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.
Saraswat V  Massiah MA  Lopez G  Amzel LM  Mildvan AS 《Biochemistry》2002,41(52):15566-15577
The MutT enzyme from E. coli, in the presence of a divalent cation, catalyzes the hydrolysis of nucleoside- and deoxynucleoside-triphosphate (NTP) substrates by nucleophilic substitution at Pbeta, to yield a nucleotide (NMP) and PPi. The best substrate of MutT is believed to be the mutagenic nucleotide 8-oxo-dGTP, on the basis of its 10(3.4)-fold lower K(m) than that of dGTP (Maki, H., and Sekiguchi, M. (1992) Nature 355, 273-275). To determine the true affinity of MutT for an 8-oxo-nucleotide and to elucidate the kinetic scheme, product inhibition by 8-oxo-dGMP and dGMP and direct binding of these nucleotides to MutT were studied. With Mg(2+)-activated dGTP hydrolysis, 8-oxo-dGMP is a noncompetitive inhibitor with K(I)(sl)(o)(pe) = 49 nM, which is 10(4.6)-fold lower than the K(I)(sl)(o)(pe)of dGMP (1.7 mM). Similarly, the K(I)(intercept) of 8-oxo-dGMP is 10(4.0)-fold lower than that of dGMP. PPi is a linear uncompetitive inhibitor, suggesting that it dissociates first from the product complex, followed by the nucleotide. Noncompetitive inhibition by dGMP and 8-oxo-dGMP indicates an "iso" mechanism in which the nucleotide product leaves an altered form of the enzyme which slowly reverts to the form which binds substrate. Consistent with this kinetic scheme, (1)H-(15)N HSQC titration of MutT with dGMP reveals weak binding and fast exchange from one site with a K(D) = 1.8 mM, in agreement with its K(I)(sl)(o)(pe). With 8-oxo-dGMP, tight binding and slow exchange (n = 1.0 +/- 0.1, K(D) < 0.25 mM) are found. Isothermal calorimetric titration of MutT with 8-oxo-dGMP yields a K(D) of 52 nM, in agreement with its K(I)(sl)(o)(pe). Changing the metal activator from Mg(2+) to Mn(2+) had little effect on the K(I)(sl)(o)(pe) of dGMP or of 8-oxo-dGMP, consistent with the second-sphere enzyme-M(2+)-H(2)O-NTP-M(2+) complex found by NMR (Lin, J., Abeygunawardana, C., Frick, D. N., Bessman, M. J., and Mildvan, A. S. (1997) Biochemistry 36, 1199-1211), but it decreased the K(I) of PPi 12-fold, suggesting direct coordination of the PPi product by the enzyme-bound divalent cation. The tight binding of 8-oxo-dGMP to MutT (DeltaG degrees = -9.8 kcal/mol) is driven by a highly favorable enthalpy ( = -32 +/- 7 kcal/mol), with an unfavorable entropy (<-TDeltaS(o)(binding)> = +22 +/- 7 kcal/mol), as determined by van't Hoff analysis of the effect of temperature on the K(I)(sl)(o)(pe) and by isothermal titration calorimetry in two buffer systems. The binding of 8-oxo-dGMP to MutT induces changes in backbone (15)N and NH chemical shifts of 62 residues widely distributed throughout the protein, while dGMP binding induces smaller changes in only 22 residues surrounding the nucleotide binding site, suggesting that the unusually high affinity of MutT for 8-oxo-nucleotides is due not only to interactions with the altered 8-oxo or 7-NH positions on guanine, but results primarily from diffuse structural changes which tighten the protein structure around the 8-oxo-nucleotide.  相似文献   

2.
The oxidized nucleotide precursors 7, 8-dihydro-8-oxo-dGTP (8-oxo-dGTP) and 1, 2-dihydro-2-oxo-dATP (2-oxo-dATP) are readily incorporated into nascent DNA strands during replication, which would cause base substitution mutations. E. coli MutT and human homologue hMTH1 hydrolyze 8-oxo-dGTP, thereby preventing mutations. In this study, we searched for hMTH1 homologues in the ascidian Ciona intestinalis using the NCBI-BLAST database. Among several candidates, we focused on one open reading frame, designated as CiMutT, because of its high degree of identity (41.7%) and similarity (58.3%) to the overall amino acid sequence of hMTH1, including the Nudix box. CiMutT significantly suppressed the mutator activity of E. coli mutT mutant. Purified CiMutT had a pyrophosphohydrolase activity that hydrolyzed 8-oxo-dGTP to 8-oxo-dGMP and inorganic pyrophosphate. It had a pH optimum of 9.5 and Mg(++) requirement with optimal activity at 5 mM. The activity of CiMutT for 8-oxo-dGTP was comparable to that of hMTH1, while it was 100-fold lower for 2-oxo-dATP than that of hMTH1. These facts indicate that CiMutT is a functional homologue of E. coli MutT. In addition, the enzyme hydrolyzed all four of the unoxidized nucleoside triphosphates, with a preference for dATP. The specific activity for 8-oxo-dGTP was greater than that for unoxidized dATP and dGTP. These results suggest that CiMutT has the potential to prevent mutations by 8-oxo-dGTP in C. intestinalis.  相似文献   

