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.
Haloalkane dehalogenase (DhlA) converts haloalkanes to their corresponding alcohols and halide ions. The rate-limiting step in the reaction of DhlA is the release of the halide ion. The kinetics of halide release have been analyzed by measuring halide binding with stopped-flow fluorescence experiments. At high halide concentrations, halide import occurs predominantly via the rapid formation of a weak initial collision complex, followed by transport of the ion to the active site. To obtain more insight in this collision complex, we determined the X-ray structure of DhlA in the presence of bromide and investigated the kinetics of mutants that were constructed on the basis of this structure. The X-ray structure revealed one bromide ion firmly bound in the active site and two bromide ions weakly bound on the surface of the enzyme. One of the weakly bound ions is close to Thr197 and Phe294, near the entrance of the earlier proposed tunnel for substrate import. Kinetic analysis of bromide import by the Thr197Ala and Phe294Ala mutants of DhlA at high halide concentration showed that the rate constants for halide binding no longer displayed a wild-type-like parabolic increase with increasing bromide concentrations. This is in agreement with an elimination or a decrease in affinity of the surface-located halide-binding site. Likewise, chloride binding kinetics of the mutants indicated significant differences with wild-type enzyme. The results indicate that Thr197 and Phe294 are involved in the formation of an initial collision complex for halide import in DhlA and provide experimental evidence for the role of the tunnel in substrate and product transport.  相似文献   

2.
Nieslanik BS  Ibarra C  Atkins WM 《Biochemistry》2001,40(12):3536-3543
Binding of a hydrophobic glutathione product conjugate to rGST A1-1 proceeds via a two-step mechanism, including rapid ligand docking, followed by a slow isomerization to the final [GST.ligand] complex, which involves the localization of the flexible C-terminal helix. These kinetically resolved steps have been observed previously by stopped-flow fluorescence with the wild-type rGST A1-1, which contains a native Trp-21 approximately 20 A from the ligand binding site at the intrasubunit domain-domain interface. To confirm this binding mechanism, as well as elucidate the effects of truncation of the C-terminus, we have further characterized the binding and dissociation of the glutathione-ethacrynic acid product conjugate (GS-EA) to wild-type, F222W:W21F, and Delta209-222 rGST A1-1 and wild-type hGST A1-1. Although modest kinetic differences were observed between the hGST A1-1 and rGST A1-1, stopped-flow binding studies with GS-EA verified that the two-step mechanism of ligand binding is not unique to the GST A1-1 isoform from rat. An F222W:W21F rGST A1-1 double mutant provides a direct fluorescence probe of changes in the environment of the C-terminal residue. The observation of two relaxation times during ligand binding and dissociation to F222W:W21F suggests that the C-terminus has an intermediate conformation following ligand docking, which is distinct from its conformation in the apoenzyme or localized helical state. For the wild-type, Delta209-222, and F222W:W21F proteins, variable-temperature stopped-flow experiments were performed and activation parameters calculated for the individual steps of the binding reaction. Activation parameters for the binding reaction coordinate illustrate that the C-terminus provides a significant entropic contribution to ligand binding, which is completely realized within the initial docking step of the binding mechanism. In contrast, the slow isomerization step is enthalpically driven. The partitioning of entropic and enthalpic components of binding energy was confirmed by isothermal titration calorimetry with wild-type and Delta209-222 rGST A1-1.  相似文献   

