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 46 毫秒
1.
Chloride binding to alkaline phosphatase. 113Cd and 35Cl NMR   总被引:1,自引:0,他引:1  
Chloride binding to alkaline phosphatase from Escherichia coli has been monitored by 35Cl NMR for the native zinc enzyme and by 113Cd NMR for two Cd(II)-substituted species, phosphorylated Cd(II)6 alkaline phosphatase and unphosphorylated Cd(II)2 alkaline phosphatase. Of the three metal binding sites per enzyme monomer, A, B, and C, only the NMR signal of 113Cd(II) at the A sites shows sensitivity to the presence of Cl-, suggesting that Cl- coordination occurs at the A site metal ion. From the differences in the chemical shift changes produced in the A site 113Cd resonance for the covalent (E-P) form of the enzyme versus the noncovalent (E . P) form of the enzyme, it is concluded that the A site metal ion can assume a five-coordinate form. The E-P form of the enzyme has three histidyl nitrogens as ligands from the protein to the A site metal ion plus either two water molecules or two Cl- ions as additional monodentate ligands. In the E . P form, there is a phosphate oxygen as a monodentate ligand and either a water molecule or a Cl- ion as the additional monodentate ligand. The shifts of the 113Cd NMR signals of the unphosphorylated Cd(II)2 enzyme induced by Cl- are very similar to those induced in the E-P derivative of the same enzyme, supporting the conclusion that the phosphoseryl residue is not directly coordinated to any of the metal ions. Specific broadening of the 35Cl resonance from bulk Cl- is induced by Zn(II)4 alkaline phosphatase, while Zn(II)2 alkaline phosphatase is even more effective, suggesting an influence by occupancy of the B site on the interaction of monodentate ligands at the A site. A reduction in this quadrupolar broadening is observed upon phosphate binding at pH values where E . P is formed, but not at pH values where E-P is the major species, confirming a specific interaction of Cl- at the A site, the site to which phosphate is bound in E . P, but not in E-P. For the zinc enzyme, a significant decrease in phosphate binding affinity can be shown to occur at pH 8 where one monomer has a higher affinity than the other.  相似文献   

2.
Alkaline phosphatase from Escherichia coli contains three metal binding sites (A, B, and C) located at sites forming a triangle with sides of 4, 5, and 7 A (Wyckoff, H.W., Handschumacher, M., Murthy, K., and Sowadski, J.M. (1983) Adv. Enzymol. 55, 453). When all three sites are occupied by Cd(II) the enzyme has a very low turnover; at least 10(3) slower than the native Zn(II) enzyme. The slow turnover number has made the Cd(II) enzyme useful in NMR studies of the mechanism of alkaline phosphatase. The binding of arsenate to two forms of Cd(II) alkaline phosphatase (Cd(II)2alkaline phosphatase and Cd(II)6alkaline phosphatase) has been studied by 113Cd NMR. Cd(II)2alkaline phosphatase, pH 6.3, binds arsenate at only one monomer of the dimeric enzyme and causes migration of Cd(II) from the A site of one monomer to the B site of the arsenylated monomer. This same migration has previously been observed to accompany metal ion-dependent phosphate binding, but is much more rapid in the case of arsenate. The acceleration of migration induced by arsenate supports the conclusion based on the phosphate data that the substrate anion binds to the A site metal ion of one monomer prior to migration and that only the metal ion at A site is required for phosphorylation (arsenylation) of serine 102. The 113Cd chemical shifts of A and B site metal ions are very sensitive to the form of the bound arsenate, i.e. covalent (E-As) or noncovalent (E X As) complex. Like the analogous phosphate derivatives, the change of chemical shift of A site (to which phosphate is coordinated in the E X P complex) is much greater than that of the B site metal ion, when the arsenate shifts between the two intermediates, suggesting that arsenate is also coordinated to A site in the E X As intermediate. The chemical shifts of A and B site 113Cd(II) ions are considerably different in the arsenate and phosphate derivatives, while the C site 113Cd(II) ions have nearly identical chemical shifts. Thus the substrate appears to interact closely with both A and B sites, while C site appears relatively unimportant in phosphomonoester hydrolysis. The analogous behavior of arsenate and phosphate at the active center as evaluated by 113Cd NMR supports the validity of using the heavier arsenate derivative in x-ray diffraction studies.  相似文献   