3.
Grochowski LL  Xu H  White RH 《Biochemistry》2008,47(9):3033-3037
Coenzyme F 420 is a hydride carrier cofactor functioning in methanogenesis. One step in the biosynthesis of coenzyme F 420 involves the coupling of 2-phospho- l-lactate (LP) to 7,8-didemethyl-8-hydroxy-5-deazaflavin, the F 420 chromophore. This condensation requires an initial activation of 2-phospho- l-lactate through a pyrophosphate linkage to GMP. Bioinformatic analysis identified an uncharacterized archaeal protein in the Methanocaldococcus jannaschii genome, MJ0887, which could be involved in this transformation. The predicted MJ0887-derived protein has domain similarity with other known nucleotidyl transferases. The MJ0887 gene was cloned and overexpressed, and the purified protein was found to catalyze the formation of lactyl-2-diphospho-5'-guanosine from LP and GTP. Kinetic constants were determined for the MJ0887-derived protein with both LP and GTP substrates and are as follows: V max = 3 micromol min (-1) mg (-1), GTP K M (app) = 56 microM, and k cat/ K M (app) = 2 x 10 (4) M (-1) s (-1) and LP K M (app) = 36 microM, and k cat/ K M (app) = 4 x 10 (4) M (-1) s (-1). The MJ0887 gene product has been designated CofC to indicate its involvement in the third step of coenzyme F 420 biosynthesis.  相似文献   

4.
Lad L  Mewies M  Raven EL 《Biochemistry》2002,41(46):13774-13781
The catalytic mechanism of recombinant soybean cytosolic ascorbate peroxidase (rsAPX) and a derivative of rsAPX in which a cysteine residue (Cys32) located close to the substrate (L-ascorbic acid) binding site has been modified to preclude binding of ascorbate [Mandelman, D., Jamal, J., and Poulos, T. L. (1998) Biochemistry 37, 17610-17617] has been examined using pre-steady-state and steady-state kinetic techniques. Formation (k1 = 3.3 +/- 0.1 x 10(7) M(-1) s(-1)) of Compound I and reduction (k(2) = 5.2 +/- 0.3 x 10(6) M(-1) s(-1)) of Compound I by substrate are fast. Wavelength maxima for Compound I of rsAPX (lambda(max) (nm) = 409, 530, 569, 655) are consistent with a porphyrin pi-cation radical. Reduction of Compound II by L-ascorbate is rate-limiting: at low substrate concentration (0-500 microM), kinetic traces were monophasic but above approximately 500 microM were biphasic. Observed rate constants for the fast phase overlaid with observed rate constants extracted from the (monophasic) dependence observed below 500 microM and showed saturation kinetics; rate constants for the slow phase were linearly dependent on substrate concentration (k(3-slow)) = 3.1 +/- 0.1 x 10(3) M(-1) s(-1)). Kinetic transients for reduction of Compound II by L-ascorbic acid for Cys32-modified rsAPX are monophasic at all substrate concentrations, and the second-order rate constant (k(3) = 0.9 +/- 0.1 x 10(3) M(-1) s(-1)) is similar to that obtained from the slow phase of Compound II reduction for unmodified rsAPX. Steady-state oxidation of L-ascorbate by rsAPX showed a sigmoidal dependence on substrate concentration and data were satisfactorily rationalized using the Hill equation; oxidation of L-ascorbic acid by Cys32-modified rsAPX showed no evidence of sigmoidal behavior. The data are consistent with the presence of two kinetically competent binding sites for ascorbate in APX.  相似文献   

