 59 mM and 40mM respectively. Bumetanide also inhibits a 22Na+ uptake component that is suppressed when external Cl- or K+ are substituted by impermeant ions. The ratio of bumetanide-sensitive 86Rb+ to 22Na+ uptake is close to 1. The cardiac Na+/K+/Cl- cotransport is a major uptake pathway for Na+ and K+. It accounts for 50% of the initial rate of 86Rb+ uptake and 17% of the initial rate of 22Na+ uptake by chick cardiac cells. It is activated two-fold by an hyperosmotic shock produced with 200 mM mannitol.  相似文献   

14.
Conditions optimum for the assay of alkaline phosphatase of marine pseudomonad B-16 (ATCC 19855) and for maintaining the activity of the enzyme have been determined. The pH for optimal activity of the cell-bound enzyme was 9.0, whereas that for the enzyme after its release from the cells exceeded 9.4. Release was effected by first washing the cells in 0.5 M NaCl and then suspending them in 0.5 M sucrose. In the absence of salts, the activity of the cell-bound enzyme decreased rapidly at 25 C and less rapidly at 4 C. This loss of activity could be arrested but not restored by adding Mg(2+). In the presence of Na(+), activity of the cell-bound enzyme dropped to about 50% of that prevailing initially, but in this case adding Mg(2+) restored enzyme activity completely. The activity of the enzyme after its release from the cells into 0.5 M sucrose was approximately 50% of that of the equivalent amount of enzyme in the original cells. This activity was relatively stable at both 25 and 4 C. Adding Mg(2+) to the released enzyme restored its activity to that of the cell-bound form. The synthesis of alkaline phosphatase by the cells was not affected by adding 50 mM inorganic phosphate to the growth medium. The K(m) of the released enzyme for p-nitrophenyl phosphate was found to be 6.1 x 10(-5) M.  相似文献   

15.
Internalization of Sucrose by Methanococcus thermolithotrophicus   总被引:1,自引:1,他引:0       下载免费PDF全文
When sucrose is present in the external medium, it is internalized by Methanococcus thermolithotrophicus. Sucrose internalization, as determined by both natural abundance (sup13)C nuclear magnetic resonance spectroscopy and [(sup14)C]sucrose uptake, is directly proportional to external sucrose levels. The uptake is energy independent and exhibits kinetic behavior consistent with a simple passive diffusion process. In the presence of 0.2 M sucrose, methanogenesis is inhibited as the NaCl concentration in the external medium is increased. Growth, as determined by protein content, is inhibited by 0.2 M sucrose when the external NaCl concentration is 1.4 M. These results are important because they show that (i) sucrose cannot be used as a noncharged solute to replace NaCl in experiments to evaluate how external osmotic strength affects the internal solute composition of M. thermolithotrophicus, and (ii) sucrose cannot be used as an impermeable marker for the extracellular volume in experiments to measure the intracellular volume of M. thermolithotrophicus.  相似文献   

16.
Growth of Escherichia coli K-12 in a modified Davis minimal medium was inhibited under high osmolarity, but it recovered remarkably with the addition of 1 mM proline. The co-existence of K+ with proline enhanced the recovery of growth under high osmolarity more than that in the presence of proline alone. The same was true for the activities of respiration and glucose uptake. A similar supplementary effect of K+ was observed for the activities of proline uptake under high osmolarity. These results suggest that K+ and proline support not only growth but respiration and uptake of the respiratory substrate glucose in the cell cytoplasm when exposed to high osmolarity. External K+ almost disappeared with 1 h of incubation at low osmolarity, indicating that active accumulation of K+ in the cells occurred. On the other hand, a gradual accumulation of K+ was recognized at high osmolarity in the presence of 1 M NaCl, especially at > 2 h of incubation. This study of L-[5-3H]proline uptake in the cell cytoplasm indicates that proline was incorporated as a substrate of protein synthesis in the absence of NaCl, but was efficiently utilized as a compatible solute in the presence of high concentrations of NaCl.  相似文献   

17.
ASCT1 protein is a member of the glutamate transporter superfamily, which shows system ASC selectivity and properties and has been characterized as a Na+-dependent neutral amino-acid exchanger. Here, by using ASCT1-expressing oocytes, the uptake of alanine and glutamate was measured to investigate ASCT1's ability to mediate a concentrative transport of alanine, ASCT1's sodium dependence, and the influence of pH on the mutual inhibition between alanine and glutamate. Alanine uptake was measured after 30 min incubation. Kinetic analysis of the Na+ dependence of alanine uptake showed an apparent K0.5 (affinity constant) value for Na+ of 23.1 +/- 4.3 mM (mean +/- SE). Concentration dependence of alanine uptake was tested at 100 and 1 mM Na+, with apparent K0.5 values of 0.16 +/- 0.04 and 1.8 +/- 0.4 mM, respectively, at pH 7.5, and 0.21 +/- 0.06 and 1.9 +/- 0.3 mM at pH 6. Vmax was not modified between 100 and 1 mM Na+ at either pH. ASCT1 actively transports alanine and accumulates it in the cytosol even when the Na+ concentration in the medium was as low as 1-3 mM. 22Na uptake studies revealed that Na+ transport was stimulated by the presence of alanine in the medium. Our results demonstrate that ASCT1 is able to mediate a concentrative transport of alanine, which is Na+-dependent but not coupled to the Na+ gradient.  相似文献   

