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 crystal structure of the mitochondrial NAD-malic enzyme from Ascaris suum, in a quaternary complex with NADH, tartronate, and magnesium has been determined to 2.0-A resolution. The structure closely resembles the previously determined structure of the same enzyme in binary complex with NAD. However, a significant difference is observed within the coenzyme-binding pocket of the active site with the nicotinamide ring of NADH molecule rotating by 198 degrees over the C-1-N-1 bond into the active site without causing significant movement of the other catalytic residues. The implications of this conformational change in the nicotinamide ring to the catalytic mechanism are discussed. The structure also reveals a binding pocket for the divalent metal ion in the active site and a binding site for tartronate located in a highly positively charged environment within the subunit interface that is distinct from the active site. The tartronate binding site, presumably an allosteric site for the activator fumarate, shows striking similarities and differences with the activator site of the human NAD-malic enzyme that has been reported recently. Thus, the structure provides additional insights into the catalytic as well as the allosteric mechanisms of the enzyme.  相似文献   

2.
Glyceraldehyde 3-phosphate dehydrogenase is a tetramer of four chemically identical subunits which requires the cofactor nicotinamide adenine dinucleotide (NAD) for activity. The structure of the holo-enzyme from Bacillus stearothermophilus has recently been refined using X-ray data to 2.4 A resolution. This has facilitated the structure determination of both the apo-enzyme and the enzyme with one molecule of NAD bound to the tetramer. These structures have been refined at 4 A resolution using the constrained-restrained parameter structure factor least-squares refinement program CORELS. When combined with individual atomic temperature factors from the holo-enzyme, these refined models give crystallographic R factors of 30.2% and 30.4%, respectively, for data to 3 A resolution. The apo-enzyme has 222 molecular symmetry, and the subunit structure is related to that of the holo-enzyme by an approximate rigid-body rotation of the coenzyme binding domain by 4.3 degrees with respect to the catalytic domains, which form the core of the tetramer. The effect of this rotation is to shield the coenzyme and active site from solvent in the holo-enzyme. In addition to the rigid-body rotation, there is a rearrangement of several residues involved in NAD binding. The structure of the 1 NAD enzyme is asymmetric. The subunit which contains the bound NAD adopts a conformation very similar to that of a holo-enzyme subunit, while the other three unliganded subunits are very similar to the apo-enzyme conformation. This result provides unambiguous evidence for ligand-induced sequential conformational changes in B. stearothermophilus glyceraldehyde 3-phosphate dehydrogenase.  相似文献   

3.
Crystal structure of S-adenosylhomocysteine hydrolase from rat liver.   总被引:5,自引:0,他引:5  
The crystal structure of rat liver S-adenosyl-L-homocysteine hydrolase (AdoHcyase, EC 3.3.1.1) which catalyzes the reversible hydrolysis of S-adenosylhomocysteine (AdoHcy) has been determined at 2.8 A resolution. AdoHcyase from rat liver is a tetrameric enzyme with 431 amino acid residues in each identical subunit. The subunit is composed of the catalytic domain, the NAD+-binding domain, and the small C-terminal domain. Both catalytic and NAD+-binding domains are folded into an ellipsoid with a typical alpha/beta twisted open sheet structure. The C-terminal section is far from the main body of the subunit and extends into the opposite subunit. An NAD+ molecule binds to the consensus NAD+-binding cleft of the NAD+-binding domain. The peptide folding pattern of the catalytic domain is quite similar to the patterns observed in many methyltransferases. Although the crystal structure does not contain AdoHcy or its analogue, there is a well-formed AdoHcy-binding crevice in the catalytic domain. Without introducing any major structural changes, an AdoHcy molecule can be placed in the catalytic domain. In the structure described here, the catalytic and NAD+-binding domains are quite far apart from each other. Thus, the enzyme appears to have an "open" conformation in the absence of substrate. It is likely that binding of AdoHcy induces a large conformational change so as to place the ribose moiety of AdoHcy in close proximity to the nicotinamide moiety of NAD+. A catalytic mechanism of AdoHcyase has been proposed on the basis of this crystal structure. Glu155 acts as a proton acceptor from the O3'-H when the proton of C3'-H is abstracted by NAD+. His54 or Asp130 acts as a general acid-base catalyst, while Cys194 modulates the oxidation state of the bound NAD+. The polypeptide folding pattern of the catalytic domain suggests that AdoHcy molecules can travel freely to and from AdoHcyase and methyltransferases to properly regulate methyltransferase activities. We believe that the crystal structure described here can provide insight into the molecular architecture of this important regulatory enzyme.  相似文献   