3.
Alkaline phosphatase. 31P NMR probes of the mechanism   总被引:1,自引:0,他引:1  
31P NMR signals from substrates and products of alkaline phosphatase have been adapted to measure the rates and product ratios for the hydrolysis and phosphotransferase reactions from pH 6 to 10. Below pH 8, glycerol is a poorer acceptor than H2O (glycerol phosphates:Pi = 0.5). Tris is a more effective acceptor below pH 8, showing a maximum acceptor efficiency at pH 8 (Tris phosphate:Pi = 2). Phosphotransferase efficiencies are in the order expected for the pKaS of the alcohol groups, Tris less than glycerol Cl, C3 less than glycerol C2. Tris and glycerol induce chemical shifts in 113Cd(II) present at the A site but not the B or C sites of the metal triad present at each active center of Cd(II)6 alkaline phosphatase, suggesting that the alcoxides of the acceptors coordinate the A site metal and become the nucleophiles attacking the phosphoseryl residue (E-P) in the second step of the mechanism. The interaction is through the oxygen of Tris. The transferase activity of the amino alcohol shows a bell-shaped pH dependency. Aliphatic alcohol acceptors show small increases in acceptor activity between pH 6 and 8, with 5-fold increases from pH 8 to 10 (at pH 10, glycerol phosphates:Pi = 2.5). 31P NMR inversion transfer has been used to measure the koff for Pi dissociation from the noncovalent enzyme complex (E . P). For the Zn(II)4 alkaline phosphatase koff is essentially pH independent at approximately 35 s-1. For Cd(II) or Mg(II) at the B site in place of Zn(II), koff less than or equal to 1 s-1 X Cl-ion, which appears to coordinate the A site metal ion, enhances koff, suggesting that both Cl- and HPO2-4 can coordinate the A site metal ion in a 5-coordinate intermediate. pH control of the alkaline phosphatase mechanism appears to reside in the stability of E-P and not the dissociation of E . P, compatible with the hypothesis that the activity-linked pKa is that of a H2O molecule coordinated to the A site metal, which in the hydroxide form becomes the nucleophile attacking the phosphoseryl group (E-P).  相似文献   

4.
113Cd nuclear magnetic resonance of Cd(II) alkaline phosphatases   总被引:1,自引:0,他引:1  
113Cd NMR spectra of 113Cd(II)-substituted Escherichia coli alkaline phosphatase have been recorded over a range of pH values, levels of metal site occupancy, and states of phosphorylation. Under all conditions resonances attributable to cadmium specifically bound at one or more of the three pairs of metal-binding sites (A, B, and C sites) are detected. By following changes in both the 113Cd and 31P NMR spectra of 113Cd(II)2 alkaline phosphatase during and after phosphorylation, it has been possible to assign the cadmium resonance that occurs between 140 and 170 ppm to Cd(II) bound to the A or catalytic site of the enzyme and the resonance occurring between 51 and 76 ppm to Cd(II) bound to B site, which from x-ray data is located 3.9 A from the A site. The kinetics of phosphorylation show that cadmium migration from the A site of one subunit to the B site of the second subunit follows and is a consequence of phosphate binding, thus precluding the migration as a sufficient explanation for half-of-the-sites reactivity. Rather, there is evidence for subunit-subunit interaction rendering the phosphate binding sites inequivalent. Although one metal ion, at A site, is sufficient for phosphate binding and phosphorylation, the presence of a second metal ion at B site greatly enhances the rate of phosphorylation. In the absence of phosphate, occupation of the lower affinity B and C sites produces exchange broadening of the cadmium resonances. Phosphorylation abolishes this exchange modulation. Magnesium at high concentration broadens the resonances to the point of undetectability. The chemical shift of 113Cd(II) in both A and B sites (but not C site) is different depending on the state of the bound phosphate (whether covalently or noncovalently bound) and gives separate resonances for each form. Care must be taken in attributing the initial distribution of cadmium or phosphate in the reconstituted enzyme to that of the equilibrium species in samples reconstituted from apoenzyme. Both 113Cd NMR and 31P NMR show that some conformational changes consequent to metal ion or phosphate binding require several days before the final equilibrium species is formed.  相似文献   

