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.
Class II alcohol dehydrogenase (ADH2) represents a highly divergent class of alcohol dehydrogenases predominantly found in liver. Several species variants of ADH2 have been described, and the rodent enzymes form a functionally distinct subgroup with interesting catalytic properties. First, as compared with other ADHs, the catalytic efficiency is low for this subgroup. Second, the substrate repertoire is unique, e.g. rodent ADH2s are not saturated with ethanol as substrate, and while ω-hydroxy fatty acids are common substrates for the human ADH1–ADH4 isoenzymes, including ADH2, these compounds function as inhibitors rather than substrates. The recently determined structure of mouse ADH2 reveals a novel substrate-pocket topography that accounts for the observed substrate specificity and may, therefore, be important for the exploration of orphan substrates of ADH2. It is possible to improve the catalytic efficiency of mouse ADH2 by an array of mutations at position 47. Residue Pro47 of the wild type ADH2 enzyme seems to strain the binding of coenzyme, which prevents a close approach between the coenzyme and substrate for efficient hydrogen transfer. Based on crystallographic and mechanistic investigations, the effects of residue replacements at position 47 are multiple, affecting the distance for hydride transfer, the pKa of the bound alcohol substrate as well as the affinity for coenzyme.  相似文献   

2.
ADH2 is a member of one of the six classes of mammalian alcohol dehydrogenases, which catalyze the reversible oxidation of alcohols using NAD(+) as a cofactor. Within the ADH2 class, the rodent enzymes form a subgroup that exhibits low catalytic activity with all substrates that were examined, as compared to other groups, such as human ADH2. The low activity can be ascribed to the rigid nature of the proline residue at position 47 as the activity can be increased by approximately 100-fold by substituting Pro47 with either His (as found in human ADH2), Ala, or Gln. Mouse ADH2 follows an ordered bi-bi mechanism, and hydride transfer is rate-limiting for oxidation of benzyl alcohols catalyzed by the mutated and wild-type enzymes. Structural studies suggest that the mouse enzyme with His47 has a more closed active site, as compared to the enzyme with Pro47, and hydride transfer can be more efficient. Oxidation of benzyl alcohol catalyzed by all forms of the enzyme is strongly pH dependent, with pK values in the range of 8.1-9.3 for turnover numbers and catalytic efficiency. These pK values probably correspond to the ionization of the zinc-bound water or alcohol. The pK values are not lowered by the Pro47 to His substitution, suggesting that His47 does not act as a catalytic base in the deprotonation of the zinc ligand.  相似文献   

3.
The structure of mouse class II alcohol dehydrogenase (ADH2) has been determined in a binary complex with the coenzyme NADH and in a ternary complex with both NADH and the inhibitor N-cyclohexylformamide to 2.2 A and 2.1 A resolution, respectively. The ADH2 dimer is asymmetric in the crystal with different orientations of the catalytic domains relative to the coenzyme-binding domains in the two subunits, resulting in a slightly different closure of the active-site cleft. Both conformations are about half way between the open apo structure and the closed holo structure of horse ADH1, thus resembling that of ADH3. The semi-open conformation and structural differences around the active-site cleft contribute to a substantially different substrate-binding pocket architecture as compared to other classes of alcohol dehydrogenase, and provide the structural basis for recognition and selectivity of alcohols and quinones. The active-site cleft is more voluminous than that of ADH1 but not as open and funnel-shaped as that of ADH3. The loop with residues 296-301 from the coenzyme-binding domain is short, thus opening up the pocket towards the coenzyme. On the opposite side, the loop with residues 114-121 stretches out over the inter-domain cleft. A cavity is formed below this loop and adds an appendix to the substrate-binding pocket. Asp301 is positioned at the entrance of the pocket and may control the binding of omega-hydroxy fatty acids, which act as inhibitors rather than substrates. Mouse ADH2 is known as an inefficient ADH with a slow hydrogen-transfer step. By replacing Pro47 with His, the alcohol dehydrogenase activity is restored. Here, the structure of this P47H mutant was determined in complex with NADH to 2.5 A resolution. His47 is suitably positioned to act as a catalytic base in the deprotonation of the substrate. Moreover, in the more closed subunit, the coenzyme is allowed a position closer to the catalytic zinc. This is consistent with hydrogen transfer from an alcoholate intermediate where the Pro/His replacement focuses on the function of the enzyme.  相似文献   