3.
Cochran JC  Gilbert SP 《Biochemistry》2005,44(50):16633-16648
The ATPase mechanism of kinesin superfamily members in the absence of microtubules remains largely uncharacterized. We have adopted a strategy to purify monomeric human Eg5 (HsKSP/Kinesin-5) in the nucleotide-free state (apoEg5) in order to perform a detailed transient state kinetic analysis. We have used steady-state and presteady-state kinetics to define the minimal ATPase mechanism for apoEg5 in the absence and presence of the Eg5-specific inhibitor, monastrol. ATP and ADP binding both occur via a two-step process with the isomerization of the collision complex limiting each forward reaction. ATP hydrolysis and phosphate product release are rapid steps in the mechanism, and the observed rate of these steps is limited by the relatively slow isomerization of the Eg5-ATP collision complex. A conformational change coupled to ADP release is the rate-limiting step in the pathway. We propose that the microtubule amplifies and accelerates the structural transitions needed to form the ATP hydrolysis competent state and for rapid ADP release, thus stimulating ATP turnover and increasing enzymatic efficiency. Monastrol appears to bind weakly to the Eg5-ATP collision complex, but after tight ATP binding, the affinity for monastrol increases, thus inhibiting the conformational change required for ADP product release. Taken together, we hypothesize that loop L5 of Eg5 undergoes an "open" to "closed" structural transition that correlates with the rearrangements of the switch-1 and switch-2 regions at the active site during the ATPase cycle.  相似文献   

4.
Divalent cation binding to the high- and low-affinity sites on G-actin   总被引:4,自引:0,他引:4  
C T Zimmerle  K Patane  C Frieden 《Biochemistry》1987,26(20):6545-6552
Metal binding to skeletal muscle G-actin has been assessed by equilibrium dialysis using 45Ca2+ and by kinetic measurements of the increase in the fluorescence of N-acetyl-N'-(5-sulfo-1-naphthyl)-ethylenediamine-labeled actin. Two classes of cation binding sites were found on G-actin which could be separated on the basis of their Ca2+ affinity: a single high-affinity site with a Kd considerably less than 1 microM and three identical moderate-affinity binding sites with a Kd of 18 microM. The data for the Mg2+-induced fluorescence enhancement of actin labeled with N-acetyl-N'-(5-sulfo-1-naphthyl)ethylenediamine support a previously suggested mechanism [Frieden, C. (1982) J. Biol. Chem. 257, 2882-2886] in which Ca2+ is replaced by Mg2+ at the moderate affinity site(s), followed by a slow actin isomerization. This isomerization occurs independently of Ca2+ release from the high-affinity site. The fluorescence data do not support a mechanism in which this isomerization is directly related to Ca2+ release from the high-affinity site. Fluorescence changes of labeled actin associated with adding metal chelators are complex and do not reflect the same change induced by Mg2+ addition. Fluorescence changes in the labeled actin have also been observed for the addition of Cd2+ or Mn2+ instead of Mg2+. It is proposed actin may undergo a host of subtle conformational changes dependent on the divalent cation bound. We have also developed a method by which progress curves of a given reaction can be analyzed by nonlinear regression fitting of kinetic simulations to experimental reaction time courses.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
6.
Frase H  Lee I 《Biochemistry》2007,46(22):6647-6657
Lon is a homo-oligomeric ATP-dependent serine protease that functions in the degradation of damaged and certain regulatory proteins. This enzyme has emerged as a novel target in the development of antibiotics because of its importance in conferring bacterial virulence. In this study, we explored the mechanism by which the proteasome inhibitor MG262, a peptidyl boronate, inhibits the peptide hydrolysis activity of Salmonella enterica serovar Typhimurium Lon. In addition, we synthesized a fluorescent peptidyl boronate inhibitor based upon the amino acid sequence of a product of peptide hydrolysis by the enzyme. Using steady-state kinetic techniques, we have shown that two peptidyl boronate variants are competitive inhibitors of the peptide hydrolysis activity of Lon and follow the same two-step, time-dependent inhibition mechanism. The first step is rapid and involves binding of the inhibitor and formation of a covalent adduct with the active site serine. This is followed by a second slow step in which Lon undergoes a conformational change or isomerization to increase the interaction of the inhibitor with the proteolytic active site to yield an overall inhibition constant of 5-20 nM. Although inhibition of serine and threonine proteases by peptidyl boronates has been detected previously, Lon is the first protease that has required the binding of ATP in order to observe inhibition.  相似文献   