4.
Glutamate dehydrogenase is found in all organisms and catalyses the oxidative deamination of l-glutamate to 2-oxoglutarate. However, only animal GDH utilizes both NAD(H) or NADP(H) with comparable efficacy and exhibits a complex pattern of allosteric inhibition by a wide variety of small molecules. The major allosteric inhibitors are GTP and NADH and the two main allosteric activators are ADP and NAD(+). The structures presented here have refined and modified the previous structural model of allosteric regulation inferred from the original boGDH.NADH.GLU.GTP complex. The boGDH.NAD(+).alpha-KG complex structure clearly demonstrates that the second coenzyme-binding site lies directly under the "pivot helix" of the NAD(+) binding domain. In this complex, phosphates are observed to occupy the inhibitory GTP site and may be responsible for the previously observed structural stabilization by polyanions. The boGDH.NADPH.GLU.GTP complex shows the location of the additional phosphate on the active site coenzyme molecule and the GTP molecule bound to the GTP inhibitory site. As expected, since NADPH does not bind well to the second coenzyme site, no evidence of a bound molecule is observed at the second coenzyme site under the pivot helix. Therefore, these results suggest that the inhibitory GTP site is as previously identified. However, ADP, NAD(+), and NADH all bind under the pivot helix, but a second GTP molecule does not. Kinetic analysis of a hyperinsulinism/hyperammonemia mutant strongly suggests that ATP can inhibit the reaction by binding to the GTP site. Finally, the fact that NADH, NAD(+), and ADP all bind to the same site requires a re-analysis of the previous models for NADH inhibition.  相似文献   

5.
MAP kinases (MAPKs), which control mitogenic signal transduction in all eukaryotic organisms, are inactivated by dual specificity MAPK phosphatases (MKPs). MKP-3, a prototypical MKP, achieves substrate specificity through its N-terminal domain binding to the MAPK ERK2, resulting in the activation of its C-terminal phosphatase domain. The solution structure and biochemical analysis of the ERK2 binding (EB) domain of MKP-3 show that regions that are essential for ERK2 binding partly overlap with its sites that interact with the C-terminal catalytic domain, and that these interactions are functionally coupled to the active site residues of MKP-3. Our findings suggest a novel mechanism by which the EB domain binding to ERK2 is transduced to cause a conformational change of the C-terminal catalytic domain, resulting in the enzymatic activation of MKP-3.  相似文献   

6.
Exotoxin A of Pseudomonas aeruginosa asserts its cellular toxicity through ADP-ribosylation of translation elongation factor 2, predicated on binding to specific cell surface receptors and intracellular trafficking via a complex pathway that ultimately results in translocation of an enzymatic activity into the cytoplasm. In early work, the crystallographic structure of exotoxin A was determined to 3.0 A resolution, revealing a tertiary fold having three distinct structural domains; subsequent work has shown that the domains are individually responsible for the receptor binding (domain I), transmembrane targeting (domain II), and ADP-ribosyl transferase (domain III) activities, respectively. Here, we report the structures of wild-type and W281A mutant toxin proteins at pH 8.0, refined with data to 1.62 A and 1.45 A resolution, respectively. The refined models clarify several ionic interactions within structural domains I and II that may modulate an obligatory conformational change that is induced by low pH. Proteolytic cleavage by furin is also obligatory for toxicity; the W281A mutant protein is substantially more susceptible to cleavage than the wild-type toxin. The tertiary structures of the furin cleavage sites of the wild-type and W281 mutant toxins are similar; however, the mutant toxin has significantly higher B-factors around the cleavage site, suggesting that the greater susceptibility to furin cleavage is due to increased local disorder/flexibility at the site, rather than to differences in static tertiary structure. Comparison of the refined structures of full-length toxin, which lacks ADP-ribosyl transferase activity, to that of the enzymatic domain alone reveals a salt bridge between Arg467 of the catalytic domain and Glu348 of domain II that restrains the substrate binding cleft in a conformation that precludes NAD+ binding. The refined structures of exotoxin A provide precise models for the design and interpretation of further studies of the mechanism of intoxication.  相似文献   