18.
Anaerobically grown glucose-fermenting E. coli cells produce molecular hydrogen, acidify the medium and uptake potassium ions. It was shown that the H2 release and the proton-potassium exchange with the fixed (2H+/K+) stoichiometry of the initial DCC-sensitive fluxes were lost in mutants with the deleted fdhF gene or the hycA-H operon responsible for the biosynthesis of formate dehydrogenase H (FDH,H) or hydrogenase 3 (H3), respectively, which are the main components of the formate hydrogen lyase FHL(H). However, both processes occurred in mutants with the deleted hycE, hycF or hycG genes encoding the major and minor components of H3, respectively. The K+ uptake was sensitive to the osmotic shock resulting from glucose addition to the medium and decreased significantly in the presence of valinomycin. The H2 release and the 2H+/K+ exchange were absent in the mutant with the deleted hycB gene encoding the corresponding minor component of H3. This mutant acidified the medium and uptook K+ with Km typical for TrkA, but the stoichiometry of the DCC-inhibited fluxes was variable, and the K+ gradient between the cytoplasm and the medium in this mutant was lower than in the mutants lacking other minor components of H3. The results obtained suggest that the hycB gene product, FdhF and HycE, form probably the FHL(H) complex that directly interacts with the H+-ATPase complex F0F1 and the TrkA(H) system of K+ uptake. Such a multienzyme association is responsible for the H2 production and 2H+/K+ exchange. The major and other minor components of H3 have probably no direct role in the H2 production and 2H+/K+ exchange. H2 production by precursor's or hycE mutant's protoplasts treated with toluene was shown to occur upon addition of the thiol reagent dithiothreitol to the medium containing ATP, potassium ions, NAD+, and NADH. H2 production was inhibited by DCC. The quantity of available thiol groups in membrane vesicles of the precursor or the hycE, hycF or hycG mutants, in which the H2 production and 2H+/K+ exchange were observed, was larger than in other mutants. The number of SH groups decreased in the presence of DCC. These results indicate a significance of the thiol groups for the function of the proposed association.  相似文献   

19.
(1) (DL)-Propranolol and Ca2+ are shown to alter the transmembrane potential difference of Ehrlich ascites tumor cells as measured by means of the cyanine dye, 3,3'-dipropyl-2,2'-thiodicarbocyanine iodide, whose fluorescent intensity changes as a function of membrane potential. (2) The changes in membrane potential elicited by these agents are dependent of the external K+ concentration in a manner which suggest that the potential changes result from a specific increase in the permeability of the plasma membrane to K+. (3) Na+-dependent amino acid transport in the presence of propranolol can be modulated by varying the external K+ concentration (K+o). The initial rate of uptake is stimulated by propranolol at low K+o and inhibited at high K+o. The change in transport rate is nearly directly proportional to the natural logarithm of [K+]o in the presence of propranolol. (4) ATP depletion of the cells by preincubation with rotenone abolishes the changes in fluorescence and amino acid uptake seen with propranolol as a function of K+o. Restoration of cellular ATP with glucose in presence of Ca2+ restores both fluorescence and amino acid transport changes which occur in response to propranolol. (5) The fluorescence changes and amino acid transport changes in response to propranolol are pH dependent, with little effect seen at pH6. (6) It is concluded that the rate of Na+-dependent amino acid uptake is a function of membrane potential and is dependent on the electrochemical potential difference for Na+.  相似文献   

20.
We have studied the effects of several cations on (1) the neuronal uptake of [3H]dopamine ([3H]DA) and (2) the specific binding of 1-[2-(diphenylmethoxy)ethyl]-4-(3-phenyl-2-[1-3H]propenyl)piperazi ne ([3H]GBR 12783) to a site associated with the neuronal carrier of DA, in preparations obtained from rat striatum. When studied under the same experimental conditions, both the uptake of [3H]DA and the binding of [3H]GBR 12783 were similarly impaired by the gradual replacement of NaCl by sucrose. In both processes, no convenient substitute for Na+ was found. Furthermore, potential substitutes of Na+ acted as inhibitors of the uptake with a rank order of potency as follows: K+ = Li+ > or = Cs+ > or = Rb+ > choline+ > Tris+ > sucrose, which was somewhat different from that observed in binding studies, i.e., Cs+ > Rb+ > choline+ > or = K+ > Li+ > Tris+ > sucrose. In the presence of either 36 mM or 136 mM Na+, [3H]DA uptake was optimal with 2 mM Mg2+, 1 mM K+, or 1 mM Ca2+. In contrast, higher concentrations of divalent cations competitively blocked the uptake process. K+ concentrations > 50 mM impaired the specific binding, whereas in the millimolar range of concentrations, K+ noncompetitively inhibited the uptake. Decreasing the Na+ concentration increased the inhibitory effect of K+, Ca2+, and Mg2+ on the specific uptake. An increase in NaCl concentration from 0 to 120 mM elicited a significant decline in the affinity of some substrates for the [3H]GBR 12783 binding site. An uptake study performed using optimal experimental conditions defined in the present study revealed that decreasing Na+ concentration reduces the affinity of DA for the neuronal transport. We propose a hypothetical model for the neuronal transport of DA in which both Na+ and K+ membrane gradients are involved.  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