5.
8-oxo-dGTP is generated in the nucleotide pool by direct oxidation of dGTP or phosphorylation of 8-oxo-dGDP. It can be incorporated into DNA during replication, which would result in mutagenic consequences. The frequency of spontaneous mutations remains low in cells owing to the action of enzymes degrading such mutagenic substrates. Escherichia coli MutT and human MTH1 hydrolyze 8-oxo-dGTP to 8-oxo-dGMP. Human NUDT5 as well as human MTH1 hydrolyze 8-oxo-dGDP to 8-oxo-dGMP. These enzymes prevent mutations caused by misincorporation of 8-oxo-dGTP into DNA. In this study, we identified a novel MutT homolog (NDX-1) of Caenorhabditis elegans that hydrolyzes 8-oxo-dGDP to 8-oxo-dGMP. NDX-1 did not hydrolyze 8-oxo-dGTP, 2-hydroxy-dATP or 2-hydroxy-dADP. Expression of NDX-1 significantly reduced spontaneous A:T to C:G transversions and mitigated the sensitivity to a superoxide-generating agent, methyl viologen, in an E. coli mutT mutant. In C. elegans, RNAi of ndx-1 did not affect the lifespan of the worm. However, the sensitivity to methyl viologen and menadione bisulfite of the ndx-1-RNAi worms was enhanced compared with that of the control worms. These facts indicate that NDX-1 is involved in sanitization of 8-oxo-dGDP and plays a critical role in defense against oxidative stress in C. elegans.  相似文献   

6.
Kinetic interactions of beta-lactam antibiotics such as penicillin-G and cefotaxime with normal, penicillin-susceptible PBP2x from Streptococcus pneumoniae and a penicillin-resistant PBP2x (PBP2x(R)) from a resistant clinical isolate (CS109) of the bacterium have been extensively characterized using electrospray mass spectrometry coupled with a fast reaction (quench flow) technique. Kinetic evidence for a two-step acylation of PBP2x by penicillin-G has been demonstrated, and the dissociation constant, K(d) of 0.9 mm, and the acylation rate constant, k(2) of 180 s(-1), have been determined for the first time. The millimolar range K(d) implies that the beta-lactam fits to the active site pocket of the penicillin-sensitive PBP rather poorly, whereas the extremely fast k(2) value indicates that this step contributes most of the binding affinity of the beta-lactam. The values of K(d) (4 mm) and k(2) (0.56 s(-1)) were also determined for PBP2x(R). The combined value of k(2)/K(d), known as overall binding efficiency, for PBP2x(R) (137 m(-1) s(-1)) was over 1000-fold slower than that for PBP2x (200,000 m(-1) s(-1)), indicating that a major part is played by the acylation steps in penicillin resistance. Most of the decreased binding efficiency of PBP2x(R) comes from the decreased ( approximately 300-fold) k(2). Kinetic studies of cefotaxime acylation of the two PBP2x proteins confirmed all of the above findings. Deacylation rate constants (k(3)) for the third step of the interactions were determined to be 8 x 10(-6) s(-1) for penicilloyl-PBP2x and 5.7 x 10(-4) s(-1) for penicilloyl-PBP2x(R), corresponding to over 70-fold increase of the deacylation rate for the resistant PBP2x(R). Similarly, over 80-fold enhancement of the deacylation rate was found for cefotaxime-PBP2x(R) complex (k(3) = 3 x 10(-4) s(-1)) as compared with that of cefotaxime-PBP2x complex (3.5 x 10(-6) s(-1)). This is the first time that such a significant increase of k(3) values was found for a beta-lactam-resistant penicillin-binding protein. These data indicate that the deacylation step also plays a role, which is much more important than previously thought, in PBP2x(R) resistance to beta-lactams.  相似文献   

7.
Coiled coils consist of two or more amphipathic a-helices wrapped around each other to form a superhelical structure stabilized at the interhelical interface by hydrophobic residues spaced in a repeating 3-4 sequence pattern. Dimeric coiled coils have been shown to often form in a single step reaction in which association and folding of peptide chains are tightly coupled. Here, we ask whether such a simple folding mechanism may also apply to the formation of a three-stranded coiled coil. The designed 29-residue peptide LZ16A was shown previously to be in a concentration-dependent equilibrium between unfolded monomer (M), folded dimer (D), and folded trimer (T). We show by time-resolved fluorescence change experiments that folding of LZ16A to D and T can be described by 2M (k1)<==>(k(-1)) D and M + D (k2)<==>(k(-2)) T. The following rate constants were determined (25 degrees C, pH 7): k1 = 7.8 x 10(4) M(-1) s(-1), k(-1) = 0.015 s(-1), k2 = 6.5 x 10(5) M(-1) s(-1), and k(-2) = 1.1 s(-1). In a separate experiment, equilibrium binding constants were determined from the change with concentration of the far-ultraviolet circular dichroism spectrum of LZ16A and were in good agreement with the kinetic rate constants according to K(D) = k1/2k(-1) and K(T) = k2/k(-2). Furthermore, pulsed hydrogen-exchange experiments indicated that only unfolded M and folded D and T were significantly populated during folding. The results are compatible with a two-step reaction in which a subpopulation of association competent (e.g., partly helical) monomers associate to dimeric and trimeric coiled coils.  相似文献   

