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 403 毫秒
1.
Hydroxymethylglutaryl-CoA synthase-catalyzed condensation of acetyl-CoA with acetoacetyl-CoA requires enolization/carbanion formation from the acetyl C-2 methyl group prior to formation of a new carbon-carbon bond. Acetyldithio-CoA, a readily enolizable analog of acetyl-CoA, was an effective competitive inhibitor of avian hydroxymethylglutaryl-CoA synthase (Ki = 28 microm). In the absence of cosubstrate, enzyme catalyzed the enolization/proton exchange from the C-2 methyl group of acetyldithio-CoA. Mutant enzymes that exhibited impaired formation of the covalent acetyl-S-enzyme reaction intermediate exhibited diminished (D159A and D203A) or undetectable (C129S) rates of enolization of acetyldithio-CoA. The results suggest that covalent thioacetylation of protein, which has not been detected previously for other enzymes that enolize this analog, occurs with hydroxymethylglutaryl-CoA synthase. Enzyme catalyzed the transfer of the thioacetyl group of this analog to 3'-dephospho-CoA suggesting the intermediacy of a covalent thioacetyl-S-enzyme species, which appears to be important for proton abstraction from C-2 of the thioacetyl group. Avian enzyme glutamate 95 is crucial to substrate condensation to form a new carboncarbon bond. Mutations of this invariant residue (avian enzyme E95A and E95Q; Staphylococcus aureus enzyme E79Q) correlated with diminished ability to catalyze enolization of acetyldithio-CoA. Enolization by E95Q was not stimulated in the presence of acetoacetyl-CoA. These observations suggest either a direct (proton abstraction) or indirect (solvent polarization) role for this active site glutamate.  相似文献   

2.
We investigated the efficiency of glutamic acid 64 and aspartic acid 64 as proton donors to the zinc-bound hydroxide in a series of site-specific mutants of human carbonic anhydrase III (HCA III). Rate constants for this intramolecular proton transfer, a step in the catalyzed dehydration of bicarbonate, were determined from the proton-transfer-dependent rates of release of H2 18O from the enzyme measured by mass spectrometry. The free energy plots representing these rate constants could be fit by the Marcus rate theory, resulting in an intrinsic barrier for the proton transfer of deltaG0++ = 2.2 +/- 0.5 kcal/mol, and a work function or thermodynamic contribution to the free energy of reaction wr = 10.8 +/- 0.1 kcal/mol. These values are very similar in magnitude to the Marcus parameters describing intramolecular proton transfer from His64 and His67 to the zinc-bound hydroxide in mutants of HCA III. That result and the equivalent efficiency of Glu64 and Asp64 as proton donors in the catalysis by CA III demonstrate a lack of specificity in proton transfer from these sites, which is indirect evidence of a number of proton conduction pathways through different structures of intervening water chains. The dominance of the thermodynamic contribution or work function for all of these proton transfers is consistent with the view that formation and breaking of hydrogen bonds in such water chains is a limiting factor for proton translocation.  相似文献   

3.
The hydration of CO(2) and the dehydration of HCO(3)(-) catalyzed by the carbonic anhydrases is accompanied by the transfer of protons between solution and the zinc-bound water molecule in the active site. This transfer is facilitated by amino acid residues of the enzyme which act as intramolecular proton shuttles; variants of carbonic anhydrase lacking such shuttle residues are enhanced or rescued in catalysis by intermolecular proton transfer from donors such as imidazole in solution. The resulting rate constants for proton transfer when compared with the values of the pK(a) of the donor and acceptor give Bronsted plots of high curvature. These data are described by Marcus theory which shows an intrinsic barrier for proton transfer from 1 to 2 kcal/mol and work terms or thermodynamic contributions to the free energy of reaction from 4 to10 kcal/mol. The interpretation of these Marcus parameters is discussed in terms of the well-studied pathway of the catalysis and structure of the enzymes.  相似文献   

