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 358 毫秒
1.
We have previously shown that intact plants and cultured plant cells can metabolize and detoxify formaldehyde through the action of a glutathione-dependent formaldehyde dehydrogenase (FDH), followed by C-1 metabolism of the initial metabolite (formic acid). The cloning and heterologous expression of a cDNA for the glutathione-dependent formaldehyde dehydrogenase from Zea mays L. is now described. The functional expression of the maize cDNA in Escherichia coli proved that the cloned enzyme catalyses the NAD+- and glutathione (GSH)-dependent oxidation of formaldehyde. The deduced amino acid sequence of 41 kDa was on average 65% identical with class III alcohol dehydrogenases from animals and less than 60% identical with conventional plant alcohol dehydrogenases (ADH) utilizing ethanol. Genomic analysis suggested the existence of a single gene for this cDNA. Phylogenetic analysis supports the convergent evolution of ethanol-consuming ADHs in animals and plants from formaldehyde-detoxifying ancestors. The high structural conservation of present-day glutathione-dependent FDH in microorganisms, plants and animals is consistent with a universal importance of these detoxifying enzymes.  相似文献   

2.
The crystal structure of NAD(+)-dependent alcohol dehydrogenase from Bacillus stearothermophilus strain LLD-R (htADH) was determined using X-ray diffraction data at a resolution of 2.35 A. The structure of homotetrameric htADH is highly homologous to those of bacterial and archaeal homotetrameric alcohol dehydrogenases (ADHs) and also to the mammalian dimeric ADHs. There is one catalytic zinc atom and one structural zinc atom per enzyme subunit. The enzyme was crystallized as a binary complex lacking the nicotinamide adenine dinucleotide (NAD(+)) cofactor but including a zinc-coordinated substrate analogue trifluoroethanol. The binary complex structure is in an open conformation similar to ADH structures without the bound cofactor. Features important for the thermostability of htADH are suggested by a comparison with a homologous mesophilic enzyme (55% identity), NAD(+)-dependent alcohol dehydrogenase from Escherichia coli. To gain insight into the conformational change triggered by NAD(+) binding, amide hydrogen-deuterium exchange of htADH, in the presence and absence of NAD(+), was studied by HPLC-coupled electrospray mass spectrometry. When the deuteron incorporation of the protein-derived peptides was analyzed, it was found that 9 of 21 peptides show some decrease in the level of deuteron incorporation upon NAD(+) binding, and another 4 peptides display slower exchange rates. With one exception (peptide number 8), none of the peptides that are altered by bound NAD(+) are in contact with the alcohol-substrate-binding pocket. Furthermore, peptides 5 and 8, which are located outside the NAD(+)-binding pocket, are notable by displaying changes upon NAD(+) binding. This suggests that the transition from the open to the closed conformation caused by cofactor binding has some long-range effects on the protein structure and dynamics.  相似文献   

3.
Abstract A fragment of Methylobacter marinus A45 DNA has been cloned and sequenced, and an open reading frame has been identified that could code for a 46-kDa polypeptide. Comparison of the deduced amino acid sequence of the polypeptide against the protein data bank has revealed strong similarity with a number of alcohol dehydrogenases, with highest similarity towards class III alcohol dehydrogenases, which recently have been shown to be identical to glutathione-dependent formaldehyde dehydrogenases. We were unable to measure appreciable levels of NAD(P)-dependent formaldehyde dehydrogenases or alcohol dehydrogenase activities using aldehydes or primary or secondary alcohols in cell-free extracts from batch cultures of M. marinus A45. However, formaldehyde dehydrogenases activity was detected on zymograms. Our data suggest that, although NAD(P)-linked formaldehyde dehydrogenase or alcohol dehydrogenase activities are undetectable in cell-free extracts of most methylotrophs employing the ribulose monophosphate pathway for formaldehyde assimilation and dissimilation, the gene encoding formaldehyde dehydrogenase is present in M. marinus A45 and may be present in more of these organisms as well.  相似文献   

