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.
Summary The effect of chloride on 4,4-dibenzamido-2,2-disulfonic stilbene (DBDS) binding to band 3 in unsealed red cell ghost membranes was studied in buffer [NaCl (0 to 500mm) + Na citrate] at constant ionic strength (160 or 600mm). pH 7.4, 25°C. In the presence of chloride, DBDS binds to a single class of sites on band 3. At 160mm ionic strength, the dissociation constant of DBDS increases linearly with chloride concentration in the range [Cl]=450mm. The observed rate of DBDS binding to ghost membranes, as measured by fluorescence stopped-flow kinetic experiments, increases with chloride concentration at both 160 and 600mm ionic strength. The equilibrium and kinetic results have been incorporated into the following model of the DBDS-band 3 interaction: The equilibrium and rate constants of the model at 600mm ionic strength areK 1=0.67±0.16 m,k 2=1.6±0.7 sec–1,k –2=0.17±0.09 sec–1,K 1=6.3±1.7 m,k 2=9±4 sec–1 andk –2=7±3 sec–1. The apparent dissociation constants of chloride from band 3,K Cl, are 40±4mm (160mm ionic strength) and 11±3mm (600mm ionic strength). Our results indicate that chloride and DBDS have distinct, interacting binding sites on band 3.  相似文献   

2.
Summary In separated outer medullary collecting duct (MCD) cells, the time course of binding of the fluorescent stilbene anion exchange inhibitor, DBDS (4,4-dibenzamido-2,2-stilbene disulfonate), to the MCD cell analog of band 3, the red blood cell (rbc) anion exchange protein, can be measured by the stopped-flow method and the reaction time constant, DBDS, can be used to report on the conformational state of the band 3 analog. In order to validate the method we have now shown that the ID50,DBDS,MCD (0.5±0.1 m) for the H2-DIDS (4,4-diisothiocyano-2,2-dihydrostilbene disulfonate) inhibition of DBDS is in agreement with the ID50,Cl ,MCD (0.94±0.07 m) for H2-DIDS inhibition of MCD cell Cl flux, thus relating DBDS directly to anion exchange. The specific cardiac glycoside cation transport inhibitor, ouabain, not only modulates DBDS binding kinetics, but also increases the time constant for Cl exchange by a factor of two, from Cl=0.30±0.02 sec to 0.56±0.06 sec (30mm NaHCO3). The ID50,DBDS,MCD for the ouabain effect on DBDS binding kinetics is 0.003±0.001 m, so that binding is about an order of magnitude tighter than that for inhibition of rbc K+ flux (K I,K +,rbc=0.017 m). These experiments indicate that the Na+,K-ATPase, required to maintain cation gradients across the MCD cell membrane, is close enough to the band 3 analog that conformational information can be exchanged. Cytochalasin E (CE), which binds to the spectrin/actin complex in rbc and other cells, modulates DBDS binding kinetics with a physiological ID50,DBDS,MCD (0.076±0.005 m); 2 m CE also more than doubles the Cl exchange time constant from 0.20±0.04 sec to 0.50±0.08 sec (30mm NaHCO3). These experiments indicate that conformational information can also be exchanged between the MCD cell band 3 analog and the MCD cell cytoskeleton.  相似文献   

3.
Summary The fluorescence enhancement of 4,4-dibenzamido-2,2-disulfonic stilbene (DBDS) upon binding to membranes was used to examine proximal tubule stilbene binding sites. Equilibrium binding studies of DBDS to renal brush border (BBMV) and basolateral membrane vesicles (BLMV) were performed using a fluorescence enhancement technique developed for red blood cells (A.S. Verkman, J.A. Dix and A.K. Solomon,J. Gen. Physiol. 81:421–449, 1983). In the absence of transportable anions, DBDS bound reversibly to a single class of sites on BLMV isolated from rabbit (K d =3.8 m) and rat (3.2 m); 100 m dihydro-4,4-diisothiocyano-2,2-disulfonic stilbene (H2DIDS) blocked >95% of binding. H2DIDS inhibitable DBDS binding was not detected using rat or rabbit BBMV. In rabbit BLMV, DBDSK d doubled with 10mm SO4, 50mm HCO3 and 100mm Cl, but was not altered by Na or pH (6–8). In stopped-flow experiments the exponential time constant for DBDS binding slowed with SO4, HCO3 and Cl, but was unaffected by Na. These results are consistent with competitive binding of DBDS and anions at an anion transport site. To relate DBDS binding data to anion transport inhibition we used35SO4 uptake to characterize several modes of rabbit BLM anion transport: H/SO4 and Na/SO4 cotransport, and Cl/SO4 countertransport. Each transport process was electroneutral and was inhibited by H2DIDS, furosemide, probenecid, chlorothiazide and DBDS. The apparentK t 's for DBDS (3–20 m) were similar toK d for DBDS binding. These studies define a class of anion transport sites on the proximal tubule basolateral membrane measureable optically by a fluorescent stilbene.  相似文献   

