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 109 毫秒
1.
The MelB permease of Salmonella typhimurium (MelB-ST) catalyzes the coupled symport of melibiose and Na(+), Li(+), or H(+). In right-side-out membrane vesicles, melibiose efflux is inhibited by an inwardly directed gradient of Na(+) or Li(+) and stimulated by equimolar concentrations of internal and external Na(+) or Li(+). Melibiose exchange is faster than efflux in the presence of H(+) or Na(+) and stimulated by an inwardly directed Na(+) gradient. Thus, sugar is released from MelB-ST externally prior to the release of cation in agreement with current models proposed for MelB of Escherichia coli (MelB-EC) and LacY. Although Li(+) stimulates efflux, and an outwardly directed Li(+) gradient increases exchange, it is striking that internal and external Li(+) with no gradient inhibits exchange. Furthermore, Trp → dansyl FRET measurements with a fluorescent sugar (2'-(N-dansyl)aminoalkyl-1-thio-β-D-galactopyranoside) demonstrate that MelB-ST, in the presence of Na(+) or Li(+), exhibits (app)K(d) values of ~1 mM for melibiose. Na(+) and Li(+) compete for a common binding pocket with activation constants for FRET of ~1 mM, whereas Rb(+) or Cs(+) exhibits little or no effect. Taken together, the findings indicate that MelB-ST utilizes H(+) in addition to Na(+) and Li(+). FRET studies also show symmetrical emission maximum at ~500 nm with MelB-ST in the presence of 2'-(N-dansyl)aminoalkyl-1-thio-β-D-galactopyranoside and Na(+), Li(+), or H(+), which implies a relatively homogeneous distribution of conformers of MelB-ST ternary complexes in the membrane.  相似文献   

2.
Fourier transform infrared difference spectroscopy has been used to obtain information about substrate-induced structural changes of the melibiose permease (MelB) from Escherichia coli reconstituted into liposomes. Binding of the cosubstrate Na(+) gives rise to several peaks in the amide I and II regions of the difference spectrum Na(+).MelB minus H(+).MelB, that denote the presence of conformational changes in all types of secondary structures (alpha-helices, beta-sheets, loops). In addition, peaks around 1400 and at 1740-1720 cm(-1) are indicative of changes in protonation/deprotonation or in environment of carboxylic groups. Binding of the cosubstrate Li(+) produces a difference spectrum that is also indicative of conformational changes, but that is at variance as compared to that induced by Na(+) binding. To analyze the following transport steps, the melibiose permease with either H(+), Na(+), or Li(+) bound was incubated with melibiose. The difference spectra obtained by subtracting the spectrum cation.MelB from the respective complex cation.melibiose.MelB were roughly similar among them, but different from those induced by cation binding, and more intense. Therefore, major conformational changes that are induced during melibiose binding/substrate translocation, like those denoted by intense peaks at 1668 and 1645 cm(-)(1), are similar for the three cotransporting cations. Changes in the protonation state and/or in the environment of given carboxylic residues were also induced by melibiose-MelB interaction in the presence of cations.  相似文献   

3.
Wild-type Citrobacter freundii cannot grow on melibiose as a sole source of carbon. The melibiose transporter gene melB was cloned from a C. freundii mutant M4 that could utilize melibiose as a sole carbon source. Although the cloned melB gene is closely similar to the melB genes of other bacteria, it is cryptic because of a frameshift mutation. Site-directed mutagenesis was used to construct a functional melB gene by deleting one nucleotide, resulting in the production of an active melibiose transporter. The active MelB transporter could utilize Na(+) and H(+) as coupling cations to melibiose transport. The amino acid sequence of the C. freundii MelB was found to be most similar to those of Salmonella typhimurium and Escherichia coli MelB. These facts are consistent with the phylogenetic relationship of bacteria and the cation coupling properties of the melibiose transporters.  相似文献   