7.
We previously reported that the kinetic profiles for the association and dissociation of functionally diverse C(8)-CoA-ligands, viz., octanoyl-CoA (substrate), octenoyl-CoA (product), and octynoyl-CoA (inactivator) with medium chain acyl-CoA dehydrogenase (MCAD), were essentially identical, suggesting that the protein conformational changes played an essential role during ligand binding and/or catalysis [Peterson, K. L., Sergienko, E. E., Wu, Y., Kumar, N. R., Strauss, A. W., Oleson, A. E., Muhonen, W. W., Shabb, J. B., and Srivastava, D. K. (1995) Biochemisry 34, 14942-14953]. To ascertain the structural basis of the above similarity, we investigated the kinetics of association and dissociation of alpha-CH-->NH-substituted C(8)-CoA, namely, 2-azaoctanoyl-CoA, with the recombinant form of human liver MCAD. The rapid-scanning and single wavelength stopped-flow data for the binding of 2-azaoctanoyl-CoA to MCAD revealed that the overall interaction proceeds via two steps. The first (fast) step involves the formation of an enzyme-ligand collision complex (with a dissociation constant of K(c)), followed by a slow isomerization step (with forward and reverse rate constants of k(f) and k(r), respectively) with concomitant changes in the electronic structure of the enzyme-bound FAD. Since the latter step involves a concurrent change in the enzyme's tryptophan fluorescence, it is suggested that the isomerization step is coupled to the changes in the protein conformation. Although the overall binding affinity (K(d)) of the enzyme-2-azaoctanoyl-CoA complex is similar to that of the enzyme-octenoyl-CoA complex, their microscopic equilibria within the collision and isomerized complexes show an opposite relationship. These results coupled with the isothermal titration microcalorimetric studies lead to the suggestion that the electrostatic interaction within the enzyme site phase modulates the microscopic steps, as well as their corresponding ground and transition states, during the course of the enzyme-ligand interaction.  相似文献   

8.
5-Aminolevulinate synthase (ALAS) is the first enzyme of the heme biosynthetic pathway in non-plant eukaryotes and the alpha-subclass of purple bacteria. The pyridoxal 5'-phosphate cofactor at the active site undergoes changes in absorptive properties during substrate binding and catalysis that have allowed us to study the kinetics of these reactions spectroscopically. Rapid scanning stopped-flow experiments of murine erythroid 5-aminolevulinate synthase demonstrate that reaction with glycine plus succinyl-CoA results in a pre-steady-state burst of quinonoid intermediate formation. Thus, a step following binding of substrates and initial quinonoid intermediate formation is rate-determining. The steady-state spectrum of the enzyme is similar to that formed in the presence of 5-aminolevulinate, suggesting that release of this product limits the overall rate. Reaction of either glycine or 5-aminolevulinate with ALAS is slow (kf = 0.15 s-1) and approximates kcat. The rate constant for reaction with glycine is increased at least 90-fold in the presence of succinyl-CoA and most likely represents a slow conformational change of the enzyme that is accelerated by succinyl-CoA. The slow rate of reaction of 5-aminolevulinate with ALAS is 5-aminolevulinate-independent, suggesting that it also represents a slow isomerization of the enzyme. Reaction of succinyl-CoA with the enzyme-glycine complex to form a quinonoid intermediate is a biphasic process and may be irreversible. Taken together, the data suggest that turnover is limited by release of 5-aminolevulinate or a conformational change associated with 5-aminolevulinate release.  相似文献   

