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.
Non-phosphorylating glyceraldehyde 3-phosphate dehydrogenase from Streptococcus mutans (GAPN) belongs to the aldehyde dehydrogenase (ALDH) family, which catalyzes the irreversible oxidation of a wide variety of aldehydes into acidic compounds via a two-step mechanism: first, the acylation step involves the formation of a covalent ternary complex ALDH-cofactor-substrate, followed by the oxidoreduction process which yields a thioacyl intermediate and reduced cofactor and second, the rate-limiting deacylation step. Structural and molecular factors involved in the chemical mechanism of GAPN have recently been examined. Specifically, evidence was put forward for the chemical activation of catalytic Cys-302 upon cofactor binding to the enzyme, through a local conformational rearrangement involving the cofactor and Glu-268. In addition, the invariant residue Glu-268 was shown to play an essential role in the activation of the water molecule in the deacylation step. For E268A/Q mutant GAPNs, nucleophilic compounds like hydrazine and hydroxylamine were shown to bind and act as substrates in this step. Further studies were focused at understanding the factors responsible for the stabilization and chemical activation of the covalent intermediates, using X-ray crystallography, site-directed mutagenesis, kinetic and physico-chemical approaches. The results support the involvement of an oxyanion site including the side-chain of Asn-169. Finally, given the strict substrate-specificity of GAPN compared to other ALDHs with wide substrate specificity, one has also initiated the characterization of the G3P binding properties of GAPN. These results will be presented and discussed from the point of view of the evolution of the catalytic mechanisms of ALDH.  相似文献   

2.
Non-phosphorylating glyceraldehyde 3-phosphate dehydrogenase from Streptococcus mutans (GAPN) belongs to the aldehyde dehydrogenase (ALDH) family, which catalyzes the irreversible oxidation of a wide variety of aldehydes into acidic compounds via a two-step mechanism: first, the acylation step involves the formation of a covalent ternary complex ALDH-cofactor-substrate, followed by the oxidoreduction process which yields a thioacyl intermediate and reduced cofactor and second, the rate-limiting deacylation step. Structural and molecular factors involved in the chemical mechanism of GAPN have recently been examined. Specifically, evidence was put forward for the chemical activation of catalytic Cys-302 upon cofactor binding to the enzyme, through a local conformational rearrangement involving the cofactor and Glu-268. In addition, the invariant residue Glu-268 was shown to play an essential role in the activation of the water molecule in the deacylation step. For E268A/Q mutant GAPNs, nucleophilic compounds like hydrazine and hydroxylamine were shown to bind and act as substrates in this step. Further studies were focused at understanding the factors responsible for the stabilization and chemical activation of the covalent intermediates, using X-ray crystallography, site-directed mutagenesis, kinetic and physico-chemical approaches. The results support the involvement of an oxyanion site including the side-chain of Asn-169. Finally, given the strict substrate-specificity of GAPN compared to other ALDHs with wide substrate specificity, one has also initiated the characterization of the G3P binding properties of GAPN. These results will be presented and discussed from the point of view of the evolution of the catalytic mechanisms of ALDH.  相似文献   

3.
One of the most striking features of several X-ray structures of CoA-independent ALDHs (aldehyde dehydrogenases) in complex with NAD(P) is the conformational flexibility of the NMN moiety. However, the fact that the rate of the acylation step is high in GAPN (non-phosphorylating glyceraldehyde-3-phosphate dehydrogenase) from Streptococcus mutans implies an optimal positioning of the nicotinamide ring relative to the hemithioacetal intermediate within the ternary GAPN complex to allow an efficient and stereospecific hydride transfer. Substitutions of serine for invariant Thr244 and alanine for Lys178 result in a drastic decrease of the efficiency of hydride transfer which becomes rate-limiting. The crystal structure of the binary complex T244S GAPN-NADP shows that the absence of the beta-methyl group leads to a well-defined conformation of the NMN part, including the nicotinamide ring, clearly different from that depicted to be suitable for an efficient hydride transfer in the wild-type. The approximately 0.6-unit increase in pK(app) of the catalytic Cys302 observed in the ternary complex for both mutated GAPNs is likely to be due to a slight difference in positioning of the nicotinamide ring relative to Cys302 with respect to the wild-type ternary complex. Taken together, the data support a critical role of the Thr244 beta-methyl group, held in position through a hydrogen-bond interaction between the Thr244 beta-hydroxy group and the epsilon-amino group of Lys178, in permitting the nicotinamide ring to adopt a conformation suitable for an efficient hydride transfer during the acylation step for all the members of the CoA-independent ALDH family.  相似文献   

