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.
The nature of a transmembrane transport process depends largely on the identity of the reaction that is rate-limiting in the transport cycle. The one-for-one exchange of two chloride ions across the red cell membrane by band 3 can be decomposed into two component reactions: 1) the binding and dissociation of chloride at the transport site, and 2) the translocation of bound chloride across the membrane. The present work utilizes 35 Cl NMR and 37 Cl NMR to set lower limits on the rates of chloride binding and dissociation at the saturated inward- and outward-facing band 3 transport sites (greater than or equal to 10(5) events site-1 s-1 in all cases). At both 0-3 and 37 degrees C, the NMR data specify that chloride binding and dissociation at the saturated transport sites are not rate-limiting, indicating that translocation of bound chloride across the membrane is the slowest step in the overall transport cycle. Using these results, it is now possible to describe many features of the kinetic equation for the ping-pong transport cycle of band 3. This transport cycle can be decomposed into two half-reactions associated with the transport of two chloride ions in opposite directions across the membrane, where each half-reaction is composed of sequential binding, translocation, and dissociation events. One half-reaction contains the rate-limiting translocation event that controls the turnover of the transport cycle; in this half-reaction, translocation must be slower than binding and dissociation. The other half-reaction contains the non-rate-limiting translocation event that in principle could be faster than binding or dissociation. However, when the following sufficient (but not necessary) condition is satisfied, both translocation events are slower than binding and dissociation: if the non-rate-limiting translocation rate is within a factor of 10(2) (0-3 degrees C) or 2 (37 degrees C) of the overall turnover rate, then translocation is rate-limiting in each saturated half-reaction. Thus, even though chloride appears to migrate through a channel that leads from the transport site to solution, the results support a picture in which the binding, dissociation, and channel migration events are rapid compared to the translocation of bound chloride across the membrane. In this case, chloride binding to the transport site can be described by a simple dissociation constant (KD = kappa OFF/kappa ON) rather than by a Michaelis-Menten constant (KM = (kappa OFF + kappa TRANSLOCATION)/KAPPA ON).  相似文献   

2.
A theoretical analysis of the time course of a ternary cotransport system in membrane vesicles has been developed by extending previous work (Weiss, S.D. et al. (1981) J. Theor. Biol. 93, 597-608; Heinz, E. and Weinstein, M. (1984) Biochim. Biophys. Acta 776, 83-91). It has been assumed that the translocation of the carrier is the rate-limiting step of the transport process. Our approach includes, in particular, the presence of isotope tracer fluxes and the generalization to the case when many solutes share the same carrier. The situation when the tracer and the solute behave differently, as in the countertransport case, is stressed. Also, the interaction of two different solutes, internal and external to vesicles, is considered. Other points regard the analysis of the solute binding to the membrane vesicles, the influence of water permeability and the possible asymmetry of the transport system. In the Appendix, the assumption of no net translocation of all carrier species is discussed.  相似文献   

3.
The uptake of citrate by renal brush-border vesicles, prepared according to the method of Vannier, occurs by Na+-linked cotransport. It is 'positive rheogenic', i.e., stimulated by an (inside) negative, and inhibited by an (inside) positive electrical potential. The question arises whether, besides Na+, other ions (e.g., K+ and H+) participate in the cotransport. As to K+, neither an inward nor an outward directed K+ gradient has a significant effect on the citrate movement, but at equal concentrations of K+ inside and outside, equilibrium exchange of citrate, and to a smaller extent, the Na+-linked net uptake of citrate, are significantly stimulated. This observation is consistent with a hypothetical model in which K+ acts by accelerating both the empty and the fully loaded translocator. As to H+, citrate uptake is also stimulated by decreasing extravesicular pH, an effect previously attributed to protonization of the citrate anion in the assumption that the resulting secondary citrate anion is more acceptable to the translocator site. It was found, however, that the pH effect is still apparent if the concentration of the secondary citrate is kept constant by adjusting the total citrate concentration. This is taken as an argument against the above assumption and as being consistent with H+-linked cotransport. After the overshoot peak citrate exits slowly, and even after several hours does not attain equilibrium distribution, presumably owing to trapping by vesicular calcium.  相似文献   

