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.
Delta 1-Piperidine 2-carboxylic acid, an alpha-imino acid, is reduced by 1,4-dihydropyridines to pipecolic acid, an alpha-amino acid, and the corresponding pyridinium ions. This nonenzymatic reaction occurs only in the direction of pipecolic acid production. Glutamate dehydrogenase catalyzes this reaction when the reductant is NADPH and gives as products L-pipecolic acid and NADP+. The reaction velocity for the enzyme-catalyzed reaction is measurable in either direction. The pH-independent equilibrium constant, Keq, for the reduction of the imino acid by NADPH to give pipecolic acid anion and NADP+ was determined from the equilibrium conditions and the pKa values of pipecolic acid (10.72) and of the cyclic imino acid (8.10). The value of Keq was found to be 175 +/- 30; the values of delta G0, delta H0 and delta S0 are -3.1 +/- 0.1 kcal/mol, 5 +/- 1 kcal/mol and 27 +/- 4 e.u., respectively. The data indicate that the reactants are far more solvated than the products and that there must be a large degree of solvent reorganization during the course of the reaction. If these thermodynamic parameters apply to the redox step of the enzyme-catalyzed glutamate reaction, then the burst phase which results upon mixing the enzyme, L-glutamate and NADP+ in stoichiometric amounts must contain a hidden nonredox step of large delta H0 value to account for the curved Arrhenius plot observed for this phase (A.H. Colen, R.T. Medary and H.F. Fisher, Biopolymers 20 (1981) 879).  相似文献   

2.
The glutamate dehydrogenase catalyzed reduction of delta 1-pyrroline-2-carboxylic acid (PCA; an alpha-imino acid) with reduced nicotinamide adenine dinucleotide phosphate (NADPH) to give L-proline and NADP+ is employed as a model for the redox step of the corresponding enzyme-catalyzed reductive amination of alpha-ketoglutarate. We demonstrate the reversibility of the model reaction and measure its equilibrium constant. The pH profiles for the model reactions show that the active substrates are the N-protonated imino acid in one direction and the proline anion with a neutral amino group in the other. The V/K value for the imino acid reduction is enhanced by a group Z of pK = 8.6 in the enzyme-NADPH complex, while that for the proline reaction is unaffected by any such group in the enzyme-NADP+ complex. The following conclusions emerge from a comparison of the pH dependence of the rates for the model reactions with that for the oxidative deamination of L-glutamate [Rife, J. E., & Cleland, W. W. (1980) Biochemistry 19, 2328]. The N-protonated form of alpha-iminoglutarate and the conjugate base of glutamate are the active substrates. The redox step is not sensitive to the protonation state of the groups that catalyze the hydrolysis of bound alpha-iminoglutarate. The group Z, which facilitates the PCA reaction, plays no role in the binding of alpha-ketoglutarate. We propose a chemical mechanism for the glutamate reaction where an unprotonated enzyme group of pK = 5.2 in enzyme-NADPH catalyzes the conversion of the alpha-iminoglutarate to the carbinolamine.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
An activation domain in p67(phox) (residues within 199-210) is essential for cytochrome b(558)-dependent activation of NADPH superoxide (O2(-.)) generation in a cell-free system (Han, C.-H., Freeman, J. L. R., Lee, T., Motalebi, S. A., and Lambeth, J. D. (1998) J. Biol. Chem. 273, 16663-16668). To determine the steady state reduction flavin in the presence of highly absorbing hemes, 8-nor-8-S-thioacetamido-FAD ("thioacetamido-FAD") was reconstituted into the flavocytochrome, and the fluorescence of its oxidized form was monitored. Thioacetamido-FAD-reconstituted cytochrome showed lower activity (7% versus 100%) and increased steady state flavin reduction (28 versus <5%) compared with the enzyme reconstituted with native FAD. Omission of p67(phox) decreased the percent steady state reduction of the flavin to 4%, but omission of p47(phox) had little effect. The activation domain on p67(phox) was critical for regulating flavin reduction, since mutations in this region that decreased O2(-.) generation also decreased the steady state reduction of flavin. Thus, the activation domain on p67(phox) regulates the reductive half-reaction for FAD. This reaction is comprised of the binding of NADPH followed by hydride transfer to the flavin. Kinetic deuterium isotope effects along with K(m) values permitted calculation of the K(d) for NADPH. (R)-NADPD but not (S)-NADPD showed kinetic deuterium isotope effects on V and V/K of about 1.9 and 1.5, respectively, demonstrating stereospecificity for the R hydride transfer. The calculated K(d) for NADPH was 40 microM in the presence of wild type p67(phox) and was approximately 55 microM using the weakly activating p67(phox)(V205A). Thus, the activation domain of p67(phox) regulates the reduction of FAD but has only a small effect on NADPH binding, consistent with a dominant effect on hydride/electron transfer from NADPH to FAD.  相似文献   