4.
Nonphosphorylating nicotinamide adenine dinucleotide (phosphate)- [NAD(P)-] dependent aldehyde dehydrogenases share a number of conserved amino acid residues, several of which are directly implicated in catalysis. In the present study, the role of Glu-268 from nonphosphorylating glyceraldehyde-3-phosphate dehydrogenase (GAPN) from Streptococcus mutans was investigated. Its substitution by Ala resulted in a k(cat) decrease by 3 orders of magnitude. Pre-steady-state analysis showed that, for both the wild-type and E268A GAPNs, the rate-limiting step of the reaction is associated with deacylation. The pH dependence of the rate of acylation of wild-type GAPN is characterized by the contributions of distinct enzyme protonic species with two pK(a)s of 6.2 and 7.5. Substitution of Glu-268 by Ala resulted in a monosigmoidal pH dependence of the rate constant of acylation with a pK(a) of 6.2, which suggested the assignment of pK(a) 7.5 to Glu-268. Moreover, the E268A substitution did not significantly affect the efficiency of acylation of GAPN, showing that Glu-268 is not critically involved in the acylation, which includes Cys-302 nucleophilic activation and hydride transfer. On the contrary, the drastic decrease of the steady-state rate constant for the E268A GAPN demonstrated the essential role of Glu-268 in the deacylation. At basic pH, the solvent isotope effect of 2.3, characterized by a unique pK(a) of 7.7, and the linearity of the proton inventory showed that the rate-limiting process for deacylation is associated with the hydrolysis step and suggested that the glutamate form of Glu-268 acts as a base catalyst in this process. Surprisingly, the double-sigmoidal form of the pH-steady-state rate constant profile, characterized by pK(a) values of 6.1 and 7.4, revealed the high efficiency of the deacylation even at pH lower than 7.4. Therefore, we propose that the major role of Glu-268 is to promote deacylation through activation and orientation of the attacking water molecule, and in addition to act as a base catalyst at basic pH. From these results in relation to those recently described [Marchal, S., and Branlant, G. (1999) Biochemistry 38, 12950-12958], a scenario for the chemical catalysis of GAPN is proposed.  相似文献   

5.
In order to address the molecular basis of the specificity of aldehyde dehydrogenase for aldehyde substrates, enzymatic characterization of the glyceraldehyde 3-phosphate (G3P) binding site of non-phosphorylating glyceraldehyde-3-phosphate dehydrogenase (GAPN) from Streptococcus mutans has been undertaken. In this work, residues Arg-124, Tyr-170, Arg-301, and Arg-459 were changed by site-directed mutagenesis and the catalytic properties of GAPN mutants investigated. Changing Tyr-170 into phenylalanine induces no major effect on k(cat) and K(m) for d-G3P in both acylation and deacylation steps. Substitutions of Arg-124 and Arg-301 by leucine and Arg-459 by isoleucine led to distinct effects on K(m), on k(cat), or on both. The rate-limiting step of the R124L GAPN remains deacylation. Pre-steady-state analysis and substrate isotope measurements show that hydride transfer remains rate-determining in acylation. Only the apparent affinity for d-G3P is decreased in both acylation and deacylation steps. Substitution of Arg-459 by isoleucine leads to a drastic effect on the catalytic efficiency by a factor of 10(5). With this R459L GAPN, the rate-limiting step is prior to hydride transfer, and the K(m) of d-G3P is increased by at least 2 orders of magnitude. Binding of NADP leads to a time-dependent formation of a charge transfer transition at 333 nm between the pyridinium ring of NADP and the thiolate of Cys-302, which is not observed with the holo-wild type. Accessibility of Cys-302 is shown to be strongly decreased within the holostructure. The substitution of Arg-301 by leucine leads to an even more drastic effect with a change of the rate-limiting step similar to that observed for R459I GAPN. Taking into account the three-dimensional structure of GAPN from S. mutans and the data of the present study, it is proposed that 1) Tyr-170 is not essential for the catalytic event, 2) Arg-124 is only involved in stabilizing d-G3P binding via an interaction with the C-3 phosphate, and 3) Arg-301 and Arg-459 participate not only in d-G3P binding via interaction with C-3 phosphate but also in positioning efficiently d-G3P relative to Cys-302 within the ternary complex GAPN.NADP.d-G3P.  相似文献   

