首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
We introduce two methods for the visualization of phosphorylated proteins using alkoxide-bridged dinuclear metal (i.e. Zn(2+) or Mn(2+)) complexes as novel phosphate-binding tag (Phos-tag) molecules. Both Zn(2+)- and Mn(2+)-Phos-tag molecules preferentially capture phosphomonoester dianions bound to Ser, Thr, and Tyr residues. One method is based on an ECL system using biotin-pendant Zn(2+)-Phos-tag and horseradish peroxidase-conjugated streptavidin. We demonstrate the electroblotting analyses of protein phosphorylation status by the phosphate-selective ECL signals. Another method is based on the mobility shift of phosphorylated proteins in SDS-PAGE with polyacrylamide-bound Mn(2+)-Phos-tag. Phosphorylated proteins in the gel are visualized as slower migration bands compared with corresponding dephosphorylated proteins. We demonstrate the kinase and phosphatase assays by phosphate affinity electrophoresis (Mn(2+)-Phos-tag SDS-PAGE).  相似文献   

3.
Arabidopsis thaliana MTP1 is a vacuolar membrane Zn(2+)/H(+) antiporter of the cation diffusion facilitator family. Here we present a structure-function analysis of AtMTP1-mediated transport and its remarkable Zn(2+) selectivity by functional complementation tests of more than 50 mutant variants in metal-sensitive yeast strains. This was combined with homology modeling of AtMTP1 based on the crystal structure of the Escherichia coli broad-specificity divalent cation transporter YiiP. The Zn(2+)-binding sites of EcYiiP in the cytoplasmic C-terminus, and the pore formed by transmembrane helices TM2 and TM5, are conserved in AtMTP1. Although absent in EcYiiP, Cys31 and Cys36 in the extended N-terminal cytosolic domain of AtMTP1 are necessary for complementation of a Zn-sensitive yeast strain. On the cytosolic side of the active Zn(2+)-binding site inside the transmembrane pore, Ala substitution of either Asn258 in TM5 or Ser101 in TM2 non-selectively enhanced the metal tolerance conferred by AtMTP1. Modeling predicts that these residues obstruct the movement of cytosolic Zn(2+) into the intra-membrane Zn(2+)-binding site of AtMTP1. A conformational change in the immediately preceding His-rich cytosolic loop may displace Asn258 and permit Zn(2+) entry into the pore. This would allow dynamic coupling of Zn(2+) transport to the His-rich loop, thus acting as selectivity filter or sensor of cytoplasmic Zn(2+) levels. Individual mutations at diverse sites within AtMTP1 conferred Co and Cd tolerance in yeast, and included deletions in N-terminal and His-rich intra-molecular cytosolic domains, and mutations of single residues flanking the transmembrane pore or participating in intra- or inter-molecular domain interactions, all of which are not conserved in the non-selective EcYiiP.  相似文献   

4.
The gamma-aminobutyric acid, type A (GABAA), receptor ion channel is lined by the second membrane-spanning (M2) segments from each of five homologous subunits that assemble to form the receptor. Gating presumably involves movement of the M2 segments. We assayed protein mobility near the M2 segment extracellular ends by measuring the ability of engineered cysteines to form disulfide bonds and high affinity Zn(2+)-binding sites. Disulfide bonds formed in alpha1beta1E270Cgamma2 but not in alpha1N275Cbeta1gamma2 or alpha1beta1gamma2K285C. Diazepam potentiation and Zn2+ inhibition demonstrated that expressed receptors contained a gamma subunit. Therefore, the disulfide bond in alpha1beta1E270Cgamma2 formed between non-adjacent subunits. In the homologous acetylcholine receptor 4-A resolution structure, the distance between alpha carbon atoms of 20' aligned positions in non-adjacent subunits is approximately 19 A. Because disulfide trapping involves covalent bond formation, it indicates the extent of movement but does not provide an indication of the energetics of protein deformation. Pairs of cysteines can form high affinity Zn(2+)-binding sites whose affinity depends on the energetics of forming a bidentate-binding site. The Zn2+ inhibition IC50 for alpha1beta1E270Cgamma2 was 34 nm. In contrast, it was greater than 100 microM in alpha1N275Cbeta1gamma2 and alpha1beta1gamma2K285C receptors. The high Zn2+ affinity in alpha1beta1E270Cgamma2 implies that this region in the beta subunit has a high protein mobility with a low energy barrier to translational motions that bring the positions into close proximity. The differential mobility of the extracellular ends of the beta and alpha M2 segments may have important implications for GABA-induced conformational changes during channel gating.  相似文献   