4.
In contrast with other animal species, humans possess three distinct genes for class I alcohol dehydrogenase and show polymorphic variation in the ADH1B and ADH1C genes. The three class I alcohol dehydrogenase isoenzymes share approximately 93% sequence identity but differ in their substrate specificity and their developmental expression. We report here the first three-dimensional structures for the ADH1A and ADH1C*2 gene products at 2.5 and 2.0 A, respectively, and the structure of the ADH1B*1 gene product in a binary complex with cofactor at 2.2 A. Not surprisingly, the overall structure of each isoenzyme is highly similar to the others. However, the substitution of Gly for Arg at position 47 in the ADH1A isoenzyme promotes a greater extent of domain closure in the ADH1A isoenzyme, whereas substitution at position 271 may account for the lower turnover rate for the ADH1C*2 isoenzyme relative to its polymorphic variant, ADH1C*1. The substrate-binding pockets of each isoenzyme possess a unique topology that dictates each isoenzyme's distinct but overlapping substrate preferences. ADH1*B1 has the most restrictive substrate-binding site near the catalytic zinc atom, whereas both ADH1A and ADH1C*2 possess amino acid substitutions that correlate with their better efficiency for the oxidation of secondary alcohols. These structures describe the nature of their individual substrate-binding pockets and will improve our understanding of how the metabolism of beverage ethanol affects the normal metabolic processes performed by these isoenzymes.  相似文献   

5.
Pseudomonas aeruginosa alcohol dehydrogenase (PaADH; ADH, EC 1.1.1.1) catalyzes the reversible oxidation of primary and secondary alcohols to the corresponding aldehydes and ketones, using NAD as coenzyme. We crystallized the ternary complex of PaADH with its coenzyme and a substrate molecule and determined its structure at a resolution of 2.3 A, using the molecular replacement method. The PaADH tetramer comprises four identical chains of 342 amino acid residues each and obeys ~222-point symmetry. The PaADH monomer is structurally similar to alcohol dehydrogenase monomers from vertebrates, archaea, and bacteria. The stabilization of the ternary complex of PaADH, the coenzyme, and the poor substrate ethylene glycol (k(cat) = 4.5 sec(-1); Km > 200 mM) was due to the blocked exit of the coenzyme in the crystalline state, combined with a high (2.5 M) concentration of the substrate. The structure of the ternary complex presents the precise geometry of the Zn coordination complex, the proton-shuttling system, and the hydride transfer path. The ternary complex structure also suggests that the low efficiency of ethylene glycol as a substrate results from the presence of a second hydroxyl group in this molecule.  相似文献   

6.
The turnover numbers and other kinetic constants for human alcohol dehydrogenase (ADH) 4 ("stomach" isoenzyme) are substantially larger (10-100-fold) than those for human class I and horse liver alcohol dehydrogenases. Comparison of the primary amino acid sequences (69% identity) and tertiary structures of these enzymes led to the suggestion that residue 317, which makes a hydrogen bond with the nicotinamide amide nitrogen of the coenzyme, may account for these differences. Ala-317 in the class I enzymes is substituted with Cys in human ADH4, and locally different conformations of the peptide backbones could affect coenzyme binding. This hypothesis was tested by making the A317C substitution in horse liver ADH1E and comparisons to the wild-type ADH1E. The steady-state kinetic constants for the oxidation of benzyl alcohol and the reduction of benzaldehyde catalyzed by the A317C enzyme were very similar (up to about 2-fold differences) to those for the wild-type enzyme. Transient kinetics showed that the rate constants for binding of NAD(+) and NADH were also similar. Transient reaction data were fitted to the full Ordered Bi Bi mechanism and showed that the rate constants for hydride transfer decreased by about 2.8-fold with the A317C substitution. The structure of A317C ADH1E complexed with NAD(+) and 2,3,4,5,6-pentafluorobenzyl alcohol at 1.2 ? resolution is essentially identical to the structure of the wild-type enzyme, except near residue 317 where the additional sulfhydryl group displaces a water molecule that is present in the wild-type enzyme. ADH is adaptable and can tolerate internal substitutions, but the protein dynamics apparently are affected, as reflected in rates of hydride transfer. The A317C substitution is not solely responsible for the larger kinetic constants in human ADH4; thus, the differences in catalytic activity must arise from one or more of the other hundred substitutions in the enzyme.  相似文献   