8.
Pollock VV  Barber MJ 《Biochemistry》2001,40(5):1430-1440
Rhodobacter sphaeroides f. sp. denitrificans biotin sulfoxide reductase catalyzes the reduction of d-biotin d-sulfoxide (BSO) to biotin. Initial rate studies of the homogeneous recombinant enzyme, expressed in Escherichia coli, have demonstrated that the purified protein utilizes NADPH as a facile electron donor in the absence of any additional auxiliary proteins. We have previously shown [Pollock, V. V., and Barber, M. J. (1997) J. Biol. Chem. 272, 3355-3362] that, at pH 8 and in the presence of saturating concentrations of BSO, the enzyme exhibits, a marked preference for NADPH (k(cat,app) = 500 s(-1), K(m,app) = 269 microM, and k(cat,app)/K(m,app) = 1.86 x 10(6) M(-1) s(-1)) compared to NADH (k(cat,app) = 47 s(-1), K(m,app) = 394 microM, and k(cat,app)/K(m,app) = 1.19 x 10(5) M(-1) s(-1)). Production of biotin using NADPH as the electron donor was confirmed by both the disk biological assay and by reversed-phase HPLC analysis of the reaction products. The purified enzyme also utilized ferricyanide as an artificial electron acceptor, which effectively suppressed biotin sulfoxide reduction and biotin formation. Analysis of the enzyme isolated from tungsten-grown cells yielded decreased reduced methyl viologen:BSO reductase, NADPH:BSO reductase, and NADPH:FR activities, confirming that Mo is required for all activities. Kinetic analyses of substrate inhibition profiles revealed that the enzyme followed a Ping Pong Bi-Bi mechanism with both NADPH and BSO exhibiting double competitive substrate inhibition. Replots of the 1/v-axes intercepts of the parallel asymptotes obtained at several low concentrations of fixed substrate yielded a K(m) for BSO of 714 and 65 microM for NADPH. In contrast, utilizing NADH as an electron donor, the replots yielded a K(m) for BSO of 132 microM and 1.25 mM for NADH. Slope replots of data obtained at high concentrations of BSO yielded a K(i) for BSO of 6.10 mM and 900 microM for NADPH. Kinetic isotope studies utilizing stereospecifically deuterated NADPD indicated that BSO reductase uses specifically the 4R-hydrogen of the nicotinamide ring. Cyanide inhibited NADPH:BSO and NADPH:FR activities in a reversible manner while diethylpyrocarbonate treatment resulted in complete irreversible inactivation of the enzyme concomitant with molybdenum cofactor release, indicating that histidine residues are involved in cofactor-binding.  相似文献   

9.
Apparent Ca(2+)-binding constant (K(app)) of Caenorhabditis elegans troponin C (CeTnC) was determined by a fluorescence titration method. The K(app) of the N-domain Ca(2+)-binding site of CeTnC was 7.9+/-1.6 x 10(5) M(-1) and that of the C-domain site was 1.2+/-0.6 x 10(6) M(-1), respectively. Mg(2+)-dependence of the K(app) showed that both Ca(2+)-binding sites did not bind competitively Mg(2+). The Ca(2+) dissociation rate constant (k(off)) of CeTnC was determined by the fluorescence stopped-flow method. The k(off) of the N-domain Ca(2+)-binding site of CeTnC was 703+/-208 s(-1) and that of the C-domain site was 286+/-33 s(-1), respectively. From these values we could calculate the Ca(2+)-binding rate constant (k(on)) as to be 5.6+/-2.8 x 10(8) M(-1) s(-1) for the N-domain site and 3.4+/-2.1 x 10(8) M(-1) s(-1) for the C-domain site, respectively. These results mean that all Ca(2+)-binding sites of CeTnC are low affinity, fast dissociating and Ca(2+)-specific sites. Evolutional function of TnC between vertebrate and invertebrate and biological functions of wild type and mutant CeTnCs are discussed.  相似文献   