6.
Nonphosphorylating nicotinamide adenine dinucleotide (phosphate)-dependent aldehyde dehydrogenases (ALDHs) catalyze the oxidation of aldehydes into either nonactivated acids or CoA-activated acids. The NADP-dependent nonphosphorylating glyceraldehyde 3-phosphate dehydrogenase (GAPN) belongs to the first subclass. It catalyzes the irreversible oxidation of glyceraldehyde 3-phosphate into 3-phosphoglycerate via a two step mechanism in which deacylation is rate-limiting. Recent studies on GAPN from Streptococcus mutans have shown that residue Glu268 plays an essential role only in the deacylation step [Marchal, S., Rahuel-Clermont, S. & Branlant, G. (2000) Biochemistry 39, 3327-3335]. The substitution of Glu268 by alanine or glutamine leads to mutants in which the attacking water molecule involved in the hydrolytic process is poorly activated. Activity can be restored by the presence of hydroxylamine and hydrazine. Neutral and protonated forms of both nucleophiles are recognized by the deacylating subsite of both mutants. pH rate profiles of deacylation show pK(a) values of 6.3 and 8.1 with hydroxylamine and hydrazine, respectively, which are those of the nucleophiles in solution. The increase in enzymatic rate is probably due to a high local concentration and not to a change of the chemical reactivity of both nucleophiles upon their binding within the active site of both mutants. The deacylation subsite of the wild-type also binds hydroxylamine and hydrazine but as inhibitors of the hydrolytic process and not as acyl acceptors. Altogether, the results point out the crucial role of the carboxyl group of Glu268 in preventing nucleophiles, other than water, from binding as efficient acyl acceptors. This may also explain why CoA-dependent ALDHs never possesses a glutamate residue at position 268.  相似文献   

7.
The NADP-dependent non-phosphorylating glyceraldehyde-3-phosphate dehydrogenase from Streptococcus mutans (abbreviated Sm-ALDH) belongs to the aldehyde dehydrogenase (ALDH) family. Its catalytic mechanism proceeds via two steps, acylation and deacylation. Its high catalytic efficiency at neutral pH implies prerequisites relative to the chemical mechanism. First, the catalytic Cys284 should be accessible and in a thiolate form at physiological pH to attack efficiently the aldehydic group of the glyceraldehyde-3-phosphate (G3P). Second, the hydride transfer from the hemithioacetal intermediate toward the nicotinamide ring of NADP should be efficient. Third, the nucleophilic character of the water molecule involved in the deacylation should be strongly increased. Moreover, the different complexes formed during the catalytic process should be stabilised.The crystal structures presented here (an apoenzyme named Apo2 with two sulphate ions bound to the catalytic site, the C284S mutant holoenzyme and the ternary complex composed of the C284S holoenzyme and G3P) together with biochemical results and previously published apo and holo crystal structures (named Apo1 and Holo1, respectively) contribute to the understanding of the ALDH catalytic mechanism.Comparison of Apo1 and Holo1 crystal structures shows a Cys284 side-chain rotation of 110 degrees, upon cofactor binding, which is probably responsible for its pK(a) decrease. In the Apo2 structure, an oxygen atom of a sulphate anion interacts by hydrogen bonds with the NH2 group of a conserved asparagine residue (Asn154 in Sm-ALDH) and the Cys284 NH group. In the ternary complex, the oxygen atom of the aldehydic carbonyl group of the substrate interacts with the Ser284 NH group and the Asn154 NH2 group. A substrate isotope effect on acylation is observed for both the wild-type and the N154A and N154T mutants. The rate of the acylation step strongly decreases for the mutants and becomes limiting. All these results suggest the involvement of Asn154 in an oxyanion hole in order to stabilise the tetrahedral intermediate and likely the other intermediates of the reaction. In the ternary complex, the cofactor conformation is shifted in comparison with its conformation in the C284S holoenzyme structure, likely resulting from its peculiar binding mode to the Rossmann fold (i.e. non-perpendicular to the plane of the beta-sheet). This change is likely favoured by a characteristic loop of the Rossmann fold, longer in ALDHs than in other dehydrogenases, whose orientation could be constrained by a conserved proline residue. In the ternary and C284S holenzyme structures, as well as in the Apo2 structure, the Glu250 side-chain is situated less than 4 A from Cys284 or Ser284 instead of 7 A in the crystal structure of the wild-type holoenzyme. It is now positioned in a hydrophobic environment. This supports the pK(a) assignment of 7.6 to Glu250 as recently proposed from enzymatic studies.  相似文献   

