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 structure of the saccharide-binding site of concanavalin A.   总被引:15,自引:1,他引:14       下载免费PDF全文
A complex of concanavalin A with methyl alpha-D-mannopyranoside has been crystallized in space group P212121 with a = 123.9 A, b = 129.1 A and c = 67.5 A. X-ray diffraction intensities to 2.9 A resolution have been collected on a Xentronics/Nicolet area detector. The structure has been solved by molecular replacement where the starting model was based on refined coordinates of an I222 crystal of saccharide-free concanavalin A. The structure of the saccharide complex was refined by restrained least-squares methods to an R-factor value of 0.19. In this crystal form, the asymmetric unit contains four protein subunits, to each of which a molecule of mannoside is bound in a shallow crevice near the surface of the protein. The methyl alpha-D-mannopyranoside molecule is bound in the C1 chair conformation 8.7 A from the calcium-binding site and 12.8 A from the transition metal-binding site. A network of seven hydrogen bonds connects oxygen atoms O-3, O-4, O-5 and O-6 of the mannoside to residues Asn14, Leu99, Tyr100, Asp208 and Arg228. O-2 and O-1 of the mannoside extend into the solvent. O-2 is hydrogen-bonded through a water molecule to an adjacent asymmetric unit. O-1 is not involved in any hydrogen bond and there is no fixed position for its methyl substituent.  相似文献   

2.
A Sadhu  J A Magnuson 《Biochemistry》1989,28(8):3197-3204
The stoichiometry of Mn2+ binding to concanavalin A was found to be influenced by temperature, pH, and the presence or absence of saccharide. Demetalized concanavalin A binds one Mn2+ (S1 site) at 5 degrees C, pH 6.5, and two Mn2+ at 25 degrees C (S1 and S2 sites). The association constants for Mn2+ are 6.2 x 10(5) and 3.7 x 10(4) M-1 for the S1 and S2 sites, respectively, at 25 degrees C. Concanavalin A with one Mn2+ bound per monomer remains in an open conformation and exhibits a relatively high water proton relaxation rate. Concanavalin A with two Mn2+ ions remains in a closed conformation characterized by a lower relaxation rate. The rate of binding of the second Mn2+ to concanavalin A as determined by ESR and the rate of conversion of open form to closed form (folding over) as determined by proton relaxation rate measurements gave an identical rate constant of 80.0 +/- 5.8 M-1 h-1 at 17 degrees C. Ca2+, Sr2+, and high levels of methyl alpha-D-mannopyranoside also induce folding of concanavalin A. Ca2+ is not catalytic but stoichiometric in causing the folding. Mn2+ in the S1 site can be displaced by Ni2+, Co2+, and Zn2+, and Mn2+ in the S2 site can be displaced by Ca2+ and Sr2+. Concanavalin A with Ni2+, Co2+, Zn2+, or Mn2+ in the S1 site and Ca2+ or Sr2+ in the S2 site has a higher affinity for methylumbelliferyl alpha-D-mannopyranoside than Ni-Mn-, Co-Mn-, Zn-Mn-, and Cd-Cd-concanavalin A.  相似文献   

3.
The interactions of Sepharose 4B-immobilized concanavalin A (ConA) with 10 glycoasparagines derived from ovalbumin were investigated quantitatively by frontal affinity chromatography. In this method, a carbohydrate solution is applied continuously to a ConA-Sepharose column and the retardation of the elution front is measured as a parameter of the strength of the interaction. The dissociation constant (Kd) for each saccharide with ConA can be determined. An analysis of the binding of p-nitrophenyl-alpha,D-mannoside has shown that the binding properties of ConA do not change essentially after immobilization on Sepharose 4B. Each of the ovalbumin glycoasparagines was labeled with tritium by the reductive methylation method for analysis. A comparison of the Kd values obtained showed that the binding of ConA varies considerably with very slight structural differences of the glycosyl chain. The results suggest that ConA recognizes a specific glycosyl chain structure, Man alpha 1-6(Man alpha 1-3)Man, in which at least one hydroxyl group at the C-3 position of C-6-linked mannose should be free. The glycoasparagines containing this structure bound strongly to ConA-Sepharose with dissociation constants below 3.4 X 10(-7) M.  相似文献   

4.
The metal-sugar distances in two metallized forms of concanavalin A have been compared by 19F magnetic resonance techniques. Using relaxation times measured at two different frequencies we have shown that the distance between the Mn2+ ion and the bound sugar in concanavalin A containing only Mn2+ is essentially identical to that found in concanavalin A containing both Mn2+ and Ca2+. Our results rule out the possibility that Mn2+ activates concanavalin A by binding at the Ca2+ site (S2) and would suggest that Mn2+ alone can induce an active saccharide binding conformation by binding at the transition metal site (S1).  相似文献   