4.
An inward-directed proton gradient energizes the transport of intact glycylsarcosine against a concentration gradient in rabbit renal brush-border membrane vesicles. Dissipation of the proton gradient abolishes the uphill transport. Generation of an inside-negative membrane potential nearly doubles the intravesicular concentration of the dipeptide at the peak of the overshoot without altering the equilibrium value. These data provide direct evidence for peptide-proton cotransport in the renal brush-border membrane.  相似文献   

5.
J E Lever 《Biochemistry》1984,23(20):4697-4702
Apical membrane vesicles isolated from a continuous renal cell line, LLC-PK1, catalyze electrogenic Na+-stimulated hexose transport and Na+-dependent binding of 3H-labeled 1-[2-(beta-D-glucopyranosyloxy)-4, 6-dihydroxyphenyl]-3-(4-hydroxyphenyl)-1-propanone [( 3H]phlorizin), a competitive ligand of this transport system. Phlorizin was not itself transported across the membrane and thus can serve as a probe of the binding step. The stoichiometry of Na+-dependent phlorizin binding in vesicles was 1:1, whereas Na+/hexose cotransport in vesicles exhibited a 2:1 stoichiometry. Na+ increased the affinity of phlorizin binding without affecting the total number of binding sites. An increased number of Na+-dependent phlorizin binding sites was observed under conditions of interior-negative membrane potential. These results are consistent with a model of the Na+/glucose cotransport cycle in which the unloaded transporter is negatively charged and its orientation influenced by membrane potential. Glucose and one sodium ion interact with the transporter, resulting in an uncharged complex. Binding of a second sodium ion triggers translocation of glucose and both sodium ions via formation of a loaded carrier complex bearing a single positive charge.  相似文献   

6.
The Na+/L-glutamate (L-aspartate) cotransport system present at the level of rat intestinal brush-border membrane vesicles is specifically activated by the ions K+ and Cl-. The presence of 100 mM K+ inside the vesicles drastically enhances the uptake rate and the transient intravesicular accumulation (overshoot) of the two acidic amino acids. It has been demonstrated that the activation of the transport system depended only in the intravesicular K+ concentration and that in the absence of any sodium gradient, an outward K+ gradient was unable to influence the Na+/acidic amino acid transport system. It was also found that Cl- could specifically activate the Na+-dependent L-glutamate (L-aspartate) uptake either in the presence or in the absence of K+. Also the effect of Cl- was observed only in the presence of an inward Na+ gradient and it was noted to be higher when chloride ion was present on both sides of the membrane vesicles. No influence (activation or accumulation) was observed in the absence of the Na+ gradient and in the presence of chloride gradient. L-Glutamate uptake measured in the presence of an imposed diffusion potential and in the presence of K+ or Cl- did not show any translocation of net charge.  相似文献   

7.
In the accompanying article (Schmidt, J.A., and Yguerabide, J. (1989) J. Biol. Chem. 264, 19790-19803), we presented a minimal quantitative kinetic model with one rate-limiting step for the transient response of rod outer segment (ROS) phosphodiesterase (PDE) to stimulating light pulses of low fractional bleach (linear response range) and showed that the model was in excellent quantitative agreement with experimental results. The model characterizes the PDE response in terms of the specific rate constant of the rate-limiting step, kL, the lifetime of photoactivated rhodopsin, tau R, and the lifetime of activated PDE, tau P, but makes no predictions on how these kinetic parameters should depend on the concentrations of the various reactive species involved in the PDE response to light and does not reveal the nature of the rate-limiting step. However, we established by curve fitting experimental data to theoretical expressions from the model that kL increases hyperbolically with [GTP], tau R decreases with [GTP], and tau P is independent of GTP. In this report we present three detailed kinetic models which make specific quantitative predictions on how the kinetic parameters of the minimal model should depend on nucleotide and G protein concentrations and test the models against experimental data. Each model consists of one rate-limiting step. The first detailed model postulates that the rate-limiting step is the dissociation of R*GT into R* and GT (T stands for GTP). The second model postulates that the rate-limiting step is the binding of GTP to R*G, and the third model postulates that the rate-limiting step is the encounter rate of R* and G on the ROS disc membrane. We find that only the first detailed model is consistent with the experimental results as characterized by the minimal model. Using this detailed model we (a) define kL and tau R in terms of more fundamental equilibrium and rate parameters, (b) develop a theory for the systematic evaluation of amplification or gain of the PDE light response from light-stimulated GTP-binding data as well as v(t) versus t graphs, and (c) clarify methods which have been used in the past to evaluate gain experimentally.  相似文献   