8.
10-Formyltetrahydrofolate dehydrogenase (FDH) catalyzes an NADP+-dependent dehydrogenase reaction resulting in conversion of 10-formyltetrahydrofolate to tetrahydrofolate and CO2. This reaction is a result of the concerted action of two catalytic domains of FDH, the amino-terminal hydrolase domain and the carboxyl-terminal aldehyde dehydrogenase domain. In addition to participation in the overall FDH mechanism, the C-terminal domain is capable of NADP+-dependent oxidation of short chain aldehydes to their corresponding acids. We have determined the crystal structure of the C-terminal domain of FDH and its complexes with oxidized and reduced forms of NADP. Compared to other members of the ALDH family, FDH demonstrates a new mode of binding of the 2'-phosphate group of NADP via a water-mediated contact with Gln600 that may contribute to the specificity of the enzyme for NADP over NAD. The structures also suggest how Glu673 can act as a general base in both acylation and deacylation steps of the reaction. In the apo structure, the general base Glu673 is positioned optimally for proton abstraction from the sulfur atom of Cys707. Upon binding of NADP+, the side chain of Glu673 is displaced from the active site by the nicotinamide ring and contacts a chain of highly ordered water molecules that may represent a pathway for translocation of the abstracted proton from Glu673 to the solvent. When reduced, the nicotinamide ring of NADP is displaced from the active site, restoring the contact between Cys707 and Glu673 and allowing the latter to activate the hydrolytic water molecule in deacylation.  相似文献   

9.
Comparisons have been made between the active center geometries of lactate dehydrogenase and glyceraldehyde-3-phosphate dehydrogenase, chymotrypsin and papain, and glyceraldehyde-3-phosphate dehydrogenase and papain. In the dehydrogenases, orientation of the nicotinamide ring about the glycosidic bond is determined by the substrate stereochemistry. The proper positioning of the carboxyamide moiety allows for the close approach of the C4 atom on the nicotinamide and the reactive carbon of the substrate. It follows that, once the conformation of the substrate or substrate intermediate has been established with respect to the functional groups in the enzyme, the A- or B-side specificity of the nicotinamide ring is predetermined. Hence, dehydrogenases which are divergently evolving from a common precursor must maintain the nicotinamide specificity if the protein fold of the catalytic domain is conserved. The tetrahedral intermediates produced during acylation of chymotrypsin and papain are found to be of opposite hand, while those of papain and glyceraldehyde-3-phosphate dehydrogenase can be regarded to be of the same hand. Thus the serine proteases, subtilisin and those of the chymotrypsin family, are of one hand while the cysteine enzymes, glyceraldehyde-3-phosphate dehydrogenase and papain, are of the other.  相似文献   

