首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 765 毫秒
1.
We have used ethylation protection experiments and modification interference using phosphorothioate nucleosides to identify phosphate groups involved in the magnesium-dependent tertiary structure and function of the VS ribozyme, a small, self-cleaving RNA. Phosphorothioate interference-rescue experiments in the presence of the thiophilic manganese ion implicate four phosphate groups in direct metal ion binding. Phosphorothioate substitution also creates a new manganese binding site that increases the cis cleavage rate of the ribozyme, possibly by disrupting an inhibitory structure. Interpreting these data in the context of a recently developed structural model shows that almost all of the important phosphate groups are located in the central core of the ribozyme. The model suggests roles for certain phosphate groups in particular steps of RNA folding and identifies a candidate region for the active site of the ribozyme.  相似文献   

2.
Phosphatidylinositol-specific phospholipase C (PI-PLC) from Bacillus thuringiensis is an allosteric enzyme with both a phospholipid activator site and an active site. The activation of PI-PLC enzyme is optimal with phosphatidylcholine (PC) binding to the activator site and anchoring the enzyme to the interface [Zhou, C., et al. (1997) Biochemistry 36, 347-355; Zhou, C., et al. (1997) Biochemistry 36, 10089-10091]. In contrast to PC, anionic short-chain phospholipids with smaller headgroups [phosphatidylmethanol (PMe) and phosphatidic acid (PA)] as well as phosphatidylglycerol (PG) can bind to both sites playing dual roles: nonessential activation and competitive inhibition of cyclic-(1, 2)-inositol phosphate hydrolysis. PG is also a substrate, albeit a poor one, for PI-PLC, and is cleaved slowly to form alpha-glycerol phosphate. Analysis of enzyme kinetics using cIP as the substrate coupled with effects of different short-chain phospholipids on enzyme intrinsic fluorescence indicates that anionic phospholipids with small headgroups bind to the two sites with different affinities. If no interface is present, all dihexanoylphospholipids bind to the activator site more strongly than to the active site. When the activator site is occupied, it is likely that the enzyme undergoes a conformational change that allows phospholipids to bind easily to the active site. Such behavior is consistent with the observation that enzyme activation is detected at low short-chain anionic phospholipid concentrations with inhibition observed at higher concentrations, and that only inhibition is seen with these phospholipids added as monomers in the presence of a PC interface that optimally activates the PI-PLC. A kinetic model is used to extract the affinity of short-chain lipids for the active site from experimental data.  相似文献   

3.
Soluble inorganic pyrophosphatase (PPase) is one of the better understood phosphoryl-transfer enzymes and is distinctive in having four divalent metal ions at the active site. Here we determine pH profiles for wild-type Saccharomyces cerevisiae PPase (Y-PPase) and for 14 of its active site variants and consider the effects of active site mutation on the pH-independent parameters and acid dissociation constants that characterize these profiles against the framework of the proposed structure of the activated complex. The results obtained (a) support the current mechanistic model in which a hydroxide ion, stabilized by binding to two metal ions at the active site and by an extended system of hydrogen bonds within the active site, is the nucleophile that attacks enzyme-bound inorganic pyrophosphate and (b) provide evidence that the acid group that is necessary for maximal activity is a water molecule coordinated to a third metal ion, as shown by the general rise in the pKa of this group that is a consequence of almost all of the mutations. We further compare the present results to those previously observed for the corresponding mutations in Escherichia coli PPase [E-PPase; Salminen et al. (1995) Biochemistry 34, 782-791]. Such comparison provides a measure of the extent to which different portions of the active site are conserved. We find that some corresponding mutations have different effects on catalytic function, demonstrating that even in the context of very similar active sites, interactions of the mutated site with less well conserved portions of the enzyme, in this case outside the active site, can lead to different outcomes. On the other hand, one region of the active site is highly conserved, suggesting that it may represent a common feature of phosphoryl-transfer enzymes or a vestige of a primitive ur-PPase active site.  相似文献   