7.
LeBrun LA  Park DH  Ramaswamy S  Plapp BV 《Biochemistry》2004,43(11):3014-3026
Histidine-51 in horse liver alcohol dehydrogenase (ADH) is part of a hydrogen-bonded system that appears to facilitate deprotonation of the hydroxyl group of water or alcohol ligated to the catalytic zinc. The contribution of His-51 to catalysis was studied by characterizing ADH with His-51 substituted with Gln (H51Q). The steady-state kinetic constants for ethanol oxidation and acetaldehyde reduction at pH 8 are similar for wild-type and H51Q enzymes. In contrast, the H51Q substitution significantly shifts the pH dependencies for steady-state and transient reactions and decreases by 11-fold the rate constant for the transient oxidation of ethanol at pH 8. Modest substrate deuterium isotope effects indicate that hydride transfer only partially limits the transient oxidation and turnover. Transient data show that the H51Q substitution significantly decreases the rate of isomerization of the enzyme-NAD(+) complex and becomes a limiting step for ethanol oxidation. Isomerization of the enzyme-NAD(+) complex is rate limiting for acetaldehyde reduction catalyzed by the wild-type enzyme, but release of alcohol is limiting for the H51Q enzyme. X-ray crystallography of doubly substituted His51Gln:Lys228Arg ADH complexed with NAD(+) and 2,3- or 2,4-difluorobenzyl alcohol shows that Gln-51 isosterically replaces histidine in interactions with the nicotinamide ribose of the coenzyme and that Arg-228 interacts with the adenosine monophosphate of the coenzyme without affecting the protein conformation. The difluorobenzyl alcohols bind in one conformation. His-51 participates in, but is not essential for, proton transfers in the mechanism.  相似文献   

8.
Rubach JK  Plapp BV 《Biochemistry》2003,42(10):2907-2915
Amino acid residues Thr-178, Val-203, and Val-292, which interact with the nicotinamide ring of the coenzyme bound to alcohol dehydrogenase (ADH), may facilitate hydride transfer and hydrogen tunneling by orientation and dynamic effects. The T178S, T178V, V203A, V292A, V292S, and V292T substitutions significantly alter the steady state and transient kinetics of the enzyme. The V292A, V292S, and V292T enzymes have decreased affinity for coenzyme (NAD+ by 30-50-fold and NADH by 35-75-fold) as compared to the wild-type enzyme. The substitutions in the nicotinamide binding site decrease the rate constant of hydride transfer for benzyl alcohol oxidation by 3-fold (for V292T ADH) to 16-fold (for V203A ADH). The modest effects suggest that catalysis does not depend critically on individual residues and that several residues in the nicotinamide binding site contribute to catalysis. The structures of the V292T ADH-NAD+-pyrazole and wild-type ADH-NAD+-4-iodopyrazole ternary complexes are very similar. Only subtle changes in the V292T enzyme cause the large changes in coenzyme binding and the small change in hydride transfer. In these complexes, one pyrazole nitrogen binds to the catalytic zinc, and the other nitrogen forms a partial covalent bond with C4 of the nicotinamide ring, which adopts a boat conformation that is postulated to be relevant for hydride transfer. The results provide an experimental basis for evaluating the contributions of dynamics to hydride transfer.  相似文献   