4.
Dynamics of the quaternary conformational change in trout hemoglobin   总被引:2,自引:0,他引:2  
The kinetics of conformational changes in trout hemoglobin I have been characterized over the temperature range 2-65 degrees C from time-resolved absorption spectra measured following photodissociation of the carbon monoxide complex. Changes in the spectra of the deoxyheme photoproduct were used to monitor changes in the protein conformation. Although the deoxyheme spectral changes are only about 8% of the total spectral change due to ligand rebinding, a combination of high-precision measurements and singular value decomposition of the data permits a detailed analysis of both their amplitudes and relaxation rates. Systematic variation of the degree of photolysis was used to alter the distribution of liganded tetramers, permitting the assignment of the spectral relaxation at 20 microseconds to the R----T quaternary conformational change of the zero-liganded and singly liganded molecules and spectral relaxations at about 50 ns and 2 microseconds to tertiary conformational changes within the R structure. Analysis of the effect of photoselection by the linearly polarized excitation pulse indicates that a major contribution to the apparent geminate rebinding in the 50-ns relaxation arises from rotational diffusion of molecules containing unphotolyzed heme-CO complexes. The activation enthalpy and activation entropy for the R0----T0 transition are +7.4 kcal/mol and -12 cal mol-1 K-1. Using the equilibrium data, delta H = +29.4 kcal/mol and delta S = +84.4 cal mol-1 K-1 [Barisas, B. G., & Gill, S. J. (1979) Biophys. Chem. 9, 235-244], the activation parameters for the T0----R0 transition are calculated to be delta H = +37 kcal/mol and delta S = +73 cal mol-1 K-1. The similarity of the equilibrium and activation parameters for the T0----R0 transition indicates that the transition state is much more R-like than T-like. This result suggests that in the path from T0 to R0 the subunits have already almost completely rearranged into the R configuration when the transition state is reached, while in the path from R0 to T0 the subunits remain in a configuration close to R in the transition state. The finding of an R-like transition state explains why the binding of ligands causes much smaller changes in the R----T rates than in the T----R rates.  相似文献   

5.
The O(2)(-) generating NADPH oxidase complex of neutrophils comprises two sets of components, namely a membrane-bound heterodimeric flavocytochrome b which contains the redox centers of the oxidase and water-soluble proteins of cytosolic origin which act as activating factors of the flavocytochrome. The NADPH oxidase can be activated in a cell-free system consisting of plasma membranes and cytosol from resting neutrophils in the presence of GTPgammaS and arachidonic acid. NADPH oxidase activation is inhibited by phenylarsine oxide (PAO), a sulfhydryl reagent for vicinal or proximal thiol groups. The site of action of PAO was localized by photolabeling in the beta-subunit of flavocytochrome b [Doussière, J., Poinas, A, Blais, C., and Vignais, P. V. (1998) Eur. J. Biochem. 251, 649-658]. Moreover, the spin state of heme b is controlled by interaction of arachidonic acid with the flavocytochrome b [Doussière, J., Gaillard, J., and Vignais, P. V. (1996) Biochemistry 35, 13400-13410]. Here we report that the promoting effect of arachidonic acid on the activation of NADPH oxidase is due to specific binding of arachidonic acid to flavocytochrome b. Elicitation of NADPH oxidase activity by arachidonic acid is in part associated with an increased affinity of flavocytochrome b for O(2), an effect that was counteracted by the methyl ester of arachidonic acid. On the other hand, the affinity for NADPH was not affected by arachidonic acid. We further demonstrate that PAO antagonizes the effect of arachidonic acid on oxidase activation by decreasing the affinity of the oxidase for O(2), but not for NADPH. PAO induced a change in the spin state of heme b, as arachidonic acid does, with, however, some differences in the constraints imposed to the heme. It is concluded that the opposite effects of arachidonic acid and PAO are exerted on the beta-subunit of flavocytochrome b at two different interacting sites.  相似文献   