4.
Summary MDCK cells, when examined by low-light level video microscopy displayed an endogenous fluorescence with two differing patterns. A low intensity emission which was punctate and associated with cell organelles was observed with emission and excitation conditions generally used to observe either fluorescein (450–500 nm excitation/>510 nm emission) or rhodamine (514 nm excitation/>530 emission) type dyes. A second 5- to 10-fold brighter emission for 450–500 nm excitation was observed, which was unusual in that each cell appeared to be outlined. Evidence obtained from spectroscopy and from using culture media of altered composition supported the conclusion that the water-soluble vitamin riboflavin accumulated in the basolateral spaces and fluid-filled domes and was the source of this fluorescent emission. Quantitative measurements showed that exposure to cultures to 10 m riboflavin resulted in accumulation in domes of 565±80 m. The transport rate was calculated to be 189±30 pmol/min-cm2. Onemm probenecid, a known inhibitor of riboflavin transport in vivo, reduced transport to 54% of control, while 10mm nearly abolished the uptake. The results demonstrate that removal of riboflavin reduces MDCK cell fluorescence to levels compatable with low-light level imaging. Furthermore, these cells actively transport riboflavin and provide a new in vitro model for this process.  相似文献   

5.
Summary The time course of binding of the fluorescent stilbene anion exchange inhibitor, DBDS (4,4-dibenzamido-2,2-stilbene disulfonate), to band 3 can be measured by the stopped-flow method. We have previously used the reaction time constant, DBDS, to obtain the kinetic constants for binding and, thus, to report on the conformational state of the band 3 binding site. To validate the method, we have now shown that the ID50 (0.3±0.1 m) for H2-DIDS (4,4-diisothiocyano-2,2-dihydrostilbene disulfonate) inhibition of DBDS is virtually the same as the ID50 (0.47±0.04 m) for H2-DIDS inhibition of red cell Cl flux, thus relating DBDS directly to band 3 anion exchange. The specific glucose transport inhibitor, cytochalasin B, causes significant changes in DBDS, which can be reversed with intracellular, but not extracellular,d-glucose. ID50 for cytochalasin B modulation of DBDS is 0.1±0.2 m in good agreement withK D =0.06±0.005 m for cytochalasin B binding to the glucose transport protein. These experiments suggest that the glucose transport protein is either adjacent to band 3, or linked to it through a mechanism, which can transmit conformational information. Ouabain (0.1 m), the specific inhibitor of red cell Na+,K+-ATPase, increases red cell Cl exchange flux in red cells by a factor of about two. This interaction indicates that the Na+,K+-ATPase, like the glucose transport protein, is either in contact with, or closely linked to, band 3. These results would be consistent with a transport proteincomplex, centered on band 3, and responsible for the entire transport process, not only the provision of metabolic energy, but also the actual carriage of the cations and anions themselves.  相似文献   