9.
Enzyme structures solved with and without bound substrate often show that substrate-induced conformational changes bring catalytic residues into alignment, alter the local environment, and position the substrate for catalysis. Although the structural data are compelling, the role of conformational changes in enzyme specificity has been controversial in that specificity is a kinetic property that is not easy to predict based upon structure alone. Recent studies on DNA polymerization have illuminated the role of substrate-induced conformational changes in enzyme specificity by showing that the rate at which the enzyme opens to release the bound substrate is a key kinetic parameter. The slow release of a correct substrate commits it to the forward reaction so that specificity is determined solely by the rate of substrate binding, including the isomerization step, and not by the slower rate of the chemical reaction. In contrast, fast dissociation of an incorrect substrate favors release rather than reaction. Thus, the conformational change acts as a molecular switch to select the right substrate and to recognize and disfavor the reaction of an incorrect substrate. A conformational switch may also favor release rather than reverse reaction of the product.  相似文献   

10.
Erythromycin and tylosin are 14- and 16-membered lactone ring macrolides, respectively. The current work shows by means of kinetic and chemical footprinting analysis that both antibiotics bind to Escherichia coli ribosomes in a two-step process. The first step established rapidly, involves a low-affinity binding site placed at the entrance of the exit tunnel in the large ribosomal subunit, where macrolides bind primarily through their hydrophobic portions. Subsequently, slow conformational changes mediated by the antibiotic hydrophilic portion push the drugs deeper into the tunnel, in a high-affinity site. Compared with erythromycin, tylosin shifts to the high-affinity site more rapidly, due to the interaction of the mycinose sugar of the drug with the loop of H35 in domain II of 23 S rRNA. Consistently, mutations of nucleosides U2609 and U754 implicated in the high-affinity site reduce the shift of tylosin to this site and destabilize, respectively, the final drug-ribosome complex. The weak interaction between tylosin and the ribosome is Mg2+ independent, unlike the tight binding. In contrast, both interactions between erythromycin and the ribosome are reduced by increasing concentrations of Mg2+ ions. Polyamines attenuate erythromycin affinity for the ribosome at both sequential steps of binding. In contrast, polyamines facilitate the initial binding of tylosin, but exert a detrimental, more pronounced, effect on the drug accommodation at its final position. Our results emphasize the role of the particular interactions that side chains of tylosin and erythromycin establish with 23 S rRNA, which govern the exact binding process of each drug and its response to the ionic environment.  相似文献   

11.
The kinetics of slow onset inhibition of Proteinase K by a proteinaceous alkaline protease inhibitor (API) from a Streptomyces sp. is presented. The kinetic analysis revealed competitive inhibition of Proteinase K by API with an IC50 value 5.5 +/- 0.5 x 10-5 m. The progress curves were time-dependent, consistent with a two-step slow tight binding inhibition. The first step involved a rapid equilibrium for formation of reversible enzyme-inhibitor complex (EI) with a Ki value 5.2 +/- 0.6 x 10-6 m. The EI complex isomerized to a stable complex (EI*) in the second step because of inhibitor-induced conformational changes, with a rate constant k5 (9.2 +/- 1 x 10-3 s-1). The rate of dissociation of EI* (k6) was slower (4.5 +/- 0.5 x 10-5 s-1) indicating the tight binding nature of the inhibitor. The overall inhibition constant Ki* for two-step inhibition of Proteinase K by API was 2.5 +/- 0.3 x 10-7 m. Time-dependent dissociation of EI* revealed that the complex failed to dissociate after a time point and formed a conformationally altered, irreversible complex EI**. These conformational states of enzyme-inhibitor complexes were characterized by fluorescence spectroscopy. Tryptophanyl fluorescence of Proteinase K was quenched as a function of API concentration without any shift in the emission maximum indicating a subtle conformational change in the enzyme, which is correlated to the isomerization of EI to EI*. Time-dependent shift in the emission maxima of EI* revealed the induction of gross conformational changes, which can be correlated to the irreversible conformationally locked EI** complex. API binds to the active site of the enzyme as demonstrated by the abolished fluorescence of 5-iodoacetamidofluorescein-labeled Proteinase K. The chemoaffinity labeling experiments lead us to hypothesize that the inactivation of Proteinase K is because of the interference in the electronic microenvironment and disruption of the hydrogen-bonding network between the catalytic triad and other residues involved in catalysis.  相似文献   