4.
Gly263 of the rat kidney Na(+),K(+)-ATPase is highly conserved within the family of P-type ATPases. Mutants in which Gly263 or the juxtaposed Arg264 had been replaced by alanine were expressed at high levels in COS-1 cells and characterized functionally. Titrations of Na(+),K(+), ATP, and vanadate dependencies of Na(+),K(+)-ATPase activity showed changes in the apparent affinities relative to wild-type compatible with a displacement of the E(1)-E(2) conformational equilibrium in favor of E(1). The level of the K(+)-occluded form was reduced in the Gly263-->Ala and Arg264-->Ala mutants, and the rate constant characterizing deocclusion of K(+) or Rb(+) was increased as much as 20-fold in the Gly263-->Ala mutant. Studies of the sensitivity of the phosphoenzyme to K(+) and ADP showed a displacement of the E(1)P-E(2)P equilibrium of the phosphoenzyme in favor of E(1)P, and dephosphorylation experiments carried out at 25 degrees C on a millisecond time scale using a quenched-flow technique demonstrated a reduction of the E(1)P to E(2)P conversion rate in the mutants. Hence, the mutations displaced the conformational equilibria of dephosphoenzyme and phosphoenzyme in parallel in favor of the E(1) and E(1)P forms. The observed effects were more pronounced in the Gly263-->Ala mutant compared with the Arg264-->Ala mutant. Leu332 mutations that likewise displaced the conformational equilibria in favor of E(1) and E(1)P were also studied. Unlike the Gly263-->Ala mutant the Leu332 mutants displayed a wild-type like rate of K(+) deocclusion. Thus, the effect of the Gly263 mutation on the E(1)-E(2) conformational equilibrium seems to be caused mainly by an acceleration of the K(+)-deoccluding step, whereas in the Leu332 mutants the rate of the reverse reaction seems to be reduced.  相似文献   

5.
Inosine 5'-monophosphate dehydrogenase (IMPDH) catalyzes the oxidation of inosine 5'-monophosphate (IMP) to xanthosine 5'-monophosphate with the concomitant reduction of NAD to NADH. Escherichia coli IMPDH is activated by K(+), Rb(+), NH(+)(4), and Cs(+). K(+) activation is inhibited by Li(+), Na(+), Ca(2+), and Mg(2+). This inhibition is competitive versus K(+) at high K(+) concentrations, noncompetitive versus IMP, and competitive versus NAD. Thus monovalent cation activation is linked to the NAD site. K(+) increases the rate constant for the pre-steady-state burst of NADH production, possibly by increasing the affinity of NAD. Three mutant IMPDHs have been identified which increase the value of K(m) for K(+): Asp13Ala, Asp50Ala, and Glu469Ala. In contrast to wild type, both Asp13Ala and Glu469Ala are activated by all cations tested. Thus these mutations eliminate cation selectivity. Both Asp13 and Glu469 appear to interact with the K(+) binding site identified in Chinese hamster IMPDH. Like wild-type IMPDH, K(+) activation of Asp50Ala is inhibited by Li(+), Na(+), Ca(2+), and Mg(2+). However, this inhibition is noncompetitive with respect to K(+) and competitive with respect to both IMP and NAD. Asp50 interacts with residues that form a rigid wall in the IMP site; disruption of this wall would be expected to decrease IMP binding, and the defect could propagate to the proposed K(+) site. Alternatively, this mutation could uncover a second monovalent cation binding site.  相似文献   

6.
C Ganea  T Pourcher  G Leblanc  K Fendler 《Biochemistry》2001,40(45):13744-13752
Electrogenic activity associated with the activity of the melibiose permease (MelB) of Escherichia coli was investigated by using proteoliposomes containing purified MelB adsorbed onto a solid-supported membrane. Transient currents were selectively recorded by applying concentration jumps of Na+ ions (or Li+) and/or of different sugar substrates of MelB (melibiose, thio-methyl galactoside, raffinose) using a fast-flow solution exchange system. Characteristically, the transient current response was fast, including a single decay exponential component (tau approximately 15 ms) on applying a Na+ (or Li+) concentration jump in the absence of sugar. On imposing a Na+ (or Li+) jump on proteoliposomes preincubated with the sugar, a sugar jump in a preparation preincubated with the cation, or a simultaneous jump of the cation and sugar substrates, the electrical transients were biphasic and comprised both the fast and an additional slow (tau approximately 350 ms) decay components. Finally, selective inactivation of the cosubstrate translocation step by acylation of MelB cysteins with N-ethyl maleimide suppressed the slow response components and had no effect on the fast transient one. We suggest that the fast transient response reflects charge transfer within MelB during cosubstrate binding while the slow component is associated with charge transfer across the proteoliposome membrane. From the time course of the transient currents, we estimate a rate constant for Na+ binding in the absence and presence of melibiose of k > 50 s(-1) and one for melibiose binding in the absence of Na+ of k approximately 10 s(-1).  相似文献   

