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.
Concanavalin A (Con A) was labeled with p-azidophenyl alpha-D-mannopyranoside under ultraviolet irradiation and the reaction products were separated by affinity chromatography on Sephadex G-100 at pH 5. One of the Con A derivatives thus obtained was characterized as a monovalent dimer at pH 5 and a divalent tetramer at pH 7 by sedimentation equilibrium and equilibrium dialysis, indicating that this photoaffinity labeling did not alter the quaternary structure of Con A. In agreement with these results, the labeled Con A did not show the capacity to precipitate glycogen at pH 5, but it formed precipitates with glycogen at pH 7. Although its hemagglutinating activity was found to be weaker than that of the native Con A, the dose-response cure of the labeled Con A in the mitogenic stimulation of human peripheral lymphocytes was almost identical to that of the native con A.  相似文献   

2.
Binding of a series of synthetic multivalent carbohydrate analogs to the Man/Glc-specific lectins concanavalin A and Dioclea grandiflora lectin was investigated by isothermal titration microcalorimetry. Dimeric analogs possessing terminal alpha-D-mannopyranoside residues, and di-, tri-, and tetrameric analogs possessing terminal 3, 6-di-O-(alpha-D-mannopyranosyl)-alpha-D-mannopyranoside residues, which is the core trimannoside of asparagine-linked carbohydrates, were selected in order to compare the effects of low and high affinity analogs, respectively. Experimental conditions were found that prevented precipitation of the carbohydrate-lectin cross-linked complexes during the isothermal titration microcalorimetry experiments. The results show that the value of n, the number of binding sites on each monomer of the lectins, is inversely proportional to the number of binding epitopes (valency) of each carbohydrate. Hence, n values close to 1.0, 0.50, and 0.25 were observed for the binding of mono-, di-, and tetravalent sugars, respectively, to the two lectins. Importantly, differences in the functional valency of a triantennary analog for concanavalin A and D. grandiflora lectin are observed. The enthalpy of binding, DeltaH, is observed to be directly proportional to the number of binding epitopes in the higher affinity analogs. For example, DeltaH of a tetravalent trimannoside analog is nearly four times greater than that of the corresponding monovalent analog. Increases in K(a) values of the multivalent carbohydrates relative to monovalent analogs, known as the "multivalency effect," are shown to be due to more positive entropy (TDeltaS) contributions to binding of the former sugars. A general thermodynamic model for distinguishing binding of multivalent ligands to a single receptor with multiple, equal subsites versus binding to separate receptor molecules is given.  相似文献   

3.
A major molecular species of concanavalin A (Con A), a mitogenic lectin from jack bean seeds, has a quaternary structure composed of four homologous subunits and a tetravalent sugar-binding ability. We show that the tetrameric Con A can be converted into a monovalent monomeric form by either photochemical alkylation or hydrogen peroxide/dioxane oxidation of about two tryptophan residues. A divalent dimeric derivative of Con A is also prepared by sulfomethylamidation of about four carboxyl groups. Chemical properties and mitogenic and hemagglutinating activities of these new Con A derivates are compared with those of the tetravalent Con A, as well as of the Con A derivatives that have appeared in the literature on cell biological studies. The significance of the lectin valences in lymphocyte activation and hemagglutination is also discussed.  相似文献   

4.
New spin labeled derivatives of phosphorylcholine have been synthesized. The compounds cause reversible inhibition of the precipitation reactions between pneumococcal C-polysaccharide and the C-reactive proteins from humans, dogfish sharks (Mustelus canis), and horseshoe crabs (Limulus polyphemus). The spin labeled phosphorylcholine derivatives also rival phosphorylcholine as a ligand for the human, dogfish, and Limulus C-reactive proteins. The interactions of the purified C-reactive proteins with the spin labeled derivatives of phosphorylcholine have been studied using electron spin resonance spectrometry. The dramatic decrease in the ESR signal of some of the spin labels is due to immobilization of the label. Only the well known phosphate spin label, 4-phosphate-2,2,6,6-tetramethylpiperidine-1-oxyl could be used for binding studies on human and Limulus C-reactive proteins. Thus, by Scatchard analysis, the human C-reactive protein bound 1 mol of phosphate spin label per mol of protein with a Ka = 3.91 X 10(3) M-1, whereas the Limulus C-reactive protein bound only 0.5 mol of phosphate spin label per mol of protein with an overall Ka = 1.95 X 10(3) M-1. Inhibition studies using purified C-polysaccharide-induced inhibition of the phosphate spin label-human C-reactive protein interaction showed competitive inhibition with a KI of 4.78 X 10(-5) M at 18 degrees C. The phosphate spin label did not bind to dogfish C-reactive protein. However, one new phosphorylcholine spin label did bind and was used for Scatchard and Hill plot analyses. The dogfish C-reactive protein, which exists as a Mr = 50,000 dimer, bound 2 mol of the phosphorylcholine spin label per mol of protein, and this binding exhibited negative cooperativity as indicated by a Hill coefficient of 0.75.  相似文献   