4.
L-threonine dehydrogenase (TDH) is an enzyme that catalyzes the oxidation of L-threonine to 2-amino-3-ketobutyrate. We solved the first crystal structure of a medium chain L-threonine dehydrogenase from a hyperthermophilic archaeon, Pyrococcus horikoshii (PhTDH), by the single wavelength anomalous diffraction method using a selenomethionine-substituted enzyme. This recombinant PhTDH is a homo-tetramer in solution. Three monomers of PhTDHs were located in the crystallographic asymmetric unit, however, the crystal structure exhibits a homo-tetramer structure with crystallographic and non-crystallographic 222 symmetry in the cell. Despite the low level of sequence identity to a medium-chain NAD(H)-dependent alcohol dehydrogenase (ADH) and the different substrate specificity, the overall folds of the PhTDH monomer and tetramer are similar to those of the other ADH. Each subunit is composed of two domains: a nicotinamide cofactor (NAD(H))-binding domain and a catalytic domain. The NAD(H)-binding domain contains the alpha/beta Rossmann fold motif, characteristic of the NAD(H)-binding protein. One molecule of PhTDH contains one zinc ion playing a structural role. This metal ion exhibits coordination with four cysteine ligands and some of the ligands are conserved throughout the structural zinc-containing ADHs and TDHs. However, the catalytic zinc ion that is coordinated at the bottom of the cleft in the case of ADH was not observed in the crystal of PhTDH. There is a significant difference in the orientation of the catalytic domain relative to the coenzyme-binding domain that results in a larger interdomain cleft.  相似文献   

5.
Sorbitol dehydrogenase (SDH) is a distant relative to the alcohol dehydrogenases (ADHs) with sequence identities around 20%. SDH is a tetramer with one zinc ion per subunit. We have crystallized rat SDH and determined the structure by molecular replacement using a tetrameric bacterial ADH as search object. The conformation of the bound coenzyme is extended and similar to NADH bound to mammalian ADH but the interactions with the NMN-part have several differences with those of ADH. The active site zinc coordination in SDH is significantly different than in mammalian ADH but similar to the one found in the bacterial tetrameric NADP(H)-dependent ADH of Clostridiim beijerinckii. The substrate cleft is significantly more polar than for mammalian ADH and a number of residues are ideally located to position the sorbitol molecule in the active site. The SDH molecule can be considered to be a dimer of dimers, with subunits A-B and C-D, where the dimer interactions are similar to those in mammalian ADH. The tetramers are composed of two of these dimers, which interact with their surfaces opposite the active site clefts, which are accessible on the opposite side. In contrast to the dimer interactions, the tetramer-forming interactions are small with only few hydrogen bonds between side-chains.  相似文献   

6.
Wang N  Chang C  Yao Q  Li G  Qin L  Chen L  Chen K 《PloS one》2011,6(6):e21454
Alcohol dehydrogenases (ADHs) are oxidoreductases catalyzing the reversible oxidation of alcohols to corresponding aldehydes or ketones accompanied by nicotinamide adenine dinucleotide (NAD) or nicotinamide adenine dinucleotide phosphate (NADP) as coenzyme. ADHs attract major scientific and industrial interest for the evolutionary perspectives, afforded by their wide occurrence in nature, and for their use in industrial synthesis. However, the low activity of ADHs under extremes of pH and temperature often limits their application. To obtain ADH with high activity, in this study, we used Bombyx mori alcohol dehydrogenases (BmADH) as foreign gene and constructed a recombinant integrative plasmid pJS700-BmADH. This pJS700-BmADH was transformed into Bacillus subtilis by double cross-over and produced an amylase inactivated mutant. The fusion protein containing BmADH was expressed on the spore surface and recognized by BmADH-specific antibody. We also assayed the alcohol dehydrogenase activity of the fusion protein together with the native BmADH at different pH and temperature levels, which indicated the recombinant enzyme exhibits activity over wider ranges of temperature and pH than its native form, perhaps due to the resistance properties of B. subtilis spores against adverse conditions.  相似文献   

7.
Sorbitol dehydrogenase (SDH) is a distant relative to the alcohol dehydrogenases (ADHs) with sequence identities around 20%. SDH is a tetramer with one zinc ion per subunit. We have crystallized rat SDH and determined the structure by molecular replacement using a tetrameric bacterial ADH as search object. The conformation of the bound coenzyme is extended and similar to NADH bound to mammalian ADH but the interactions with the NMN-part have several differences with those of ADH. The active site zinc coordination in SDH is significantly different than in mammalian ADH but similar to the one found in the bacterial tetrameric NADP(H)-dependent ADH of Clostridiim beijerinckii. The substrate cleft is significantly more polar than for mammalian ADH and a number of residues are ideally located to position the sorbitol molecule in the active site. The SDH molecule can be considered to be a dimer of dimers, with subunits A–B and C–D, where the dimer interactions are similar to those in mammalian ADH. The tetramers are composed of two of these dimers, which interact with their surfaces opposite the active site clefts, which are accessible on the opposite side. In contrast to the dimer interactions, the tetramer-forming interactions are small with only few hydrogen bonds between side-chains.  相似文献   