7.
Residue Asp-189 plays an important dual role in thrombin: it defines the primary specificity for Arg side chains and participates indirectly in the coordination of Na(+). The former role is shared by other proteases with trypsin-like specificity, whereas the latter is unique to Na(+)-activated proteases in blood coagulation and the complement system. Replacement of Asp-189 with Ala, Asn, Glu, and Ser drastically reduces the specificity toward substrates carrying Arg or Lys at P1, whereas it has little or no effect toward the hydrolysis of substrates carrying Phe at P1. These findings confirm the important role of Asp-189 in substrate recognition by trypsin-like proteases. The substitutions also affect significantly and unexpectedly the monovalent cation specificity of the enzyme. The Ala and Asn mutations abrogate monovalent cation binding, whereas the Ser and Glu mutations change the monovalent cation preference from Na(+) to the smaller cation Li(+) or to the larger cation Rb(+), respectively. The observation that a single amino acid substitution can alter the monovalent cation specificity of thrombin from Na(+) (Asp-189) to Li(+) (Ser-189) or Rb(+) (Glu-189) is unprecedented in the realm of monovalent cation-activated enzymes.  相似文献   

8.
Glu(282) located in the NH(2)-terminal part of transmembrane helix M3 of the Na(+),K(+)-ATPase was replaced by alanine, glycine, leucine, lysine, aspartate, or glutamine, and the effects of the mutations on the overall and partial reactions of the enzyme were analyzed. The mutations affected at least 3 important functions of the Na(+),K(+)-ATPase: (i) the conformational transitions between E(1) and E(2) forms of dephospho- and phosphoenzyme, (ii) Na(+) binding at the cytoplasmically facing sites of E(1), and (iii) long-range interaction controlling dephosphorylation. In mutants Glu(282) --> Lys and Glu(282) --> Asp, the E(1) form was favored during ATP hydrolysis, whereas the E(2) form was favored in Glu(282) --> Ala and Glu(282) --> Gly. Regardless of the change of conformational equilibrium, all the mutants displayed a reduced apparent affinity for Na(+), at least 3-fold for Glu(282) --> Lys and Glu(282) --> Asp, suggesting a direct effect on the Na(+) binding properties of E(1). Glu(282) --> Ala and Glu(282) --> Gly exhibited an extraordinary high rate of ATP hydrolysis in the mere presence of Na(+) without K(+) ("Na(+)-ATPase activity"), because of an increased rate of dephosphorylation of E(2)P. These results are in accordance with the hypothesis that Glu(282) is involved in the communication between the cation binding pocket and the catalytic site and in control of the cytoplasmic entry pathway for Na(+).  相似文献   

9.
The melibiose permease of Salmonella enterica serovar Typhimurium (MelBSt) catalyzes symport of melibiose with Na+, Li+, or H+. Bioinformatics and mutational analyses indicate that a conserved Gly117 (helix IV) is a component of the Na+-binding site. In this study, Gly117 was mutated to Ser, Asn, or Cys. All three mutations increase the maximum rate (Vmax) for melibiose transport in Escherichia coli DW2 and greatly decrease Na+ affinity, indicating that intracellular release of Na+ is facilitated. Rapid melibiose transport, particularly by the G117N mutant, triggers osmotic lysis in the lag phase of growth. The findings support the previous conclusion that Gly117 plays an important role in cation binding and translocation. Furthermore, a spontaneous second-site mutation (P148L between loop4-5 and helix V) in the G117C mutant prevents cell lysis. This mutation significantly decreases Vmax with little effect on cosubstrate binding in G117C, G117S, and G117N mutants. Thus, the P148L mutation specifically inhibits transport velocity and thereby blocks the lethal effect of elevated melibiose transport in the Gly117 mutants.  相似文献   