7.
The interaction of 3-aminopyridine-adenine dinucleotide, an NAD + 2 analogue which is fluorescent at the pyridine end of the molecule, with rabbit muscle glyceraldehyde-3-phosphate dehydrogenase was investigated. The fluorescence properties of the AAD+ molecule were used to monitor the nicotinamide subsites ou the GPDHase tetramer, the fluorescent aminopyridine moiety of the molecule serving as an intrinsic probe. Although the binding of AAD+ wag found to be negatively co-operative, no conformational changes induced at the nicotinamide subsite upon coenzyme binding were found to be transmitted to neighboring subunits. These findings, in conjunction with our earlier findings and with the observation that different NAD+ analogues which differ in the chemistry of the pyridine moiety bind with different extents of co-operativity, enable us to offer specific roles for the nicotinamide and the adenine subsites in generating the negative co-operativity.It is suggested that the structure of the pyridine moiety of the coenzyme controls the mode of binding of the pyridine moiety to the nicotinamide subsite. This, in turn, controls the orientation of the adenine moiety with respect to its subsite, thereby determining the mode of the interactions between the adenine and its binding domain. As the propagation of conformational changes caused by these interactions to neighboring subunits is believed to be the cause of the negative co-operativity exhibited by this enzyme towards coenzyme binding, the structure of the pyridine moiety controls this phenomenon.  相似文献   

8.
ADARs are adenosine deaminases responsible for RNA-editing reactions that occur within duplex RNA. Currently little is known regarding the nature of the protein-RNA interactions that lead to site-selective adenosine deamination. We previously reported that ADAR2 induced changes in 2-aminopurine fluorescence of a modified substrate, consistent with a base-flipping mechanism. Additional data have been obtained using full-length ADAR2 and a protein comprising only the RNA binding domain (RBD) of ADAR2. The increase in 2-aminopurine fluorescence is specific to the editing site and dependent on the presence of the catalytic domain. Hydroxyl radical footprinting demonstrates that the RBD protects a region of the RNA duplex around the editing site, suggesting a significant role for the RBD in identifying potential ADAR2 editing sites. Nucleotides near the editing site on the non-edited strand become hypersensitive to hydrolytic cleavage upon binding of ADAR2 RBD. Therefore, the RBD may assist base flipping by increasing the conformational flexibility of nucleotides in the duplex adjacent to its binding site. In addition, an increase in tryptophan fluorescence is observed when ADAR2 binds duplex RNA, suggesting a conformational change in the catalytic domain of the enzyme. Furthermore, acrylamide quenching experiments indicate that RNA binding creates heterogeneity in the solvent accessibility of ADAR2 tryptophan residues, with one out of five tryptophans more solvent-accessible in the ADAR2.RNA complex.  相似文献   