5.
In addition to binding Ca(2+), the S100 protein S100B binds Zn(2+) with relatively high affinity as confirmed using isothermal titration calorimetry (ITC; K(d) = 94 +/- 17 nM). The Zn(2+)-binding site on Ca(2+)-bound S100B was examined further using NMR spectroscopy and site-directed mutagenesis. Specifically, ITC measurements of S100B mutants (helix 1, H15A and H25A; helix 4, C84A, H85A, and H90A) were found to bind Zn(2+) with lower affinity than wild-type S100B (from 2- to >25-fold). Thus, His-15, His-25, Cys-84, His-85, and perhaps His-90 of S100B are involved in coordinating Zn(2+), which was confirmed by NMR spectroscopy. Previous studies indicate that the binding of Zn(2+) enhances calcium and target protein-binding affinities, which may contribute to its biological function. Thus, chemical shift perturbations observed here for residues in both EF-hand domains of S100B during Zn(2+) titrations could be detecting structural changes in the Ca(2+)-binding domains of S100B that are pertinent to its increase in Ca(2+)-binding affinity in the presence of Zn(2+). Furthermore, Zn(2+) binding causes helix 4 to extend by one full turn when compared to Ca(2+)-bound S100B. This change in secondary structure likely contributes to the increased binding affinity that S100B has for target peptides (i.e., TRTK peptide) in the presence of Zn(2+).  相似文献   

6.
Recently, we have described a distance constraint in the unknown tertiary structure of the human dopamine transporter (hDAT) by identification of two histidines, His(193) in the second extracellular loop and His(375) at the top of transmembrane (TM) 7, that form two coordinates in an endogenous, high affinity Zn(2+)-binding site. To achieve further insight into the tertiary organization of hDAT, we set out to identify additional residues involved in Zn(2+) binding and subsequently to engineer artificial Zn(2+)-binding sites. Ten aspartic acids and glutamic acids, predicted to be on the extracellular side, were mutated to asparagine and glutamine, respectively. Mutation of Glu(396) (E396Q) at the top of TM 8 increased the IC(50) value for Zn(2+) inhibition of [(3)H]dopamine uptake from 1.1 to 530 microM and eliminated Zn(2+)-induced potentiation of [(3)H]WIN 35,428 binding. These data suggest that Glu(396) is involved in Zn(2+) binding to hDAT. Importantly, Zn(2+) sensitivity was preserved following substitution of Glu(396) with histidine, indicating that the effect of mutating Glu(396) is not an indirect effect because of the removal of a negatively charged residue. The common participation of Glu(396), His(193), and His(375) in binding the small Zn(2+) ion implies their proximity in the unknown tertiary structure of hDAT. The close association between TM 7 and 8 was further established by engineering of a Zn(2+)-binding site between His(375) and a cysteine inserted in position 400 in TM 8. Summarized, our data define an important set of proximity relationships in hDAT that should prove an important template for further exploring the molecular architecture of Na(+)/Cl(-)-dependent neurotransmitter transporters.  相似文献   