12.
Slow inactivation involves a local rearrangement of the outer mouth of voltage-gated potassium channels, but nothing is known regarding rearrangements in the cavity between the activation gate and the selectivity filter. We now report that the cavity undergoes a conformational change in the slow-inactivated state. This change is manifest as altered accessibility of residues facing the aqueous cavity and as a marked decrease in the affinity of tetraethylammonium for its internal binding site. These findings have implications for global alterations of the channel during slow inactivation and putative coupling between activation and slow-inactivation gates.  相似文献   

13.
This is the first report that describes the inhibition mechanism of xylanase from Thermomonospora sp. by pepstatin A, a specific inhibitor toward aspartic proteases. The kinetic analysis revealed competitive inhibition of xylanase by pepstatin A with an IC50 value 3.6 +/- 0.5 microm. The progress curves were time-depended, consistent with a two-step slow tight binding inhibition. The inhibition followed a rapid equilibrium step to form a reversible enzyme-inhibitor complex (EI), which isomerizes to the second enzyme-inhibitor complex (EI*), which dissociated at a very slow rate. The rate constants determined for the isomerization of EI to EI* and the dissociation of EI* were 15 +/- 1 x 10(-5) and 3.0 +/- 1 x 10(-8) s(-1), respectively. The Ki value for the formation of EI complex was 1.5 +/- 0.5 microm, whereas the overall inhibition constant Ki* was 28.0 +/- 1 nm. The conformational changes induced in Xyl I by pepstatin A were monitored by fluorescence spectroscopy, and the rate constants derived were in agreement with the kinetic data. Thus, the conformational alterations were correlated to the isomerization of EI to EI*. Pepstatin A binds to the active site of the enzyme and disturbs the native interaction between the histidine and lysine, as demonstrated by the abolished isoindole fluorescence of o-phthalaldehyde-labeled xylanase. Our results revealed that the inactivation of xylanase is due to the interference in the electronic microenvironment and disruption of the hydrogen-bonding network between the essential histidine and other residues involved in catalysis, and a model depicting the probable interaction between pepstatin A with xylanase has been proposed.  相似文献   

14.
Eukaryotic cell cycle progression is controlled by the ordered action of cyclin-dependent kinases, activation of which occurs through the binding of the cyclin to the Cdk followed by phosphorylation of a conserved threonine in the T-loop of the Cdk by Cdk-activating kinase (CAK). Despite our understanding of the structural changes, which occur upon Cdk/cyclin formation and activation, little is known about the dynamics of the molecular events involved. We have characterized the mechanism of Cdk2/cyclin A complex formation and activation at the molecular and dynamic level by rapid kinetics and demonstrate here that it is a two-step process. The first step involves the rapid association between the PSTAIRE helix of Cdk2 and helices 3 and 5 of the cyclin to yield an intermediate complex in which the threonine in the T-loop is not accessible for phosphorylation. Additional contacts between the C-lobe of the Cdk and the N-terminal helix of the cyclin then induce the isomerization of the Cdk into a fully mature form by promoting the exposure of the T-loop for phosphorylation by CAK and the formation of the substrate binding site. This conformational change is selective for the cyclin partner.  相似文献   

15.
The x-ray crystal structure of human myeloperoxidase has been extended to 1.8 A resolution, using x-ray data recorded at -180 degrees C (r = 0.197, free r = 0.239). Results confirm that the heme is covalently attached to the protein via two ester linkages between the carboxyl groups of Glu(242) and Asp(94) and modified methyl groups on pyrrole rings A and C of the heme as well as a sulfonium ion linkage between the sulfur atom of Met(243) and the beta-carbon of the vinyl group on pyrrole ring A. In the native enzyme a bound chloride ion has been identified at the amino terminus of the helix containing the proximal His(336). Determination of the x-ray crystal structure of a myeloperoxidase-bromide complex (r = 0.243, free r = 0.296) has shown that this chloride ion can be replaced by bromide. Bromide is also seen to bind, at partial occupancy, in the distal heme cavity, in close proximity to the distal His(95), where it replaces the water molecule hydrogen bonded to Gln(91). The bromide-binding site in the distal cavity appears to be the halide-binding site responsible for shifts in the Soret band of the absorption spectrum of myeloperoxidase. It is proposed that halide binding to this site inhibits the enzyme by effectively competing with H(2)O(2) for access to the distal histidine, whereas in compound I, the same site may be the halide substrate-binding site.  相似文献   