5.
The structures of two forms of a recombinant flavoprotein have been determined at high resolution and compared. These proteins are (1) the flavocytochrome c p-cresol methylhydroxylase (rPCMH, 1.85 A resolution) and (2) the cytochrome-free flavoprotein subunit of rPCMH (PchF, 1.30 A resolution). A significant conformational difference is observed in a protein segment that is in contact with the re face of the isoalloxazine ring of FAD when the structure of PchF is compared to the subunit in the intact flavocytochrome. This structural change is important for optimum catalytic function of the flavoprotein, which has been shown to be dependent on the presence of the cytochrome subunit. This change results in different protein-flavin and apparently different protein-substrate interactions that have a "tuning effect" on the electronic and redox properties of bound p-cresol and the covalently bound FAD. The conformational change in the segment in the cofactor-binding site is induced by a small rearrangement in the flavoprotein-cytochrome interface region of the flavoprotein.  相似文献   

6.
The carbohydrate-recognition domain of rat serum mannose-binding protein A has been subjected to random cassette mutagenesis. Mutant domains, expressed in bacteria, were initially screened for binding to invertase-coated nitrocellulose and then analyzed further for Ca2+ affinity, saccharide binding, resistance to proteolysis, and oligomerization. The results are consistent with previous evolutionary and structural studies. Six out of seven completely inactive mutants have changes in residues directly involved in ligating Ca2+. Most changes in conserved residues which form part of the hydrophobic core characteristic of Ca(2+)-dependent (C-type) animal lectins result in decreased affinity for Ca2+, even though these residues are distant from the Ca2+ sites. Changes can be made in large portions of the surface without affecting saccharide binding. The results indicate that the precise arrangement of the regular portion of the domain containing the hydrophobic core is necessary for formation of a stable Ca(2+)-ligated structure under physiological conditions. The data also suggest that the saccharide-binding site is likely to be in close proximity to the bound Ca2+.  相似文献   

7.
A temperature-dependent conformational change occurs following the binding of only one Mn(II) to a concanavalin A monomer. This change is independent of Ca(II) near pH 7 and is characterized by an activation energy of 22.3 kcal mol?1, a value similar to that attributed to a cis-trans peptide isomerization. Two conformations have been detected in magnetic resonance experiments on solvent water protons where spin lattice relaxation times are influenced by bound Mn(II). Both conformations possess saccharide binding activity and Ca(II) stoichiometrically enhances the rate of conversion to the final, more stable conformation.  相似文献   

8.
X-ray absorption near-edge structure (XANES) spectroscopy and molecular dynamics (MD) simulations have been jointly applied to the study of the Cu,Zn superoxide dismutase from Haemophilus ducreyi (HdSOD) in interaction with the carbon monoxide molecule. The configurational flexibility of the Fe(II)-heme group, intercalated between the two subunits, has been sampled by MD simulations and included in the XANES data analysis without optimization in the structural parameter space. Our results provide an interpretation of the observed discrepancy in the Fe-heme distances as detected by extended X-ray absorption fine structure (EXAFS) spectroscopy and the classical XANES analysis, in which the structural parameters are optimized in a unique structure. Moreover, binding of the CO molecule to the heme induces a long range effect on the Cu,Zn active site, as evidenced by both MD simulations and in vitro experiments. MD simulation of the CO bound system, in fact, highlighted a structural rearrangement of the protein-protein hydrogen bond network in the region of the Cu,Zn active site, correlated with an increase in water accessibility at short distance from the copper atom. In line, in vitro experiments evidenced an increase of copper accessibility to a chelating agent when the CO molecule binds to the heme group, as compared to a heme deprived HdSOD. Altogether, our results support the hypothesis that the HdSOD is a heme-sensor protein, in which binding to small gaseous molecules modulates the enzyme superoxide activity as an adaptive response to the bacterial environment.  相似文献   