6.
Glutamate dehydrogenase catalyzes the reversible oxidation of L-proline and L-pipecolic acid to the corresponding cyclic alpha-imino acids. The active substrates are the amino acid anion in one direction and the iminium ion in the other. The oxidation of the ester, amide, and N-methyl derivatives of L-proline by enzyme-NADP+ and the reduction of N-methyl-delta 1-tetrahydropyridinium ion by enzyme-NADPH do not proceed to a detectable extent under the experimental conditions. The methyliminium ion, however, undergoes facile nonenzymatic reduction by NADPH. If it is assumed that the nonenzymatic reaction reflects the structural requirements of the redox step of the enzymatic reaction, then the lack of reactivity of the tetrahydropyridinium ion toward enzyme-NADPH must be due to the instability of the central complex. It appears that the alpha-carboxylate and NH groups in the amino acid anion and in the alpha-imino acid are involved in binding the substrates to the enzyme-coenzyme complexes. We conclude that each of these active substrates binds to its appropriate enzyme-coenzyme complex through a hydrogen bond between its NH group and a basic enzyme group; there is also an ionic bond in the central complex between the alpha-carboxylate group of the active substrate and a positively charged enzyme group. The five-membered amino acid anions are more reactive toward enzyme-NADP+ than the six-membered ones. The same reactivity order is seen for the reduction of imino acids by enzyme-NADPH. Since these effects are also present in the nonenzymatic reduction by NADPH we ascribe the ring size effects on V/Ksubstrate primarily to those on the hydride-transfer step.  相似文献   

7.
Hicks SN  Smiley RD  Hamilton JB  Howell EE 《Biochemistry》2003,42(36):10569-10578
R67 dihydrofolate reductase (DHFR), which catalyzes the NADPH dependent reduction of dihydrofolate to tetrahydrofolate, belongs to a type II family of R-plasmid encoded DHFRs that confer resistance to the antibacterial drug trimethoprim. Crystal structure data reveals this enzyme is a homotetramer that possesses a single active site pore. Only two charged residues in each monomer are located near the pore, K32 and K33. Site-directed mutants were constructed to probe the role of these residues in ligand binding and/or catalysis. As a result of the 222 symmetry of this enzyme, mutagenesis of one residue results in modification at four related sites. All mutants at K32 affected the quaternary structure, producing an inactive dimer. The K33M mutant shows only a 2-4-fold effect on K(m) values. Salt effects on ligand binding and catalysis for K33M and wildtype R67 DHFRs were investigated to determine if these lysines are involved in forming ionic interactions with the negatively charged substrates, dihydrofolate (overall charge of -2) and NADPH (overall charge of -3). Binding studies indicate that two ionic interactions occur between NADPH and R67 DHFR. In contrast, the binding of folate, a poor substrate, to R67 DHFR.NADPH appears weak as a titration in enthalpy is lost at low ionic strength. Steady-state kinetic studies for both wild type (wt) and K33M R67 DHFRs also support a strong electrostatic interaction between NADPH and the enzyme. Interestingly, quantitation of the observed salt effects by measuring the slopes of the log of ionic strength versus the log of k(cat)/K(m) plots indicates that only one ionic interaction is involved in forming the transition state. These data support a model where two ionic interactions are formed between NADPH and symmetry related K32 residues in the ground state. To reach the transition state, an ionic interaction between K32 and the pyrophosphate bridge is broken. This unusual scenario likely arises from the constraints imposed by the 222 symmetry of the enzyme.  相似文献   