5.
Differential scanning calorimetry of Cd(II) alkaline phosphatases   总被引:1,自引:0,他引:1  
Differential scanning calorimetry has been employed to monitor structural alterations induced in the dimeric enzyme alkaline phosphatase on binding of Cd(II) (to the metal-free apoenzyme) and phosphate (Pi) (to the Cd(II) enzyme). Cd(II) addition to the apoenzyme at pH 6.5 results in an increased transition temperature, suggesting a stabilizing effect of the bound metal ion. Two distinct structural forms of the protein are detected as discrete calorimetric transitions (Tm = 69-84 degrees C; 87-94 degrees C, respectively). Distribution of the enzyme between these forms is found to depend on the exogenous Cd(II) concentration and the protocol of Cd(II) addition. These results indicate that conversion between the conformational forms is a slow process which appears to require specific levels of metal ion site occupancy. These studies, in which the exogenous Cd(II) concentration was varied from 10(-5) M to 10(-3) M suggest a structural basis for previously observed hysteretic phenomena observed on Cd(II) binding to the enzyme. Even at a minimum stoichiometry of Cd(II) (2 eq/mol of dimer) a single equivalent of Pi is sufficient to accelerate assumption of a stabilized form of the protein (Tm = 90 degrees C). This is followed by a slow structural change paralleling the time course of formation of the functional 2 Cd(II) phosphoryl enzyme which displays two calorimetric transitions (Tm = 65 degrees C, 88 degrees C). The low temperature transition does not appear if Pi is initially present at millimolar concentrations and is abolished on addition of Pi at concentrations in excess of 0.1 mM. These observations suggest the presence of a second, distinct Pi binding site on the 2 Cd(II) phosphoryl enzyme. This is supported by the changes observed in the 31P NMR chemical shift of Pi added to comparable enzyme samples. These data, including assessment of the effect of the presence of Mg(II), are discussed in terms of the mechanism of metal ion association to the enzyme and rearrangement of bound metal ions induced by Pi binding.  相似文献   

6.
Gene 32 protein (g32P), the single-stranded DNA binding protein from bacteriophage T4, contains 1 mol of Zn(II)/mol of protein. This intrinsic zinc is retained within the DNA-binding core fragment, g32P-(A+B) (residues 22-253), obtained by limited proteolysis of the intact protein. Ultraviolet circular dichroism provides evidence that Zn(II) binding causes significant changes in the conformation of the peptide chain coupled with alterations in the microenvironments of tryptophan and tyrosine side chains. NMR spectroscopy of the 113Cd(II) derivative of g32P-(A+B) at both 44.4 and 110.9 MHz shows a single 113Cd resonance, delta 637, a chemical shift consistent with coordination to three of the four sulfhydryl groups in the protein. In vitro mutagenesis of Cys166 to Ser166 creates a mutant g32P that still contains 1 Zn(II)/molecule. This mutant protein when substituted with 113Cd(II) shows a 113Cd signal with a delta and a line width the same as those observed for the wild-type protein. Thus, the S-ligands to the metal ion appear to be contributed by Cys77, Cys87, and Cys90. Relaxation data suggest that chemical shift anisotropy is the dominant, but not exclusive, mechanism of relaxation of the 113Cd nucleus in g32P, since a dipolar modulation from ligand protons is observed at 44.4 MHz but not at 110.9 MHz. Complexation of core 113Cd g32P with d(pA)6 or Co(II) g32P with poly(dT) shows only minor perturbation of the NMR signal or d-d electronic transitions, respectively, suggesting that the metal ion in g32P does not add a ligand from the bound DNA.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
Cd-substituted forms of the Bacillus cereus metallo-beta-lactamases (BCII) were studied by perturbed angular correlation of gamma-rays (PAC) spectroscopy. At very low [Cd]:[apo-beta-lactamase] ratios, two nuclear quadrupole interactions (NQI) were detected. For [Cd]:[apo-beta-lactamase] ratios between 0.8 and 3.0, two new NQIs appear, and the spectra show that up to 2 cadmium ions can be bound per molecule of apoenzyme. These results show the existence of two interacting Cd-binding sites in BCII. The relative populations of the two NQIs found at low [Cd]:[apo-beta-lactamase] ratios yielded a 1:3 ratio for the microscopic dissociation constants of the two different metal sites (when only one cadmium ion is bound). X-ray diffraction data at pH 7.5 demonstrate that also for Zn(II) two binding sites exist, which may be bridged by a solvent molecule. The measured NQIs could be assigned to the site with three histidines as metal ligands (three-His site) and to the site with histidine, cysteine, and aspartic acid as metal ligands (Cys site), respectively, by PAC measurements on the Cys168Ala mutant enzyme. This assignment shows that cadmium ions preferentially bind to the Cys site. This is in contrast to the preference of Zn(II) in the hybrid Zn(II)Cd(II) enzyme, where an analysis of the corresponding PAC spectrum showed that Cd(II) occupied the Cys site, whereby Zn(II) occupied the site with three histidines. The difference between Zn(II) and Cd(II) in affinity for the two sites is combined with the kinetics of hydrolysis of nitrocefin for different metal ion substitutions (Zn(2)E, ZnE, Cd(2)E, CdE, and ZnCdE) to study the function of the two metal ion binding sites.  相似文献   