4.
Dopamine beta-monooxygenase requires copper ions for catalytic activity. The stoichiometry of copper activation has been a matter of discussion, but most of the recent literature agrees on a model with two copper ions per active site. We have now reinvestigated this problem with kinetic experiments at high and low protein levels. The apoenzyme (metal free) is rapidly activated by adding copper. Incremental addition of copper to high levels (up to 10 microM subunits) of enzyme raised the catalytic activity until the stoichiometric relationship between copper and enzyme subunits was 1:1. No increase in activity was observed upon addition of copper in excess of this up to levels of 3 Cu/subunit. Experiments at low protein levels (0.12 microM subunits) revealed that copper activation is described by a hyperbolic, Michaelis-Menten-type curve. This is to be expected for the 1 Cu/subunit model, whereas the 2/1 model predicts sigmoid curves. With an incremental addition of apoenzyme (high level) to a fixed level of copper, a sharp break was again observed at 1 Cu/subunit, and excess apoenzyme showed no evidence of the inhibition predicted by the 2 Cu/subunit model. Steady-state kinetics experiments with variation of the concentrations of copper and the three substrates supported an equilibrium-ordered mechanism, whereby a single activating copper ion is trapped in the active site by the substrates. Treatment of enzyme containing more than 1 Cu/subunit [both the Cu(I) and Cu(II) states were examined] with a chelating column resulted in loss of all copper in excess of 1 Cu/subunit. Reactivation of apoenzyme by vanadyl ions was studied, both as dopamine beta-monooxygenase-catalyzed electron transfer and hydroxylation. The maximal velocity with vanadium was 70% of that with copper, and the activation curve was clearly hyperbolic, again supporting the requirement of only one metal ion per active site. In conclusion, our results support the view that the copper ions bound to dopamine beta-monooxygenase in excess of 1 Cu/subunit are not required for catalysis.  相似文献   

5.
Different pathways of the metal-induced isomerization of D-xylose to D-xylulose are investigated and compared in detail using energy minimization and molecular dynamics simulation. Two theoretical models are constructed for the reaction: in vacuum and in the enzyme D-xylose isomerase. The vacuum model is constructed based on the X-ray structure of the active site of D-xylose isomerase. It contains the atoms directly involved in the reaction and is studied using a semi-empirical molecular orbital method (PM3). The model in the enzyme includes the effects of the enzyme environment on the reaction using a combined quantum mechanical and molecular mechanical potential. For both models, the structures of the reactants, products, and intermediate complexes along the isomerization pathway are optimized. The effects of the position of the "catalytic Mg2+ ion" on the energies of the reactions are studied. The results indicate: 1) in vacuum, the isomerization reaction is favored when the catalytic metal cation is at site A, which is remote from the substrate; 2) in the enzyme, the catalytic metal cation, starting from site A, moves and stays at site B, which is close to the substrate; analysis of the charge redistribution of the active site during the catalytic process shows that the metal ion acts as a Lewis acid to polarize the substrate and catalyze the hydride shift; these results are consistent with previous experimental observations; and 3) Lys183 plays an important role in the isomerization reaction. The epsilon-NH3+ group of its side chain can provide a proton to the carboxide ion of the substrate to form a hydroxyl group after the hydride shift step. This role of Lys183 has not been suggested before. Based on our calculations, we believe that this is a reasonable mechanism and consistent with site-directed mutation experiments.  相似文献   

6.
The responses of the canine colonic epithelium to histamine are potentiated by O-alkylhydroxylamines. A study of a series of such compounds suggested that active compounds had the structure R-O-NH2, substitution of a nitrogen led to total loss of activity. The locus of the potentiation effect was traced to the inhibition of diamine oxidase. A new series of aliphatic and aromatic O-alkylhydroxylamines were synthesized to explore further the structure-activity relations of this effect. The potentiating effects of these compounds were determined by examining the changes in short circuit current (Isc) produced by histamine and from the activity of a soluble preparation of diamine oxidase. We found that 1) branched compounds are less active than their straight chain counterparts, 2) greater steric bulk of the aliphatic substituent decreased activity, 3) the presence of a double bond had no significant effect though a triple bond reduced activity, 4) longer straight chain compounds were less active than the shorter chain derivatives and 5) all benzylic compounds were less active than the straight chain aliphatics. O-1-benzyl was inactive however the meta or para oxygen substituted compounds as well as the O-(1-E-Cinnamyl) derivative were active. A current model for the action of diamine oxidase proposes a crucial role for a trihydroxyphenylalanine quinone cofactor as part of the active site together with a copper atom. Using molecular modeling based on our inhibition data we are able to define the region of space that is just beyond the reactive carbonyl of the trihydroxyphenylalanine residue at the active site of diamine oxidase. We suggest that a negatively charged species, such as an aspartate or a glutamate, resides in a trough about 7 to 8 A from the trihydroxyphenylalanine carbonyl carbon and this species aids in the strong selective binding of substrates such as putrescine and histamine.  相似文献   