8.
Sodium-alanine cotransport was investigated in single isolated proximal tubule cells from rabbit kidney with the whole-cell current recording technique. Addition of L-alanine at the extracellular side induced an inward-directed sodium current and a cell depolarization. The sodium-alanine cotransport current was stereospecific and sodium dependent. Competition experiments suggested a common cotransport system for L-alanine and L-phenylalanine. Sodium-alanine cotransport current followed simple Michaelis-Menten kinetics, with an apparent Km of 6.6 mM alanine and 11.6 mM sodium and a maximal cotransport current of 0.98 pA/pF at -60 mV clamp potential. Hill plots of cotransport current suggested a potential-independent coupling ratio of one sodium and one alanine. The apparent Km for sodium and the maximal cotransport current were potential dependent, whereas the apparent Km for L-alanine was not affected by transmembrane potential. The increase in Km for alanine with decreasing inward-directed sodium gradients suggested a simultaneous transport mechanism. These results are consistent with a cotransport model with potential-dependent binding or unbinding of sodium (high-field access channel) and a potential-dependent translocation step.  相似文献   

9.
The uptake of citrate by renal brush-border vesicles, prepared according to the method of Vannier, occurs by Na+-linked cotransport. It is ‘positive rheogenic’, i.e., stimulated by an (inside) negative, and inhibited by an (inside) positive electrical potential. The question arises whether, besides Na+, other ions (e.g., K+ and H+) participate in the cotransport. As to K+, neither an inward nor an outward directed K+ gradient has a significant effect on the citrate movement, but at equal concentrations of K+ inside and outside, equilibrium exchange of citrate, and to a smaller extent, the Na+-linked net uptake of citrate, are significantly stimulated. This observation is consistent with a hypothetical model in which K+ acts by accelerating both the empty and the fully loaded translocator. As to H+, citrate uptake is also stimulated by decreasing extravesicular pH, an effect previously attributed to protonization of the citrate anion in the assumption that the resulting secondary citrate anion is more acceptable to the translocator site. It was found, however, that the pH effect is still apparent if the concentration of the secondary citrate is kept constant by adjusting the total citrate concentration. This is taken as an argument against the above assumption and as being consistent with H+-linked cotransport. After the overshoot peak citrate exits slowly, and even after several hours does not attain equilibrium distribution, presumably owing to trapping by vesicular calcium.  相似文献   

10.
In order to clarify contradictory reports on the electrogenicity of sodium/L-glutamate cotransport, this cotransport was studied using brush-border membrane vesicles isolated from rabbit renal cortex. Beforehand, the claim that the symport of L-glutamate with Na+ is linked to simultaneous antiport with K+ has been confirmed by the demonstration that equilibrium exchange of L-glutamate is inhibited by potassium. Concerning the electrogenicity of the system, the following results are reported: net uptake of sodium-dependent L-glutamate uptake was stimulated when the transmembranal electrical potential difference was increased by replacing a sodium sulfate gradient by a sodium nitrate gradient. At 100 mM Na+ the 'relative electrogenicity' of the initial uptake in the presence of intravesicular potassium was 2-times higher than in its absence. At a sodium concentration of 20 mM, when overall uptake was reduced, the relative electrogenicity in the presence of K+ was even 3-fold higher than in K+-free media. The relative electrogenicity of sodium/D-glucose cotransport measured under the same experimental conditions was not affected by K+. These results are discussed in terms of a model where the apparent electrogenicity of a cotransport system is dependent on the extent to which the charge translocating step is rate limiting ('rate limitancy'). It is proposed that potassium antiport, while decreasing charge stoichiometry of Na+/glutamate transport, increases the relative rate limitancy of the transport step translocating three cations (probably two Na+, one H+) together with one glutamate. Thereby the positive electrogenicity of glutamate uptake increases, in complete contrast to what would be expected from simple considerations of charge stoichiometry.  相似文献   