4.
In Rhodobacter sphaeroides, transfer of the first electron in quinol oxidation by the bc(1) complex shows kinetic features (a slow rate (approx. 1.5 x 10(3)/s), high activation energy (approx. 65 kJ/mol) and reorganization energy, lambda (2.5 V)) that are unexpected from Marcus theory and the distances shown by the structures. Reduction of the oxidized iron-sulfur protein occurs after formation of the enzyme-substrate complex, and involves a H-transfer in which the electron transfer occurs through the approx. 7 A of a bridging histidine forming a H-bond with quinol and a ligand to 2Fe-2S. The anomalous kinetic features can be explained by a mechanism in which the electron transfer is constrained by coupled transfer of the proton. We discuss this in the context of mutant strains with modified E(m,7) and pK for the iron-sulfur protein, and Marcus theory for proton-coupled electron transfer. We suggest that transfer of the second proton and electron involve movement of semiquinone in the Q(o) site, and rotation of the Glu of the conserved -PEWY- sequence. Mutational studies show a key role for the domain proximal to heme b(L). The effects of mutation at Tyr-302 (Tyr-279 in bovine sequence) point to a possible linkage between conformational changes in the proximal domain, and changes leading to closure of the iron-sulfur protein access channel at the distal domain.  相似文献   

5.
The delta 5-3-ketosteroid isomerase (EC 5.3.3.1) of Pseudomonas testosteroni catalyzes the conversion of androst-5-ene-3,17-dione to androst-4-ene-3,17-dione by a stereospecific transfer of the 4 beta-proton to the 6 beta-position. The reaction involves two steps: (a) a rate-limiting concerted enolization, comprising protonation of the 3-carbonyl oxygen by the phenolic hydroxyl group of Tyr-14 and abstraction of the 4 beta-proton by the carboxylate group of Asp-38, and (b) rapid reketonization of the dienol, which may or may not be concerted. The active-site mutant D38N, which lacks the base responsible for proton transfer, is about 10(6.0)-fold less active catalytically than the wild-type enzyme. With the D38N mutant it was demonstrated spectroscopically that the enzymatic reaction involves the conversion of the substrate to both the dienol and its anion as tightly enzyme-bound intermediates, which are then converted much more slowly to the alpha,beta-unsaturated product. In contrast to the mechanism of the wild-type enzyme, the enolization reaction promoted by the D38N mutant is not stereospecific with respect to removal of the 4 beta-proton and shows primary kinetic isotope effects on enolization when either 4 alpha or 4 beta or both of these protons are replaced by deuterium. Kinetic isotope effects obtained with deuterated substrates, solvent, or combinations of the two indicate that, unlike in the wild-type enzyme, protonation of the carbonyl oxygen and removal of the C-4 proton are not concerted in the D38N mutant.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
We have examined the kinetics and thermodynamics of the exchange of a fluorescent amphiphile derived from a phospholipid, NBD-DMPE, between serum albumin and the serum lipoproteins of high density (HDL2 and HDL3), LDL, and VLDL. Binding of the fluorescent lipid amphiphile to bovine serum albumin is characterized, at 35 degrees C, by an equilibrium binding constant of approximately 3 x 10(6) M(-1) and a characteristic time < or = 0.1 s. Association of NBD-DMPE with the lipoprotein particles, if considered as a partitioning of amphiphile monomers between the aqueous phase and the lipoprotein particles, is characterized by an equilibrium partition coefficient between 10(5) and 10(6), being highest for LDL and lowest for HDL. The association of NBD-DMPE monomers with lipoprotein particles can be described by insertion rate constants on the order of 10(5) M(-1) s(-1) for VLDL and LDL and 10(4) M(-1) s(-1) for HDL. The desorption rate constants are on the order of 10(-5) s(-1) for all particles. The study was performed as a function of temperature between 15 and 35 degrees C. This permitted the calculation of the equilibrium thermodynamic parameters (deltaG(o), deltaH(o), and deltaS(o)) as well as the activation parameters (deltaG++(o), deltaH++(o), and deltaS++(o)) for the insertion and desorption processes. The association equilibrium is dominated by the entropic contribution to the free energy in all cases. The results are discussed in relation to phospholipid and amphiphile exchange phenomena involving the lipoproteins.  相似文献   