6.
Summary The kinetics of the initial phases of d-glucose binding to the glucose transport protein (GLUT1) of the human red cell can be followed by stopped-flow measurements of the time course of tryptophan (trp) fluorescence enhancement. A number of control experiments have shown that the trp fluorescence kinetics are the result of conformational changes in GLUT1. One shows that nontransportable l-glucose has no kinetic response, in contrast to d-glucose kinetics. Other controls show that d-glucose binding is inhibited by cytochalasin B and by extracellular d-maltose. A typical time course for a transportable sugar, such as d-glucose, consists of a zero-time displacement, too fast for us to measure, followed by three rapid reactions whose exponential time courses have rate constants of0.5–100 sec+–1 at 20°C. It is suggested that the zero-time displacement represents the initial bimolecular ligand/GLUT1 association. Exponential 1 appears to be located at, or near, the external membrane face where it is involved in discriminating among the sugars. Exponential 3 is apparently controlled by events at the cytosolic face. Trp kinetics distinguish the K d of the epimer, d-galactose, from the K dfor d-glucose, with results in agreement with determinations by other methods. Trp kinetics distinguish between the binding of the - and -d-glucose anomers. The exponential 1 activation energy of the -anomer, 13.6 ± 1.4 kcal mol+–1, is less than that of -d-glucose, 18.4 ± 0.8 kcal mol+–1, and the two Arrhenius lines cross at 23.5°C. The temperature dependence of the kinetic response following -d-glucose binding illustrates the interplay among the exponentials and the increasing dominance of exponential 2 as the temperature increases from 22.3 to 36.6°C. The existence of these interrelations means that previously acceptable approximations in simplified reaction schemes for sugar transport will now have to be justified on a point-to-point basis.We should like to express our thanks to Michael R. Toon for his important contributions. This work was supported in part by a grant-in-aid from the American Heart Association, by the Squibb Institute for Medical Research and by The Council for Tobacco Research.  相似文献   

7.
Summary In jejunal brush-border membrane vesicles, an out-wardly directed OH gradient (in>out) stimulates DIDS-sensitive, saturable folate (F) uptake (Schron, C.M., 1985).J. Clin. Invest. 76:2030–2033), suggesting carrier-mediated folate: OH exchange (or phenomenologically indistiguishable H+: folate cotransport). In the present study, the precise role of pH in the transport process was elucidated by examinin F uptake at varying pH. For pH gradients of identical magnitude, F uptake (0.1 M) was geater at lower (pHint/pHext:5.5/4.5) compared with higher (6.5/5.5) pH ranges. In the absence of a pH gradient, internal Ftrans stimulated DIDS-sensitive3H-folate uptake only at pH6.0. Since setepwise increments ininternal pH (4.57.5; pHext=4.5) stimulated F uptake, an inhibitory effect of higherinternal pH was excluded. In contrast, with increasing external pH(4.356.5; pHint=7.8), a 50-fold decrement in F uptake was observed (H+ K m =12.8±1.2m). Hill plots of these data suggest involvement of at least one H+ (OH) at high pH (divalent F–2 predominates). Since an inside-negative electrical potential did not affect F uptake at either pHext 4.55 or 5.8, transport of F and F–2 is electroneutral. Kinetic parameters for F and F–2 were calculated from uptake data at pHext 4.55 and 5.0. Comparision of predictedvs. experimentally determined kinetic parameters at pHext 5.8 (K m =1.33vs. 1.70 m;V max=12.8vs. 58.0 pmol/mg prot min) suggest that increasing external pH lowers theV max, but does not affect thatK m, for carrier-mediated F transport. These data are consistent with similarK i's for sulfasalazine (competitive inhibitor) at pHext 5.35 and 5.8 (64.7 and 58.5 m, respectively). In summary, the jejunal F carrier mediates electroneutral transport of mono- and divalen F and is sensitive to extermal pH with a H+ K m (or OH IC50) corresponding to pH 4.89. External pH affects theV max, but not theK m for carriermediated F uptake suggesting a reaction mechanism involving a ternary complex between the outward-facing conformation of the carrier and the transported ions (F and either OH or H+) rather than competitive binding that is mutually exclusive.  相似文献   