9.
The crystal structure of Escherichia coli aspartate carbamoyltransferase complexed with adenosine 5'-triphosphate (ATP) has been solved by molecular replacement and has been refined to a crystallographic residual of 0.17 at 2.6-A resolution by using the computer program X-PLOR. The unit cell dimensions of this crystal form are a = b = 122.2 A and c = 143.3 A and the space group is P321. Although the c-axis unit cell dimension is approximately 1 A longer than the corresponding dimension of the CTP-ligated P321 crystal form (c = 142.2 A), the ATP-ligated enzyme adopts a T-like quaternary structure. The base moiety of ATP interacts with residues Glu10, Ile12, and Lys60 while the ribose is near Asp19 and Lys60; the triphosphate entity is bound to Lys94, although His20 and Arg96 are nearby. We observe a higher occupancy for ATP in the allosteric site of the R1 regulatory chain in comparison to the occupancy of the R6 allosteric site. These crystallographically independent sites are related by a molecular 2-fold axis. There are other violations of the noncrystallographic symmetry that are similar to those observed in the refined CTP-ligated aspartate carbamoyltransferase structure. These infringements on the molecular symmetry might be the result of intermolecular interactions in the crystal. To ensure the most meaningful comparison with the ATP-ligated structure, we refined the previously reported CTP-bound and unligated structures to crystallographic residuals between 0.17 and 0.18 using X-PLOR. These X-PLOR refined structures are not significantly different from the initial structures that had been crystallographically refined by a restrained least-squares method. After making all possible comparisons between the CTP- and ATP-ligated and the unligated T-state structures, we find that the most significant differences are located at the allosteric sites and in small changes in the quaternary structures. At the allosteric site, the binding of CTP and ATP successively enlarges the nucleotide binding cavity, particularly in the vicinity of the base. The changes in the quaternary structure can be characterized by an increase in the separation of the catalytic trimers by approximately 0.5 A as ATP binds to the unligated T structure. On the basis of these structural studies, we discuss the relationships between the conformational differences in the allosteric site and the small changes in the quaternary structure within the T form to the possible mechanisms for CTP inhibition and ATP activation.  相似文献   

10.
The interaction between the nicotinamide adenine dinucleotide binding domain of calf spleen NAD glycohydrolase and its ligands has been studied. The use of competitive inhibitors, structurally related to different portions of the NAD molecule (i.e. adenosine and nicotinamide moieties), revealed the considerable importance of the binding between the pyrophosphate linkage and probably an arginyl residue of the active site. This interaction allows the positioning of the substrate in a conformation which permits catalysis to occur. The binding between the 2'-hydroxyl of the adenosine moiety and a residue of the active site, which exists in NAD-linked dehydrogenases, is probably missing in the calf spleen NAD glycohydrolase, based on the inhibition by salicylates, 2'-deoxyadenosine 5'-monophosphate and the hydrolysis of the 2'-deoxyadenosine analogue of NAD. The NAD glycohydrolase could be completely inactivated by 2,3-butanedione, an arginyl-modifying reagent. The reaction followed pseudo-first-order kinetics and the modification was found to be reversible. Woodward's reagent K, a reagent for carboxyl residues, partially inactivated the enzyme, which resulted in a change of the NAD glycohydrolase kinetic parameters Km and V. The inactivation rate was complicated by a parallel decomposition of the reagent.  相似文献   

11.
Horse liver alcohol dehydrogenase is a homodimer, the protomer having a coenzyme-binding domain and a catalytic domain. Using all available x-ray structures and 50 ns of molecular dynamics simulations, we investigated the mechanism of NAD+-induced domain closure. When the well-known loop at the domain interface was modeled to its conformation in the closed structure, the NAD+-induced domain closure from the open structure could be simulated with remarkable accuracy. Native interactions in the closed structure between Arg369, Arg47, His51, Ala317, Phe319, and NAD+ were seen to form at different stages during domain closure. Removal of the Arg369 side-chain charge resulted in the loss of the tendency to close, verifying that specific interactions do help drive the domains closed. Further simulations and a careful analysis of x-ray structures suggest that the loop prevents domain closure in the absence of NAD+, and a cooperative mechanism operates between the subunits for domain closure. This cooperative mechanism explains the role of the loop as a block to closure because in the absence of NAD+ it would prevent the occurrence of an unliganded closed subunit when the other subunit closes on NAD+. Simulations that started with one subunit open and one closed supported this.  相似文献   

