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 omega-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 pK(a) of the bound alcohol substrate as well as the affinity for coenzyme.  相似文献   

2.
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   

3.
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   

4.
The 1.7A resolution crystal structure of recombinant family G/11 beta-1,4-xylanase (rXynA) from Bacillus subtilis 1A1 shows a jellyroll fold in which two curved beta-sheets form the active-site and substrate-binding cleft. The onset of thermal denaturation of rXynA occurs at 328 K, in excellent agreement with the optimum catalytic temperature. Molecular dynamics simulations at temperatures of 298-328 K demonstrate that below the optimum temperature the thumb loop and palm domain adopt a closed conformation. However, at 328 K these two domains separate facilitating substrate access to the active-site pocket, thereby accounting for the optimum catalytic temperature of the rXynA.  相似文献   

5.
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   

6.
The crystal structure analysis of horse liver alcohol dehydrogenase has been extended to 2.4 Å resolution. From the corresponding electron density map of the apoenzyme we have determined the positions of the 374 amino acids in the polypeptide chain of each subunit.The coenzyme binding domain of the subunit comprises residues 176 to 318. 45% of these residues are helical and 32% are in the central six-stranded pleated sheet structure. The positions and orientations of the helices with respect to the pleated sheet indicate a possible folding mechanism for this part of the subunit structure. The coenzyme analogue ADP-ribose binds to this domain in a position and orientation very similar to coenzyme binding to lactate dehydrogenase. The adenine part binds in a hydrophobic pocket, the adenosine ribose is hydrogen-bonded to the side chain of Asp223, the pyrophosphate is positioned by interaction with Arg47 and the nicotinamide ribose is 6Å away from the catalytic zinc atom.The catalytic domain is mainly built up from three distinct antiparallel pleated-sheet regions. Residues within this domain provide ligands to the catalytic zinc atom; Cys46, His67 and Cys174. An approximate tetrahedral coordination of this zinc is completed by a water molecule or hydroxyl ion depending on the pH. Residues 95 to 113 form a lobe that binds the second zinc atom of the subunit. This zinc is liganded in a distorted tetrahedral arrangement by four sulphur atoms from the cysteine residues 97, 100, 103 and 111. The lobe forms one side of a significant cleft in the enzyme surface suggesting that this region might constitute a second catalytic centre of unknown function.The two domains of the subunit are separated by a crevice that contains a wide and deep hydrophobic pocket. The catalytic zinc atom is at the bottom of this pocket, with the zinc-bound water molecule projecting out into the pocket. This water molecule is hydrogen-bonded to the side chain of Ser48 which in turn is hydrogen-bonded to His51. The pocket which in all probability is the binding site for the substrate and the nicotinamide moiety of the coenzyme, is lined almost exclusively with hydrophobic side chains. Both subunits contribute residues to each of the two substrate binding pockets of the molecule. The only accessible polar groups in the vicinity of the catalytic centre are Ser48 and Thr178 apart from zinc and the zinc-bound water molecule.  相似文献   

7.
Sanghani PC  Bosron WF  Hurley TD 《Biochemistry》2002,41(51):15189-15194
Human glutathione-dependent formaldehyde dehydrogenase plays an important role in the metabolism of glutathione adducts such as S-(hydroxymethyl)glutathione and S-nitrosoglutathione. The role of specific active site residues in binding these physiologically important substrates and the structural changes during the catalytic cycle of glutathione-dependent formaldehyde dehydrogenase was examined by determining the crystal structure of a ternary complex with S-(hydroxymethyl)glutathione and the reduced coenzyme to 2.6 A resolution. The formation of the ternary complex caused the movement of the catalytic domain toward the coenzyme-binding domain. This represents the first observation of domain closure in glutathione-dependent formaldehyde dehydrogenase in response to substrate binding. A water molecule adjacent to the 2'-ribose hydroxyl of NADH suggests that the alcohol proton is relayed to solvent directly from the coenzyme, rather than through the action of the terminal histidine residue as observed in the proton relay system for class I alcohol dehydrogenases. S-(Hydroxymethyl)glutathione is directly coordinated to the active site zinc and forms interactions with the highly conserved residues Arg114, Asp55, Glu57, and Thr46. The active site zinc has a tetrahedral coordination environment with Cys44, His66, and Cys173 as the three protein ligands in addition to S-(hydroxymethyl)glutathione. This is in contrast to zinc coordination in the binary coenzyme complex where all of the ligands were contributed by the enzyme and included Glu67 as the fourth protein ligand. This change in zinc coordination is accomplished by an approximately 2.3 A movement of the catalytic zinc.  相似文献   