8.
Arg-127 stabilizes the oxyanion of the tetrahedral intermediate formed during Zn2+ carboxypeptidase A-catalyzed hydrolysis. Mutant carboxypeptidases lacking Arg-127 exhibit substantially reduced rates of hydrolysis with the change manifest almost entirely in kcat (kcat/Km is decreased by 10(4) for R127A). Therefore, Arg-127 stabilizes the enzyme-transition state complex but not the ground state enzyme-substrate complex (Phillips, M.A., Fletterick, R., & Rutter, W.J., 1990, J. Biol. Chem. 265, 20692-20698). The addition of guandine, methylguanidine, or ethylguanidine to R127A increases the kcat for hydrolysis of Bz-gly(o)phe by 10(2) without changing the Km. Dissociation constants (Kd) for the guanidine derivatives range from 0.1 to 0.5 M. The binding affinity for the transition state analog Cbz-phe-alaP(o)ala is increased similarly by 10(2); in contrast, the binding affinity of the ground state inhibitor benzylsuccinic acid is not altered. Thus, guanidine derivatives mimic Arg-127 in stabilizing the rate-limiting transition state. Hydrolysis of Bz-gly-(o)phe by wild-type carboxypeptidase, R127K, or R127M is not substantially affected by guanidine derivatives. Additionally, primary amines do not change the activity of R127A. These observations imply that guanidine binds in the cavity vacated by Arg-127 specifically and in a productive conformation for catalysis.  相似文献   

9.
Bailey J  Powell L  Sinanan L  Neal J  Li M  Smith T  Bell E 《The FEBS journal》2011,278(17):3140-3151
Bovine glutamate dehydrogenase is potently inhibited by zinc and the major impact is on V(max) suggesting a V-type effect on catalysis or product release. Zinc inhibition decreases as glutamate concentrations decrease suggesting a role for subunit interactions. With the monocarboxylic amino acid norvaline, which gives no evidence of subunit interactions, zinc does not inhibit. Zinc significantly decreases the size of the pre-steady state burst in the reaction but does not affect NADPH binding in the enzyme-NADPH-glutamate complex that governs the steady state turnover, again suggesting that zinc disrupts subunit interactions required for catalytic competence. While differential scanning calorimetry suggests zinc binds and induces a slightly conformationally more rigid state of the protein, limited proteolysis indicates that regions in the vicinity of the antennae regions and the trimer-trimer interface become more flexible. The structures of glutamate dehydrogenase bound with zinc and europium show that zinc binds between the three dimers of subunits in the hexamer, a region shown to bind novel inhibitors that block catalytic turnover, which is consistent with the above findings. In contrast, europium binds to the base of the antenna region and appears to abrogate the inhibitory effect of zinc. Structures of various states of the enzyme have shown that both regions are heavily involved in the conformational changes associated with catalytic turnover. These results suggest that the V-type inhibition produced with glutamate as the substrate results from disruption of subunit interactions necessary for efficient catalysis rather than by a direct effect on the active site conformation.  相似文献   

10.
Luba J  Charrier V  Claiborne A 《Biochemistry》1999,38(9):2725-2737
An unusual flavoprotein disulfide reductase, which catalyzes the NADPH-dependent reduction of CoASSCoA, has recently been purified from the human pathogen Staphylococcus aureus [delCardayré, S. B., Stock, K. P., Newton, G. L., Fahey, R. C., and Davies, J. E. (1998) J. Biol. Chem. 273, 5744-5751]. Coenzyme A-disulfide reductase (CoADR) lacks the redox-active protein disulfide characteristic of the disulfide reductases; instead, NADPH reduction yields 1 protein-SH and 1 CoASH. Furthermore, the CoADR sequence reveals the presence of a single putative active-site Cys (Cys43) within an SFXXC motif also seen in the Enterococcus faecalis NADH oxidase and NADH peroxidase, which use a single redox-active cysteine-sulfenic acid in catalysis. In this report, we provide a detailed examination of the equilibrium properties of both wild-type and C43S CoADRs, focusing on the role of Cys43 in the catalytic redox cycle, the behavior of both enzyme forms on reduction with dithionite and NADPH, and the interaction of NADP+ with the corresponding reduced enzyme species. The results of these analyses, combined with electrospray mass spectrometric data for the two oxidized enzyme forms, fully support the catalytic redox role proposed for Cys43 and confirm that this is the attachment site for bound CoASH. In addition, we provide evidence indicating dramatic thermodynamic inequivalence between the two active sites per dimer, similar to that documented for the related enzymes mercuric reductase and NADH oxidase; only 1 FAD is reduced with NADPH in wild-type CoADR. The EH2.NADPH/EH4.NADP+ complex which results is reoxidized quantitatively in titrations with CoASSCoA, supporting a possible role for the asymmetric reduced dimer in catalysis.  相似文献   