9.
Phylogenetic relationship and the rates of evolution of mammalian alcohol dehydrogenases (ADHs) have been studied by using the amino acid sequences from the human (ADH alpha, ADH beta, and ADH gamma), rat, mouse, and horse (ADH E and ADH S). With the maize ADH1 and ADH2 used as references, the patterns of the amino acid replacements in the beta-sheets, alpha-helices, and random coils in each of the catalytic and coenzyme-binding domains were analyzed separately. The phylogenetic trees based on the different sets of amino acid substitutions consistently showed that (1) multiple ADHs in human and horse have arisen after mammalian radiation, (2) the common ancestor of human ADHs alpha and beta diverged from the ancestor of ADH gamma first and the former two ADHs diverged from each other more recently, and (3) the human ADHs are more closely related to the rodent ADHs than to the horse ADHs. Furthermore, the estimated branch lengths showed that the rodent ADHs are evolving faster than the other ADHs. This difference in evolutionary rate between the two groups of organisms is explainable either in terms of the difference in the number of cell generations per year or in terms of reduction of functional constraints.  相似文献   

10.
1. Starch gel electrophoresis of adult shrew (Suncus murinus) liver extracts revealed five forms of alcohol dehydrogenase (ADH 1-5) and four of them were purified. 2. ADH-4 and ADH-5 resemble human class I ADH in terms of electrophoretic mobility, substrate specificity and sensitivity to pyrazole inhibition. 3. ADH-2 does not belong to any of the three classes of human ADHs but rather with catalytic properties similar to those of the class B ADH found in guinea pig liver. 4. ADH-1 prefers secondary alcohol over primary alcohol substrates and between the enantiomers tested, the enzyme favors the S isomers.  相似文献   

11.
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   

12.
Studies in knockout mice support the involvement of alcohol dehydrogenases ADH1 and ADH4 in retinoid metabolism, although kinetics with retinoids are not known for the mouse enzymes. Moreover, a role of alcohol dehydrogenase (ADH) in the eye retinoid interconversions cannot be ascertained due to the lack of information on the kinetics with 11-cis-retinoids. We report here the kinetics of human ADH1B1, ADH1B2, ADH4, and mouse ADH1 and ADH4 with all-trans-, 7-cis-, 9-cis-, 11-cis- and 13-cis-isomers of retinol and retinal. These retinoids are substrates for all enzymes tested, except the 13-cis isomers which are not used by ADH1. In general, human and mouse ADH4 exhibit similar activity, higher than that of ADH1, while mouse ADH1 is more efficient than the homologous human enzymes. All tested ADHs use 11-cis-retinoids efficiently. ADH4 shows much higher k(cat)/K(m) values for 11-cis-retinol oxidation than for 11-cis-retinal reduction, a unique property among mammalian ADHs for any alcohol/aldehyde substrate pair. Docking simulations and the kinetic properties of the human ADH4 M141L mutant demonstrated that residue 141, in the middle region of the active site, is essential for such ADH4 specificity. The distinct kinetics of ADH4 with 11-cis-retinol, its wide specificity with retinol isomers and its immunolocalization in several retinal cell layers, including pigment epithelium, support a role of this enzyme in the various retinol oxidations that occur in the retina. Cytosolic ADH4 activity may complement the isomer-specific microsomal enzymes involved in photopigment regeneration and retinoic acid synthesis.  相似文献   

13.
The quinone‐dependent alcohol dehydrogenase (PQQ‐ADH, E.C. 1.1.5.2) from the Gram‐negative bacterium Pseudogluconobacter saccharoketogenes IFO 14464 oxidizes primary alcohols (e.g. ethanol, butanol), secondary alcohols (monosaccharides), as well as aldehydes, polysaccharides, and cyclodextrins. The recombinant protein, expressed in Pichia pastoris, was crystallized, and three‐dimensional (3D) structures of the native form, with PQQ and a Ca2+ ion, and of the enzyme in complex with a Zn2+ ion and a bound substrate mimic were determined at 1.72 Å and 1.84 Å resolution, respectively. PQQ‐ADH displays an eight‐bladed β‐propeller fold, characteristic of Type I quinone‐dependent methanol dehydrogenases. However, three of the four ligands of the Ca2+ ion differ from those of related dehydrogenases and they come from different parts of the polypeptide chain. These differences result in a more open, easily accessible active site, which explains why PQQ‐ADH can oxidize a broad range of substrates. The bound substrate mimic suggests Asp333 as the catalytic base. Remarkably, no vicinal disulfide bridge is present near the PQQ, which in other PQQ‐dependent alcohol dehydrogenases has been proposed to be necessary for electron transfer. Instead an associated cytochrome c can approach the PQQ for direct electron transfer.  相似文献   