9.
The binding of the ligands Mn2+, Ca2+, and methyl alpha-D-glucopyranoside to concanavalin A, purified as described (A.J. Sophianopoulos and J.A. Sophianopoulos (1981) Prep. Biochem. 11, 413-435), was studied by ultrafiltration in 0.2 M NaCl, pH 5.2 and pH 6.5 to 7, and at 23 to 25 degrees C. The association constant (Ka) of methyl alpha-D-glucopyranoside to concanavalin A was (2 +/- 0.2) X 10(3) M-1, both at pH 5.2 and 7. At pH 5.2 and in the absence of Ca2+, the Ka of Mn2+ to concanavalin A was (5 +/- 1) X 10(3) M-1, and in the presence of 1 mM Ca2+, the Ka was (9.1 +/- 2.1) X 10(5) M-1. At pH 6.5 Mn2+ bound to concanavalin A with a Ka of (7.3 +/- 1.8) X 10(5) M-1, and the binding affinity was virtually independent of the presence of Ca2+. Experiments of binding of 4-methylumbelliferyl alpha-D-mannopyranoside to concanavalin A indicated that at pH 5.2, binding of a single Mn2+ per concanavalin A monomer was sufficient to induce a fully active saccharide binding site. Ca2+ is not necessary for such activation, but rather it increases the affinity of concanavalin A for binding Mn2+.  相似文献   

10.
Binding of concanavalin A to bleached bovine retinal rod disc membranes is governed by a single association constant of 6.4 × 107 M?1, at 26 °C. α-Methyl glucoside competes with the saccharide ligand of the membrane for the saccharide-binding site of concanavalin A. The surface density of the saccharide ligand is calculated to be one ligand per 5 × 104Å2 of membrane surface.  相似文献   

11.
Li L  Singh BR 《Biochemistry》2000,39(21):6466-6474
Clostridial botulinum neurotoxins (BoNTs) cause neuroparalysis by blocking neurotransmitter release at the neuromuscular junctions. While the toxin's heavy chain (HC) is involved in binding and internalization, the light chain (LC) acts as a unique Zn(2+)-endopeptidase against a target protein in the exocytotic docking/fusion machinery. During the translocation of the LC to the cytosol, it is exposed to the endosomal low pH. Low pH showed a dramatic change in the BoNT/A LC polypeptide folding as indicated by differential heat denaturation. Furthermore, binding of 1-anilinonaphthalenesulfonate (ANS) revealed exposure of hydrophobic domains of BoNT/A LC at low pH. Low-pH-induced structural (and by implication the endopeptidase activity) changes were completely reversible. Exposure of BoNT/A LC to low pH (4.7) did not, however, evoke the loss of Zn(2+) bound to its active site. Implications of these observations to the delivery of active BoNT/A LC to the nerve cell are discussed. We further analyzed the nature of low-pH-induced change in the polypeptide folding of BoNT/A LC by Trp fluorescence measurements. The Trp fluorescence peak was observed at 322 nm, and the two fluorescence lifetime components estimated at 2.1 ns (88%) and 0.6 ns (12%) did not change much at low pH. These observations suggested that the two Trp residues are buried and constrained in a hydrophobic environment, and it is likely that the core of the BoNT/A LC protein matrix does not participate in the low-pH-induced structural alteration. This conclusion was further supported by the near-UV circular dichroism spectra under two pH conditions.  相似文献   

12.
We have determined the crystal structure of demetallized concanavalin A, at a resolution of 3.2 Å, by molecular replacement using the known structure of native concanavalin A. Refinement of the initial model using a constraint-restraint reciprocal-space least-squares procedure caused the conventional crystallographic agreement (R) factor to decrease from 0.47 to a final value of 0.26. There are significant conformational changes in the metal-binding region involving residues Asp 19 and His24, which are substantially closer to each other than in native concanavalin A. These residues form an internal salt bridge which does not exist when the metal ions are attached to the protein. The binding site for transitionmetal ions is still intact, but the calcium site is not, since one of its two carboxylic ligands, Asp 19, is unavailable. Flexibility is observed for one of the transitionmetal ligands, Glu8, as well as for some segments of the backbone. The latter could account for the increased susceptibility of demetallizcd concanavalin A to proteolysis.  相似文献   

13.
A D Sherry  A D Newman  C G Gutz 《Biochemistry》1975,14(10):2191-2196
Divalent cadmium and lead and the trivalent lanthanides bind in the trasition metal site (S1) of concamavanlin A and induce saccharide binding to the protein in the presence of calcium. Partial activation of the protein in the presence of lanthanides alone indicates these ions bind into both transition metal (S1) and calcium sites (S2). The activity of a lanthanide-protein derivative may be increased by the addition of either calcium or a transition metal ion. The saccharide binding activity decreases in the order Zn2+ is greater than Ni2+ is greater than Co2+ is greater than Mn2+ is greater than Cd2+ reflecting the order of binding constants for these ions in the transition metal site. Like the lanthanides, divalent cadmium substitutes for both the transition metal ion and calcium ion to partially activate the protein. Divalent lead substitutes only for the transition metal ion and partially activates the protein upon addingcalcium. The data are consistent with a model in which saccharide binding activity is independent of the metal size in S1 but critically dependent upon metal size in S2.  相似文献   