10.
Kinetic and binding studies involving a model DNA cytosine-5-methyltransferase, M.HhaI, and a 37-mer DNA duplex containing a single hemimethylated target site were applied to characterize intermediates on the reaction pathway. Stopped-flow fluorescence studies reveal that cofactor S-adenosyl-l-methionine (AdoMet) and product S-adenosyl-l-homocysteine (AdoHcy) form similar rapidly reversible binary complexes with the enzyme in solution. The M.HhaI.AdoMet complex (k(off) = 22 s(-)1, K(D) = 6 microm) is partially converted into products during isotope-partitioning experiments, suggesting that it is catalytically competent. Chemical formation of the product M.HhaI.(Me)DNA.AdoHcy (k(chem) = 0.26 s(-)1) is followed by a slower decay step (k(off) = 0.045 s(-)1), which is the rate-limiting step in the catalytic cycle (k(cat) = 0.04 s(-)1). Analysis of reaction products shows that the hemimethylated substrate undergoes complete (>95%) conversion into fully methylated product during the initial burst phase, indicating that M.HhaI exerts high binding selectivity toward the target strand. The T250N, T250D, and T250H mutations, which introduce moderate perturbation in the catalytic site, lead to substantially increased K(D)(DNA(ternary)), k(off)(DNA(ternary)), K(M)(AdoMet(ternary)) values but small changes in K(D)(DNA(binary)), K(D)(AdoMet(binary)), k(chem), and k(cat). When the target cytosine is replaced with 5-fluorocytosine, the chemistry step leading to an irreversible covalent M.HhaI.DNA complex is inhibited 400-fold (k(chem)(5FC) = 0.7 x 10(-)3 s(-)1), and the Thr-250 mutations confer further dramatic decrease of the rate of the covalent methylation k(chem). We suggest that activation of the pyrimidine ring via covalent addition at C-6 is a major contributor to the rate of the chemistry step (k(chem)) in the case of cytosine but not 5-fluorocytosine. In contrast to previous reports, our results imply a random substrate binding order mechanism for M.HhaI.  相似文献   

11.
In a search for a plant antimutator MutT protein, an Arabidopsis thaliana Nudix hydrolase with homology to the mammalian GFG protein was expressed as a hexahistidine fusion polypeptide in Escherichia coli and purified to homogeneity. Unlike the GFG protein, the A. thaliana homolog could not complement the mutT mutation in a MutT-deficient E. coli strain nor was it able to hydrolyze 8-oxo-dGTP, the main substrate of the MutT protein. Instead the recombinant protein hydrolyzed a variety of nucleoside diphosphate derivatives showing a preference for ADP-ribose, with Km and k(cat) values of 1.2 mM and 2.7 s(-1) respectively. The products of ADP-ribose hydrolysis were AMP and ribose-5-phosphate. The optimal activity was at alkaline pH (8.5) with Mg2+ (5 mM) ions as the cofactor. The protein exists as a dimmer in solution.  相似文献   

12.
Human MTH1 and Escherichia coli MutT proteins hydrolyze 7, 8-dihydro-8-oxo-dGTP (8-oxo-dGTP) to monophosphate, thus avoiding the incorporation of 8-oxo-7,8-dihydroguanine into nascent DNA. Although only 30 amino acid residues (23%) are identical between MTH1 and MutT, there is a highly conserved region consisting of 23 residues (MTH1, Gly(36)-Gly(58)) with 14 identical residues. A chimeric protein MTH1-Ec, in which the 23-residue sequence of MTH1 was replaced with that of MutT, retains its capability to hydrolyze 8-oxo-dGTP, thereby indicating that the 23-residue sequences of MTH1 and MutT are functionally and structurally equivalent and constitute functional modules. By saturation mutagenesis of the module in MTH1, 14 of the 23 residues proved to be essential to exert 8-oxo-dGTPase activity. For the other 9 residues (40, 42, 44, 46, 47, 49, 50, 54, and 58), positive mutants were obtained, and Arg(50) can be replaced with hydrophobic residues (Val, Leu, or Ile), with a greater stability and higher specific activity of the enzyme. Indispensabilities of Val(39), Ile(45), and Leu(53) indicate that an amphipathic property of alpha-helix I consisting of 14 residues of the module (Thr(44)-Gly(58)) is essential to maintain the stable catalytic surface for 8-oxo-dGTPase.  相似文献   