8.
Summary The effects of glucose on cellular respiration were examined in suspensions of rabbit cortical tubules. When glucose was removed from the bathing fluid, oxygen consumption (QO2) decreased from 18.6±0.8 to 15.7±0.5 nmol O2/mg protein·min (P<0.01). The transported but nonmetabolized analogue of glucose, -methyl-d-glucoside (MG), was found to support QO2 to the same extent as glucose. These observations were also evident in the presence of butyrate, a readily oxidized substrate of the renal cortex. Additional studies with nystatin and ouabain indicated that glucose-related changes in QO2 were the result of changes in Na, K-ATPase associated respiration. The effect of glucose was localized to the luminal membrane since phlorizin (10–5 m), a specific inhibitor of liminalk glucose-sodium cotransport, also significantly reduced QO2 by 10±1%. Phlorizin inhibition of QO2 was also evident in the presence of MG but was abolished when glucose was removed from the bathing medium. Finally, measurement of NADH fluorescence showed that addition of glucose (5mm) to a tubule suspension causes an oxidation of NAD. These data are all consistent with glucose acting to increase respiration by stimulating sodium entry at the luminal membrane (via glucose-sodium cotransport) followed by increased sodium pump activity and its associated increase in mitochondrial respiration.  相似文献   

9.
Summary Cell K activity,a k, was measured in the short-circuited frog skin by simultaneous cell punctures from the apical surface with open-tip and K-selective microelectrodes. Strict criteria for acceptance of impalements included constancy of the open-tip microelectrode resistance, agreement within 3% of the fractional apical voltage measured with open-tip and K-selective microelectrodes, and constancy of the differential voltage recorded between the open-tip and the K microelectrodes 30–60 sec after application of amiloride or substitution of apical Na. Skins were bathed on the serosal surface with NaCl Ringer and, to reduce paracellular Cl conductance and effects of amiloride on paracellular conductance, with NaNO3 Ringer on the apical surface.Under control conditionsa k r was nearly constant among skins (mean±SD=92±8mM, 14 skins) in spite of a wide range of cellular currents (5 to 70 A/cm2). Cell current (and transcellular Na transport) was inhibited by either apical addition of amiloride or substitution of Na by other cations. Although in some experiments the expected small increase ina k r after inhibition of cell current was observed, on the average the change was not significant (98±11mM after amiloride, 101±12mM after Na substitution), even 30 min after the inhibition of cell current. The membrane potential, which in the control state ranged from –42 to –77 mV, hyperpolarized after inhibition of cell current, initially to –109±5mV, then depolarizing to a stable value (–88±5mV) after 15–25 min. At this time K was above equilibrium (E k=98±2mV), indicating that the active pump mechanism is still operating after inhibition of transcellular Na transport.The measurement ofa k r permitted the calculation of the passive K current and pump current under control conditions. assuming a constant current source with almost all of the basolateral conductance attributable to K. We found a significant correlation between pump current and cell current with a slope of 0.31, indicating that about one-third of the cell current is carried by the pump, i.e., a pump stoichiometry of 3Na/2K.  相似文献   

10.
Summary Addition of glucose or the nonmetabolizable analogue -methyl-d-glucoside to rabbit proximal tubules suspended in a glucoseand alanine-free buffer caused a sustained increase in intracellular Na+ content (+43±7 nmol · (mg protein)–1) and a concomitant but larger decrease in K+ content (–72±11 nmol· (mg protein)–1). A component of the net K+ efflux was Ba2+ insensitive, and was inhibited by high (1mm) but not low (10 m) concentrations of the diuretics, furosemide and bumetanide. The increase in intracellular Na+ content is consistent with the view that the increased rates of Na+ and water transport seen in the proximal tubule in the presence of glucose can be attributed (at least in part) to a stimulation of basolateral pump activity by an increased [Na+] i .  相似文献   