11.
We expressed the mouse gamma-aminobutyric acid (GABA) transporter GAT4 (homologous to rat/ human GAT-3) in Xenopus laevis oocytes and examined its functional and pharmacological properties by using electrophysiological and tracer uptake methods. In the coupled mode of transport (Na+/ Cl-/GABA cotransport), there was tight coupling between charge flux and GABA flux across the plasma membrane (2 charges/GABA). Transport was highly temperature-dependent with a temperature coefficient (Q10) of 4.3. The GAT4 turnover rate (1.5 s(-l); -50 mV, 21 degrees C) and temperature dependence suggest physiological turnover rates of 15-20 s(-1). No uncoupled current was observed in the presence of Na+. In the absence of external Na+, GAT4 exhibited two distinct uncoupled currents. (i) A Cl- leak current (ICl(leak)) was observed when Na+ was replaced with choline or tetraethylammonium. The reversal potential of (ICl(leak)) followed the Cl- Nernst potential. (ii) A Li+ leak current (ILi(leak)) was observed when Na+ was replaced with Li+. Both leak currents were inhibited by Na+, and both were temperature-independent (Q10 approximately 1). The two leak modes appeared not to coexist, as Li+ inhibited (ICl(leak)). The results suggest the existence of cation- and anion-selective channel-like pathways in GAT4. Flufenamic acid inhibited GAT4 Na+/Cl-/GABA cotransport, ILi(leak), and ICl(leak), (Ki approximately 30 microM), and the voltage-induced presteady-state charge movements (Ki approximately 440 microM). Flufenamic acid exhibited little or no selectivity for GAT1, GAT2, or GAT3. Sodium and GABA concentration jicroumps revealed that slow Na+ binding to the transporter is followed by rapid GABA-induced translocation of the ligands across the plasma membrane. Thus, Na+ binding and associated conformational changes constitute the rate-limiting steps in the transport cycle.  相似文献   

12.
How Drugs Interact with Transporters: SGLT1 as a Model   总被引:1,自引:0,他引:1  
Drugs are transported by cotransporters with widely different turnover rates. We have examined the underlying mechanism using, as a model system, glucose and indican (indoxyl-beta-D: -glucopyranoside) transport by human Na(+)/glucose cotransporter (hSGLT1). Indican is transported by hSGLT1 at 10% of the rate for glucose but with a fivefold higher apparent affinity. We expressed wild-type hSGLT1 and mutant G507C in Xenopus oocytes and used electrical and optical methods to measure the kinetics of glucose (using nonmetabolized glucose analogue alpha-methyl-D: -glucopyranoside, alphaMDG) and indican transport, alone and together. Indican behaved as a competitive inhibitor of alphaMDG transport. To examine protein conformations, we recorded SGLT1 capacitive currents (charge movements) and fluorescence changes in response to step jumps in membrane voltage, in the presence and absence of indican and/or alphaMDG. In the absence of sugar, voltage jumps elicited capacitive SGLT currents that decayed to steady state with time constants (tau) of 3-20 ms. These transient currents were abolished in saturating alphaMDG but only slightly reduced (10%) in saturating indican. SGLT1 G507C rhodamine fluorescence intensity increased with depolarizing and decreased with hyperpolarizing voltages. Maximal fluorescence increased approximately 150% in saturating indican but decreased approximately 50% in saturating alphaMDG. Modeling indicated that the rate-limiting step for indican transport is sugar translocation, whereas for alphaMDG it is dissociation of Na(+) from the internal binding sites. The inhibitory effects of indican on alphaMDG transport are due to its higher affinity and a 100-fold lower translocation rate. Our results indicate that competition between substrates and drugs should be taken into consideration when targeting transporters as drug delivery systems.  相似文献   