10.
The melB gene coding for the melibiose carrier of Klebsiella pneumoniae was cloned and sequenced. There were two potential translation initiation sites. It was predicted that the melibiose carrier consists of 471 (or 467) amino acid residues. Seventy-eight percent of the 471 amino acids were identical to the Escherichia coli melibiose carrier. Sugar transport characteristics were studied using an E. coli mel- mutant expressing cloned K. pneumoniae melB gene. Accumulation of melibiose via the K. pneumoniae melibiose carrier was not stimulated by adding NaCl or LiCl which stimulates melibiose accumulation via the E. coli melibiose carrier. Lactose was accumulated only in the presence of LiCl. TMG (methyl-1-thio-beta-D-galactopyranoside) was accumulated in the absence of added NaCl or LiCl. The accumulation was stimulated by LiCl but not by NaCl. Rapid H+ uptake was observed when melibiose or TMG was added to cell suspensions. These results suggest that the preferred cation couplings via K. pneumoniae melibiose carrier are H(+)-melibiose, Li(+)-lactose, and H+/Li(+)-TMG. This coupling spectrum is quite different from that of the E. coli melibiose carrier. It is of special interest that the K. pneumoniae melibiose carrier seems to be lacking the ability to recognize Na+ which is a preferred coupling cation of the E. coli melibiose carrier for all known sugar substrates. Further investigation of these two carriers may give us insight into the Na+ recognition site.  相似文献   

11.
Li(+) interacts with the Na(+)/Cl(-)-dependent GABA transporter, GAT1, under two conditions: in the absence of Na(+) it induces a voltage-dependent leak current; in the presence of Na(+) and GABA, Li(+) stimulates GABA-induced steady-state currents. The amino acids directly involved in the interaction with the Na(+) and Li(+) ions at the so-called "Na2" binding site have been identified, but how Li(+) affects the kinetics of GABA cotransport has not been fully explored. We expressed GAT1 in Xenopus oocytes and applied the two-electrode voltage clamp and (22)Na uptake assays to determine coupling ratios and steady-state and presteady-state kinetics under experimental conditions in which extracellular Na(+) was partially substituted by Li(+). Three novel findings are: 1) Li(+) reduced the coupling ratio between Na(+) and net charge translocated during GABA cotransport; 2) Li(+) increased the apparent Na(+) affinity without changing its voltage dependence; 3) Li(+) altered the voltage dependence of presteady-state relaxations in the absence of GABA. We propose an ordered binding scheme for cotransport in which either a Na(+) or Li(+) ion can bind at the putative first cation binding site (Na2). This is followed by the cooperative binding of the second Na(+) ion at the second cation binding site (Na1) and then binding of GABA. With Li(+) bound to Na2, the second Na(+) ion binds more readily GAT1, and despite a lower apparent GABA affinity, the translocation rate of the fully loaded carrier is not reduced. Numerical simulations using a nonrapid equilibrium model fully recapitulated our experimental findings.  相似文献   