14.
The structural framework of cod liver alcohol dehydrogenase is similar to that of horse and human alcohol dehydrogenases. In contrast, the substrate pocket differs significantly, and main differences are located in three loops. Nevertheless, the substrate pocket is hydrophobic like that of the mammalian class I enzymes and has a similar topography in spite of many main-chain and side-chain differences. The structural framework of alcohol dehydrogenase is also present in a number of related enzymes like glucose dehydrogenase and quinone oxidoreductase. These enzymes have completely different substrate specificity, but also for these enzymes, the corresponding loops of the substrate pocket have significantly different structures. The domains of the two subunits in the crystals of the cod enzyme further differ by a rotation of the catalytic domains by about 6 degrees. In one subunit, they close around the coenzyme similarly as in coenzyme complexes of the horse enzyme, but form a more open cleft in the other subunit, similar to the situation in coenzyme-free structures of the horse enzyme. The proton relay system differs from the mammalian class I alcohol dehydrogenases. His 51, which has been implicated in mammalian enzymes to be important for proton transfer from the buried active site to the surface is not present in the cod enzyme. A tyrosine in the corresponding position is turned into the substrate pocket and a water molecule occupies the same position in space as the His side chain, forming a shorter proton relay system.  相似文献   

15.
Rubach JK  Plapp BV 《Biochemistry》2002,41(52):15770-15779
The relationship between substrate mobility and catalysis was studied with wild-type and Phe93Ala (F93A) horse liver alcohol dehydrogenase (ADH). Wild-type ADH binds 2,3,4,5,6-pentafluorobenzyl alcohol in one position as shown by X-ray results, and (19)F NMR shows five resonances for the fluorines of the bound alcohol. The two meta-fluorines exchange positions with a rate constant of about 4 s(-1), indicating that mobility (ring flipping) of the benzyl alcohol is relatively restricted. The wild-type enzyme binds 2,3-difluorobenzyl alcohol in two alternative conformations that are related by a ring flip and a small translation of the fluorinated benzene ring, and the (19)F NMR spectrum shows three resonances for the two bound fluorines, consistent with the two orientations. Phe-93 interacts with the bound benzyl alcohols, and the F93A substitution decreases the rate constants for hydride transfer for benzyl alcohol oxidation and benzaldehyde reduction by 7.4- and 130-fold, respectively. The structure of F93A ADH crystallized with NAD(+) and 2,3,4,5,6-pentafluorobenzyl alcohol is similar to the structure of the wild-type enzyme complex except that the pentafluorobenzyl alcohol is not found in one position. The (19)F NMR spectrum of the F93A ADH-NAD(+)-pentafluorobenzyl alcohol complex shows three resonances for the bound fluorines. Line shape analysis of the spectrum suggests the bound pentafluorobenzyl ring undergoes rapid ring-flipping at about 20 000 s(-1). The F93A substitution greatly increases the mobility of the benzyl alcohol but modestly and differentially decreases the probability that the substrate is preorganized for hydride transfer.  相似文献   