8.
Abstract: Alcohol dehydrogenase (ADH) is a key enzyme for the production of butanol, ethanol, and isopropanol by the solvent-producing clostridia. Initial studies of ADH in extracts of several strains of Clostridium acetobutylicum and C. beijerinckii gave conflicting molecular properties. A more coherent picture has emerged because of the following results: (i) identification of ADHs with different coenzyme specificities in these species; (ii) discovery of structurally conserved ADHs (type 3) in three solvent-producing species; (iii) isolation of mutants with deficiencies in butanol production and restoration of butanol production with a cloned alcohol/aldehyde dehydrogenase gene; and (iv) resolution of various ' C. acetobutylicum ' cultures into four species. The three ADH isozymes of C. beijerinckii NRRL B592 have high sequence similarities to ADH-1 of Clostridium sp. NCP 262 (formerly C. acetobutylicum P262) and to the ADH domain of the alcohol/aldehyde dehydrogenase of C. acetobutylicum ATCC 824/DSM 792. The NADH-dependent activity of the ADHs from C. beijerinckii NRRL B592 and the BDHs from C. acetobutylicum ATCC 824 is profoundly affected by the pH of the assay, and the relative importance of NADH and NADPH to butanol production may be misappraised when NAD(P)H-dependent activities were measured at different pH values. The primary/secondary ADH of isopropanol-producing C. beijerinckii is a type-1 enzyme and is highly conserved in Thermoanaerobacter brockii (formerly Thermoanaerobium brockii ) and Entamoeba histolytica . Several solvent-forming enzymes (primary ADH, aldehyde dehydrogenase, and 3-hydroxybutyryl-CoA dehydrogenase) are very similar between C. beijerinckii and the species represented by Clostridium sp. NCP 262 and NRRL B643. The realization of such relationships will facilitate the elucidation of the roles of different ADHs because each type of ADH can now be studied in an organism most amenable to experimental manipulations.  相似文献   

9.
甲醛脱氢酶(formaldehyde dehydrogenase,ADH)与甲酸脱氢酶(formate dehydrogenase,FDH)是甲醛氧化途径的两个关键酶.恶臭假单胞菌(Pseudomonas putida)的PADH是一种不依赖谷胱甘肽可以把游离甲醛直接氧化为甲酸的脱氢酶,博伊丁假丝酵母菌(Candida boidinii)的FDH在有NAD+存在时可以把甲酸氧化为二氧化碳.以基因组DNA为模板用PCR方法,从P.putida中扩增出PADH基因的编码区(padh),从C.boidinii中扩增出FDH的编码区(fdh),然后亚克隆到pET-28a(+)中分别构建这两个基因的原核表达载体pET-28a-padh和pET-28a-fdh,转化大肠杆菌,利用IPTG诱导重组蛋白PADH和FDH的表达.通过优化条件使重组蛋白的表达量占菌体总蛋白的70%以上,通过亲和层析法纯化出可溶性PADH和FDH重组蛋白.对重组蛋白的生化特性分析结果表明:PADH在最适反应温度50℃的活性为1.95 U/mg;FDH在最适反应温度40℃的活性为0.376 U/mg.所表达的重组蛋白与之前报道过的相比,具有更好的热稳定性和更广的温度适应范围.将PADH、FDH两个重组蛋白及辅因子NAD+固定到聚丙烯酰胺载体基质上,对固定化酶甲醛吸收效果的初步分析结果显示固定化酶对空气中的甲醛有一定的吸收效果,说明这两种酶被固定后具有开发成治理甲醛污染环保产品的潜力.  相似文献   

10.
The crystal structure of a medium-chain NAD(H)-dependent alcohol dehydrogenase (ADH) from an archaeon has been solved by multiwavelength anomalous diffraction, using a selenomethionine-substituted enzyme. The protein (SsADH), extracted from the hyperthermophilic organism Sulfolobus solfataricus, is a homo-tetramer with a crystallographic 222 symmetry. Despite the low level of sequence identity, the overall fold of the monomer is similar to that of the other homologous ADHs of known structure. However, a significant difference is the orientation of the catalytic domain relative to the coenzyme-binding domain that results in a larger interdomain cleft. At the bottom of this cleft, the catalytic zinc ion is coordinated tetrahedrally and lacks the zinc-bound water molecule that is usually found in ADH apoform structures. The fourth coordination position is indeed occupied by a Glu residue, as found in bacterial tetrameric ADHs. Other differences are found in the architecture of the substrate pocket whose entrance is more restricted than in other ADHs. SsADH is the first tetrameric ADH X-ray structure containing a second zinc ion playing a structural role. This latter metal ion shows a peculiar coordination, with a glutamic acid residue replacing one of the four cysteine ligands that are highly conserved throughout the structural zinc-containing dimeric ADHs.  相似文献   