11.
Summary The role of adenosine 3,5-monophosphate (cAMP) dependent protein kinase (PK-A) on the Cl conductance has been studied in the apical membrane vesicles purified from the chorionic villi of human placenta. In order to phosphorylate the cytosolic side of the membranes, vesicles have been hypotonically lysed, loaded with 100nm catalytic subunit of PK-A purified from human placenta and 1mm of the phosphatase resistant adenosine 5-thiotriphosphate (ATP-gamma-S) and resealed. Cl conductance has been measured by the quenching of the fluorescent probe 6-methoxy-N-(3-sulfopropyl) quinolinium (SPQ) at 23°C with membrane potential clamped at 0 mV. The actual volume of the resealed vesicles was measured in each experiment by trapping an impermeable radioactive molecule ([14C]-sucrose) and included in each Cl flux calculation. In 19 independent experiments, the mean Cl conductance in placental membranes in the absence of phosphorylation was 3.67±3.18 whereas with the addition of PK-A and ATP-gamma-S it was 1.97±1.75 nmol·sec–1·(mg protein)–1 (mean±sd). PK-A dependent phosphorylation reduced the Cl conductance in 14/19 experiments. The same protocol applied to the apical membranes of bovine trachea, where PK-A is known to activate the Cl channels, confirmed that the PK-A dependent phosphorylation increased the Cl conductance in 11/13 experiments, from 1.01±0.61 to 1.85±0.99 nmol·sec–1·(mg protein)–1(mean±sd). These studies indicate that the PK-A dependent phosphorylation inhibits one or more Cl channel(s) of the apical membranes of human placenta.  相似文献   

12.
Summary Sodium (22Na) transport was studied in a basolateral membrane vesicle preparation from rabbit parotid. Sodium uptake was markedly dependent on the presence of both K+ and Cl in the extravesicular medium, being reduced 5 times when K+ was replaced by a nonphysiologic cation and 10 times when Cl was replaced by a nonphysiologic anion. Sodium uptake was stimulated by gradients of either K+ or Cl (relative to nongradient conditions) and could be driven against a sodium concentration gradient by a KCl gradient. No effect of membrane potentials on KCl-dependent sodium flux could be detected, indicating that this is an electroneutral process. A KCl-dependent component of sodium flux could also be demonstrated under equuilibrium exchange conditions, indicating a direct effect of K+ and Cl on the sodium transport pathway. KCl-dependent sodium uptake exhibited a hyperbolic dependence on sodium concentration consistent with the existence of a single-transport system withK m =3.2mm at 80mm KCl and 23°C. Furosemide inhibited this transporter withK 0.5=2×10–4 m (23°C). When sodium uptake was measured as a function of potassium and chloride concentrations a hyperbolic dependence on [K] (Hill coefficient =1.31±0.07) were observed, consistent with a Na/K/Cl stoichiometry of 112. Taken together these data provide strong evidence for the electroneutral coupling of sodium and KCl movements in this preparation and strongly support the hypothesis that a Na+/K+/Cl cotransport system thought to be associated with transepithelial chloride and water movements in many exocrine glands is present in the parotid acinar basolateral membrane.  相似文献   

13.
Summary In jejunal brush-border membrane vesicles, an outwardly directed OH gradient (in>out) stimulates DIDS-sensitive, saturable folate (F) uptake (Schron, C.M. 1985.J. Clin. Invest. 76:2030–2033), suggesting carrier-mediated folate: OH exchange (or phenomenologically indistinguishable H+: folate cotransport). In the present study, the precise role of pH in the transport process was elucidated by examining F uptake at varying pH. For pH gradients of identical magnitude, F uptake (0.1 M) was greater at lower (pHint/pHext: 5.5/4.5) compared with higher (6.5/5.5) pH ranges. In the absence of a pH gradient, internal Ftrans stimulated DIDS-sensitive3H-folate uptake only at pH6.0. Since stepwise increments ininternal pH (4.57.5; pHext=4.5) stimulated F uptake, an inhibitory effect of higherinternal pH was excluded. In contrast, with increasing external pH (4.356.5; pHint=7.8), a 50-fold decrement in F uptake was observed (H+ K m =12.8±1.2 M). Hill plots of these data suggest involvement of at least one H+ (OH) at low pH (monovalent F predominates) and at least 2 H+ (OH) at high pH (divalent F–2 predominates). Since an inside-negative electrical potential did not affect F uptake at either pHext 4.55 or 5.8, transport of F and F–2 is electroneutral. Kinetic parameters for F and F–2 were calculated from uptake data at pHext 4.55 and 5.0. Comparison of predictedvs. experimentally determined kinetic parameters at pHext5.8 (K m =1.33vs. 1.70 M;V max=123.8vs. 58.0 pmol/mg prot min) suggest that increasing external pH lowers theV max, but does not affect theK m for carrier-mediated F transport. These data are consistent with similarK i ' s for sulfasalazine (competitive inhibitor) at pHext 5.35 and 5.8 (64.7 and 58.5 M, respectively). In summary, the jejunal F carrier mediates electroneutral transport of mono- and divalent F and is sensitive to external pH with a H+ K m (or OH lC50) corresponding to pH 4.89. External pH effects theV max, but not theK m for carriermediated F uptake suggesting a reaction mechanism involving a ternary complex between the outward-facing conformation of the carrier and the transported ions (F and either OH or H+),rather than competitive binding that is mutually exclusive.  相似文献   