10.
The overall chemical mechanism of the reaction catalyzed by the hydrolytic aldehyde dehydrogenases (ALDHs) involves three main steps: (1) nucleophilic attack of the thiol group of the catalytic cysteine on the carbonyl carbon of the aldehyde substrate; (2) hydride transfer from the tetrahedral thiohemiacetal intermediate to the pyridine ring of NAD(P)(+); and (3) hydrolysis of the resulting thioester intermediate (deacylation). Crystal structures of different ALDHs from several organisms-determined in the absence and presence of bound NAD(P)(+), NAD(P)H, aldehydes, or acid products-showed specific details at the atomic level about the catalytic residues involved in each of the catalytic steps. These structures also showed the conformational flexibility of the nicotinamide half of the cofactor, and of the catalytic cysteinyl and glutamyl residues, the latter being the general base that activates the hydrolytic water molecule in the deacylation step. The architecture of the ALDH active site allows for this conformational flexibility, which, undoubtedly, is crucial for catalysis in these enzymes. Focusing in the deacylation step of the ALDH-catalyzed reaction, here we review and systematize the crystallographic evidence of the structural features responsible for the conformational flexibility of the catalytic glutamyl residue, and for the positioning of the hydrolytic water molecule inside the ALDH active site. Based on the analysis of the available crystallographic data and of energy-minimized models of the thioester reaction intermediate, as well as on the results of theoretical calculations of the pK(a) of the carboxyl group of the catalytic glutamic acid in its three different conformations, we discuss the role that the conformational flexibility of this residue plays in the activation of the hydrolytic water. We also propose a critical participation in the water activation process of the peptide bond to which the catalytic glutamic acid in the intermediate conformation is hydrogen bonded.  相似文献   

11.
Catalysis by the NADP-dependent non-phosphorylating glyceraldehyde-3-phosphate dehydrogenase (GAPN) from Streptococcus mutans, a member of the aldehyde dehydrogenase (ALDH) family, relies on a local conformational reorganization of the active site. This rearrangement is promoted by the binding of NADP and is strongly kinetically favored by the formation of the ternary complex enzyme.NADP.substrate. Adiabatic differential scanning calorimetry was used to investigate the effect of ligands on the irreversible thermal denaturation of GAPN. We showed that phosphate binds to GAPN, resulting in the formation of a GAPN.phosphate binary complex characterized by a strongly decreased thermal stability, with a difference of at least 15 degrees C between the maximum temperatures of the thermal transition peaks. The kinetics of phosphate association and dissociation are slow, allowing both free and GAPN.phosphate complexes to be observed by differential scanning calorimetry and to be separated by native polyacrylamide electrophoresis run in phosphate buffer. Analysis of a set of mutants of GAPN strongly suggests that phosphate is bound to the substrate C-3 subsite. In addition, the substrate analog glycerol-3-phosphate has similar effects as does phosphate on the thermal behavior of GAPN. Based on the current knowledge on the catalytic mechanism of GAPN and other ALDHs, we propose that ligand-induced thermal destabilization is a mechanism that provides to ALDHs the required flexibility for an efficient catalysis.  相似文献   

12.
S Marchal  G Branlant 《Biochemistry》1999,38(39):12950-12958
Nonphosphorylating glyceraldehyde 3-phosphate dehydrogenase (GAPN) from Streptococcus mutans which catalyzes the irreversible oxidation of D-glyceraldehyde-3 phosphate (D-G3P) into 3-phosphoglycerate (3-PGA) in the presence of NADP belongs to the aldehyde dehydrogenase (ALDH) superfamily. Oxidation of D-G3P into 3-PGA by GAPN involves the formation of a covalent enzyme intermediate via the nucleophilic attack of the invariant Cys-302. Titration of Cys-302 in the apo-enzyme by two different kinetic probes, iodoacetamide and 2,2'-dipyridyl disulfide, shows a pK(app) of 8.5 and a chemical reactivity surprisingly low compared to a reactive and accessible thiolate. Binding of NADP causes a strong increase of the reactivity of Cys-302-which is time dependent-with a pK(app) shift from 8.5 to 6.1. Concomitant with the increase in the Cys-302 reactivity, an additional protein fluorescence quenching is observed. These data suggest that cofactor binding induces at least a local conformational rearrangement within the active site. The efficiency of the rearrangement depends on the structure of the cofactors and on the protonation of an amino acid with a pK(app)( )()of 5.7. The rate of the rearrangement also strongly increases when temperature decreases. The data on the conformational rearrangement also reveal an amino acid with a pK(app) of 7.6 whose deprotonation increases the reactivity of the thiolate of Cys-302 by a 3-fold factor. The nature of the amino acid involved-which should be located close to Cys-302 in the holo-active form-is likely the invariant Glu-268. Changing Glu-268 into Ala or Cys-302 into Ala leads to mutants in which the rearrangement is only efficient in the presence of saturating concentrations of both NADP and G3P. The structural aspects of the conformational rearrangement occurring during the catalytic process in the wild-type GAPN should include at least reorientation of both Cys-302 and Glu-268 side chains and repositioning of the nicotinamide ring of the cofactor to permit the chemical activation of Cys-302 and the formation of an efficient ternary complex. Thus, it is likely that the conformation of the active site in the reported X-ray structures of ALDHs determined so far in the presence of cofactor, in which the side chains of Cys-302 and Glu-268 are 6.7 A apart from each other, does not represent the biological active form.  相似文献   