5.
A model is developed for ionic conduction in the sheep cardiac sarcoplasmic reticulum ryanodine receptor channel based on Eyring rate theory. A simple scheme is proposed founded on single-ion occupancy and an energy profile with four barriers and three binding sites. The model is able to quantitatively predict a large number of conduction properties of the purified and native receptor with monovalent and divalent cations as permeant species. It suggests that discrimination between divalent and monovalent cations is due to a high affinity central binding site and a process that favors the passage of divalent cations between binding sites. Furthermore, differences in conductance among the group Ia cations and among the alkaline earths are largely explained by differing affinity at this putative central binding site.  相似文献   

6.
The binding isotherms of Mn2+ to bovine plasma protein C (PC), des(1-41)-light chain protein C (GDPC), and activated GDPC (GDAPC) have been measured. PC contains 14-16 total Mn2+ binding sites, a value that is reduced to approximately 7-8 in the presence of NaCl. The average Kd of the latter sites is 230 +/- 30 microM. Upon removal of a 41-residue peptide from the amino terminus of the light chain of PC, and, concomitantly, all of the gamma-carboxyglutamic acid residues, the resulting protein, GDPC, possesses a single Mn2+ site of Kd = 120 +/- 20 microM. Activation of GDPC to GDAPC results in a slight lowering of the Kd for the single Mn2+ binding site to 53 +/- 8 microM, a value that is essentially unchanged in the presence of monovalent cations, a competitive inhibitor of the enzyme, or an active site directed affinity label. The Mn2+ on GDAPC is displaced by Ca2+, suggesting that the protein binding site for these two divalent cations is the same. These studies establish that Mn2+ is a suitable spectroscopic probe for the Ca2+ binding site of GDAPC, and that the divalent cation site is separate from the monovalent cation site(s) and the active site of the enzyme.  相似文献   

7.
Stimulation of polymorphonuclear leukocytes (PMN) by tetravalent concanavalin A (alpha-ConA) induces membrane depolarization preceding the onset of superoxide anion (O2-) production. Both divalent and monovalent ConA analogues were studied to evaluate the role of valence. Monovalent ConA (m-ConA) was inactive in stimulating O2- production and divalent derivatives were less active than native alpha-ConA. Similarly, membrane depolarization was dependent on the valency of ConA. m-ConA did not induce a marked change in membrane potential, whereas sustained depolarization occurred with multivalent ConA. The formation of multiple linked interactions between surface receptors may be an important early event in the activation of PMN by ConA.  相似文献   

8.
Concanavalin A (Con A) and agglutinins from the pea (PSA), lentil (LCH), and fava bean (VFA) constitute a group of D-mannose/D-glucose binding legume lectins. In addition to their sugar binding specificity, these lectins also contain sites that bind hydrophobic ligands. The present study explores a class of nonpolar binding sites reportedly present adjacent to the carbohydrate binding site in PSA, LCH, and VFA. A series of 2-O- and 3-O-substituted nitrobenzoyl and nitrobenzyl derivatives of methyl alpha-D-glucopyranoside and methyl alpha-D-mannopyranoside were synthesized. Evaluation of their binding to Con A, PSA, LCH, and VFA was carried out by the technique of hapten inhibition of precipitation reaction. The hapten inhibition assay results reveal that the presence of a methyl or methylene group at the O-2 or O-3 position of the sugar is essential for hydrophobic interaction with PSA, LCH, and VFA. The substitution of methyl by nitrobenzyl leads to enhanced binding (1.7-16.7 times for the 2-O-substituted compounds and 7.9-40.5 times for the 3-O-substituted compounds) with the m-nitrobenzyl group contributing to maximum binding. A hydrophobic interaction is also involved between Con A and 2-O-nitrobenzyl derivatives, resulting in enhanced binding, but the corresponding 3-O-isomers bind poorly due probably to steric reasons. These results may be rationalized on the basis of the recently published X-ray data of Con A and VFA. The nitrobenzyl derivatives, after transformation to their azido analogs, have potential applications in the photoaffinity labeling of these lectins.  相似文献   