11.
An homology model of Candida methylica formate dehydrogenase (cmFDH) was constructed based on the Pseudomonas sp. 101 formate dehydrogenase (psFDH) structure. An aspartic acid residue in the model was predicted to interact with the adenine ribose of the NAD cofactor, in common with many NAD-dependent oxoreductases. Replacement of this aspartic acid residue by serine in cmFDH removed the absolute requirement for NAD over NADP shown by the wild type enzyme. Taken with similar results shown by d- and l-lactate dehydrogenases, this suggests that an aspartic acid in this position is a major determinant of coenzyme specificity in NAD/NADP-dependent dehydrogenases.  相似文献   

12.
Glu-60 of the zinc-dependent Thermoanaerobacter brockii alcohol dehydrogenase (TbADH) is a strictly conserved residue in all members of the alcohol dehydrogenase (ADH) family. Unlike most other ADHs, the crystal structures of TbADH and its analogs, ADH from Clostridium beijerinckii (CbADH), exhibit a unique zinc coordination environment in which this conserved residue is directly coordinated to the catalytic zinc ion in the native form of the enzymes. To explore the role of Glu-60 in TbADH catalysis, we have replaced it by alanine (E60A-TbADH) and aspartate (E60D-TbADH). Steady-state kinetic measurements show that the catalytic efficiency of these mutants is only four- and eightfold, respectively, lower than that of wild-type TbADH. We applied X-ray absorption fine-structure (EXAFS) and near-UV circular dichroism to characterize the local environment around the catalytic zinc ion in the variant enzymes in their native, cofactor-bound, and inhibited forms. We show that the catalytic zinc site in the studied complexes of the variant enzymes exhibits minor changes relative to the analogous complexes of wild-type TbADH. These moderate changes in the kinetic parameters and in the zinc ion environment imply that the Glu-60 in TbADH does not remain bound to the catalytic zinc ion during catalysis. Furthermore, our results suggest that a water molecule replaces this residue during substrate turnover.  相似文献   

13.
Polyhydric alcohols are widely found in nature and can be accumulated to high concentrations as a protection against a variety of environmental stresses. It is only recently, however, that these molecules have been shown to be active in protection against heat stress, specifically in the use of sorbitol by the silverleaf whitefly, Bemisia argentifolii. We have determined the structure of the enzyme responsible for production of sorbitol in Bemisia argentifolii, NADP(H)-dependent ketose reductase (BaKR), to 2.3 A resolution. The structure was solved by multiwavelength anomalous diffraction (MAD) using the anomalous scattering from two zinc atoms bound in the structure, and was refined to an R factor of 21.9 % (R(free)=25.1 %). BaKR belongs to the medium-chain dehydrogenase family and its structure is the first for the sorbitol dehydrogenase branch of this family. The enzyme is tetrameric, with the monomer having a very similar fold to the alcohol dehydrogenases (ADHs). Although the structure determined is for the apo form, a phosphate ion in the active site marks the likely position for the adenyl phosphate of NADP(H). The catalytic zinc ion is tetrahedrally coordinated to Cys41, His66, Glu67 and a water molecule, in a modification of the zinc site usually found in ADHs. This modified zinc site seems likely to be a conserved feature of the sorbitol dehydrogenase sub-family. Comparisons with other members of the ADH family have also enabled us to model a ternary complex of the enzyme, and suggest how structural differences may influence coenzyme binding and substrate specificity in the reduction of fructose to sorbitol.  相似文献   

