首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The fatty aldehyde dehydrogenase (Vh-ALDH) isolated from the luminescent bacterium, Vibrio harveyi, differs from other aldehyde dehydrogenases in its high affinity for NADP(+). The binding of NADP(+) appears to arise from the interaction of the 2'-phosphate of the adenosine moiety of NADP(+) with a threonine (T175) in the nucleotide recognition site just after the beta(B) strand as well as with an arginine (R210) that pi stacks over the adenosine moiety. The active site of Vh-ALDH contains the usual suspects of a cysteine (C289), two glutamates (E253 and E377) and an asparagine (N147) involved in the aldehyde dehydrogenase mechanism. However, Vh-ALDH has one polar residue in the active site that distinguishes it from other ALDHs; a histidine (H450) is in close contact with the cysteine nucleophile. As a glutamate has been implicated in promoting the nucleophilicity of the active site cysteine residue in ALDHs, the close contact of a histidine with the cysteine nucleophile in Vh-ALDH raises the possibility of alternate routes to increase the reactivity of the cysteine nucleophile. The effects of mutation of these residues on the different functions catalyzed by Vh-ALDH including acylation, (thio)esterase, reductase and dehydrogenase activities should help define the specific roles of the residues in the active site of ALDHs.  相似文献   

2.
The fatty aldehyde dehydrogenase (Vh-ALDH) isolated from the luminescent bacterium, Vibrio harveyi, differs from other aldehyde dehydrogenases in its high affinity for NADP+. The binding of NADP+ appears to arise from the interaction of the 2′-phosphate of the adenosine moiety of NADP+ with a threonine (T175) in the nucleotide recognition site just after the βB strand as well as with an arginine (R210) that pi stacks over the adenosine moiety. The active site of Vh-ALDH contains the usual suspects of a cysteine (C289), two glutamates (E253 and E377) and an asparagine (N147) involved in the aldehyde dehydrogenase mechanism. However, Vh-ALDH has one polar residue in the active site that distinguishes it from other ALDHs; a histidine (H450) is in close contact with the cysteine nucleophile. As a glutamate has been implicated in promoting the nucleophilicity of the active site cysteine residue in ALDHs, the close contact of a histidine with the cysteine nucleophile in Vh-ALDH raises the possibility of alternate routes to increase the reactivity of the cysteine nucleophile. The effects of mutation of these residues on the different functions catalyzed by Vh-ALDH including acylation, (thio)esterase, reductase and dehydrogenase activities should help define the specific roles of the residues in the active site of ALDHs.  相似文献   

3.
Relationships within the aldehyde dehydrogenase extended family   总被引:2,自引:0,他引:2       下载免费PDF全文
One hundred-forty-five full-length aldehyde dehydrogenase-related sequences were aligned to determine relationships within the aldehyde dehydrogenase (ALDH) extended family. The alignment reveals only four invariant residues: two glycines, a phenylalanine involved in NAD binding, and a glutamic acid that coordinates the nicotinamide ribose in certain E-NAD binary complex crystal structures, but which may also serve as a general base for the catalytic reaction. The cysteine that provides the catalytic thiol and its closest neighbor in space, an asparagine residue, are conserved in all ALDHs with demonstrated dehydrogenase activity. Sixteen residues are conserved in at least 95% of the sequences; 12 of these cluster into seven sequence motifs conserved in almost all ALDHs. These motifs cluster around the active site of the enzyme. Phylogenetic analysis of these ALDHs indicates at least 13 ALDH families, most of which have previously been identified but not grouped separately by alignment. ALDHs cluster into two main trunks of the phylogenetic tree. The largest, the "Class 3" trunk, contains mostly substrate-specific ALDH families, as well as the class 3 ALDH family itself. The other trunk, the "Class 1/2" trunk, contains mostly variable substrate ALDH families, including the class 1 and 2 ALDH families. Divergence of the substrate-specific ALDHs occurred earlier than the division between ALDHs with broad substrate specificities. A site on the World Wide Web has also been devoted to this alignment project.  相似文献   