12.
The binding of a spin-labeled AMP analog to tetrameric glyceraldehyde-3-phosphate dehydrogenase from rabbit muscle is described. The spin label, perdeuterated and 15N-substituted 4-amino-2,2,6,6-tetramethylpiperidine-1-oxyl, was attached to C-8 of AMP (C8-SL-AMP). Up to 8 equivalents of C8-SL-AMP bind per enzyme tetramer, i.e., 2 per monomer. Combining sites are the adenine subsite of the coenzyme-binding domain and the phosphate site. Glyceraldehyde 3-phosphate causes a conformational change in the enzyme that brings C8-SL-AMP molecules bound to adjacent R-axis-related subunits closer to one another by 0.2-0.3 nm and allows for spin-spin interaction between the nitroxide radicals. Similar, but less pronounced structural changes take place upon lowering the pH from 8 to 7. Addition of a single equivalent of NAD+ to a complex of the enzyme with 7.6 equivalents of C8-SL-AMP leads to the release of almost 4 C8-SL-AMP molecules. This supports our previous findings that binding of just one NAD+ molecule induces conformational changes in all four subunits.  相似文献   

13.
W Maret  M Zeppezauer 《Biochemistry》1986,25(7):1584-1588
The conformational change of horse liver alcohol dehydrogenase induced by binding of NAD+ was studied by electronic absorption spectroscopy using cobalt as a spectroscopic probe in the active site. The complex of the enzyme with NAD+ exists in an acidic and an alkaline form. The transition between the two forms proceeds through several intermediates and is controlled by an apparent pKa of 6.9. Only at pH values below this pKa can a complex between enzyme, NAD+, and Cl- be formed. The spectral changes indicate that chloride displaces the cobalt-bound water molecule in a tetracoordinate structure. We conclude that a negative charge at the active site is necessary to stabilize the closed conformation of the enzyme in the presence of NAD+. Spectral correlations are given which strongly support the postulation of a metal-bound alkoxide in the closed structure of the enzyme as an essential feature of the catalytic mechanism of horse liver alcohol dehydrogenase.  相似文献   

14.
Certain pathogenic trypanosomatids are highly dependent on glycolysis for ATP production, and hence their glycolytic enzymes, including glycerol-3-phosphate dehydrogenase (GPDH), are considered attractive drug targets. The ternary complex structure of Leishmania mexicana GPDH (LmGPDH) with dihydroxyacetone phosphate (DHAP) and NAD(+) was determined to 1.9A resolution as a further step towards understanding this enzyme's mode of action. When compared with the apo and binary complex structures, the ternary complex structure shows an 11 degrees hinge-bending motion of the C-terminal domain with respect to the N-terminal domain. In addition, residues in the C-terminal domain involved in catalysis or substrates binding show significant movements and a previously invisible five-residue loop region becomes well ordered and participates in NAD(+) binding. Unexpectedly, DHAP and NAD(+) appear to form a covalent bond, producing an adduct in the active site of LmGPDH. Modeling a ternary complex glycerol 3-phosphate (G3P) and NAD(+) with LmGPDH identified ten active site residues that are highly conserved among all GPDHs. Two lysine residues, Lys125 and Lys210, that are presumed to be critical in catalysis, were mutated resulting in greatly reduced catalytic activity. Comparison with other structurally related enzymes found by the program DALI suggested Lys210 as a key catalytic residue, which is located on a structurally conserved alpha-helix. From the results of site-directed mutagenesis, molecular modeling and comparison with related dehydrogenases, a catalytic mechanism of LmGPDH and a possible evolutionary scenario of this group of dehydrogenases are proposed.  相似文献   

