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.
Zhang L  Ahvazi B  Szittner R  Vrielink A  Meighen E 《Biochemistry》2000,39(47):14409-14418
Aldehyde dehydrogenases (ALDHs) catalyze the transfer to NAD(P) of a hydride ion from a thiohemiacetal derivative of the aldehyde coupled with a cysteine residue in the active site. In Vibrio harveyi aldehyde dehydrogenase (Vh-ALDH), a histidine residue (H450) is in proximity (3.8 A) to the cysteine nucleophile (C289) and is thus capable of increasing its reactivity in sharp contrast to other ALDHs in which more distantly located glutamic acid residues are proposed to act as the general base. Mutation of H450 in Vh-ALDH to Gln and Asn resulted in loss of dehydrogenase, (thio)esterase, and acyl-CoA reductase activities; the residual activity of H450Q was higher than that of the H450N mutant in agreement with the capability of Gln but not Asn to partially replace the epsilon-imino group of H450. Coupled with a change in the rate-limiting step, these results indicate that H450 increases the reactivity of C289. Moreover, for the first time, the acylated enzyme intermediate could be directly monitored after reaction with [(3)H]tetradecanoyl-CoA showing that the H450Q mutant was acylated more rapidly than the H450N mutant. Inactivation of the wild-type enzyme with N-ethylmaleimide was much more rapid than the H450Q mutant which in turn was faster than the H450N mutant, demonstrating directly that the nucleophilicity of C289 was affected by H450. As the glutamic acid residue implicated as the general base in promoting cysteine nucleophilicity in other ALDHs is conserved in Vh-ALDH, elucidation of why a histidine residue has evolved to assist in this function in Vh-ALDH will be important to understand the mechanism of ALDHs in general, as well as help delineate the specific roles of the active site glutamic acid residues.  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.
Klebsiella pneumoniae produces 3-hydroxypropionic acid (3-HP) from glycerol with oxidation of 3-hydroxypropionaldehyde (3-HPA) to 3-HP in a reaction catalyzed by aldehyde dehydrogenase (ALDH). In the present study, two putative ALDHs of K. pneumoniae, YneI and YdcW were identified and characterized. Recombinant YneI was specifically active on 3-HPA and preferred NAD+ as a cofactor, whereas YdcW exhibited broad substrate specificity and preferred NADP+ as a cofactor. Overexpression of ALDHs in the glycerol oxidative pathway-deficient mutant K. pneumoniae AK resulted in a significant increase in 3-HP production upon shake-flask culture. The final titers of 3-HP were 2.4 and 1.8 g L?1 by recombinants overexpressing YneI and YdcW, respectively. Deletion of the ALDH gene from K. pneumoniae did not affect the extent of 3-HP synthesis, implying non-specific activity of ALDHs on 3-HPA. The ALDHs might play major roles in detoxifying the aldehyde generated in glycerol metabolism.  相似文献   

5.
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   

6.
Dehydrogenase activity of the cytoplasmic (E1) isozyme of human liver aldehyde dehydrogenase (EC 1.2.1.3) was almost totally abolished (3% activity remaining) by preincubation with dicyclohexylcarbodiimide (DCC), while esterase activity with p-nitrophenyl acetate as substrate remained intact. The esterase reaction of the modified enzyme exhibited a hysteretic burst prior to achieving steady-state velocity; addition of NAD+ abolished the burst. TheK m for p-nitrophenyl acetate was increased, but physicochemical properties remained unchanged. The selective inactivation of dehydrogenase activity was the result of covalent bond formation. Protection by NAD+ and chloral, saturation kinetics, and the stoichiometry and specificity of interaction indicated that the reaction of DCC occurred at the active site of the E1 isozyme. The results suggested that some amino acid other than aspartate or glutamate, possibly a cysteine residue, located on a large tryptic peptide of the E1 enzyme, may have reacted with DCC.  相似文献   