13.
A preparation of purified erythrocyte membrane ATPase whose activation by Ca2+ is or is not dependent on calmodulin depending on the enzyme dilution was used in the low dilution state for these studies. In appropriate conditions, the purified ATPase in the absence of calmodulin exhibited a Ca2+ concentration dependence identical to that of the native enzyme in the erythrocyte membrane ghost in the presence of calmodulin. Accordingly, an apparent Kd approximately equal to 1 X 10(-7) M was derived for cooperative calcium binding to the activating and transport sites of the nonphosphorylated enzyme. The kinetics of enzyme phosphorylation in the transient state following addition of ATP to enzyme activated with calcium were then resolved by rapid kinetic methods, demonstrating directly that phosphoenzyme formation precedes Pi production, consistent with the phosphoenzyme role as an intermediate in the catalytic cycle. Titration of a low affinity site (Kd approximately equal to 2 X 10(-3) M) with calcium produced inhibition of phosphoenzyme cleavage and favored reversal of the catalytic cycle, indicating that calcium dissociation from the transport sites precedes hydrolytic cleavage of the phosphoenzyme. The two different calcium dissociation constants of the nonphosphorylated and phosphorylated enzyme demonstrate that a phosphorylation-induced reduction of calcium affinity is the basic coupling mechanism of catalysis and active transport, with an energy expenditure of approximately 6 kcal/mol of calcium in standard conditions. From the kinetic point of view, a rate-limiting step is identified with the slow dissociation of calcium from the phosphoenzyme; another relatively slow step following hydrolytic cleavage and preceding recycling of the enzyme is suggested by the occurrence of a presteady state phosphoenzyme overshoot.  相似文献   

14.
15.
After termination of protein synthesis in bacteria, ribosomes are recycled from posttermination complexes by the combined action of elongation factor G (EF-G), ribosome recycling factor (RRF), and initiation factor 3 (IF3). The functions of the factors and the sequence in which ribosomal subunits, tRNA, and mRNA are released from posttermination complexes are unclear and, in part, controversial. Here, we study the reaction by rapid kinetics monitoring fluorescence. We show that RRF and EF-G with GTP, but not with GDPNP, promote the dissociation of 50S subunits from the posttermination complex without involving translocation or a translocation-like event. IF3 does not affect subunit dissociation but prevents reassociation, thereby masking the dissociating effect of EF-G-RRF under certain experimental conditions. IF3 is required for the subsequent ejection of tRNA and mRNA from the small subunit. The latter step is slower than subunit dissociation and constitutes the rate-limiting step of ribosome recycling.  相似文献   

16.
The decay of delta pH across vesicular membranes by nigericin-mediated H+ and metal ion (M+) transports has been studied at 25 degrees C after creating delta pH by temperature jump (T-jump). In these experiments K+ or Na+ were chosen as M+ for the compensating flux. Theoretical expressions derived to analyse these data suggest a method for estimating the intrinsic rate constants for the translocation of nig-H (k1) and for the translocation of nig-M (k2) across membrane, from the pH dependence of the delta pH decay. The following could be inferred from the analysis of data. (a) At pH approximately 7.5 and 250 mM ion concentrations, nigericin-mediated H+ and M+ transport rates are lower in a medium of K+ than in a medium of Na+, although ionophore selectivity of nigericin towards K+ is 25-45-times higher than that towards Na+. However, at lower [M+] (approximately 50 mM) the transport rates are higher in a medium of K+ than in a medium of Na+. Such behaviours can be understood with the help of parameters determined in this work. (b) The intrinsic rate constants k1 and k2 associated with the translocations of nig-H and nig-K or nig-Na across membrane are similar in magnitude. (c) At pH approximately 7.5 translocation of nig-H is the dominant rate-limiting step in a medium containing K+. In contrast with this, at this pH, translocation of nig-M is the dominant rate-limiting step when metal ion is Na+. (d)k1 approximately k2 approximately 6.10(3) s-1 could be estimated at 25 degrees C in vesicles prepared from soyabean phospholipid, and lipid mixtures of 80% phosphatidylcholine (PC) + 20% phosphatidylethanolamine and 92% PC + 8% phosphatidic acid. (e) The apparent dissociation constants of nig-M in vesicles were estimated to be approximately 1.5.10(-3) M for K+ and 6.4.10(-2) M for Na+ (at 50 mM ion concentrations) using approximately 10(-8.45) M for the apparent dissociation constant of nig-H.  相似文献   