13.
The reaction of mushroom (Agaricus bisporus) tyrosinase with dioxygen in the presence of several o-diphenolic substrates has been studied by steady-state and transient-phase kinetics in order to elucidate the rate-limiting step and to provide new insights into the mechanism of oxidation of these substrates. A kinetic analysis has allowed for the first time the determination of individual rate constants for several of the partial reactions that comprise the catalytic cycle. Mushroom tyrosinase rapidly reacts with dioxygen with a second-order rate constant k(+8) = 2.3 x 10(7) M(-)(1) s(-)(1), which is similar to that reported for hemocyanins [(1.3 x 10(6))-(5.7 x 10(7)) M(-)(1) s(-)(1)]. Deoxytyrosinase binds dioxygen reversibly at the binuclear Cu(I) site with a dissociation constant K(D)(O)()2 = 46.6 microM, which is similar to the value (K(D)(O)()2 = 90 microM) reported for the binding of dioxygen to Octopus vulgaris deoxyhemocyanin [Salvato et al. (1998) Biochemistry 37, 14065-14077]. Transient and steady-state kinetics showed that o-diphenols such as 4-tert-butylcatechol react significantly faster with mettyrosinase (k(+2) = 9.02 x 10(6) M(-)(1) s(-)(1)) than with oxytyrosinase (k(+6) = 5.4 x 10(5) M(-)(1) s(-)(1)). This difference is interpreted in terms of differential steric and polar effects that modulate the access of o-diphenols to the active site for these two forms of the enzyme. The values of k(cat) for several o-diphenols are also consistent with steric and polar factors controlling the mobility, orientation, and thence the reactivity of substrates at the active site of tyrosinase.  相似文献   

14.
Cdc25A is a dual-specific protein phosphatase involved in the regulation of the kinase activity of Cdk-cyclin complexes in the eukaryotic cell cycle. To understand the mechanism of this important regulator, we have generated highly purified biochemical reagents to determine the kinetic constants for human Cdc25A with respect to a set of peptidic, artificial, and natural substrates. Cdc25A and its catalytic domain (dN25A) demonstrate very similar kinetics toward the artificial substrates p-nitrophenyl phosphate (k(cat)/K(m) = 15-25 M(-1) s(-1)) and 3-O-methylfluorescein phosphate (k(cat)/K(m) = 1.1-1.3 x 10(4) M(-1) s(-1)). Phospho-peptide substrates exhibit extremely low second-order rate constants and a flat specificity profile toward Cdc25A and dN25A (k(cat)/K(m) = 1 to 10 M(-1) s(-1)). In contrast to peptidic substrates, Cdc25A and dN25A are highly active phosphatases toward the natural substrate, T14- and Y15-bis-phosphorylated Cdk2/CycA complex (Cdk2-pTpY/CycA) with k(cat)/K(m) values of 1.0-1.1 x 10(6) M(-1) s(-1). In the context of the Cdk2-pTpY/CycA complex, phospho-threonine is preferred over phospho-tyrosine by more than 10-fold. The highly homologous catalytic domain of Cdc25c is essentially inactive toward Cdk2-pTpY/CycA. Taken together these data indicate that a significant degree of the specificity of Cdc25 toward its Cdk substrate resides within the catalytic domain itself and yet is in a region(s) that is outside the phosphate binding site of the enzyme.  相似文献   

15.
MutT distinguishes substrate 8-oxo-dGTP from dGTP and also 8-oxo-dGMP from dGMP despite small differences of chemical structures between them. In this paper we show by the method of molecular dynamics simulation that the transition between conformational substates of MutT is a key mechanism for a high-resolution molecular recognition of the differences between the very similar chemical compounds. (1) The native state MutT has two conformational substates with similar free energies, each characterized by either open or closed of two loops surrounding the substrate binding active site. Between the two substates, the open substate is more stable in free MutT and in dGMP-MutT complex, and the closed substate is more stable in 8-oxo-dGMP-MutT complex. (2) Conformational fluctuation of the open substate is much larger than that of the closed substate. An estimate of associated entropy difference was found to be consistent with the experimentally found difference of entropy contribution to the binding free energies of the two molecules. (3) A hydrogen bond between H7 atom of 8-oxo-dGMP and the sidechain of Asn119 plays a crucial role for maintaining the closed substate in 8-oxo-dGMP-MutT complex. When this hydrogen bond is absent in the H7-deficient dGMP-MutT complex, the closed substate is no more maintained and transition to the more entropically-favored open substate is induced. (4) Thus, this mechanism of the hydrogen bond controlling the relative stabilities of the drastically different two conformational substates enhances the resolution to recognize the small difference of the chemical structures between the two molecules, dGMP and 8-oxo-dGMP.  相似文献   