9.
The polyamide N-formamido imidazole-pyrrole-imidazole (f-ImPyIm) binds with an exceptionally high affinity for its cognate site 5'-ACGCGT-3' as a stacked, staggered, and noncovalent cooperative dimer. Investigations are presented into its sequence specificity and binding affinity when linked covalently as an H-pin "dimer". Five f-ImPyIm cross-linked analogues with six to nine methylene linkers and an eight-linked ethylene glycol linker were examined to investigate the effect of linkage and linker length on DNA binding. Thermal denaturation studies on short DNA hairpins showed preferential binding by both f-ImPyIm (DeltaTm = 7.8 degrees C) and its cross-linked derivatives (DeltaTm > 30 degrees C) at 5'-ACGCGT-3', indicating sequence specificity was retained on linkage. DNase I footprinting confirmed strict cognate site selectivity and demonstrated that affinity increased with linker length (f-ImPyIm-9 = f-ImPyIm-8 = f-ImPyIm-EG-8 > f-ImPyIm-7 > f-ImPyIm-6). The eight- and nine-linked derivatives bound at 100-fold lower concentrations at the cognate site relative to f-ImPyIm-6, and with 10-fold higher affinity than unlinked f-ImPyIm. Use of an ethylene glycol linkage in f-ImPyIm-EG-8 to improve solubility slightly increased the cognate site affinity relative to those of f-ImPyIm-8 and f-ImPyIm-9, although some selectivity was lost at high ligand concentration. CD demonstrated that cognate site binding by eight and nine-linked compounds occurred in the minor groove. SPR analysis gave a binding affinity (K) for f-ImPyIm-EG-8 at the cognate site of 2 x 10(10) M-1, representing a 100-fold increase relative to that of f-ImPyIm. This study demonstrates that the high-affinity cooperative binding of f-ImPyIm can be enhanced significantly by suitable covalent linkage, while maintaining its strict cognate site selectivity.  相似文献   

10.
Competitive adsorption to a negative surface between monovalent and divalent cations is studied in a closed system. A self-consistent theory is presented for the cases when the divalent cation binds to two negative sites (1:2 binding) and to one negative site (1:1 binding). It is demonstrated that these two cases are distinguishable when the relative difference in total divalent concentrations determined at fixed bulk divalent concentrations is plotted as a function of the bulk monovalent concentration. The 1:2 binding case yields a sigmoidal curve while the 1:1 binding curve is hyperbolic. The comparison between the 1:2 and 1:1 binding cases of the divalent cation is extended to include; (1) the existence of a surface charge to which no binding occurs, and (2) the case when an additional non-binding monovalent cation is present.  相似文献   

11.
Li Y  Zhang X  Chu S  Yu K  Guan H 《Carbohydrate research》2004,339(4):873-879
The Ugi four-component reaction (U-4CR) was utilized to prepare divalent and trivalent cluster mannosides with different scaffolds. The glycoclusters obtained were tested for their relative inhibitory potency against the binding of yeast mannan to concanavalin A by solid-phase enzyme-linked lectin assays (ELLA) using methyl alpha-D-mannopyranoside as a standard. Among them, a divalent mannoside containing aromatic groups showed the strongest binding affinity to concanavalin A.  相似文献   

12.
Alkaline conditions are generally preferred for sanitization of chromatography media by cleaning-in-place (CIP) protocols in industrial biopharmaceutical processes. The use of such rigorous conditions places stringent demands on the stability of ligands intended for use in affinity chromatography. Here, we describe efforts to meet these requirements for a divalent proteinaceous human serum albumin (HSA) binding ligand, denoted ABD*dimer. The ABD*dimer ligand was constructed by genetic head-to-tail linkage of two copies of the ABD* moiety, which is a monovalent and alkali-stabilized variant of one of the serum albumin-binding motifs of streptococcal protein G. Dimerization was performed to investigate whether a higher HSA-binding capacity could be obtained by ligand multimerization. We also investigated the influence on alkaline stability and HSA-binding capacity of three variants (VDANS, VDADS and GGGSG) of the inter-domain linker. Biosensor binding studies showed that divalent ligands coupled using non-directed chemistry demonstrate an increased molar HSA-binding capacity compared with monovalent ligands. In contrast, equal molar binding capacities were observed for both types of ligands when using directed ligand coupling chemistry involving the introduction and recruitment of a unique C-terminal cysteine residue. Significantly higher molar binding capacities were also detected when using the directed coupling chemistry. These results were confirmed in affinity chromatography binding capacity experiments, using resins containing thiol-coupled ligands. Interestingly, column sanitization studies involving exposure to 0.1 M NaOH solution (pH 13) showed that of all the tested constructs, including the monovalent ligand, the divalent ligand construct containing the VDADS linker sequence was the most stable, retaining 95% of its binding capacity after 7 h of alkaline treatment.  相似文献   