14.
Five different immobilized NAD+ derivatives were employed to compare the behavior of four amino acid dehydrogenases chromatographed using kinetic-based enzyme capture strategies (KBECS): S6-, N6-, N1-, 8'-azo-, and pyrophosphate-linked immobilized NAD+. The amino acid dehydrogenases were NAD+-dependent phenylalanine (EC 1.4.1.20), alanine (EC 1.4.1.1), and leucine (EC 1.4.1.9) dehydrogenases from various microbial species and NAD(P)+-dependent glutamate dehydrogenase from bovine liver (GDH; EC 1.4.1.3). KBECS for bovine heart L-lactate dehydrogenase (EC 1.1.1.27) and yeast alcohol dehydrogenase (EC 1.1.1.1) were also applied to assist in a preliminary assessment of the immobilized cofactor derivatives. Results confirm that the majority of the enzymes studied retained affinity for NAD+ immobilized through an N6 linkage, as opposed to an N1 linkage, replacement of the nitrogen with sulfur to produce an S6 linkage, or attachment of the cofactor through the C8 position or the pyrophosphate group of the cofactor. The one exception to this was the dual-cofactor-specific GDH from bovine liver, which showed no affinity for N6-linked NAD+ but was biospecifically adsorbed to S6-linked NAD+ derivatives in the presence of its soluble KBEC ligand. The molecular basis for this is discussed together with the implications for future development and application of KBECS.  相似文献   

15.
High NAD-dependent alcohol dehydrogenase (ADH) activity was found in the cytoplasm when a membrane-bound, quinoprotein, ADH-deficient mutant strain of Acetobacter pasteurianus SKU1108 was grown on ethanol. Two NAD-dependent ADHs were separated and purified from the supernatant fraction of the cells. One (ADH I) is a trimer, consisting of an identical subunit of 42 kDa, while the other (ADH II) is a homodimer, having a subunit of 31 kDa. One of the two ADHs, ADH II, easily lost the activity during the column chromatographies, which could be stabilized by the addition of DTT and MgCl2 in the column buffer. ADH I but not ADH II contained approximately one zinc atom per subunit. The N-terminal amino acid analysis indicated that ADH I and ADH II have homology to the long-chain and short-chain ADH families, respectively. ADH I showed a preference for primary alcohols, while ADH II had a preference for secondary alcohols. The two ADHs showed clear difference in their kinetics on ethanol, acetaldehyde, NAD, and NADH. The physiological function of both ADH I and ADH II are also discussed.  相似文献   

16.
The amphibian enzyme ADH8, previously named class IV-like, is the only known vertebrate alcohol dehydrogenase (ADH) with specificity towards NADP(H). The three-dimensional structures of ADH8 and of the binary complex ADH8-NADP(+) have been now determined and refined to resolutions of 2.2A and 1.8A, respectively. The coenzyme and substrate specificity of ADH8, that has 50-65% sequence identity with vertebrate NAD(H)-dependent ADHs, suggest a role in aldehyde reduction probably as a retinal reductase. The large volume of the substrate-binding pocket can explain both the high catalytic efficiency of ADH8 with retinoids and the high K(m) value for ethanol. Preference of NADP(H) appears to be achieved by the presence in ADH8 of the triad Gly223-Thr224-His225 and the recruitment of conserved Lys228, which define a binding pocket for the terminal phosphate group of the cofactor. NADP(H) binds to ADH8 in an extended conformation that superimposes well with the NAD(H) molecules found in NAD(H)-dependent ADH complexes. No additional reshaping of the dinucleotide-binding site is observed which explains why NAD(H) can also be used as a cofactor by ADH8. The structural features support the classification of ADH8 as an independent ADH class.  相似文献   

17.

Background

Alcohol dehydrogenases (ADHs) catalyze the reversible oxidation of alcohol using NAD+ or NADP+ as cofactor. Three ADH homologues have been identified in Komagataella phaffii GS115 (also named Pichia pastoris GS115), ADH1, ADH2 and ADH3, among which adh3 is the only gene responsible for consumption of ethanol in Komagataella phaffii GS115. However, the relationship between structure and function of mitochondrial alcohol dehydrogenase isozyme III from Komagataella phaffii GS115 (KpADH3) is still not clear yet.

Methods

KpADH3 was purified, identified and characterized by multiple biophysical techniques (Nano LC-MS/MS, Enzymatic activity assay, X-ray crystallography).

Results

The crystal structure of KpADH3, which was the first ADH structure from Komagataella phaffii GS115, was solved at 1.745?Å resolution. Structural analysis indicated that KpADH3 was the sole dimeric ADH structure with face-to-face orientation quaternary structure from yeast. The major structural different conformations located on residues 100–114 (the structural zinc binding loop) and residues 337–344 (the loop between α12 and β15 which covered the catalytic domain). In addition, three channels were observed in KpADH3 crystal structure, channel 2 and channel 3 may be essential for substrate specific recognition, ingress and egress, channel 1 may be the pass-through for cofactor.