7.
Sims PA  Larsen TM  Poyner RR  Cleland WW  Reed GH 《Biochemistry》2003,42(27):8298-8306
The pH dependence of enolase catalysis was studied to understand how enolase is able to utilize both general acid and general base catalysis in each direction of the reaction at near-neutral pHs. Wild-type enolase from yeast was assayed in the dehydration reaction (2-phospho-D-glycerate --> phosphoenolpyruvate + H(2)O) at different pHs. E211Q, a site-specific variant of enolase that catalyzes the exchange of the alpha-proton of 2-phospho-D-glycerate but not the complete dehydration, was assayed in a (1)H/(2)H exchange reaction at different pDs. Additionally, crystal structures of E211Q and E168Q were obtained at 2.0 and 1.8 A resolution, respectively. Analysis of the pH profile of k(cat)/K(Mg) for wild-type enolase yielded macroscopic pK(a) estimates of 7.4 +/- 0.3 and 9.0 +/- 0.3, while the results of the pD profile of the exchange reaction of E211Q led to a pK(a) estimate of 9.5 +/- 0.1. These values permit estimates of the four microscopic pK(a)s that describe the four relevant protonation states of the acid/base catalytic groups in the active site. The analysis indicates that the dehydration reaction is catalyzed by a small fraction of enzyme that is reverse-protonated (i.e., Lys345-NH(2), Glu211-COOH), whereas the hydration reaction is catalyzed by a larger fraction of the enzyme that is typically protonated (i.e., Lys345-NH(3)(+), Glu211-COO(-)). These two forms of the enzyme coexist in a constant, pH-independent ratio. The structures of E211Q and E168Q both show virtually identical folds and active-site architectures (as compared to wild-type enolase) and thus provide additional support to the conclusions reported herein. Other enzymes that require both general acid and general base catalysis likely require reverse protonation of catalytic groups in one direction of the reaction.  相似文献   

8.
Flavocytochrome b2 (or L-lactate dehydrogenase) from baker's yeast is thought to operate by the initial formation of a carbanion, as do the evolutionarily related alpha-hydroxy acid-oxidizing FMN-dependent oxidases. Previous work has shown that, in the active site of the unligated reduced flavocytochrome b2, the group that has captured the substrate alpha-proton has a high pKapp, calculated to lie around 15 through the use of Eigen's equation. A detailed inspection of the now known three-dimensional structure of the enzyme leads to the conclusion that the high pKa belongs to His 373, an active site group that plays the role of general base in the forward reaction and of general acid in the reverse direction. Moreover, consideration of the kinetics of proton transfer during the catalytic cycle suggests that the pKa of the reduced FMN N5 position should be lowered by several pH units compared to its pKa of 20 or more when free. The features of the three-dimensional structure possibly responsible for these pK shifts are analyzed; they are proposed to consist of a network of hydrogen bonds with the solvent and of a mutual electrostatic stabilization of anionic reduced flavin and the imidazolium ion. Finally, it is suggested that similar pK shifts affect the active sites of the alpha-hydroxy acid-oxidizing flavooxidases, which are homologous to flavocytochrome b2. The functional significance of these pK shifts in terms of catalysis and semiquinone stabilization is discussed.  相似文献   

9.
The rate-limiting reaction of the bc(1) complex from Rhodobacter sphaeroides is transfer of the first electron from ubihydroquinone (quinol, QH(2)) to the [2Fe-2S] cluster of the Rieske iron-sulfur protein (ISP) at the Q(o)-site. Formation of the ES-complex requires participation of two substrates (S), QH(2) and ISP(ox). From the variation of rate with [S], the binding constants for both substrates involved in formation of the complex can be estimated. The configuration of the ES-complex likely involves the dissociated form of the oxidized ISP (ISP(ox)) docked at the b-interface on cyt b, in a complex in which N(epsilon) of His-161 (bovine sequence) forms a H-bond with the quinol -OH. A coupled proton and electron transfer occurs along this H-bond. This brief review discusses the information available on the nature of this reaction from kinetic, structural and mutagenesis studies. The rate is much slower than expected from the distance involved, likely because it is controlled by the low probability of finding the proton in the configuration required for electron transfer. A simplified treatment of the activation barrier is developed in terms of a probability function determined by the Br?nsted relationship, and a Marcus treatment of the electron transfer step. Incorporation of this relationship into a computer model allows exploration of the energy landscape. A set of parameters including reasonable values for activation energy, reorganization energy, distances between reactants, and driving forces, all consistent with experimental data, explains why the rate is slow, and accounts for the altered kinetics in mutant strains in which the driving force and energy profile are modified by changes in E(m) and/or pK of ISP or heme b(L).  相似文献   