14.
Two recombinant human isopeptidase T isoforms, ISOT-S and ISOT-L, differing by an insertion of 23 amino acids in ISOT-L, were previously classified as thiol proteases. Both contain one Zn2+-binding site of high-affinity, which is part of a cryptic nitrilo-triacetate-resistant pocket (site 1). A second Zn2+ site (site 2) was disclosed when both isoforms of the holoenzyme were incubated with an excess of Zn2+. The firmly bound Zn2+ of site 1 could be removed either slowly by dialysis against 1,10-phenanthroline at pH 5.5 or rapidly by treatment at pH 3.0 in the presence of 6 M urea followed by gel filtration at neutral pH. Zn2+ in site 1, but not in site 2, is essential for proteolytic activity because apoproteins were inactive. Inhibition of the catalytic activity was not due to a loss of ubiquitin binding capacity. CD spectra of both isoforms disclosed no major structural differences between the apo- and holoenzymes. The reconstitution of apoenzyme with Zn2+ under nondenaturing conditions at pH 5.5 completely restored enzymatic activity, which was indistinguishable from the reconstitution carried out in urea at pH 3.0. Thus, both human ISOTs are either thiol proteases with a local structural Zn2+ or monozinc metalloproteases that might use in catalysis a Zn2+-activated hydroxide ion polarized by Cys335.  相似文献   

15.
Phage lysozyme has catalytic activity similar to that of hen egg white lysozyme, but the amino acid sequences of the two enzymes are completely different.The binding to phage lysozyme of several saccharides including N-acetylglucosamine (GlcNAc), N-acetylmuramic acid (MurNAc) and (GlcNAc)3 have been determined crystallographically and shown to occupy the pronounced active site cleft. GlcNAc binds at a single location analogous to the C site of hen egg white lysozyme. MurNAc binds at the same site. (GlcNAc)3 clearly occupies sites B and C, but the binding in site A is ill-defined.Model building suggests that, with the enzyme in the conformation seen in the crystal structure, a saccharide in the normal chair configuration cannot be placed in site D without incurring unacceptable steric interference between sugar and protein. However, as with hen egg white lysozyme, the bad contacts can be avoided by assuming the saccharide to be in the sofa conformation. Also Asp20 in T4 lysozyme is located 3 Å from carbon C(1) of saccharide D, and is in a position to stabilize the developing positive charge on a carbonium ion intermediate. Prior genetic evidence had indicated that Asp20 is critically important for catalysis. This suggests that in phage lysozyme catalysis is promoted by a combination of steric and electronic effects, acting in concert, The enzyme shape favors the binding in site D of a saccharide with the geometry of the transition state, while Asp20 stabilizes the positive charge on the oxocarbonium ion of this intermediate. Tn phage lysozyme, the identity of the proton donor is uncertain. In contrast to hen egg white lysozyme, where Glu35 is 3 Å from the glycosidic DOE bond, and is in a non-polar environment, phage lysozyme has an ion pair, Glull … Arg145, 5 Å away from the glycosidic oxygen. Possibly Glull undergoes a conformational adjustment in the presence of bound substrate, and acts as the proton donor. Alternatively, the proton might come from a bound water molecule.  相似文献   

16.
L-N(omega),N(omega)-dimethylarginine dimethylaminohydrolase-1 (DDAH-1) is a Zn(II)-containing enzyme that, through hydrolysis of side-chain methylated l-arginines, regulates the activity of nitric-oxide synthase. Herein we report the structural and functional properties of the Zn(II)-binding site in DDAH-1 from bovine brain. Activity measurements of the native and metal-free enzyme have revealed that the endogenously bound Zn(II) inhibits the enzyme. Native DDAH-1 could be fully or partially activated using various concentrations of phosphate, imidazole, histidine, and histamine, a process that is paralleled by the release of Zn(II). The slow activation of the enzyme by the bulky complexing agents EDTA and 1,10-phenantroline suggests that the Zn(II)-binding site is partially buried in the protein structure. The apparent Zn(II)-dissociation constant of 4.2 nm, determined by 19F NMR using the chelator 5F-BAPTA (1,2-bis(2-amino-5-fluorophenoxy)ethane-N,N,N',N'-tetraacetic acid), lies in the range of intracellular free Zn(II) concentrations. These results suggest a regulatory role for the Zn(II)-binding site. The coordination environment of the Zn(II) in DDAH-1 has been examined by Zn K-edge x-ray absorption spectroscopy. The extended x-ray absorption fine structure observed is consistent with Zn(II) being coordinated by 2 S and 2 N (or O) atoms. The biological implications of these findings are discussed.  相似文献   