7.
We compare the low free energy structures of ten small, polar ligands in solution to their conformations in their respective receptor active sites. The solution conformations are generated by a systematic search and the free energies of representative structures are computed with a continuum solvation model. Based on the values of torsion angles, we find little similarity between low energy solution structures of small ligands and their active site conformations. However, in nine out of ten cases, the positions of 'anchor points' (key atoms responsible for tight binding) in the lowest energy solution structures are very similar to the positions of these atoms in the active site conformations. A metric that more closely captures the essentials of binding supports the basic premise underlying pharmacophore mapping, namely that active site conformations of small flexible ligands correspond to their low energy structures in solution. This work supports the efforts of building pharmacophore models based on the information present in solution structures of small isolated ligands.  相似文献   

8.
The enantiomers of the potent nonsteroidal inhibitor of aromatase fadrozole hydrochloride 3 have been separated and their absolute configuration determined by X-ray crystallography. On the basis of a molecular modeling comparison of the active enantiomer 4 and one of the most potent steroidal inhibitors reported to date, (19R)-10-thiiranylestr-4-ene-3,17-dione, 7, a model describing the relative binding modes of the azole-type and steroidal inhibitors of aromatase at the active site of the enzyme is proposed. It is suggested that the cyanophenyl moiety present in the most active azole inhibitors partially mimics the steroid backbone of the natural substrate for aromatase, androst-4-ene-3,17-dione, 1. The synthesis and biological testing of novel analogues of 3 used to define the accessible and nonaccessible volumes to ligands in the model of the active site of aromatase are reported.  相似文献   

9.
10.
The chloroplastic NADP malate dehydrogenase is completely inactive in its oxidized form and is activated by thiol/disulfide interchange with reduced thioredoxin. To elucidate the molecular mechanism underlying the absence of activity of the oxidized enzyme, we used site-directed mutagenesis to delete or substitute the two most C-terminal residues (C-terminal Val, penultimate Glu, both bearing negative charges). We also combined these mutations with the elimination of one or both of the possible regulatory N-terminal disulfides by mutating the corresponding cysteines. Proteins mutated at the C-terminal residues had no activity in the oxidized form but were partially inhibited when pretreated with the histidine-specific reagent diethyl pyrocarbonate before activation, showing that the active site was partially accessible. Proteins missing both N-terminal regulatory disulfides reached almost full activity without activation upon elimination of the negative charge of the penultimate Glu. These results strongly support a model where the C-terminal extension is docked into the active site through a negatively charged residue, acting as an internal inhibitor. They show also that the reduction of both N-terminal bridges is necessary to release the C-terminal extension from the active site. This is the first report for a thiol-activated enzyme of a regulatory mechanism resembling the well known intrasteric inhibition of protein kinases.  相似文献   

11.
The two main catalytic residues Cys25 and His159 of the monomeric cysteine protease papain are located on different walls of a cleft formed by two domains. This topology suggests a possible relationship between relative domain organization and catalytic mechanism. The effect on enzymatic parameters of structural modifications at various locations of the two-domain interface of papain was examined by individual or double replacements by Ala of pairs of interacting residues. Most modifications had no effect on enzyme activity. However, the enzyme's substrate turnover (kcat) decreased following simultaneous alteration of the two most conserved residues, forming an apolar contact located 15 A away from the active site. The pH activity profile of the double mutant was unchanged, indicating a conserved ionization state of the active site thiolate-imidazolium ion pair. This state is strongly dependent on the distance separating the two residues, thus suggesting that the active site geometry has not been significantly altered. Efficient enzymatic activity in papain requires more than a correct active site geometry and is influenced by domain packing properties in a region remote from the active site.  相似文献   