10.
To elucidate a detailed catalytic mechanism for nitrile hydratases (NHases), the pH and temperature dependence of the kinetic constants k(cat) and K(m) for the cobalt-type NHase from Pseudonocardia thermophila JCM 3095 (PtNHase) were examined. PtNHase was found to exhibit a bell-shaped curve for plots of relative activity versus pH at pH 3.2-11 and was found to display maximal activity between pH 7.2 and 7.8. Fits of these data provided pK(E)(S1) and pK(E)(S2) values of 5.9 +/- 0.1 and 9.2 +/- 0.1 (k(cat)' = 130 +/- 1 s(-1)), respectively, and pK(E)(1) and pK(E)(2) values of 5.8 +/- 0.1 and 9.1 +/- 0.1 (k(cat)'/K(m)' = (6.5 +/- 0.1) x 10(3) s(-1) mm(-1)), respectively. Proton inventory studies indicated that two protons are transferred in the rate-limiting step of the reaction at pH 7.6. Because PtNHase is stable at 60 degrees C, an Arrhenius plot was constructed by plotting ln(k(cat)) versus 1/T, providing E(a) = 23.0 +/- 1.2 kJ/mol. The thermal stability of PtNHase also allowed DeltaH(0) ionization values to be determined, thus helping to identify the ionizing groups exhibiting the pK(E)(S1) and pK(E)(S2) values. Based on DeltaH(0)(ion) data, pK(E)(S1) is assigned to betaTyr(68), whereas pK(E)(S2) is assigned to betaArg(52), betaArg(157), or alphaSer(112) (NHases are alpha(2)beta(2)-heterotetramers). A combination of these data with those previously reported for NHases and synthetic model complexes, along with sequence comparisons of both iron- and cobalt-type NHases, allowed a novel catalytic mechanism for NHases to be proposed.  相似文献   

11.
R67 is a Type II dihydrofolate reductase (DHFR) that catalyzes the reduction of dihydrofolate (DHF) to tetrahydrofolate by facilitating the addition of a proton to N5 of DHF and the transfer of a hydride ion from NADPH to C6. Because this enzyme is a plasmid-encoded DHFR from trimethoprim-resistant bacteria, extensive studies on R67 with various methods have been performed to elucidate its reaction mechanism. Here, Raman difference measurements, conducted on the ternary complex of R67.NADP(+).DHF believed to be an accurate mimic of the productive DHFR.NADPH.DHF complex, show that the pK(a) of N5 in the complex is less than 4. This is in clear contrast to the behavior observed in Escherichia coli DHFR, a substantially more efficient enzyme, where the pK(a) of bound DHF at N5 is increased to 6.5 compared with its solution value of 2.6. A comparison of the ternary complexes in R67 and E. coli DHFRs suggests that enzymic raising of the pK(a) at N5 can significantly increase the catalytic efficiency of the hydride transfer step. However, R67 shows that even without such a strategy an effective DHFR can still be designed.  相似文献   

12.
Cytochrome c oxidase from Rhodobacter sphaeroides has two proton-input pathways leading from the protein surface towards the catalytic site, located within the membrane-spanning part of the enzyme. One of these pathways, the D-pathway, contains a highly conserved Glu residue [E(I-286)], which plays an important role in proton transfer through the pathway. In a recent study, we showed that a mutant enzyme in which E(I-286) was re-located to the opposite side of the D-pathway [EA(I-286)/IE(I-112) double mutant enzyme] was able to pump protons, although with a stoichiometry that was lower than that of the wild-type enzyme (approximately 0.6 H(+)/e(-)) (Aagaard et al. (2000) Biochemistry 39, 15847-15850). These results showed that the residue must not necessarily be located at a specific place in the amino-acid sequence, but rather at a specific location in space. In this study, we have investigated the effect of moving E(I-286) on the kinetics of specific reaction steps of the catalytic cycle in the pH range 6-11. Our results show that during the reaction of the four-electron reduced enzyme with O(2), the rates of the two first transitions (up to formation of the 'peroxy' intermediate, P(r)) are the same for the double mutant as for the wild-type enzyme, but formation of the oxo-ferryl (F) and fully oxidized (O) states, associated with proton uptake from the bulk solution, are slowed by factors of approximately 30 and approximately 400, respectively. Thus, in spite of the dramatically reduced transition rates, the proton-pumping stoichiometry is reduced only by approximately 40%. The apparent pK(a) values in the pH-dependencies of the rates of the P(R)-->F and F-->O transitions were >3 and approximately 2 units lower than those of the corresponding transitions in the wild-type enzyme, respectively. The relation between the modified pK(a)s, the transition rates between oxygen intermediates and the pumping stoichiometry is discussed.  相似文献   