13.
S-Adenosylmethionine synthetase from Escherichia coli   总被引:16,自引:0,他引:16  
Adenosylmethionine (AdoMet) synthetase has been purified to homogeneity from Escherichia coli. For this purification, a strain of E. coli which was derepressed for AdoMet synthetase and which harbors a plasmid containing the structural gene for AdoMet synthetase was constructed. This strain produces 80-fold more AdoMet synthetase than a wild type E. coli. AdoMet synthetase has a molecular weight of 180,000 and is composed of four identical subunits. In addition to the synthetase reaction, the purified enzyme catalyzes a tripolyphosphatase reaction that is stimulated by AdoMet. Both enzymatic activities require a divalent metal ion and are markedly stimulated by certain monovalent cations. AdoMet synthesis also takes place if adenyl-5'yl imidodiphosphate (AMP-PNP) is substituted for ATP. The imidotriphosphate (PPNP) formed is not hydrolyzed, permitting dissociation of AdoMet formation from tripolyphosphate cleavage. An enzyme complex is formed which contains one equivalent (per subunit) of adenosylmethionine, monovalent cation, imidotriphosphate, and presumably divalent cation(s). The rate of product dissociation from this complex is 3 orders of magnitude slower than the rate of AdoMet formation from ATP. Studies with the phosphorothioate derivatives of ATP (ATP alpha S and ATP beta S) in the presence of Mg2+, Mn2+, or Co2+ indicate that a divalent ion is bound to the nucleotide during the reaction and provide information on the stereochemistry of the metal-nucleotide binding site.  相似文献   

14.
Summary (1) When salts are added to buffered suspensions of membrane fragments containing the fluorochrome 1-anilino-8-naphthalenesulfonate (ANS), there is an increased fluorescence. This is caused by increased binding of the fluorochrome; the intrinsic fluorescence characteristics of the bound dye remain unaltered. These properties make ANS a sensitive and versatile indicator of ion association equilibria with membranes. (2) Alkali metal and alkylammonium cations bind to membranes in a unique manner. Cs+ binds most strongly to rat brain microsomal material, with the other alkali metals in the order Cs+>Rb+>K+>Na+>Li+. The reaction is endothermic and entropy driven. Monovalent cations are displaced by other monovalent cations. Divalent cations and some drugs (e. g., cocaine) displace monovalent cations more strongly. (3) Divalent cations bind to membranes (and to lecithin micelles) at four distinct sites, having apparent association constants between 50 and 0.2mm –1. The characteristics of the titration suggest that only one species of binding site is present at any one time, and open the possibility that structural transitions of the unassociated coordination sites may be induced by divalent cation binding. Divalent cation binding at the weakest site (like monovalent cation binding) is endothermic and entropy driven. At the next stronger site, the reaction is exothermic. Monovalent cations affect divalent cation binding by reducing the activity coefficient: they do not appear to displace divalent cations from their binding sites.  相似文献   

15.
The structure of a ribonuclease III processing signal from bacteriophage T7 was examined by NMR spectroscopy, optical melting, and chemical and enzymatic modification. A 41 nucleotide variant of the T7 R1.1 processing signal has two Watson-Crick base-paired helices separated by an internal loop, consistent with its predicted secondary structure. The internal loop is neither rigidly structured nor completely exposed to solvent, and seems to be helical. The secondary structure of R1.1 RNA is largely insensitive to the monovalent cation concentration, which suggests that the monovalent cation sensitivity of secondary site cleavage by RNase III is not due to a low salt-induced RNA conformational change. However, spectroscopic data show that Mg2+ affects the conformation of the internal loop, suggesting a divalent cation binding site(s) within this region. The Mg(2+)-dependence of RNase III processing of some substrates may reflect not only a requirement for a divalent cation as a catalytic cofactor, but also a requirement for a local RNA conformation which is divalent cation-stabilized.  相似文献   