8.
Rabbit 113Cd7-metallothionein-2a (MT) contains two metal-thiolate clusters of three (cluster B) and four (cluster A) metal ions. The 113Cd-n.m.r. spectrum of 113Cd6-MT, isolated from 113Cd7-MT upon treatment with EDTA, is similar to that of 113Cd7-MT, but the cluster B resonances are lower in intensity, suggesting its co-operative metal depletion. (Zn1,113Cd6)-MT, formed upon addition of the Zn(II) ions to 113Cd6-MT, shows 113Cd-n.m.r. features characteristic of cluster B populations containing both Cd(II) and Zn(II) ions. The overall intensity gain of the mixed cluster B resonances per Cd as to those in 113Cd6- and 113Cd7-MT suggests a stabilization effect of the bound Zn(II) ions upon the previously established intramolecular 113Cd exchange within this cluster.  相似文献   

9.
The rate constants which characterize the formation and breakdown of the noncovalent (E.P) and covalent (E-P) enzyme-phosphate intermediates on the alkaline phosphatase reaction pathway are known to be sensitive to the nature of the metal ion bound to the enzyme. 31P NMR saturation transfer has been demonstrated to provide a simple and sensitive method for measuring the metal ion dependence of these rates under equilibrium conditions. When the native Zn2+ was replaced by Cd2+, the 31P NMR spectrum at high pH revealed a new resonance at 12.6 ppm which has been assigned to the noncovalent enzyme.phosphate complex. Reconstituting the enzyme with enriched 113Cd2+ caused this unusually downfield-shifted resonance to appear as a doublet due to 113Cd-31P spin coupling (2J31P-O-113Cd = 30 Hz). This result provides the first unequivocal evidence for direct metal-phosphate interaction in alkaline phosphatase.  相似文献   

10.
Inactivation of Helicobacter pylori cadA, encoding a putative transition metal ATPase, was only possible in one of four natural competent H. pylori strains, designated 69A. All tested cadA mutants showed increased growth sensitivity to Cd(II) and Zn(II). In addition, some of them showed both reduced 63Ni accumulation during growth and no or impaired urease activity, which was not due to lack of urease enzyme subunits. Gene complementation experiments with plasmid (pY178)-derived H. pylori cadA failed to correct the deficiencies, whereas resistance to Cd(II) and Zn(II) was restored. Moreover, pY178 conferred increased Co(II) resistance to both the cadA mutants and the wild-type strain 69A. Heterologous expression of H. pylori cadA in an Escherichia coli zntA mutant resulted in an elevated resistance to Cd(II) and Zn(II). Expression of cadA in E. coli SE5000 harbouring H. pylori nixA, which encodes a divalent cation importer along with the H. pylori urease gene cluster, led to about a threefold increase in urease activity compared with E. coli control cells lacking the H. pylori cadA gene. These results suggest that H. pylori CadA is an essential resistance pump with ion specificity towards Cd(II), Zn(II) and Co(II). They also point to a possible role of H. pylori CadA in high-level activity of H. pylori urease, an enzyme sensitive to a variety of metal ions.  相似文献   