13.
Molecular dynamics (MD) simulation of the rat class 3 aldehyde dehydrogenase (ALDH) with nicotinamide dinucleotide (NAD) cofactors and explicit water molecules are reported. Our results demonstrate that MD simulation using the latest methodologies can maintain the crystal structure of the enzyme, as well as closely reproduce the short timescale dynamics of the enzyme. Furthermore, the examination of the distance between the nucleophilic Cys-243 and the NAD cofactor reveal important fluctuations that could be linked to ALDH catalysis. Finally, our quantum mechanical model of benzaldehyde in the active site of ALDH demonstrates that the enzyme requires only minor conformational changes to be poised for nucleophilic attack on the substrate.  相似文献   

14.
Ho KK  Hurley TD  Weiner H 《Biochemistry》2006,45(31):9445-9453
Random mutagenesis followed by a filter-based screening assay has been used to identify a mutant of human class 1 aldehyde dehydrogenase (ALDH1) that was no longer inhibited by Mg(2+) ions but was activated in their presence. Several mutants possessed double, triple, and quadruple amino acid substitutions with a total of seven different residues being altered, but each had a common T244S change. This point mutation proved to be responsible for the Mg(2+) ion activation. An ALDH1 T244S mutant was recombinantly expressed and was used for mechanistic studies. Mg(2+) ions have been shown to increase the rate of deacylation. Consistent with the rate-limiting step for ALDH1 being changed from coenzyme dissociation to deacylation was finding that chloroacetaldehyde was oxidized more rapidly than acetaldehyde. Furthermore, Mg(2+) ions only in the presence of NAD(H) increased the rate of hydrolysis of p-nitrophenyl acetate showing that the metal only affects the binary complex. Though the rate-limiting step for the T244S mutant was different from that of the native enzyme, the catalytic efficiency of the mutant was just 20% that of the native enzyme. The basis for the change in the rate-limiting step appears to be related to NAD(+) binding. Using the structure of a sheep class 1 ALDH, it was possible to deduce that the interaction between the side chain of T244 and its neighboring residues with the nicotinamide ring of NAD(+) were an essential determinant in the catalytic action of ALDH1.  相似文献   

15.
Molecular dynamics (MD) simulation of the rat class 3 aldehyde dehydrogenase (ALDH) with nicotinamide dinucleotide (NAD) cofactors and explicit water molecules are reported. Our results demonstrate that MD simulation using the latest methodologies can maintain the crystal structure of the enzyme, as well as closely reproduce the short timescale dynamics of the enzyme. Furthermore, the examination of the distance between the nucleophilic Cys-243 and the NAD cofactor reveal important fluctuations that could be linked to ALDH catalysis. Finally, our quantum mechanical model of benzaldehyde in the active site of ALDH demonstrates that the enzyme requires only minor conformational changes to be poised for nucleophilic attack on the substrate.  相似文献   

16.
Structural dynamics associated with cofactor binding have been shown to play key roles in the catalytic mechanism of hydrolytic NAD(P)-dependent aldehyde dehydrogenases (ALDH). By contrast, no information is available for their CoA-dependent counterparts. We present here the first crystal structure of a CoA-dependent ALDH. The structure of the methylmalonate semialdehyde dehydrogenase (MSDH) from Bacillus subtilis in binary complex with NAD(+) shows that, in contrast to what is observed for hydrolytic ALDHs, the nicotinamide ring is well defined in the electron density due to direct and H(2)O-mediated hydrogen bonds with the carboxamide. The structure also reveals that a conformational isomerization of the NMNH is possible in MSDH, as shown for hydrolytic ALDHs. Finally, the adenine ring is substantially more solvent-exposed, a result that could be explained by the presence of a Val residue at position 229 in helix α(F) that reduces the depth of the binding pocket and the absence of Gly-225 at the N-terminal end of helix α(F). Substitution of glycine for Val-229 and/or insertion of a glycine residue at position 225 resulted in a significant decrease of the rate constant associated with the dissociation of NADH from the NADH/thioacylenzyme complex, thus demonstrating that the weaker stabilization of the adenine ring is a key factor in triggering the early NADH release in the MSDH-catalyzed reaction. This study provides for the first time structural insights into the mechanism whereby the cofactor binding mode is responsible at least in part for the different kinetic behaviors of the hydrolytic and CoA-dependent ALDHs.  相似文献   