12.
A family of aryl isothiouronium derivatives was designed as probes for cation binding sites of Na(+),K(+)-ATPase. Previous work showed that 1-bromo-2,4,6-tris(methylisothiouronium)benzene (Br-TITU) acts as a competitive blocker of Na(+) or K(+) occlusion. In addition to a high-affinity cytoplasmic site (K(D) < 1 microM), a low-affinity site (K(D) approximately 10 microM) was detected, presumably extracellular. Here we describe properties of Br-TITU as a blocker at the extracellular surface. In human red blood cells Br-TITU inhibits ouabain-sensitive Na(+) transport (K(D) approximately 30 microM) in a manner antagonistic with respect to extracellular Na(+). In addition, Br-TITU impairs K(+)-stimulated dephosphorylation and Rb(+) occlusion from phosphorylated enzyme of renal Na(+),K(+)-ATPase, consistent with binding to an extracellular site. Incubation of renal Na(+),K(+)-ATPase with Br-TITU at pH 9 irreversibly inactivates Na(+),K(+)-ATPase activity and Rb(+) occlusion. Rb(+) or Na(+) ions protect. Preincubation of Br-TITU with red cells in a K(+)-free medium at pH 9 irreversibly inactivates ouabain-sensitive (22)Na(+) efflux, showing that inactivation occurs at an extracellular site. K(+), Cs(+), and Li(+) ions protect against this effect, but the apparent affinity for K(+), Cs(+), or Li(+) is similar (K(D) approximately 5 mM) despite their different affinities for external activation of the Na(+) pump. Br-TITU quenches tryptophan fluorescence of renal Na(+),K(+)-ATPase or of digested "19 kDa membranes". After incubation at pH 9 irreversible loss of tryptophan fluorescence is observed and Rb(+) or Na(+) ions protect. The Br-TITU appears to interact strongly with tryptophan residue(s) within the lipid or at the extracellular membrane-water interface and interfere with cation occlusion and Na(+),K(+)-ATPase activity.  相似文献   

13.
Previous studies have suggested that the conformation of the activation peptide of protein C is influenced by the binding of Ca(2+). To provide direct evidence for the linkage between Ca(2+) binding and the conformation of the activation peptide, we have constructed a protein C mutant in the gamma-carboxyglutamic acid-domainless form in which the P1 Arg(169) of the activation peptide is replaced with the fluorescence reporter Trp. Upon binding of Ca(2+), the intrinsic fluorescence of the mutant decreases approximately 30%, as opposed to only 5% for the wild-type, indicating that Trp(169) is directly influenced by the divalent cation. The K(d) of Ca(2+) binding for the mutant protein C was impaired approximately 4-fold compared with wild-type. Interestingly, the conformation of the activation peptide was also found to be sensitive to the binding of Na(+), and the affinity for Na(+) binding increased approximately 5-fold in the presence of Ca(2+). These findings suggest that Ca(2+) changes the conformation of the activation peptide of protein C and that protein C is also capable of binding Na(+), although with a weaker affinity compared with the mature protease. The mutant protein C can no longer be activated by thrombin but remarkably it can be activated efficiently by chymotrypsin and by the thrombin mutant D189S. Activation of the mutant protein C by chymotrypsin proceeds at a rate comparable to the activation of wild-type protein C by the thrombin-thrombomodulin complex.  相似文献   

14.
M Quick  S St?lting  H Jung 《Biochemistry》1999,38(41):13523-13529
The Na+/proline transporter of Escherichia coli (PutP) is a member of a large family of Na+/solute symporters. To investigate the role of Arg residues which are conserved within this family, Arg40 at the cytoplasmic end of transmembrane domain (TM) II and Arg117 in cytoplasmic loop 4 of PutP are subjected to amino acid substitution analysis. Removal of the positive charge at position 40 (PutP-R40C, Q, E) leads to a dramatic decrease of the V(max) of Na(+)-coupled proline uptake (1-10% of PutP-wild-type). The reduced transport rates are accompanied by decreased apparent affinities of the transporter for Na+ and Li+ while the apparent affinity for proline is only slightly altered. Furthermore, single Cys PutP-R40C reacts with N-ethylmaleimide (NEM), and this reaction is partially inhibited by proline and more efficiently by Na+ ions. Remarkably, NEM modification of Cys40 inhibits Na(+)-driven proline uptake almost completely while facilitated influx of proline into deenergized cells is stimulated by this reaction, suggesting an at least partially uncoupled phenotype under these conditions. These results suggest that Arg40 is located close to the site of ion binding and is important for the coupling of ion and proline transport. The observations confirm the functional importance of TM II described in earlier studies [M. Quick and H. Jung (1997) Biochemistry 36, 4631-4636]. In contrast to Arg40, Arg117 is apparently not important for function of the mature protein. The low transport rates observed upon substitution of Arg117 (PutP-R117C, K, Q) can at least partially be attributed to reduced amounts of PutP in the membrane. However, once inserted into the membrane, PutP containing Arg117 replacements shows a stability comparable to the wild-type as indicated by pulse-chase experiments. These observations suggest that Arg117 plays a crucial role at a stage prior to complete functional insertion of PutP into the membrane, i. e., by stabilizing a folding intermediate.  相似文献   