11.
12.
13.
Rat matrix-induced alkaline phosphatase is an enzyme which requires magnesium and zinc ions for its maximal activity. Two Zn(II) ions and one Mg(II) ion are bound to each subunit of native dimeric enzyme. The presence of magnesium ion (10-100 microM) or zinc ion (7-20 nM) alone is sufficient to stimulate apoenzyme activity. However maximal activity (264 U/mg) requires the presence of both ions. Binding of Zn(II) ions to the Mg(II) binding site causes a strong inhibition of the apoenzyme while the binding of Mg(II) on Zn(II) binding site is not sufficient to stimulate PNPPase activity of the apoenzyme. Binding of both ions to the enzyme molecule did not change the apparent dissociation constant for PNPP hydrolysis.  相似文献   

14.
The aim of this study is to investigate complexation of metallothionein (MT) with cadmium and zinc ions. An oligopeptide (i.e. Lys-Cys-Thr-Cys-Cys-Ala), a fragment of MT was covalently immobilized onto piezoelectric crystals, which were first treated with ethylene diamine plasma in a glow-discharge apparatus, and then were chemically reacted with glutaraldehyde. Complexation of the immobilized MT with Zn(II) and Cd(II) ions in aqueous media was followed by recording the changes of the frequency shifts of the piezoelectric quartz crystals. The amount of Cd(II) ions interacted with the immobilized MT molecules was the highest at pH 7.4, and decreased with an increase in the pH of the medium, in parallel to the decrease in the amount of immobilized MT. The number of Zn(II) ions interacted with the immobilized MT molecules was higher than the number of Cd(II) ions when the adsorption was from solutions containing a single-metal ion with the same ion concentrations. In consecutive adsorption studies, we observed that the type of metal ions used in the first interaction is important. These experiments showed also that there is an exchange between the metal ions, and competition provokes adsorption of both ions due to synergistic-antagonistic effects.  相似文献   

15.
16.
The complex formation equilibria of Zn(II) and Cd(II) with cephalexin have been studied through potentiometric titrations. Experimental data were analyzed using the least squares computer program SUPERQUAD. The stability constants were 1g beta ZnCEX+ = 2.40, 1g beta Zn(CEX)(OH) = -4.54, 1g beta CdCEX+ = 2.18, and 1g beta Cd(CEX)(OH) = -5.18 (I = 0.1 M NaNO3), CEX complexes of formulae Zn(CEX)2(3)H2O and Cd(CEX)(OH)H2O have been synthesized and characterized by elemental analysis, IR spectra, conductivity measurements, and electronic and NMR spectra. The thermal behavior of the synthesized compounds were studied by TGA and DTA. We conclude that the metal ion interacts with the amido group of CEX.  相似文献   

17.
The binding of Cd(II) and Zn(II) to human serum albumin (HSA) and dog serum albumin (DSA) has been studied by equilibrium dialysis and 113Cd(II)-NMR techniques at physiological pH. Scatchard analysis of the equilibrium dialysis data indicate the presence of at least two classes of binding sites for Cd(II) and Zn(II). On analysis of the high-affinity class of sites, HSA is shown to bind 2.08 +/- 0.09 (log K = 5.3 +/- 0.6) and 1.07 +/- 0.12 (log K = 6.4 +/- 0.8) moles of Cd(II) and Zn(II) per mole of protein, respectively. DSA bound 2.02 +/- 0.19 (log K = 5.1 +/- 0.8), and 1.06 +/- 0.15 (log K = 6.0 +/- 0.2) moles of Cd(II) and Zn(II) per mole of protein, respectively. Competition studies indicate the presence of one high-affinity Cd(II) site on both HSA and DSA that is not affected by Zn(II) or Cu(II), and one high-affinity Zn(II) site on both HSA and DSA that is not affected by Cd(II) or Cu(II). 113Cadmium-HSA spectra display three resonances corresponding to three different sites of complexation. In site I, Cd(II) is most probably coordinated to two or three histidyl residues, site II to one histidyl residue and three oxygen ligands (carboxylate), while for the most upfield site III, four oxygens are likely to be involved in the binding of the metal ion. The 113Cd(II)-DSA spectra display only two resonances corresponding to two different sites of complexation. The environment around Cd(II) at sites I and II on DSA is similar to sites I and II, respectively, on HSA. No additional resonances are observed in any of these experiments and in particular in the low field region where sulfur coordination occurs. Overall, our results are consistent with the proposal that the physiologically important high-affinity Zn(II) and Cd(II) binding sites of albumins are located not at the Cu(II)-specific NH2-terminal site, but at internal sites, involving mostly nitrogen and oxygen ligands and no sulphur ligand.  相似文献   