14.
Summary When bathed on both sides with identical chloride-containing salines thein vitro preparation of the plaice intestine maintains a negative (serosa to mucosa) short-circuit current of 107±11 A/cm2, a transepithelial potential difference of 5.5±0.6 mV (serosa negative), and a mean mucosal membrane potential of –45.4±0.6 mV. Under these conditions the intracellular chloride activity is 32mm.If chloride in the bathing media is partially, or completely substituted by thiocyanate the measured electrical parameters do not change but transepithelial flux determinations show a reduction in chloride fluxes and the presence of a significant thiocyanate flux. The addition of piretanide (10–4 m) reduced the short-circuit current and the mucosa-to-serosa fluxes of chloride and thiocyanate; this inhibition is similar to the effect of piretanide on chloride transport in this tissue.The results indicate that thiocyanate is transported in this tissue via the piretanide-sensitive chloride pathway and are compared with the effects of thiocyanate on other tissues reported in the literature.  相似文献   

15.
Summary We previously reported that3H-folate uptake by rabbit jejunal brush-border membrane (BBM) vesicles was markedly stimulated by an outwardly directed OH gradient (pHin 7.7, pHout 5.5), inhibited by anion exchange inhibitors (DIDS, SITS, furosemide), and saturable (folateK m=0.19 m) suggesting carrier-mediated folate/OH exchange (or H+/folate cotransport). In the present study, the anion specificity of this transport process was examined. Under conditions of an outwardly directed OH gradient, DIDS-sensitive folate uptake wascis inhibited (>90%) by reduced folate analogues: dihydrofolate (IC50=0.40 m), folinic acid (IC50=0.50 m), 5-methyltetrahydrofolate (IC50=0.53 m), and (+)amethopterin (IC50=0.93 M). In contrast, 10 m (–)amethopterin had only a modest effect on folate uptake (18% inhibition) suggesting stereospecificity of the folate/OH exchanger. The nonpteridine compounds which are transported by the folate carrier in L1210 leukemic cells (phthalate, thiamine pyrophosphate, and PO 4 –3 ) did not inhibit jejunal folate uptake. Furthermore, folate uptake was not inhibited by SO 4 –2 (4mm) or oxalate (4mm) thereby distinguishing this carrier from the previously described intestinal SO 4 –2 /OH and oxalate/Cl exchangers. After BBM vesicles were loaded with3H-folate, the initial velocity of3H-folate efflux was stimulated by unlabeled folate in the efflux medium. The transstimulation of3H-folate efflux by unlabeled folate was furosemide (or DIDS) inhibitable and temperature sensitive. Half-maximal stimulation of furosemide-sensitive3H-folate efflux was observed with 0.25±0.05 m unlabeled folate, a concentration similar to theK m for folate uptake. These data suggest that folate-stimulated3H-folate efflux is mediated by the folate/OH exchanger. With the exception of (–) amethopterin, reduced folate analogues also transstimulated furosemide-sensitive3H-folate efflux in a concentration-dependent manner suggesting stereospecific transport of these analogues by the folate/OH exchanger. In summary, folate transport by the jejunal folate/OH exchanger demonstrates bothcis inhibition and transstimulation by reduced folate analogues, but not by other inorganic or organic anions suggesting bidirectional transport of folate and a high degree of anion specificity.  相似文献   