17.
cis-Biphenyl-2,3-dihydrodiol-2,3-dehydrogenase (BphB) is involved in the aerobic biodegradation of polychlorinated biphenyls (PCBs). The crystal structure of the NAD+-enzyme complex was determined by molecular replacement and refined to an R-value of 17.9% at 2.0 A. As a member of the short-chain alcohol dehydrogenase/reductase (SDR) family, the overall protein fold and positioning of the catalytic triad in BphB are very similar to those observed in other SDR enzymes, although small differences occur in the cofactor binding site. Modeling studies indicate that the substrate is bound in a deep hydrophobic cleft close to the nicotinamide moiety of the NAD+ cofactor. These studies further suggest that Asn143 is a key determinant of substrate specificity. A two-step reaction mechanism is proposed for cis-dihydrodiol dehydrogenases.  相似文献   

18.
19.
Quantitative structure activity analysis of the substrate types Ala-Ala-AR and Ala-Pro-AR containing different substituents in the aryl ring showed that the rate-limiting step in the hydrolysis of the alanine substrates by dipeptidyl peptidase IV occurs in th acylation reaction (kcat approximately k2). Probably, the tetrahedral intermediate of the acylation process has a real life time. The positive q-value of the Hammett-equation in k'cat suggests that the N-atom of the arylamide is charged more negatively in the transition state TI not equal to than in the original state TI. The analysis of the quantitative conformation activity relationship (QCAR) gives information on the steric situation in the tetrahedral intermediate of the acylation step near the transition state. The rate limiting step in the hydrolysis of the substrates of the proline type occurs in the deacylation reaction.  相似文献   

20.
Phenylpropenes, a large group of plant volatile compounds that serve in multiple roles in defense and pollinator attraction, contain a propenyl side chain. Eugenol synthase (EGS) catalyzes the reductive displacement of acetate from the propenyl side chain of the substrate coniferyl acetate to produce the allyl-phenylpropene eugenol. We report here the structure determination of EGS from basil (Ocimum basilicum) by protein x-ray crystallography. EGS is structurally related to the short-chain dehydrogenase/reductases (SDRs), and in particular, enzymes in the isoflavone-reductase-like subfamily. The structure of a ternary complex of EGS bound to the cofactor NADP(H) and a mixed competitive inhibitor EMDF ((7S,8S)-ethyl (7,8-methylene)-dihydroferulate) provides a detailed view of the binding interactions within the EGS active site and a starting point for mutagenic examination of the unusual reductive mechanism of EGS. The key interactions between EMDF and the EGS-holoenzyme include stacking of the phenyl ring of EMDF against the cofactor's nicotinamide ring and a water-mediated hydrogen-bonding interaction between the EMDF 4-hydroxy group and the side-chain amino moiety of a conserved lysine residue, Lys132. The C4 carbon of nicotinamide resides immediately adjacent to the site of hydride addition, the C7 carbon of cinnamyl acetate substrates. The inhibitor-bound EGS structure suggests a two-step reaction mechanism involving the formation of a quinone-methide prior to reduction. The formation of this intermediate is promoted by a hydrogen-bonding network that favors deprotonation of the substrate's 4-hydroxyl group and disfavors binding of the acetate moiety, akin to a push-pull catalytic mechanism. Notably, the catalytic involvement in EGS of the conserved Lys132 in preparing the phenolic substrate for quinone methide formation through the proton-relay network appears to be an adaptation of the analogous role in hydrogen bonding played by the equivalent lysine residue in other enzymes of the SDR family.  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