7.
Saturation transfer difference (STD) NMR experiments on Escherichia coli and Drosophila melanogaster succinic semialdehyde dehydrogenase (SSADH, EC1.2.1.24) suggest that only the aldehyde forms and not the gem-diol forms of the specific substrate succinic semialdehyde (SSA), of selected aldehyde substrates, and of the inhibitor 3-tolualdehyde bind to these enzymes. Site-directed mutagenesis of the active site cysteine311 to alanine in D. melanogaster SSADH leads to an inactive product binding both SSA aldehyde and gem-diol. Thus, the residue cysteine311 is crucial for their discrimination. STD experiments on SSADH and NAD+/NADP+ indicate differential affinity in agreement with the respective cosubstrate properties. Epitope mapping by STD points to a strong interaction of the NAD+/NADP+ adenine H2 proton with SSADH. Adenine H8, nicotinamide H2, H4, and H6 also show STD signals. Saturation transfer to the ribose moieties is limited to the anomeric protons of E. coli SSADH suggesting that the NAD+/NADP+ adenine and nicotinamide, but not the ribose moieties are important for the binding of the coenzymes.  相似文献   

8.
Aldehyde dehydrogenases (ALDHs) have been well established in all three domains of life and were shown to play essential roles, e.g., in intermediary metabolism and detoxification. In the genome of Sulfolobus solfataricus, five paralogs of the aldehyde dehydrogenases superfamily were identified, however, so far only the non-phosphorylating glyceraldehyde-3-phosphate dehydrogenase (GAPN) and α-ketoglutaric semialdehyde dehydrogenase (α-KGSADH) have been characterized. Detailed biochemical analyses of the remaining three ALDHs revealed the presence of two succinic semialdehyde dehydrogenase (SSADH) isoenzymes catalyzing the NAD(P)+-dependent oxidation of succinic semialdehyde. Whereas SSO1629 (SSADH-I) is specific for NAD+, SSO1842 (SSADH-II) exhibits dual cosubstrate specificity (NAD(P)+). Physiological significant activity for both SSO-SSADHs was only detected with succinic semialdehyde and α-ketoglutarate semialdehyde. Bioinformatic reconstructions suggest a major function of both enzymes in γ-aminobutyrate, polyamine as well as nitrogen metabolism and they might additionally also function in pentose metabolism. Phylogenetic studies indicated a close relationship of SSO-SSALDHs to GAPNs and also a convergent evolution with the SSADHs from E. coli. Furthermore, for SSO1218, methylmalonate semialdehyde dehydrogenase (MSDH) activity was demonstrated. The enzyme catalyzes the NAD+- and CoA-dependent oxidation of methylmalonate semialdehyde, malonate semialdehyde as well as propionaldehyde (PA). For MSDH, a major function in the degradation of branched chain amino acids is proposed which is supported by the high sequence homology with characterized MSDHs from bacteria. This is the first report of MSDH as well as SSADH isoenzymes in Archaea.  相似文献   

9.
The aldehyde dehydrogenases (ALDHs) are a superfamily of multimeric enzymes which catalyse the oxidation of a broad range of aldehydes into their corresponding carboxylic acids with the reduction of their cofactor, NAD or NADP, into NADH or NADPH. At present, the only known structures concern NAD-dependent ALDHs. Three structures are available in the Protein Data Bank: two are tetrameric and the other is a dimer. We solved by molecular replacement the first structure of an NADP-dependent ALDH isolated from Streptococcus mutans, in its apo form and holo form in complex with NADP, at 1.8 and 2.6 A resolution, respectively. Although the protein sequence shares only approximately 30 % identity with the other solved tetrameric ALDHs, the structures are very similar. However, a large local conformational change in the region surrounding the 2' phosphate group of the adenosine moiety is observed when the enzyme binds NADP, in contrast to the NAD-dependent ALDHs.Structure and sequence analyses reveal several properties. A small number of residues seem to determine the oligomeric state. Likewise, the nature (charge and volume) of the residue at position 180 (Thr in ALDH from S. mutans) determines the cofactor specificity in comparison with the structures of NAD-dependent ALDHs. The presence of a hydrogen bond network around the cofactor not only allows it to bind to the enzyme but also directs the side-chains in a correct orientation for the catalytic reaction to take place. Moreover, a specific part of this network appears to be important in substrate binding. Since the enzyme oxidises the same substrate, glyceraldehyde-3-phosphate (G3P), as NAD-dependent phosphorylating glyceraldehyde-3-phosphate dehydrogenases (GAPDH), the active site of GAPDH was compared with that of the S. mutans ALDH. It was found that Arg103, Arg283 and Asp440 might be key residues for substrate binding.  相似文献   