11.
Experimental probes of the acidity of weak carbon acids have been developed and used to determine the carbon acid pK(a)s of glycine, glycine derivatives and iminium ion adducts of glycine to the carbonyl group, including 5'-deoxypyridoxal (DPL). The high reactivity of the DPL-stabilized glycyl carbanion towards nucleophilic addition to both DPL and the glycine-DPL iminium ion favors the formation of Claisen condensation products at enzyme active sites. The formation of the iminium ion between glycine and DPL is accompanied by a 12-unit decrease in the pK(a) of 29 for glycine. The complicated effects of formation of glycine iminium ions to DPL and other aromatic and aliphatic aldehydes and ketones on carbon acid pK(a) are discussed. These data provide insight into the contribution of the individual pyridine ring substituents to the catalytic efficiency of DPL. It is suggested that the 5'-phosphodianion group of PLP may play an important role in enzymatic catalysis of carbon deprotonation by providing up to 12 kcal/mol of binding energy that is utilized to stabilize the transition state for the enzymatic reaction. This article is part of a Special Issue entitled: Pyridoxal Phospate Enzymology.  相似文献   

12.
Huang HH  Arscott LD  Ballou DP  Williams CH 《Biochemistry》2008,47(48):12769-12776
Thioredoxin reductase (TrxR) catalyzes the reduction of thioredoxin (Trx) by NADPH. Because dipteran insects such as Drosophila melanogaster lack glutathione reductase, their TrxRs are particularly important for antioxidant protection; reduced Trx reacts nonenzymatically with oxidized glutathione to maintain a high glutathione/glutathione disulfide ratio. Like other members of the pyridine nucleotide-disulfide oxidoreductase family, TrxR is a homodimer; in the enzyme from D. melanogaster (DmTrxR), each catalytically active unit consists of three redox centers: FAD and an N-terminal Cys-57-Cys-62 redox-active disulfide from one monomer and a Cys-489'-Cys-490' C-terminal redox-active disulfide from the second monomer. A dyad of His-464' and Glu-469' in TrxR acts as the acid-base catalyst of the dithiol-disulfide interchange reactions required in catalysis [Huang, H.-H., et al. (2008) Biochemistry 47, 1721-1731]. In this investigation, the role of Glu-469' in catalysis by DmTrxR has been studied. The E469'A and E469'Q DmTrxR variants retain 28 and 35% of the wild-type activity, respectively, indicating that this glutamate residue is important but not critical to catalysis. The pH dependence of V(max) for both glutamate variants yields pK(a) values of 6.0 and 8.7, compared to those in the wild-type enzyme of 6.4 and 9.3, respectively, indicating that the basicity of His-464' in TrxR in complex with its substrate, DmTrx-2, is significantly lower in the glutamate variants than in wild-type enzyme. The rates of some steps in the reductive half-reactions in both glutamate variants are much slower than those of the wild-type enzyme. On the basis of our observations, it is proposed that the function of Glu-469' is to facilitate the positioning of His-464' toward the interchange thiol, Cys-57, as suggested for the analogous residue in glutathione reductase.  相似文献   

13.
A stable covalent complex was prepared by cross-linking adrenodoxin reductase with adrenodoxin using 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide. The covalent complex was purified extensively until free components were removed completely. The major component of the complex had a molecular weight of 63 kDa, which corresponds to a 1:1 stoichiometric complex between adrenodoxin reductase and adrenodoxin. NADPH-cytochrome c reduction activity of the covalent complex was comparable to that of an equimolar mixture of adrenodoxin reductase and adrenodoxin (native complex), and the NADPH-ferricyanide reduction activity of the complex was equal to that of the native one. In contrast to the native complex, the covalent complex produced much less superoxide upon NADPH-oxidation, and the covalent complex was found to be more stable than the native complex, suggesting that the complex state is more favorable for catalysis. From these results, we conclude that the adrenodoxin molecule does not need to dissociate from the complex during electron transfer from NADPH to cytochrome c.  相似文献   