4.
L Zhang  B Ahvazi  R Szittner  A Vrielink  E Meighen 《Biochemistry》1999,38(35):11440-11447
The fatty aldehyde dehydrogenase from the luminescent bacterium, Vibrio harveyi (Vh-ALDH), is unique with respect to its high specificity for NADP(+) over NAD(+). By mutation of a single threonine residue (Thr175) immediately downstream of the beta(B) strand in the Rossmann fold, the nucleotide specificity of Vh-ALDH has been changed from NADP(+) to NAD(+). Replacement of Thr175 by a negatively charged residue (Asp or Glu) resulted in an increase in k(cat)/K(m) for NAD(+) relative to that for NADP(+) of up to 5000-fold due to a decrease for NAD(+) and an increase for NADP(+) in their respective Michaelis constants (K(a)). Differential protection by NAD(+) and NADP(+) against thermal inactivation and comparison of the dissociation constants of NMN, 2'-AMP, 2'5'-ADP, and 5'-AMP for these mutants and the wild-type enzyme clearly support the change in nucleotide specificity. Moreover, replacement of Thr175 with polar residues (N, S, or Q) demonstrated that a more efficient NAD(+)-dependent enzyme T175Q could be created without loss of NADP(+)-dependent activity. Analysis of the three-dimensional structure of Vh-ALDH with bound NADP(+) showed that the hydroxyl group of Thr175 forms a hydrogen bond to the 2'-phosphate of NADP(+). Replacement with glutamic acid or glutamine strengthened interactions with NAD(+) and indicated why threonine would be the preferred polar residue at the nucleotide recognition site in NADP(+)-specific aldehyde dehydrogenases. These results have shown that the size and the structure of the residue at the nucleotide recognition site play the key roles in differentiating between NAD(+) and NADP(+) interactions while the presence of a negative charge is responsible for the decrease in interactions with NADP(+) in Vh-ALDH.  相似文献   

5.
PaBADH (Pseudomonas aeruginosa betaine aldehyde dehydrogenase) catalyses the irreversible NAD(P)+-dependent oxidation of betaine aldehyde to its corresponding acid, the osmoprotector glycine betaine. This reaction is involved in the catabolism of choline and in the response of this important pathogen to the osmotic and oxidative stresses prevalent in infection sites. The crystal structure of PaBADH in complex with NADPH showed a novel covalent adduct between the C2N of the pyridine ring and the sulfur atom of the catalytic cysteine residue, Cys286. This kind of adduct has not been reported previously either for a cysteine residue or for a low-molecular-mass thiol. The Michael addition of the cysteine thiolate in the 'resting' conformation to the double bond of the α,β-unsaturated nicotinamide is facilitated by the particular conformation of NADPH in the active site of PaBADH (also observed in the crystal structure of the Cys286Ala mutant) and by an ordered water molecule hydrogen bonded to the carboxamide group. Reversible formation of NAD(P)H-Cys286 adducts in solution causes reversible enzyme inactivation as well as the loss of Cys286 reactivity towards thiol-specific reagents. This novel covalent modification may provide a physiologically relevant regulatory mechanism of the irreversible PaBADH-catalysed reaction, preventing deleterious decreases in the intracellular NAD(P)+/NAD(P)H ratios.  相似文献   

6.
A nucleophilic group in the active site of aldehyde dehydrogenase, which covalently binds the aldehyde moiety during the enzyme-catalyzed oxidation of aldehydes to acids, was acylated with the chromophoric aldehyde trans-4-(N,N-dimethylamino)cinnamaldehyde (DACA). Acyl-enzyme trapped by precipitation with perchloric acid was digested with trypsin, and the peptide associated with the chromophoric group was isolated and shown to be Gln-Ala-Phe-Gln-Ile-Gly-Ser-Pro-Trp-Arg. After redigestion with thermolysin, the chromophore was associated with the C-terminal hexaresidue part. If the chromophore is attached to this peptide, serine would be expected to bind the aldehyde and lead to the required acylated derivative. Differential labeling experiments were performed in which all free thiol groups on the acylated enzyme were blocked by carboxymethylation. The acyl chromophore was then removed by controlled hydrolysis and the protein reacted with [14C]iodoacetamide. No 14C-labeled tryptic peptides were isolated, suggesting that the sulfur of a cysteine cannot be the acylated residue in the precipitated acyl-enzyme.  相似文献   