10.
The recently identified benzoate oxidation (box) pathway in Burkholderia xenovorans LB400 (LB400 hereinafter) assimilates benzoate through a unique mechanism where each intermediate is processed as a coenzyme A (CoA) thioester. A key step in this process is the conversion of 3,4-dehydroadipyl-CoA semialdehyde into its corresponding CoA acid by a novel aldehyde dehydrogenase (ALDH) (EC 1.2.1.x). The goal of this study is to characterize the biochemical and structural properties of the chromosomally encoded form of this new class of ALDHs from LB400 (ALDHC) in order to better understand its role in benzoate degradation. To this end, we carried out kinetic studies with six structurally diverse aldehydes and nicotinamide adenine dinucleotide (phosphate) (NAD + and NADP +). Our data definitively show that ALDHC is more active in the presence of NADP + and selective for linear medium-chain to long-chain aldehydes. To elucidate the structural basis for these biochemical observations, we solved the 1.6-Å crystal structure of ALDHC in complex with NADPH bound in the cofactor-binding pocket and an ordered fragment of a polyethylene glycol molecule bound in the substrate tunnel. These data show that cofactor selectivity is governed by a complex network of hydrogen bonds between the oxygen atoms of the 2′-phosphoryl moiety of NADP + and a threonine/lysine pair on ALDHC. The catalytic preference of ALDHC for linear longer-chain substrates is mediated by a deep narrow configuration of the substrate tunnel. Comparative analysis reveals that reorientation of an extended loop (Asn478-Pro490) in ALDHC induces the constricted structure of the substrate tunnel, with the side chain of Asn478 imposing steric restrictions on branched-chain and aromatic aldehydes. Furthermore, a key glycine (Gly104) positioned at the mouth of the tunnel allows for maximum tunnel depth required to bind medium-chain to long-chain aldehydes. This study provides the first integrated biochemical and structural characterization of a box-pathway-encoded ALDH from any organism and offers insight into the catalytic role of ALDHC in benzoate degradation.  相似文献   

11.
The mitochondrial isozyme of horse liver aldehyde dehydrogenase was labeled with brominated [5-(3-acetylpyridinio)pentyl]diphosphoadenosine. Specific labeling of a coenzyme binding region was proven by an enzymatic activity of the isozyme with the nonbrominated coenzyme derivative, optical properties of the complex, stoichiometry of incorporation, and protection against inactivation. A cysteine residue was selectively modified by the brominated coenzyme analogue and was identified in a 35-residue tryptic peptide. This cysteine residue corresponds to Cys-302 of the cytoplasmic isozyme and has earlier been implicated in disulfiram binding, confirming a position close to the active site. In contrast, the butyl homologue of the coenzyme analogue labels another residue of the mitochondrial isozyme. Thus, in the same isozyme, two residues are selectively reactive. They are concluded to be close together in the tertiary structure and to be close enough to the coenzyme binding site to be differentially labeled by coenzyme analogues differing only by a single methylene group.  相似文献   

12.
Aldehyde dehydrogenases (ALDHs) are a diverse family of enzymes that catalyze the NAD(P)+-dependent detoxification of toxic aldehyde compounds. ALDHs are also involved in non-enzymatic ligand binding to endobiotics and xenobiotics. Here, the enzyme crucial non-canonical and non-catalytic interaction with kolaflavanone, a component of kolaviron, and a major bioflavonoid isolated from Garcinia kola (Bitter kola) was characterized by various spectroscopic and in silico approaches under simulated physiological condition. Kolaflavanone quenched the intrinsic fluorescence of ALDH in a concentration dependent manner with an effective quenching constant (Ksv) of 1.14 × 103 L.mol−1 at 25 °C. The enzyme has one binding site for kolaflavanone with a binding constant (Ka) of 2.57 × 104 L.mol−1 and effective Forster resonance energy transfer (FRET) of 4.87 nm. The bonding process was enthalpically driven. The reaction was not spontaneous and was predominantly characterized by Van der Waals forces and hydrogen bond. The flavonoid bonding slightly perturbed the secondary and tertiary structures of ALDH that was ‘tryptophan-gated’. The interaction was regulated by both diffusion and ionic strength. Molecular docking showed the binding of kolaflavanone was at the active site of ALDH and the participation of some amino acid residues in the complex formation with −9.6 kcal mol−1 binding energy. The profiles of atomic fluctuations indicated the rigidity of the ligand-binding site during the simulation. With these, ALDH as a subtle nano-particle determinant of kolaviron bioavailability and efficacy is hereby proposed.  相似文献   