14.
R67 dihydrofolate reductase (DHFR) is a novel protein that possesses 222 symmetry. A single active site pore traverses the length of the homotetramer. Although the 222 symmetry implies that four symmetry-related binding sites should exist for each substrate as well as each cofactor, isothermal titration calorimetry (ITC) studies indicate only two molecules bind. Three possible combinations include two dihydrofolate molecules, two NADPH molecules, or one substrate with one cofactor. The latter is the productive ternary complex. To evaluate the roles of A36, Y46, T51, G64, and V66 residues in binding and catalysis, a site-directed mutagenesis approach was employed. One mutation per gene produces four mutations per active site pore, which often result in large cumulative effects. Conservative mutations at these positions either eliminate the ability of the gene to confer trimethoprim resistance or have no effect on catalysis. This result, in conjunction with previous mutagenesis studies on K32, K33, S65, Q67, I68, and Y69 [Strader, M. B., et al. (2001) Biochemistry 40, 11344-11352; Hicks, S. N., et al. (2003) Biochemistry 42, 10569-10578; Park, H., et al. (1997) Protein Eng. 10, 1415-1424], allows mapping of the active site surface. Residues for which conservative mutations have large effects on binding and catalysis include K32, Q67, I68, and Y69. These residues form a stripe that establishes the ligand binding surface. Residues that accommodate conservative mutations that do not greatly affect catalysis include K33, Y46, T51, S65, and V66. Isothermal titration calorimetry studies were also conducted on many of the mutants described above to determine the enthalpy of folate binding to the R67 DHFR.NADPH complex. A linear correlation between this DeltaH value and log k(cat)/K(m) is observed. Since structural tightness appears to be correlated with the exothermicity of the binding interaction, this leads to the hypothesis that enthalpy-driven formation of the ternary complex in these R67 DHFR variants plays a strong role in catalysis. Use of the alternate cofactor, NADH, extends this correlation, indicating preorganization of the ternary complex determines the efficiency of the reaction. This hypothesis is consistent with data suggesting R67 DHFR uses an endo transition state (where the nicotinamide ring of cofactor overlaps the more bulky side of the substrate's pteridine ring).  相似文献   

15.
Two mechanisms have been postulated for the formation of bound alpha-iminoglutarate intermediate during the glutamate dehydrogenase-catalyzed reductive amination of alpha-ketoglutarate; one involves the nucleophilic attack of ammonia on a covalently bound Schiff base in the enzyme-NADPH-alpha-ketoglutarate complex, and the other involves the reaction of ammonia with the carbonyl group of alpha-ketoglutarate in the ternary complex. We have measured the rates of carbonyl oxygen exchange in the complex to unambiguously distinguish between these two mechanisms. We find that the loss of label in the carbonyl oxygen-labeled ternary complex is at least 10(5) times slower than the rate of the reductive amination reaction. Therefore, the former mechanism cannot be operative. We also find that (i) the carbonyl oxygen exchange in free alpha-ketoglutarate proceeds without any significant catalysis by its gamma-carboxylate group; (ii) this exchange reaction has energy parameters which are comparable to those observed for the hydration of simple aliphatic ketones; and (iii) the carbonyl oxygen exchange in bound alpha-ketoglutarate is slower than that in the free keto acid over a wide pH range. We conclude that the oxygen exchange in the free and bound alpha-ketoglutarate must occur via a gem-diol intermediate. The observation that the enzyme inhibits the reaction of water with alpha-ketoglutarate while it catalyzes the reaction of ammonia with the same keto acid points to an extraordinary recognition of ammonia by the enzyme. We interpret this observation by assuming that the enzyme-NADPH-alpha-ketoglutarate complex exists in two forms, a predominant form which is produced rapidly upon mixing the components together and an unstable form which is produced in trace amounts from the predominant form via a gem-diol intermediate. These two forms are presumed to differ in the spatial relationship of the carbonyl group to the enzyme functional groups. The carbonyl group in the unstable form is assumed to be surrounded by the same enzyme groups as the iminium ion is in the bound iminoglutarate complex. We ascribe the remarkable catalysis of the ammonia reaction and the inhibition of the water reaction by the enzyme to the opposing interactions of the iminium and carbonyl groups with these surrounding enzyme groups.  相似文献   