13.
Co-ordination of catalytic Zn2+ in sorbitol/xylitol dehydrogenases of the medium-chain dehydrogenase/reductase superfamily involves direct or water-mediated interactions from a glutamic acid residue, which substitutes a homologous cysteine ligand in alcohol dehydrogenases of the yeast and liver type. Glu154 of xylitol dehydrogenase from the yeast Galactocandida mastotermitis (termed GmXDH) was mutated to a cysteine residue (E154C) to revert this replacement. In spite of their variable Zn2+ content (0.10-0.40 atom/subunit), purified preparations of E154C exhibited a constant catalytic Zn2+ centre activity (kcat) of 1.19+/-0.03 s(-1) and did not require exogenous Zn2+ for activity or stability. E154C retained 0.019+/-0.003% and 0.74+/-0.03% of wild-type catalytic efficiency (kcat/K(sorbitol)=7800+/-700 M(-1) x s(-1)) and kcat (=161+/-4 s(-1)) for NAD+-dependent oxidation of sorbitol at 25 degrees C respectively. The pH profile of kcat/K(sorbitol) for E154C decreased below an apparent pK of 9.1+/-0.3, reflecting a shift in pK by about +1.7-1.9 pH units compared with the corresponding pH profiles for GmXDH and sheep liver sorbitol dehydrogenase (termed slSDH). The difference in pK for profiles determined in 1H2O and 2H2O solvent was similar and unusually small for all three enzymes (approximately +0.2 log units), suggesting that the observed pK in the binary enzyme-NAD+ complexes could be due to Zn2+-bound water. Under conditions eliminating their different pH-dependences, wild-type and mutant GmXDH displayed similar primary and solvent deuterium kinetic isotope effects of 1.7+/-0.2 (E154C, 1.7+/-0.1) and 1.9+/-0.3 (E154C, 2.4+/-0.2) on kcat/K(sorbitol) respectively. Transient kinetic studies of NAD+ reduction and proton release during sorbitol oxidation by slSDH at pH 8.2 show that two protons are lost with a rate constant of 687+/-12 s(-1) in the pre-steady state, which features a turnover of 0.9+/-0.1 enzyme equivalents as NADH was produced with a rate constant of 409+/-3 s(-1). The results support an auxiliary participation of Glu154 in catalysis, and possible mechanisms of proton transfer in sorbitol/xylitol dehydrogenases are discussed.  相似文献   

14.
The nature of the proton donor to the C-3 of the enolate of pyruvate, the intermediate in the reaction catalyzed by yeast pyruvate kinase, was investigated by site-directed mutagenesis and physical and kinetic analyses. Thr-298 is correctly located to function as the proton donor. T298S and T298A were constructed and purified. Both mutants are catalytically active with a decrease in k(cat) and k(cat)/K(m)(,PEP). Mn(2+)-activated T298S and T298A do not exhibit homotropic kinetic cooperativity with phosphoenolpyruvate (PEP) in the absence of fructose 1,6-bisphosphate, although PEP binding to enzyme-Mn(2+) is cooperative. The pH dependence of k(cat) for T298A indicates the loss of pK(a)(,2) = 6.4-6.9. Thr-298 affects the ionization (pK(a) approximately 6.5) responsible for modulation of k(cat). Fluorescence studies show altered dissociation constants of ligands to each enzyme complex upon Thr-298 mutations. The rates of the phosphoryl transfer and proton transfer steps in the pyruvate kinase-catalyzed reaction are altered; pyruvate enolization is affected to a greater extent. Proton inventory studies demonstrate solvent isotope effects on k(cat) and k(cat)/K(m)(,PEP). Fractionation factors are metal-dependent and significantly <1. The data suggest that a water molecule in a water channel is the direct proton donor to enolpyruvate and that Thr-298 affects a late step in catalysis.  相似文献   