16.
One-electron oxidation of (6R)-5,6,7,8-tetrahydrobiopterin (H(4)B) by the azide radical generates the radical cation (H(4)B(*)(+)) which rapidly deprotonates at physiological pH to give the neutral trihydrobiopterin radical (H(3)B(*)); pK(a) (H(4)B(*)(+) <==> H(3)B(*) + H(+)) = (5.2 +/- 0.1). In the absence of ascorbate both the H(4)B(*)(+) and H(3)B(*) radicals undergo disproportionation to form quinonoid dihydrobiopterin (qH(2)B) and the parent H(4)B with rate constants k(H(4)B(*)(+) + H(4)B(*)(+)) = 6.5 x 10(3) M(-1) s(-1) and k(H(3)B(*) + H(3)B(*)) = 9.3 x 10(4) M(-1) s(-1), respectively. The H(3)B(*) radical is scavenged by ascorbate (AscH(-)) with an estimated rate constant of k(H(3)B(*) + AscH(-)) similar 1.7 x 10(5) M(-1) s(-1). At physiological pH the pterin rapidly scavenges a range of biological oxidants often associated with cellular oxidative stress and nitric oxide synthase (NOS) dysfunction including hydroxyl ((*)OH), nitrogen dioxide (NO(2)(*)), glutathione thiyl (GS(*)), and carbonate (CO(3)(*-)) radicals. Without exception these radicals react appreciably faster with H(4)B than with AscH(-) with k(*OH + H(4)B) = 8.8 x 10(9) M(-1) s(-1), k(NO(2)(*) + H(4)B) = 9.4 x 10(8) M(-1) s(-1), k(CO(3)(*-) + H(4)B) = 4.6 x 10(9) M(-1) s(-1), and k(GS(*) + H(4)B) = 1.1 x 10(9) M(-1) s(-1), respectively. The glutathione disulfide radical anion (GSSG(*-)) rapidly reduces the pterin to the tetrahydrobiopterin radical anion (H(4)B(*-)) with a rate constant of k(GSSG(*-) + H(4)B) similar 4.5 x 10(8) M(-1) s(-1). The results are discussed in the context of the general antioxidant properties of the pterin and the redox role played by H(4)B in NOS catalysis.  相似文献   

17.
Hsc66 from Escherichia coli is a constitutively expressed hsp70 class molecular chaperone whose activity is coupled to ATP binding and hydrolysis. To better understand the mechanism and regulation of Hsc66, we investigated the kinetics of ATP hydrolysis and the interactions of Hsc66 with nucleotides. Steady-state experiments revealed that Hsc66 has a low affinity for ATP (K(m)(ATP) = 12.7 microM) compared with other hsp70 chaperones. The kinetics of nucleotide binding were determined by analyzing changes in the Hsc66 absorbance spectrum using stopped-flow methods at 23 degrees C. ATP binding results in a rapid, biphasic increase of Hsc66 absorbance at 280 nm; this is interpreted as arising from a two-step process in which ATP binding (k(a)(ATP) = 4.2 x 10(4) M(-1) s(-1), k(d)(ATP) = 1.1 s(-1)) is followed by a slow conformational change (k(conf) = 0. 1 s(-1)). Under single turnover conditions, the ATP-induced transition decays exponentially with a rate (k(decay) = 0.0013 s(-1)) similar to that observed in both steady-state and single turnover ATP hydrolysis experiments (k(hyd) = 0.0014 s(-1)). ADP binding to Hsc66 results in a monophasic transition in the absence (k(a)(ADP) = 7 x 10(5) M(-1) s(-1), k(d)(ADP) = 60 s(-1)) and presence of physiological levels of inorganic phosphate (k(a)(ADP(P(i)) = 0.28 x 10(5) M(-1) s(-1), k(d)(ADP(P(i)) = 9.1 s(-1)). These results indicate that ATP hydrolysis is the rate-limiting step under steady-state conditions and is >10(3)-fold slower than the rate of ADP/ATP exchange. Thus, in contrast to DnaK and eukaryotic forms of hsp70 that have been characterized to date, the R if T equilibrium balance for Hsc66 is shifted in favor of the low peptide affinity T state, and regulation of the reaction cycle is expected to occur at the ATP hydrolysis step rather than at nucleotide exchange.  相似文献   