12.
Interactions of mammalian pancreatic cholesterol esterases from pig and rat with a family of aryl carbamates CnH2n+1NHCOOAr [n = 4-9; Ar = phenyl, p-X-phenyl (X = acetamido, bromo, fluoro, nitro, trifluoromethyl), 2-naphthyl, 2-tetrahydronaphthyl, estronyl] have been investigated, with an aim of delineating the ligand structural features which lead to effective molecular recognition by the active site of the enzyme. These carbamates inhibit the catalytic activity of CEase by rapid carbamylation of the active site, a process that shows saturation kinetics. Subsequent slow decarbamylation usually leads to full restoration of activity, and therefore aryl carbamates are transient inhibitors, or pseudo-substrates, of CEase. Structural variation of carbamate inhibitors allowed molecular recognition in the fatty acid binding and steroid binding loci of the extended active site to be probed, and the electronic nature of the carbamylation transition state to be characterized. Optimal inhibitory activity is observed when the length of the carbamyl function is n = 6 and n = 7 for porcine and rat cholesterol esterases, respectively, equivalent to eight- and nine-carbon fatty acyl chains. In contrast, inhibitory activity increases progressively as the partial molecular volume of the aromatic fragment increases. Hammett plots for p-substituted phenyl-N-hexyl carbamates indicate that the rate-determining step for carbamate inhibition is phenolate anion expulsion. Effects of the bile salt activator taurocholate on the kinetically resolved phases of the pseudo-substrate turnover of aryl carbamates were also studied. Taurocholate increases the affinity of the carbamate for the active site of cholesterol esterase in the reversible, noncovalent complex that precedes carbamylation and increases the rate constants of the serial carbamylation and decarbamylation steps. Structural variation of the N-alkyl chain and of the aryl fused-ring system provides an accounting of bile salt modulation of the fatty acid and steroid binding sites, respectively. In that pseudo-substrate turnover of aryl carbamates proceeds by a three-step mechanism that is analogous to that for rapid turnover of lipid ester substrates, these investigations illuminate details of ligand recognition by the extended active site of cholesterol esterase that are prominent determinants of the substrate specificity and catalytic power of the enzyme.  相似文献   

13.
Cytochrome c oxidase catalyzes the reduction of oxygen to water that is accompanied by pumping of four protons across the mitochondrial or bacterial membrane. Triggered by the results of recent x-ray crystallographic analyses, published data concerning the coupling of individual electron transfer steps to proton pumping are reanalyzed: Conversion of the conventional oxoferryl intermediate F to the fully oxidized form O is connected to pumping of only one proton. Most likely one proton is already pumped during the double reduction of O, and only three protons during conversion of the "peroxy" forms P to O via the oxoferryl form F. Based on the available structural, spectroscopic, and mutagenesis data, a detailed mechanistic model, carefully considering electrostatic interactions, is presented. In this model, each of the four reductions of heme a during the catalytic cycle is coupled to the uptake of one proton via the D-pathway. These protons, but never more than two, are temporarily stored in the regions of the heme a and a3 propionates and are driven to the outside ("pumped") by electrostatic repulsion from protons entering the active site during turnover. The first proton is pumped by uptake of one proton via the K-pathway during reduction, the second and third proton during the P --> F transition when the D-pathway and the active site become directly connected, and the fourth one upon conversion of F to O. Atomic structures are assigned to each intermediate including F' with an alternative route to O.  相似文献   