7.
Publication of the rhodopsin X-ray structure has facilitated the development of homology models of other G protein-coupled receptors. However, possible shifts of transmembrane (TM) alpha helices, expected variations in helical distortions, and differences in loop size necessitate experimental verification of these comparative models. To refine a rhodopsin-based homology model of the mu-opioid receptor (MOR), we experimentally determined structural-distance constraints from intrinsic and engineered metal-binding sites in the rat MOR. Investigating the relatively high intrinsic affinity of MOR for Zn(2+) (IC(50) approximately 30microM), we observed that mutation of His(319) (TM7) abolished Zn(2+) inhibition of ligand binding, while mutation of Asp(216) (extracellular loop 2) decreased the effect of Zn(2+), suggesting these residues participate in the intrinsic Zn(2+)-binding center of MOR. To verify the relative orientation of TM5 and TM6 and to examine whether a rhodopsin-like alpha aneurism is present in TM5, we engineered Zn(2+)-binding centers by mutating residues of TM5 and TM6 to Cys or His, making use of the native His(297) in TM6 as an additional Zn(2+)-coordination site. Inhibition of opioid ligand binding by Zn(2+) suggests that residues Ile(234) and Phe(237) in TM5 face the binding-site crevice and form a metal-binding center with His(297) and Val(300) in TM6. This observation is inconsistent with a rhodopsin-like structure, which would locate Ile(234) on the lipid-exposed side of TM5, too distant from other residues making up the Zn(2+)-binding site. Subsequent distance geometry refinement of the MOR model indicates that the rhodopsin-like alpha aneurism is likely absent in TM2 but present in TM5.  相似文献   

8.
The vertebrate calmodulin is configured with two structurally independent globular lobes in N- and C-terminus, and a flexible central linker. Distinctly, two lobes of calmodulin from Saccharomyces cerevisiae (yCaM) interact and influence the Ca(2+)-binding profile of each other. We explored this further using the mutant proteins with eliminated Ca(2+)-binding ability in one of the lobes and found that the Ca(2+)-bound N-lobe associates with the Ca(2+)-free C-lobe to gain the Ca(2+) affinity of a wild-type level. Next, analysing series of C-terminal residue truncation mutant, we found that the truncation of C-terminal three residues induce the hyper Ca(2+) affinity. These residues are also important for the general structural behaviour of calmodulin, such as Ca(2+)-induced slow mobility shift in polyacrylamide gel electrophoresis and for the ability to activate Cmk1p (yeast calmodulin kinase). These suggest: (i) when Ca(2+) occupies only N-lobe, two lobes interact and form the stable intermediate leading to a proper level of Ca(2+) affinity; (ii) the C-terminal three residues are required to prohibit abnormal stabilization of the intermediate promoting abnormally high Ca(2+) affinity and for recognition of target enzymes. A model for Ca(2+) and target bindings of yCaM is proposed. Evolutional aspect concerning the biological significance of this model was discussed.  相似文献   

9.
Metal ions are indispensable cofactors for chemical catalysis by a plethora of enzymes. Porphobilinogen synthases (PBGSs), which catalyse the second step of tetrapyrrole biosynthesis, are grouped according to their dependence on Zn(2+). Using site-directed mutagenesis, we embarked on transforming Zn(2+)-independent Pseudomonas aeruginosa PBGS into a Zn(2+)-dependent enzyme. Nine PBGS variants were generated by permutationally introducing three cysteine residues and a further two residues into the active site of the enzyme to match the homologous Zn(2+)-containing PBGS from Escherichia coli. Crystal structures of seven enzyme variants were solved to elucidate the nature of Zn(2+) coordination at high resolution. The three single-cysteine variants were invariably found to be enzymatically inactive and only one (D139C) was found to bind detectable amounts of Zn(2+). The double mutant A129C/D139C is enzymatically active and binds Zn(2+) in a tetrahedral coordination. Structurally and functionally it mimics mycobacterial PBGS, which bears an equivalent Zn(2+)-coordination site. The remaining two double mutants, without known natural equivalents, reveal strongly distorted tetrahedral Zn(2+)-binding sites. Variant A129C/D131C possesses weak PBGS activity while D131C/D139C is inactive. The triple mutant A129C/D131C/D139C, finally, displays an almost ideal tetrahedral Zn(2+)-binding geometry and a significant Zn(2+)-dependent enzymatic activity. Two additional amino acid exchanges further optimize the active site architecture towards the E.coli enzyme with an additional increase in activity. Our study delineates the potential evolutionary path between Zn(2+)-free and Zn(2+)-dependent PBGS enyzmes showing that the rigid backbone of PBGS enzymes is an ideal framework to create or eliminate metal dependence through a limited number of amino acid exchanges.  相似文献   