7.
Studies of pH-dependent kinetics implicate two ionizable groups in the dehydrogenase and esterase reactions catalysed by high-Km aldehyde dehydrogenase from rat liver mitochondria. Sensitized photooxidation completely arrests the bifunctional activities of the dehydrogenase. Carboxamidomethylation abolishes the dehydrogenase activity, whereas acetimidination eliminates the esterase activity. These results suggest that histidine (pKa near 6) and cysteine (pKa near 10) are likely the catalytic residues for the dehydrogenase activity, while the esterase activity is functionally related to histidine (pKa near 7) and a residue with the pKa value of 10-11. The two residues, a carboxyl group and an arginine, that discriminate between NAD+ and NADP+ are present at the coenzyme binding site of the mitochondrial high-Km aldehyde dehydrogenase from rat liver.  相似文献   

8.
Concentrated urine formation in the kidney is accompanied by conditions that favor the accumulation of reactive oxygen species (ROS). Under hyperosmotic conditions, medulla cells accumulate glycine betaine, which is an osmolyte synthesized by betaine aldehyde dehydrogenase (BADH, EC 1.2.1.8). All BADHs identified to date have a highly reactive cysteine residue at the active site, and this cysteine is susceptible to oxidation by hydrogen peroxide. Porcine kidney BADH incubated with H(2)O(2) (0-500 μM) lost 25% of its activity. However, pkBADH inactivation by hydrogen peroxide was limited, even after 120 min of incubation. The presence of coenzyme NAD(+) (10-50 μM) increased the extent of inactivation (60%) at 120 min of reaction, but the ligands betaine aldehyde (50 and 500 μM) and glycine betaine (100 mM) did not change the rate or extent of inactivation as compared to the reaction without ligand. 2-Mercaptoethanol and dithiothreitol, but not reduced glutathione, were able to restore enzyme activity. Mass spectrometry analysis of hydrogen peroxide inactivated BADH revealed oxidation of M278, M243, M241 and H335 in the absence and oxidation of M94, M327 and M278 in the presence of NAD(+). Molecular modeling of BADH revealed that the oxidized methionine and histidine residues are near the NAD(+) binding site. In the presence of the coenzyme, these oxidized residues are proximal to the betaine aldehyde binding site. None of the oxidized amino acid residues participates directly in catalysis. We suggest that pkBADH inactivation by hydrogen peroxide occurs via disulfide bond formation between vicinal catalytic cysteines (C288 and C289).  相似文献   

9.
The enzyme, 10-formyltetrahydrofolate dehydrogenase (FDH), converts 10-formyltetrahydrofolate (10-formyl-THF) to tetrahydrofolate in an NADP(+)-dependent dehydrogenase reaction or an NADP(+)-independent hydrolase reaction. The hydrolase reaction occurs in a 310-amino acid long amino-terminal domain of FDH (N(t)-FDH), whereas the dehydrogenase reaction requires the full-length enzyme. The amino-terminal domain of FDH shares some sequence identity with several other enzymes utilizing 10-formyl-THF as a substrate. These enzymes have two strictly conserved residues, aspartate and histidine, in the putative catalytic center. We have shown recently that the conserved aspartate is involved in FDH catalysis. In the present work we studied the role of the conserved histidine, His(106), in FDH function. Site-directed mutagenesis experiments showed that replacement of the histidine with alanine, asparagine, aspartate, glutamate, glutamine, or arginine in N(t)-FDH resulted in expression of insoluble proteins. Replacement of the histidine with another positively charged residue, lysine, produced a soluble mutant with no hydrolase activity. The insoluble mutants refolded from inclusion bodies adopted a conformation inherent to the wild-type N(t)-FDH, but they did not exhibit any hydrolase activity. Substitution of alanine for three non-conserved histidines located close to the conserved one did not reveal any significant changes in the hydrolase activity of N(t)-FDH. Expressed full-length FDH with the substitution of lysine for the His(106) completely lost both the hydrolase and dehydrogenase activities. Thus, our study showed that His(106), besides being an important structural residue, is also directly involved in both the hydrolase and dehydrogenase mechanisms of FDH. Modeling of the putative hydrolase catalytic center/folate-binding site suggested that the catalytic residues, aspartate and histidine, are unlikely to be adjacent to the catalytic cysteine in the aldehyde dehydrogenase catalytic center. We hypothesize that 10-formyl-THF dehydrogenase reaction is not an independent reaction but is a combination of hydrolase and aldehyde dehydrogenase reactions.  相似文献   