16.
Proton-linked sugar transport systems in bacteria   总被引:12,自引:0,他引:12  
The cell membranes of various bacteria contain proton-linked transport systems ford-xylose,l-arabinose,d-galactose,d-glucose,l-rhamnose,l-fucose, lactose, and melibiose. The melibiose transporter ofE. coli is linked to both Na+ and H+ translocation. The substrate and inhibitor specificities of the monosaccharide transporters are described. By locating, cloning, and sequencing the genes encoding the sugar/H+ transporters inE. coli, the primary sequences of the transport proteins have been deduced. Those for xylose/H+, arabinose/H+, and galactose/H+ transport are homologous to each other. Furthermore, they are just as similar to the primary sequences of the following: glucose transport proteins found in a Cyanobacterium, yeast, alga, rat, mouse, and man; proteins for transport of galactose, lactose, or maltose in species of yeast; and to a developmentally regulated protein of Leishmania for which a function is not yet established. Some of these proteins catalyze facilitated diffusion of the sugar without cation transport. From the alignments of the homologous amino acid sequences, predictions of common structural features can be made: there are likely to be twelve membrane-spanning -helices, possibly in two groups of six, there is a central hydrophilic region, probably comprised largely of -helix; the highly conserved amino acid residues (40–50 out of 472–522 total) form discrete patterns or motifs throughout the proteins that are presumably critical for substrate recognition and the molecular mechanism of transport. Some of these features are found also in other transport proteins for citrate, tetracycline, lactose, or melibiose, the primary sequences of which are not similar to each other or to the homologous series of transporters. The glucose/Na+ transporter of rabbit and man is different in primary sequence to all the other sugar transporters characterized, but it is homologous to the proline/Na+ transporter ofE. coli, and there is evidence for its structural similarity to glucose/H+ transporters in Plants.In vivo andin vitro mutagenesis of the lactose/H+ and melibiose/Na+ (H+) transporters ofE. coli has identified individual amino acid residues alterations of which affect sugar and/or cation recognition and parameters of transport. Most of the bacterial transport proteins have been identified and the lactose/H+ transporter has been purified. The directions of future investigations are discussed.  相似文献   

17.
Summary Na+–H+ exchange activity in renal brush border membrane vesicles isolated from hyperthyroid rats was increased. When examined as a function of [Na+], treatment altered the initial rate of Na+ uptake by increasingV m (hyperthyroid, 18.9±1.1 nmol Na+ · mg–1 · 2 sec–1; normal, 8.9±0.3 nmol Na+ · mg–1 · 2 sec–1), and not the apparent affinityK Na + (hyperthyroid, 7.3±1.7mm; normal, 6.5±0.9mm). When examined as a function of [H+] and at a subsaturating [Na+] (1mm), hyperthyroidism resulted in the proportional increase in Na+ uptake at every intravesicular pH measured. A positive cooperative effect on Na+ uptake was found with increased intravesicular acidity in vesicles from both normal and hyperthyroid rats. When the data were analyzed by the Hill equation, it was found that hyperthyroidism did not change then (hyperthyroid, 1.2±0.06; normal, 1.2±0.07) or the [H+]0.5 (hyperthyroid, 0.39±0.08 m; normal, 0.44±0.07 m) but increased the apparentV m (hyperthyroid, 1.68±0.14 nmol Na+ · mg–1 · 2 sec–1; normal 0.96±0.10 nmol Na+ · mg–1 · 2 sec–1). The uptake of Na+ in exchange for H+ in membrane vesicles from normal and hyperthyroid animals was not influenced by membrane potential. H+ translocation or debinding was rate limiting for Na+–H+ exchange since Na+–Na+ exchange activity was greater than Na+–H+ exchange activity. Hyperthyroidism caused a proportional increase and hypothyroidism caused a proportional decrease in Na+–Na+ and Na+–H+ exchange. We conclude that hyperthyroidism leads to either an increase in the number of functional exchangers in the membrane or exactly proportional increases in the rate-limiting steps for Na+–Na+ and Na+–H+ exchange activity.  相似文献   