10.
The display of peptide sequences on the surface of bacteria is a technology that offers exciting applications in biotechnology and medical research. Type 1 fimbriae are surface organelles of Escherichia coli which mediate D-mannose-sensitive binding to different host surfaces by virtue of the FimH adhesin. FimH is a component of the fimbrial organelle that can accommodate and display a diverse range of peptide sequences on the E. coli cell surface. In this study we have constructed a random peptide library in FimH. The library, consisting of approximately 40 million individual clones, was screened for peptide sequences that conferred on recombinant cells the ability to bind Zn(2+). By serial selection, sequences that exhibited various degrees of binding affinity and specificity toward Zn(2+) were enriched. None of the isolated sequences showed similarity to known Zn(2+)-binding proteins, indicating that completely novel Zn(2+)-binding peptide sequences had been isolated. By changing the protein scaffold system, we demonstrated that the Zn(2+)-binding seems to be uniquely mediated by the peptide insert and to be independent of the sequence of the carrier protein. These findings might be applied in the design of biomatrices for bioremediation purposes or in the development of sensors for detection of heavy metals.  相似文献   

11.
When high affinity Ca(2+)-binding proteins like calmodulin, or proteins with a high Ca(2+)-binding capacity like calsequestrin, underwent sodium dodecyl sulfate-gel electrophoresis in Laemmli systems, their electrophoretic migration rates were much higher in gels containing 1 mM Ca2+ than in gels containing ethylene glycol bis(beta-aminoethyl ether) N,N'-tetraacetic acid (EGTA). Replacement of EGTA by Ca2+ in the gel, combined with the blotting of electrophoretically separated proteins on polyvinylidene difluoride membranes and subsequent 45Ca2+ overlay, proved a very effective means of detecting Ca(2+)-binding proteins. This combined approach is important since artifacts occur in both techniques when used separately. We found that the usual procedure of adding Ca2+ to the sample before electrophoresis without including it in the gel itself (C.B. Klee, T. H. Crouch, and M. H. Krinks, 1979, Proc. Natl. Acad. Sci. USA 76, 6270-6273) permitted the detection of only very high affinity Ca(2+)-binding proteins.  相似文献   

12.
S100 family proteins are characterized by short individual N and C termini and a conserved central part, harboring two Ca(2+)-binding EF-hands, one of them highly conserved among EF-hand family proteins and the other characteristic for S100 proteins. In addition to Ca(2+), several members of the S100 protein family, including S100A2, bind Zn(2+). Two regions in the amino acid sequences of S100 proteins, namely the helices of the N-terminal EF-hand motif and the very C-terminal loop are believed to be involved in Zn(2+)-binding due to the presence of histidine and/or cysteine residues. Human S100A2 contains four cysteine residues, each of them located at positions that may be important for Zn(2+) binding. We have now constructed and purified 10 cysteine-deficient mutants of human S100A2 by site-directed mutagenesis and investigated the contribution of the individual cysteine residues to Zn(2+) binding. Here we show that Cys(1(3)) (the number in parentheses indicating the position in the sequence of S100A2) is the crucial determinant for Zn(2+) binding in association with conformational changes as determined by internal tyrosine fluorescence. Solid phase Zn(2+) binding assays also revealed that the C-terminal residues Cys(3(87)) and Cys(4(94)) mediated a second type of Zn(2+) binding, not associated with detectable conformational changes in the molecule. Cys(2(22)), by contrast, which is located within the first EF hand motif affected neither Ca(2+) nor Zn(2+) binding, and a Cys "null" mutant was entirely incapable of ligating Zn(2+). These results provide new information about the mechanism and the site(s) of zinc binding in S100A2.  相似文献   