13.
In sorbitol dehydrogenase only one cysteine residue, Cys-43, is reactive in both anionic buffer (phosphate) and zinc-liganding buffer (imidazole) upon carboxymethylation. This is in contrast to the situation in the structurally related liver alcohol dehydrogenase, with either of two alternative Cys residues being reactive, and is compatible with differences in zinc-binding and active site relationships between these two metalloenzymes. Unrelated aldehyde dehydrogenase, upon carboxamidomethylation, shows a third pattern, now less well defined but confirming the presence of a thiol function of Cys-302 close to the active site.  相似文献   

14.
The cofactor-binding site of the NAD+-dependent Arabidopsis thaliana aldehyde dehydrogenase ALDH3H1 was analyzed to understand structural features determining cofactor-specificity. Homology modeling and mutant analysis elucidated important amino acid residues. Glu149 occupies a central position in the cofactor-binding cleft, and its carboxylate group coordinates the 2′- and 3′-hydroxyl groups of the adenosyl ribose ring of NAD+ and repels the 2′-phosphate moiety of NADP+. If Glu149 is mutated to Gln, Asp, Asn or Thr the binding of NAD+ is altered and rendered the enzyme capable of using NADP+. This change is attributed to a weaker steric hindrance and elimination of the electrostatic repulsion force of the 2′-phosphate of NADP+. Simultaneous mutations of Glu149 and Ile200, which is situated opposite of the cofactor binding cleft, improved the enzyme efficiency with NADP+. The double mutant ALDH3H1Glu149Thr/Ile200Val showed a good catalysis with NADP+. Subsequently a triple mutation was generated by replacing Val178 by Arg in order to create a “closed” cofactor binding site. The cofactor specificity was shifted even further in favor of NADP+, as the mutant ALDH3H1E149T/V178R/I200V uses NADP+ with almost 7-fold higher catalytic efficiency compared to NAD+. Our experiments suggest that residues occupying positions equivalent to 149, 178 and 200 constitute a group of amino acids in the ALDH3H1 protein determining cofactor affinity.  相似文献   

15.
We have proposed developing rat hepatoma cell lines as an in vitro model for studying the regulation of changes in aldehyde dehydrogenase activity occurring duringhepatocarcinogenesis. Aldehyde dehydrogenase purified in a single step from HTC rat hepatoma cells is identical to the aldehyde dehydrogenase isolated from rat hepatocellular carcinomas. HTC aldehyde dehydrogenase is a 110 kDa dimer composed of 54-kDa subunits, prefers NADP+ as coenzyme, and preferentially oxidizes benzaldehyde-like aromatic aldehydes but not phenylacetaldehyde. The substrate and coenzyme specificity, effects of disulfiram, pH profile and isoelectric point of HTC aldehyde dehydrogenase are also identical to these same properties of the tumor aldehyde dehydrogenase. In immunodiffusions, both isozymes are recognized with complete identity by anti-HTC aldehyde dehydrogenase antibodies. Having established that HTC aldehyde dehydrogenase is very similar, if not identical, to the aldehyde dehydrogenase found in hepatocellular carcinomas, simplifies the development of molecular probes for examination of the regulation of tumor aldehyde dehydrogenase activity in vivo and in vitro.  相似文献   