8.
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   

9.
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   

10.
L A LeBrun  B V Plapp 《Biochemistry》1999,38(38):12387-12393
The rate of association of NAD(+) with wild-type horse liver alcohol dehydrogenase (ADH) is maximal at pH values between pK values of about 7 and 9, and the rate of NADH association is maximal at a pH below a pK of 9. The catalytic zinc-bound water, His-51 (which interacts with the 2'- and 3'-hydroxyl groups of the nicotinamide ribose of the coenzyme in the proton relay system), and Lys-228 (which interacts with the adenosine 3'-hydroxyl group and the pyrophosphate of the coenzyme) may be responsible for the observed pK values. In this study, the Lys228Arg, His51Gln, and Lys228Arg/His51Gln (to isolate the effect of the catalytic zinc-bound water) mutations were used to test the roles of the residues in coenzyme binding. The steady state kinetic constants at pH 8 for the His51Gln enzyme are similar to those for wild-type ADH. The Lys228Arg and Lys228Arg/His51Gln substitutions decrease the affinity for the coenzymes up to 16-fold, probably due to altered interactions with the arginine at position 228. As determined by transient kinetics, the rate constant for association of NAD(+) with the mutated enzymes no longer decreases at high pH. The pH profile for the Lys228Arg enzyme retains the pK value near 7. The His51Gln and Lys228Arg/His51Gln substitutions significantly decrease the rate constants for NAD(+) association, and the pH dependencies show that these enzymes bind NAD(+) most rapidly at a pH above pK values of 8. 0 and 9.0, respectively. It appears that the pK of 7 in the wild-type enzyme is shifted up by the H51Q substitutions, and the resulting pH dependence is due to the deprotonation of the catalytic zinc-bound water. Kinetic simulations suggest that isomerization of the enzyme-NAD(+) complex is substantially altered by the mutations. In contrast, the pH dependencies for NADH association with His51Gln, Lys228Arg, and Lys228Arg/His51Gln enzymes were the same as for wild-type ADH, suggesting that the binding of NAD(+) and the binding of NADH are controlled differently.  相似文献   

11.
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   

12.
L-lactate dehydrogenase (LDH) from Bacillus stearothermophilus is a redox enzyme which has a strong preference for NADH over NADPH as coenzyme. To exclude NADPH from the coenzyme-binding pocket, LDH contains a conserved aspartate residue at position 52. However, this residue is probably not solely responsible for the NADH specificity. In this report we examine the possibilities of altering the coenzyme specificity of LDH by introducing a range of different point mutations in the coenzyme-binding domain. Furthermore, after choosing the mutant with the highest selectivity for NADPH, we also investigated the possibility of further altering the coenzyme specificity by adding an organic solvent to the reaction mixture. The LDH mutant, I51K:D52S, exhibited a 56-fold increased specificity to NADPH over the wild-type LDH in a reaction mixture containing 15% methanol. Furthermore, the NADPH turnover number of this mutant was increased almost fourfold as compared with wild-type LDH. To explain the altered coenzyme specificity exhibited by the D52SI51K double mutant, molecular dynamics simulations were performed.  相似文献   

13.
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   

14.
A novel short-chain (S)-1-phenyl-1,2-ethanediol dehydrogenase (SCR) from Candida parapsilosis exhibits coenzyme specificity for NADPH over NADH. It catalyzes an anti-Prelog type reaction to reduce 2-hydroxyacetophenone into (S)-1-phenyl-1,2-ethanediol. The coding gene was overexpressed in Escherichia coli and the purified protein was crystallized. The crystal structure of the apo-form was solved to 2.7 Å resolution. This protein forms a homo-tetramer with a broken 2-2-2 symmetry. The overall fold of each SCR subunit is similar to that of the known structures of other homologous alcohol dehydrogenases, although the latter usually form tetramers with perfect 2-2-2 symmetries. Additionally, in the apo-SCR structure, the entrance of the NADPH pocket is blocked by a surface loop. In order to understand the structure–function relationship of SCR, we carried out a number of mutagenesis–enzymatic analyses based on the new structural information. First, mutations of the putative catalytic Ser-Tyr-Lys triad confirmed their functional role. Second, truncation of an N-terminal 31-residue peptide indicated its role in oligomerization, but not in catalytic activity. Similarly, a V270D point mutation rendered the SCR as a dimer, rather than a tetramer, without affecting the enzymatic activity. Moreover, the S67D/H68D double-point mutation inside the coenzyme-binding pocket resulted in a nearly 10-fold increase and a 20-fold decrease in the kcat/KM value when NADH and NADPH were used as cofactors, respectively, with kcat remaining essentially the same. This latter result provides a new example of a protein engineering approach to modify the coenzyme specificity in SCR and short-chain dehydrogenases/reductases in general.  相似文献   