16.
Glutamine synthetase (GS) catalyzes the ATP-dependent condensation of ammonia and glutamate to yield glutamine, ADP, and inorganic phosphate in the presence of divalent cations. Bacterial GS is an enzyme of 12 identical subunits, arranged in two rings of 6, with the active site between each pair of subunits in a ring. In earlier work, we have reported the locations within the funnel-shaped active site of the substrates glutamate and ATP and of the two divalent cations, but the site for ammonia (or ammonium) has remained elusive. Here we report the discovery by X-ray crystallography of a binding site on GS for monovalent cations, Tl+ and Cs+, which is probably the binding site for the substrate ammonium ion. Fourier difference maps show the following. (1) Tl+ and Cs+ bind at essentially the same site, with ligands being Glu 212, Tyr 179, Asp 50', Ser 53' of the adjacent subunit, and the substrate glutamate. From its position adjacent to the substrate glutamate and the cofactor ADP, we propose that this monovalent cation site is the substrate ammonium ion binding site. This proposal is supported by enzyme kinetics. Our kinetic measurements show that Tl+, Cs+, and NH4+ are competitive inhibitors to NH2OH in the gamma-glutamyl transfer reaction. (2) GS is a trimetallic enzyme containing two divalent cation sites (n1, n2) and one monovalent cation site per subunit. These three closely spaced ions are all at the active site: the distance between n1 and n2 is 6 A, between n1 and Tl+ is 4 A, and between n2 and Tl+ is 7 A. Glu 212 and the substrate glutamate are bridging ligands for the n1 ion and Tl+. (3) The presence of a monovalent cation in this site may enhance the structural stability of GS, because of its effect of balancing the negative charges of the substrate glutamate and its ligands and because of strengthening the "side-to-side" intersubunit interaction through the cation-protein bonding. (4) The presence of the cofactor ADP increases the Tl+ binding to GS because ADP binding induces movement of Asp 50' toward this monovalent cation site, essentially forming the site. This observation supports a two-step mechanism with ordered substrate binding: ATP first binds to GS, then Glu binds and attacks ATP to form gamma-glutamyl phosphate and ADP, which complete the ammonium binding site. The third substrate, an ammonium ion, then binds to GS, and then loses a proton to form the more active species ammonia, which attacks the gamma-glutamyl phosphate to yield Gln. (5) Because the products (Glu or Gln) of the reactions catalyzed by GS are determined by the molecule (water or ammonium) attacking the intermediate gamma-glutamyl phosphate, this negatively charged ammonium binding pocket has been designed naturally for high affinity of ammonium to GS, permitting glutamine synthesis to proceed in aqueous solution.  相似文献   

17.
Three photoreactive substrate analogues, o-azidophenethyl pyrophosphate, p-azidophenethyl pyrophosphate, and 3-azido-1-butyl pyrophosphate, have been synthesized as site-directed probes to label the catalytic site of prenyltransferase. Due to the relatively poor affinity of p-azidophenethyl pyrophosphate and 3-azido-1-butyl pyrophosphate for the enzyme, only o-azidophenethyl pyrophosphate (aryl azide) was utilized for photoaffinity labeling. This aryl azide has a UV absorption maximum at 250 nm. In the absence of activating light, binding studies demonstrate that the o-aryl azide competes for binding with both the natural substrates, isopentenyl pyrophosphate and geranyl pyrophosphate. More than 90% enzymatic activity is lost when enzyme is irradiated in the presence of the aryl azide as compared to irradiation in the absence of the azide, and the protein loses its capacity for substrate binding in direct proportion to photolabeling. A stoichiometry of 2 mol of affinity label covalently bound per mol of enzyme dimer was established with [1-3H]-o-azidophenethyl pyrophosphate. Since there are two catalytic sites per enzyme dimer, the o-aryl azide appears specifically to label the enzyme at its catalytic sites. Additional evidence that the reagent was specific for the catalytic site came from the observation that farnesyl pyrophosphate afforded complete protection against photoinactivation, while isopentenyl pyrophosphate provided partial protection. Gel isoelectric focusing verified this stoichiometry and indicated that the labeled enzyme has a more acidic isoelectric point than the native enzyme.  相似文献   