15.
Comparison of crystal structures of S-adenosylhomocysteine (AdoHcy) hydrolase in the substrate-free, NAD(+) form [Hu, Y., Komoto, J., Huang, Y., Gomi, T., Ogawa, H., Takata, Y., Fujioka, M., and Takusagawa, F. (1999) Biochemistry 38, 8323-8333] and a substrate-bound, NADH form [Turner, M. A., Yuan, C.-S., Borchardt, R. T., Hershfield, M. S., Smith, G. D., and Howell, P. L. (1998) Nat. Struct. Biol. 5, 369-376] indicates large differences in the spatial arrangement of the catalytic and NAD(+) binding domains. The substrate-free, NAD(+) form exists in an "open" form with respect to catalytic and NAD(+) binding domains, whereas the substrate-bound, NADH form exists in a closed form with respect to those domains. To address whether domain closure is induced by substrate binding or its subsequent oxidation, we have measured the rotational dynamics of spectroscopic probes covalently bound to Cys(113) and Cys(421) within the catalytic and carboxyl-terminal domains. An independent domain motion is associated with the catalytic domain prior to substrate binding, suggesting the presence of a flexible hinge element between the catalytic and NAD(+) binding domains. Following binding of substrates (i.e., adenosine or neplanocin A) or a nonsubstrate (i.e., 3'-deoxyadenosine), the independent domain motion associated with the catalytic domain is essentially abolished. Likewise, there is a substantial decrease in the average hydrodynamic volume of the protein that is consistent with a reduction in the overall dimensions of the homotetrameric enzyme following substrate binding and oxidation observed in earlier crystallographic studies. Thus, the catalytic and NAD(+) binding domains are stabilized to form a closed active site through interactions with the substrate prior to substrate oxidation.  相似文献   

16.
The three-dimensional crystal structure of the NAD(+)-linked glutamate dehydrogenase from Clostridium symbiosum has been solved to 1.96 A resolution by a combination of isomorphous replacement and molecular averaging and refined to a conventional crystallographic R factor of 0.227. Each subunit in this multimeric enzyme is organised into two domains separated by a deep cleft. One domain directs the self-assembly of the molecule into a hexameric oligomer with 32 symmetry. The other domain is structurally similar to the classical dinucleotide binding fold but with the direction of one of the strands reversed. Difference Fourier analysis on the binary complex of the enzyme with NAD+ shows that the dinucleotide is bound in an extended conformation with the nicotinamide moiety deep in the cleft between the two domains. Hydrogen bonds between the carboxyamide group of the nicotinamide ring and the side chains of T209 and N240, residues conserved in all hexameric GDH sequences, provide a positive selection for the syn conformer of this ring. This results in a molecular arrangement in which the A face of the nicotinamide ring is buried against the enzyme surface and the B face is exposed, adjacent to a striking cluster of conserved residues including K89, K113, and K125. Modeling studies, correlated with chemical modification data, have implicated this region as the glutamate/2-oxoglutarate binding site and provide an explanation at the molecular level for the B type stereospecificity of the hydride transfer of GDH during the catalytic cycle.  相似文献   

17.
The fluorescence of the natural coenzyme, NADH, is used to monitor the environment of the nicotinamide moiety at the active centre of rabbit muscle glyceraldehyde-3-phosphate dehydrogenase (EC 1.2.1.12). Changes of the fluorescence quantum yield and polarization of a small amount of NADH, totally bound by an excess of enzyme, show that at half-saturation of the oligomer with NAD a conformational change is induced which affects the active centre regions of the remaining subunits. This conformational transition is not effected by adenosine diphosphoribose, suggesting that the binding of the nicotinamide moiety of NAD to two subunits is essential for the change of tertiary structure of the remaining subunits that causes the observed changes of the fluorescence properties of the ADH "tracer probe". It is suggested that this conformational transition of the oligomer is responsible for the major decrease of affinity for NAD which occurs at half-saturation, and possibly for the activation by NAD+ of the reductive dephosphorylation reaction catalysed by the enzyme. It is also suggested, by analogy with haemoglobin, that the molecular basis of the negative cooperativity may be the creation of additional intersubunit bonds during the binding of the first two NAD molecules to the tetramer, and a change from a "relaxed" quaternary structure to a "tense" structure at half-saturation.  相似文献   

18.
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   

19.
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   

20.
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   

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