16.
The alpha subunit is bound with negative cooperativity to the holo beta 2 subunit of tryptophan synthase in phosphate buffer. Thus it is feasible to measure separately the rates of formation both of the stable alpha beta 2 subcomplex from beta 2, and of the mature alpha 2 beta 2 complex from alpha beta 2, using stopped-flow techniques. Addition of each alpha subunit proceeds in two steps; an initial alpha beta protomer is formed rapidly, which subsequently isomerizes slowly to the equilibrium state. The rates of dissociation of both the alpha beta 2 and alpha 2 beta 2 complexes were measured by trapping released alpha subunit with enzymically inactive reduced beta 2 subunit. The reversal of the slow isomerization both determines the rate of dissociation, and accounts for the high overall affinity of the beta protomer for the alpha subunit. The data fit to a sequential assembly mechanism consisting of seven protein species and yields values for most of the rate constants and all of the microscopic equilibrium constants. Negative cooperativity arises from a weaker initial binding of the second alpha subunit, as expressed by its larger off-constant, possibly due to steric hindrance. The kinetics of binding of L-serine and indolepropanol phosphate during the assembly process shows that the beta protomer is already partially activated in the initial alpha beta complex. Full activation is achieved in the slow isomerization reaction. In contrast, the alpha subunit gains high affinity for indolepropanol phosphate only in the isomerization reaction. These observations indicate that the isomerization involves synchronous conformation changes of both alpha and beta protomers.  相似文献   

17.
As shown by X-ray crystallography, horse liver alcohol dehydrogenase undergoes a global conformational change upon binding of NAD+ or NADH, involving a rotation of the catalytic domain relative to the coenzyme binding domain and the closing up of the active site to produce a catalytically efficient enzyme. The conformational change requires a complete coenzyme and is affected by various chemical or mutational substitutions that can increase the catalytic turnover by altering the kinetics of the isomerization and rate of dissociation of coenzymes. The binding of NAD+ is kinetically limited by a unimolecular isomerization (corresponding to the conformational change) that is controlled by deprotonation of the catalytic zinc-water to produce a negatively-charged zinc-hydroxide, which can attract the positively-charged nicotinamide ring. The deprotonation is facilitated by His-51 acting through a hydrogen-bonded network to relay the proton to solvent. Binding of NADH also involves a conformational change, but the rate is very fast. After the enzyme binds NAD+ and closes up, the substrate displaces the hydroxide bound to the catalytic zinc; this exchange may involve a double displacement reaction where the carboxylate group of a glutamate residue first displaces the hydroxide (inverting the tetrahedral coordination of the zinc), and then the exogenous ligand displaces the glutamate. The resulting enzyme-NAD+-alcoholate complex is poised for hydrogen transfer, and small conformational fluctuations may bring the reactants together so that the hydride ion is transferred by quantum mechanical tunneling. In the process, the nicotinamide ring may become puckered, as seen in structures of complexes of the enzyme with NADH. The conformational changes of alcohol dehydrogenase demonstrate the importance of protein dynamics in catalysis.  相似文献   