18.
MDCK cell monolayers grown on glass coverslips were used to examine the Na+ concentration in individual lateral intercellular spaces (LIS) by video fluorescence microscopy. The LIS was filled with the Na+-sensitive fluorescent dye SBFO by incubation of the monolayers for 75–90 min with 250 m of the membrane impermeant form of the dye. After dye loading, the monolayers were perfused at 37°C with solutions buffered with HEPES or bicarbonate/CO2 containing 142 mm Na+. Ratios of the fluorescence images after sequential excitation with 340 nm and 380 nm light were performed and in situ calibration of LIS Na+ was accomplished after blocking the Na+ pump with 5 × 10–4 m ouabain. Measurements of Na+ along the basolateral-to-apical axis of the LIS at 1.0 or 1.5 m intervals did not reveal a Na+ gradient when the perfusate was either HEPES or bicarbonate/CO2 solutions. In bicarbonate solutions, the mean Na+ concentration (mm) was 157.2 ± 2.3, 15 mm higher than the bath Na+ concentration. In HEPES solutions, however, the Na+ concentration was not different from the bath concentration (142.7 ± 3.1 mm). The time course of Na+ changes in LIS was investigated by rapidly switching the perfusate from 142 to 80 mm Na+ and measuring the Na+ changes at one focal plane.We would like to thank P.H. Tran and C. Gibson for their technical and computational assistance as well as Dr. B.-E. Persson (University of Uppsala, Sweden) for his contribution in the early phases of the study.  相似文献   

19.
Summary The developmental maturation of Na+–H+ antiporter was determined using a well-validated brush-border membrane vesicles (BBMV's) technique. Na+ uptake represented transport into an osmotically sensitive intravesicular space as evidenced by an osmolality study at equilibrium. An outwardly directed pH gradient (pH inside/pH outside=5.2/7.5) significantly stimulated Na+ uptake compared with no pH gradient conditions at all age groups; however, the magnitude of stimulation was significantly different between the age groups. Moreover, the imposition of greater pH gradient across the vesicles resulted in marked stimulation of Na+ uptake which increased with advancing age. Na+ uptake represented an electroneutral process.The amiloride sensitivity of the pH-stimulated Na+ uptake was investigated using [amiloride] 10–2–10–5 m. At 10–3 m amiloride concentration, Na+ uptake under pH gradient conditions was inhibited 80, 45, and 20% in BBMV's of adolescent, weanling and suckling rats, respectively. Kinetic studies revealed aK m for amiloride-sensitive Na+ uptake of 21.8±6.4, 24.9±10.9 and 11.8±4.17mm andV max of 8.76±1.21, 5.38±1.16 and 1.99±0.28 nmol/mg protein/5 sec in adolescent, weanling and suckling rats, respectively. The rate of pH dissipation, as determined by the fluorescence quenching of acridine orange, was similar across membrane preparation of all age groups studied. These findings suggest for the first time the presence of an ileal brush-border membrane Na+–H+ antiporter system in all ages studied. This system exhibits changes in regard to amiloride sensitivity and kinetic parameters.  相似文献   

20.
Summary We examined the direct effects of isoproterenol (ISO) andl-norepinephrine (NE) on electrolyte transport in isolated rabbit cortical collecting tubules (CCT) perfusedin vitro. The addition of either ISO (10–6 m) or NE (10–6 m) to the bath decreased transepithelial potential difference (PD), on average by 51 and 25%, respectively. These effects of ISO and NE were abolished by prior addition of the -adrenergic blocker,l-propranolol. ISO (10–5 m) had no effect from lumen. Also, osmotic water permeability was not influenced by ISO. Ouabain and ISO had additive effects on PD. Elimination of chloride from both perfusate and bath, or addition of acetazolamide, abolished the effect of ISO on PD. Although isotopic sodium flux from lumen to bath was not influenced by ISO, chemical net chloride absorption increased from 1.1±0.4 to 2.7±0.6 peq·cm–1·sec–1 (n=8,p<0.005). In conclusion, both ISO and NE are capable of decreasing PD in rabbit CCT perfusedin vitro. This effect is mediated by -adrenergic receptors and is accompanied by the increase in net chloride absorption. Although the mechanism responsible for this decrease in PD with ISO is unclear, active chloride absorption, active hydrogen secretion, or membrane chloride permeability changes may account for the effects of ISO.  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