Conclusions

KpADH3 plays an important role in the metabolism of alcohols in Komagataella phaffii GS115, and its crystal structure is the only dimeric medium-chain ADH from yeast described so far.

General significance

Knowledge of the relationship between structure and function of KpADH3 is crucial for understanding the role of KpADH3 in Komagataella phaffii GS115 mitochondrial metabolism.  相似文献   

18.
The escalation in genome sequencing has presented a mass of potentially useful new alcohol dehydrogenases (ADHs) in the form of putative open reading frame (ORF). To take advantage of such available resources, PCR primers based on the genome sequence of Bacillus thuringiensis serovar israelensis were used to clone a gene encoding a hypothetical alcohol dehydrogenase (named as BtADH). Activity studies of the translation product revealed that the alcohol dehydrogenases catalyse the inter-conversion of aliphatic aldehydes and corresponding primary alcohol with chain length of two to ten carbons. The required co-factor for such inter-conversion was found to be NAD(H). The ADH gene was engineered for heterologous expression in Escherichia coli, and the enzyme was produced in a soluble form. The recombinant enzyme was purified to homogeneity and physical, spectral and catalytical properties were determined.The findings lead us to propose that BtADH represents a novel primary–secondary alcohol dehydrogenase that acts on primary alcohols of medium chain lengths and simple ketones. Besides, BtADH shares high sequence similarity with well known ADHs from thermophilic origins. Such biochemical characterisation of BtADH provides valuable information for the study of sequence–function relationship including source of thermal stability, cofactor and substrate preferences.  相似文献   

19.
The kinetic locking-on strategy utilizes soluble analogues of the target enzymes' specific substrate to promote selective adsorption of individual NAD(+)-dependent dehydrogenases on their complementary immobilized cofactor derivative. Application of this strategy to the purification of NAD(+)-dependent dehydrogenases from crude extracts has proven that it can yield bioaffinity systems capable of producing one-chromatographic-step purifications with yields approaching 100%. However, in some cases the purified enzyme preparation was found to be contaminated with other proteins weakly bound to the immobilized cofactor derivative through binary complex formation and/or nonspecific interactions, which continuously "dribbled" off the matrix during the chromatographic procedure. The fact that this problem can be overcome by including a short pulse of 5'-AMP (stripping ligand) in the irrigant a couple of column volumes prior to the discontinuation of the specific substrate analogue (locking-on ligand) is clear from the results presented in this report. The general effectiveness of this auxiliary tactic has been assessed using model studies and through incorporation into an actual purification from a crude cellular extract. The results confirm the usefulness of the stripping-ligand tactic for the resolution and purification of NAD(+)-dependent dehydrogenases when using the locking-on strategy. These studies have been carried out using bovine liver glutamate dehydrogenase (GDH, EC 1.4.1.3), yeast alcohol dehydrogenase (YADH, EC 1.1.1.1), porcine heart mitochondrial malate dehydrogenase (mMDH, EC 1.1.1.37), and bovine heart L-lactate dehydrogenase (l-LDH, EC 1.1.1.27).  相似文献   

20.
Summary Sequences of 47 members of the Zn-containing alcohol dehydrogenase (ADH) family were aligned progressively, and an evolutionary tree with detailed branch order and branch lengths was produced. The alignment shows that only 9 amino acid residues (of 374 in the horse liver ADH sequence) are conserved in this family; these include eight Gly and one Val with structural roles. Three residues that bind the catalytic Zn and modulate its electrostatic environment are conserved in 45 members. Asp 223, which determines specificity for NAD, is found in all but the two NADP-dependent enzymes, which have Gly or Ala. Ser or Thr 48, which makes a hydrogen bond to the substrate, is present in 46 members. The four Cys ligands for the structural zinc are conserved except in -crystallin, the sorbitol dehydrogenases, and two bacterial enzymes. Analysis of the evolutionary tree gives estimates of the times of divergence for different animal ADHs. The human class II () and class III () ADHs probably diverged about 630 million years ago, and the newly identified human ADH6 appeared about 520 million years ago, implying that these classes of enzymes may exist or have existed in all vertebrates. The human class I ADH isoenzymes (, , and ) diverged about 80 million years ago, suggesting that these isoenzymes may exist or have existed in all primates. Analysis of branch lengths shows that these plant ADHs are more conserved than the animal ones and that class III ADHs are more conserved than class I ADHs. The rate of acceptance of point mutations (PAM units) shows that selection pressure has existed for ADHs, implying that these enzymes play definite metabolic roles.Offprint requests to: B.V. Plapp  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