15.
The accessibility of Escherichia coli melibiose permease to aqueous solvent was studied following hydrogen-deuterium exchange kinetics monitored by attenuated total reflection-Fourier transform infrared spectroscopy under four distinct conditions where MelB forms different complexes with its substrates (H(+), Na(+), melibiose). Analysis of the amide II band upon (2)H(2)O exposure discloses a significant sugar protection of the protein against aqueous solvent, resulting in an 8% less exchange of the corresponding H(+)*melibiose*MelB complex compared with the protein in the absence of sugar. Investigation of the amide I exchange reveals clear substrate effects on beta-sheet accessibility, with the complex H(+)*melibiose*MelB being the most protected state against exchange, followed by Na(+)*melibiose*MelB. Although of smaller magnitude, similar changes in alpha-helices plus non-ordered structures are detected. Finally, no differences are observed when analyzing reverse turn structures. The results suggest that sugar binding induces a remarkable compactness of the carrier's structure, affecting mainly beta-sheet domains of the transporter, which, according to secondary structure predictions, may include cytoplasmic loops 4-5 and 10-11. A possible catalytic role of these two loops in the functioning of MelB is hypothesized.  相似文献   

16.
Arg-52 of the Escherichia coli melibiose carrier was replaced by Ser (R52S), Gln (R52Q), or Val (R52V). While the level of carrier in the membrane for each mutant remained similar to that for the wild type, analysis of melibiose transport showed an uncoupling of proton cotransport and a drastic reduction in Na(+)-coupled transport. Second-site revertants were selected on MacConkey plates containing melibiose, and substitutions were found at nine distinct locations in the carrier. Eight revertant substitutions were isolated from the R52S strain: Asp-19-->Gly, Asp-55-->Asn, Pro-60-->Gln, Trp-116-->Arg, Asn-244-->Ser, Ser-247-->Arg, Asn-248-->Lys, and Ile-352-->Val. Two revertants were also isolated from the R52V strain: Trp-116-->Arg and Thr-338-->Arg revertants. The R52Q strain yielded an Asp-55-->Asn substitution and a first-site revertant, Lys-52 (R52K). The R52K strain had transport properties similar to those of the wild type. Analysis of melibiose accumulation showed that proton-driven accumulation was still defective in the second-site revertant strains, and only the Trp-116-->Arg, Ser-247-->Arg, and Asn-248-->Lys revertants regained significant Na(+)-coupled accumulation. In general, downhill melibiose transport in the presence of Na(+) was better in the revertant strains than in the parental mutants. Three revertant strains, Asp-19-->Gly, Asp-55-->Asn, and Thr-338-->Arg strains, required a high Na(+) concentration (100 mM) for maximal activity. Kinetic measurements showed that the N248K and W116R revertants lowered the K(m) for melibiose, while other revertants restored transport velocity. We suggest that the insertion of positive charges on membrane helices is compensating for the loss of Arg-52 and that helix II is close to helix IV and VII. We also suggest that Arg-52 is salt bridged to Asp-55 (helix II) and Asp-19 (helix I).  相似文献   

17.
Electrogenic events associated with the activity of the melibiose permease (MelB), a transporter from Escherichia coli, were investigated. Proteoliposomes containing purified MelB were adsorbed to a solid supported lipid membrane, activated by a substrate concentration jump, and transient currents were measured. When the transporter was preincubated with Na(+) at saturating concentrations, a charge translocation in the protein upon melibiose binding could still be observed. This result demonstrates that binding of the uncharged substrate melibiose triggers a charge displacement in the protein. Further analysis showed that the charge displacement is neither related to extra Na(+) binding to the transporter, nor to the displacement of already bound Na(+) within the transporter. The electrogenic melibiose binding process is explained by a conformational change with concomitant displacement of charged amino acid side chains and/or a reorientation of helix dipoles. A kinetic model is suggested, in which Na(+) and melibiose binding are distinct electrogenic processes associated with approximately the same charge displacement. These binding reactions are fast in the presence of the respective cosubstrate (k > 50 s(-1)).  相似文献   