16.
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   

17.
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   

18.
Aldehyde dehydrogenases (ALDHs) couple the oxidation of aldehydes to the reduction of NAD(P)+. These enzymes have gained importance as they have been related to the detoxification of aldehydes generated in several diseases involving oxidative stress. It has been determined that tetrameric ALDHs work only with two of their four active sites (half‐of‐the‐sites reactivity), but the mechanistic reason for this feature remains unknown. In this study, tetrameric human aldehyde dehydrogenase class 1A1 (ALDH1A1) was dimerized to study the correlation of the oligomeric structure with the presence of half‐of‐the‐sites reactivity. Stable dimers from ALDH1A1 were generated by combining the mutation of two residues of the dimer–dimer interface in the tetramer (previously shown to render a low‐active and unstable enzyme) and the fusion of green fluorescent protein (GFP) in the C‐terminus of the mutant. Some kinetic properties of the GFP‐fusion mutant resembled those of human aldehyde dehydrogenase class 3A1, a native dimer, in that the fusion dimer did not show burst in the generation of nicotinamide adenine dinucleotide (NADH) and was less sensitive to the action of specific modulators. The presence of primary isotope effect indicated that the rate‐limiting step changed from NADH release to hydride transfer. The mutant showed higher activity with malondialdehyde and acrolein and was more resistant to inactivation by acrolein compared with the wild type. The mutant kinetic profile showed two hyperbolic components when the substrates were varied, suggesting the presence of two active sites with different affinities and catalytic capacities. In conclusion, the ALDH1A1–GFP dimeric mutant exhibits full site reactivity, suggesting that only the tetrameric structure induces the half‐of‐the‐sites reactivity. Proteins 2013; 81:1330–1339. © 2013 Wiley Periodicals, Inc.  相似文献   

19.
The aldehyde dehydrogenase activity of the sulfate-reducing bacterium Desulfovibrio simplex strain DSM 4141 was characterized in cell-free extracts. Oxygen-sensitive, constitutive aldehyde dehydrogenase activity was found in cells grown on l(+)-lactate, hydrogen, or vanillin with sulfate as the electron acceptor. A 1.83- to 2.6-fold higher specific activity was obtained in cells grown in media supplemented with 1 μM WO4 2–. The aldehyde dehydrogenase in cell-free extracts catalyzed the oxidation of aliphatic (K m < 20 μM) and aromatic aldehydes (K m < 0.32 mM) using methyl viologen as the electron acceptor. Flavins (FMN and FAD) were also active and are proposed to be the natural cofactors, while no activity was obtained with NAD+ or NADP+. 185WO4 2– was incorporated in vivo into D. simplex; it was found exclusively in the soluble fraction (≥ 98%). Anionic-exchange chromatography demonstrated coelution of 185W with two distinct peaks, the first one containing hydrogenase and formate dehydrogenase activities, and the second one aldehyde dehydrogenase activity. Received: 7 February 1997 / Accepted: 6 June 1997  相似文献   

20.
A gene (pegC) encoding aldehyde dehydrogenase (ALDH) was located 3.4 kb upstream of a gene encoding polyethylene glycol (PEG) dehydrogenase (pegA) in Sphingomonas macrogoltabidus strain 103. ALDH was expressed in Escherichia coli and purified on a Ni-nitrilotriacetic acid agarose column. The recombinant enzyme was a homotetramer consisting of four 46.1-kDa subunits. The alignment of the putative amino acid sequence of the cloned enzyme showed high similarity with a group of NAD(P)-dependent ALDHs (identity 36–52%); NAD-binding domains (Rossmann fold and four glycine residues) and catalytic residues (Glu225 and Cys259) were well conserved. The cofactor, which was extracted from the purified enzyme, was tightly bound to the enzyme and identified as NADP. The enzyme contained 0.94 mol NADP per subunit. The enzyme was activated by Ca2+, but by no other metals; no metal (Zn, Fe, Mg, or Mn) was detected in the purified recombinant enzyme. Activity was inhibited by p-chloromercuric benzoate, and heavy metals such as Hg, Cu, Pb and Cd, indicating that a cysteine residue is involved in the activity. Enzyme activity was independent of N,N-dimethyl-p-nitrosoaniline as an electron acceptor. Trans-4-(N,N-dimethylamino)-cinnamaldehyde was not oxidized as a substrate, but the compound worked as an inhibitor for the enzyme, as did pyrazole. The enzyme acted on n-aldehydes C2–C14) and PEG-aldehydes. Thus the enzyme was concluded to be a novel Ca2+-activating nicotinoprotein (NADP-containing) PEG-aldehyde dehydrogenase involved in the degradation of PEG in S. macrogoltabidus strain 103.  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