14.
The mammalian ubiquitin conjugating enzyme known as E2-25K catalyzes the synthesis of polyubiquitin chains linked exclusively through K48-G76 isopeptide bonds. The properties of truncated and chimeric forms of E2-25K suggest that the polyubiquitin chain synthesis activity of this E2 depends on specific interactions between its conserved 150-residue core domain and its unique 50-residue tail domain [Haldeman, M. T., Xia, G., Kasperek, E. M., and Pickart, C. M. (1997) Biochemistry 36, 10526-10537]. In the present study, we provide strong support for this model by showing that a point mutation in the core domain (S86Y) mimics the effect of deleting the entire tail domain: the ability to form an E2 approximately ubiquitin thiol ester is intact, while conjugation activity is severely inhibited (>/=100-fold reduction in kcat/Km). The properties of E2-25K enzymes carrying the S86Y mutation indicate that this mutation strengthens the interaction between the core and tail domains: both free and ubiquitin-bound forms of S86Y-25K are completely resistant to tryptic cleavage at K164 in the tail domain, whereas wild-type enzyme is rapidly cleaved at this site. Other properties of S86Y-26K suggest that the active site of this mutant enzyme is more occluded than the active site of the wild-type enzyme. (1) Free S86Y-25K is alkylated by iodoacetamide 2-fold more slowly than the wild-type enzyme. (2) In assays of E2 approximately ubiquitin thiol ester formation, S86Y-25K shows a 4-fold reduced affinity for E1. (3) The ubiquitin thiol ester adduct of S86Y-25K undergoes (uncatalyzed) reaction with dithiothreitol 3-fold more slowly than the wild-type thiol ester adduct. One model to accommodate these findings postulates that an enhanced interaction between the core and tail domains, induced by the S86Y mutation, causes a steric blockade at the active site which prevents access of the incoming ubiquitin acceptor to the thiol ester bond. Consistent with this model, the S86Y mutation inhibits ubiquitin transfer to macromolecular acceptors (ubiquitin and polylysine) more strongly than transfer to small-molecule acceptors (free lysine and short peptides). These results suggest that unique residues proximal to E2 active sites may influence specific function by mediating intramolecular interactions.  相似文献   

15.
The structure of the complex of the chorismate mutase from the yeast Saccharomyces cerevisiae with a transition state analog is constructed using a suite of docking tools. The construction finds the best location for the active site in the enzyme, and the best orientation of the analog compound in the active site. The resulting complex shows extensive salt links and hydrogen bonds between the enzyme and the compound, including those mediated by water molecules. A network of polar interactions between amino acid residues is found to solidify the active site of the enzyme. The enzymatic mechanism suggested for a bacterial chorismate mutase, that the active site is by design capable of selecting an active conformer of the substrate, and of stabilizing the transition state, is apparently intact in the yeast enzyme. No direct evidence is found to support an alternative mechanism which involves specific catalytic groups, although the possibility is not eliminated. This finding reinforces the notion of a function being evolutionarily conserved via a common mechanism, rather than via sequential or structural homology.  相似文献   

16.
Hydration of protein cavities influences protein stability, dynamics, and function. Protein active sites usually contain water molecules that, upon ligand binding, are either displaced into bulk solvent or retained to mediate protein-ligand interactions. The contribution of water molecules to ligand binding must be accounted for to compute accurate values of binding affinities. This requires estimation of the extent of hydration of the binding site. However, it is often difficult to identify the water molecules involved in the binding process when ligands bind on the surface of a protein. Cytochrome P450cam is, therefore, an ideal model system because its substrate binds in a buried active site, displacing partially disordered solvent, and the protein is well characterized experimentally. We calculated the free energy differences for having five to eight water molecules in the active site cavity of the unliganded enzyme from molecular dynamics simulations by thermodynamic integration employing a three-stage perturbation scheme. The computed free energy differences between the hydration states are small (within 12 kJ mol-1) but distinct. Consistent with the crystallographic determination and studies employing hydrostatic pressure, we calculated that, although ten water molecules could in principle occupy the volume of the active site, occupation by five to six water molecules is thermodynamically most favorable.  相似文献   

17.
The diffusional encounter between substrate and enzyme, and hence catalytic efficiency, can be enhanced by mutating charged residues on the surface of the enzyme. In this paper we present a simple method for screening such mutations. This is based on our earlier result that electrostatic enhancement of the enzyme-substrate binding rate constant can be accounted for just by the interaction potential within the active site. Assuming that catalytic and structural integrity is maintained, the catalytic efficiency can be optimized by surface charge mutations which lead to stronger interaction potential within the active site. Application of the screening method on superoxide dismutase shows that only charge mutations close to the active site will have practical effect on the catalytic efficiency. This rationalizes a large number of findings obtained in previous simulation and experimental studies.  相似文献   