10.
Lipid peroxidation is one of the consequences of environmental stress in plants and leads to the accumulation of highly toxic, reactive aldehydes. One of the processes to detoxify these aldehydes is their oxidation into carboxylic acids catalyzed by NAD(P)+-dependent ALDHs (aldehyde dehydrogenases). We investigated kinetic parameters of two Arabidopsis thaliana family 3 ALDHs, the cytosolic ALDH3H1 and the chloroplastic isoform ALDH3I1. Both enzymes had similar substrate specificity and oxidized saturated aliphatic aldehydes. Catalytic efficiencies improved with the increase of carbon chain length. Both enzymes were also able to oxidize α,β-unsaturated aldehydes, but not aromatic aldehydes. Activity of ALDH3H1 was NAD+-dependent, whereas ALDH3I1 was able to use NAD+ and NADP+. An unusual isoleucine residue within the coenzyme-binding cleft was responsible for the NAD+-dependence of ALDH3H1. Engineering the coenzyme-binding environment of ALDH3I1 elucidated the influence of the surrounding amino acids. Enzyme activities of both ALDHs were redox-sensitive. Inhibition was correlated with oxidation of both catalytic and non-catalytic cysteine residues in addition to homodimer formation. Dimerization and inactivation could be reversed by reducing agents. Mutant analysis showed that cysteine residues mediating homodimerization are located in the N-terminal region. Modelling of the protein structures revealed that the redox-sensitive cysteine residues are located at the surfaces of the subunits.  相似文献   

11.
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   

12.
NAD+-dependent 15-hydroxyprostaglandin dehydrogenase (15-PGDH), a member of the short-chain dehydrogenase/reductase (SDR) family, catalyzes the first step in the catabolic pathways of prostaglandins and lipoxins. This enzyme oxidizes the C-15 hydroxyl group of prostaglandins and lipoxins to produce 15-keto metabolites which exhibit greatly reduced biological activities. A three-dimensional (3D) structure of 15-PGDH based on the crystal structures of the levodione reductase and tropinone reductase-II was generated and used for docking study with NAD+ coenzyme and PGE2 substrate. Three well-conserved residues among SDR family which correspond to Ser-138, Tyr-151, and Lys-155 of 15-PGDH have been shown to participate in the catalytic reaction. Based on the molecular interactions observed from 3D structure of 15-PGDH, we further propose that Gln-148 in 15-PGDH is important in properly positioning the 15-hydroxyl group of PGE2 by hydrogen bonding with the side-chain oxygen atom of Gln-148. This residue is found to be less conserved and replaceable by glutamyl, histidinyl, and asparaginyl residues in SDR family. Accordingly, site-directed mutagenesis of Gln-148 of 15-PGDH to alanine, glutamic acid, histidine, and asparagine (Q148A, Q148E, Q148H, and Q148N) was carried out. The activity of mutant Q148A was not detectable, whereas those of mutants Q148E, Q148H, and Q148N were comparable to or higher than the wild type. This indicates that the side-chain oxygen or nitrogen atom at position 148 of 15-PGDH plays an important role in anchoring C-15 hydroxyl group of PGE2 through hydrogen bonding for catalytic reaction.  相似文献   

13.
Zheng R  Dam TK  Brewer CF  Blanchard JS 《Biochemistry》2004,43(22):7171-7178
Pantothenate synthetase (EC 6.3.2.1) catalyzes the formation of pantothenate from ATP, D-pantoate, and beta-alanine in bacteria, yeast, and plants. The three-dimensional structural determination of pantothenate synthetase from Mycobacterium tuberculosis has indicated specific roles for His44, His47, Asn69, Gln72, Lys160, and Gln164 residues in the binding of substrates and the pantoyl adenylate intermediate. To evaluate the functional roles of these strictly conserved residues, we constructed six Ala mutants and determined their catalytic properties. The substitution of alanine for H44, H47, N69, Q72, and K160 residues in M. tuberculosis pantothenate synthetase caused a greater than 1000-fold reduction in enzyme activity, while the Q164A mutant exhibited 50-fold less activity. The rate of the isolated adenylation reaction in single turnover studies was also reduced 40-1000-fold by the replacement of one of these six amino acids with alanine, suggesting that these residues are essential for the formation of the pantoyl adenylate intermediate. The rate of pantothenate formation from the adenylate and beta-alanine in the second half reaction could not be measured for the H44A, H47A, N69A, Q72A, and K160A mutants and was reduced 40-fold in the Q164A mutants. The activity of the K160C mutant enzyme was markedly enhanced by the alkylation of cysteine with bromoethylamine, further supporting the critical role of the K160 residue in pantoyl adenylate formation. Isothermal titration microcalorimetry analysis demonstrated that the substitution of either H47 or K160 for Ala resulted in a decreased affinity of the enzyme for ATP. These results indicate that the highly conserved His44, His47, Asn69, Gln72, Lys160 and residues are essential for the formation and stabilization of pantoyl adenylate intermediate in the pantothenate synthetase reaction.  相似文献   