15.
The structure of a triclinic complex between liver alcohol dehydrogenase, reduced coenzyme NADH, and the inhibitor dimethylsulfoxide has been determined to 2.9 Å resolution using isomorphous replacement methods. The heavy-atom positions were derived by molecular replacement methods using phase angles derived from a model of the orthorhombic apoenzyme structure previously determined to 2.4 Å resolution. A model of the present holoenzyme molecule was built on a Vector General 3400 display system using the RING system of programs. This model gave a crystallographic R-value of 37.9%.There are extensive conformational differences between the protein molecules in the two forms. The conformational change involves a rotation of 7.5 ° of the catalytic domains relative to the coenzyme binding domains. A hinge region for this rotation is defined within a hydrophobic core between two helices. The internal structures of the domains are preserved with the exception of a movement of a small loop in the coenzyme binding domain. A cleft between the domains is closed by this coenzyme-induced conformational change, making the active site less accessible from solution and thus more hydrophobic.The two crystallographically independent subunits are very similar and bind both coenzyme and inhibitor in an identical way within the present limits of error. The coenzyme molecule is bound in an extended conformation with the two ends in hydrophobic crevices on opposite sides of the central pleated sheet of the coenzyme binding domain. There are hydrogen bonds to oxygen atoms of the ribose moities from Asp223, Lys228 and His51. The pyrophosphate group is in contact with the side-chains of Arg47 and Arg369.No new residues are brought into the active site compared to the apoenzyme structure. The active site zinc atom is close to the hinge region, where the smallest structural changes occur. Small differences in the co-ordination geometry of the ligands Cys46, His67 and Cysl74 are not excluded and may account for the ordered mechanism. The oxygen atom of the inhibitor dimethylsulfoxide is bound directly to zinc confirming the structural basis for the suggested mechanism of action based on studies of the apoenzyme structure.  相似文献   

16.
The structure of the recombinant medium chain alcohol dehydrogenase (ADH) from the hyperthermophilic archaeon Aeropyrum pernix has been solved by the multiple anomalous dispersion technique using the signal from the naturally occurring zinc ions. The enzyme is a tetramer with 222 point group symmetry. The ADH monomer is formed from a catalytic and a cofactor-binding domain, with the overall fold similar to previously solved ADH structures. The 1.62 A resolution A.pernix ADH structure is that of the holo form, with the cofactor NADH bound into the cleft between the two domains. The electron density found in the active site has been interpreted to be octanoic acid, which has been shown to be an inhibitor of the enzyme. This inhibitor is positioned with its carbonyl oxygen atom forming the fourth ligand of the catalytic zinc ion. The structural zinc ion of each monomer is present at only partial occupancy and in its absence a disulfide bond is formed. The enhanced thermal stability of the A.pernix ADH is thought to arise primarily from increased ionic and hydrophobic interactions on the subunit interfaces.  相似文献   

17.
W F Bosron  S J Yin  F E Dwulet  T K Li 《Biochemistry》1986,25(8):1876-1881
The beta 1 beta 1 and beta 2 beta 2 human liver alcohol dehydrogenase isoenzymes differ by only one residue at the coenzyme-binding site; Arg-47 in beta 1 is replaced by His in the beta 2 subunit. Since Arg-47 is thought to facilitate the carboxymethylation of Cys-46 in horse liver alcohol dehydrogenase by binding halo acids in a Michaelis-Menten complex prior to inactivation, the specificity and kinetics of modification of the two human liver beta beta isoenzymes with iodoacetate were compared. Both of the beta beta isoenzymes were inactivated by treatment with iodo[14C]acetate, and one Cys per subunit was carboxymethylated. Cys-174, which is a ligand to the active-site zinc atom in horse liver alcohol dehydrogenase, was selectively carboxymethylated in each of the human beta beta isoenzymes; less than 15% of the iodo[14C]acetate incorporated into the enzyme appeared in Cys-46. Therefore, the three-dimensional structure of the basic amino acids in the anion-binding site of the human beta beta isoenzymes appears to be different from that of horse liver alcohol dehydrogenase. The kinetics of alkylation are consistent with the formation of a Michaelis-Menten complex before inactivation of the isoenzymes. The average Ki values for iodoacetate were 10 and 16 mM for beta 1 beta 1 and beta 2 beta 2, respectively, and maximal rate constants for inactivation were 0.22 and 0.17 min-1, respectively. From these data, it can be concluded that there is a relatively minor effect of the substitution of His for Arg at position 47 on the kinetics of inactivation.  相似文献   