18.
Three derivatives of ouabain have been synthesized which alkylate the digitalis receptor. These derivatives were formed through reductive amination of p-nitrophenyltriazene (NPT) ethylenediamine to the periodate-oxidized rhamnose moiety of ouabain. The non-covalent binding of the ouabain derivatives (NPT-ouabain, designated I, II, and III) was followed (i) by their ability to inhibit the activity of sodium- and potassium-activated ATPase ((Na+,K+)-ATPase) purified from the electric organ of Electrophorus electricus, (ii) by the binding of [3H]NPT-ouabain I to the enzyme, and (iii) by the inhibition of [3H]ouabain binding with unlabeled NPT-ouabain I. Covalent modification of the digitalis site of (Na+,K+)-ATPase occurs after long periods of time. At pH 7.5 (25 degrees C) the best alkylating derivative, NPT-ouabain I, gives maximum covalent labeling after 6 h. Only the large polypeptide chain (Mr = 93,000) of the purified enzyme is specifically labeled with [3H]NPT-ouabain I while the glycoprotein chain (Mr = 47,000) is not significantly labeled. Labeling of a microsomal fraction of the electric organ with [3H]NPT-ouabain I gave the same type of gel pattern as that observed with the purified enzyme. [3H]NPT-ouabain I was also used to label the digitalis receptor in highly purified axonal membranes and in cardiac membranes prepared from embryonic chick heart. Although the (Na+,K+)-ATPase in both types of membranes has a low affinity for ouabain, [3H]NPT-ouabain I proved to be a very efficient affinity label for the digitalis receptor. In the complex mixture of polypeptides found in these membrane preparations, only a single polypeptide chain having a Mr = 93,000 is specifically labeled by [3H]NPT-ouabain I.  相似文献   

19.
The prenyl transferase undecaprenyl pyrophosphate synthetase was partially purified from the cytosolic fraction of Escherichia coli. Its enzymic products were characterized as a family of cis-polyprenyl phosphates, which ranged in carbon number from C55 to C25. The enzyme is constituted of two subunits of approximately 30,000 molecular weight. A radiolabeled photolabile analogue of t,t-farnesyl pyrophosphate, [3H]2-diazo-3-trifluoropropionyloxy geranyl pyrophosphate, was shown to label Lactobacillus plantarum and E. coli undecaprenyl pyrophosphate synthetase on UV irradiation in the presence of isopentenyl pyrophosphate and divalent cation. The only labeled polypeptide migrated on electrophoresis in a sodium dodecyl sulfate-polyacrylamide gel at a molecular weight of approximately 30,000. No protein was radiolabeled when the natural substrate, t,t-farnesyl pyrophosphate was included in the irradiation mixture. Irradiation in the presence of MgCl2 without isopentenyl pyrophosphate gave less labeling of the polypeptide. Irradiation with only isopentenyl pyrophosphate gave little labeling of the polypeptide. When the enzyme was irradiated with 3H-photoprobe, [14C]isopentenyl pyrophosphate, and MgCl2, the labeled polypeptide gave a ratio of 14C/3H that indicated the product must also bind to the enzyme on irradiation. These results demonstrate the ability to radiolabel the allylic pyrophosphate binding site and possibly product binding site of undecaprenyl pyrophosphate synthetase by a process which is favored when both cosubstrate and divalent cation are present.  相似文献   

20.
Bacterial cell division occurs through the formation of a protein ring (division ring) at the site of division, with FtsZ being its main component in most bacteria. FtsZ is the prokaryotic ortholog of eukaryotic tubulin; it shares GTPase activity properties and the ability to polymerize in vitro. To study the mechanism of action of FtsZ, we used molecular dynamics simulations of the behavior of the FtsZ dimer in the presence of GTP-Mg2+ and monovalent cations. The presence of a K+ ion at the GTP binding site allows the positioning of one water molecule that interacts with catalytic residues Asp235 and Asp238, which are also involved in the coordination sphere of K+. This arrangement might favor dimer stability and GTP hydrolysis. Contrary to this, Na+ destabilizes the dimer and does not allow the positioning of the catalytic water molecule. Protonation of the GTP gamma-phosphate, simulating low pH, excludes both monovalent cations and the catalytic water molecule from the GTP binding site and stabilizes the dimer. These molecular dynamics predictions were contrasted experimentally by analyzing the GTPase and polymerization activities of purified Methanococcus jannaschii and Escherichia coli FtsZ proteins in vitro.  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