15.
The catalytic mechanism underlying the aminopeptidase from Streptomyces griseus (SGAP) was investigated. pH-dependent activity profiles revealed the enthalpy of ionization for the hydrolysis of leucine-para-nitroanilide by SGAP. The value obtained (30 +/- 5 kJ.mol(-1)) is typical of a zinc-bound water molecule, suggesting that the zinc-bound water/hydroxide molecule acts as the reaction nucleophile. Fluoride was found to act as a pure noncompetitive inhibitor of SGAP at pH values of 5.9-8 with a K(i) of 11.4 mM at pH 8.0, indicating that the fluoride ion interacts equally with the free enzyme as with the enzyme-substrate complex. pH-dependent pK(i) experiments resulted in a pK(a) value of 7.0, suggesting a single deprotonation step of the catalytic water molecule to an hydroxide ion. The number of proton transfers during the catalytic pathway was determined by monitoring the solvent isotope effect on SGAP and its general acid-base mutant SGAP(E131D) at different pHs. The results indicate that a single proton transfer is involved in catalysis at pH 8.0, whereas two proton transfers are implicated at pH 6.5. The role of Glu131 in binding and catalysis was assessed by determining the catalytic constants (K(m), k(cat)) over a temperature range of 293-329 degrees K for both SGAP and the E131D mutant. For the binding step, the measured and calculated thermodynamic parameters for the reaction (free energy, enthalpy and entropy) for both SGAP and the E131D mutant were similar. By contrast, the E131D point mutation resulted in a four orders of magnitude decrease in k(cat), corresponding to an increase of 9 kJ.mol(-1) in the activation energy for the E131D mutant, emphasizing the crucial role of Glu131 in catalysis.  相似文献   

16.
The pyruvate dehydrogenase from Escherichia coli showed a primary kinetic isotope effect when its overall reaction or the partial reaction of the pyruvate dehydrogenase component were tested in deuterium oxide. The Michaelis constants for pyruvate were nearly unchanged, but the maximum velocities in water and deuterium oxide differed, their ratio being DV = 1.7 for the overall reaction and DV = 2.1 for the E1p reaction. The pH profile and, accordingly, the delta pK1 and delta pK2 values were shifted by 0.6 units to higher pL values. A linear proton inventory curve was obtained when varying the atom fractions of protons relative to deuterons from 100 to 0%. This is an indication for a single proton transfer. It is proposed that this relatively weak primary isotope effect may be caused by the protonation of the N1' nitrogen at the pyrimidine ring of the cofactor by an adjacent glutamate residue. The proton of its carboxylic group exchanges very fast with deuterons of the solvent.  相似文献   

17.
The role of two glutamate residues (E164 and E144) in the active site of enoyl-CoA hydratase has been probed by site-directed mutagenesis. The catalytic activity of the E164Q and E144Q mutants has been determined using 3'-dephosphocrotonyl-CoA. Removal of the 3'-phosphate group reduces the affinity of the substrate for the enzyme, thereby facilitating the determination of K(m) and simplifying the analysis of the enzymes' pH dependence. k(cat) for the hydration of 3'-dephosphocrotonyl-CoA is reduced 7700-fold for the E144Q mutant and 630000-fold for the E164Q mutant, while K(m) is unaffected. These results indicate that both glutamate residues play crucial roles in the hydration chemistry catalyzed by the enzyme. Previously, we reported that, in contrast to the wild-type enzyme, the E164Q mutant was unable to exchange the alpha-proton of butyryl-CoA with D(2)O [D'Ordine, R. L., Bahnson, B. J., Tonge, P. J. , and Anderson, V. E. (1994) Biochemistry 33, 14733-14742]. Here we demonstrate that E144Q is also unable to catalyze alpha-proton exchange even though E164, the glutamate that is positioned to abstract the alpha-proton, is intact in the active site. The catalytic function of each residue has been further investigated by exploring the ability of the wild-type and mutant enzymes to eliminate 2-mercaptobenzothiazole from 4-(2-benzothiazole)-4-thiabutanoyl-CoA (BTTB-CoA). As expected, reactivity toward BTTB-CoA is substantially reduced (690-fold) for the E164Q enzyme compared to wild-type. However, E144Q is also less active than wild-type (180-fold) even though elimination of 2-mercaptobenzothiazole (pK(a) 6.8) should require no assistance from an acid catalyst. Clearly, the ability of E164 to function as an acid-base in the active site is affected by mutation of E144 and it is concluded that the two glutamates act in concert to effect catalysis.  相似文献   