18.
Bacterial 3alpha-hydroxysteroid dehydrogenase, which belongs to a short-chain dehydrogenase/reductase family and forms a dimer composed of two 26-kDa subunits, catalyzes the oxidoreduction of hydroxysteroids in a coenzyme-dependent manner. This enzyme also catalyzes the oxidoreduction of nonsteroid compounds that play an important role in xenobiotic metabolism of bacteria. We performed an x-ray analysis on the crystal of Ps3alphaHSD, the enzyme from Pseudomonas sp. B-0831 complexed with NADH. The resulting crystal structure at 1.8A resolution showed that Ps3alphaHSD exists as a structural heterodimer composed of apo- and holo-subunits. A distinct structural difference between them was found in the 185-207-amino acid region, where the structure in the apo-subunit is disordered whereas that in the holo-subunit consists of two alpha-helices. This fact proved that the NADH binding allows the helical structures to form the substrate binding pocket even in the absence of the substrate, although the region corresponds to the so-called "substrate-binding loop." The induction of alpha-helices in solution by the coenzyme binding was also confirmed by the CD experiment. In addition, the CD experiment revealed that the helix-inducing ability of NADH is stronger than that of NAD. We discuss the negative cooperativity for the coenzyme binding, which is caused by the effect of the structural change transferred between the subunits of the heterodimer.  相似文献   

19.
l-Gulonate 3-dehydrogenase (GDH) is a bifunctional dimeric protein that functions not only as an NAD+-dependent enzyme in the uronate cycle but also as a taxon-specific λ-crystallin in rabbit lens. Here we report the first crystal structure of GDH in both apo form and NADH-bound holo form. The GDH protomer consists of two structural domains: the N-terminal domain with a Rossmann fold and the C-terminal domain with a novel helical fold. In the N-terminal domain of the NADH-bound structure, we identified 11 coenzyme-binding residues and found 2 distinct side-chain conformers of Ser124, which is a putative coenzyme/substrate-binding residue. A structural comparison between apo form and holo form and a mutagenesis study with E97Q mutant suggest an induced-fit mechanism upon coenzyme binding; coenzyme binding induces a conformational change in the coenzyme-binding residues Glu97 and Ser124 to switch their activation state from resting to active, which is required for the subsequent substrate recruitment. Subunit dimerization is mediated by numerous intersubunit interactions, including 22 hydrogen bonds and 104 residue pairs of van der Waals interactions, of which those between two cognate C-terminal domains are predominant. From a structure/sequence comparison within GDH homologues, a much greater degree of interprotomer interactions (both polar and hydrophobic) in the rabbit GDH would contribute to its higher thermostability, which may be relevant to the other function of this enzyme as λ-crystallin, a constitutive structural protein in rabbit lens. The present crystal structures and amino acid mutagenesis studies assigned the role of active-site residues: catalytic base for His145 and substrate binding for Ser124, Cys125, Asn196, and Arg231. Notably, Arg231 participates in substrate binding from the other subunit of the GDH dimer, indicating the functional significance of the dimeric state. Proper orientation of the substrate-binding residues for catalysis is likely to be maintained by an interprotomer hydrogen-bonding network of residues Asn196, Gln199, and Arg231, suggesting a network-based substrate recognition of GDH.  相似文献   

20.
Structural comparisons of sorbitol dehydrogenase with zinc-containing 'long' alcohol dehydrogenases reveal distant but clear relationships. An alignment suggests 93 positional identities with horse liver alcohol dehydrogenase (25% of 374 positions) and 73 identities with yeast alcohol dehydrogenase (20%). Sorbitol dehydrogenase forms a link between these distantly related alcohol dehydrogenases and is in some regions more similar to one of them that they are to each other. 43 residues (11%) are common to all three enzymes and include a heavy over-representation of glycine (half of all glycine residues in sorbitol dehydrogenase), showing the importance of space restrictions in protein structures. Four regions are well conserved, two in each domain of horse liver alcohol dehydrogenase. They are two segments close to the active-site zinc atom of the catalytic domain, and two in the central beta-pleated sheet strands of the coenzyme-binding domain. These similarities demonstrate the general importance of internal and central building units in proteins. Large variations affect a region adjacent to the third protein ligand to the active-site zinc atom in horse liver alcohol dehydrogenase. Such changes at active sites of related enzymes are unusual. Other large differences concern the segment around the non-catalytic zinc atom of horse liver alcohol dehydrogenase; three of its four cysteine ligands are absent from sorbitol dehydrogenase. Three segments with several exchanges correspond to a continuous region with superficial areas, inter-domain contacts and inter-subunit interactions in the catalytic domain of alcohol dehydrogenase. They may correlate with the altered quaternary structure of sorbitol dehydrogenase. Regions corresponding to top and bottom beta-strands in the coenzyme-binding domain of the alcohol dehydrogenase are also little conserved. Within sorbitol dehydrogenase, a large segment shows an internal similarity. The two distantly related alcohol dehydrogenases and sorbitol dehydrogenase form a triplet of enzymes illustrating basic protein relationships. They are ancestrally close enough to establish similarities, yet sufficiently divergent to illustrate changes in all but fundamental properties.  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