17.
We have developed a mathematical model of adenine nucleotide translocase (ANT) function on the basis of the structural and kinetic properties of the transporter. The model takes into account the effect of membrane potential, pH, and magnesium concentration on ATP and ADP exchange velocity. The parameters of the model have been estimated from experimental data. A satisfactory model should take into account the influence of the electric potential difference on both ternary complex formation and translocation processes. To describe the dependence of translocation constants on electric potential we have supposed that ANT molecules carry charged groups. These groups are shifted during the translocation. Using the model we have evaluated the translocator efficiency and predicted the behavior of ANT under physiological conditions.  相似文献   

18.
The thylakoid membrane, located inside the chloroplast, requires proteins transported across it for plastid biogenesis and functional photosynthetic electron transport. The chloroplast Tat translocator found on thylakoids transports proteins from the plastid stroma to the thylakoid lumen. Previous studies have shown that the chloroplast Tat pathway is independent of NTP hydrolysis as an energy source and instead depends on the thylakoid transmembrane proton gradient to power protein translocation. Because of its localization on the same membrane as the proton motive force-dependent F(0)F(1) ATPase, we believed that the chloroplast Tat pathway also made use of the thylakoid electric potential for transporting substrates. By adjusting the rate of photosynthetic proton pumping and by utilizing ionophores, we show that the chloroplast Tat pathway can also utilize the transmembrane electric potential for protein transport. Our findings indicate that the chloroplast Tat pathway is likely dependent on the total protonmotive force (PMF) as an energy source. As a protonmotive-dependent device, certain predictions can be made about structural features expected to be found in the Tat translocon, specifically, the presence of a proton well, a device in the membrane that converts electrical potential into chemical potential.  相似文献   

19.
Hass MA  Christensen HE  Zhang J  Led JJ 《Biochemistry》2007,46(50):14619-14628
Exchange on the microsecond time scale between the protonated and deprotonated forms of His92 in the copper site of reduced plastocyanin from the cyanobacteria Anabaena variabilis was monitored using 15N NMR relaxation measurements. On the basis of the dependence of the kinetics on pH and phosphate buffer concentration, we propose a two-step model for the protonation of the copper site in agreement with previous crystallographic studies. It is shown that the proton transfer is the rate-limiting step in the reaction at low buffer concentrations, whereas at high buffer concentrations, another step becomes rate-limiting. We suggest that the latter step is a concerted dissociation of His92 from the Cu(I) ion and a 180 degrees rotation of the imidazole ring, which precede the protonation. The first-order rate constant for the dissociation of His92 from the Cu(I) ion is estimated to be 2.4 x 10(4) s(-1). Also, a cooperative effect of the protonation of the remote His61 on the protonation of His92 and the redox properties of the protein was investigated by substituting His61 with asparagine. The mutation causes a modest change in both the pKa value of His92 and the redox potential of the protein.  相似文献   

20.
After a brief review of Na+-cotransport systems which also accept other ions as co-ions or modifiers, modulation of the Na+-L-glutamate transport system in rabbit renal brush border membranes by K+ and H+ is discussed in more detail. Intravesicular K+ increases the initial uptake rate and electrogenicity of the cotransport. This effect of K+ is attributed to the formation of a K+-carrier complex that moves much more rapidly than do the other complexes. The resulting shift in rate limitancy (relative increase in overall rate over the relative increase in rate of step under consideration) from an electroneutral towards a charge-translocating pathway unmasks the electrogenicity of the initial L-glutamate uptake. A positive correlation between relative rate limitancy of the electrogenic pathway and electrogenicity is demonstrated supporting this model. Protons, in addition to acting as co-ions, modify Na+-glutamate cotransport by increasing both the initial rate and the electrogenicity of uptake. This phenomenon is assumed to represent a transition of the transport system from a carrier-like to an open channel-like translocation mode. Thus, the intrinsic properties of Na+-cotransport systems may vary under the influence of other ions. This holds true in particular for the electrogenicity of the initial transport rate which may change independently of alterations in charge stoichiometry.  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