14.
Dong S  Ybe JA  Hecht MH  Spiro TG 《Biochemistry》1999,38(11):3379-3385
Type I Cu proteins maintain a trigonal N2S coordination group (with weak axial ligation) in both oxidation states of the Cu2+/+ ion, thereby reducing the reorganization energy for electron transfer. Requirements for maintaining this coordination group were investigated in poplar plastocyanin (Pcy) by mutation of a conserved element of the type 1 architecture, an asparagine residue (Asn38) adjacent to one of the ligating histidines. The side chain of this asparagine forms an active site clasp via two H-bonds with the residue (Ser85) adjacent to the ligating cysteine (Cys84). In addition, the main chain NH of Asn38 donates an H-bond to the thiolate ligand. We have investigated the importance of these interactions by mutating Asn38 to Gln, Thr, and Leu. The mutant proteins are capable of folding and binding Cu2+, but the blue color fades; the rate of fading increases in the order Gln < Thr < Leu. The color is not restored by ferricyanide, showing that the protein is modified irreversibly, probably by oxidation of Cys84. The more stable mutants N38Q and N38T were characterized spectroscopically. The wild-type properties are slightly perturbed for N38Q, but N38T shows remarkable similarity to another type 1 Cu protein, azurin (Azu) from Pseudomonas aeruginosa. The Cu-S(Cys) bond is longer in Azu than in Pcy, and the NH H-bond to the ligating S atom is shorter. Molecular modeling suggests a similar effect for N38T because the threonine residue shifts toward Ser85 in order to avoid a steric clash and to optimize H-bonding. These results demonstrate that H-bonding adjacent to the type 1 site stabilizes an architecture which both modulates the electronic properties of the Cu, and suppresses side reactions of the cysteine ligand.  相似文献   

15.
[NiFe] hydrogenases contain a highly conserved histidine residue close to the [NiFe] active site which is altered by a glutamine residue in the H(2)-sensing [NiFe] hydrogenases. In this study, we exchanged the respective glutamine residue of the H(2) sensor (RH) of Ralstonia eutropha, Q67 of the RH large subunit HoxC, by histidine, asparagine and glutamate. The replacement by histidine and asparagine resulted in slightly unstable RH proteins which were hardly affected in their regulatory and enzymatic properties. The exchange to glutamate led to a completely unstable RH protein. The purified wild-type RH and the mutant protein with the Gln/His exchange were analysed by continuous-wave and pulsed electron paramagnetic resonance (EPR) techniques. We observed a coupling of a nitrogen nucleus with the [NiFe] active site for the mutant protein which was absent in the spectrum of the wild-type RH. A combination of theoretical calculations with the experimental data provided an explanation for the observed coupling. It is shown that the coupling is due to the formation of a weak hydrogen bond between the protonated N(epsilon) nucleus of the histidine with the sulfur of a conserved cysteine residue which coordinates the metal atoms of the [NiFe] active site as a bridging ligand. The effect of this hydrogen bond on the local structure of the [NiFe] active site is discussed.  相似文献   

16.
NADP(+)-specific glutamate dehydrogenase of Salmonella typhimurium was previously shown to react irreversibly at the coenzyme site with the nucleotide analogue 2-((4-bromo-2,3-dioxobutyl)thio)-1,N6-ethenoadenosine 2',5'-bisphosphate (2-BDB-T epsilon A 2',5'-DP) yielding a partially active enzyme, and inactivation was attributed to modification of the peptide Leu282-Cys-Glu-Ile-Lys286 (Bansal, A., Dayton, M.A., Zalkin, H., and Colman, R.F. (1989) J. Biol. Chem. 264, 9827-9835). Three mutant enzymes have now been engineered, expressed in Escherichia coli, and purified: the single mutants C283I and E284Q and the double mutant C283I:E284Q. The wild-type and mutant enzymes have similar specific activities and Km values for alpha-ketoglutarate, ammonium ion, and NADPH, indicating that neither cysteine 283 nor glutamic acid 284 is essential for activity. The mutant enzyme E284Q, like wild-type glutamate dehydrogenase, is substantially inactivated by 2-BDB-T epsilon A 2',5'-DP. In contrast, the two cysteine mutant enzymes, C283I and C283I:E284Q, are not inactivated by 2-BDB-T epsilon A 2',5'-DP. Modified tryptic peptides with the sequence Leu-X-Glu(Gln)-Ile-Lys were isolated from wild-type or E284Q enzymes inactivated by 2-BDB-T epsilon A 2',5'-DP. This peptide was absent from digests of active wild-type enzyme modified in the presence of the protectant NADPH and from digests of active C283I enzyme after incubation with 2-BDB-T epsilon A 2',5'-DP. Although it is not required for catalytic activity, cysteine 283 is implicated by the results of the affinity labeling experiments as the reaction target of the nucleotide analogue and is located in the region of the coenzyme binding site.  相似文献   