18.
The equilibria and kinetics of the interactions of proflavine (PR) and its platinum-containing derivative [PtCl(tmen)(2)HNC(13)H(7)(NHCH(2)CH(2))(2)](+) (PRPt) with double-stranded poly(A) have been investigated by spectrophotometry and Joule temperature-jump relaxation at ionic strength 0.1 M, 25 degrees C, and pH 5.2. Spectrophotometric measurements indicate that base-dye interactions are prevailing. T-jump experiments with polarized light showed that effects due to field-induced alignment could be neglected. Both of the investigated systems display two relaxation effects. The kinetic features of the reaction are discussed in terms of a two-step series mechanism in which a precursor complex DS(I) is formed in the fast step, which is then converted to a final complex in the slow step. The rate constants of the fast step are k(1) = (2.5 +/- 0.4) x 10(6) M(-1) s(-1), k(-1) = (2.4 +/- 0.1) x 10(3) s(-1) for poly(A)-PR and k(1) = (2.3 +/- 0.1) x 10(6) M(-1) s(-1), k(-1) = (1.6 +/- 0.2) x 10(3) s(-1) for poly(A)-PRPt. The rate constants for the slow step are k(2) = (4.5 +/- 0.5) x 10(2) s(-1), k(-2) = (1.7 +/- 0.1) x 10(2) s(-1) for poly(A)-PR and k(2) = 9.7 +/- 1.2 s(-1), k(-2) = 10.6 +/- 0.2 s(-1) for poly(A)-PRPt. Spectrophotometric measurements yield for the equilibrium constants and site size the values K = (4.5 +/- 0.1) x 10(3) M(-1), n = 1.3 +/- 0.5 for poly(A)-PR and K = (2.9 +/- 0.1) x 10(3) M(-1), n = 2.3 +/- 0.6 for poly(A)-PRPt. The values of k(1) are similar and lower than expected for diffusion-limited reactions. The values of k(-1) are similar as well. It is suggested that the formation of DS(I) involves only the proflavine residues in both systems. In contrast, the values of k(2) and k(-2) in poly(A)-PRPt are much lower than in poly(A)-PR. The results suggest that in the complex DS(II) of poly(A)-PRPt both proflavine and platinum residues are intercalated. In addition, a very slow process was detected and ascribed to the covalent binding of Pt(II) to the adenine.  相似文献   

19.
20.
Catalytic reaction pathway for the mitogen-activated protein kinase ERK2   总被引:2,自引:0,他引:2  
Prowse CN  Hagopian JC  Cobb MH  Ahn NG  Lew J 《Biochemistry》2000,39(20):6258-6266
The structural, functional, and regulatory properties of the mitogen-activated protein kinases (MAP kinases) have long attracted considerable attention owing to the critical role that these enzymes play in signal transduction. While several MAP kinase X-ray crystal structures currently exist, there is by comparison little mechanistic information available to correlate the structural data with the known biochemical properties of these molecules. We have employed steady-state kinetic and solvent viscosometric techniques to characterize the catalytic reaction pathway of the MAP kinase ERK2 with respect to the phosphorylation of a protein substrate, myelin basic protein (MBP), and a synthetic peptide substrate, ERKtide. A minor viscosity effect on k(cat) with respect to the phosphorylation of MBP was observed (k(cat) = 10 +/- 2 s(-1), k(cat)(eta) = 0.18 +/- 0.05), indicating that substrate processing occurs via slow phosphoryl group transfer (12 +/- 4 s(-1)) followed by the faster release of products (56 +/- 4 s(-1)). At an MBP concentration extrapolated to infinity, no significant viscosity effect on k(cat)/K(m(ATP)) was observed (k(cat)/K(m(ATP)) = 0.2 +/- 0.1 microM(-1) s(-1), k(cat)/K(m(ATP))(eta) = -0.08 +/- 0.04), consistent with rapid-equilibrium binding of the nucleotide. In contrast, at saturating ATP, a full viscosity effect on k(cat)/K(m) for MBP was apparent (k(cat)/K(m(MBP)) = 2.4 +/- 1 microM(-1) s(-1), k(cat)/K(m(MBP))(eta) = 1.0 +/- 0.1), while no viscosity effect was observed on k(cat)/K(m) for the phosphorylation of ERKtide (k(cat)/K(m(ERKtide)) = (4 +/- 2) x 10(-3) microM(-1) s(-1), k(cat)/K(m(ERKtide))(eta) = -0.02 +/- 0.02). This is consistent with the diffusion-limited binding of MBP, in contrast to the rapid-equilibrium binding of ERKtide, to form the ternary Michaelis complex. Calculated values for binding constants show that the estimated value for K(d(MBP)) (/= 1.5 mM). The dramatically higher catalytic efficiency of MBP in comparison to that of ERKtide ( approximately 600-fold difference) is largely attributable to the slow dissociation rate of MBP (/=56 s(-1)), from the ERK2 active site.  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