16.
A mutant of the thermostable NAD+-dependent (S)-stereospecific alcohol dehydrogenase from Sulfolobus solfataricus (SsADH) which has a single substitution, Trp95Leu, located at the substrate binding pocket, was fully characterized to ascertain the role of Trp95 in discriminating between chiral secondary alcohols suggested by the wild-type SsADH crystallographic structure. The Trp95Leu mutant displays no apparent activity with short-chain primary and secondary alcohols and poor activity with aromatic substrates and coenzyme. Moreover, the Trp → Leu substitution affects the structural stability of the archaeal ADH, decreasing its thermal stability without relevant changes in secondary structure. The double mutant Trp95Leu/Asn249Tyr was also purified to assist in crystallographic analysis. This mutant exhibits higher activity but decreased affinity toward aliphatic alcohols, aldehydes as well as NAD+ and NADH compared to the wild-type enzyme. The crystal structure of the Trp95Leu/Asn249Tyr mutant apo form, determined at 2.0 Å resolution, reveals a large local rearrangement of the substrate site with dramatic consequences. The Leu95 side-chain conformation points away from the catalytic metal center and the widening of the substrate site is partially counteracted by a concomitant change of Trp117 side chain conformation. Structural changes at the active site are consistent with the reduced activity on substrates and decreased coenzyme binding.  相似文献   

17.
We report a high-throughput phage selection method to identify mutants of Sfp phosphopantetheinyl transferase with altered substrate specificities from a large library of the Sfp enzyme. In this method, Sfp and its peptide substrates are co-displayed on the M13 phage surface as fusions to the phage capsid protein pIII. Phage-displayed Sfp mutants that are active with biotin-conjugated coenzyme A (CoA) analogues would covalently transfer biotin to the peptide substrates anchored on the same phage particle. Affinity selection for biotin-labeled phages would enrich Sfp mutants that recognize CoA analogues for carrier protein modification. We used this method to successfully change the substrate specificity of Sfp and identified mutant enzymes with more than 300-fold increase in catalytic efficiency with 3′-dephospho CoA as the substrate. The method we developed in this study provides a useful platform to display enzymes and their peptide substrates on the phage surface and directly couples phage selection with enzyme catalysis. We envision this method to be applied to engineering the catalytic activities of other protein posttranslational modification enzymes.  相似文献   

18.
Among strains of Drosophila melanogaster each derived from a single fertilized female taken from natural populations, there is variation in both alcohol dehydrogenase (ADH) activity and the amount of ADH protein. The correlation between ADH activity and number of molecules over all strains examined is 0.87 or 0.96 in late third instar larvae depending on whether the substrate is 2-propanol or ethanol. With respect to the two common electrophoretic allozymic forms, F and S, segregating in these populations, the FF strains on the whole have higher ADH activities and numbers of ADH molecules than the SS strains. Over all strains examined, enzyme extracts from FF strains have a mean catalytic efficiency per enzyme molecule higher than that of enzyme extracts from SS strains when ethanol is the substrate, and much higher when 2-propanol is the substrate. One FF strain had an ADH activity/ADH protein ratio characteristic of SS strains.  相似文献   

19.
Mammalian alcohol dehydrogenase (ADH) constitutes a complex system with different forms and extensive multiplicity (ADH1–ADH6) that catalyze the oxidation and reduction of a wide variety of alcohols and aldehydes. The ADH1 enzymes, the classical liver forms, are involved in several metabolic pathways beside the oxidation of ethanol, e.g. norepinephrine, dopamine, serotonin and bile acid metabolism. This class is also able to further oxidize aldehydes into the corresponding carboxylic acids, i.e. dismutation. ADH2, can be divided into two subgroups, one group consisting of the human enzyme together with a rabbit form and another consisting of the rodent forms. The rodent enzymes almost lack ethanol-oxidizing capacity in contrast to the human form, indicating that rodents are poor model systems for human ethanol metabolism. ADH3 (identical to glutathione-dependent formaldehyde dehydrogenase) is clearly the ancestral ADH form and S-hydroxymethylglutathione is the main physiological substrate, but the enzyme can still oxidize ethanol at high concentrations. ADH4 is solely extrahepatically expressed and is probably involved in first pass metabolism of ethanol beside its role in retinol metabolism. The higher classes, ADH5 and ADH6, have been poorly investigated and their substrate repertoire is unknown. The entire ADH system can be seen as a general detoxifying system for alcohols and aldehydes without generating toxic radicals in contrast to the cytochrome P450 system.  相似文献   

20.
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   

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