13.
Zinc (II) modulates the function of many integral membrane proteins. To identify the Zn(2+)-binding site responsible for allosteric modulation of the D(2) dopamine receptor, we first demonstrated that the binding site is likely located in extracellular loops or in transmembrane regions that are accessible from the extracellular milieu. We mutated every histidine in these regions to alanine; two mutants, H394A and H399A, exhibited a reduced response to Zn(2+). Combined mutation of H394 and H399 caused a larger effect of zinc than did either single mutation. Mutation of other potential Zn(2+)-binding residues predicted to be in proximity to H394 or H399 did not substantially alter the potency of Zn(2+). The double mutant H394A/H399A was similar to D(2) in affinity for [(3)H]spiperone and ability to inhibit cyclic AMP accumulation. We conclude that binding of Zn(2+) to H394 and H399 on the dopamine D(2) receptor contributes to allosteric regulation of antagonist binding.  相似文献   

14.
Histidine-rich Ca(2+)-binding protein (HRC) is a 170 kDa protein that can be identified in the isolated sarcoplasmic reticulum from rabbit skeletal muscle by its ability to bind [125I]low-density lipoprotein on blots after SDS-PAGE and that appears to be bound to the junctional membrane through calcium bridges. Molecular cDNA cloning of this protein predicts the existence of a Ca(2+)-binding domain and of a distinct heavy-metal binding domain at the cystein-rich COOH-terminus. Here we demonstrate, using radioactive ligand blot techniques, that HRC protein binds 45Ca at low affinity, as well as being able to bind 65Zn, but at different sites, that are largely inhibitable by prior reductive alkylation of the protein. In contrast to Ca(2+)-binding protein calsequestrin not having detectable 65Zn-binding sites, HRC protein bound selectively to immobilized Zn2+ on IDA-agarose affinity columns. Our results also indicate that rabbit and human 140 kDa HRC protein have common properties.  相似文献   

15.
A heterotropic allosteric effect involves an effector molecule that is distinct from the substrate or ligand of the protein. How heterotropic allostery originates is an unanswered question. We have previously created several heterotropic allosteric enzymes by recombining the genes for TEM1 beta-lactamase (BLA) and maltose binding protein (MBP) to create BLAs that are positively or negatively regulated by maltose. We show here that one of these engineered enzymes has approximately 10(6) M(-1) affinity for Zn(2+), a property that neither of the parental proteins possesses. Furthermore, Zn(2+) is a negative effector that noncompetitively switches off beta-lactam hydrolysis activity. Mutagenesis experiments indicate that the Zn(2+)-binding site does not involve a histidine or a cysteine, which is atypical of natural Zn(2+)-binding sites. These studies also implicate helices 1 and 12 of the BLA domain in allosteric signal propagation. These results support a model for the evolution of heterotropic allostery in which effector affinity and allosteric signaling emerge simultaneously.  相似文献   

16.
The metabolic pathway for histidine biosynthesis is interesting from an evolutionary perspective because of the diversity of gene organizations and protein structures involved. Hydrolysis of phosphoribosyl-AMP, the third step in the histidine biosynthetic pathway, is carried out by PR-AMP cyclohydrolase, the product of the hisI gene. The three-dimensional structure of PR-AMP cyclohydrolase from Methanobacterium thermoautotrophicum was solved and refined to 1.7 A resolution. The enzyme is a homodimer. The position of the Zn(2+)-binding site that is essential for catalysis was inferred from the positions of bound Cd(2+) ions, which were part of the crystallization medium. These metal binding sites include three cysteine ligands, two from one monomer and the third from the second monomer. The enzyme remains active when Cd(2+) is substituted for Zn(2+). The likely binding site for Mg(2+), also necessary for activity in a homologous cyclohydrolase, was also inferred from Cd(2+) positions and is comprised of aspartic acid side chains. The putative substrate-binding cleft is formed at the interface between the two monomers of the dimer. This fact, combined with the localization of the Zn(2+)-binding site, indicates that the enzyme is an obligate dimer.  相似文献   