17.
The gene coding for thermophilic xylose (glucose) isomerase of Clostridium thermosulfurogenes was isolated and its complete nucleotide sequence was determined. The structural gene (xylA) for xylose isomerase encodes a polypeptide of 439 amino acids with an estimated molecular weight of 50,474. The deduced amino acid sequence of thermophilic C. thermosulfurogenes xylose isomerase displayed higher homology with those of thermolabile xylose isomerases from Bacillus subtilis (70%) and Escherichia coli (50%) than with those of thermostable xylose isomerases from Ampullariella (22%), Arthrobacter (23%), and Streptomyces violaceoniger (24%). Several discrete regions were highly conserved throughout the amino acid sequences of all these enzymes. To identify the histidine residue of the active site and to elucidate its function during enzymatic xylose or glucose isomerization, histidine residues at four different positions in the C. thermosulfurogenes enzyme were individually modified by site-directed mutagenesis. Substitution of His101 by phenylalanine completely abolished enzyme activity whereas substitution of other histidine residues by phenylalanine had no effect on enzyme activity. When His101 was changed to glutamine, glutamic acid, asparagine, or aspartic acid, approximately 10-16% of wild-type enzyme activity was retained by the mutant enzymes. The Gln101 mutant enzyme was resistant to diethylpyrocarbonate inhibition which completely inactivated the wild-type enzyme, indicating that His101 is the only essential histidine residue involved directly in enzyme catalysis. The constant Vmax values of the Gln101, Glu101, Asn101, and Asp101 mutant enzymes over the pH range of 5.0-8.5 indicate that protonation of His101 is responsible for the reduced Vmax values of the wild-type enzyme at pH below 6.5. Deuterium isotope effects by D-[2-2H]glucose on the rate of glucose isomerization indicated that hydrogen transfer and not substrate ring opening is the rate-determining step for both the wild-type and Gln101 mutant enzymes. These results suggest that the enzymatic sugar isomerization does not involve a histidine-catalyzed proton transfer mechanism. Rather, essential histidine functions to stabilize the transition state by hydrogen bonding to the C5 hydroxyl group of the substrate and this enables a metal-catalyzed hydride shift from C2 to C1.  相似文献   

18.
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   

19.
Genetically altered transacylase (E2b) proteins of the bovine branched-chain alpha-keto acid dehydrogenase complex were overexpressed in Escherichia coli and characterized. Deletion by PstI or Bal31 digestion of the amino-terminal region of the inner-core domain (residues 175-421) beyond residue 209 resulted in a complete loss of transacylase activity. The enzyme assay was carried out using [1-14C]isovaleryl-CoA and exogenous dihydrolipoamide as substrates. The removal of 4 residues (Thr-Ile-Pro-Ile) (residues 175-178) from the amino terminus of the inner-core domain significantly reduced the level of transacylase activity. The results establish that the segment between residues 175 and 209 is an integral part of the active site of E2b. The residue His-391 in the recombinant inner-core domain (E2b delta 167) was changed to Asn or Gln by site-directed mutagenesis. The wild-type and the two mutant inner-core domains were assembled into 24-mers as determined by gel filtration. However, both Asn and Gln mutations were accompanied by a complete loss of the enzymatic activity. Titration of the natural branched-chain alpha-keto dehydrogenase complex from pH 8 to 6 sharply reduced transacylase activity. The above data support the hypothesis that a conserved histidine residue in E2 acts as a general base for the transacylation reaction by analogy with E. coli chloramphenicol acetyltransferases.  相似文献   

20.
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   

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