18.
We detected inducible melibiose transport activity in cells of Enterobacter cloacae IID977. H+, but not Na+, was found to be the coupling cation for this transporter. We cloned and sequenced the gene encoding the melibiose transporter. A homology search of a protein sequence database revealed that this melibiose transporter has high sequence similarity with the lactose transporter (LacY) and the raffinose transporter (RafB) and has some similarity with the melibiose transporter (MelB) of Escherichia coli.  相似文献   

19.
Thr(214) of the highly conserved (214)TGES sequence in domain A of the Na(+),K(+)-ATPase was replaced with alanine, and the mutant was compared functionally with the previously characterized domain A mutant Gly(263) --> Ala. Thr(214) --> Ala displayed a conspicuous 150-fold reduction of the apparent vanadate affinity for inhibition of ATPase activity, which could not simply be explained by the observed shifts of the conformational equilibria in favor of E(1) and E(1)P. The intrinsic vanadate affinity of the E(2) form and the effect on the apparent vanadate affinity of displacement of the E(1)-E(2) equilibrium were determined in a phosphorylation assay that allows the enzyme-vanadate complex to be formed under equilibrium conditions. When the E(2) form prevailed, Thr(214) --> Ala retained a reduced vanadate affinity relative to wild type, whereas the affinity of Gly(263) --> Ala became wild type-like. Thus, mutation of Thr(214) affected the intrinsic affinity of E(2) for vanadate. Furthermore, Thr(214) --> Ala showed at least a 5-fold reduced E(2)P dephosphorylation rate relative to wild type in the presence of saturating concentrations of K(+) and Mg(2+). Because vanadate is a phosphoryl transition state analog, it is proposed that defective binding of the phosphoryl transition state complex (transition state destabilization) causes the inability to catalyze E(2)P dephosphorylation properly. By contrast, the phosphorylation site in the E(1) form was unaffected in Thr(214) --> Ala. Replacement of the glutamate, Glu(216), of (214)TGES with alanine was incompatible with cell viability, indicating a very low transport activity or expression level. Our results support the hypothesis that domain A is isolated in the E(1) form, but contributes to make up the catalytic site in the E(2) and E(2)P conformations.  相似文献   

20.
Fedosova NU  Esmann M 《Biochemistry》2004,43(14):4212-4218
Correlation between the Na,K-ATPase affinity to ADP and the cation (its nature and concentration) present in the medium was investigated. In buffer with low ionic strength (I approximately 1 mM) high-affinity ADP binding was not observed, while a stepwise increase in the concentrations of added cation (Na(+), Tris(+), imidazole(+), N-methylglucamine(+), choline(+)) induced an increase in the ADP affinity. The effect was fully saturated at 30-50 mM for all of the cations tested. The maximal affinity for ADP was slightly higher in the presence of Na(+), Tris(+), or imidazole(+) than in the presence of N-methylglucamine(+) or choline(+) (equilibrium dissociation constant K(d) 0.2-0.3 vs 0.7 microM). The ADP dissociation rates from its complex with enzyme in the presence of Na(+) or Tris(+) were similar, implying identity of the nucleotide-binding enzyme conformations, which therefore are assigned to E(1). The ability to compete with K(+) clearly distinguished Na(+) from other cations, which speaks against the sole involvement of the transport sites in the induction of the ADP-binding E(1) conformation. Since the cations are similar in their mode of induction of the high ADP affinity but they demonstrate a pronounced difference in ability to compete with K(+), their effects cannot be combined within any scheme with only one type of cation-binding sites. We suggest that the high affinity toward nucleotide is induced by cation interactions within the protein or lipid and that these nucleotide-domain-related sites coexist with the transport sites, which bind only Na(+) or K(+).  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