18.
A stopped-flow technique was used to study the spectral changes occurring in bilirubin-albumin following a pH jump as well as following binding of bilirubin at 25 degrees C. The changes were studied in two wavelength ranges, 280-310 nm (tyrosine residues) and 400-510 nm (bound bilirubin). The changes were analyzed according to a scheme of consecutive unimolecular reactions. Spectral monitoring of a pH jump from 11.3 to 11.8 reveals that the bilirubin-albumin complex changes its structure in several steps. The UV absorption spectra show that 3.8 tyrosine residues ionize in the first step, 2.5 in the second, none in the third, and 0.8 in the fourth and following steps. The visible absorption spectrum of bound bilirubin changes in the second, third, and fourth steps. The bilirubin spectra of the different bilirubin-albumin complexes occurring in the transition show a common isosbestic point at 445 nm, indicating a change of the dihedral angle between the two bilirubin chromophores in a three-step reaction. It is suggested that 1 tyrosine residue is located close to the bilirubin site and is externalized in the second step. Bilirubin binding to albumin was monitored at two pH values, 11.3 and 11.8. At pH 11.3 the complex changes its structure in a three-step relaxation sequence. A change of the dihedral angle between the bilirubin chromophores can explain the spectral changes observed in the second and third relaxations. Protonation of 0.7 tyrosine residues occurs in the third relaxation, suggesting internalization of a tyrosine residue as a late consequence of bilirubin binding. At pH 11.8 a two-step relaxation sequence follows bilirubin binding. No tyrosine protonation occurs. Bilirubin is probably bound more superficially at pH 11.8 than at pH 11.3.  相似文献   

19.
Galletto R  Jezewska MJ  Bujalowski W 《Biochemistry》2004,43(34):11002-11016
Kinetics of the Escherichia coli PriA helicase interactions with the ssDNA has been studied, using the fluorescence stopped-flow technique. Experiments have been performed with a series of fluorescent etheno derivatives of ssDNA adenosine oligomers, differing in the number of nucleotide residues. The PriA helicase binds the ssDNA in the sequential process defined by [reaction: see text]. In the first step, the enzyme associates fast with the ssDNA without inducing conformational changes in the DNA. The dependence of the partial equilibrium constant, characterizing the first step, upon the length of the ssDNA strictly reflects the statistical relationship between the size of the DNA-binding site and the number of potential binding sites on the ssDNA. Only the DNA-binding site that encompasses 6.3 +/- 1 residues is directly involved in interactions. The site is located on a structural domain allowing the enzyme to efficiently search and recognize small patches of the ssDNA. Intramolecular steps are independent of the ssDNA length and accompanied by changes in the DNA structure. Salt and glycerol effects on the studied kinetics indicate a very different nature of the intermediates. While the bimolecular step is characterized by net ion release and water uptake, net ion uptake and water release accompany intramolecular transitions. Specific ion binding stabilizes the helicase-ssDNA complex in (P)(2) and (P)(3) intermediates. However, magnesium and AMP-PNP do not affect the mechanism of enzyme-ssDNA interactions. The sequential character of the mechanism indicates that the enzyme does not exist in a preequilibrium conformational transition prior to the DNA binding.  相似文献   

20.
The serpin, antithrombin, requires allosteric activation by a sequence-specific pentasaccharide unit of heparin or heparan sulfate glycosaminoglycans to function as an anticoagulant regulator of blood clotting proteases. Surprisingly, X-ray structures have shown that the pentasaccharide produces similar induced-fit changes in the heparin binding site of native and latent antithrombin despite large differences in the heparin affinity and global conformation of these two forms. Here we present kinetic evidence for similar induced-fit mechanisms of pentasaccharide binding to native and latent antithrombins and kinetic simulations which together support a three-step mechanism of allosteric activation of native antithrombin involving two successive conformational changes. Equilibrium binding studies of pentasaccharide interactions with native and latent antithrombins and the salt dependence of these interactions suggest that each conformational change is associated with distinct spectroscopic changes and is driven by a progressively better fit of the pentasaccharide in the binding site. The observation that variant antithrombins that cannot undergo the second conformational change bind the pentasaccharide like latent antithrombin and are partially activated suggests that both conformational changes contribute to allosteric activation, in agreement with a recently proposed model of allosteric activation.  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