17.
In recent years, protein translocation has been implicated as the mechanism that controls assembly of signaling complexes and induction of signaling cascades. Several members of the multifunctional Ca(2+)- (Zn(2+)- and Cu(2+))-binding S100 proteins appear to translocate upon cellular stimulation, and some are even secreted from cells, exerting extracellular functions. We transfected cells with S100B-green fluorescent fusion proteins and followed the relocation in real time. A small number of cells underwent translocation spontaneously. However, the addition of thapsigargin, which increases Ca(2+) levels, intensified ongoing translocation and secretion or induced these processes in resting cells. On the other hand, EGTA or BAPTA (1,2-bis(2-aminophenoxy)ethane-N,N,N',N'-tetraacetic acid), the Ca(2+)-chelating agents, inhibited these processes. In contrast, relocation of S100B seemed to be negatively dependent on Zn(2+) levels. Treatment of cells with TPEN (N,N,N',N'-tetrakis(2-pyridylmethyl)ethylenediamine), a Zn(2+)-binding drug, resulted in a dramatic redistribution and translocation of S100B. Secretion of S100B, when measured by ELISA, was dependent on cell density. As cells reached confluence the secretion drastically declined. However, an increase in Ca(2+) levels, and even more so, a decrease in Zn(2+) concentration, reactivated secretion of S100B. On the other hand, secretion did not decrease by treatment with brefeldin A, supporting the view that this process is independent of the endoplasmic reticulum-Golgi classical secretion pathway.  相似文献   

18.
The protease domain within the RUBV (rubella virus) NS (non-structural) replicase proteins functions in the self-cleavage of the polyprotein precursor into the two mature proteins which form the replication complex. This domain has previously been shown to require both zinc and calcium ions for optimal activity. In the present study we carried out metal-binding and conformational experiments on a purified cysteine-rich minidomain of the RUBV NS protease containing the putative Zn(2+)-binding ligands. This minidomain bound to Zn(2+) with a stoichiometry of approximately 0.7 and an apparent dissociation constant of <500 nM. Fluorescence quenching and 8-anilinonaphthalene-1-sulfonic acid fluorescence methods revealed that Zn(2+) binding resulted in conformational changes characterized by shielding of hydrophobic regions from the solvent. Mutational analyses using the minidomain identified residues Cys(1175), Cys(1178), Cys(1225) and Cys(1227) were required for the binding of Zn(2+). Corresponding mutational analyses using a RUBV replicon confirmed that these residues were necessary for both proteolytic activity of the NS protease and viability. The present study demonstrates that the CXXC(X)(48)CXC Zn(2+)-binding motif in the RUBV NS protease is critical for maintaining the structural integrity of the protease domain and essential for proteolysis and virus replication.  相似文献   

19.
Zinc ion (Zn(2+)) can be coordinated with four or three amino acid residues to stabilize a protein's structure or to form a catalytic active center. We used phage display selection of a dodecamer random peptide library with Zn(2+) to identify structural zinc sites. The binding specificity for Zn(2+) of selected sequences was confirmed using enzyme-linked immunosorbent and competitive inhibition assays. Circular dichroism spectra indicated that the interaction with Zn(2+) induced a change in conformation, which means the peptide acts as a structural zinc site. Furthermore, a search of protein databases revealed that two selected sequences corresponded to parts of natural zinc sites of copper/zinc superoxide dismutase and zinc-containing ferredoxin. We demonstrated that Zn(2+)-binding sequences selected from the random combinatorial library would be candidates for artificial structural zinc sites.  相似文献   

20.
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