18.
The structure of the isocitrate dehydrogenase (IDH) complex with bound alpha-ketoglutarate, Ca2+, and NADPH was solved at 2.7-A resolution. The alpha-ketoglutarate binds in the active site at the same position and orientation as isocitrate, with a difference between the two bound molecules of about 0.8 A. The Ca2+ metal is coordinated by alpha-ketoglutarate, three conserved aspartate residues, and a pair of water molecules. The largest motion in the active site relative to the isocitrate enzyme complex is observed for tyrosine 160, which originally forms a hydrogen bond to the labile carboxyl group of isocitrate and moves to form a new hydrogen bond to Asp 307 in the complex with alpha-ketoglutarate. This triggers a number of significant movements among several short loops and adjoining secondary structural elements in the enzyme, most of which participate in dimer stabilization and formation of the active-site cleft. These rearrangements are similar to the ligand-binding-induced movements observed in globins and insulin and serve as a model for an enzymatic mechanism which involves local shifts of secondary structural elements during turnover, rather than large-scale domain closures or loop transitions induced by substrate binding such as those observed in hexokinase or triosephosphate isomerase.  相似文献   

19.
We report the x-ray crystal structure of the methylesterase CheB, a phosphorylation-activated response regulator involved in reversible modification of bacterial chemotaxis receptors. Methylesterase CheB and methyltransferase CheR modulate signaling output of the chemotaxis receptors by controlling the level of receptor methylation. The structure of CheB, which consists of an N-terminal regulatory domain and a C-terminal catalytic domain joined by a linker, was solved by molecular replacement methods using independent search models for the two domains. In unphosphorylated CheB, the N-terminal domain packs against the active site of the C-terminal domain and thus inhibits methylesterase activity by directly restricting access to the active site. We propose that phosphorylation of CheB induces a conformational change in the regulatory domain that disrupts the domain interface, resulting in a repositioning of the domains and allowing access to the active site. Structural similarity between the two companion receptor modification enzymes, CheB and CheR, suggests an evolutionary and/or functional relationship. Specifically, the phosphorylated N-terminal domain of CheB may facilitate interaction with the receptors, similar to the postulated role of the N-terminal domain of CheR. Examination of surfaces in the N-terminal regulatory domain of CheB suggests that despite a common fold throughout the response regulator family, surfaces used for protein-protein interactions differ significantly. Comparison between CheB and other response regulators indicates that analogous surfaces are used for different functions and conversely, similar functions are mediated by different molecular surfaces.  相似文献   

20.
The active site gorge of acetylcholinesterase (AChE) contains two sites of ligand binding, an acylation site near the base of the gorge with a catalytic triad characteristic of serine hydrolases, and a peripheral site at the mouth of the gorge some 10-20 A from the acylation site. Many ligands that bind exclusively to the peripheral site inhibit substrate hydrolysis at the acylation site, but the mechanistic interpretation of this inhibition has been unclear. Previous interpretations have been based on analyses of inhibition patterns obtained from steady-state kinetic models that assume equilibrium ligand binding. These analyses indicate that inhibitors bound to the peripheral site decrease acylation and deacylation rate constants and/or decrease substrate affinity at the acylation site by factors of up to 100. Conformational interactions have been proposed to account for such large inhibitory effects transmitted over the distance between the two sites, but site-specific mutagenesis has failed to reveal residues that mediate the proposed conformational linkage. Since examination of individual rate constants in the AChE catalytic pathway reveals that assumptions of equilibrium ligand binding cannot be justified, we introduce here an alternative nonequilibrium analysis of the steady-state inhibition patterns. This analysis incorporates a steric blockade hypothesis which assumes that the only effect of a bound peripheral site ligand is to decrease the association and dissociation rate constants for an acylation site ligand without altering the equilibrium constant for ligand binding to the acylation site. Simulations based on this nonequilibrium steric blockade model were in good agreement with experimental data for inhibition by the peripheral site ligands propidium and gallamine at low concentrations of either acetylthiocholine or phenyl acetate if binding of these ligands slows substrate association and dissociation rate constants by factors of 5-70. Direct measurements with the acylation site ligands huperzine A and m-(N,N, N-trimethylammonio)trifluoroacetophenone showed that bound propidium decreased the association rate constants 49- and 380-fold and the dissociation rate constants 10- and 60-fold, respectively, relative to the rate constants for these acylation site ligands with free AChE, in reasonable agreement with the nonequilibrium steric blockade model. We conclude that this model can account for the inhibition of AChE by small peripheral site ligands such as propidium without invoking any conformational interaction between the peripheral and acylation sites.  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-23

京公网安备 11010802026262号