16.
GDP-mannose hydrolase (GDPMH) catalyzes the hydrolysis of GDP-alpha-d-sugars by nucleophilic substitution with inversion at the anomeric C1 atom of the sugar, with general base catalysis by H124. Three lines of evidence indicate a mechanism with dissociative character. First, in the 1.3 A X-ray structure of the GDPMH-Mg(2+)-GDP.Tris(+) complex [Gabelli, S. B., et al. (2004) Structure 12, 927-935], the GDP leaving group interacts with five catalytic components: R37, Y103, R52, R65, and the essential Mg(2+). As determined by the effects of site-specific mutants on k(cat), these components contribute factors of 24-, 100-, 309-, 24-, and >/=10(5)-fold, respectively, to catalysis. Both R37 and Y103 bind the beta-phosphate of GDP and are only 5.0 A apart. Accordingly, the R37Q/Y103F double mutant exhibits partially additive effects of the two single mutants on k(cat), indicating cooperativity of R37 and Y103 in promoting catalysis, and antagonistic effects on K(m). Second, the conserved residue, D22, is positioned to accept a hydrogen bond from the C2-OH group of the sugar undergoing substitution at C1, as was shown by modeling an alpha-d-mannosyl group into the sugar binding site. The D22A and D22N mutations decreased k(cat) by factors of 10(2.1) and 10(2.6), respectively, for the hydrolysis of GDP-alpha-d-mannose, and showed smaller effects on K(m), suggesting that the D22 anion stabilizes a cationic oxocarbenium transition state. Third, the fluorinated substrate, GDP-2F-alpha-d-mannose, for which a cationic oxocarbenium transition state would be destabilized by electron withdrawal, exhibited a 16-fold decrease in k(cat) and a smaller, 2.5-fold increase in K(m). The D22A and D22N mutations further decreased the k(cat) with GDP-2F-alpha-d-mannose to values similar to those found with GDP-alpha-d-mannose, and decreased the K(m) of the fluorinated substrate. The choice of histidine as the general base over glutamate, the preferred base in other Nudix enzymes, is not due to the greater basicity of histidine, since the pK(a) of E124 in the active complex (7.7) exceeded that of H124 (6.7), and the H124E mutation showed a 10(2.2)-fold decrease in k(cat) and a 4.0-fold increase in K(m) at pH 9.3. Similarly, the catalytic triad detected in the X-ray structure (H124- - -Y127- - -P120) is unnecessary for orienting H124, since the Y127F mutation had only 2-fold effects on k(cat) and K(m) with either H124 or E124 as the general base. Hence, a neutral histidine rather than an anionic glutamate may be necessary to preserve electroneutrality in the active complex.  相似文献   

17.
Xu D  Enroth C  Lindqvist Y  Ballou DP  Massey V 《Biochemistry》2002,41(46):13627-13636
An active site residue in phenol hydroxylase (PHHY), Pro364, was mutated to serine to investigate its role in enzymatic catalysis. In the presence of phenol, the reaction between the reduced flavin of P364S and oxygen is very fast, but only 13% of the flavin is utilized to hydroxylate the substrate, compared to nearly 100% for the wild-type enzyme. The oxidative half-reaction of PHHY using m-cresol as a substrate is similarly affected by the mutation. Pro364 was suggested to be important in stabilizing the transition state of the oxygen transfer step by forming a hydrogen bond between its carbonyl oxygen and the C4a-hydroperoxyflavin [Ridder, L., Mullholland, A. J., Rietjens, I. M. C. M., and Vervoort, J. (2000) J. Am. Chem. Soc. 122, 8728-8738]. The P364S mutation may weaken this interaction by increasing the flexibility of the peptide chain; hence, the transition state would be destabilized to result in a decreased level of hydroxylation of phenol. However, when the oxidative half-reaction was studied using resorcinol as a substrate, the P364S mutant form was not significantly different from the wild-type enzyme. The rate constants for all the reaction steps as well as the hydroxylation efficiency (coupling between NADPH oxidation and resorcinol consumption) are comparable to those of the wild-type enzyme. It is suggested that the function of Pro364 in catalysis, stabilization of the transition state, is not as important in the reaction with resorcinol, possibly because the position of hydroxylation is different with resorcinol than with phenol and m-cresol.  相似文献   