18.
Johnson CM  Huang B  Roderick SL  Cook PF 《Biochemistry》2004,43(49):15534-15539
The pH dependence of kinetic parameters was determined in both reaction directions to obtain information about the acid-base chemical mechanism of serine acetyltransferase from Haemophilus influenzae (HiSAT). The maximum rates in both reaction directions, as well as the V/K(serine) and V/K(OAS), decrease at low pH, exhibiting a pK of approximately 7 for a single enzyme residue that must be unprotonated for optimum activity. The pH-independent values of V(1)/E(t), V(1)/K(serine)E(t), V/K(AcCoA)E(t), V(2)/E(t), V(2)/K(OAS)E(t), and V/K(CoA)E(t) are 3300 +/- 180 s(-1), (9.6 +/- 0.4) x 10(5) M(-1) s(-1), 3.3 x 10(6) M(-1) s(-1), 420 +/- 50 s(-1), (2.1 +/- 0.5) x 10(4) M(-1) s(-1), and (4.2 +/- 0.7) x 10(5) M(-1) s(-1), respectively. The K(i) values for the competitive inhibitors glycine and l-cysteine are pH-independent. The solvent deuterium kinetic isotope effects on V and V/K in the direction of serine acetylation are 1.9 +/- 0.2 and 2.5 +/- 0.4, respectively, and the proton inventories are linear for both parameters. Data are consistent with a single proton in flight in the rate-limiting transition state. A general base catalytic mechanism is proposed for the serine acetyltransferase. Once acetyl-CoA and l-serine are bound, an enzymic general base accepts a proton from the l-serine side chain hydroxyl as it undergoes a nucleophilic attack on the carbonyl of acetyl-CoA. The same enzyme residue then functions as a general acid, donating a proton to the sulfur atom of CoASH as the tetrahedral intermediate collapses, generating the products OAS and CoASH. The rate-limiting step in the reaction at limiting l-serine levels is likely formation of the tetrahedral intermediate between serine and acetyl-CoA.  相似文献   

19.
20.
J S Williams  P R Rosevear 《Biochemistry》1991,30(26):6412-6416
The Escherichia coli truncated methionyl-tRNA synthetase (delta MTS) was shown to catalyze alpha-carbon hydrogen-deuterium exchange of L-selenomethionine, L-methionine, L-ethionine, and L-norleucine in the presence of deuterium oxide. The rate of alpha-proton exchange for L-methionine was shown to be linear with respect to delta MTS concentration. The exchange reaction showed saturation kinetics with apparent Km values of 21 and 4 mM in the absence and presence of saturating adenosine concentrations, respectively. As expected, delta MTS did not catalyze alpha-proton exchange of D-methionine since the enzyme has been shown to be specific for L-amino acids. In the absence of enzyme or in the presence of an equivalent concentration of Zn2+, no hydrogen-deuterium exchange was detected. The exchange reaction was not observed with L-methioninol, an analogue of L-methionine lacking the carboxylate group. These results suggest that the alpha-carboxylate group is a requirement for the delta MTS-catalyzed exchange reaction. The E. coli methionyl-tRNA synthetase (MTS) has previously been shown to be a zinc metalloprotein [Posorske, L. H., Cohn, M., Yanagisawa, N., & Auld, D. S. (1979) Biochim. Biophys. Acta 576, 128]. On the basis of the structural and mechanistic information available on MTS, we propose that the enzyme-bound zinc coordinates the carboxylate of the amino acid, while a base on the enzyme is responsible for exchange of the alpha-proton. The role of the enzyme-bound metal is to render the alpha-proton more acidic through coordination of the carboxylate group.(ABSTRACT TRUNCATED AT 250 WORDS)  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