17.
ATP-binding cassette superfamily of periplasmic metal transporters are known to be vital for maintaining ion homeostasis in several pathogenic and non-pathogenic bacteria. We have determined crystal structure of the high-affinity zinc transporter ZnuA from Escherichia coli at 1.8 A resolution. This structure represents the first native (non-recombinant) protein structure of a periplasmic metal binding protein. ZnuA reveals numerous conformational features, which occur either in Zn(2+) or in Mn(2+) transporters, and presents a unique conformational state. A comprehensive comparison of ZnuA with other periplasmic ligand binding protein structures suggests vital mechanistic differences between bound and release states of metal transporters. The key new attributes in ZnuA include a C-domain disulfide bond, an extra alpha-helix proximal to the highly charged metal chelating mobile loop region, alternate conformations of secondary shell stabilizing residues at the metal binding site, and domain movements potentially controlled by salt bridges. Based on in-depth structural analyses of five metal binding transporters, we present here a mechanistic model termed as "partial domain slippage" for binding and release of Zn(2+).  相似文献   

18.
Co(III) has been stoichiometrically incorporated into jack bean concanavalin A. The Co(III) protein still possesses a binding site for an additional divalent transition metal ion which together with Ca(II) can induce the sugar binding ability. No H2O2 oxidation of Co(II) occurs with demetallized concanavalin A activated with Ca(II) and Co(II) unless Co(II) is present in a stoichiometric excess. Evidence is presented to indicate that kinetically stable Co(III) is bound to a completely different location than the thermodynamically stable Co(II) protein site.  相似文献   

19.
The association of thrombin with thrombomodulin, a non-enzymatic endothelial cell surface receptor, alters the substrate specificity of thrombin. Complex formation converts thrombin from a procoagulant to an anticoagulant enzyme. Structure-function analysis of this change in specificity is facilitated by the availability of two soluble proteolytic derivatives of thrombomodulin, one consisting of the six repeated growth factor-like domains of thrombomodulin (GF1-6) and the other containing only the fifth and sixth such domains (GF5-6). Both derivatives can bind to thrombin and block fibrinogen clotting activity, though only the larger GF1-6 can stimulate the activation of protein C. To ascertain whether the substrate specificity change from fibrinogen to protein C is accompanied by structural changes in the active site of the enzyme, fluorescent dyes were positioned at different locations within the active site. A 5-dimethylaminonaphthalene-1-sulfonyl (dansyl) dye was covalently attached to the active site serine to form dansyl-thrombin, while either a fluorescein or an anilinonaphthalene-6-sulfonic acid (ANS) dye was attached covalently to the active site histidine of thrombin via a D-Phe-Pro-Arg linkage. The environment of the dansyl dye was altered in a similar fashion when either GF1-6 or GF5-6 bound to thrombin, since a similar reduction in dansyl emission intensity was elicited by these two thrombomodulin derivatives (25 and 32%, respectively). These spectral changes, and all others in this study, were saturable and reached a maximum when the ratio of thrombomodulin derivative to thrombin was close to 1. The environments of the fluorescein and ANS dyes were also altered when GF1-6 bound to thrombin because binding resulted in emission intensity changes of -13% and +18%, respectively. In contrast, no fluorescence changes were observed when the fluorescein and ANS thrombin derivatives were titrated with GF5-6. Thus, the structure of the active site was altered by thrombomodulin both immediately adjacent to the active site serine and also more than 15 A away from it. However, the structural change far from Ser-195 was only elicited by thrombomodulin species that stimulate thrombin-dependent activation of protein C.  相似文献   

20.
Protein engineering was used previously to convert maltose-binding protein (MBP) into a zinc biosensor. Zn(2+) binding by the engineered MBP was thought to require a large conformational change from "open" to "closed", similar to that observed when maltose is bound by the wild-type protein. We show that although this re-designed MBP molecule binds Zn(2+) with high affinity as previously reported, it does not adopt a closed conformation in solution as assessed by small-angle X-ray scattering. High-resolution crystallographic studies of the engineered Zn(2+)-binding MBP molecule demonstrate that Zn(2+) is coordinated by residues on the N-terminal lobe only, and therefore Zn(2+) binding does not require the protein to adopt a fully closed conformation. Additional crystallographic studies indicate that this unexpected Zn(2+) binding site can also coordinate Cu(2+) and Ni(2+) with only subtle changes in the overall conformation of the protein. This work illustrates that the energetic barrier to domain closure, which normally functions to maintain MBP in an open concentration in the absence of ligand, is not easily overcome by protein design. A comparison to the mechanism of maltose-induced domain rearrangement is discussed.  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