18.
A ferrous heme-NO complex builds up in rat neuronal NO synthase during catalysis and lowers its activity. Mutation of a tryptophan located directly below the heme (Trp(409)) to Phe or Tyr causes hyperactive NO synthesis and less heme-NO complex buildup in the steady state (Adak, S., Crooks, C., Wang, Q., Crane, B. R., Tainer, J. A., Getzoff, E. D., and Stuehr, D. J. (1999) J. Biol. Chem. 274, 26907-26911). To understand the mechanism, we used conventional and stopped flow spectroscopy to compare kinetics of heme-NO complex formation, enzyme activity prior to and after complex formation, NO binding affinity, NO complex stability, and its reaction with O(2) in mutants and wild type nNOS. During the initial phase of NO synthesis, heme-NO complex formation was 3 and 5 times slower in W409F and W409Y, and their rates of NADPH oxidation were 50 and 30% that of wild type, probably due to slower heme reduction. NO complex formation slowed NADPH oxidation in the wild type by 7-fold but reduced mutant activities less than 2-fold, giving mutants higher final activities. NO binding kinetics were similar among mutants and wild type, although in ferrous W409Y (and to a lesser extent W409F) the 436-nm NO complex converted to a 417-nm NO complex with time. Oxidation of the ferrous heme-NO complex to ferric enzyme was 7 times faster in Trp(409) mutants than in wild type. Thus, mutant hyperactivity derives from slower formation and faster decay of the heme-NO complex. Together, these minimize partitioning into the NO-bound form.  相似文献   

19.
Stabilization of an oxyanion transition state is important to catalysis of peptide bond hydrolysis in all proteases. For subtilisin BPN', a bacterial serine protease, structural data suggest that two hydrogen bonds stabilize the tetrahedral-like oxyanion intermediate: one from the main chain NH of Ser221 and another from the side chain NH2 of Asn155. Molecular dynamic studies (Rao, S., N., Singh, U., C. Bush, P. A., and Kollman, P. A. (1987) Nature 328, 551-554) have indicated the gamma-hydroxyl of Thr220 may be a third hydrogen bond donor even though it is 4A away in the static x-ray structure. We have probed the role of Thr220 by replacing it with serine, cysteine, valine, or alanine by site-directed mutagenesis. These substitutions were intended to alter the size and hydrogen bonding ability of residue 220. Removal of the gamma-hydroxyl group reduced the transition state stabilization energy (delta delta GT) by 1.8-2.1 kcal/mol depending upon the substitution. By comparison, removal of the gamma-methyl group in the Thr220 to serine mutation only decreased delta GT by 0.5 kcal/mol. The gamma-hydroxyl of Thr220 is most important for catalysis, not substrate binding, because virtually all of the effects were on kcat and not KM. The role of the Thr220 hydroxyl is functionally independent from the amide NH2 of Asn155 because the free energy effects of double alanine mutants at these two positions are additive. These data indicate that a distal hydrogen bond donor, namely the hydroxyl of Thr220, plays a functionally important role in stabilizing the oxyanion transition state in subtilisin which is independent of Asn155.  相似文献   

20.
The thermodynamic-compensation law observed for the deacylation of a series of acyl-alpha-chymotrypsins has been re-examined. From consideration of the effect of small changes in delta delta H+(+)delta T along a homologous series, it is suggested that the high Tc value of 420 K observed for the process has its origin in the solvation of the acyl-group-catalyst transition state.  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