18.
Gene 32 protein (g32P), the replication accessory single-stranded nucleic acid binding protein from bacteriophage T4, contains 1 mol of Zn(II)/mol of protein. Zinc coordination provides structural stability to the DNA-binding core domain of the molecule, termed g32P-(A+B) (residues 22-253). Optical absorption studies with the Co(II)-substituted protein and 113Cd NMR spectroscopy of 113Cd(II)-substituted g32P-(A+B) show that the metal coordination sphere in g32P is characterized by approximately tetrahedral ligand symmetry and ligation by the Cys-S- atoms of Cys77, Cys87, and Cys90. These studies predicted the involvement of a fourth protein-derived non-thiol ligand to complete the tetrahedral complex, postulated to be His81 on the basis of primary structure prediction and modeling [Giedroc, D.P., Johnson, B.A., Armitage, I.M., & Coleman, J.E. (1989) Biochemistry 28, 2410-2418]. To test this model, we have employed site-directed mutagenesis to substitute each of the two histidine residues in g32P (His64 and His81), accompanied by purification and structural characterization of these single-site mutant proteins. We show that g32P's containing any of three substitutions at residue 64 (H64Q, H64N, and H64L) are isolated from Escherichia coli in a Zn(II)-free form [less than or equal to 0.03 g.atom Zn(II)]. All derivatives show extremely weak affinity for the ssDNA homopolymer poly(dT). All are characterized by a far-UV-CD spectrum reduced in negative intensity relative to the wild-type protein. These structural features parallel those found for the known metal ligand mutant Cys87----Ser87 (C87S) g32P. In contrast, g32P-(A+B) containing a substitution of His81 with glutamine (H81Q), alanine (H81A) or cysteine (H81C), contains stoichiometric Zn(II) as isolated and binds to polynucleotides with an affinity comparable to the wild-type g32P-(A+B). Spin-echo 1H NMR spectra recorded for wild-type and H81Q g32P-(A+B) as a function of pH allow the assignment of His81 ring proteins to delta = 6.81 and 6.57 ppm, respectively, at pH 7.8, corresponding to the C and D histidyl protons of 1H-His-g32P-(A+B) [Pan, T., Giedroc, D.P., & Coleman, J.E. (1989) Biochemistry 28, 8828-8832]. These resonances shift downfield as the pH is reduced from 7.8 to 6.6 without metal dissociation, a result incompatible with His81 donating a ligand to the Zn(II) in wild-type g32P. Likewise, Cys81 in Zn(II) H81C g32P is readily reactive with 5,5'-dithiobis(2-nitrobenzoic acid), unlike metal ligands Cys77, Cys87, and Cys90.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

19.
20.
B Mitra  R Sharma 《Biochemistry》2001,40(25):7694-7699
Soft metal-translocating P1-type ATPases have a distinctive amino-terminal domain that contains one to six copies of the conserved metal-binding motif, GXXCXXC. ZntA from Escherichia coli, a Pb(II)-, Zn(II)-, and Cd(II)-transporting ATPase, has an approximately 120 residue amino-terminal domain with one copy of the GXXCXXC motif as well as four additional cysteine residues. The function of this domain was investigated by constructing a mutant of ZntA lacking the first approximately 100 residues. The mutant, DeltaN-ZntA, was able to confer resistance to Pb(II), Zn(II), and Cd(II) salts, in a manner similar to ZntA. The soft metal dependent ATP hydrolysis activity of purified DeltaN-ZntA was characterized. Purified DeltaN-ZntA and ZntA were both inactivated by oxidation. The K(m) for MgATP was unchanged for DeltaN-ZntA relative to ZntA. DeltaN-ZntA displayed the same metal ion specificity as ZntA. Thiolates increased the activities of both ZntA and DeltaN-ZntA. The V(max) values for DeltaN-ZntA were approximately 3-fold lower than for ZntA for all three metal ions. Thus, the amino-terminal domain is not essential for the function of ZntA or for conferring specificity toward particular soft metals. Its function may be to increase the overall catalytic rate by increasing the rate of metal ion binding to the transporter. Residues involved in the ATP-dependent soft metal ion-translocating mechanism as well as those responsible for recognition of specific metal ions must be part of the core structure of the P1-type ATPases.  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